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.
K. Yamamoto  A. Niwa 《Amino acids》1996,10(3):263-271
Summary A subline growing in medium without arginine and ornithine was established from a rat Reuber hepatoma cell line (R-Y121B·cho). The subline designated R-Y117B·cho was able to grow in glutamine, arginine and ornithine-free, glutamate-supplemented medium. Arginine synthesis from glutamate requires four urea cycle enzymes and another two enzymes, glutamate semialdehyde dehydrogenase and ornithine aminotransferase. Since R-Y121B·cho cells have all the urea cycle enzymes, two other enzyme activities were determined. The activities of ornithine aminotransferase and glutamate semialdehyde dehydrogenase were similar in R-Y117B·cho and its parental R-Y121B·cho cells, but R-Y117B·cho cells had higher conversion of glutamate to arginine than parental cells.  相似文献   

2.
Tyrosine 984 in the juxtamembrane region of the insulin receptor, between the transmembrane helix and the cytoplasmic tyrosine kinase domain, is conserved among all insulin receptor-like proteins from hydra to humans. Crystallographic studies of the tyrosine kinase domain and proximal juxtamembrane region reveal that Tyr-984 interacts with several other conserved residues in the N-terminal lobe of the kinase domain, stabilizing a catalytically nonproductive position of alpha-helix C. Steady-state kinetics measurements on the soluble kinase domain demonstrate that replacement of Tyr-984 with phenylalanine results in a 4-fold increase in kcat in the unphosphorylated (basal state) enzyme. Moreover, mutation of Tyr-984 in the full-length insulin receptor results in significantly elevated receptor phosphorylation levels in cells, both in the absence of insulin and following insulin stimulation. These data demonstrate that Tyr-984 plays an important structural role in maintaining the quiescent, basal state of the insulin receptor. In addition, the structural studies suggest a possible target site for small molecule activators of the insulin receptor, with potential use in the treatment of noninsulin-dependent diabetes mellitus.  相似文献   

3.
We measured the metabolism of ornithine in Neurospora during the transition from minimal medium to arginine-supplemented medium. Within an hour after arginine supplementation, the amount of intracellular ornithine (95% of which had been stored in vesicles) dropped by 65%, even though the catabolism of arginine produces as much ornithine as had been produced on minimal medium. The arginine level in the cell rose 10-fold. Ornithine flux through the catabolic enzyme ornithine aminotransferase increased fivefold, but flux through the mitochondrial enzyme ornithine transcarbamylase (leading to arginine synthesis) was only 20% of the rate seen on minimal medium. During this transition to arginine catabolism, the enzymes of the arginine pathway operate as an ornithine cycle, but at a restricted rate. We suggest the hypothesis that high levels of arginine may inhibit the movement of ornithine into the vesicles and into the mitochondria.  相似文献   

4.
Metabolism of arginine in lactating rat mammary gland.   总被引:3,自引:1,他引:2       下载免费PDF全文
Significant activities of the four enzymes needed to convert arginine into proline and glutamate (arginase, ornithine aminotransferase, pyrroline-5-carboxylate reductase and pyrroline-5-carboxylate dehydrogenase) develop co-ordinately in lactating rat mammary glands in proportion to the increased production of milk. No enzymes were detected to carry out the reactions of proline oxidation or reduction of glutamate to pyrroline-5-carboxylate. Minces of the gland converted ornithine into proline and into glutamate plus glutamine. These conversions increased during the cycle of lactation in proportion to the increased milk production and to the content of the necessary enzymes. The minced gland did not convert labelled ornithine into citrulline, confirming the absence from the gland of a functioning urea cycle, and did not convert labelled proline or glutamate into ornithine. A metabolic flow of labelled arginine to proline and glutamate in mammary gland was confirmed in intact animals with experiments during which the specific radioactivity of proline in plasma remained below that of the proline being formed from labelled arginine within the gland. It was concluded that arginase in this tissue had a metabolic role in the biosynthesis of extra proline and glutamate needed for synthesis of milk proteins.  相似文献   

5.
4-Aminobutyrate aminotransferase (GABAT) from Pseudomonas aeruginosa was purified 64-fold to apparent electrophoretic homogeneity from cells grown with 4-aminobutyrate as the only source of carbon and nitrogen. Purified GABAT catalyzed the transamination of 4-aminobutyrate, N2-acetyl-L-ornithine, L-ornithine, putrescine, L-lysine, and cadaverine with 2-oxoglutarate (listed in order of decreasing activity). The enzyme is induced in cells grown on 4-guanidinobutyrate, 4-aminobutyrate, or putrescine as the only carbon and nitrogen source. Cells grown on arginine or on glutamate contained low levels of the enzyme. The regulation of the synthesis of GABAT as well as the properties of the mutant with an inactive N2-acetyl-L-ornithin 5-aminotransferase suggest that GABAT functions in the biosynthesis of arginine by convertine N2-acetyl-L-glutamate 5-semialdehyde to N2-acetyl-Lornithine as well as in catabolic reactions during growth on putrescine or 4-guanidinobutyrate but not during growth on arginine.  相似文献   

6.
Daubner SC  Fitzpatrick PF 《Biochemistry》1999,38(14):4448-4454
The active site of tyrosine hydroxylase consists of a hydrophobic cleft with an iron atom near the bottom. Within the cleft are several charged residues which are conserved across the family of pterin-dependent hydroxylases. We have studied four of these residues, glutamates 326 and 332, aspartate 328, and arginine 316 in tyrosine hydroxylase, by site-directed substitution with alternate amino acid residues. Replacement of arginine 316 with lysine results in a protein with a Ktyr value that is at least 400-fold greater and a V/Ktyr value that is 4000-fold lower than those found in the wild-type enzyme; substitution with alanine, serine, or glutamine yields insoluble enzyme. Arginine 316 is therefore critical for the binding of tyrosine. Replacement of glutamate 326 with alanine has no effect on the KM value for tyrosine and results in a 2-fold increase in the KM value for tetrahydropterin. The Vmax for DOPA production is reduced 9-fold, and the Vmax for dihydropterin formation is reduced 4-fold. These data suggest that glutamate 326 is not directly involved in catalysis. Replacement of aspartate 328 with serine results in a 26-fold higher KM value for tyrosine, a 8-fold lower Vmax for dihydropterin formation, and a 13-fold lower Vmax for DOPA formation. These data suggest that aspartate 328 has a role in tyrosine binding. Replacement of glutamate 332 with alanine results in a 10-fold higher KM value for 6-methyltetrahydropterin with no change in the KM value for tyrosine, a 125-fold lower Vmax for DOPA formation, and an only 3.3-fold lower Vmax for tetrahydropterin oxidation. These data suggest that glutamate 332 is required for productive tetrahydropterin binding.  相似文献   

7.
1. Partially purified preparations of rat brain 4-aminobutyrate aminotransferase were inhibited in a time-dependent manner by ethanolamine O-sulphate. The inhibition was not reversed by dialysis. 2. The inhibitor formed an initial reversible complex with the enzyme (K(i)=4.4x10(-4)m) and the rate of inactivation followed pseudo-first-order kinetics (k=7.15x10(-4)s(-1)). The inclusion of 4-aminobutyrate markedly slowed the rate of inactivation. 3. Ethanolamine O-sulphate did not inhibit glutamate decarboxylase, alanine aminotransferase or aspartate aminotransferase. 4. Intracisternal injection of ethanolamine O-sulphate into rats led to rapid inactivation of 4-aminobutyrate aminotransferase in vivo.  相似文献   

8.
Aromatic amino acids are protein building blocks and precursors to a number of plant natural products, such as the structural polymer lignin and a variety of medicinally relevant compounds. Plants make tyrosine and phenylalanine by a different pathway from many microbes; this pathway requires prephenate aminotransferase (PAT) as the key enzyme. Prephenate aminotransferase produces arogenate, the unique and immediate precursor for both tyrosine and phenylalanine in plants, and also has aspartate aminotransferase (AAT) activity. The molecular mechanisms governing the substrate specificity and activation or inhibition of PAT are currently unknown. Here we present the X‐ray crystal structures of the wild‐type and various mutants of PAT from Arabidopsis thaliana (AtPAT). Steady‐state kinetic and ligand‐binding analyses identified key residues, such as Glu108, that are involved in both keto acid and amino acid substrate specificities and probably contributed to the evolution of PAT activity among class Ib AAT enzymes. Structures of AtPAT mutants co‐crystallized with either α‐ketoglutarate or pyridoxamine 5′‐phosphate and glutamate further define the molecular mechanisms underlying recognition of keto acid and amino acid substrates. Furthermore, cysteine was identified as an inhibitor of PAT from A. thaliana and Antirrhinum majus plants as well as the bacterium Chlorobium tepidum, uncovering a potential new effector of PAT.  相似文献   

9.
Effects of heat shock on amino Acid metabolism of cowpea cells   总被引:8,自引:4,他引:8       下载免费PDF全文
When cowpea (Vigna unguiculata) cells maintained at 26°C are transferred to 42°C, rapid accumulation of γ-aminobutyrate (>10-fold) is induced. Several other amino acids (including β-alanine, alanine, and proline) are also accumulated, but less extensively than γ-aminobutyrate. Total free amino acid levels are increased approximately 1.5-fold after 24 hours at 42°C. Heat shock also leads to release of amino acids into the medium, indicating heat shock damage to the integrity of the plasmalemma. Some of the changes in metabolic rates associated with heat shock were estimated by monitoring the 15N labeling kinetics of free intracellular, extracellular and protein-bound amino acids of cultures supplied with 15NH4+, and analyzing the labeling data by computer simulation. Preliminary computer simulation models of nitrogen flux suggest that heat shock induces an increase in the γ-aminobutyrate synthesis rate from 12.5 nanomoles per hour per gram fresh weight in control cells maintained at 26°C, to as high as 800 nanomoles per hour per gram fresh weight within the first 2 hours of heat shock. This 64-fold increase in the γ-aminobutyrate synthesis rate greatly exceeds the expected (Q10) change of metabolic rate of 2.5- to 3-fold due to a 16°C increase in temperature. We suggest that this metabolic response may in part involve an activation of glutamate decarboxylase in vivo, perhaps mediated by a transient cytoplasmic acidification. Proline appears to be synthesized from glutamate and not from ornithine in cowpea cells. Proline became severalfold more heavily labeled than ornithine, citrulline and arginine in both control and heat-shocked cultures. Proline synthesis rate was increased 2.7-fold by heat shock. Alanine, β-alanine, valine, leucine, and isoleucine synthesis rates were increased 1.6-, 3.5-, 2.0-, 5.0-, and 6.0-fold, respectively, by heat shock. In contrast, the phenylalanine synthesis rate was decreased by 50% in response to heat shock. The differential effects of heat stress on metabolic rates lead to flux and pool size redistributions throughout the entire network of amino acid metabolism.  相似文献   

10.
C S Gibbs  M J Zoller 《Biochemistry》1991,30(22):5329-5334
"Charged-to alanine" scanning mutagenesis of the catalytic subunit of the Saccharomyces cerevisiae cAMP-dependent protein kinase (C1) identified three glutamate residues, E171, E214, and E274, that are involved in the recognition of a peptide substrate, kemptide (Leu1Arg2Arg3Ala4Ser5Leu6Gly7). These glutamate residues are conserved or conservatively substituted with asparate in the serine/threonine protein kinases that have a requirement for basic residues on the N-terminal side of their phosphorylation sites. Alanine replacement mutants in C1 were subjected to kinetic analysis using alanine-substituted peptides as substrates. The additivity or nonadditivity of the effects of the alanine substitutions on the catalytic efficiency (kcat/Km) was analyzed. This allowed the identification of electrostatic interactions between the three glutamate residues in the enzyme and the two arginine residues present in the peptide substrate. The data suggest that E171 interacts with Arg2 in the substrate and that E214 and E274 both interact with Arg3. This may be a general method for identifying simple intermolecular interactions involving proteins when there is no three-dimensional structure available of the complex of interacting species. The identification of these interactions provides the potential for rational protein engineering of enzymes with alternative specificities.  相似文献   

11.
Ungerminated pumpkin (Cucurbita moschata Poir.) cotyledons contained 30 % of their dry weight as lipid and 26 % as protein, of which 93 % was globulin. There was a rapid degradation of these reserves 4 to 8 days after planting when the cotyledons had their maximum metabolic activity. About half of the mole percent of amino acids found in the globulin reserve was in arginine, glutamate, aspartate, and their amides. The cotyledons had a large soluble pool of arginine, glutamine, glutamate, and leucine. Most amino acids increased steadily in amount in the cotyledons during germination, except glutamine, ornithine, alanine, serine, glycine, and γ-aminobutyrate and these appeared in large amounts in the translocation stream to the axis tissue. Little arginine or proline was translocated. By 10 days, when translocation had decreased, amino acids accumulated. Ornithine, γ-aminobutyrate, and aspartate were rapidly utilized in the hypocotyl, while glutamine, glycine, and alanine accumulated there. Cysteine and methionine levels were low in the reserve, trans-location stream and soluble fractions. γ-Aminobutyrate-U?14C injected into cotyledons or incubated with hypocotyls was utilized in a similar fashion. The label appeared in citric acid cycle acids and in the amino acids closely related to this cycle, but the bulk of the label appeared in CO2. The labeling pattern suggests that γ-aminobutyrate was utilized via succinate, and thus entered the citric acid cycle. A close relationship between arginine, ornithine, glutamate, and γ-aminobutyrate exists in the cotyledon with all but arginine being translocated rapidly to the axis tissue where these amino acids are rapidly metabolized.  相似文献   

12.
Eight enzymes involved in the conversion of acetylglutamate to arginine in Neurospora crassa were studied. The data indicate that of three enzymes early in the sequence, only the first, acetylglutamate kinase, is a nonorganellar enzyme. The next two, N-acetyl-gamma-glutamyl-phosphate reductase and acetylornithine aminotransferase, are in the mitochondrion, which was previously shown to contain the subsequent enzymes: acetylornithine-glutamate acetyltransferase, ornithine carbamyltransferase, and carbamyl-phosphate synthetase A (arginine specific). The last two enzymes of the pathway, argininosuccinate synthetase and argininosuccinate lyase, were previously shown to be cytosolic. All enzymes but one have low amplitudes or repression. Their levels respond little to arginine excess and are about twofold elevated (threefold for ornithine carbamyltransferase) as a result of arginine limitation in the arg-12-8 strain. No restriction of the incorporation of mitochondrial enzymes into mitochondria could be detected when the levels of these enzymes were elevated. Two enzymes, acetylglutamate kinase and carbamyl-phosphate synthetase A, which initiate the synthesis of the ornithine and guanidino moieties of arginine, respectively, show the lowest specific activities in crude extract. These enzymes display special regulatroy features. Acetylglutamate kinase, which has a typically low amplitude of repression, is subject to feedback inhibition. Carbamyl-phosphate synthetase A is wholly insensitive to arginine or citrulline in vitro or in vivo, but displays a very large amplitude of repression (about 60-fold). It is unique in that it can be almost completely repressed by growth of mycelia in excess arginine. These data suggest that mitochondrial localization may be incompatible with a mechanism of feedback inhibition by a cytosolic effector, arginine. Further, they suggest that the high repressibility of carbamyl-phosphate synthetase A compensates for its feedback insensitivity.  相似文献   

13.
Pseudouridine synthases introduce the most common RNA modification and likely use the same catalytic mechanism. Besides a catalytic aspartate residue, the contributions of other residues for catalysis of pseudouridine formation are poorly understood. Here, we have tested the role of a conserved basic residue in the active site for catalysis using the bacterial pseudouridine synthase TruB targeting U55 in tRNAs. Substitution of arginine 181 with lysine results in a 2500-fold reduction of TruB’s catalytic rate without affecting tRNA binding. Furthermore, we analyzed the function of a second-shell aspartate residue (D90) that is conserved in all TruB enzymes and interacts with C56 of tRNA. Site-directed mutagenesis, biochemical and kinetic studies reveal that this residue is not critical for substrate binding but influences catalysis significantly as replacement of D90 with glutamate or asparagine reduces the catalytic rate 30- and 50-fold, respectively. In agreement with molecular dynamics simulations of TruB wild type and TruB D90N, we propose an electrostatic network composed of the catalytic aspartate (D48), R181 and D90 that is important for catalysis by fine-tuning the D48-R181 interaction. Conserved, negatively charged residues similar to D90 are found in a number of pseudouridine synthases, suggesting that this might be a general mechanism.  相似文献   

14.
Cells of the unicellular cyanobacterium Synechocystis sp. strain PCC 6803 supplemented with micromolar concentrations of L-[(14)C]arginine took up, concentrated, and catabolized this amino acid. Metabolism of L-[(14)C]arginine generated a set of labeled amino acids that included argininosuccinate, citrulline, glutamate, glutamine, ornithine, and proline. Production of [(14)C]ornithine preceded that of [(14)C]citrulline, and the patterns of labeled amino acids were similar in cells incubated with L-[(14)C]ornithine, suggesting that the reaction of arginase, rendering ornithine and urea, is the main initial step in arginine catabolism. Ornithine followed two metabolic pathways: (i) conversion into citrulline, catalyzed by ornithine carbamoyltransferase, and then, with incorporation of aspartate, conversion into argininosuccinate, in a sort of urea cycle, and (ii) a sort of arginase pathway rendering glutamate (and glutamine) via Delta(1)pyrroline-5-carboxylate and proline. Consistently with the proposed metabolic scheme (i) an argF (ornithine carbamoyltransferase) insertional mutant was impaired in the production of [(14)C]citrulline from [(14)C]arginine; (ii) a proC (Delta(1)pyrroline-5-carboxylate reductase) insertional mutant was impaired in the production of [(14)C]proline, [(14)C]glutamate, and [(14)C]glutamine from [(14)C]arginine or [(14)C]ornithine; and (iii) a putA (proline oxidase) insertional mutant did not produce [(14)C]glutamate from L-[(14)C]arginine, L-[(14)C]ornithine, or L-[(14)C]proline. Mutation of two open reading frames (sll0228 and sll1077) putatively encoding proteins homologous to arginase indicated, however, that none of these proteins was responsible for the arginase activity detected in this cyanobacterium, and mutation of argD (N-acetylornithine aminotransferase) suggested that this transaminase is not important in the production of Delta(1)pyrroline-5-carboxylate from ornithine. The metabolic pathways proposed to explain [(14)C]arginine catabolism also provide a rationale for understanding how nitrogen is made available to the cell after mobilization of cyanophycin [multi-L-arginyl-poly(L-aspartic acid)], a reserve material unique to cyanobacteria.  相似文献   

15.
Gabaculine, a potent suicide inhibitor of ornithine aminotransferase (OAT), at a dose of 50 mg/kg inhibited this enzyme in mouse tissues and dramatically increased tissue ornithine concentrations, whether or not arginine was present in the diet. Thus even under arginine deprivation there is catabolism of ornithine which involves OAT. This was confirmed by administration of [14C]ornithine to arginine-deprived mice. Gabaculine (3-amino-2,3-dihydrobenzoic acid) drastically decreased the release of 14CO2 and increased the radioactivity in the basic amino acids in the tissues. When [1-14C]glutamate was injected into mice deprived of arginine, a significant amount of radioactivity was recovered in tissue ornithine and arginine, and gabaculine decreased this labelling by about two-thirds, indicating that ornithine was synthesized in vivo from glutamate via OAT. In addition, we failed to detect in liver and small intestine alpha-N-acetylornithine, N-acetylglutamate kinase or N-acetylornithine aminotransferase, which are obligatory components of a potential route of ornithine synthesis from N-acetylglutamate. Our results indicate that at least 45 mumol of ornithine was synthesized and catabolized daily via OAT in the mouse deprived of arginine.  相似文献   

16.
Delta1-pyrroline-5-carboxylate dehydrogenase (P5CDh) catalyzes the conversion of Delta1-pyrroline-5-carboxylate to glutamate in a reaction requiring NADP+ as a cofactor. Delta1-pyrroline-5-carboxylate is formed in liver from proline by proline oxidase (EC number not assigned) or from ornithine via ornithine aminotransferase. A spectrophotometric assay for P5CDh was shown to be valid if rotenone was included in the assay to prevent reoxidation of NADH. Using this new assay, liver was fractionated using differential centrifugation and the distribution of P5CDh was compared to that of appropriate marker enzymes. P5CDh is enriched only in the mitochondrial fractions, as are the mitochondrial enzymes, succinate cytochrome c reductase, proline oxidase, glutaminase, and ornithine aminotransferase. Thus, it can be concluded that P5CDh occurs only in mitochondria, not in both mitochondria and cytoplasm, as had previously been reported.  相似文献   

17.
The pyridoxal-P binding sites of the two isoforms of human glutamate decarboxylase (GAD65 and GAD67) were modeled by using PROBE (a recently developed algorithm for multiple sequence alignment and database searching) to align the primary sequence of GAD with pyridoxal-P binding proteins of known structure. GAD's cofactor binding site is particularly interesting because GAD activity in the brain is controlled in part by a regulated interconversion of the apo- and holoenzymes. PROBE identified six motifs shared by the two GADs and four proteins of known structure: bacterial ornithine decarboxylase, dialkylglycine decarboxylase, aspartate aminotransferase, and tyrosine phenol-lyase. Five of the motifs corresponded to the alpha/beta elements and loops that form most of the conserved fold of the pyridoxal-P binding cleft of the four enzymes of known structure; the sixth motif corresponded to a helical element of the small domain that closes when the substrate binds. Eight residues that interact with pyridoxal-P and a ninth residue that lies at the interface of the large and small domains were also identified. Eleven additional conserved residues were identified and their functions were evaluated by examining the proteins of known structure. The key residues that interact directly with pyridoxal-P were identical in ornithine decarboxylase and the two GADs, thus allowing us to make a specific structural prediction of the cofactor binding site of GAD. The strong conservation of the cofactor binding site in GAD indicates that the highly regulated transition between apo- and holoGAD is accomplished by modifications in this basic fold rather than through a novel folding pattern.  相似文献   

18.
Function of arginase in lactating mammary gland   总被引:5,自引:3,他引:2  
The potential for a considerable formation of ornithine exists in lactating mammary gland because of its arginase content. Late in lactation arginase reaches an activity in the gland higher than that present in any rat tissue except liver. Occurrence of the urea cycle can be excluded since two enzymes for the further reaction of ornithine in the cycle, carbamoyl phosphate synthetase I and ornithine carbamoyltransferase, are both absent from this tissue. Instead, carbamoyl phosphate synthetase II appears early in lactation, associated with accumulation of aspartate carbamoyltransferase and DNA, consistent with the proposed role of these enzymes in pyrimidine synthesis. The facts require another physiological role for arginase apart from its known function in the urea cycle. Significant activity of ornithine aminotransferase develops in mammary gland in close parallel with the arginase. By this reaction, ornithine can be converted into glutamic semialdehyde and subsequently into proline. The enzymic composition of the lactating mammary gland is therefore appropriate for the major conversion of arginine into proline that is known to occur in the intact gland.  相似文献   

19.
The active site of Sulfolobus solfataricus aspartate aminotransferase   总被引:1,自引:0,他引:1  
Aspartate aminotransferase from the archaebacterium Sulfolobus solfataricus binds pyridoxal 5' phosphate, via an aldimine bond, with Lys-241. This residue has been identified by reducing the enzyme in the pyridoxal form with sodium cyanoboro[3H]hydride and sequencing the specifically labeled peptic peptides. The amino acid sequence centered around the coenzyme binding site is highly conserved between thermophilic aspartate aminotransferases and differs from that found in mesophilic isoenzymes. An alignment of aspartate aminotransferase from Sulfolobus solfataricus with mesophilic isoenzymes, attempted in spite of the low degree of similarity, was confirmed by the correspondence between pyridoxal 5' phosphate binding residues. Using this alignment it was possible to insert the archaebacterial aspartate aminotransferase into a subclass, subclass I, of pyridoxal 5' phosphate binding enzymes comprising mesophilic aspartate aminotransferases, tyrosine aminotransferases and histidinol phosphate aminotransferases. These enzymes share 12 invariant amino acids most of which interact with the coenzyme or with the substrates. Some enzymes of subclass I and in particular aspartate aminotransferase from Sulfolobus solfataricus, lack a positively charged residue, corresponding to Arg-292, which in pig cytosolic aspartate aminotransferase interacts with the distal carboxylate of the substrates (and determines the specificity towards dicarboxylic acids). It was confirmed that aspartate aminotransferase from Sulfolobus solfataricus does not possess any arginine residue exposed to chemical modifications responsible for the binding of omega-carboxylate of the substrates. Furthermore, it has been found that aspartate aminotransferase from Sulfolobus solfataricus is fairly active when alanine is used as substrate and that this activity is not affected by the presence of formate. The KM value of the thermophilic aspartate aminotransferase towards alanine is at least one order of magnitude lower than that of the mesophilic analogue enzymes.  相似文献   

20.
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   

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