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.
2.
The allosteric transition of threonine-sensitive aspartokinase I-homoserine dehydrogenase I from Escherichia coli has been studied by time-resolved fluorescence spectroscopy. Fluorescence decay can be resolved into 2 distinct classes of tryptophan emitters: a fast component, with a lifetime of about 1.5 ns; and a slow component, with a lifetime of about 4.5 ns. The fluorescence properties of the slow component are modified by the allosteric transition. In the T-form of the enzyme stabilized by threonine, the lifetime of the slow component is longer, with a red-shifted spectrum; its accessibility to quenching by acrylamide becomes slightly higher without any decrease of fluorescence anisotropy. These results indicate a change in polarity of the slow component environment. The quaternary structure change associated with the allosteric transition probably involves global movements of structural domains without leading to any local mobility on the nanosecond time-scale. We suggest that the slow component corresponds to the unique tryptophan of the buried kinase domain.  相似文献   

3.
The preparation of immunoadsorbents against aspartokinase I-homoserine dehydrogenase I from E.coli is described. In the presence of aspartate, considerably less enzyme is bound by the fixed antibodies. The fixed protein can be displaced by a protein extracted from a nonsense mutant.  相似文献   

4.
M Veron  Y Guillou  G N Cohen 《FEBS letters》1985,181(2):381-384
A proteolytic fragment (Mr approximately 25 000) carrying only the aspartokinase activity has been purified by chromatofocusing after limited proteolysis of aspartokinase I-homoserine dehydrogenase I from E.coli K12. The NH2-terminal sequence shows that it corresponds to the amino terminal peptide of the native enzyme. The results confirm a previous hypothesis about the organization of native aspartokinase I-homoserine dehydrogenase I.  相似文献   

5.
6.
Limited proteolysis of aspartokinase I-homoserine dehydrogenase I from Escherichia coli by type VI protease from Streptomyces griseus yields five proteolytic fragments, three of which are dimeric, the other two being monomeric. One of the monomeric fragments (27 kilodaltons) exhibits residual aspartokinase activity, while the second one (33 kilodaltons) possesses residual homoserine dehydrogenase activity. The smallest of the dimeric species (2 X 25 kilodaltons) is inactive; the two other dimers exhibit either only homoserine dehydrogenase activity (2 X 59 kilodaltons) or both activities (hybrid fragment, 89 + 59 kilodaltons). This characterization of the proteolytic species in terms of molecular weight, subunit structure, and activity leads to the proposal of a triglobular model for the native enzyme. In addition, the time course of the formation of the various fragments was followed by measuring enzymatic activity and performing gel electrophoretic analysis of the protein mixture at defined time intervals during proteolysis. On the basis of the results of these studies, a reaction scheme describing the succession of events during proteolysis is given.  相似文献   

7.
Based on crystal structure analysis of the Serratia nuclease and a sequence alignment of six related nucleases, conserved amino acid residues that are located in proximity to the previously identified catalytic site residue His89 were selected for a mutagenesis study. Five out of 12 amino acid residues analyzed turned out to be of particular importance for the catalytic activity of the enzyme: Arg57, Arg87, His89, Asn119 and Glu127. Their replacement by alanine, for example, resulted in mutant proteins of very low activity, < 1% of the activity of the wild-type enzyme. Steady-state kinetic analysis of the mutant proteins demonstrates that some of these mutants are predominantly affected in their kcat, others in their Km. These results and the determination of the pH and metal ion dependence of selected mutant proteins were used for a tentative assignment for the function of these amino acid residues in the mechanism of phosphodiester bond cleavage by the Serratia nuclease.  相似文献   

8.
A thermostable mutant of kanamycin nucleotidyltransferase (KNTase) with a single amino acid replacement of Asp at position 80 by Tyr has been isolated by a novel screening method in a previous study [Matsumura, M. & Aiba, S. (1985) J. Biol. Chem. 260, 15298-15303]. To elucidate the role of Tyr80 in stabilizing the enzyme, the KNTase gene was modified by site-directed mutagenesis so that the codon for Asp80 of the wild type was replaced by that for Ser, Thr, Ala, Val, Leu, Phe and Trp, respectively. The eight mutant KNTases including Tyr80 were all purified, as well as the wild-type enzyme. The heat-inactivation rate constants were determined at 58 degrees C and the half-life values were found to be correlated with the hydrophobicity of the amino acid residues replaced at the unique position. The Gibbs energy change of unfolding in water of KNTase assessed from urea denaturation (25 degrees C, pH 7.0) was also found to be correlated with hydrophobicity. The results suggest that different amino acids at position 80 of KNTase contribute to the stability of the protein by hydrophobic interactions. In the case of tyrosine at position 80 the unusually high stability of the enzyme compared to the Phe80 enzyme suggests that the hydroxyl group also contributes to the conformational stability.  相似文献   

9.
10.
The sequence of the first 25 residues of the homoserine dehydrogenase fragment, produced by limited proteolysis of aspartokinase I-homoserine dehydrogenase I with substilisin, has been determined. The sequence of a cyanogen bromide peptide (CB5, 59 residues), isolated from the entire protein, is also presented. Residues 1 to 18 of the subtilisin homoserine dehydrogenase fragment match the sequence 42 to 59 of peptide CB5.  相似文献   

11.
12.
The threonine sensitive aspartokinase-homoserine dehydrogenase devoid of aspartokinase activity has been extracted from a missense mutant of E. coli K12 and some of its properties have been investigated. The genetic localization of the corresponding mutation indicated that the amino acid replacement lies in the kinase region of the molecule. The cooperativity of threonine inhibition of the homoserine dehydrogenase activity is lowered. The measurement of the molecular weight of the enzyme in presence or absence of threonine indicates that the molecule dissociates more easily than the wild type enzyme. These results are discussed in view of the recent structural model proposed for aspartokinase I-homoserine dehydrogenase I.  相似文献   

13.
The nucleotide sequence of the Serratia marcescens threonine operon (thrA1A2BC) was determined. Three long open reading frames were identified; these open reading frames code for aspartokinase I (AKI)-homoserine dehydrogenase I (HDI), homoserine kinase, and threonine synthase, in that order. The predicted amino acid sequences of these enzymes were similar to the amino acid sequences of the corresponding enzymes in Escherichia coli. The AKI-HDI protein is apparently a tetramer composed of monomer polypeptides that are 819 amino acids long. A deletion analysis revealed that the central and C-terminal region was responsible for threonine-resistant HDI activity, a monomeric fragment extending from the N terminus to residue 306 was responsible for threonine-resistant AKI activity, and an N-terminal portion containing 468 residues was responsible for threonine-sensitive AKI activity. The thrA(1)1A(2)1 and thrA(1)5A(2)5 mutations of threonine-excreting strains HNr21 and TLr156, which result in the loss of threonine-mediated feedback inhibition of both AKI activity and HDI activity, cause single amino acid substitutions (Gly to Asp at position 330 and Ser to Phe at position 352, respectively) in the central region of the AKI-HDI protein. The thrA1+A(2)2 mutation of strain HNr59, which results in a threonine-sensitive AKI and a threonine-resistant HDI, also causes a single amino acid substitution (Ala to Thr at position 479).  相似文献   

14.
S-2-Aminoethyl cysteine (AEC) reduced both growth rate and final growth level of Serratia marcescens Sr41. The growth inhibition was completely reversed by lysine. AEC inhibited the activity of lysine-sensitive aspartokinase to a lesser extent than lysine. The AEC addition to the medium lowered not only the level of lysine-sensite aspartokinase but also those of homoserine dehydrogenase and threonine deaminase, whereas lysine repressed the aspartokinase alone. To select mutations releasing lysine-sensitive aspartokinase from feedback controls, AEC-resistant colonies were isolated from strains HNr31 and HNr53, both of which were previously found to excrete threonine on the minimal plates but not on the plates containing excess lysine. Two of 280 resistant colonies excreted large amounts of threonine. Strains AECr174 and AECr301, derived from strains HNr31 and HNr53, respectively, lacked both feedback inhibition and repression of lysine-sensitive aspartokinase. These strains produced about 7 mg of threonine per ml in the medium containing glucose and urea.  相似文献   

15.
S-2-Aminoethyl cysteine (AEC) reduced both growth rate and final growth level of Serratia marcescens Sr41. The growth inhibition was completely reversed by lysine. AEC inhibited the activity of lysine-sensitive aspartokinase to a lesser extent than lysine. The AEC addition to the medium lowered not only the level of lysine-sensite aspartokinase but also those of homoserine dehydrogenase and threonine deaminase, whereas lysine repressed the aspartokinase alone. To select mutations releasing lysine-sensitive aspartokinase from feedback controls, AEC-resistant colonies were isolated from strains HNr31 and HNr53, both of which were previously found to excrete threonine on the minimal plates but not on the plates containing excess lysine. Two of 280 resistant colonies excreted large amounts of threonine. Strains AECr174 and AECr301, derived from strains HNr31 and HNr53, respectively, lacked both feedback inhibition and repression of lysine-sensitive aspartokinase. These strains produced about 7 mg of threonine per ml in the medium containing glucose and urea.  相似文献   

16.
17.
The catalytically essential amino acid, histidine 176, in the active site of Escherichia coli glyceraldehyde-3-phosphate dehydrogenase (GAPDH) has been replaced with an asparagine residue by site-directed mutagenesis. The role of histidine 176 as a chemical activator, enhancing the reactivity of the thiol group of cysteine 149, has been demonstrated, with iodoacetamide as a probe. The esterolytic properties of GAPDH, illustrated by the hydrolysis of p-nitrophenyl acetate, have been also studied. The kinetic results favor a role for histidine 176 not only as a chemical activator of cysteine 149 but also as a hydrogen donor facilitating the formation of tetrahedral intermediates. These results support the hypothesis that histidine 176 plays a similar role during the oxidative phosphorylation of glyceraldehyde 3-phosphate.  相似文献   

18.
19.
In earlier attempts to shift the substrate specificity of glutamate dehydrogenase (GDH) in favour of monocarboxylic amino-acid substrates, the active-site residues K89 and S380 were replaced by leucine and valine, respectively, which occupy corresponding positions in leucine dehydrogenase. In the GDH framework, however, the mutation S380V caused a steric clash. To avoid this, S380 has been replaced with alanine instead. The single mutant S380A and the combined double mutant K89L/S380A were satisfactorily overexpressed in soluble form and folded correctly as hexameric enzymes. Both were purified successfully by Remazol Red dye chromatography as routinely used for wild-type GDH. The S380A mutant shows much lower activity than wild-type GDH with glutamate. Activities towards monocarboxylic substrates were only marginally altered, and the pH profile of substrate specificity was not markedly altered. In the double mutant K89L/S380A, activity towards glutamate was undetectable. Activity towards L-methionine, L-norleucine and L-norvaline, however, was measurable at pH 7.0, 8.0 and 9.0, as for wild-type GDH. Ala163 is one of the residues that lines the binding pocket for the side chain of the amino-acid substrate. To explore its importance, the three mutants A163G, K89L/A163G and K89L/S380A/A163G were constructed. All three were abundantly overexpressed and showed chromatographic behaviour identical with that of wild-type GDH. With A163G, glutamate activity was lower at pH 7.0 and 8.0, but by contrast higher at pH 9.0 than with wild-type GDH. Activities towards five aliphatic amino acids were remarkably higher than those for the wild-type enzyme at pH 8.0 and 9.0. In addition, the mutant A163G used L-aspartate and L-leucine as substrates, neither of which gave any detectable activity with wild-type GDH. Compared with wild-type GDH, the A163 mutant showed lower catalytic efficiencies and higher K(m ) values for glutamate/2-oxoglutarate at pH 7.0, but a similar k(cat)/K(m) value and lower K(m) at pH 8.0, and a nearly 22-fold lower S(0.5) (substrate concentration giving half-saturation under conditions where Michaelis-Menten kinetics does not apply) at pH 9.0. Coupling the A163G mutation with the K89L mutation markedly enhanced activity (100-1000-fold) over that of the single mutant K89L towards monocarboxylic amino acids, especially L-norleucine and L-methionine. The triple mutant K89L/S380A/A163G retained a level of activity towards monocarboxylic amino acids similar to that of the double mutant K89L/A163G, but could no longer use glutamate as substrate. In terms of natural amino-acid substrates, the triple mutant represents effective conversion of a glutamate dehydrogenase into a methionine dehydrogenase. Kinetic parameters for the reductive amination reaction are also reported. At pH 7 the triple mutant and K89L/A163G show 5 to 10-fold increased catalytic efficiency, compared with K89L, towards the novel substrates. In the oxidative deamination reaction, it is not possible to estimate k(cat) and K(m) separately, but for reductive amination the additional mutations have no significant effect on k(cat) at pH 7, and the increase in catalytic efficiency is entirely attributable to the measured decrease in K(m). At pH 8 the enhancement of catalytic efficiency with the novel substrates was much more striking (e.g. for norleucine approximately 2000-fold compared with wild-type or the K89L mutant), but it was not established whether this is also exclusively due to more favourable Michaelis constants.  相似文献   

20.
The structural gene for pyruvate kinase from Bacillus stearothermophilus has been cloned in Escherichia coli and sequenced. The open reading frame from the ATG start codon to the TAG stop codon is 1482 base-pairs and encodes a peptide of relative molecular mass 52,967. In the expression vector pKK223-3, containing the synthetic tac promoter, the gene is overexpressed in E. coli cells to an estimated level of 30% total soluble cell protein. A purification procedure for the overexpressed protein has been established. The construction and characterization of a pair of mutant proteins has given insight into the structural basis of allosteric regulation in the tetrameric enzyme. Substituting tryptophan for tyrosine at position 466 (mutant Trp466-->Tyr) resulted in an activated form of the enzyme, having a reduced K1/2 for the substrate phosphoenolpyruvate. We propose that the characteristics of this mutant might be the result of bulk removal releasing steric inhibition to the formation of an interdomain salt bridge between Asp356 and Arg444. The regulatory behaviour of the double mutant produced by making the additional substitution aspartate for glutamate at position 356 (Trp466-->Tyr/Asp356-->Glu) corroborates this. The position of the salt bridge is such that it might be pivotal to the conformation of a pocket that is proposed to open up when the active R-conformation is adopted. We suggest that the mechanism of activation of B. stearothermophilus pyruvate kinase by ribose-5-phosphate might hinge on an interaction with, or indirectly through, residue Trp466, removing it from the vicinity of the potential salt bridge between Asp356 and Arg444 and thus effecting a closing together of the protein structure concomitant with an opening up of the pocket region.  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