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.
Deregulation of allosteric inhibition of enzymes is a challenge for strain engineering and has been achieved so far primarily by random mutation and trial-and-error. In this work, we used aspartokinase, an important allosteric enzyme for industrial amino acids production, to demonstrate a predictive approach that combines protein dynamics and evolution for a rational reengineering of enzyme allostery. Molecular dynamic simulation of aspartokinase III (AK3) from Escherichia coli and statistical coupling analysis of protein sequences of the aspartokinase family allowed to identify a cluster of residues which are correlated during protein motion and coupled during the evolution. This cluster of residues forms an interconnected network mediating the allosteric regulation, including most of the previously reported positions mutated in feedback insensitive AK3 mutants. Beyond these mutation positions, we have successfully constructed another twelve targeted mutations of AK3 desensitized toward lysine inhibition. Six threonine-insensitive mutants of aspartokinase I-homoserine dehydrogenase I (AK1-HD1) were also created based on the predictions. The proposed approach can be widely applied for the deregulation of other allosteric enzymes.  相似文献   

2.
The activity of three enzymes, aspartokinase, homoserine dehydrogenase, and homoserine kinase, has been studied in the industrial strainSaccharomyces cerevisiae IFI256 and in the mutants derived from it that are able to overproduce methionine and/or threonine. Most of the mutants showed alteration of the kinetic properties of the enzymes aspartokinase, which was less inhibited by threonine and increased its affinity for aspartate, and homoserine dehydrogenase and homoserine kinase, which both lost affinity for homoserine. Furthermore, they showed in vitro specific activities for aspartokinase and homoserine kinase that were higher than those of the wild type, resulting in accumulation of aspartate, homoserine, threonine, and/or methionine/S-adenosyl-methionine (Ado-Met). Together with an increase in the specific activity of both aspartokinase and homoserine kinase, there was a considerable and parallel increase in methionine and threonine concentration in the mutants. Those which produced the maximal concentration of these amino acids underwent minimal aspartokinase inhibition by threonine. This supports previous data that identify aspartokinase as the main agent in the regulation of the biosynthetic pathway of these amino acids. The homoserine kinase in the mutants showed inhibition by methionine together with a lack or a reduction of the inhibition by threonine that the wild type undergoes, which finding suggests an important role for this enzyme in methionine and threonine regulation. Finally, homoserine dehydrogenase displayed very similar specific activity in the mutants and the wild type in spite of the changes observed in amino acid concentrations; this points to a minor role for this enzyme in amino acid regulation.  相似文献   

3.
Aspartokinase II from Bacillus subtilis was shown by immunochemical methods to be regulated by degradation in response to starvation of cells for various nutrients. Ammonium starvation induced the fastest aspartokinase II decline (t1/2 = 65 min), followed by amino acid starvation (t1/2 = 80 min) and glucose limitation (t1/2 = 120 min). Loss of enzyme activity was closely correlated with the disappearance of the alpha subunit; degradation of the beta subunit was somewhat delayed or slower under some conditions. Pulse-chase experiments demonstrated that aspartokinase II was stable during exponential growth; the synthesis of the enzyme rapidly declined in response to nutrient exhaustion. The degradation of aspartokinase II was interrupted by inhibitors of energy production and protein synthesis but was not changed in a mutant lacking a major intracellular protease. Mutants lacking a normal stringent response displayed only a slight decrease in the rate of aspartokinase II degradation, even though aspartate transcarbamylase was degraded more slowly in the same mutant cells. These results indicate that although energy-dependent degradation of biosynthetic enzymes is a general phenomenon in nutrient-starved B. subtilis cells, the degradation of specific enzymes probably involves different pathways.  相似文献   

4.
Structure of the yeast HOM3 gene which encodes aspartokinase   总被引:5,自引:0,他引:5  
The yeast HOM3 gene has been cloned molecularly by complementation of a HOM3 mutant. The gene is located about 8 kilobase pairs from HIS1 and is present as a single copy in the yeast genome. Mutations in HOM3 result in a requirement for threonine and methionine (or homoserine) for growth and a lack of detectable aspartokinase activity. The nucleotide sequence of HOM3 predicts an enzyme 414 amino acids long that shows homology to the three Escherichia coli aspartokinases, indicating that it is the structural gene for yeast aspartokinase. An approximately 1800-base pair mRNA is transcribed from the HOM3 gene, initiating at several start sites, 80 and 70 base pairs downstream, respectively, from two TATA boxes. Upstream of the TATA boxes is a single TGACTC sequence. This sequence has been shown to be essential for regulation of several genes that encode amino acid biosynthetic enzymes by the general control system. However, no increase in aspartokinase mRNA is observed under general control derepressing conditions.  相似文献   

5.
Two aspartokinase (ATP:L-aspartate 4-phosphotrasferase, EC 2.7.2.4) enzyme activities have been identified and partially purified from Bacillus brevis. Aspartokinase I is subject to both inhibition and repression by lysine, and has a molecular weight in the region of 110 000. Aspartokinase II is a lysine-stabilised enzyme, inhibited multivalently by lysine plus theonine and has a molecular weight in the region of 95 000. This attern of aspartokinase activity has not been described previously and is unusual in that one end product (lysine) regulates two isoenzymes catalysing the first reaction of a branced biosynthetic pathway. In the absence of lysine, aspartokinase II changes to a more unstable non-inhibitable enzyme. Both enzymes are stabilised by sulphydryl reducing agents and have similar affinities for ATP, aspartate and lysine. However, there is no evidence for a view that they are products of a common gene. Problem concerned with the regulation of aspartokinase activities in Bacillus species are discussed.  相似文献   

6.
The enzymes aspartokinase and homoserine dehydrogenase catalyze the reaction at key branching points in the aspartate pathway of amino acid biosynthesis. Enterococcus faecium has been found to contain two distinct aspartokinases and a single homoserine dehydrogenase. Aspartokinase isozymes eluted on gel filtration chromatography at molecular weights greater than 250,000 and about 125,000. The molecular weight of homoserine dehydrogenase was determined to be 220,000. One aspartokinase isozyme was slightly inhibited by meso-diaminopimelic acid. Another aspartokinase was repressed and inhibited by lysine. Although the level of diaminopimelate-sensitive (DAPs) enzyme was not much affected by growth conditions, the activity of lysine-sensitive (Lyss) aspartokinase disappeared rapidly during the stationary phase and was depressed in rich media. The synthesis of homoserine dehydrogenase was controlled by threonine and methionine. Threonine also inhibited the specific activity of this enzyme. The regulatory properties of aspartokinase isozymes and homoserine dehydrogenase from E. faecium are discussed and compared with those from Bacillus subtilis.  相似文献   

7.
We have purified homoserine dehydrogenase to homogeneity and subjected polypeptide fragments derived from digests of the protein to amino acid sequencing. The amino acid sequence of homoserine dehydrogenase from carrot (Daucus carota) indicates that in carrot both aspartokinase and homoserine dehydrogenase activities reside on the same protein. Additional evidence that aspartokinase and homoserine dehydrogenase reside on a bifunctional protein is provided by coelution of activities during purification steps and by enzyme-specific gel staining techniques. Highly purified fractions containing aspartokinase activity were stained for aspartokinase activity, homoserine dehydrogenase activity, and protein. These gels confirmed that aspartokinase activity and homoserine dehydrogenase activity were present on the same protein. This arrangement of aspartokinase and homoserine dehydrogenase activities residing on the same protein is also found in Escherichia coli, which has two bifunctional enzymes, aspartokinase I-homoserine dehydrogenase I and aspartokinase II-homoserine dehydrogenase II. The amino acid sequence of the major form of homoserine dehydrogenase from carrot cell suspension cultures most closely resembles that of the E. coli ThrA gene product aspartokinase I-homoserine dehydrogenase I.  相似文献   

8.
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   

9.
To characterize the mobilization and uptake of iron by cyanobacteria, 14 species were screened for ability to scavenge iron in a competitive system. The cyanobacteria exhibited a range of growth responses to iron limitation which could be separated into three groups, and a representative species from each group was chosen for further study. Effects of iron-limitation on growth and siderophore production of Anacystis nidulans R2, Anabaena variabilis ATCC 29413, and Plectonema boryanum UTEX 581 were determined. Both A. nidulans R2 and A. variabilis showed a reduced rate of growth with decreased available iron concentration (PFe 17–19). Growth rates increased with further reduction in the level of available iron (pFe 20 to pFe 21). The increase in growth rate occurred at the same available iron concentration as the initiation of extracellular siderophore production. In contrast, the growth of P. boryanum decreased with decreasing available iron levels. No siderophore production was detected from P. boryanum cultures. The growth kinetics of siderophore-producing species differ from traditional nutrient-limited growth kinetics and clearly reflect the presence of a high affinity, siderophore-mediated iron transport system in A. nidulans R2 and A. variabilis. Iron-limited growth kinetics more similar to traditional nutrient-limited growth kinetics were found in P. boryanum. The available nitrogen source influenced amount of siderophore produced and concentration of available iron which induced siderophore production. Siderophores were produced at high iron concentrations (pFe 18) when A. variablilis cultures were grown in the absence of combined nitrogen source. When nitrate was supplied to the culture, iron concentrations had to be reduced to pFe 20 before siderophores were produced. Cells grown on nitrogen also produced greater than two times the amount of siderophore compared with nitrate grown cells. This may be indicative of an increased demand for iron by nitrogen fixing A. variabilis Cultures.  相似文献   

10.
Metabolism of aspartate in Mycobacterium smegmatis   总被引:2,自引:0,他引:2  
Mycobacterium smegmatis grows best on L-asparagine as a sole nitrogen source; this was confirmed. [14C]Aspartate was taken up rapidly (46 nmol.mg dry cells-1.h-1 from 1 mM L-asparagine) and metabolised to CO2 as well as to amino acids synthesised through the aspartate pathway. Proportionately more radioactivity appeared in the amino acids in bacteria grown in medium containing low nitrogen. Activities of aspartokinase and homoserine dehydrogenase, the initial enzymes of the aspartate pathway, were carried by separate proteins. Aspartokinase was purified as three isoenzymes and represented up to 8% of the soluble protein of M. smegmatis. All three isoenzymes contained molecular mass subunits of 50 kDa and 11 kDa which showed no activity individually; full enzyme activity was recovered on pooling the subunits. Km values for aspartate were: aspartokinases I and III, 2.4 mM; aspartokinase II, 6.4 mM. Aspartokinase I was inhibited by threonine and homoserine and aspartokinase III by lysine, but aspartokinase II was not inhibited by any amino acids. Aspartokinase activity was repressed by methionine and lysine with a small residue of activity attributable to unrepressed aspartokinase I. Homoserine dehydrogenase activity was 96% inhibited by 2 mM threonine; isoleucine, cysteine and valine had lesser effects and in combination gave additive inhibition. Homoserine dehydrogenase was repressed by threonine and leucine. Only amino acids synthesised through the aspartate pathway were tested for inhibition and repression. Of these, only one, meso-diaminopimilate, had no discernable effect on either enzyme activity.  相似文献   

11.
James CL  Viola RE 《Biochemistry》2002,41(11):3726-3731
The direct channeling of an intermediate between enzymes that catalyze consecutive reactions in a pathway offers the possibility of an efficient, exclusive, and protected means of metabolite delivery. Aspartokinase-homoserine dehydrogenase I (AK-HDH I) from Escherichia coli is an unusual bifunctional enzyme in that it does not catalyze consecutive reactions. The potential channeling of the intermediate beta-aspartyl phosphate between the aspartokinase of this bifunctional enzyme and aspartate semialdehyde dehydrogenase (ASADH), the enzyme that catalyzes the intervening reaction, has been examined. The introduction of increasing levels of inactivated ASADH has been shown to compete against enzyme-enzyme interactions and direct intermediate channeling, leading to a decrease in the overall reaction flux through these consecutive enzymes. These same results are obtained whether these experiments are conducted with aspartokinase III, a naturally occurring monofunctional isozyme, with an artificially produced monofunctional aspartokinase I, or with a fusion construct of AK I-ASADH. These results provide definitive evidence for the channeling of beta-aspartyl phosphate between aspartokinase and aspartate semialdehyde dehydrogenase in E. coli and suggest that ASADH may provide a bridge to channel the intermediates between the non-consecutive reactions of AK-HDH I.  相似文献   

12.
The aspartate pathway of amino acid biosynthesis in bacteria serves as paradigm for the evolution of patterns of enzyme regulation in response to specific physiological requirements. InBacillus species, the first step in the pathway is catalyzed by multiple forms of aspartokinase, which differ in their structure and feedback regulation. One form of aspartokinase (V-type) functions primarily during cell growth, another form (S-type) during sporulation. The V-type aspartokinase fromBacillus subtilis andBacillus polymyxa is discussed in some detail on account of its complex pattern of regulation by the pathway endproducts lysine and threonine and its unusual subunit structure. The enzyme is composed of two dissimilar subunits, the smaller of which corresponds to the carboxyl-terminal domain of the larger subunit. The coding sequence for the subunits ofBacillus subtilis aspartokinase has recently been cloned inEscherichia coli. The study of its structure and mode of expression has revealed that the two aspartokinase subunits are encoded by in-phase overlapping genes. These unusual features of aspartokinase suggest that important aspects of the regulation of the aspartate pathway are yet to be discovered.  相似文献   

13.
The levels of two aspartokinase isozymes, a lysine-sensitive enzyme and an aspartokinase that is inhibited synergistically by lysine plus threonine, differ strikingly in different strains of Bacillus subtilis. In derivatives of B. subtilis 168 growing in minimal medium, the predominant isozyme is the lysine-sensitive aspartokinase. In B. subtilis ATCC 6051, the Marburg strain, the level of the lysine-sensitive aspartokinase is much lower during growth in minimal medium, and the major aspartokinase activity is the lysine-plus-threonine-sensitive isozyme. Molecular cloning and nucleotide sequence determination of the genes for the lysine-sensitive isozymes from the two B. subtilis strains and their upstream control regions showed these genes to be identical. Evidence that the lysine-sensitive aspartokinase, referred to as aspartokinase II, is distinct from the threonine-plus-lysine-sensitive aspartokinase comes from the observation that disruption of the aspartokinase II gene by recombinational insertion had no effect on the latter. Mutants were obtained from the aspartokinase II-negative strain that also lacked the threonine-plus-lysine-sensitive aspartokinase, which will be referred to as aspartokinase III. Aspartokinase II could be selectively restored to these mutants by transformation with plasmids carrying the aspartokinase II gene. Study of the growth properties of the various mutant strains showed that the loss of either aspartokinase II or aspartokinase III had no effect on growth in minimal medium but that the loss of both enzymes interfered with growth unless the medium was supplemented with the three major end products of the aspartate pathway. It appears, therefore, that aspartokinase I alone cannot provide adequate supplies of precursors for the synthesis of lysine, threonine, and methionine by exponentially growing cells.  相似文献   

14.
James CL  Viola RE 《Biochemistry》2002,41(11):3720-3725
The bifunctional enzyme aspartokinase-homoserine dehydrogenase I from Escherichia coli catalyzes non-consecutive reactions in the aspartate pathway of amino acid biosynthesis. Both catalytic activities are subject to allosteric regulation by the end product amino acid L-threonine. To examine the kinetics and regulation of the enzymes in this pathway, each of these catalytic domains were separately expressed and purified. The separated catalytic domains remain active, with each of their catalytic activities enhanced in comparison to the native enzyme. The allosteric regulation of the kinase activity is lost, and regulation of the dehydrogenase activity is dramatically decreased in these separate domains. To create a new bifunctional enzyme that can catalyze consecutive metabolic reactions, the aspartokinase I domain was fused to the enzyme that catalyzes the intervening reaction in the pathway, aspartate semialdehyde dehydrogenase. A hybrid bifunctional enzyme was also created between the native monofunctional aspartokinase III, an allosteric enzyme regulated by lysine, and the catalytic domain of homoserine dehydrogenase I with its regulatory interface domain still attached. In this hybrid the kinase activity remains sensitive to lysine, while the dehydrogenase activity is now regulated by both threonine and lysine. The dehydrogenase domain is less thermally stable than the kinase domain and becomes further destabilized upon removal of the regulatory domain. The more stable aspartokinase III is further stabilized against thermal denaturation in the hybrid bifunctional enzyme and was found to retain some catalytic activity even at temperatures approaching 100 degrees C.  相似文献   

15.
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   

16.
Comprehensive studies were made with Lemna paucicostata Hegelm. 6746 of the effects of combinations of lysine, methionine, and threonine on growth rates, soluble amino acid contents, aspartokinase activities, and fluxes of 4-carbon moieties from aspartate through the aspartokinase step into the amino acids of the aspartate family. These studies show that flux in vitro through the aspartokinase step is insensitive to inhibition by lysine or threonine, and confirm previous in vitro data in establishing that aspartokinase in vivo is present in two orders of magnitude excess of its requirements. No evidence of channeling of the products of the lysine- and threonine-sensitive aspartokinases was obtained, either form of the enzyme alone being more than adequate for the combined in vivo flux through the aspartokinase step. The marked insensitivity of flux through the aspartokinase step to inhibition by lysine or threonine strongly suggests that inhibition of aspartokinase by these amino acids is not normally a major factor in regulation of entry of 4-carbon units into the aspartate family of amino acids. Direct measurement of fluxes of 4-carbon units demonstrated that: (a) Lysine strongly feedback regulates its own synthesis, probably at the step catalyzed by dihydrodipicolinate synthase. (b) Threonine alone does not regulate its own synthesis in vivo, thereby confirming previous studies of the metabolism of [14C]threonine and [14C]homoserine in Lemna. This finding excludes not only aspartokinases as an important regulatory determinant of threonine synthesis, but also two other enzymes (homoserine dehydrogenase and threonine synthase) suggested to fulfill this role. Complete inhibition of threonine synthesis was observed only in the combined presence of accumulated threonine and lysine. The physiological significance of this single example of apparent regulation of flux at the aspartokinase step, albeit under unusually stringent conditions of aspartokinase inhibition, remains to be determined. (c) Isoleucine strongly inhibits its own synthesis, probably at threonine dehydratase, without causing compensatory reduction in threonine synthesis. A fundamentally changed scheme for regulation of synthesis of the aspartate family of amino acids is presented that has important implications for improvement of the nutritional contents of these amino acids in plants.  相似文献   

17.
In the green alga Pediastrum boryanum NIES-301, plastocyanin accumulates under copper-sufficient conditions and cytochrome c6 accumulates under copper-deficient conditions. We cloned the cDNA which encodes pre-apoplastocyanin from P. boryanum cultured under the copper-sufficient condition. The deduced amino acid sequence of the pre-apoplastocyanin protein consists of 151 amino acid residues including a putative bipartite presequence of 53 amino acid residues. Southern blot analysis of P. boryanum genomic DNA indicated that pre-apoplastocyanin is encoded by a single nuclear gene. Northern blot analysis showed that copper-deficient cells accumulated a shorter form of the mRNA of pre-apoplastocyanin, which did not generate pre-apoplastocyanin in the wheat-germ translation system. The difference in size was ascribed to the absence of the 5' region in the mRNA of pre-apoplastocyanin obtained from the copper-deficient cells, which accounts for the absence of plastocyanin under these conditions. This phenomenon represents a novel regulatory mechanism, although details of the mechanism are not yet known.  相似文献   

18.
The presence of a single aspartokinase was demonstrated in Rhodospirillum tenue. The enzyme has been purified about 60-fold. No physical association exists in this species between aspartokinase and homoserine dehydrogenase. The general properties of the enzyme are described. Inhibition by l-lysine, by l-threonine, and concerted inhibition by these two end products are regulatory characters which have also been found in many other species. In R. tenue, aspartokinase is also subject to a hitherto not encountered type of concerted feedback inhibition, by l-threonine plus l-methionine. The inhibition caused by lysine can be reversed either by glycine, l-isoleucine, l-methionine, or l-phenylalanine. The concerted inhibition by lysine plus threonine is reversed by glycine, l-isoleucine, or l-phenylalanine, but not by l-methionine, which exerts in conjunction with threonine the independent concerted inhibition referred to above. Addition of single or several metabolites to cultures of R. tenue caused inhibition of growth and reversal of growth inhibition, compatible with the effects observed in vitro on aspartokinase activity. The regulation of this enzyme in relation to that of other bacterial aspartokinases is discussed.  相似文献   

19.
The gene coding for the subunits of aspartokinase II from Bacillus subtilis has been identified in a B. subtilis DNA library and cloned in a bacterial plasmid (Bondaryk, R. P., and Paulus, H. (1984) J. Biol. Chem. 259, 585-591). The introduction of a plasmid carrying the aspartokinase II gene into an auxotrophic Escherichia coli strain lacking all three aspartokinases restored its ability to grow in the absence of L-lysine, L-threonine, and L-methionine. The B. subtilis aspartokinase gene could thus be functionally expressed in E. coli and substitute for the E. coli aspartokinases. Measurement of aspartokinase levels in extracts of aspartokinaseless E. coli transformed with the B. subtilis aspartokinase II gene revealed an enzyme level comparable to that in a genetically derepressed B. subtilis strain. In spite of the high level of aspartokinase, the growth of the transformed E. coli strain was severely inhibited by the addition of L-lysine but could be restored by also adding L-homoserine. This apparently paradoxical sensitivity to lysine was due to the allosteric inhibition of B. subtilis aspartokinase II by that amino acid, a property which was also observed in extracts of the transformed E. coli strain. The synthesis and degradation of the aspartokinase II subunits were measured by labeling experiments in E. coli transformed with the B. subtilis aspartokinase II gene. In contrast to exponentially growing cells of B. subtilis which contained equimolar amounts of the aspartokinase alpha and beta subunits, the transformed E. coli strain contained a 3-fold molar excess of beta subunit. Pulse-chase experiments showed that the disproportionate level of beta subunit was not due to more rapid turnover of alpha subunit, both subunits being quite stable, but presumably to a more rapid rate of synthesis. After the addition of rifampicin, the synthesis of alpha subunit declined much more rapidly than that of beta subunit, indicating that the two subunits were translated independently from mRNA species that differ in functional stability. In conjunction with the results described in the preceding paper which demonstrated that the aspartokinase subunits are encoded by a single DNA sequence, these observations imply that the alpha and beta subunits of B. subtilis aspartokinase II are the products of in-phase overlapping genes.  相似文献   

20.
The relationship between aspartokinase activity and fruiting body formation in Myxococcus xanthus was investigated. Two required amino acids, methionine and isoleucine, which stimulated the enzyme in vitro also inhibited fruiting body formation when added to 0.1% Casitone agar. Threonine, a potent feedback inhibitor of the aspartokinase, completely reversed the effects of methionine and isoleucine both on enzyme activity and fruiting body formation. A mutant, M. xanthus FB-S, which had the unusual property of forming fruiting bodies on 1.0% Casitone agar, also exhibited an altered regulation of aspartokinase activity. Spermidine, which is a strong stimulator of the enzyme in vitro, interfered with the developmental cycle of both M. xanthus FB and FS-S. During glycerol induction of myxospores the level of aspartokinase dropped more than 75% during the first hour. These data indicate a strong correlation between aspartokinase activity and the induction of the developmental cycle in M. xanthus. It is suggested that the decrease in aspartokinase activity results in diaminopimelic acid starvation, blockage of cell wall growth, and subsequent induction of the developmental cycle.  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