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 0 毫秒
1.
The enzymic interconversion of Escherichia coli isocitrate dehydrogenase (ICDH) between the catalytically active and inactive forms is mediated through the activities of ICDH-kinase/phosphatase in response to changes in the metabolic environment. In this study, the use of mutant strains devoid of isocitrate lyase ( aceA:: Tn10 ) and pyruvate dehydrogenase activities revealed that the signal which triggers the reversible inactivation of ICDH in vivo is not directly related to acetate itself, but rather to the need to maintain high intracellular levels of isocitrate and free co-enzyme A. The use of these mutants also revealed, rather unexpectedly, that acetate grown cells contain more ICDH protein than those grown with other carbon sources and that the catalytic activity of ICDH kinase/phosphatase is in excess of cellular demands. Furthermore, this study also revealed the presence of a 50-kDa (±2 kDa) acetate-specific polypeptide, the identity of which has yet to be established.  相似文献   

2.
Abstract Cell-free extracts of the photosynthetic eubacterium Rhodomicrobium vannielii contained both NADP and NAD-linked isocitrate dehydrogenase activities. Apparent K m values of 12 μM for NADP, 0.75 mM for NAD, 9.3 μM for isocitrate (NADP utilising) and 8.2 μM for isocitrate (NAD utilising) were determined in such extracts. Four lines of evidence indicated that one enzyme was responsible for the two activities; (i) non-additivity of reaction rates in the presence of both NADP and NAD (ii) the presence of one band which stained for activity with both cofactors on non-denaturing polyacrylamide gels (iii) identical heat inactivation kinetics for both activities (iv) co-elution of both activities after ion-exchange and hydrophobic interaction chromatography. This is the first report of a eubacterial isocitrate dehydrogenase with dual cofactor specificity.  相似文献   

3.
A gene encoding NADP-dependent Ds-threo-isocitrate dehydrogenase was isolated from Haloferax volcanii genomic DNA by using a combination of polymerase chain reaction and screening of a lambda EMBL3 library. Analysis of the nucleotide sequence revealed an open reading frame of 1260 bp encoding a protein of 419 amino acids with 45837 Da molecular mass. This sequence is highly similar to previously sequenced isocitrate dehydrogenases. In the alignment of the amino acid sequences with those from several archaeal and mesophilic NADP-dependent isocitrate dehydrogenases, the residues involved in dinucleotide binding and isocitrate binding are well conserved. We have developed methods for the expression in Escherichia coli and purification of the enzyme from H. volcanii. This expression was carried out in E. coli as inclusion bodies using the cytoplasmic expression vector pET3a. The enzyme was refolded by solubilisation in 8 M urea followed by dilution into a buffer containing EDTA, MgCl(2) and 3 M NaCl. Maximal activity was obtained after several hours incubation at room temperature.  相似文献   

4.
Isocitrate dehydrogenase from Escherichia coli is regulated by a reversible phosphorylation mechanism. We report here the amino acid sequence round the phosphorylation site; this is the first such sequence to be reported for a bacterial protein kinase. The sequence does not resemble sequences phosphorylated by cyclic AMP-dependent protein kinase.  相似文献   

5.
Sequence alignment of pig mitochondrial NADP-dependent isocitrate dehydrogenase with eukaryotic (human, rat, and yeast) and Escherichia coli isocitrate dehydrogenases reveals that Tyr316 is completely conserved and is equivalent to the E. coli Tyr345, which interacts with the 2'-phosphate of NADP in the crystal structure [Hurley et al., Biochemistry 30 (1991) 8671-8678]. Lys321 is also completely conserved in the five isocitrate dehydrogenases. Either an arginine or lysine residue is found among the enzymes from other species at the position corresponding to the pig enzyme Arg314. While Arg323 is not conserved among all species, its proximity to the coenzyme site makes it a good candidate for investigation. The importance of these four amino acids to the function of pig mitochondrial NADP-isocitrate dehydrogenase was studied by site-directed mutagenesis. Mutants (R314Q, Y316F, Y316L, K321Q, and R323Q) were generated by a megaprimer polymerase chain reaction method. Wild-type and mutant enzymes were expressed in E. coli and purified to homogeneity. All mutant and wild-type enzymes exhibited comparable molecular weights indicative of the dimeric enzyme. Mutations do not cause an appreciable change in enzyme secondary structure as revealed by circular dichroism measurements. The kinetic parameters (V(max) and K(M) values) of K321Q and R323Q are similar to those of wild-type, indicating that Lys321 and Arg323 are not involved in enzyme function. R314Q exhibits a 10-fold increase in K(M) for NADP as compared to that of wild-type, while they have comparable V(max) values. These results suggest that Arg314 contributes to the affinity between the enzyme and NADP. The hydroxyl group of Tyr316 is not required for enzyme function since Y316F exhibits similar kinetic parameters to those of wild-type. Y316L shows a 4-fold increase in K(M) for NADP and a decrease in V(max) as compared to wild-type, suggesting that the aromatic ring of the Tyr of isocitrate dehydrogenase contributes to the affinity for coenzyme, as well as to catalysis. The K(i) for NAD of R314Q, Y316F, and Y316L is comparable to that of wild-type, indicating that the Arg314 and Tyr316 may be located near the 2'-phosphate of enzyme-bound NADP.  相似文献   

6.
Thr(373), Lys(374), Asp(375), and Lys(260) were chosen as site-directed mutagenesis targets within porcine NADP-dependent isocitrate dehydrogenase based on structurally corrected sequence alignment among prokaryotic and eukaryotic NADP-isocitrate dehydrogenases. Wild-type and all mutant enzymes were expressed in Escherichia coli and purified to homogeneity. These mutations do not alter the secondary structure or dimerization state of the mutants. The D375N and K260Q mutants exhibit, respectively, a 15- and 28-fold increase in K(m) for NADP, along with marked decreases in V(max) as compared to wild-type enzyme. In contrast, replacing Lys(374), which was previously proposed to contribute to apparent coenzyme affinity, does not change the enzyme's kinetic parameters. T373S exhibits similar kinetic parameters to those of wild-type while T373A and T373V mutations reduce the V(max) values of the resulting enzymes to 1 and 20%, respectively of that of wild-type. We conclude that a hydroxyl group at position 373 is required for effective enzyme function and that Asp(375) and Lys(260) are critical amino acids contributing to coenzyme affinity as well as catalysis by porcine NADP-isocitrate dehydrogenase.  相似文献   

7.
Isocitrate lyase was purified to homogeneity from Escherichia coli ML308. Its subunit Mr and native Mr were 44,670 +/- 460 and 17,000-180,000 respectively. The kinetic mechanism of the enzyme was investigated by using product and dead-end inhibitors of the cleavage and condensation reactions. The data indicated a random-order equilibrium mechanism, with formation of a ternary enzyme-isocitrate-succinate complex. In an attempt to predict the properties of isocitrate lyase in intact cells, the effects of pH, inorganic anions and potential regulatory metabolites on the enzyme were studied. The Km of the enzyme for isocitrate was 63 microM at physiological pH and in the absence of competing anions. Chloride, phosphate and sulphate ions inhibited competitively with respect to isocitrate. Phosphoenolpyruvate inhibited non-competitively with respect to isocitrate, but the Ki value suggested that this effect was unlikely to be significant in intact cells. 3-Phosphoglycerate was a competitive inhibitor. At the concentration reported to occur in intact cells, this metabolite would have a significant effect on the activity of isocitrate lyase. The available data suggest that the Km of isocitrate lyase for isocitrate is similar to the concentration of isocitrate in E. coli cells growing on acetate, about one order of magnitude higher than the Km determined in vitro in the absence of competing anions.  相似文献   

8.
Bradyrhizobium japonicum, the nitrogen-fixing symbiotic partner of soybean, was grown on various carbon substrates and assayed for the presence of the glyoxylate cycle enzymes, isocitrate lyase and malate synthase. The highest levels of isocitrate lyase [165–170 nmol min–1 (mg protein)–1] were found in cells grown on acetate or β-hydroxybutyrate, intermediate activity was found after growth on pyruvate or galactose, and very little activity was found in cells grown on arabinose, malate, or glycerol. Malate synthase activity was present in arabinose- and malate-grown cultures and increased by only 50–80% when cells were grown on acetate. B. japonicum bacteroids, harvested at four different nodule ages, showed very little isocitrate lyase activity, implying that a complete glyoxylate cycle is not functional during symbiosis. The apparent K m of isocitrate lyase for d,l-isocitrate was fourfold higher than that of isocitrate dehydrogenase (61.5 and 15.5 μM, respectively) in desalted crude extracts from acetate-grown B. japonicum. When isocitrate lyase was induced, neither the V max nor the d,l-isocitrate K m of isocitrate dehydrogenase changed, implying that isocitrate dehydrogenase is not inhibited by covalent modification to facilitate operation of the glyoxylate cycle in B. japonicum. Received: 10 October 1997 / Accepted: 16 January 1998  相似文献   

9.
We constructed mutant genes of Caldococcus noboribetus isocitrate dehydrogenase containing ancestral amino acid residues that were inferred using the maximal likelihood method and a composite phylogenetic tree of isocitrate dehydrogenase and 3-isopropylmalate dehydrogenase. The mutant genes were expressed in Escherichia coli and the protein products purified. Thermostabilities, reported as the half-inactivation temperatures, for the purified enzymes were determined and compared with that of the wild-type enzyme. Four of the five mutant enzymes have greater thermal stabilities than wild-type isocitrate dehydrogenase. The results are compatible with the hyperthermophilic universal ancestor (commonote) hypothesis. Incorporation of ancestral residues into a modern-day protein sequence can be used to improve protein thermostability.  相似文献   

10.
The activity and hormonal regulation of NAD- and NADP-linked isocitrate dehydrogenase (EC 1.1.1.41 and 1.1.1.42, respectively) in the brain and liver of rats of various ages were investigated. The activity of NAD-linked isocitrate dehydrogenase of the brain was greater than cytoplasmic or mitochondrial NADP-linked isocitrate dehydrogenase. In contrast, the cytoplasmic NADP-isocitrate dehydrogenase of the liver predominates over both NAD- and mitochondrial NADP-isocitrate dehydrogenases at the three ages studied. The activity of NAD-isocitrate dehydrogenase increased in the brain (139%) and liver (17%) of rats upt o 33 weeks of age and decreased (57 and 39%, respectively) in old rats (85-week-old). The activity of cytoplasmic NADP-isocitrate dehydrogenase was maximum in immature (6-week-old) rat brain and decreased as the age of the rats increased; whereas, in liver, the activity of this enzyme was found to be maximum in adult rats (33-week-old). Brain mitochondrial NADP-isocitrate dehydrogenase activity increased (64%) in adult rats, but in liver it decreased (45 and 33% in 33- and 85-week-old rats, respectively). In both tissues, adrenalectomy and hydrocortisone treatment showed differential age-dependent response. Hydrocortisone-mediated induction of the level of enzymes was inhibited by actinomycin D.  相似文献   

11.
By fractional extraction of minced bovine heart muscle with iso-osmotic sucrose and phosphate buffer solutions, it is shown that less than 4% of the total citrate synthase in the tissue is in the cytosol. Using citrate synthase as a marker for broken mitochondria, two methods of fractionation of 750 × g supernatants from homogenates of bovine heart muscle show that 10% of the total fumarase and NADP-linked isocitrate dehydrogenase activities are present in the cytoplasm. Homogenates prepared by sonication and osmotic shock and by sand-grinding gave closely similar results as regards enzyme distributions and extent of mitochondrial breakage. The results are compared with those reported for other tissues.  相似文献   

12.
Phosphorylated NADP+-isocitrate dehydrogenase (EC 1.1.1.42) has been purified to electrophoretic homogeneity from in vivo 32P-labeled Escherichia coli. The cells used as the source of phosphorylated enzyme were harvested 1 h after the addition of 5 mCi of [32P]orthophosphoric acid and 25 mM sodium acetate to cultures grown to early stationary phase on a low phosphate medium with limiting glucose. Double immunodiffusion and autoradiography demonstrated immunological identity between the 32P-labeled NADP+-isocitrate dehydrogenase and the enzyme isolated from glucose-grown E. coli. The phosphoenzyme had an apparent subunit molecular weight of 51,000 as determined by denaturing acrylamide gel electrophoresis in the presence of sodium dodecyl sulfate, and the radioactivity co-electrophoresed with NADP+-isocitrate dehydrogenase activity when purified enzyme was subjected to nondenaturing gel electrophoresis. [32P]Phosphoserine was identified following partial acid hydrolysis of the purified phosphoenzyme.  相似文献   

13.
Yeast peroxisomal NADP+-specific isocitrate dehydrogenase (IDP3) contains a canonical type I peroxisomal targeting sequence (a carboxyl-terminal Cys-Lys-Leu tripeptide), and provides the NADPH required for β-oxidation of some fatty acids in that organelle. Cytosolic yeast IDP2 carrying a PTS1 (IDP2+CKL) was only partially localized to peroxisomes, and the enzyme was able to function in lieu of either peroxisomal IDP3 or cytosolic IDP2. The analogous isocitrate dehydrogenase enzyme (IDPA) from Aspergillus nidulans, irrespective of the presence or absence of a putative PTS1, was found to exhibit patterns of dual compartmental distribution and of dual function in yeast similar to those observed for IDP2+CKL. To test a potential cellular limit on peroxisomal levels, authentic yeast IDP3, which is normally strictly peroxisomal, was over-expressed. This also resulted in dual distribution and function of the enzyme in both the cytosol and in peroxisomes, supporting the possibility of a restriction on organellar amounts of IDP.  相似文献   

14.
To probe the functions of multiple forms of isocitrate dehydrogenase in Saccharomyces cerevisiae, mutants lacking three of the isozymes were constructed and analyzed. Results show that, while the mitochondrial NAD+-dependent enzyme, IDH (composed of Idh1p and Idh2p subunits) is not the major contributor to total isocitrate dehydrogenase activity under any growth condition, loss of IDH produces the most dramatic growth phenotypes. These include reduced growth in the absence of glutamate, as well as an increase in expression of Idp2p (the cytosolic NADP+-dependent enzyme) under some growth conditions. In this study, we have focused on another phenotype associated with loss of IDH, an elevated frequency of petite mutations indicating loss of functional mtDNA. Using mutant forms of IDH with altered active site residues, a correlation was observed between the high frequency of petite mutations and the loss of catalytic activity. Loss of Idp1p (the mitochondrial NADP+-dependent enzyme) and Idp2p contributes to the loss of functional mtDNA, but only in an IDH dysfunctional background. Surprisingly, overexpression of Idp1p, but not of Idp2p, was found to result in an elevated petite frequency independent of the functional state of IDH. This is the first phenotype associated with altered Idp1p. Finally, throughout this study we examined effects of loss of mitochondrial citrate synthase (Cit1p) on isocitrate dehydrogenase mutants, since defects in the CIT1 gene were previously shown to enhance growth of IDH dysfunctional strains on nonfermentable carbon sources. Loss of Cit1p was found to suppress the petite phenotype of strains lacking IDH, suggesting that these phenotypes may be linked.  相似文献   

15.
NADP+-specific isocitrate dehydrogenase (EC 1.1.1.42) was purified to homogeneity from the sulfate-reducing bacterium Desulfobacter vibrioformis, and shown to be a monomeric protein with a molecular mass of 80 kDa. The pH and temperature optima were 8.5 and 45°C, respectively. The N-terminal amino acid sequence (Thr, Glu, Thr, Ile, Arg, Trp, Thr, X, Thr, Asp, Glu, Ala, Pro, Leu, Leu, Ala, Thr) showed similarity with that of other known monomeric isocitrate dehydrogenases. Catalytically active isocitrate dehydrogenase from D. vibrioformis was obtained by activity staining after SDS-PAGE and removal of SDS from the gel. This technique revealed a NADP+-dependent monomeric enzyme in other Desulfobacter spp., Desulfuromonas acetoxidans and Chlorobium tepidium. These findings imply that monomeric isocitrate dehydrogenases are present in distantly related bacteria and indicate an early evolution of monomeric isocitrate dehydrogenases in the bacterial lineage.  相似文献   

16.
A NADP+-specific isocitrate dehydrogenase (EC 1.1.1.42) was isolated and purified over 400-fold from Anacystis nidulans. The enzyme activity responded slowly to rapid changes in ligand (NADP+, isocitrate, Mg2+-ions) or enzyme concentration as well as to rapid changes in temperature. These are properties characteristic of the hysteretic enzymes. In addition, the enzyme activity was subject to product (-ketoglutarate) inhibition. ATP, ADP and CDP also inhibited the enzyme. Unlike several other cyanobacterial enzymes, the isocitrate dehydrogenase of Anacystis is not under redox control.  相似文献   

17.
Yeast mutants lacking mitochondrial NAD+-specific isocitrate dehydrogenase (idhΔ) or aconitase (aco1Δ) were found to share several growth phenotypes as well as patterns of specific protein expression that differed from the parental strain. These shared properties of idhΔ and aco1Δ strains were eliminated or moderated by co-disruption of the CIT1 gene encoding mitochondrial citrate synthase. Gas chromatography/mass spectrometry analyses indicated a particularly dramatic increase in cellular citrate levels in idhΔ and aco1Δ strains, whereas citrate levels were substantially lower in idhΔcit1Δ and aco1Δcit1Δ strains. Exogenous addition of citrate to parental strain cultures partially recapitulated effects of high endogenous levels of citrate in idhΔ and aco1Δ strains. Finally, effects of elevated cellular citrate in idhΔ and aco1Δ mutant strains were partially alleviated by addition of iron or by an increase in pH of the growth medium, suggesting that detrimental effects of citrate are due to elevated levels of the ionized form of this metabolite.  相似文献   

18.
Staphylococcus aureus a natural inhabitant of nasopharyngeal tract mainly survives as biofilms and possess complete Krebs cycle which plays major role in its pathogenesis. This TCA cycle is regulated by Isocitrate dehydrogenase (IDH) we have earlier cloned, sequenced (HM067707), expressed and characterized this enzyme from S. aureus ATCC12600. We have observed only one type of IDH in all the strains of S. aureus which dictates the flow of carbon thereby controlling the virulence and biofilm formation, this phenomenon is variable among bacteria. Therefore in the present study comparative structural and functional analysis of IDH was undertaken. As the crystal structure of S. aureus IDH was not available therefore using the deduced amino sequence of complete gene the 3D structure of IDH was built in Modeller 9v8. The PROCHECK and ProSAweb analysis showed the built structure was close to the crystal structure of Bacillus subtilis. This structure when superimposed with other bacterial IDH structures exhibited extensive structural variations as evidenced from the RMSD values correlating with extensive sequential variations. Only 24% sequence identity was observed with both human NADP dependent IDHs (PDB: 1T09 and 1T0L) and the structural comparative studies indicated extensive structural variations with an RMSD values of 14.284Å and 10.073Å respectively. Docking of isocitrate to both human IDHs and S. aureus IDH structures showed docking scores of -11.6169 and -10.973 respectively clearly indicating higher binding affinity of isocitrate to human IDH.  相似文献   

19.
The 18.1-kDa protein IIIGlc from Escherichia coli acts as both a phosphocarrier protein in the phosphoenolpyruvate:glycose phosphotransferase system (PTS) and as a signal-transducing protein with respect to the uptake of non-PTS sugars. Phosphorylation of IIIGlc at the N epsilon (N3) position of His-90 was effected through a regeneration system that included MgCl2, DTT, excess PEP, and catalytic amounts of Enzyme I and HPr. NH, 15N, and 13C alpha signal assignments for P-IIIGlc were made through comparison of 15N-1H correlation spectra (HSQC) of uniformly 15N-labeled preparations of phosphorylated and unphosphorylated protein and through analysis of three-dimensional triple-resonance HNCA spectra of P-IIIGlc uniformly labeled with both 15N and 13C. Backbone and side-chain 1H and 13C beta signals were assigned using 3D heteronuclear HCCH-COSY and HCCH-TOCSY spectra of P-IIIGlc. Using this approach, the assignments were made without reference to nuclear Overhauser effect data or assumptions regarding protein structure. The majority of NH, 15N, H alpha, and 13C alpha chemical shifts measured for P-IIIGlc were identical to those obtained for the unphosphorylated protein [Pelton, J. G., Torchia, D. A., Meadow, N. D., Wong, C.-Y., & Roseman, S. (1991) Biochemistry 30, 10043]. Those signals that exhibited shifts corresponded to residues within four segments (1) Leu-87-Gly-100, (2) Val-36-Val-46, (3) His-75-Ser-78, and (4) Ala-131-Val-138. These four segments are in close proximity to the active site residues His-75 and His-90 in the unphosphorylated protein [Worthylake, D., Meadow, N. D., Roseman, S., Liao, D., Hertzberg, O., & Remington, S.J. (1991) Proc. Natl. Acad. Sci. U.S.A. 88, 10382], and the chemical shift data provide strong evidence that if any structural changes accompany phosphorylation, they are confined to residues in these four segments. This conclusion is confirmed by comparing NOEs observed in 3D 15N/13C NOESY-HMQC spectra of the two forms of the protein. No NOE differences are seen for residues having the same chemical shifts in IIIGlc and P-IIIGlc. Furthermore, with the exception of residues Ala-76, Asp-94, and Val-96, the NOEs of residues (in the four segments) which exhibited chemical shift differences also had the same NOEs in IIIGlc and P-IIIGlc. In the case of residues Ala-76, Asp-94, and Val-96, minor differences in NOEs, corresponding to interproton distances changes of less than 1.5 A, were observed.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

20.
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