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 46 毫秒
1.
Regulation of valine catabolism in Pseudomonas putida   总被引:12,自引:10,他引:2       下载免费PDF全文
The activities of six enzymes which take part in the oxidation of valine by Pseudomonas putida were measured under various conditions of growth. The formation of four of the six enzymes was induced by growth on d- or l-valine: d-amino acid dehydrogenase, branched-chain keto acid dehydrogenase, 3-hydroxyisobutyrate dehydrogenase, and methylmalonate semialdehyde dehydrogenase. Branched-chain amino acid transaminase and isobutyryl-CoA dehydrogenase were synthesized constitutively. d-Amino acid dehydrogenase and branched-chain keto acid dehydrogenase were induced during growth on valine, leucine, and isoleucine, and these enzymes were assumed to be common to the metabolism of all three branched-chain amino acids. The segment of the pathway required for oxidation of isobutyrate was induced by growth on isobutyrate or 3-hydroxyisobutyrate without formation of the preceding enzymes. d-Amino acid dehydrogenase was induced by growth on l-alanine without formation of other enzymes required for the catabolism of valine. d-Valine was a more effective inducer of d-amino acid dehydrogenase than was l-valine. Therefore, the valine catabolic pathway was induced in three separate segments: (i) d-amino acid dehydrogenase, (ii) branched-chain keto acid dehydrogenase, and (iii) 3-hydroxyisobutyrate dehydrogenase plus methylmalonate semialdehyde dehydrogenase. In a study of the kinetics of formation of the inducible enzymes, it was found that 3-hydroxyisobutyrate and methylmalonate semialdehyde dehydrogenases were coordinately induced. Induction of enzymes of the valine catabolic pathway was studied in a mutant that had lost the ability to grow on all three branched-chain amino acids. Strain PpM2106 had lowered levels of branched-chain amino acid transaminase and completely lacked branched-chain keto acid dehydrogenase when grown in medium which contained valine. Addition of 2-ketoisovalerate, 2-ketoisocaproate, or 2-keto-3-methylvalerate to the growth medium of strain PpM2106 resulted in induction of normal levels of branched-chain keto acid dehydrogenase; therefore, the branched-chain keto acids were the actual inducers of branched-chain keto acid dehydrogenase.  相似文献   

2.
The Escherichia coli aroE gene encoding shikimate dehydrogenase was sequenced. The deduced amino acid sequence was confirmed by N-terminal amino acid sequencing and amino acid analysis of the overproduced protein. The complete polypeptide chain has 272 amino acid residues and has a calculated Mr of 29,380. E. coli shikimate dehydrogenase is homologous to the shikimate dehydrogenase domain of the fungal arom multifunctional enzymes and to the catabolic quinate dehydrogenase of Neurospora crassa.  相似文献   

3.
Two experiments were performed to examine the effects of intramuscular estradiol administration on the hepatic specific activities of some enzymes of lipid, carbohydrate and amino acid metabolism in the immature fowl. Estradiol increased the specific activities of the hepatic lipogenic enzymes, ATP citrate lyase and malate dehydrogenase (decarboxylating) (NADP), but had no effects on the activities of the glycolytic, gluconeogenic and amino acid metabolising enzymes except for pyruvate kinase and glutamate dehydrogenase which were reduced in activity in both experiments. The results indicate that the estrogen-induced increase in hepatic lipid biosynthesis is due to a specific effect on lipid metabolism and not to a general increase in liver metabolism.  相似文献   

4.
D-Mannitol-1-phosphate dehydrogenase (EC 1.1.1.17) and D-glucitol-6-phosphate dehydrogenase (EC 1.1.1.140) were purified to apparent homogeneity in good yields from Escherichia coli. The amino acid compositions, N-terminal amino acid sequences, sensitivities to chemical reagents, and catalytic properties of the two enzymes were determined. Both enzymes showed absolute specificities for their substrates. The subunit molecular weights of mannitol-1-phosphate and glucitol-6-phosphate dehydrogenases were 40,000 and 26,000, respectively; the apparent molecular weights of the native proteins, determined by gel filtration, were 40,000 and 117,000, respectively. It is therefore concluded that whereas mannitol-1-phosphate dehydrogenase is a monomer, glucitol-6-phosphate dehydrogenase is probably a tetramer. These two proteins differed in several fundamental respects.  相似文献   

5.
In this study, we report cDNA sequences of the cytosolic NADP-dependent isocitrate dehydrogenase for humans, mice, and two species of voles (Microtus mexicanus and Microtus ochrogaster). Inferred amino acid sequences from these taxa display a high level of amino acid sequence conservation, comparable to that of myosin beta heavy chain, and share known structural features. A Caenorhabditis elegans enzyme that was previously identified as a protein similar to isocitrate dehydrogenase is most likely the NADP-dependent cytosolic isocitrate dehydrogenase enzyme equivalent, based on amino acid similarity to mammalian enzymes and phylogenetic analysis. We also suggest that NADP-dependent isocitrate dehydrogenases characterized from alfalfa, soybean, and eucalyptus are most likely cytosolic enzymes. The phylogenetic tree of various isocitrate dehydrogenases from eukaryotic sources revealed that independent gene duplications may have given rise to the cytosolic and mitochondrial forms of NADP-dependent isocitrate dehydrogenase in animals and fungi. There appears to be no statistical support for a hypothesis that the mitochondrial and cytosolic forms of the enzyme are orthologous in these groups. A possible scenario of the evolution of NADP-dependent isocitrate dehydrogenases is proposed.   相似文献   

6.
Malate dehydrogenase: a model for structure, evolution, and catalysis.   总被引:11,自引:0,他引:11       下载免费PDF全文
Malate dehydrogenases are widely distributed and alignment of the amino acid sequences show that the enzyme has diverged into 2 main phylogenetic groups. Multiple amino acid sequence alignments of malate dehydrogenases also show that there is a low degree of primary structural similarity, apart from in several positions crucial for nucleotide binding, catalysis, and the subunit interface. The 3-dimensional structures of several malate dehydrogenases are similar, despite their low amino acid sequence identity. The coenzyme specificity of malate dehydrogenase may be modulated by substitution of a single residue, as can the substrate specificity. The mechanism of catalysis of malate dehydrogenase is similar to that of lactate dehydrogenase, an enzyme with which it shares a similar 3-dimensional structure. Substitution of a single amino acid residue of a lactate dehydrogenase changes the enzyme specificity to that of a malate dehydrogenase, but a similar substitution in a malate dehydrogenase resulted in relaxation of the high degree of specificity for oxaloacetate. Knowledge of the 3-dimensional structures of malate and lactate dehydrogenases allows the redesign of enzymes by rational rather than random mutation and may have important commercial implications.  相似文献   

7.
8.
Formaldehyde dehydrogenase (EC 1.2.1.1) is a widely occurring enzyme which catalyzes the oxidation of S-hydroxymethylglutathione, formed from formaldehyde and glutathione, into S-formyglutathione in the presence of NAD. We determined the amino acid sequences for 5 tryptic peptides (containing altogether 57 amino acids) from electrophoretically homogeneous rat liver formaldehyde dehydrogenase and found that they all were exactly homologous to the sequence of rat liver class III alcohol dehydrogenase (ADH-2). Formaldehyde dehydrogenase was found to be able at high pH values to catalyze the NAD-dependent oxidation of long-chain aliphatic alcohols like n-octanol and 12-hydroxydodecanoate but ethanol was used only at very high substrate concentrations and pyrazole was not inhibitory. The amino acid sequence homology and identical structural and kinetic properties indicate that formaldehyde dehydrogenase and the mammalian class III alcohol dehydrogenases are identical enzymes.  相似文献   

9.
We have purified homoserine dehydrogenase to homogeneity and subjected polypeptide fragments derived from digests of the protein to amino acid sequencing. The amino acid sequence of homoserine dehydrogenase from carrot (Daucus carota) indicates that in carrot both aspartokinase and homoserine dehydrogenase activities reside on the same protein. Additional evidence that aspartokinase and homoserine dehydrogenase reside on a bifunctional protein is provided by coelution of activities during purification steps and by enzyme-specific gel staining techniques. Highly purified fractions containing aspartokinase activity were stained for aspartokinase activity, homoserine dehydrogenase activity, and protein. These gels confirmed that aspartokinase activity and homoserine dehydrogenase activity were present on the same protein. This arrangement of aspartokinase and homoserine dehydrogenase activities residing on the same protein is also found in Escherichia coli, which has two bifunctional enzymes, aspartokinase I-homoserine dehydrogenase I and aspartokinase II-homoserine dehydrogenase II. The amino acid sequence of the major form of homoserine dehydrogenase from carrot cell suspension cultures most closely resembles that of the E. coli ThrA gene product aspartokinase I-homoserine dehydrogenase I.  相似文献   

10.
The biochemical properties of the enzyme alcohol dehydrogenase of two different Drosophila species, Drosophila simulans and Drosophila virilis, were studied and compared with those of Drosophila melanogaster Adhs enzyme. All of them consist of two identical subunits of molecular weight 27800 and share significant similarities in function. The substrate specificities of these enzymes were characterized and Km(app.) and Vmax.(app.) values were calculated. All these alcohol dehydrogenases show greater affinity for secondary rather than for primary alcohols. The amino acid compositions of the three enzymes were determined, and there is a close similarity between the D. simulans and the D. melanogaster enzymes, but there are significant differences from the alcohol dehydrogenase of D. virilis. The N-terminal amino acid is blocked and the C-terminal amino acid is the same for all three alcohol dehydrogenases. The enzymes from the three species were carboxymethylated and digested with trypsin. The peptide 'maps' reveal, as expected, more homologies between the enzymes of D. simulans and D. melanogaster than with the enzyme of D. virilis.  相似文献   

11.
Exposure of L929 murine fibroblasts to ozone resulted in K+ leakage and inhibition of several enzymes. Most sensitive to ozone exposure were glyceraldehyde-3-phosphate dehydrogenase and pyruvate kinase. The activities of another cytosolic enzyme, lactate dehydrogenase, the mitochondrial enzymes glutamate dehydrogenase, succinate dehydrogenase, cytochrome c oxidase and the activity of the lysosomal enzymes acid phosphatase and beta-glucuronidase were, initially, not or only slightly affected. The localization of the lysosomal enzymes did not change during ozone exposure. After prolonged exposure complete deterioration of the cells was observed and all enzyme activities declined. The activity of the enzymes was also monitored during ozone exposure of a sonicated cell suspension and it was shown that all these enzymes are in fact susceptible to ozone. These observations clearly demonstrate that, besides the structure and amino acid composition of an enzyme, the localization in the cell plays an important role in its susceptibility to ozone. The intracellular levels of reduced and oxidized glutathione were affected as well. The ATP content, however, proved to be insensitive to ozone exposure.  相似文献   

12.
The nucleotide sequence of a 2493 base pair (bp) region, spanning the coding regions for the meta-cleavage pathway enzymes 2-hydroxymuconic semialdehyde dehydrogenase (HMSD) and 2-hydroxymuconic semialdehyde hydrolase (HMSH), was determined. The deduced protein sequence for HMSD is 486 amino acid residues long with an Mr of 51,682. HMSD has homology with a number of aldehyde dehydrogenases from various eukaryotic sources. The deduced protein sequence for HMSH is 283 amino acids long with an Mr of 30,965. The amino acid composition of this enzyme is similar to that of isofunctional enzymes from toluene and m-cresol catabolic pathways.  相似文献   

13.
The polypeptide chain of the allosteric L-lactate dehydrogenase (EC 1.1.1.27) of Lactobacillus casei consists of 325 amino acid residues. Despite the strikingly different enzymatic characteristics of the allosteric L-lactate dehydrogenase of L. casei and of the non-allosteric vertebrate enzymes, the sequence of the allosteric enzyme shows a distinct homology with that of the non-allosteric vertebrate enzymes (average identity: 37%). An especially high sequence homology can be identified within the active center (average identity: 70%). A clear deviation of the L. casei enzyme from the vertebrate enzyme is the lack of the first 12 amino acid residues at the N terminus and an additional 7 amino acid residues at the C terminus. The localization of the binding site of the allosteric effector D-fructose 1,6-bisphosphate and pH and effector-induced changes of the spectroscopic properties are discussed on the basis of the primary structure.  相似文献   

14.
We constructed mutant genes of Caldococcus noboribetus isocitrate dehydrogenase containing ancestral amino acid residues that were inferred using the maximal likelihood method and a composite phylogenetic tree of isocitrate dehydrogenase and 3-isopropylmalate dehydrogenase. The mutant genes were expressed in Escherichia coli and the protein products purified. Thermostabilities, reported as the half-inactivation temperatures, for the purified enzymes were determined and compared with that of the wild-type enzyme. Four of the five mutant enzymes have greater thermal stabilities than wild-type isocitrate dehydrogenase. The results are compatible with the hyperthermophilic universal ancestor (commonote) hypothesis. Incorporation of ancestral residues into a modern-day protein sequence can be used to improve protein thermostability.  相似文献   

15.
The effect of 24-hr starvation on the amino acid pool composition and its concentration ratios with respect to blood and plasma as well as the activities of alanine, aspartate and branched chain amino acid transaminases, glutamate dehydrogenase, glutamine synthetase and adenylate deaminase have been studied in rat brown adipose tissue. Starvation induced a considerable decrease of pool amino acid concentration. Alanine and taurine were the amino acids in which the decrease was more marked. Small changes were observed in the activities of the enzymes studied, with decreases only in glutamate dehydrogenase and adenylate deaminase. These changes agree with a decrease in amino acid utilization in this tissue induced by starvation.  相似文献   

16.
D-Methionine was converted to L-methionine in a reaction system where four enzymes were used. D-amino acid oxidase (D-AAO) from Arthrobacter protophormiae was used for the complete conversion of D-methionine to 2-oxo-4-methylthiobutyric acid. Catalase was added to prevent 2-oxo-4-methylthiobutyric acid decarboxylation. In the second reaction step, L-phenylalanine dehydrogenase (L-PheDH) from Rhodococcus sp. was used to convert 2- oxo-4-methylthiobutyric acid to L-methionine, and formate dehydrogenase (FDH) from Candida boidinii was added for NADH regeneration. Enzyme kinetics of all enzymes was analyzed in detail. Mathematical models for separate reactions steps, as well as for the complete system were developed and validated in the batch reactor experiments. Complete conversion of D-methionine to L-methionine was achieved. Considering that both enzymes act on different substrates, such a system could be easily employed for the synthesis of other amino acids from D-isomer, as well as from the racemate of a certain amino acid (DL-amino acid).  相似文献   

17.
The molecular basis for the specificity of aldehyde dehydrogenases (ALDHs) for retinal, the precursor of the morphogen retinoic acid, is still poorly understood. We have expressed in Escherichia coli both retinal dehydrogenase (RALDH), a cytosolic aldehyde dehydrogenase originally isolated from rat kidney, and the highly homologous phenobarbital-induced aldehyde dehydrogenase (PB-ALDH). Oxidation of propanal was observed with both enzymes. On the other hand, recombinant RALDH efficiently catalyzed oxidation of 9-cis- and all-trans-retinal, whereas PB-ALDH was inactive with all-trans-retinal and poorly active with 9-cis-retinal. A striking difference between PB-ALDH and all other class I ALDHs is the identity of the amino acid immediately preceding the active nucleophile Cys(302) (Ile(301) instead of Cys(301)). Nevertheless, these amino acids could be exchanged in either RALDH or PB-ALDH without affecting substrate specificity. Characterization of chimeric enzymes demonstrates that distinct groups of amino acids control the differential activity of RALDH and PB-ALDH with all-trans- and 9-cis-retinal. Of 52 divergent amino acids, the first 17 are crucial for activity with all-trans-retinal, whereas the next 25 are important for catalysis of 9-cis-retinal oxidation. Recombinant enzymes with specificity for all-trans- or 9-cis-retinal were obtained, which should provide useful tools to study the relative importance of local production of all-trans- versus 9-cis-retinoic acid in development and tissue differentiation.  相似文献   

18.
The amino acid sequence of D-beta-hydroxybutyrate dehydrogenase (BDH), a phosphatidyl-choline-dependent enzyme, has been determined for the enzyme from rat liver by a combination of nucleotide sequencing of cDNA clones and amino acid sequencing of the purified protein. This represents the first report of the primary structure of this enzyme. The largest clone contained 1435 base pairs and encoded the entire amino acid sequence of mature BDH and the leader peptide of precursor BDH. Hybridization of poly(A+) rat liver mRNA revealed two bands with estimated sizes of 3.2 and 1.7 kb. A computer-based comparison of the amino acid sequence of BDH with other reported sequences reveals a homology with the superfamily of short-chain alcohol dehydrogenases, which are distinct from the classical zinc-dependent alcohol dehydrogenases. This protein family, initially discerned from Drosophila alcohol dehydrogenase and bacterial ribitol dehydrogenase, is now known to include at least 20 enzymes catalyzing oxidations of distinct substrates.  相似文献   

19.
Changes in oxygen and/or glucose availability may result in altered levels of ATP production and amino acid levels, and alteration in lactic acid production. However, under certain metabolic insults, the retina demonstrates considerable resilience and maintains ATP production, and/or retinal function. We wanted to investigate whether this resilience would be reflected in alterations in the activity of key enzymes of retinal metabolism, or enzymes associated with amino acid production that may supply their carbon skeleton for energy production. Enzymatic assays were conducted to determine the activity of key retinal metabolic enzymes total ATPase and Na(+)/K(+)-ATPase, aspartate aminotransferase and lactate dehydrogenase. In vitro anoxia led to an increase in retinal lactate dehydrogenase activity and to a decrease in retinal aspartate aminotransferase activity, without significant changes in Na(+)/K(+)-ATPase activity. In vivo inhibition of glutamine synthetase resulted in a short-term significant decrease in retinal aspartate aminotransferase activity. An increase in retinal aspartate aminotransferase and lactate dehydrogenase activities was accompanied by altered levels of amino acids in neurons and glia after partial inhibition of glial metabolism, implying that short- and long-term up- and down-regulation of key metabolic enzymes occurs to supply carbon skeletons for retinal metabolism. ATPase activity does not appear to fluctuate under the metabolic stresses employed in our experimental procedures.  相似文献   

20.
Alanopine dehydrogenase (AlDH) and three isoforms of strombine/alanopine dehydrogenase (St/AlDH) were purified from muscle tissue of the polychaete rockworm Marphysa sanguinea. The four enzymes, which can be distinguished by the isoelectric point, are monomeric 42 kDa proteins, possess similar pH-activity profiles, and display specificity for pyruvate and NAD(H). The three isoforms of St/AlDH show equivalent Km and Vmax for glycine and L-alanine and for D-strombine and meso-alanopine. Free amino acid levels in the muscle and D-strombine accumulation in vivo during muscle activity suggest that St/AlDHs function physiologically as StDH. AlDH shows specificity for L-alanine and meso-alanopine, but not for glycine or D-strombine. The amino acid sequences of AlDH and one of the St/AlDH isoforms were determined by a combination of amino acid sequence analysis and cDNA cloning. St/AlDH cDNA consisted of 1586 bp nucleotides that encode a 399-residue protein (43,346.70 Da), and AlDH cDNA consisted of 1587 bp nucleotides that encode a 399-residue protein (43,886.68 Da). The two amino acid sequences deduced from the cDNA displayed 67% amino acid identity, with greatest similarity to that of tauropine dehydrogenase from the polychaete Arabella iricolor.  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