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.
Aspartate kinase (AK) is the first and committed enzyme of the biosynthetic pathway producing aspartate family amino acids, lysine, threonine, and methionine. AK from Corynebacterium glutamicum (CgAK), a bacterium used for industrial fermentation of amino acids, including glutamate and lysine, is inhibited by lysine and threonine in a concerted manner. To elucidate the mechanism of this unique regulation in CgAK, we determined the crystal structures in several forms: an inhibitory form complexed with both lysine and threonine, an active form complexed with only threonine, and a feedback inhibition-resistant mutant (S301F) complexed with both lysine and threonine. CgAK has a characteristic α2β2-type heterotetrameric structure made up of two α subunits and two β subunits. Comparison of the crystal structures between inhibitory and active forms revealed that binding inhibitors causes a conformational change to a closed inhibitory form, and the interaction between the catalytic domain in the α subunit and β subunit (regulatory subunit) is a key event for stabilizing the inhibitory form. This study shows not only the first crystal structures of α2β2-type AK but also the mechanism of concerted inhibition in CgAK.  相似文献   

2.
Samples of midseason Florida citrus molasses from 10 different manufacturing plants were studied during 2 years of production for suitability as fermentation substrates. They were analyzed for degrees Brix, pH, percentage of protein, and free amino acid content. The following ranges of values were obtained: degrees Brix, 65.4 to 73.0; pH, 4.2 to 6.4; and protein, 3.3 to 5.1%. Although the general characteristics of all samples were quite similar, the total free amino acid content and the number of amino acids differed from year to year and from plant to plant. The free amino acids consistently present were asparagine, aspartic acid, alanine, γ-amino-butyric, glycine, glutamic acid, proline, serine, and threonine. Occasionally found were arginine, histidine, isoleucine, leucine, lysine, phenylalanine, tyrosine, and valine. An extremely high level of proline was detected in all samples.  相似文献   

3.
Kinraide TB 《Plant physiology》1981,68(6):1327-1333
Data from published experiments were analyzed to determine the number and specificities of amino acid transport channels in cells of higher plants. Each experiment measured the uptake of a labeled amino acid in the presence of unlabeled amino acids, used one at a time, in the incubating medium. The observed interamino acid inhibitions can be accounted for by two transport channels, each with characteristic affinities that were computed from the observed interamino acid inhibitions. The first channel is a general transport system with the following relative affinities for the amino acids: methionine 75, alanine 75, phenylalanine 64, tyrosine 64, leucine 63, cysteine 58, serine 57, glycine 56, tryptophan 54, glutamine 51, threonine 49, valine 44, isoleucine 44, glutamic acid 44, proline 43, histidine 33, lysine 32, asparagine 22, arginine 22, aspartic acid 18. The second channel is a basic amino acid tranport system with relative affinities for arginine, lysine, and histidine of 66, 39, and 21, respectively. The affinities for the other acids in the second channel are lower. Despite considerable diversity in the species, tissues, and solute concentrations employed in the experiments, multiple regression equations (Y = α + β1X1 + βX2, in which Y is the observed transport inhibition and X1 and X2 are the relative transport affinities of the two channels) account for 50 to 99% of the variance in all but six experiments, five of which employed unusually high solute concentrations.  相似文献   

4.
Vacuoles of internodal cells of Chara australis (or Chara corallina) were loaded with a 10 millimolar amount of various amino acids by a perfusion method and incubated under continuous light. After 20 to 24 hours, the cell sap was collected, and free amino acids in it and the rest of the cell (cytoplasm) were analyzed. The only amino acid metabolized completely was alanine. About 40 to 80% of the aspartic acid, glutamine, serine, and glycine were metabolized, whereas less than 30% of the threonine, asparagine, isoasparagine, isoleucine, phenylalanine, γ-aminobutyric acid, lysine, and arginine were metabolized. The figure for glutamic acid fluctuated between 10 and 100%. The main metabolites of alanine were glutamine, glycine and ammonia, which accumulated in the vacuole. Alanine utilization was not affected by l-methionine-d,l-sulfoximine or azaserine, but was strongly inhibited by aminooxyacetate. The cell extract contained enough alanine aminotransferase activity to account for the rate of alanine metabolism.  相似文献   

5.
α-Amylase has been purified from de-embryonated seeds of barley (Hordeum vulgare L. cv. Betzes) which have been incubated on 10−6 m gibberellic acid (GA3) following 3 days of imbibition in buffer. Incubation of the half-seeds in up to 10−2 m 5-fluorouracil (5-FU) during the entire incubation period, including imbibition, had no effect on any of the following characteristics of purified α-amylase: thermal stability in the absence of calcium, molecular weight of the enzyme, isozyme composition, specific activity, or the amount of α-amylase synthesized by the aleurone tissue. The synthesis of rRNA and tRNA was strongly inhibited by 5-FU, indicating that the analog had entered the aleurone cells. These results are not in agreement with those of Carlson (Nature New Biology 237: 39-41 [1972]) who found that treatment of barley aleurone with 10−4 m 5-FU prior to the addition of GA3 resulted in decreased thermal stability of GA3-induced α-amylase and who interpreted this as evidence that the mRNA for α-amylase was synthesized during the imbibition of the aleurone tissue and independently of gibberellin action. Results of the present experiments indicate that the thermal stability of highly purified α-amylase is not altered by treatment of barley half-seeds with 5-FU, and that 5-FU cannot be used as a probe to examine the timing of α-amylase mRNA synthesis.  相似文献   

6.
The host-specific toxin of Helminthosporium carbonum (C32H50N6O10) was hydrolyzed by 6 n HCl to yield a number of α-amino acids. The common amino acids, proline and alanine, occurred in a ratio of 1:2. Two other unstable α-amino acids that produced lower color values with ninhydrin were also produced. One of these was tentatively identified as 2-amino-2,3-dehydro-3-methylpentanoic acid by electrolytic reduction to isoleucine. Additional ninhydrin-reacting substances were produced in low yield and probably represented secondary hydrolysis products of the unstable amino acids. The finding of an α,β-unsaturated linkage in H. carbonum toxin explains the instability of the compound and may also account for its specific toxicity.  相似文献   

7.
Methionine γ-lyase (MGL) catalyzes the γ-elimination of l-methionine and its derivatives as well as the β-elimination of l-cysteine and its analogs. These reactions yield α-keto acids and thiols. The mechanism of chemical conversion of amino acids includes numerous reaction intermediates. The detailed analysis of MGL interaction with glycine, l-alanine, l-norvaline, and l-cycloserine was performed by pre-steady-state stopped-flow kinetics. The structure of side chains of the amino acids is important both for their binding with enzyme and for the stability of the external aldimine and ketimine intermediates. X-ray structure of the MGL·l-cycloserine complex has been solved at 1.6 Å resolution. The structure models the ketimine intermediate of physiological reaction. The results elucidate the mechanisms of the intermediate interconversion at the stages of external aldimine and ketimine formation.  相似文献   

8.
Properties of an Aminotransferase of Pea (Pisum sativum L.)   总被引:2,自引:2,他引:0  
A transaminase (aminotransferase, EC 2.6.1) fraction was partially purified from shoot tips of pea (Pisum sativum L. cv. Alaska) seedlings. With α-ketoglutarate as co-substrate, the enzyme transaminated the following aromatic amino acids: d,l-tryptophan, d,l-tyrosine, and d,l-phenylalanine, as well as the following aliphatic amino acids: d,l-alanine, d,l-methionine, and d,l-leucine. Of other α-keto acids tested, pyruvate and oxalacetate were more active than α-ketoglutarate with d,l-tryptophan. Stoichiometric yields of indolepyruvate and glutamate were obtained with d,l-tryptophan and α-ketoglutarate as co-substrates. The specific activity was three times higher with d-tryptophan than with l-tryptophan.  相似文献   

9.
Clostridium thermosulfurogenes EM1 formed blebs, i.e., protrusions still in contact with the cytoplasmic membrane, that originated from the cytoplasmic membrane during growth in batch culture and continuous culture. They could be observed squeezed between the cell wall and cytoplasmic membrane in cells with seemingly intact wall layers (surface layer and peptidoglycan layer) as well as in cells with wall layers in different states of degradation caused by phosphate limitation or high dilution rates. Blebs were found to turn into membrane vesicles by constriction in cases when the cell wall was heavily degraded. Bleb and vesicle formation was also observed in the absence of substrates that induce α-amylase and pullulanase synthesis. No correlations existed between bleb formation and the presence of active enzyme. Similar blebs could also be observed in a number of other gram-positive bacteria not producing these enzymes, but they were not observed in gram-negative bacteria. For immunoelectron-microscopic localization of α-amylase and pullulanase in C. thermosulfurogenes EM1, two different antisera were applied. One was raised against the enzymes isolated from the culture fluid; the other was produced against a peptide synthesized, as a defined epitope, in analogy to the N-terminal amino acid sequence (21 amino acids) of the native extracellular α-amylase. By using these antisera, α-amylase and pullulanase were localized at the cell periphery in samples taken from continuous culture or batch culture. In samples prepared for electron microscopy by freeze substitution followed by ultrathin sectioning, blebs could be seen, and the immunolabel pinpointing α-amylase enzyme particles was seen not only randomly distributed in the cell periphery, but also lining the surface of the cytoplasmic membrane and the blebs. Cells exhibiting high or virtually no enzyme activity were labeled similarly with both antisera. This finding strongly suggests that α-amylase and pullulanase may occur in both active and inactive forms, depending on growth conditions.  相似文献   

10.
Promotion of seed germination by cyanide   总被引:2,自引:2,他引:0  
Potassium cyanide at 3 μm to 10 mm promotes germination of Amaranthus albus, Lactuca sativa, and Lepidium virginicum seeds. l-Cysteine hydrogen sulfide lyase, which catalyzes the reaction of HCN with l-cysteine to form β-l cyanoalanine, is active in the seeds. β-l-Cyanoalanine is the most effective of the 23 α-amino acids tested for promoting germination of A. albus seeds. Aspartate, which is produced by enzymatic hydrolysis of asparagine formed by hydrolysis from β-cyanoalanine, is the second most effective of the 23 amino acids. Uptake of aspartate-4-14C is much lower than of cyanide.  相似文献   

11.
Amino acids emitted and extracted from surface-sterilized larvae and adults of Heterodera glycines were identified by paper chromatography and quantitatively analyzed by column chromatography. Five amino acids (alanine, aspartic acid, glutamic acid, glycine and serine) were emitted by H. glycines larvae and eight others (asparagine, glutamine, leucine/isoleucine, lysine, methionine sulfoxide, threonine, tyrosine, valine/methionine) were found in extracts from crushed larvae.In addition to the amino acids emitted or extracted from larvae, four others were emitted by adults (γ-aminobutyric acid, histidine, phenylalanine, and proline). Four different amino acids (arginine, cystathionine, hydroxyproline, and ornithine) were found only in the extract from crushed adults. Greater quantities of alanine, aspartic acid and glycine were emitted than could be detected in nematode extracts suggesting selective emission.Subsamples of nematode populations were taken from growing plants 19, 26, 33, and 40 days after inoculation and extracted to determine whether changes in specific amino acid content correlated with aging. Proline content shifted most, increasing from 4.1% to 21.5% of the total amino acid complement from the 19th to the 40th days.  相似文献   

12.
Polyclonal antibodies raised against barley (1→3,1→4)-β-d-glucanase, α-amylase and carboxypeptidase were used to detect precursor polypeptides of these hydrolytic enzymes among the in vitro translation products of mRNA isolated from the scutellum and aleurone of germinating barley. In the scutellum, mRNA encoding carboxypeptidase appeared to be relatively more abundant than that encoding α-amylase or (1→3,1→4)-β-d-glucanase, while in the aleurone α-amylase and (1→3,1→4)-β-d-glucanase mRNAs predominated. The apparent molecular weights of the precursors for (1→3,1→4)-β-d-glucanase, α-amylase, and carboxypeptidase were 33,000, 44,000, and 35,000, respectively. In each case these are slightly higher (1,500-5,000) than molecular weights of the mature enzymes. Molecular weights of precursors immunoprecipitated from aleurone and scutellum mRNA translation products were identical for each enzyme.  相似文献   

13.
To define the upstream and downstream signaling specificities of metabotropic glutamate receptors (mGluR), we have examined the ability of representative mGluR of group I, II, and III to be activated by endogenous amino acids and catalyze activation of G proteins coupled to phospholipase C (PLC), or activation of Gi/o proteins coupled to the ion channel TRPC4β. Fluorescence-based assays have allowed us to observe interactions not previously reported or clearly identified. We have found that the specificity for endogenous amino acids is remarkably stringent. Even at millimolar levels, structurally similar compounds do not elicit significant activation. As reported previously, the clear exception is l-serine-O-phosphate (l-SOP), which strongly activates group III mGluR, especially mGluR4,-6,-8 but not group I or II mGluR. Whereas l-SOP cannot activate mGluR1 or mGluR2, it acts as a weak antagonist for mGluR1 and a potent antagonist for mGluR2, suggesting that co-recognition of l-glutamate and l-SOP arose early in evolution, and was followed later by divergence of group I and group II mGluR versus group III in l-SOP responses. mGluR7 has low affinity and efficacy for activation by both l-glutamate and l-SOP. Molecular docking studies suggested that residue 74 corresponding to lysine in mGluR4 and asparagine in mGluR7 might play a key role, and, indeed, mutagenesis experiments demonstrated that mutating this residue to lysine in mGluR7 enhances the potency of l-SOP. Experiments with pertussis toxin and dominant-negative Gαi/o proteins revealed that mGluR1 couples strongly to TRPC4β through Gαi/o, in addition to coupling to PLC through Gαq/11.  相似文献   

14.
1. The amino acid requirements of freshly explanted chick embryonic heart tissues cultivated in completely synthetic media have been determined, employing a nutritional depletion technique. Arginine, histidine, lysine, tyrosine, tryptophan, phenylalanine, cystine, methionine, threonine, leucine, and valine were found to be essential. Serine, isoleucine, glycine, and glutamine were found to be non-essential. Glutamic acid, aspartic acid, α-alanine, proline, and hydroxyproline were found to be inhibitory in this test system. 2. A total amino acid level of approximately 100 mg. per cent was found to be optimal and DL-amino acids were found to be non-toxic, unless used in high concentrations. 3. A comparison has been made of the amino acid requirements of various types of tissue cultures, of the chick, and of man and certain differences in these requirements have been discussed.  相似文献   

15.
16.
Dwarf maize (Zea mays L.), a mutant deficient in gibberellin synthesis, provides an excellent model to study the influence of gibberellin on biochemical processes related to plant development. Alterations in the chemical structure of the cell wall mediated by gibberellin were examined in seedlings of this mutant. The composition of the walls of roots, mesocotyl, coleoptile, and primary leaves of dwarf maize was similar to that of normal maize and other cereal grasses. Glucuronoarabinoxylans constituted the principal hemicelluloses, but walls also contained substantial amounts of xyloglucan and mixed-linkage β-d-glucan. Root growth in dwarf maize was essentially normal, but growth of mesocotyl and primary leaves was severely retarded. Injection of the gibberellin into the cavity of the coleoptile resulted in a marked increase in elongation of the primary leaves. This elongation was accompanied by increases in total wall mass, but the proportion of β-d-glucan decreased from 20% to 15% of the hemicellulosic polysaccharide. During leaf expansion, the proportion decreased further to only 10%. Through 4 days incubation, the proportion of β-d-glucan in leaves of control seedlings without gibberellin was nearly constant. Extraction of exo- and endo-β-d-glucan hydrolases from purified cell walls and assay against a purified oat bran β-d-glucan demonstrated that gibberellin increased the activity of the endo-β-d-glucan hydrolase. These and other data support the hypothesis that β-d-glucan metabolism is central to control of cell expansion in cereal grasses.  相似文献   

17.
The peptidyl nucleoside arginomycin is active against Gram-positive bacteria and fungi but displays much lower toxicity to mice than its analog blasticidin S. It features a rare amino acid, β-methylarginine, which is attached to the deoxyhexose moiety via a 4′-aminoacyl bond. We here report cloning of the complete biosynthetic gene cluster for arginomycin from Streptomyces arginensis NRRL 15941. Among the 14 putative essential open reading frames, argM, encoding an aspartate aminotransferase (AAT), and adjacent argN, encoding an S-adenosyl methionine (SAM)-dependent methyltransferase, are coupled to catalyze arginine and yield β-methylarginine in Escherichia coli. Purified ArgM can transfer the α-amino group of l-arginine to α-ketoglutaric acid to give glutamate and thereby converts l-arginine to 5-guanidino-2-oxopentanoic acid, which is methylated at the C-3 position by ArgN to form 5-guanidino-3-methyl-2-oxopentanoic acid. Iteratively, ArgM specifically catalyzes transamination from the donor l-aspartate to the resulting 5-guanidino-3-methyl-2-oxopentanoic acid, generating β-methylarginine. The complete and concise biosynthetic pathway for the rare and bioactive amino acid revealed by this study may pave the way for the production of β-methylarginine either by enzymatic conversion or by engineered living cells.  相似文献   

18.
Transamination is the first step in the conversion of amino acids into aroma compounds by lactic acid bacteria (LAB) used in food fermentations. The process is limited by the availability of α-ketoglutarate, which is the best α-keto donor for transaminases in LAB. Here, uptake of α-ketoglutarate by the citrate transporter CitP is reported. Cells of Lactococcus lactis IL1403 expressing CitP showed significant levels of transamination activity in the presence of α-ketoglutarate and one of the amino acids Ile, Leu, Val, Phe, or Met, while the same cells lacking CitP showed transamination activity only after permeabilization of the cell membrane. Moreover, the transamination activity of the cells followed the levels of CitP in a controlled expression system. The involvement of CitP in the uptake of the α-keto donor was further demonstrated by the increased consumption rate in the presence of l-lactate, which drives CitP in the fast exchange mode of transport. Transamination is the only active pathway for the conversion of α-ketoglutarate in IL1403; a stoichiometric conversion to glutamate and the corresponding α-keto acid from the amino acids was observed. The transamination activity by both the cells and the cytoplasmic fraction showed a remarkably flat pH profile over the range from pH 5 to pH 8, especially with the branched-chain amino acids. Further metabolism of the produced α-keto acids into α-hydroxy acids and other flavor compounds required the coupling of transamination to glycolysis. The results suggest a much broader role of the citrate transporter CitP in LAB than citrate uptake in the citrate fermentation pathway alone.  相似文献   

19.
Suspension cultured cells of tomato (Lycopersicon esculentum Mill. cv VFNT Cherry) adapted to water stress induced with polyethylene glycol 6000 (PEG), exhibit marked alterations in free amino acid pools (Handa et al. 1983 Plant Physiol 73: 834-843). Using computer simulation models the in vivo rates of synthesis and utilization and compartmentation of free amino acid pools were determined from 15N labeling kinetics after substituting [15N]ammonium and [15N]nitrate for the 14N salts in the culture medium of cell lines adapted to 0% and 25% PEG. The 300-fold elevated proline pool in 25% PEG adapted cells is primarily the consequence of a 10-fold elevated rate of proline synthesis via the glutamate pathway. Ornithine was insufficiently labeled to serve as a major precursor for proline. Our calculations suggest that the rate of proline synthesis only slightly exceeds the rate required to sustain both protein synthesis and proline pool maintenance with growth. Mechanisms must operate to restrict proline oxidation in adapted cells. The kinetics of labeling of proline in 25% PEG adapted cells are consistent with a single, greatly enlarged metabolic pool of proline. The depletion of glutamine in adapted cells appears to be a consequence of a selective depletion of a large, metabolically inactive storage pool present in unadapted cultures. The labeling kinetics of the amino nitrogen groups of glutamine and glutamate are consistent with the operation of the glutamine synthetase-glutamate synthase cycle in both cell lines. However, we could not conclusively discriminate between the exclusive operation of the glutamine synthetase-glutamate synthase cycle and a 10 to 20% contribution of the glutamate dehydrogenase pathway of ammonia assimilation. Adaptation to water stress leads to increased nitrogen flux from glutamate into alanine and γ-aminobutyrate, suggesting increased pyruvate availability and increased rates of glutamate decarboxylation. Both alanine and γ-aminobutyrate are synthesized at rates greatly in excess of those simply required to maintain the free pools with growth, indicating that these amino acids are rapidly turned over. Thus, both synthesis and utilization rates for alanine and γ-aminobutyrate are increased in adapted cells. Adaptation to stress leads to increased rates of synthesis of valine and leucine apparently at the expense of isoleucine. Remarkably low 15N flux via the aspartate family amino acids was observed in these experiments. The rate of synthesis of threonine appeared too low to account for threonine utilization in protein synthesis, pool maintenance, and isoleucine biosynthesis. It is possible that isoleucine may be deriving carbon skeletons from sources other than threonine. Tentative models of the nitrogen flux of these two contrasting cell lines are discussed in relation to carbon metabolism, osmoregulation, and nitrogenous solute compartmentation.  相似文献   

20.
A lysine racemase (lyr) gene was isolated from a soil metagenome by functional complementation for the first time by using Escherichia coli BCRC 51734 cells as the host and d-lysine as the selection agent. The lyr gene consisted of a 1,182-bp nucleotide sequence encoding a protein of 393 amino acids with a molecular mass of about 42.7 kDa. The enzyme exhibited higher specific activity toward lysine in the l-lysine-to-d-lysine direction than in the reverse reaction.Amino acids are the building blocks of proteins and play an important role in the regulation of the metabolism of living organisms. Among two enantiomers of naturally occurring amino acids, l-amino acids are predominant in living organisms, while d-amino acids are found in both free and bound states in various organisms like bacteria (36), yeasts (35), plants (47), insects (11), mammals (17), bivalves (39), and fish (28). The d-amino acids are mostly endogenous and produced by racemization from their counterparts by the action of a racemase. Thus, the amino acid racemases are involved in d-amino acid metabolism (29, 46). Since the discovery of alanine racemase in 1951 (42), several racemases toward amino acids, such as those for glutamate, threonine, serine, aspartate, methionine, proline, arginine, and phenylalanine, have been reported in bacteria, archaea, and eukaryotes, including mammals (1, 2, 15, 30, 31, 44). They are classified into two groups: pyridoxal 5′-phosphate (PLP)-dependent and PLP-independent enzymes (9, 36).Lysine racemase (Lyr, EC 5.1.1.5) was first reported in Proteus vulgaris ATCC 4669 (19) and proposed to be involved in the lysine degradation of bacterial cells (5, 19). Catabolism of lysine occurs via two parallel pathways. In one of the pathways, δ-aminovalerate is the key metabolite, whereas in the other l-lysine is racemized to d-lysine, and l-pipecolate and α-aminoadipate (AMA) are the key metabolites (5). d-Lysine catabolism proceeds through a series of cyclized intermediates which are necessary to regenerate an α-amino acid and comprise the following metabolites (AMA pathway): d-lysine→α-keto-ɛ-amino caproate→Δ1-piperideine-2-carboxylate→pipecolate→Δ1-piperideine-6-carboxylate→α-amino-δ-formylcaproate→α-AMA→α-ketoadipate (6, 7, 12, 27). The final product is converted to α-ketoglutarate via a series of coenzyme A derivatives and subsequently participates as an intermediate in the Krebs cycle. This pathway suggests that the biological function of d-lysine in the bacteria is that of a carbon or nitrogen source. Racemization of added l-lysine to d-lysine by whole cells of Proteus spp. and Escherichia spp. (19) and by the cell extract of Pseudomonas putida ATCC 15070 (5, 20) has been found. However, the enzyme has not been purified to homogeneity, and thus, its molecular and catalytic characteristics, including its gene structure, have not been elucidated. In this study, we explored a metagenomic library constructed from a garden soil to isolate a novel Lyr enzyme. After expression in Escherichia coli, the purified enzyme was characterized in terms of optimal pH and temperature, thermal stability, and racemization activity.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号