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 31 毫秒
1.
With L-aspartate (L-Asp) as the amino donor, L-phenylalanine (L-Phe) can be prepared from phenylpyruvate (PPA) via an amination reaction mediated by aminotransferase (encoded by aspC). On the other hand, L-Asp can be produced by an aspartase (encoded by aspA) -catalyzed reaction using fumaric acid as substrate. To overproduce aspartase in Escherichia coli, the aspA gene was cloned and overexpressed 180 times over the wild-type level. The use of AspA-overproducing E. coli strain for L-Asp production exhibited an 83% conversion, approaching to the theoretical yield, whereas the wild-type strain obtained scarcely L-Asp. Furthermore, the recombinant strain overproducing both AspA and AspC was able to produce L-Asp and L-Phe simultaneously by using fumaric acid and PPA as substrates. As a result, the conversion yields obtained for L-Asp and L-Phe were 78% and 85%, respectively. In sharp contrast, the wild-type strain attained a conversion of L-Phe less than 15% and an undetectable level of L-Asp. This result illustrates a potential and attractive process to yield both L-Asp and L-Phe by coupling AspA and AspC. A further study on the repeated use of the recombinant strain immobilized with calcium alginate showed that after eight batch runs L-Asp conversion maintained roughly constant (around 75%), whereas L-Phe conversion dropped to 65% from 81%. This result indicates the stability of AspA being superior to AspC.  相似文献   

2.
In this paper we describe the cloning and sequence analysis of the tyrB and aspC genes from Escherichia coli K12, which encode the aromatic aminotransferase and aspartate aminotransferase respectively. The tyrB gene was isolated from a cosmid carrying the nearby dnaB gene, identified by its ability to complement a dnaB lesion. Deletion and linker insertion analysis located the tyrB gene to a 1.7-kilobase NruI-HindIII-digest fragment. Sequence analysis revealed a gene encoding a 43 000 Da polypeptide. The gene starts with a GTG codon and is closely followed by a structure resembling a rho independent terminator. The aspC gene was cloned by screening gene banks, prepared from a prototrophic E. coli K12 strain, for plasmids able to complement the aspC tyrB lesions in the aminotransferase-deficient strain HW225. Sub-cloning and deletion analysis located the aspC gene on a 1.8-kilobase HincII-StuI-digest fragment. Sequence analysis revealed the presence of a gene encoding a 43 000 Da protein, the sequence of which is identical with that previously obtained for the aspartate aminotransferase from E. coli B. Considerable overproduction of the two enzymes was demonstrated. We compared the deduced protein sequences with those of the pig mitochondrial and cytoplasmic aspartate aminotransferases. From the extensive homology observed we are able to propose that the two E. coli enzymes possess subunit structures, subunit interactions and coenzyme-binding and substrate-binding sites that are very similar both to each other and to those of the mammalian enzymes and therefore must also have very similar catalytic mechanisms. Comparison of the aspC and tyrB gene sequences reveals that they appear to have diverged as much as is possible within the constraints of functionality and codon usage.  相似文献   

3.
The tyrB gene of E. coli K-12, which encodes aromatic amino acid aminotransferase (EC 2.6.1.57) was cloned. The nucleotide sequence of about 2 kilobase pairs containing the gene was determined. The coding region of the tyrB gene and the deduced amino acid sequence revealed that the aromatic amino acid aminotransferase of E. coli is homologous with the aspartate aminotransferase.  相似文献   

4.
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   

5.
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   

6.
The nucleotide sequence of the aspartate aminotransferase [EC 2.6.1.1] structural gene, aspC, of Escherichia coli K-12 was determined. The coding region of the aspC gene contained 1,188 nucleotide residues and encoded 396 amino acid residues. The amino acid sequence deduced from the nucleotide sequence agreed perfectly with that of the protein recently determined for the aspartate aminotransferase of E. coli B (Kondo, K., Wakabayashi, S., Yagi, T., & Kagamiyama, H. (1984) Biochem. Biophys. Res. Commun. 122, 62-67).  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.
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   

9.
The pathway construction for biosynthesis of aromatic amino acids in Escherichia coli is atypical of the phylogenetic subdivision of gram-negative bacteria to which it belongs (R. A. Jensen, Mol. Biol. Evol. 2:92-108, 1985). Related organisms possess second pathways to phenylalanine and tyrosine which depend upon the expression of a monofunctional chorismate mutase (CM-F) and cyclohexadienyl dehydratase (CDT). Some enteric bacteria, unlike E. coli, possess either CM-F or CDT. These essentially cryptic remnants of an ancestral pathway can be a latent source of biochemical potential under certain conditions. As one example of advantageous biochemical potential, the presence of CM-F in Salmonella typhimurium increases the capacity for prephenate accumulation in a tyrA auxotroph. We report the finding that a significant fraction of the latter prephenate is transaminated to L-arogenate. The tyrA19 mutant is now the organism of choice for isolation of L-arogenate, uncomplicated by the presence of other cyclohexadienyl products coaccumulated by a Neurospora crassa mutant that had previously served as the prime biological source of L-arogenate. Prephenate aminotransferase activity was not conferred by a discrete enzyme, but rather was found to be synonymous with the combined activities of aspartate aminotransferase (aspC), aromatic aminotransferase (tyrB), and branched-chain aminotransferase (ilvE). This conclusion was confirmed by results obtained with combinations of aspC-, tyrB-, and ilvE-deficient mutations in E. coli. An example of disadvantageous biochemical potential is the presence of a cryptic CDT in Klebsiella pneumoniae, where a mutant carrying multiple enzyme blocks is the standard organism used for accumulation and isolation of chorismate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The pathway construction for biosynthesis of aromatic amino acids in Escherichia coli is atypical of the phylogenetic subdivision of gram-negative bacteria to which it belongs (R. A. Jensen, Mol. Biol. Evol. 2:92-108, 1985). Related organisms possess second pathways to phenylalanine and tyrosine which depend upon the expression of a monofunctional chorismate mutase (CM-F) and cyclohexadienyl dehydratase (CDT). Some enteric bacteria, unlike E. coli, possess either CM-F or CDT. These essentially cryptic remnants of an ancestral pathway can be a latent source of biochemical potential under certain conditions. As one example of advantageous biochemical potential, the presence of CM-F in Salmonella typhimurium increases the capacity for prephenate accumulation in a tyrA auxotroph. We report the finding that a significant fraction of the latter prephenate is transaminated to L-arogenate. The tyrA19 mutant is now the organism of choice for isolation of L-arogenate, uncomplicated by the presence of other cyclohexadienyl products coaccumulated by a Neurospora crassa mutant that had previously served as the prime biological source of L-arogenate. Prephenate aminotransferase activity was not conferred by a discrete enzyme, but rather was found to be synonymous with the combined activities of aspartate aminotransferase (aspC), aromatic aminotransferase (tyrB), and branched-chain aminotransferase (ilvE). This conclusion was confirmed by results obtained with combinations of aspC-, tyrB-, and ilvE-deficient mutations in E. coli. An example of disadvantageous biochemical potential is the presence of a cryptic CDT in Klebsiella pneumoniae, where a mutant carrying multiple enzyme blocks is the standard organism used for accumulation and isolation of chorismate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
Strains of Escherichia coli that lack the branched-chain amino acid amino-transferase because of mutations in the ilvE gene had no growth requirement for leucine when the cells contained the aromatic amino acid aminotransferase that is the product of the tyrB gene. The presence of leucine increased the generation time of these cells and decreased the specific activity of the aromatic amino acid aminotransferase. It is concluded that this enzyme functions efficiently in leucine biosynthesis and can be repressed by leucine as well as by tyrosine.  相似文献   

12.
Modeling the three-dimensional structures of bacterial aminotransferases   总被引:1,自引:0,他引:1  
M Seville  M G Vincent  K Hahn 《Biochemistry》1988,27(22):8344-8349
The refined crystallographic structure of the "closed" conformation of chicken mitochondrial aspartate aminotransferase has been used as a template for the construction of models of the two Escherichia coli aminotransferases encoded by the tyrB and aspC genes. The main results are as follows: (1) Only minor changes are required in the coordinates of the backbone atoms to accommodate the large number of substituted side chains. (2) All deletions and insertions required to allow maximum primary sequence alignment are on the solvent-accessible surface. (3) Charged residues are all located on the surface, in contact with solvent, except for certain conserved active site residues. (4) The close packing within the hydrophobic core is maintained. (5) The interactions between the subunits are maintained. (6) Modeling of tyrosine as an external aldimine into the active sites points to several residues that could be involved in determining the substrate specificities of these aminotransferases.  相似文献   

13.
For the simultaneous synthesis of enatiomerically pure (S)-amino acids and (R)-amines from corresponding alpha-keto acids and racemic amines, an alpha/omega-transaminase coupled reaction system was designed using favorable reaction equilibrium shift led by omega-transaminase reaction. Cloned tyrB, aspC and avtA, and omegataA were co-expressed in E. coli BL21(DE3) using pET23b(+) and pET24ma, respectively. The coupled reaction produced the (S)-amino acids with 73-90% (> 99% ee(S)) of conversion yield and resolved the racemic amines with 83-99% ee(R) for 5 to 10 hours. In designing the coupled reactions in the cell, alanine and pyruvate were efficiently used in the cell as an amine donor for the alanine transaminase and an amino acceptor for the omega-transaminase, respectively, resulting in an alanine-pyruvate shuttling system. The common problem of the low equilibrium constant of the alpha-transaminase can be efficiently overcome by the coupling with the omega-transaminase. However, overcoming the product inhibition of omega-transaminase by the ketone by-product and increasing the decarboxylation rate of the oxaloacetate produced during the transaminase reaction become barriers to further improving the overall reaction rate and the yield of the coupled reactions.  相似文献   

14.
In lactococci, transamination is the first step of the enzymatic conversion of aromatic and branched-chain amino acids to aroma compounds. In previous work we purified and biochemically characterized the major aromatic aminotransferase (AraT) of a Lactococcus lactis subsp. cremoris strain. Here we characterized the corresponding gene and evaluated the role of AraT in the biosynthesis of amino acids and in the conversion of amino acids to aroma compounds. Amino acid sequence homologies with other aminotransferases showed that the enzyme belongs to a new subclass of the aminotransferase I subfamily gamma; AraT is the best-characterized representative of this new aromatic-amino-acid-specific subclass. We demonstrated that AraT plays a major role in the conversion of aromatic amino acids to aroma compounds, since gene inactivation almost completely prevented the degradation of these amino acids. It is also highly involved in methionine and leucine conversion. AraT also has a major physiological role in the biosynthesis of phenylalanine and tyrosine, since gene inactivation weakly slowed down growth on medium without phenylalanine and highly affected growth on every medium without tyrosine. However, another biosynthesis aromatic aminotransferase is induced in the absence of phenylalanine in the culture medium.  相似文献   

15.
The nucleotide sequences of mRNAs for the mouse mitochondrial and cytosolic aspartate aminotransferase isoenzymes (mAspAT and cAspAT) (EC 2.6.1.1) were determined from complementary DNAs. The mAspAT mRNA comprises minimally 2460 nucleotides and codes for a polypeptide of 430 amino acid residues corresponding to the precursor form of the mAspAT (pre-mAspAT). The cAspAT mRNA comprises minimally 2086 nucleotides and codes for a polypeptide of 413 amino acid residues. The region coding for the mature mAspAT and that for the cAspAT show about 53% overall homology. The former shares 49% and the latter 48% of homology, respectively, with that of the Escherichia coli aspC gene, which has been shown to code for the E. coli AspAT (Kuramitsu, S., Okuno, S., Ogawa, T., Ogawa, H., and Kagamiyama, H. (1985) J. Biochem. (Tokyo) 97, 1259-1262). When the deduced amino acid sequence of the mouse pre-mAspAT was compared with that of the pig pre-mAspAT polypeptide, we found that they share a 94% homology and that the mouse pre-mAspAT yields a presequence consisting of 29 amino acid residues and a mature mAspAT, consisting of 401 amino acid residues. These numbers and the amino acid residues present at the putative cleavage site are all in complete agreement in these two species. The deduced amino acid sequence of the mouse cAspAT shares 91% homology with that of the pig cAspAT. Comparisons of the nucleotide and deduced amino acid sequences between the mouse and E. coli AspATs suggest that the mammalian mAspAT gene is more closely related to the E. coli aspC gene than is the mammalian cAspAT gene.  相似文献   

16.
G J Lees 《Life sciences》1977,20(10):1749-1762
Using low concentrations of substrates and cofactors, a comparison was made of the relative rates by which aminotransferases catalysed transaminations between aromatic amino acids and aromatic or aliphatic keto acids. Tryptophan aminotransferase in homogenates of rat midbrain and liver transaminated phenylpyruvate at a rate 70 to 150-fold greater than the rate with α-ketoglutarate at low concentrations of substrates. Phenylalanine aminotransferase in liver and midbrain also was more active with aromatic keto acids than with aliphatic keto acids. However, tyrosine aminotransferase in dialysed homogenates of midbrain transaminated α-ketoglutarate and phenylpyruvate at approximately equal rates. Fresh homogenates of midbrain contained an inhibitor which markedly decreased tyrosine aminotransferase activity with α-ketoglutarate but not with phenylpyruvate. Tyrosine aminotransferase in homogenates of rat liver transaminated α-ketoglutarate and phenylpyruvate at equal rates below 10 μM keto acid, but above 10 μM, transamination of α-ketoglutarate was favoured. With homogenates of liver, transamination of α-ketoglutarate, but not phenylpyruvate, by tyrosine was increased 650% by exogenous pyridoxal phosphate. Since tryptophan aminotransferase in the brain may compete with tryptophan hydroxylase for available tryptophan, a comparison was made of the relative activities of tryptophan hydroxylase and tryptophan aminotransferase. At concentrations above 7.5 μM phenylpyruvate, transamination was 8 to 17-fold greater than the rate of hydroxylation of 50 μM tryptophan.  相似文献   

17.
Although several high-resolution X-ray crystallographic structures have been determined for Escherichia coli aspartate aminotransferase (eAATase), efforts to crystallize E. coli tyrosine aminotransferase (eTATase) have been unsuccessful. Sequence alignment analyses of eTATase and eAATase show 43% sequence identity and 72% sequence similarity, allowing for conservative substitutions. The high similarity of the two sequences indicates that both enzymes must have similar secondary and tertiary structures. Six active site residues of eAATase were targeted by homology modeling as being important for aromatic amino acid reactivity with eTATase. Two of these positions (Thr 109 and Asn 297) are invariant in all known aspartate aminotransferase enzymes, but differ in eTATase (Ser 109 and Ser 297). The other four positions (Val 39, Lys 41, Thr 47, and Asn 69) line the active site pocket of eAATase and are replaced by amino acids with more hydrophobic side chains in eTATase (Leu 39, Tyr 41, Ile 47, and Leu 69). These six positions in eAATase were mutated by site-directed mutagenesis to the corresponding amino acids found in eTATase in an attempt to redesign the substrate specificity of eAATase to that of eTATase. Five combinations of the individual mutations were obtained from mutagenesis reactions. The redesigned eAATase mutant containing all six mutations (Hex) displays second-order rate constants for the transamination of aspartate and phenylalanine that are within an order of magnitude of those observed for eTATase. Thus, the reactivity of eAATase with phenylalanine was increased by over three orders of magnitude without sacrificing the high transamination activity with aspartate observed for both enzymes.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Klebsiella aerogenes utilized aromatic amino acids as sole sources of nitrogen but not as sole sources of carbon. K. aerogenes abstracted the alpha-amino group of these compounds by transamination and excreted the arylpyruvate portions into the medium. When tryptophan was utilized as the sole source of nitrogen by K. aerogenes, indolepyruvate was excreted into the medium, where it polymerized non-enzymatically to form a brick red pigment. At least four separate aromatic aminotransferase activities were found in K. aerogenes. One activity (aromatic aminotransferase I) appeared to be solely responsible for the aminotransferase reaction necessary for the growth of K. aerogenes when tryptophan was the source of nitrogen; the loss of this activity by mutation (tut) prevented the growth of cells on media containing this and other aromatic amino acids. None of the other aminotransferase activities in the cells could substitute for aromatic aminotransferase in this regard. Tryptophan-dependent pigment formation in K. aerogenes was positively controlled by the intracellular level of glutamine synthetase. Nevertheless, the aromatic aminotransferase activity in cells varied less than 2-fold in response to 10-fold or greater changes in the levels of glutamine synthetase. Glutamine synthetase affected the ability of the cells to take up tryptophan from the medium.  相似文献   

19.
L-2-Aminobutyric acid was synthesised in a transamination reaction from L-threonine and L-aspartic acid as substrates in a whole cell biotransformation using recombinant Escherichia coli K12. The cells contained the cloned genes tyrB, ilvA and alsS which respectively encode tyrosine aminotransferase of E. coli, threonine deaminase of E. coli and alpha-acetolactate synthase of B. subtilis 168. The 2-aminobutyric acid was produced by the action of the aminotransferase on 2-ketobutyrate and L-aspartate. The 2-ketobutyrate is generated in situ from L-threonine by the action of the deaminase, and the pyruvate by-product is eliminated by the acetolactate synthase. The concerted action of the three enzymes offers significant yield and purity advantages over the process using the transaminase alone with an eight to tenfold increase in the ratio of product to the major impurity.  相似文献   

20.
Site-directed mutagenesis was performed to change the substrate specificity of Escherichia coli aspartate aminotransferase (AAT). A double mutant, R292E/L18H, with a 12.9-fold increase in the specific activity toward L-lysine and 2-oxo-4-phenylbutanoic acid (OPBA) was identified. E. coli cells expressing this mutant enzyme could convert OPBA to L-homophenylalanine (L-HPA) with 97% yield and more than 99.9% ee using L-lysine as amino donor. The transamination product of L-lysine, 2-keto-6-aminocaproate, was cyclized nonenzymatically to form Delta(1)-piperideine 2-carboxylic acid in the reaction mixture. The low solubility of L-HPA and spontaneous cyclization of 2-keto-6-aminocaproate drove the reaction completely toward L-HPA production. This is the first aminotransferase process using L-lysine as inexpensive amino donor for the L-HPA production to be reported.  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