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.
Lysine decarboxylase (LDC, EC 4.1.1.18) from Selenomonas ruminantium has decarboxylating activities towards both L-lysine and L-ornithine with similar K(m) and Vmax. Here, we identified four amino acid residues that confer substrate specificity upon S. ruminantium LDC and that are located in its catalytic domain. We have succeeded in converting S. ruminantium LDC to an enzyme with a preference in decarboxylating activity for L-ornithine when the four-residue of LDC were replaced by the corresponding residues of mouse ornithine decarboxylase (EC 4.1.1.17).  相似文献   

2.
Lysine decarboxylase (LDC; EC 4.1.1.18) of Selenomonas ruminantium is a constitutive enzyme and is involved in the synthesis of cadaverine, which is an essential constituent of the peptidoglycan for normal cell growth. We purified the S. ruminantium LDC by an improved method including hydrophobic chromatography and studied the fine characteristics of the enzyme. Kinetic study of LDC showed that S. ruminantium LDC decarboxylated both L-lysine and L-ornithine with similar Km and the decarboxylase activities towards both substrates were competitively and irreversibly inhibited by DL-alpha-difluoromethylornithine, which is a specific inhibitor of ornithine decarboxylase (EC 4.1.1.17). We also showed a drastic descent of LDC activity owing to the degradation of LDC at entry into the stationary phase of cell growth.  相似文献   

3.
The kinetics of L-lysine production by a fluoropyruvate sensitive (FPs) mutant of B.lactofermentum have been analysed in batch and continuous culture. Under optimal conditions (viz. T = 30° C, pH = 7.0, DO = 10 % air saturation) it was established that the initial phosphate level could be used to extend the batch culture to give an L-lysine.HCl concentration of 68.0 g/l and a yield of 0.43 g/g. In continuous culture the yield increased at low dilution rates under conditions similar to those for extended batch culture.  相似文献   

4.
To identify potential L-lysine export limitations by Corynebacterium glutamicum in the L-lysine production process, the excretion of L-lysine was studied in continuous and fed-batch operated stirred tank reactors. A structured biochemical model of the L-lysine excretion mechanism was used to determine the activity of the export carrier and to calculate a cell-specific concentration of the export carrier. For the biochemical characterization of this specific carrier concentration a standardized L-lysine efflux test was developed. Carrier activity, cell-specific carrier concentration, and the specific L-lysine export rate were identified as a function of pH value and L-lysine concentration in the reactors. Also, the correlation of these parameters to the metabolic state of C. glutamicum was determined. The pH value in the reactor governs the carrier activity (maximum at pH 6.5) and the specific carrier concentration (maximum at pH 8.0). The specific L-lysine export rate, as the product of carrier activity and specific carrier concentration, revealed a maximum at pH 7.0. Decreasing L-lysine productivities also correlated with decreasing specific carrier concentrations. The L-lysine concentration in the reactor had no effect on the specific carrier concentration but strongly inhibited the carrier activity. The specific export rate was reduced to 50% at 400 mM L-lysine compared to the specific export rate at 80 mM L-lysine. (c) 1996 John Wiley & Sons, Inc.  相似文献   

5.
The yeast Candida albicans is able to utilize L-lysine as the sole nitrogen and carbon source accompanied by intracellular accumulation of alpha-aminoadipate-delta-semialdehyde. A novel yeast amino acid dehydrogenase catalysing the oxidative deamination of the epsilon-group of L-lysine was found in this yeast. The enzyme, L-lysine epsilon-dehydrogenase, is strongly induced in cells grown on L-lysine as the sole nitrogen source. The enzyme is specific for both L-lysine and NADP+. The Km values were determined to be 0.87 mM for L-lysine and 0.071 mM for NADP+. An apparent Mr of 87,000 was estimated by gel filtration. The enzyme has maximum activity at pH 9.5 and a temperature optimum of 32 degrees C under our assay conditions.  相似文献   

6.
Polyamines are fundamental molecules of life, and their deep evolutionary history is reflected in extensive biosynthetic diversification. The polyamines putrescine, agmatine, and cadaverine are produced by pyridoxal 5′-phosphate-dependent L-ornithine, L-arginine, and L-lysine decarboxylases (ODC, ADC, LDC), respectively, from both the alanine racemase (AR) and aspartate aminotransferase (AAT) folds. Two homologous forms of AAT-fold decarboxylase are present in bacteria: an ancestral form and a derived, acid-inducible extended form containing an N-terminal fusion to the receiver-like domain of a bacterial response regulator. Only ADC was known from the ancestral form and limited to the Firmicutes phylum, whereas extended forms of ADC, ODC, and LDC are present in Proteobacteria and Firmicutes. Here, we report the discovery of ancestral form ODC, LDC, and bifunctional O/LDC and extend the phylogenetic diversity of functionally characterized ancestral ADC, ODC, and LDC to include phyla Fusobacteria, Caldiserica, Nitrospirae, and Euryarchaeota. Using purified recombinant enzymes, we show that these ancestral forms have a nascent ability to decarboxylate kinetically less preferred amino acid substrates with low efficiency, and that product inhibition primarily affects preferred substrates. We also note a correlation between the presence of ancestral ODC and ornithine/arginine auxotrophy and link this with a known symbiotic dependence on exogenous ornithine produced by species using the arginine deiminase system. Finally, we show that ADC, ODC, and LDC activities emerged independently, in parallel, in the homologous AAT-fold ancestral and extended forms. The emergence of the same ODC, ADC, and LDC activities in the nonhomologous AR-fold suggests that polyamine biosynthesis may be inevitable.  相似文献   

7.
Lysine decarboxylase (LDC; EC 4.1.1.18) from Selenomonas ruminantium comprises two identical monomeric subunits of 43 kDa and has decarboxylating activities toward both L-lysine and L-ornithine with similar K(m) and V(max) values (Y. Takatsuka, M. Onoda, T. Sugiyama, K. Muramoto, T. Tomita, and Y. Kamio, Biosci. Biotechnol. Biochem. 62:1063-1069, 1999). Here, the LDC-encoding gene (ldc) of this bacterium was cloned and characterized. DNA sequencing analysis revealed that the amino acid sequence of S. ruminantium LDC is 35% identical to those of eukaryotic ornithine decarboxylases (ODCs; EC 4.1.1.17), including the mouse, Saccharomyces cerevisiae, Neurospora crassa, Trypanosoma brucei, and Caenorhabditis elegans enzymes. In addition, 26 amino acid residues, K69, D88, E94, D134, R154, K169, H197, D233, G235, G236, G237, F238, E274, G276, R277, Y278, K294, Y323, Y331, D332, C360, D361, D364, G387, Y389, and F397 (mouse ODC numbering), all of which are implicated in the formation of the pyridoxal phosphate-binding domain and the substrate-binding domain and in dimer stabilization with the eukaryotic ODCs, were also conserved in S. ruminantium LDC. Computer analysis of the putative secondary structure of S. ruminantium LDC showed that it is approximately 70% identical to that of mouse ODC. We identified five amino acid residues, A44, G45, V46, P54, and S322, within the LDC catalytic domain that confer decarboxylase activities toward both L-lysine and L-ornithine with a substrate specificity ratio of 0.83 (defined as the k(cat)/K(m) ratio obtained with L-ornithine relative to that obtained with L-lysine). We have succeeded in converting S. ruminantium LDC to form with a substrate specificity ratio of 58 (70 times that of wild-type LDC) by constructing a mutant protein, A44V/G45T/V46P/P54D/S322A. In this study, we also showed that G350 is a crucial residue for stabilization of the dimer in S. ruminantium LDC.  相似文献   

8.
Three Trp variants of lysyl-tRNA synthetase from Bacillus stearothermophilus, in which either one or both of the two Trp residues within the enzyme (Trp314 and Trp332) were substituted by a Phe residue, were produced by site-directed mutagenesis without appreciable loss of catalytic activity. The following two phenomena were observed with W332F and with the wild-type enzyme, but not with W314F: (1) the addition of L-lysine alone decreased the protein fluorescence of the enzyme, but the addition of ATP alone did not; (2) the subsequent addition of ATP after the addition of excess L-lysine restored the fluorescence to its original level. Fluorometry under various conditions and UV-absorption spectroscopy revealed that Trp314, which was about 20A away from the lysine binding site and was shielded in a non-polar environment, was solely responsible for the fluorescence changes of the enzyme in the L-lysine activation reaction. Furthermore, the microenvironmental conditions around the residue were made more polar upon the binding of L-lysine, though its contact with the solvent was still restricted. It was suggested that Trp314 was located in a less polar environment than was Trp332, after comparison of the wavelengths at the peaks of fluorescence emission and of the relative fluorescence quantum yields. Trp332 was thought, based on the fluorescence quenching by some perturbants and the chemical modification with N-bromosuccinimide, to be on the surface of the enzyme, whereas Trp314 was buried inside. The UV absorption difference spectra induced by the L-lysine binding indicated that the state of Trp314, including its electrostatic environment, changed during the process, but Trp332 did not change. The increased fluorescence from Trp314 at acidic pH compared with that at neutral pH suggests that carboxylate(s) are in close proximity to the Trp314 residue.  相似文献   

9.
Summary The effect of several parameters (pH, time of reaction, temperature, enzyme concentration) on trypsin immobilization onto glutaraldehyde-activated amine-Spherosil was investigated. This activated support could be stored over long periods of time without any important loss of capacity for trypsin coupling. When increasing the amount of trypsin bound to the carrier, enzymatic activity shows an optimal value, beyond which an augmentation of Spherosil enzyme content results in a lowered activity. The influence of the number of available reactive aldehyde groups on silica was investigated by coupling L-lysine to activated support either prior to or simulataneously with trypsin immobilization. In both cases, the activity of trypsin derivatives is decreased when L-lysine concentration is increased, yet the activity of trypsin derivatives is never equal to zero, even in presence of a large excess of L-lysine. This suggests the presence of two types of reactive groups on the activated support.  相似文献   

10.
用响应面法对蜂房哈夫尼菌(Hafnia alvei)L-赖氨酸脱羧酶产酶诱导条件进行优化。首先通过单因素实验对产酶体系的pH、震荡培养时间、静置培养时间、诱导物添加量和Ⅷ添加量进行优化。在此基础上,用部分因子重复试验筛选出对酶活影响显著的3个因素(静置培养时间,诱导物添加量,VB6添加量),再通过Box-behnken实验对这三个因素进行优化,得出最优值。最终得到产酶最佳诱导条件为:震荡培养阶段培养基pH6.5,静置培养阶段pH5.5;摇床震荡培养11h后静置培养7.5h,诱导物L一赖氨酸加入量为5.18dL,维生素B6加入量为1.38g/L时酶活最高,达到71.2U/mL,为优化前(1.74u/mL)的41.8倍,在单因素的基础上提高了19%。  相似文献   

11.
The possibility of producing L-lysine from chemically synthesized DL-lysine has been investigated. Optical resolution of racemic DK-lysine may be achieved by using the stereospecific esterasic activity of trypsin on DL-lysine methyl ester, which gives L-lysine and unchanged D-lysine methyl ester. SL-lysine methyl ester spontaneous hydrolysis may be neglected when operating at pH 5.5 and 30 degrees C. Effect of pH and substrate concentration on hydrolysis rate has been investigated when using as a catalyst either soluble or immobilized trypsin. For this purpose, trypsin was coupled onto an amine porous silica, Spherosil, activated with glutaraldehyde. The optimal pH is 5.8 for soluble trypsin and 6.0 for immobilized trypsin. It was yet possible to lower the parent optimal pH of immobilized trypsin, and thus increase its activity at 5.5, by co-grafting onto Spherosil an aminosilane, for enzyme coupling via glutaraldehyde activation and a positively charged diethyl amino ethyl (DEAE) silane, for decreasing the pH of trypsin microenvironment.  相似文献   

12.
The degradation of mammalian ornithine decarboxylase (ODC) (EC 4.1.1.17) by 26 S proteasome, is accelerated by the ODC antizyme (AZ), a trigger protein involved in the specific degradation of eukaryotic ODC. In prokaryotes, AZ has not been found. Previously, we found that in Selenomonas ruminantium, a strictly anaerobic and Gram-negative bacterium, a drastic degradation of lysine decarboxylase (LDC; EC 4.1.1.18), which has decarboxylase activities toward both L-lysine and L-ornithine with similar K(m) values, occurs upon entry into the stationary phase of cell growth by protease together with a protein of 22 kDa (P22). Here, we show that P22 is a direct counterpart of eukaryotic AZ by the following evidence. (i) P22 synthesis is induced by putrescine but not cadaverine. (ii) P22 enhances the degradation of both mouse ODC and S. ruminantium LDC by a 26 S proteasome. (iii) S. ruminantium LDC degradation is also enhanced by mouse AZ replacing P22 in a cell-free extract from S. ruminantium. (iv) Both P22 and mouse AZ bind to S. ruminantium LDC but not to the LDC mutated in its binding site for P22 and AZ. In this report, we also show that P22 is a ribosomal protein of S. ruminantium.  相似文献   

13.
L-Lysine dehydrogenase, which catalyzes the oxidative deamination of L-lysine in the presence of NAD, was found in the thermophilic bacterium Geobacillus stearothermophilus UTB 1103 and then purified about 3,040-fold from a crude extract of the organism by using four successive column chromatography steps. This is the first report showing the presence of a thermophilic NAD-dependent lysine dehydrogenase. The product of the enzyme catalytic activity was determined to be Delta1-piperideine-6-carboxylate, indicating that the enzyme is L-lysine 6-dehydrogenase (LysDH) (EC 1.4.1.18). The molecular mass of the purified protein was about 260 kDa, and the molecule was determined to be a homohexamer with subunit molecular mass of about 43 kDa. The optimum pH and temperature for the catalytic activity of the enzyme were about 10.1 and 70 degrees C, respectively. No activity was lost at temperatures up to 65 degrees C in the presence of 5 mM L-lysine. The enzyme was relatively selective for L-lysine as the electron donor, and either NAD or NADP could serve as the electron acceptor (NADP exhibited about 22% of the activity of NAD). The Km values for L-lysine, NAD, and NADP at 50 degrees C and pH 10.0 were 0.73, 0.088, and 0.48 mM, respectively. When the gene encoding this LysDH was cloned and overexpressed in Escherichia coli, a crude extract of the recombinant cells had about 800-fold-higher enzyme activity than the extract of G. stearothermophilus. The nucleotide sequence of the LysDH gene encoded a peptide containing 385 amino acids with a calculated molecular mass of 42,239 Da.  相似文献   

14.
Summary The amino acid L-lysine was produced from auxotrophic-regulatory mutants ofBacillus stearothermophilus at a temperature of 60–65°C. One of the mutants (AEC 12 A5, S-(2-aminoethyl)-cysteiner, homoserine), produced L-lysine at the concentration of 7.5 g/l in shaken flasks in minimal medium containing 5% glucose. Culture conditions for optimizing L-lysine production were not investigated. The aspartokinase activity of the wild strainB. stearothermophilus Zu 183 was inhibited by lysine alone and by threonine plus lysine. AEC resistant mutants showed an aspartokinase activity genetically desensitized to the feedback inhibition. Optimal temperature and pH of aspartokinase were 45°C and 9.5, respectively. The data provide significant evidence that mutants of the speciesB. stearothermophilus have a potential value for amino acid production.  相似文献   

15.
Lysine epsilon-dehydrogenase, which has been purified to homogeneity from the extract of Agrobacterium tumefaciens ICR 1600, had a molecular weight of approximately 78,000 and consisted of two subunits identical in molecular weight (about 39,000). The enzyme showed a high substrate specificity. In addition to L-lysine, S-(beta-aminoethyl)-L-cysteine was deaminated by the enzyme, but to a far lesser extent. NAD+ and some NAD+ analogs (deamino-NAD+ and 3-acetylpyridine-NAD+) served as a cofactor. The pH optimum was at about 9.7 for the deamination of L-lysine. Although the NAD+ saturation curve was hyperbolic, a sigmoid saturation curve for L-lysine was obtained with the diluted enzyme solution, in which the dimeric enzyme was predominant. The reversible association of the enzyme to the tetramer was induced either by increasing the enzyme concentration or by addition of L-lysine. The preincubation of the enzyme with 5 mM L-lysine resulted in a 2-fold increase in the activity and gave a hyperbolic saturation curve for L-lysine. Upon modification of SH groups of the enzyme with DTNB, neither the interconversion between the dimer and the tetramer nor the activation by L-lysine occurred. These results indicated that the dimeric enzyme was activated by L-lysine and the activation resulted from the association of two dimeric enzymes to form a tetramer.  相似文献   

16.
Changes in lysine decarboxylase (LDC) activity caused by Sitobion avenae (F.) feeding on two winter triticale cultivars (cvs) were studied. The aphid fecundity and values of intrinsic rate of natural increase showed that cv Witon was less susceptible to S. avenae than cv Tornado. The grain aphid feeding on more susceptible triticale caused a decrease in the LDC activity, with exceptions of root tissues after two weeks of the feeding. In case of less susceptible cv Witon reduction of the LDC activity was observed only during initial period of S. avenae feeding. Later the aphid infestation induced activity of the LDC within tissues of cv Witon.  相似文献   

17.
Selenomonas ruminantium synthesizes cadaverine and putrescine from L-lysine and L-ornithine as the essential constituents of its peptidoglycan by a constitutive lysine/ornithine decarboxylase (LDC/ODC). S. ruminantium grew normally in the presence of the specific inhibitor for LDC/ODC, DL-alpha-difluoromethylornithine, when arginine was supplied in the medium. In this study, we discovered the presence of arginine decarboxylase (ADC), the key enzyme in agmatine pathway for putrescine synthesis, in S. ruminantium. We purified and characterized ADC and cloned its gene (adc) from S. ruminantium chromosomal DNA. ADC showed more than 60% identity with those of LDC/ODC/ADCs from Gram-positive bacteria, but no similarity to that from Gram-negative bacteria. In this study, we also cloned the aguA and aguB genes, encoding agmatine deiminase (AguA) and N-carbamoyl-putrescine amidohydrolase (AguB), both of which are involved in conversion from agmatine into putrescine. AguA and AguB were expressed in S. ruminantium. Hence, we concluded that S. ruminantium has both ornithine and agmatine pathways for the synthesis of putrescine.  相似文献   

18.
The activity of 6-phosphogluconate dehydrogenase, aspartate kinase and phosphoenolpyruvate carboxylase has been studied at different dilution rates in aerobic continuous culture of Corynebacterium glutamicum. 6-Phosphogluconate dehydrogenase and aspartate kinase reached their maximum values at the lower dilution rates (0.02–0.06 h–1), when L-lysine was produced. The phosphoenolpyruvate carboxylase activity seemed to be independent of metabolite synthesis. The production of L-lysine was also studied in non-growing cells in batch cultures. In these conditions, statistical analysis revealed significant differences in L-lysine titres when glucose or gluconic acid were used as carbon sources. Higher L-lysine concentration obtained with gluconic acid was found to be associated with a high 6-phosphogluconate dehydrogenase activity.  相似文献   

19.
20.
An IgG1 monoclonal antibody (MAB) was isolated from hybridoma culture supernatant by affinity precipitation with an Eudragit S-100-based heterobifunctional ligand. Affinity binding was performed in a homogeneous aqueous phase at pH 7.5 followed by precipitation of the bound affinity complex by lowering the pH to 4.8. After two washing steps, elution of specifically bound MAB was achieved by incubating the precipitate with 0.1 M glycine.HCl pH 2.5. The influence of elution volume and time on the recovery of active MAB and the overall purification factor were studied. The best conditions enabled the recovery of 50.2% of active MAB with a purification factor of 6.2. A further dialysis against 50 mM Tris.HCl pH 8.0 increased the activity yield and the purification factor to 68.4% and 8.3, respectively. This result showed that part of the antibody activity loss during affinity precipitation was due to a reversible inactivation process, being easily recovered after a refining dialysis step.  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