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.
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   

2.
l-Hydroxyproline (4-hydroxyproline) mainly exists in collagen, and most bacteria cannot metabolize this hydroxyamino acid. Pseudomonas putida and Pseudomonas aeruginosa convert l-hydroxyproline to α-ketoglutarate via four hypothetical enzymatic steps different from known mammalian pathways, but the molecular background is rather unclear. Here, we identified and characterized for the first time two novel enzymes, d-hydroxyproline dehydrogenase and Δ1-pyrroline-4-hydroxy-2-carboxylate (Pyr4H2C) deaminase, involved in this hypothetical pathway. These genes were clustered together with genes encoding other catalytic enzymes on the bacterial genomes. d-Hydroxyproline dehydrogenases from P. putida and P. aeruginosa were completely different from known bacterial proline dehydrogenases and showed similar high specificity for substrate (d-hydroxyproline) and some artificial electron acceptor(s). On the other hand, the former is a homomeric enzyme only containing FAD as a prosthetic group, whereas the latter is a novel heterododecameric structure consisting of three different subunits (α4β4γ4), and two FADs, FMN, and [2Fe-2S] iron-sulfur cluster were contained in αβγ of the heterotrimeric unit. These results suggested that the l-hydroxyproline pathway clearly evolved convergently in P. putida and P. aeruginosa. Pyr4H2C deaminase is a unique member of the dihydrodipicolinate synthase/N-acetylneuraminate lyase protein family, and its activity was competitively inhibited by pyruvate, a common substrate for other dihydrodipicolinate synthase/N-acetylneuraminate lyase proteins. Furthermore, disruption of Pyr4H2C deaminase genes led to loss of growth on l-hydroxyproline (as well as d-hydroxyproline) but not l- and d-proline, indicating that this pathway is related only to l-hydroxyproline degradation, which is not linked to proline metabolism.  相似文献   

3.
Formation of the peptidoglycan stem pentapeptide requires the insertion of both l and d amino acids by the ATP-dependent ligase enzymes MurC, -D, -E, and -F. The stereochemical control of the third position amino acid in the pentapeptide is crucial to maintain the fidelity of later biosynthetic steps contributing to cell morphology, antibiotic resistance, and pathogenesis. Here we determined the x-ray crystal structure of Staphylococcus aureus MurE UDP-N-acetylmuramoyl-l-alanyl-d-glutamate:meso-2,6-diaminopimelate ligase (MurE) (E.C. 6.3.2.7) at 1.8 Å resolution in the presence of ADP and the reaction product, UDP-MurNAc-l-Ala-γ-d-Glu-l-Lys. This structure provides for the first time a molecular understanding of how this Gram-positive enzyme discriminates between l-lysine and d,l-diaminopimelic acid, the predominant amino acid that replaces l-lysine in Gram-negative peptidoglycan. Despite the presence of a consensus sequence previously implicated in the selection of the third position residue in the stem pentapeptide in S. aureus MurE, the structure shows that only part of this sequence is involved in the selection of l-lysine. Instead, other parts of the protein contribute substrate-selecting residues, resulting in a lysine-binding pocket based on charge characteristics. Despite the absolute specificity for l-lysine, S. aureus MurE binds this substrate relatively poorly. In vivo analysis and metabolomic data reveal that this is compensated for by high cytoplasmic l-lysine concentrations. Therefore, both metabolic and structural constraints maintain the structural integrity of the staphylococcal peptidoglycan. This study provides a novel focus for S. aureus-directed antimicrobials based on dual targeting of essential amino acid biogenesis and its linkage to cell wall assembly.  相似文献   

4.
An α-l-arabinofuranosidase has been purified 1043-fold from radish (Raphanus sativus L.) seeds. The purified enzyme was a homogeneous glycoprotein consisting of a single polypeptide with an apparent molecular weight of 64,000 and an isoelectric point value of 4.7, as evidenced by denaturing gel electrophoresis and reversed-phase or size-exclusion high-performance liquid chromatography and isoelectric focusing. The enzyme characteristically catalyzes the hydrolysis of p-nitrophenyl α-l-arabinofuranoside and p-nitrophenyl β-d-xylopyranoside in a constant ratio (3:1) of the initial velocities at pH 4.5, whereas the corresponding α-l-arabinopyranoside and β-d-xylofuranoside are unsusceptible. The following evidence was provided to support that a single enzyme with one catalytic site was responsible for the specificity: (a) high purity of the enzyme preparation, (b) an invariable ratio of the activities toward the two substrates throughout the purification steps, (c) a parallelism of the activities in activation with bovine serum albumin and in heat inactivation of the enzyme as well as in the inhibition with heavy metal ions and sugars such as Hg2+, Ag+, l-arabino-(1→4)-lactone, and d-xylose, and (d) results of the mixed substrate kinetic analysis using the two substrates. The enzyme was shown to split off α-l-arabinofuranosyl residues in sugar beet arabinan, soybean arabinan-4-galactan, and radish seed and leaf arabinogalactan proteins. Arabinose and xylose were released by the action of the enzyme on oat-spelt xylan. Synergistic action of α-l-arabinofuranosidase and β-d-galactosidase on radish seed arabinogalactan protein resulted in the extensive degradation of the carbohydrate moiety.  相似文献   

5.
1. Human uterine cervical stroma was found to contain a Ca2+-independent neutral proteinase against casein and N-benzoyl-dl-arginine p-nitroanilide (Bz-dl-Arg-Nan). This enzyme was tightly bound to an insoluble material (20000g pellet) and was solubilized by high concentrations of NaCl or KCl. High concentrations of them in the reaction system, however, inhibited reversibly the activity of this enzyme. 2. The neutral proteinase was partially purified by extraction with NaCl, gel filtration on Sephadex G-200 and affinity chromatography on casein–Sepharose. 3. The optimal pH of this partially purified enzyme was 7.4–8.0 against casein and Bz-dl-Arg-Nan. The molecular weight of the enzyme was found to be about 1.4×105 by gel filtration on Sephadex G-200. 4. The enzyme was significantly inhibited by di-isopropyl phosphorofluoridate (0.1mm). High concentration of phenylmethanesulphonyl fluoride (5mm), 7-amino-1-chloro-3-l-tosylamidoheptan-2-one (0.5mm), antipain (10μm) or leupeptin (10μm) was also found to be inhibitory, but chymostatin (40μg/ml), soya-bean trypsin inhibitor (2.5mg/ml), human plasma (10%, v/v), p-chloromercuribenzoate (1mm), EDTA (10mm) and 1-chloro-4-phenyl-3-l-tosylamidobutan-2-one (1mm) had no effect on the enzyme. 5. The neutral proteinase hydrolysed casein, Bz-dl-Arg-Nan and heat-denatured collagen, but was inactive towards native collagen and several synthetic substrates, such as 4-phenylazobenzyloxycarbonyl-Pro-Leu-Gly-Pro-d-Arg, 3-carboxypropionyl-Ala-Ala-Ala p-nitroanilide and 2,4-dinitrophenyl-Pro-Gln-Gly-Ile-Ala-Gly-Gln-d-Arg, and also proteoglycan. The enzyme did not act as a plasminogen activator. 6. These properties suggested that a neutral proteinase in the human uterine cervix was different from enzymes previously reported.  相似文献   

6.
The uncharacterized gene previously proposed as a mannose-6-phosphate isomerase from Bacillus subtilis was cloned and expressed in Escherichia coli. The maximal activity of the recombinant enzyme was observed at pH 7.5 and 40°C in the presence of 0.5 mM Co2+. The isomerization activity was specific for aldose substrates possessing hydroxyl groups oriented in the same direction at the C-2 and C-3 positions, such as the d and l forms of ribose, lyxose, talose, mannose, and allose. The enzyme exhibited the highest activity for l-ribulose among all pentoses and hexoses. Thus, l-ribose, as a potential starting material for many l-nucleoside-based pharmaceutical compounds, was produced at 213 g/liter from 300-g/liter l-ribulose by mannose-6-phosphate isomerase at 40°C for 3 h, with a conversion yield of 71% and a volumetric productivity of 71 g liter−1 h−1.l-Ribose is a potential starting material for the synthesis of many l-nucleoside-based pharmaceutical compounds, and it is not abundant in nature (5, 19). l-Ribose has been produced mainly by chemical synthesis from l-arabinose, l-xylose, d-glucose, d-galactose, d-ribose, or d-mannono-1,4-lactone (2, 17, 23). Biological l-ribose manufacture has been investigated using ribitol or l-ribulose. Recently, l-ribose was produced from ribitol by a recombinant Escherichia coli containing an NAD-dependent mannitol-1-dehydrogenase (MDH) with a 55% conversion yield when 100 g/liter ribitol was used in a 72-h fermentation (18). However, the volumetric productivity of l-ribose in the fermentation is 28-fold lower than that of the chemical method synthesized from l-arabinose (8). l-Ribulose has been biochemically converted from l-ribose using an l-ribose isomerase from an Acinetobacter sp. (9), an l-arabinose isomerase mutant from Escherichia coli (4), a d-xylose isomerase mutant from Actinoplanes missouriensis (14), and a d-lyxose isomerase from Cohnella laeviribosi (3), indicating that l-ribose can be produced from l-ribulose by these enzymes. However, the enzymatic production of l-ribulose is slow, and the enzymatic production of l-ribose from l-ribulose has been not reported.Sugar phosphate isomerases, such as ribose-5-phosphate isomerase, glucose-6-phosphate isomerase, and galactose-6-phosphate isomerase, work as general aldose-ketose isomerases and are useful tools for producing rare sugars, because they convert the substrate sugar phosphates and the substrate sugars without phosphate to have a similar configuration (11, 12, 21, 22). l-Ribose isomerase from an Acinetobacter sp. (9) and d-lyxose isomerase from C. laeviribosi (3) had activity with l-ribose, d-lyxose, and d-mannose. Thus, we can apply mannose-6-phosphate (EC 5.3.1.8) isomerase to the production of l-ribose, because there are no sugar phosphate isomerases relating to l-ribose and d-lyxose. The production of the expensive sugar l-ribose (bulk price, $1,000/kg) from the rare sugar l-ribulose by mannose-6-phosphate isomerase may prove to be a valuable industrial process, because we have produced l-ribulose from the cheap sugar l-arabinose (bulk price, $50/kg) using the l-arabinose isomerase from Geobacillus thermodenitrificans (20) (Fig. (Fig.11).Open in a separate windowFIG. 1.Schematic representation for the production of l-ribulose from l-arabinose by G. thermodenitrificans l-arabinose isomerase and the production of l-ribose from l-ribulose by B. subtilis mannose-6-phosphate isomerase.In this study, the gene encoding mannose-6-phosphate isomerase from Bacillus subtilis was cloned and expressed in E. coli. The substrate specificity of the recombinant enzyme for various aldoses and ketoses was investigated, and l-ribulose exhibited the highest activity among all pentoses and hexoses. Therefore, mannose-6-phosphate isomerase was applied to the production of l-ribose from l-ribulose.  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.
The oxidation of d- and l-glycerate by rat liver   总被引:1,自引:1,他引:0  
1. The interconversion of hydroxypyruvate and l-glycerate in the presence of NAD and rat-liver l-lactate dehydrogenase has been demonstrated. Michaelis constants for these substrates together with an equilibrium constant have been determined and compared with those for pyruvate and l-lactate. 2. The presence of d-glycerate dehydrogenase in rat liver has been confirmed and the enzyme has been purified 16–20-fold from the supernatant fraction of a homogenate, when it is free of l-lactate dehydrogenase, with a 23–29% recovery. The enzyme catalyses the interconversion of hydroxypyruvate and d-glycerate in the presence of either NAD or NADP with almost equal efficiency. d-Glycerate dehydrogenase also catalyses the reduction of glyoxylate, but is distinct from l-lactate dehydrogenase in that it fails to act on pyruvate, d-lactate or l-lactate. The enzyme is strongly dependent on free thiol groups, as shown by inhibition with p-chloromercuribenzoate, and in the presence of sodium chloride the reduction of hydroxypyruvate is activated. Michaelis constants for these substrates of d-glycerate dehydrogenase and an equilibrium constant for the NAD-catalysed reaction have been calculated. 3. An explanation for the lowered Vmax. with d-glycerate as compared with dl-glycerate for the rabbit-kidney d-α-hydroxy acid dehydrogenase has been proposed.  相似文献   

9.
The mechanism of hexose transport into plasma membrane vesicles isolated from mature sugarbeet leaves (Beta vulgaris L.) was investigated. The initial rate of glucose uptake into the vesicles was stimulated approximately fivefold by imposing a transmembrane pH gradient (ΔpH), alkaline inside, and approximately fourfold by a negative membrane potential (ΔΨ), generated as a K+-diffusion potential, negative inside. The -fold stimulation was directly related to the relative ΔpH or ΔΨ gradient imposed, which were determined by the uptake of acetate or tetraphenylphosphonium, respectively. ΔΨ- and ΔpH-dependent glucose uptake showed saturation kinetics with a Km of 286 micromolar for glucose. Other hexose molecules (e.g. 2-deoxy-d-glucose, 3-O-methyl-d-glucose, and d-mannose) were also accumulated into plasma membrane vesicles in a ΔpH-dependent manner. Inhibition constants of a number of compounds for glucose uptake were determined. Effective inhibitors of glucose uptake included: 3-O-methyl-d-glucose, 5-thio-d-glucose, d-fructose, d-galactose, and d-mannose, but not 1-O-methyl-d-glucose, d- and l-xylose, l-glucose, d-ribose, and l-sorbose. Under all conditions of proton motive force magnitude and glucose and sucrose concentration tested, there was no effect of sucrose on glucose uptake. Thus, hexose transport on the sugarbeet leaf plasma membrane was by a H+-hexose symporter, and the carrier and possibly the energy source were not shared by the plasma membrane H+-sucrose symporter.  相似文献   

10.
Bacillus licheniformis l-arabinose isomerase (l-AI) is distinguished from other l-AIs by its high degree of substrate specificity for l-arabinose and its high turnover rate. A systematic strategy that included a sequence alignment-based first screening of residues and a homology model-based second screening, followed by site-directed mutagenesis to alter individual screened residues, was used to study the molecular determinants for the catalytic efficiency of B. licheniformis l-AI. One conserved amino acid, Y333, in the substrate binding pocket of the wild-type B. licheniformis l-AI was identified as an important residue affecting the catalytic efficiency of B. licheniformis l-AI. Further insights into the function of residue Y333 were obtained by replacing it with other aromatic, nonpolar hydrophobic amino acids or polar amino acids. Replacing Y333 with the aromatic amino acid Phe did not alter catalytic efficiency toward l-arabinose. In contrast, the activities of mutants containing a hydrophobic amino acid (Ala, Val, or Leu) at position 333 decreased as the size of the hydrophobic side chain of the amino acid decreased. However, mutants containing hydrophilic and charged amino acids, such as Asp, Glu, and Lys, showed almost no activity with l-arabinose. These data and a molecular dynamics simulation suggest that Y333 is involved in the catalytic efficiency of B. licheniformis l-AI.l-Arabinose isomerase (l-AI) is an enzyme that mediates in vivo isomerization between l-arabinose and l-ribulose as well as in vitro isomerization of d-galactose and d-tagatose (20). l-Ribulose (l-erythro-pentulose) is a rare and expensive ketopentose sugar (1) that can be used as a precursor for the production of other rare sugars of high market value, such as l-ribose. Despite being a common metabolic intermediate in different organisms, l-ribulose is scarce in nature. The market for rare and unnatural sugars has been growing, especially in the sweetener and pharmaceutical industries. For example, several modified nucleosides derived from l-sugars have been shown to act as potent antiviral agents and are also useful in antigen therapy. Derivatives of rare sugars have also been used as agents against hepatitis B virus and human immunodeficiency virus (2, 22).For these reasons, interest in the enzymology of rare sugars has also been increasing. Various forms of l-AI from a variety of organisms have been characterized, and some have shown potential for industrial use. Several highly thermotolerant enzyme forms from Thermotoga maritima (12), Thermotoga neapolitana (10), Bacillus stearothermophilus (18), Thermoanaerobacter mathranii (9), and Lactobacillus plantarum (5) have been characterized previously. All of these reported l-AIs tend to have broad specificity, although a few l-AIs with high degrees of substrate specificity for l-arabinose have also been documented.The enzyme properties of l-AIs have been examined by engineering several forms by error-prone PCR and site-directed mutagenesis. Galactose conversion was reportedly enhanced 20% following site-directed introduction of a double mutation (C450S-N475K) into l-AI (16). Error-prone PCR manipulation of l-AI from Geobacillus stearothermophilus resulted in a shift in temperature specificity from 60 to 65°C and increased isomerization activity (11). All of these previously reported mutational studies have been aimed at improving enzymatic properties for industrial application. However, even though the three-dimensional (3D) structure of Escherichia coli l-AI has been determined previously (15), few new structural studies have been performed to decipher the reaction mechanism of this enzyme. Rhimi et al. (19) have reported an important role for D308, F329, E351, and H446 in catalysis, as indicated by findings from site-directed mutagenesis. Nonetheless, detailed analysis of the important molecular determinants controlling the catalytic activities of the l-AIs is still lacking.Previously, we have reported the cloning and characterization of a novel l-AI from Bacillus licheniformis (17). This enzyme can be distinguished from other l-AIs by its wide pH range, high degree of substrate specificity for l-arabinose, and extremely high turnover rate. In the present paper, we report the identification of an important amino acid residue responsible for the catalytic efficiency of l-AIs, as determined by a systematic screening process composed of sequence alignment and molecular dynamics (MD) simulation, followed by site-directed mutagenesis. Using the crystal structure of E. coli l-AI as a template, we have built a 3D model of B. licheniformis l-AI. Analysis of the 3D model of B. licheniformis l-AI docked with l-arabinose, followed by a systematic screening process, showed that Y333 interacted with the substrate, suggesting that this residue in B. licheniformis l-AI may be essential for catalysis. We further characterized the role of Y333 in B. licheniformis l-AI binding of and catalytic efficiency for l-arabinose.  相似文献   

11.
The first enzyme in the pathway for l-arabinose catabolism in eukaryotic microorganisms is a reductase, reducing l-arabinose to l-arabitol. The enzymes catalyzing this reduction are in general nonspecific and would also reduce d-xylose to xylitol, the first step in eukaryotic d-xylose catabolism. It is not clear whether microorganisms use different enzymes depending on the carbon source. Here we show that Aspergillus niger makes use of two different enzymes. We identified, cloned, and characterized an l-arabinose reductase, larA, that is different from the d-xylose reductase, xyrA. The larA is up-regulated on l-arabinose, while the xyrA is up-regulated on d-xylose. There is however an initial up-regulation of larA also on d-xylose but that fades away after about 4 h. The deletion of the larA gene in A. niger results in a slow growth phenotype on l-arabinose, whereas the growth on d-xylose is unaffected. The l-arabinose reductase can convert l-arabinose and d-xylose to their corresponding sugar alcohols but has a higher affinity for l-arabinose. The Km for l-arabinose is 54 ± 6 mm and for d-xylose 155 ± 15 mm.  相似文献   

12.
The biological sulphation of l-tyrosyl peptides   总被引:3,自引:3,他引:0       下载免费PDF全文
1. A rat-liver supernatant preparation can achieve the biological O-sulphation of l-tyrosylglycine and l-tyrosyl-l-alanine at pH7·0. 2. The optimum concentrations of l-tyrosylglycine and l-tyrosyl-l-alanine in this system are 50mm and 60mm respectively. 3. l-Tyrosylglycine yields two sulphated products, whereas l-tyrosyl-l-alanine yields three sulphated products, when used as acceptor for sulphate in the rat-liver system. 4. With both substrates, one of the sulphated products has been identified as the O-sulphate ester of the corresponding parent peptide.  相似文献   

13.
14.
Peptidoglycan hydrolases (PGHs) are responsible for bacterial cell lysis. Most PGHs have a modular structure comprising a catalytic domain and a cell wall-binding domain (CWBD). PGHs of bacteriophage origin, called endolysins, are involved in bacterial lysis at the end of the infection cycle. We have characterized two endolysins, Lc-Lys and Lc-Lys-2, identified in prophages present in the genome of Lactobacillus casei BL23. These two enzymes have different catalytic domains but similar putative C-terminal CWBDs. By analyzing purified peptidoglycan (PG) degradation products, we showed that Lc-Lys is an N-acetylmuramoyl-l-alanine amidase, whereas Lc-Lys-2 is a γ-d-glutamyl-l-lysyl endopeptidase. Remarkably, both lysins were able to lyse only Gram-positive bacterial strains that possess PG with d-Ala4d-Asx-l-Lys3 in their cross-bridge, such as Lactococcus casei, Lactococcus lactis, and Enterococcus faecium. By testing a panel of L. lactis cell wall mutants, we observed that Lc-Lys and Lc-Lys-2 were not able to lyse mutants with a modified PG cross-bridge, constituting d-Ala4l-Ala-(l-Ala/l-Ser)-l-Lys3; moreover, they do not lyse the L. lactis mutant containing only the nonamidated d-Asp cross-bridge, i.e. d-Ala4d-Asp-l-Lys3. In contrast, Lc-Lys could lyse the ampicillin-resistant E. faecium mutant with 3→3 l-Lys3-d-Asn-l-Lys3 bridges replacing the wild-type 4→3 d-Ala4-d-Asn-l-Lys3 bridges. We showed that the C-terminal CWBD of Lc-Lys binds PG containing mainly d-Asn but not PG with only the nonamidated d-Asp-containing cross-bridge, indicating that the CWBD confers to Lc-Lys its narrow specificity. In conclusion, the CWBD characterized in this study is a novel type of PG-binding domain targeting specifically the d-Asn interpeptide bridge of PG.  相似文献   

15.
Organisms that overproduced l-cysteine and l-cystine from glucose were constructed by using Escherichia coli K-12 strains. cysE genes coding for altered serine acetyltransferase, which was genetically desensitized to feedback inhibition by l-cysteine, were constructed by replacing the methionine residue at position 256 of the serine acetyltransferase protein with 19 other amino acid residues or the termination codon to truncate the carboxy terminus from amino acid residues 256 to 273 through site-directed mutagenesis by using PCR. A cysteine auxotroph, strain JM39, was transformed with plasmids having these altered cysE genes. The serine acetyltransferase activities of most of the transformants, which were selected based on restored cysteine requirements and ampicillin resistance, were less sensitive than the serine acetyltransferase activity of the wild type to feedback inhibition by l-cysteine. At the same time, these transformants produced approximately 200 mg of l-cysteine plus l-cystine per liter, whereas these amino acids were not detected in the recombinant strain carrying the wild-type serine acetyltransferase gene. However, the production of l-cysteine and l-cystine by the transformants was very unstable, presumably due to a cysteine-degrading enzyme of the host, such as cysteine desulfhydrase. Therefore, mutants that did not utilize cysteine were derived from host strain JM39 by mutagenesis with N-methyl-N′-nitro-N-nitrosoguanidine. When a newly derived host was transformed with plasmids having the altered cysE genes, we found that the production of l-cysteine plus l-cystine was markedly increased compared to production in JM39.l-Cysteine, one of the important amino acids used in the pharmaceutical, food, and cosmetics industries, has been obtained by extracting it from acid hydrolysates of the keratinous proteins in human hair and feathers. The first successful microbial process used for industrial production of l-cysteine involved the asymmetric conversion of dl-2-aminothiazoline-4-carboxylic acid, an intermediate compound in the chemical synthesis of dl-cysteine, to l-cysteine by enzymes from a newly isolated bacterium, Pseudomonas thiazoliniphilum (11). Yamada and Kumagai (13) also described enzymatic synthesis of l-cysteine from beta-chloroalanine and sodium sulfide in which Enterobacter cloacae cysteine desulfhydrase (CD) was used. However, high level production of l-cysteine from glucose with microorganisms has not been studied.Biosynthesis of l-cysteine in wild-type strains of Escherichia coli and Salmonella typhimurium is regulated through feedback inhibition by l-cysteine of serine acetyltransferase (SAT), a key enzyme in l-cysteine biosynthesis, and repression of expression of a series of enzymes used for sulfide reduction from sulfate by l-cysteine (4), as shown in Fig. Fig.1.1. Denk and Böck reported that a small amount of l-cysteine was excreted by a revertant of a cysteine auxotroph of E. coli. In this revertant, SAT encoded by the cysE gene was desensitized to feedback inhibition by l-cysteine, and the methionine residue at position 256 in SAT was replaced by isoleucine (2). These results indicate that it may be possible to construct organisms that produce high levels of l-cysteine by amplifying an altered cysE gene. Although the residue at position 256 is supposedly part of the allosteric site for cysteine binding, no attention has been given to the effect of an amino acid substitution at position 256 in SAT on feedback inhibition by l-cysteine and production of l-cysteine. It is also not known whether isoleucine is the best residue for desensitization to feedback inhibition. Open in a separate windowFIG. 1Biosynthesis and regulation of l-cysteine in E. coli. Abbreviations: APS, adenosine 5′-phosphosulfate; PAPS, phosphoadenosine 5′-phosphosulfate; Acetyl CoA, acetyl coenzyme A. The open arrow indicates feedback inhibition, and the dotted arrows indicate repression.On the other hand, l-cysteine appears to be degraded by E. coli cells. Therefore, in order to obtain l-cysteine producers, a host strain with a lower level of l-cysteine degradation activity must be isolated. In this paper we describe high-level production of l-cysteine plus l-cystine from glucose by E. coli resulting from construction of altered cysE genes. The methionine residue at position 256 in SAT was replaced by other amino acids or the termination codon in order to truncate the carboxy terminus from amino acid residues 256 to 273 by site-directed mutagenesis. A newly derived cysteine-nondegrading E. coli strain with plasmids having the altered cysE genes was used to investigate production of l-cysteine plus l-cystine.  相似文献   

16.
In comparison to other pseudomonads, Pseudomonas aeruginosa grows poorly in l-lysine as a sole source of nutrient. In this study, the ldcA gene (lysine decarboxylase A; PA1818), previously identified as a member of the ArgR regulon of l-arginine metabolism, was found essential for l-lysine catabolism in this organism. LdcA was purified to homogeneity from a recombinant strain of Escherichia coli, and the results of enzyme characterization revealed that this pyridoxal-5-phosphate-dependent decarboxylase takes l-lysine, but not l-arginine, as a substrate. At an optimal pH of 8.5, cooperative substrate activation by l-lysine was depicted from kinetics studies, with calculated Km and Vmax values of 0.73 mM and 2.2 μmole/mg/min, respectively. Contrarily, the ldcA promoter was induced by exogenous l-arginine but not by l-lysine in the wild-type strain PAO1, and the binding of ArgR to this promoter region was demonstrated by electromobility shift assays. This peculiar arginine control on lysine utilization was also noted from uptake experiments in which incorporation of radioactively labeled l-lysine was enhanced in cells grown in the presence of l-arginine but not l-lysine. Rapid growth on l-lysine was detected in a mutant devoid of the main arginine catabolic pathway and with a higher basal level of the intracellular l-arginine pool and hence elevated ArgR-responsive regulons, including ldcA. Growth on l-lysine as a nitrogen source can also be enhanced when the aruH gene encoding an arginine/lysine:pyruvate transaminase was expressed constitutively from plasmids; however, no growth of the ldcA mutant on l-lysine suggests a minor role of this transaminase in l-lysine catabolism. In summary, this study reveals a tight connection of lysine catabolism to the arginine regulatory network, and the lack of lysine-responsive control on lysine uptake and decarboxylation provides an explanation of l-lysine as a poor nutrient for P. aeruginosa.Decarboxylation of amino acids, including lysine, arginine, and glutamate, is important for bacterial survival under low pH (2, 7, 19). Lysine is abundant in the rhizosphere where fluorescent Pseudomonas preferentially resides, and serves as a nitrogen and carbon source to these organisms (28). In microbes, lysine catabolism can be initiated either through monooxygenase, decarboxylase, or transaminase activities. The monooxygenase pathway has been considered the major route for l-lysine utilization in Pseudomonas putida, and davBATD encoding enzymes for the first four steps of the pathway have been characterized (25, 26). In contrast, Pseudomonas aeruginosa cannot use exogenous l-lysine efficiently for growth (5, 24). It has been reported that enzymatic activities for the first two steps of the monooxygenase pathway are not detectable in P. aeruginosa, and no davBA orthologs can be identified from this organism (24, 25).Mutants of P. aeruginosa with improved growth on l-lysine and a high level of lysine decarboxylase activity can be isolated by repeated subcultures in l-lysine (5). This suggests that in P. aeruginosa, l-lysine utilization might be mediated by the lysine decarboxylase pathway with cadaverine and 5-aminovalerate as intermediates (Fig. (Fig.1).1). Alternatively, conversion of l-lysine into 5-aminovalerate may also be accomplished by a coupled reaction catalyzed by AruH and AruI. The AruH and AruI enzymes were reported as arginine:pyruvate transaminase and 2-ketoarginine decarboxylase, respectively (36). Interestingly, transamination by AruH using l-lysine as an amino group donor can also be detected in vitro (35). The reaction product α-keto-ɛ-aminohexanonate can potentially be decarboxylated into 5-aminovalerate by AruI, providing an alternative route for lysine degradation.Open in a separate windowFIG. 1.Lysine catabolic pathways. l-lysine decarboxylase pathway is shown at center. Broken arrows represent lysine monooxygenase pathway from P. putida which is not present in P. aeruginosa.In this study, we showed that the lysine decarboxylase pathway is the main route for lysine utilization under arginine control. Expression of the ldcAB operon encoding l-lysine decarboxylase and a putative lysine/cadaverine antiporter was analyzed regarding its response to l-lysine, l-arginine, and the arginine-responsive regulator ArgR. Enzyme characterization was performed to verify the function of LdcA as l-lysine decarboxylase. Arginine control on lysine incorporation was also investigated by genetic studies and uptake experiments. The peculiar role of ArgR controlling arginine and lysine uptake and catabolism provides the explanation for poor growth in lysine, and it implies a higher level of complexity in metabolic networks of pseudomonads.  相似文献   

17.
An enzyme catalyzing the formation of δ-aminolevulinic acid by transamination of γ,δ-dioxovaleric acid with l-α-alanine, l-glutamic acid, or l-phenylalanine has been detected in extracts of Chlorella vulgaris. The activity of this enzyme does not appear to parallel changes in chlorophyll content in a Chlorella mutant which requires light for chlorophyll production. The role of this enzyme in δ-aminolevulinic acid metabolism in plants is not clearly understood.  相似文献   

18.
Mechanically isolated Asparagus sprengeri Regel mesophyll cells cause alkalinization of the suspension medium on the addition of l-glutamate or its analog l-methionine-d,l-sulfoximine. Using a radiolabeled pH probe, it was found that both compounds caused internal acidification whereas l-aspartate did not. Fusicoccin stimulated H+ efflux from the cells by 111% and the uptake of l-[U-14C]glutamate by 55%. Manometric experiments demonstrated that, unlike l-methionine-d,l-sulfoximine, l-glutamate stimulated CO2 evolution from nonilluminated cells. Simultaneous measurements of medium alkalinization and 14CO2 evolution upon the addition of labeled l-glutamate showed that alkalinization was immediate and reached a maximum value after 45 minutes whereas 14CO2 evolution exhibited a lag before its appearance and continued in a linear manner for at least 100 minutes. Rates of alkalinization and uptake of l-[U-14C]glutamate were higher in the light while rates of 14CO2 evolution were higher in the dark. The major labeled product of glutamate decarboxylation, γ-aminobutyric acid, was found in the cells and the suspension medium. Its addition to the cell suspension did not result in medium alkalinization and evidence indicates that it is lost from the cell to the medium. The data suggest that the origin of medium alkalinization is co-transport not metabolism, and that the loss of labeled CO2 and γ-aminobutyric acid from the cell result in an overestimation of the stoichiometry of the H+/l-glutamate uptake process.  相似文献   

19.
Studies of inhibition of rat spermidine synthase and spermine synthase   总被引:5,自引:4,他引:1  
1. S-Adenosyl-l-methionine, S-adenosyl-l-homocysteine, 5′-methylthioadenosine and a number of analogues having changes in the base, sugar or amino acid portions of the molecule were tested as potential inhibitors of spermidine synthase and spermine synthase from rat ventral prostate. 2. S-Adenosyl-l-methionine was inhibitory to these reactions, as were other nucleosides containing a sulphonium centre. The most active of these were S-adenosyl-l-ethionine, S-adenosyl-4-methylthiobutyric acid, S-adenosyl-d-methionine and S-tubercidinylmethionine, which were all comparable in activity with S-adenosylmethionine itself, producing 70–98% inhibition at 1mm concentrations. Spermine synthase was somewhat more sensitive than spermidine synthase. 3. 5′-Methylthioadenosine, 5′-ethylthioadenosine and 5′-methylthiotubercidin were all powerful inhibitors of both enzymes, giving 50% inhibition of spermine synthase at 10–15μm and 50% inhibition of spermidine synthase at 30–45μm. 4. S-Adenosyl-l-homocysteine was a weak inhibitor of spermine synthase and practically inactive against spermidine synthase. Analogues of S-adenosylhomocysteine lacking either the carboxy or the amino group of the amino acid portion were somewhat more active, as were derivatives in which the ribose ring had been opened by oxidation. The sulphoxide and sulphone derivatives of decarboxylated S-adenosyl-l-homocysteine and the sulphone of S-adenosyl-l-homocysteine were quite potent inhibitors and were particularly active against spermidine synthase (giving 50% inhibition at 380, 50 and 20μm respectively). 5. These results are discussed in terms of the possible regulation of polyamine synthesis by endogenous nucleosides and the possible value of some of the inhibitory substances in experimental manipulations of polyamine concentrations. It is suggested that 5′-methylthiotubercidin and the sulphone of S-adenosylhomocysteine or of S-adenosyl-3-thiopropylamine may be particularly valuable in this respect.  相似文献   

20.
A Mg2+-dependent, alkaline phosphatase has been isolated from mature pollen of Lilium longiflorum Thunb., cv. Ace and partially purified. It hydrolyzes 1l- and 1d-myo-inositol 1-phosphate, myo-inositol 2-phosphate, and β-glycerophosphate at rates decreasing in the order named. The affinity of the enzyme for 1l- and 1d-myo-inositol 1-phosphate is approximately 10-fold greater than its affinity for myo-inositol 2-phosphate. Little or no activity is found with phytate, d-glucose 6-phosphate, d-glucose 1-phosphate, d-fructose 1-phosphate, d-fructose 6-phosphate, d-mannose 6-phosphate, or p-nitrophenyl phosphate. 3-Phosphosphoglycerate is a weak competitive inhibitor. myo-Inositol does not inhibit the reaction. Optimal activity is obtained at pH 8.5 and requires the presence of Mg2+. At 4 millimolar, Co2+, Fe2+ or Mn2+ are less effective. Substantial inhibition is obtained with 0.25 molar Li+. With β-glycerophosphate as substrate the Km is 0.06 millimolar and the reaction remains linear at least 2 hours. In 0.1 molar Tris, β-glycerophosphate yields equivalent amounts of glycerol and inorganic phosphate, evidence that transphosphorylation does not occur.  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