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.
Two new mutations are described which, together, eliminate essentially all the aminotransferase activity required for de novo biosynthesis of tyrosine, phenylalanine, and aspartic acid in a K-12 strain of Escherichia coli. One mutation, designated tyrB, lies at about 80 min on the E. coli map and inactivates the "tyrosine-repressible" tyrosine/phenylalanine aminotransferase. The second mutation, aspC, maps at about 20 min and inactivates a nonrespressible aspartate aminotransferase that also has activity on the aromatic amino acids. In ilvE- strains, which lack the branched-chain amino acid aminotransferase, the presence of either the tyrosine-repressible aminotransferase or the aspartate aminotransferase is sufficient for growth in the absence of exogenous tyrosine, phenylalanine, or aspartate; the tyrosine-repressible enzyme is also active in leucine biosynthesis. The ilvE gene product alone can reverse a phenylalanine requirement. Biochemical studies on extracts of strains carrying combinations of these aminotransferase mutations confirm the existence of two distinct enzymes with overlapping specificities for the alpha-keto acid analogues of tyrosine, phenylalanine, and aspartate. These enzymes can be distinguished by electrophoretic mobilities, by kinetic parameters using various substrates, and by a difference in tyrosine repressibility. In extracts of an ilvE- tyrB- aspC- triple mutant, no aminotransferase activity for the alpha-keto acids of tyrosine, phenylalanine, or aspartate could be detected.  相似文献   

3.
Aminotransferases, which catalyze the last step of biosynthesis of most amino acids and the first step of their catabolism, may be involved in the growth of Lactococcus lactis in milk. Previously, we isolated two aminotransferases from L. lactis, AraT and BcaT, which are responsible for the transamination of aromatic amino acids, branched-chain amino acids, and methionine. In this study, we demonstrated that double inactivation of AraT and BcaT strongly reduced the growth of L. lactis in milk. Supplementation of milk with amino acids and keto acids that are substrates of both aminotransferases did not improve the growth of the double mutant. On the contrary, supplementation of milk with isoleucine or a dipeptide containing isoleucine almost totally inhibited the growth of the double mutant, while it did not affect or only slightly affected the growth of the wild-type strain. These results suggest that AraT and BcaT play a major role in the growth of L. lactis in milk by degrading the intracellular excess isoleucine, which is responsible for the growth inhibition. The growth inhibition by isoleucine is likely to be due to CodY repression of the proteolytic system, which is necessary for maximal growth of L. lactis in milk, since the growth of the CodY mutant was not affected by addition of isoleucine to milk. Moreover, we demonstrated that AraT and BcaT are part of the CodY regulon and therefore are regulated by nutritional factors, such as the carbohydrate and nitrogen sources.  相似文献   

4.
Aminotransferases, which catalyze the last step of biosynthesis of most amino acids and the first step of their catabolism, may be involved in the growth of Lactococcus lactis in milk. Previously, we isolated two aminotransferases from L. lactis, AraT and BcaT, which are responsible for the transamination of aromatic amino acids, branched-chain amino acids, and methionine. In this study, we demonstrated that double inactivation of AraT and BcaT strongly reduced the growth of L. lactis in milk. Supplementation of milk with amino acids and keto acids that are substrates of both aminotransferases did not improve the growth of the double mutant. On the contrary, supplementation of milk with isoleucine or a dipeptide containing isoleucine almost totally inhibited the growth of the double mutant, while it did not affect or only slightly affected the growth of the wild-type strain. These results suggest that AraT and BcaT play a major role in the growth of L. lactis in milk by degrading the intracellular excess isoleucine, which is responsible for the growth inhibition. The growth inhibition by isoleucine is likely to be due to CodY repression of the proteolytic system, which is necessary for maximal growth of L. lactis in milk, since the growth of the CodY mutant was not affected by addition of isoleucine to milk. Moreover, we demonstrated that AraT and BcaT are part of the CodY regulon and therefore are regulated by nutritional factors, such as the carbohydrate and nitrogen sources.  相似文献   

5.
Tryptophan was found to be degraded in Saccharomyces cerevisiae mainly to tryptophol. Upon chromatography on DEAE-cellulose two aminotransferases were identified: Aromatic aminotransferase I was constitutively synthesized and was active in vitro with tryptophan, phenylalanine or tyrosine as amino donors and pyruvate, phenylpyruvate or 2-oxoglutarate as amino acceptors. The enzyme was six times less active with and had a twenty times lower affinity for tryptophan (K m=6 mM) than phenylalanine or tyrosine. It was postulated thus that aromatic aminotransferase I is involved in vivo in the last step of tyrosine and phenylalanine biosynthesis. Aromatic aminotransferase II was inducible with tryptophan but also with the other two aromatic amino acids either alone or in combinations. With tryptophan as amino donor the enzyme was most active with phenylpyruvate and not active with 2-oxoglutarate as amino acceptor; its affinity for tryptophan was similar as for the other aromatic amino acids (K m=0.2–0.4 mM). Aromatic aminotransferase II was postulated to be involved in vivo mainly in the degradation of tryptophan, but may play also a role in the degradation of the other aromatic amino acids.A mutant strain defective in the aromatic aminotransferase II (aat2) was isolated and its influence on tryptophan accumulation and pool was studied. In combination with mutations trp2 fbr, aro7 and cdr1-1, mutation aat2 led to a threefold increase of the tryptophan pool as compared to a strain with an intact aromatic aminotransferase II.  相似文献   

6.
Bacterial indoleacetic acid (IAA) production, which has been proposed to play a role in the Rhizobium-legume symbiosis, is a poorly understood process. Previous data have suggested that IAA biosynthesis in Rhizobium meliloti can occur through an indolepyruvate intermediate derived from tryptophan by an aminotransferase activity. To further examine this biosynthetic pathway, the aromatic aminotransferase (AAT) activity of Rhizobium meliloti 102F34 (F34) was characterized. At least four proteins were detected on nondenaturing gels of F34 protein extracts that exhibited AAT activity. All four of these AATs were constitutively produced and utilized the aromatic amino acids tryptophan, phenylalanine, and tyrosine as amino substrates. Two AATs were also capable of using aspartate. Plasmids from an F34 gene bank were identified that coded for the synthesis of at least three of these proteins, and the respective gene sequences were localized by transposon mutagenesis. Selected transposon insertions were recombined into the F34 genome to produce strains defective in two of these proteins (AAT1 and AAT2). Characterization of the mutants revealed that neither was essential for the biosynthesis of IAA in the absence of exogenous tryptophan, but that both contributed to IAA biosynthesis when high levels of exogenous tryptophan were present. AAT1 and AAT2 were also not required for the production of a minimal level of aromatic amino acids, but both were able to scavenge nitrogen from the aromatic amino acids during nitrogen deprivation. Neither AAT1 nor AAT2 was essential for symbiosis with alfalfa.  相似文献   

7.
The genes encoding aromatic aminotransferase II (AroAT II) and aspartate aminotransferase (AspAT) from Pyrococcus furiosus have been identified, expressed in Escherichia coli and the recombinant proteins characterized. The AroAT II enzyme was specific for the transamination reaction of the aromatic amino acids, and uses a-ketoglutarate as the amino acceptor. Like the previously characterized AroAT I, AroAT II has highest efficiency for phenylalanine (k(cat)/Km = 923 s(-1) mM(-1)). Northern blot analyses revealed that AroAT I was mainly expressed when tryptone was the primary carbon and energy source. Although the expression was significantly lower, a similar trend was observed for AroAT II. These observations suggest that both AroATs are involved in amino acid degradation. Although AspAT exhibited highest activity with aspartate and alpha-ketoglutarate (k(cat) approximately 105 s(-1)), it also showed significant activity with alanine, glutamate and the aromatic amino acids. With aspartate as the amino donor, AspAT catalyzed the amination of alpha-ketoglutarate, pyruvate and phenyl-pyruvate. No activity was detected with either branched-chain amino acids or alpha-keto acids. The AspAT gene (aspC) was expressed as a polycistronic message as part of the aro operon, with expression observed only when the aromatic amino acids were absent from the growth medium, indicating a role in the biosynthesis of the aromatic amino acids.  相似文献   

8.
The enzymatic degradation of amino acids in cheese is believed to generate aroma compounds and therefore to be involved in the complex process of cheese flavor development. In lactococci, transamination is the first step in the degradation of aromatic and branched-chain amino acids which are precursors of aroma compounds. Here, the major aromatic amino acid aminotransferase of a Lactococcus lactis subsp. cremoris strain was purified and characterized. The enzyme transaminates the aromatic amino acids, leucine, and methionine. It uses the ketoacids corresponding to these amino acids and alpha-ketoglutarate as amino group acceptors. In contrast to most bacterial aromatic aminotransferases, it does not act on aspartate and does not use oxaloacetate as second substrate. It is essential for the transformation of aromatic amino acids to flavor compounds. It is a pyridoxal 5'-phosphate-dependent enzyme and is composed of two identical subunits of 43.5 kDa. The activity of the enzyme is optimal between pH 6.5 and 8 and between 35 and 45 degrees C, but it is still active under cheese-ripening conditions.  相似文献   

9.
Abstract Epimastigotes of Trypanosoma cruzi in culture produce and excrete into the medium small amounts of phenyllactic acid and p -hydrocyphenyllactic acids, presumbly arising from the catabolism of the aromatic amino acids phenylalanine and tyrosine, respectively. This production might constitute a minor pathway for the reoxidation of cytosolic NADH, through the concerted action of tyrosine aminotransferase and aromatic α-hydroxyacid dehydrogenase.  相似文献   

10.
A cell line of Eschscholtzia californica selected for meta-fluorotyrosine (MFT) tolerance was found to have 10-fold increased levels of phenylalanine and tyrosine compared to the parent line, while most other amino acids were only increased 2-fold. Tracer experiments with shikimic acid in the presence of MFT showed that the biosynthesis of the aromatic amino acids was not impaired in the tolerant line. Feeding experiments with phenylalanine, tyrosine, or shikimic acid also revealed a reduced turnover of the pools of the aromatic amino acids in the variant. Thus undisturbed de novo biosynthesis of the aromatic amino acids and dilution of toxic effects of MFT by the enlarged pool sizes seemed to be the main reason for the acquired tolerance. Despite the enlarged availability of the precursor tyrosine, formation of the benzophenanthridine alkaloids was enhanced neither in the growth nor in the production medium.  相似文献   

11.
A subfamily I aminotransferase gene homologue containing an open reading frame encoding 381 amino acid residues (Mr=42,271) has been identified in the process of the genome project of an extremely thermophilic bacterium, Thermus thermophilus HB8. Alignment of the predicted amino acid sequence using FASTA shows that this protein is a member of aminotransferase subfamily Igamma. The protein shows around 40% identity with both T. thermophilus aspartate aminotransferase [EC 2.6.1.1] and mammalian glutamine:phenylpyruvate aminotransferase [EC 2.6.1.64]. The recombinant protein expressed in Escherichia coli is a homodimer with a subunit molecular weight of 42,000, has one pyridoxal 5'-phosphate per subunit, and is highly active toward glutamine, methionine, aromatic amino acids, and corresponding keto acids, but has no preference for alanine and dicarboxylic amino acids. These substrate specificities are similar to those described for mammalian glutamine: phenylpyruvate aminotransferase. This is the first enzyme reported so far that has the glutamine aminotransferase activity in non-eukaryotic cells. As the presence of aromatic amino acid:2-oxoglutarate aminotransferase [EC 2.6.1.57] has not been reported in T. thermophilus, this enzyme is expected to catalyze the last transamination step of phenylalanine and tyrosine biosynthesis. It may also be involved in the methionine regeneration pathway associated with polyamine biosynthesis. The enzyme shows a strikingly high pKa value (9.3) of the coenzyme Schiff base in comparison with other subfamily I aminotransferases. The origin of this unique pKa value and the substrate specificity is discussed based on the previous crystallographic data of T. thermophilus and E. coli aspartate aminotransferases.  相似文献   

12.
An aminotransferase which catalyzes the final step in methionine recycling from methylthioadenosine, the conversion of alpha-ketomethiobutyrate to methionine, has been purified from Klebsiella pneumoniae and characterized. The enzyme was found to be a homodimer of 45-kDa subunits, and it catalyzed methionine formation primarily using aromatic amino acids and glutamate as the amino donors. Histidine, leucine, asparagine, and arginine were also functional amino donors but to a lesser extent. The N-terminal amino acid sequence of the enzyme was determined and found to be almost identical to the N-terminal sequence of both the Escherichia coli and Salmonella typhimurium tyrosine aminotransferases (tyrB gene products). The structural gene for the tyrosine aminotransferase was cloned from K. pneumoniae and expressed in E. coli. The deduced amino acid sequence displayed 83, 80, 38, and 34% identity to the tyrosine aminotransferases from E. coli, S. typhimurium, Paracoccus denitrificans, and Rhizobium meliloti, respectively, but it showed less than 13% identity to any characterized eukaryotic tyrosine aminotransferase. Structural motifs around key invariant residues placed the K. pneumoniae enzyme within the Ia subfamily of aminotransferases. Kinetic analysis of the aminotransferase showed that reactions of an aromatic amino acid with alpha-ketomethiobutyrate and of glutamate with alpha-ketomethiobutyrate proceed as favorably as the well-known reactions of tyrosine with alpha-ketoglutarate and tyrosine with oxaloacetate normally associated with tyrosine aminotransferases. The aminotransferase was inhibited by the aminooxy compounds canaline and carboxymethoxylamine but not by substrate analogues, such as nitrotyrosine or nitrophenylalanine.  相似文献   

13.
Two enzymes which transaminate tyrosine and phenylalanine in Bacillus subtilis were each purified over 200-fold and partially characterized. One of the enzymes, termed histidinol phosphate aminotransferase, is also active with imidazole acetyl phosphate as the amino group recipient. Previous studies have shown that mutants lacking this enzyme require histidine for growth. Mutants in the other enzyme termed aromatic aminotransferase are prototrophs. Neither enzyme is active on any other substrate involved in amino acid synthesis. The two enzymes can be distinguished by a number of criteria. Gel filtration analysis indicate the aromatic and histidinol phosphate aminotransferases have molecular weights of 63,500 and 33,000, respectively. Histidinol phosphate aminotransferase is heat-sensitive, whereas aromatic aminotransferase is relatively heat-stable, particularly in the presence of alpha-ketoglutarate. Both enzymes display typical Michaelis-Menten kinetics in their rates of reaction. The two enzymes have similar pH optima and employ a ping-pong mechanism of action. The Km values for various substrates suggest that histidinol phosphate aminotransferase is the predominant enzyme responsible for the transamaination reactions in the synthesis of tyrosine and phenylalanine. This enzyme has a 4-fold higher affinity for tyrosine and phenylalanine than does the aromatic aminotransferase. Competitive substrate inhibition was observed between tyrosine, phenylalanine, and histidinol phosphate for histidinol phosphate aminotransferase. The significance of the fact that an enzyme of histidine synthesis plays an important role in aromatic amino acid synthesis is discussed.  相似文献   

14.
This paper reports the first isolation of Saccharomyces cerevisiae mutants lacking aromatic aminotransferase I activity (aro8), and of aro8 aro9 double mutants which are auxotrophic for both phenylalanine and tyrosine, because the second mutation, aro9, affects aromatic aminotransferase II. Neither of the single mutants displays any nutritional requirement on minimal ammonia medium. In vitro, aromatic aminotransferase I is active not only with the aromatic amino acids, but also with methionine, α-aminoadipate, and leucine when phenylpyruvate is the amino acceptor, and in the reverse reactions with their oxo-acid analogues and phenylalanine as the amino donor. Its contribution amounts to half of the glutamate:2-oxoadipate activity detected in cell-free extracts and the enzyme might be identical to one of the two known α-aminoadipate aminotransferases. Aromatic aminotransferase I has properties of a general aminotransferase which, like several aminotransferases of Escherichia coli, may be able to play a role in several otherwise unrelated metabolic pathways. Aromatic aminotransferase II also has a broader substrate specificity than initially described. In particular, it is responsible for all the measured kynurenine aminotransferase activity. Mutants lacking this activity grow very slowly on kynurenine medium.  相似文献   

15.
In histidine biosynthesis, histidinol-phosphate aminotransferase catalyzes the transfer of the amino group from glutamate to imidazole acetol-phosphate producing 2-oxoglutarate and histidinol phosphate. In some organisms such as the hyperthermophile Thermotoga maritima, specific tyrosine and aromatic amino acid transaminases have not been identified to date, suggesting an additional role for histidinol-phosphate aminotransferase in other transamination reactions generating aromatic amino acids. To gain insight into the specific function of this transaminase, we have determined its crystal structure in the absence of any ligand except phosphate, in the presence of covalently bound pyridoxal 5'-phosphate, of the coenzyme histidinol phosphate adduct, and of pyridoxamine 5'-phosphate. The enzyme accepts histidinol phosphate, tyrosine, tryptophan, and phenylalanine, but not histidine, as substrates. The structures provide a model of how these different substrates could be accommodated by histidinol-phosphate aminotransferase. Some of the structural features of the enzyme are more preserved between the T. maritima enzyme and a related threonine-phosphate decarboxylase from S. typhimurium than with histidinol-phosphate aminotransferases from different organisms.  相似文献   

16.
Tyrosine, added to the growth medium of a strain of Escherichia coli K-12 lacking transaminase B, repressed the tyrosine, phenylalanine, and tryptophan aminotransferase activities while leaving the aspartate aminotransferase activity unchanged. This suggested that the aspartate and the aromatic aminotransferase activities, previously believed to reside in the same protein, viz. transaminase A, are actually nonidentical. Further experiments showed that, upon incubation at 55 C, the aspartate aminotransferase of crude extracts was almost completely stable, whereas the tyrosine and phenylalanine activities were rapidly inactivated. Apoenzyme formation was faster, and apoenzyme degradation proceeded more slowly with aspartate aminotransferase than with tyrosine aminotransferase. Electrophoresis in polyacrylamide gels separated the aminotransferases. A more rapidly moving band contained tyrosine, phenylalanine, and tryptophan aminotransferases, and a slower band contained aspartate aminotransferase. A mutant of E. coli K-12 with low levels of aspartate aminotransferase exhibited unchanged levels of tyrosine aminotransferase. Thus, transaminase A appears to be made up of at least two proteins: one of broad specificity whose synthesis is repressed by tyrosine and another, specific for aspartate, which is not subject to repression by amino acids. The apparent molecular weights of both the aspartate and the aromatic aminotransferases, determined by gel filtration, were about 100,000.  相似文献   

17.
This paper reports the first isolation of Saccharomyces cerevisiae mutants lacking aromatic aminotransferase I activity (aro8), and of aro8 aro9 double mutants which are auxotrophic for both phenylalanine and tyrosine, because the second mutation, aro9, affects aromatic aminotransferase II. Neither of the single mutants displays any nutritional requirement on minimal ammonia medium. In vitro, aromatic aminotransferase I is active not only with the aromatic amino acids, but also with methionine, α-aminoadipate, and leucine when phenylpyruvate is the amino acceptor, and in the reverse reactions with their oxo-acid analogues and phenylalanine as the amino donor. Its contribution amounts to half of the glutamate:2-oxoadipate activity detected in cell-free extracts and the enzyme might be identical to one of the two known α-aminoadipate aminotransferases. Aromatic aminotransferase I has properties of a general aminotransferase which, like several aminotransferases of Escherichia coli, may be able to play a role in several otherwise unrelated metabolic pathways. Aromatic aminotransferase II also has a broader substrate specificity than initially described. In particular, it is responsible for all the measured kynurenine aminotransferase activity. Mutants lacking this activity grow very slowly on kynurenine medium. Received: 21 October 1996 / Accepted: 23 September 1997  相似文献   

18.
We describe the complete purification of aromatic aminotransferase I, the enzyme responsible for the ability of Klebsiella aerogenes to use tryptophan and phenylalanine as sole sources of nitrogen, as well as the partial purification of aromatic aminotransferase IV. An examination of the properties of these enzymes revealed that aminotransferase I had much greater affinity for the aromatic amino acids than aminotransferase IV, explaining the essential role of aminotransferase I in the utilization of exogenously supplied aromatic amino acids. The properties of aminotransferase IV suggest that this enzyme is actually an aspartate aminotransferase (EC 2.6.1.1), corresponding to the product of the aspC gene of Escherichia coli.  相似文献   

19.
Incubating chloridazon-degrading bacteria with L-phenylalanine leads to the accumulation of L-2,3-dihydroxyphenylalanine, o-tyrosine and m-tyrosine in the medium. Incubating the bacteria with N-acetyl-L-phenylalanine leads to N-acetyl-(2,3-dihydroxyphenyl)alanine. Using phenylacetic acid as substrate leads to the accumulation of malonic acid. The products are isolated by gel chromatography and high performance liquid chromatography. 2,3-Dihydroxy-L-phenylalanine is attacked by a catechol 2,3-dioxygenase in the presence of Fe2. An unstable yellow compound is formed in this reaction. This meta-cleavage-product is again cleaved by a hydrolase, leading to aspartic acid and 4-hydroxy-2-oxovaleric acid. Both products were isolated fromthe reaction buffer by amino acid analysis and high performance liquid chromatography. The dioxygenase and hydrolase were partially purified and characterized. A new degradation pathway for phenylalanine is discussed and compared with known pathways. The enzymes chorismate mutase, prephenate dehydratase and prephenate dehydrogenase are characterized and inhibition as well as repression are investigated. Only prephenate dehydrogenase is inhibited by phenylalanine, tyrosine and tryptophane. Chorismate mutase is repressed by phenylalanine, prephenate dehydrogenase by phenylalanine and tyrosine. Prephenate dehydratase is not repressed by aromatic amino acids. Regulation of aromatic amino acid biosynthesis in connection with phenylalanine degradation is discussed.  相似文献   

20.
Mutants have been isolated which correspond to every step concerned with the biosynthesis of the aromatic amino acids in Bacillus subtilis. Each mutant has been characterized, and the lesion it bore was analyzed by deoxyribonucleic acid transformation and PBS-1 mediated transduction. The biochemical analysis revealed that each of the mutations appears to have affected a single enzyme, except for two groups of pleiotropic mutations. All aroF mutants (chorismic acid synthetase) lack dehydroquinic acid synthetase (aroB) activity. The gene that specifies aroB is closely linked to the gene coding for the aroF enzyme. Both genes are a part of the aro cluster. Mutants lacking chorismate mutase activity also lack d-arabino-heptulosonic acid-7-phosphate synthetase and shikimate kinase activity, presumably as a result of these three activities forming a multi-enzyme complex. Another mutant, previously undescribed, had been isolated. The affected gene codes for the tyrosine and phenylalanine aminotransferase activity. All of the mutations have been located on the B. subtilis genome except those in the genes specifying shikimate kinase activity and tyrosine-phenylalanine aminotransferase activity.  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