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.
M Mevarech  W Leicht  M M Werber 《Biochemistry》1976,15(11):2383-2387
Ammonium sulfate fractionation of proteins from extremely halophilic bacteria on Sepharose 4B, carboxymethylcellulose, diethylaminoethylcellulose, and hexamethylenediamine-Agarose is described. Halophilic proteins are absorbed on these gels at 2.5 M ammonium sulfate and eluted by decreasing concentration gradients of this salt. The method has enabled the separation of malate dehydrogenase from glutamate dehydrogenase and aspartate aminotransferase on Sepharose 4B and the additional 15-fold purification of glutamate dehydrogenase on DEAE-cellulose. The technique is simple and convenient, operates at low cost, and possesses great power of resolution. The mechanism of adsorption is discussed and compared to previous instances of "hydrophobic chromatography". It is concluded that the retention of halophilic proteins on the polysaccharide gels at 2.5 M ammonium sulfate is due to hydrophobic interactions.  相似文献   

2.
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   

3.
The quantitative distributions of aspartate aminotransferase and glutaminase were mapped in subregions of olfactory bulb and cochlear nucleus of rat, and were compared with similar data for retina and with the distributions of their substrate and product amino acids aspartate, glutamate, and glutamine. The distributions of both enzymes paralleled that of aspartate in the olfactory bulb and that of glutamate in the cochlear nucleus. In retina (excluding inner segments), there were similarities between aspartate aminotransferase and both glutamate and aspartate distributions. The distribution of -aminobutyrate (GABA) was similar to those of both enzymes in olfactory bulb, to aspartate aminotransferase in cochlear nucleus, and to glutaminase in retina (excluding inner segments). The results are consistent with significant involvement of aspartate aminotransferase, especially the cytosolic isoenzyme, and glutaminase in accumulation of the neurotransmitter amino acids glutamate, aspartate, and GABA, although with preferential accumulation of different amino acids in different brain regions.  相似文献   

4.
Chicken liver mitochondrial aspartate aminotransferase was found to be located in the intermembrane space and bound to the inner mitochondrial membrane. Purification of two mitochondrial fractions containing aspartate aminotransferase activity was performed. Both fractions showed similar chromatographic behaviour and identical isoelectric point and molecular weight values. There were no significant differences in the general kinetic mechanism, Km values, substrates inhibition and effect of various anions on the activity of mitochondrial aspartate aminotransferase purified from both fractions.  相似文献   

5.
In the adult rat kidney, alanine aminotransferase (EC 2.6.1.2), aspartate aminotransferase (EC 2.6.1.1) and D-amino acid oxidase (EC 1.4.3.3) were measured in glomeruli, 4 parts of the proximal tubule, 2 parts of the distal tubule and in patches from the thin limb area and the papilla. These enzymes were measured in more limited parts of the nephron during postnatal development. Adult aspartate aminotransferase activities (percentage of the highest) ranged from 100 in the distal straight segment to 25 in the late part of the proximal straight segment to 10 in the thin limb and papillary area. Alanine aminotransferase (lower by a factor of 100 in absolute terms) was distributed as the mirror image of aspartate aminotransferase within proximal and distal tubules. D-Amino acid oxidase was 850-fold higher in proximal straight segments than in medullary structures. During development alanine aminotransferase increased 6-fold and D-amino acid oxidase, 4.5-fold in proximal straight tubules but aspartate aminotransferase increased in distal straight tubles 8-fold.  相似文献   

6.
SYNOPSIS. Culture stages (promastigotes) of Leishmania tarentolae were tested for alanine aminotransferase (E.C.2.6.1.2) and aspartate aminotransferase (E.C.2.6.1.1.). Neither enzyme was detected in crude cell extracts. After starch block electrophoresis, however, both transaminase activities were found in proteins migrating toward the anode. Only one species of each enzyme was found. Using coupled enzyme assay systems, the following physical and kinetic properties were seen: 1) aspartate aminotransferase was inhibited by α-ketoglutarate concentrations above 1.68 × 10?2 M and alanine aminotransferase was inhibited by concentrations higher than 1.34 × 10?2 M; 2) the Michaelis constant (Km[α-ketoglutarate]) was 5.4 × 10?4 M for aspartate aminotransferase and 3.0 × 10?4 M for alanine aminotransferase; 3) maximum activity was found at ?pH 8.5 (broad range between pH 7.75–9.0) for aspartate aminotransferase whereas maximum activity for alanine aminotransferase was ?pH 7.2 (range between pH 7.0–7.5); 4) both enzymes lost half of their activity after 4 days at 8 C; 5) aspartate aminotransferase was most active at 35 C and completely inactivated at 59.5 C, alanine aminotransferase exhibited maximum activity at 29.5 C and was completely inactivated at 61 C; and 6) neither enzyme showed enhanced activity with added pyridoxal phosphate.  相似文献   

7.
The transaminase inhibitor l-2-amino-4-methoxy-trans-3-butenoic acid (AMB) decreased aspartate aminotransferase activity by approximately two-thirds in isolated rat liver mitohondria incubated with succinate, ammonia, and ornithine. Aspartate production by the mitochondria was unaffected over the 30-min incubation period, indicating that mitochondrial aspartate aminotransferase activity is normally far in excess of that required for maximal rates of aspartate production. In rat hepatocytes incubated with lactate, ammonia, and ornithine the inhibition of both the cytosolic and mitochondrial isozymes of aspartate aminotransferase by AMB was partially blocked by the presence of ammonia and ornithine. When pyruvate was substituted for lactate as a carbon source with isolated hepatocytes, the presence of ammonia and ornithine blocked the inhibition by AMB of the mitochondrial but not the cytosolic isozyme of aspartate aminotransferase. Urea formation by cells incubated with lactate, ammonia, and ornithine was unaffected by AMB unless the cells were preincubated with the inhibitor prior to the addition of substrates. However, urea formation by cells incubated in the presence of pyruvate, ammonia, and ornithine was inhibited strongly by AMB even without preincubation. The results suggest that the stimulation of ureogenesis from ammonia and ornithine by pyruvate involves the cytosolic isozyme of aspartate aminotransferase. In contrast, the stimulation of ureogenesis elicited by lactate primarily involved mitochondrial aspartate aminotransferase.  相似文献   

8.
Monoamine oxidase (MAO) and aspartate aminotransferase activities in cellular fractions of thyroid gland in rats after hypophysectomy and TSH treatment were investigated. MAO and aspartate aminotransferase activities in thyroid mitochondria were decreased after hypophysectomy and significantly increased after daily injection of TSH during five days to hypophysectomized rats. In microsomes after hypophysectomy a similar decrease of MAO activity was found but TSH was without effect on this activity. In conclusion: it is evident that MAO and aspartate aminotransferase in thyroid mitochondria the enzymes which could be a source for hydrogen peroxide to catalyze thyroid hormone synthesis are under the regulatory influence of TSH.  相似文献   

9.
A mutant of Rhizobium meliloti, 4R3, which is unable to grow on aspartate has been isolated. The defect is specific to aspartate utilization, since 4R3 is not an auxotroph and grows as well as its parent strain on other carbon and nitrogen sources. The defect was correlated with an inability to fix nitrogen within nodules formed on alfalfa. Transport of aspartate into the mutant cells was found to be normal. Analysis of enzymes involved in aspartate catabolism showed a significantly lower level of aspartate aminotransferase activity in cell extracts of 4R3 than in the wild type. Two unrelated regions identified from a genomic cosmid bank each complemented the aspartate catabolism and symbiotic defects in 4R3. One of the cosmids was found to encode an aspartate aminotransferase enzyme and resulted in restoration of aspartate aminotransferase activity in the mutant. Analysis of the region cloned in this cosmid by transposon mutagenesis showed that mutations within this region generate the original mutant phenotypes. The second type of cosmid was found to encode an aromatic aminotransferase enzyme and resulted in highly elevated levels of aromatic aminotransferase activity. This enzyme apparently compensated for the mutation by its ability to partially utilize aspartate as a substrate. These findings demonstrate that R. meliloti contains an aspartate aminotransferase activity required for symbiotic nitrogen fixation and implicate aspartate as an essential substrate for bacteria in the nodule.  相似文献   

10.
In previous studies it was found that: (a) aspartate aminotransferase increases the aspartate dehydrogenase activity of glutamate dehydrogenase; (b) the pyridoxamine-P form of this aminotransferase can form an enzyme-enzyme complex with glutamate dehydrogenase; and (c) the pyridoxamine-P form can be dehydrogenated to the pyridoxal-P form by glutamate dehydrogenase. It was therefore concluded (Fahien, L.A., and Smith, S.E. (1974) J. Biol. Chem 249, 2696-2703) that in the aspartate dehydrogenase reaction, aspartate converts the aminotransferase into the pyridoxamine-P form which is then dehydrogenated by glutamate dehydrogenase. The present results support this mechanism and essentially exclude the possibility that aspartate actually reacts with glutamate dehydrogenase and the aminotransferase is an allosteric activator. Indeed, it was found that aspartate is actually an activator of the reaction between glutamate dehydrogenase and the pyridoxamine-P form of the aminotransferase. Aspartate also markedly activated the alanine dehydrogenase reaction catalyzed by glutamate dehydrogenase plus alanine aminotransferase and the ornithine dehydrogenase reaction catalyzed by ornithine aminotransferase plus glutamate dehydrogenase. In these latter two reactions, there is no significant conversion of aspartate to oxalecetate and other compounds tested (including oxalacetate) would not substitute for aspartate. Thus aspartate is apparently bound to glutamate dehydrogenase and this increases the reactivity of this enzyme with the pyridoxamine-P form of aminotransferases. This could be of physiological importance because aspartate enables the aspartate and ornithine dehydrogenase reactions to be catalyzed almost as rapidly by complexes between glutamate dehydrogenase and the appropriate mitochondrial aminotransferase in the absence of alpha-ketoglutarate as they are in the presence of this substrate. Furthermore, in the presence of aspartate, alpha-ketoglutarate can have little or no affect on these reactions. Consequently, in the mitochondria of some organs these reactions could be catalyzed exclusively by enzyme-enzyme complexes even in the presence of alpha-ketoglutarate. Rat liver glutamate dehydrogenase is essentially as active as thebovine liver enzyme with aminotransferases. Since the rat liver enzyme does not polymerize, this unambiguously demonstrates that monomeric forms of glutamate dehydrogenase can react with aminotransferases.  相似文献   

11.
Using analytical subcellular fractionation techniques, 12% of the total L-alanine aminotransferase activity and 26% of the total L-aspartate aminotransferase activity was localized in enterocyte mitochondria. Alanine and aspartate were products from the oxidation of glutamine and glutamate by enterocyte mitochondria. At low concentrations, malate stimulated aspartate synthesis but was inhibitory at higher concentrations. The malate inhibition of aspartate synthesis, which increased in the presence of pyruvate, was accompanied by an increase in alanine synthesis. With glutamine as substrate in the presence of pyruvate and malate, alanine synthesis was increased by 127% on addition of purified L-alanine aminotransferase, in spite of large amounts of glutamate generated. It was concluded that when pyruvate is available the important route for glutamine or glutamate oxidation by transamination was via L-alanine:2-oxoglutarate aminotransferase and not via L-aspartate:2-oxoglutarate aminotransferase. Results suggested that mitochondria may account for 50% of alanine production from glutamine in the enterocyte despite the relatively low activity of L-alanine aminotransferase therein.  相似文献   

12.
Physical interactions between pyridoxal kinase and aspartate aminotransferase were detected by means of emission anisotropy and affinity chromatography techniques. Binding of aspartate aminotransferase (apoenzymes) to pyridoxal kinase tagged with a fluorescent probe was detected by emission anisotropy measurements at pH 6.8 (150 mM KCl). Upon saturation of the kinase with the aminotransferase, the emission anisotropy increases 22%. The protein complex is characterized by a dissociation constant of 3 microM. Time-dependent emission anisotropy measurements conducted with the mixture 5-naphthylamine-1-sulfonic acid-kinase aspartate aminotransferase (apoenzyme), revealed the presence of two rotational correlation times of phi 1 = 36 and phi 2 = 62 ns. The longer correlation time is attributed to the stable protein complex. By immobilizing one enzyme (pyridoxal kinase) through interactions with pyridoxal-Sepharose, it was possible to demonstrate that aspartate aminotransferase releases pyridoxal kinase. A test of compartmentation of pyridoxal-5-phosphate within the protein complex using alkaline phosphatase as trapping agent, indicates that the cofactor generated by the catalytic action of the kinase is channeled to the apotransaminase. The main function of the stable complex formed by the kinase and the aminotransferase is to hinder the release of free pyridoxal-5-phosphate into the bulk solvent.  相似文献   

13.
To study the effect of facilitated diffusion of the intermediate metabolite, oxaloacetate, on the coupled reaction of aspartate aminotransferase (L-aspartate: 2-oxoglutarate aminotransferase, EC 2.6.1.1) and malate dehydrogenase (L-malate:NAD+ oxidoreductase, EC 1.1.1.37), these enzymes were co-immobilized on the surface of a collagen film. The kinetic properties of the immobilized enzymes were compared with those observed with the enzymes in solution. Since the reactions correspond to the cytosolic enzymes, they have been studied in the direction aspartate aminotransferase toward malate dehydrogenase. Coupled enzymes in solution showed classical behaviour. A lag-time was observed before they reached a steady state and this lag-time was dependent on the kinetic properties of the second enzyme, malate dehydrogenase. The same lag-time was observed when malate dehydrogenase in solution was coupled with aspartate aminotransferase bound to the film. When aspartate aminotransferase in solution was coupled with malate dehydrogenase bound to the collagen film, a very long lag-time was observed. Theoretical considerations showed that in the latter case, the lag-time was dependent on the kinetic properties of the second enzyme and the transport coefficient of the intermediate substrate through the boundary layer near the surface of the film. Then both enzymes were co-immobilized on the collagen film. The coupled activity of aspartate aminotransferase and malate dehydrogenase was compared for films with an activity ratio of 5 and 0.8. In both cases, a highly efficient coupling was observed. In the former case, where malate dehydrogenase was rate-limiting, 81% of this limiting activity was observed. In the latter case, aspartate aminotransferase was rate-limiting and 82% of its rate was obtained for the final product formation. The linear increase of product formation with time corresponded fairly well to the theoretical equations developed in the paper. To interpret these rate equations, one should assume that the intermediate substrate oxaloacetate formed by aspartate aminotransferase was used by malate dehydrogenase in the diffusion layer near the film, before diffusing in the bulk solution.  相似文献   

14.
The streptozotocin diabetic rat was selected as a model to study how insulin deficiency alters vitamin B6 utilization by focusing on pyridoxal phosphate levels and aspartate aminotransferase activities in liver tissues. Diabetes of 15 weeks' duration lowered plasma pyridoxal phosphate levels by 84%. Normal plasma pyridoxal phosphate was 480 pmole/ml. Fractionation of liver into mitochondrial and extramitochondrial compartments demonstrated that diabetes caused a 43% diminution in mitochondrial pyridoxal phosphate per gram of liver. There was no cytoplasmic change in these diabetic rats. Mitochondrial aspartate aminotransferase activity was decreased 53% per gram of diabetic liver and cytoplasmic aspartate aminotransferase activity was elevated 3.4-fold. Damage to diabetic mitochondria during preparation procedures could not account for the rise in cytoplasmic aspartate aminotransferase activity. Electrophoresis showed that in the diabetic cytoplasm both cathodal and anodal forms of the enzyme were elevated. Speculations concerning mitochondrial loss and cytoplasmic gain of enzyme activity as well as those on the reduction of plasma pyridoxal phosphate in the diabetic rat are presented.  相似文献   

15.
Affinity chromatography of yeast aspartate aminotransferase [l-aspartate: 2-oxoglutarate aminotransferase, EC 2.6.1.1] on N′(ω-aminohexyl) pyridoxamine-5-phosphate Sepharose 4B is reported. The specific activity of the enzyme obtained, fully activated with pyridoxal-5-phosphate, was higher than that of previous preparations but the yield of purified enzyme was poor. Purification using DEAE-cellulose gave a higher yield of enzyme with lower specific activity. This preparation contained an appreciable amount of the holoenzyme. Use of sodium borohydride permitted the preparation of apoenzyme containing only 1.4% of the holo-form. Four coenzyme analogues were synthesized. These were the N′-acetyl-, the N′-methyl- and the N′-benzyloxycarbonylglycyl-pyridoxamine-5-phosphate and the O-acetylpyridoxal-5-phosphate. The three N′-substituted pyridoxamine-5-phosphate derivatives were all effective inhibitors of the enzyme, while the O-acetylpyridoxal-5-phosphate bound to the apoenzyme and gave an active enzyme.  相似文献   

16.
Kinetic measurements of the uptake of native mitochondrial aspartate aminotransferase and malate dehydrogenase into mitochondria in vitro were carried out. The uptake of both the enzymes is essentially complete in 1 min and shows saturation characteristics. The rate of uptake of aspartate aminotransferase into mitochondria is decreased by malate dehydrogenase, and vice versa. The inhibition is exerted by isoenzyme remaining outside the mitochondria rather than by isoenzyme that has been imported. The thiol compound beta-mercaptoethanol decreases the rate of uptake of the tested enzymes; inhibition is a result of interaction of beta-mercaptoethanol with the mitochondria and not with the enzymes themselves. The rate of uptake of aspartate aminotransferase is inhibited non-competitively by malate dehydrogenase, but competitively by beta-mercaptoethanol. The rate of uptake of malate dehydrogenase is inhibited non-competitively by aspartate aminotransferase and by beta-mercaptoethanol. beta-Mercaptoethanol prevents the inhibition of the rate of uptake of malate dehydrogenase by aspartate aminotransferase. These results are interpreted in terms of a model system in which the two isoenzymes have separate but interacting binding sites within a receptor in the mitochondrial membrane system.  相似文献   

17.
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   

18.
Aspartate: 2-oxoglutarate aminotransferase from the anaerobic protozoon Trichomonas vaginalis was purified to homogeneity and characterized. It is a dimeric protein of overall Mr approx. 100000. Only a single isoenzyme was found in T. vaginalis. The overall molecular and catalytic properties have features in common with both the vertebrate cytoplasmic and mitochondrial isoenzymes. The purified aspartate aminotransferase from T. vaginalis showed very high rates of activity with aromatic amino acids as donors and 2-oxoglutarate as acceptor. This broad-spectrum activity was restricted to aromatic amino acids and aromatic 2-oxo acids, and no significant activity was seen with other common amino acids, other than with the substrates and products of the aspartate: 2-oxoglutarate aminotransferase reaction. Co-purification and co-inhibition, by the irreversible inhibitor gostatin, of the aromatic amino acid aminotransferase and aspartate aminotransferase activities, in conjunction with competitive substrate experiments, strongly suggest that a single enzyme is responsible for both activities. Such high rates of aromatic amino acid aminotransferase activity have not been reported before in eukaryotic aspartate aminotransferase.  相似文献   

19.
Activity of aspartate aminotransferase, an enzyme which catalyzes the interconversion of the excitatory transmitter candidates, glutamate and aspartate, has been measured in fiber tracts of rat, with an emphasis on sensory and motor systems of the brain. Most tracts had significantly higher activities than the cholinergic facial nerve root, consistent with the possibility that a component of aspartate aminotransferase activity might serve as a marker for neurons using glutamate and/or aspartate as neurotransmitter. Highest activity was in the auditory nerve root. On the other hand, a close correlation was found between aspartate aminotransferase and malate dehydrogenase activities in the fiber tracts, raising the question whether aspartate aminotransferase activity may be more closely related to energy metabolism than to transmitter metabolism.  相似文献   

20.
The gene coding for aspartate aminotransferase (EC 2.6.1.1) has been cloned from the extreme thermoacidophilic archaebacterium Sulfolobus solfataricus strain MT4. Partial sequence data obtained directly from the purified protein and from the two cyanogen-bromide-generated peptides confirm the primary structure of aspartate aminotransferase inferred from the nucleotide sequence of its gene. A comparison of the enzyme with other aminotransferases revealed an interesting similarity with tyrosine aminotransferase from rat liver (EC 2.6.1.5) and allowed some tentative assignments of the residues implied in the catalysis. The aspartate aminotransferase gene-flanking regions were compared to those of other archaebacterial genes already described in the literature with the aim of identifying potential regulatory sites.  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