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 17 毫秒
1.
A general radiochemical method for estimating the activity of amino acid decarboxylases is reported. This method utilizes ion-exchange cartridges to separate unreacted radiolabeled amino acid substrates from product amines, which can then readily be quantitated by liquid scintillation counting. The assay is simple, rapid, and more sensitive than standard 14CO2 trapping procedures if uniformly labeled amino acid substrates are utilized. Acidic, basic, and aromatic amino acid decarboxylases can be assayed with the appropriate choice of cation or anion exchangers. The utility of the method is demonstrated for aspartate-alpha-decarboxylase, tyrosine decarboxylase, and lysine decarboxylase where kinetic parameters are comparable to values obtained by standard radiochemical 14CO2 trapping assays.  相似文献   

2.
A spectrophotometric method for assaying the activity of three amino acid decarboxylases is reported. This method makes use of the coupled reaction of the decarboxylase with phosphoenolpyruvate carboxylase and malate dehydrogenase. The assay is simple and rapid and allows continuous monitoring of the reaction progress. The kinetic parameters obtained using this method for diaminopimelate decarboxylase, lysine decarboxylase, and arginine decarboxylase are comparable to values obtained by radiochemical methods.  相似文献   

3.
Amino acid decarboxylases catalyze decarboxylation of amino acids into amines that possess wide industrial applications. As key enzymes in biobased production of industrially important amines such as cadaverine, putrescine and β-alanine, lysine decarboxylase, ornithine decarboxylase and aspartic acid decarboxylase have attracted increasing attention. To develop enzyme variants with superior catalytic properties, there is a great need for high-throughput assay of these decarboxylases. Here we report the development of assays based on the color change of pH indicator – chlorophenol red (CPR) or bromothymol blue (BTB) – in decarboxylation reactions, in which one proton was consumed per carboxylic group decarboxylated resulting in an increase in pH. First, two buffer-indicator pairs, 4-morpholineethanesulfonic acid (MES)-CPR and 3-morpholinopropanesulfonic acid (MOPS)-BTB, were chosen on the basis of their similar pKa values at approximately pH 6.0 and 7.0, both of which are physiologically relevant. Next, the effects of buffer strength and indicator concentration on absorbance changes were examined in assay mixtures with NaOH titration, which mimicked proton consumption in decarboxylation reactions. Finally, high-throughput quantification of lysine decarboxylase, ornithine decarboxylase and aspartic acid decarboxylase was achieved using a microplate format. These results suggest that our indicator assay system may have potential applications for screening diverse decarboxylases.  相似文献   

4.
The lysA gene encodes meso-diaminopimelate (DAP) decarboxylase (E.C.4.1.1.20), the last enzyme of the lysine biosynthetic pathway in bacteria. We have determined the nucleotide sequence of the lysA gene from Pseudomonas aeruginosa. Comparison of the deduced amino acid sequence of the lysA gene product revealed extensive similarity with the sequences of the functionally equivalent enzymes from Escherichia coli and Corynebacterium glutamicum. Even though both P. aeruginosa and E. coli are Gram-negative bacteria, sequence comparisons indicate a greater similarity between enzymes of P. aeruginosa and the Gram- positive bacterium C. glutamicum than between those of P. aeruginosa and E. coli enzymes. Comparison of DAP decarboxylase with protein sequences present in data bases revealed that bacterial DAP decarboxylases are homologous to mouse (Mus musculus) ornithine decarboxylase (E.C.4.1.1.17), the key enzyme in polyamine biosynthesis in mammals. On the other hand, no similarity was detected between DAP decarboxylases and other bacterial amino acid decarboxylases.   相似文献   

5.
A reliable and reproducible assay was developed for measuring mitochondrial α-keto acid decarboxylase activity using ferricyanide as the electron acceptor. This method permitted the functional isolation and investigation of the decarboxylase step of the branched-chain α-keto acid dehydrogenases in rat liver mitochondria. Pyruvate and α-ketoglutarate decarboxylases are known to be separate and distinct enzymes from the branched-chain α-keto acid decarboxylases and were studied as controls. The relative specific activities of rat liver mitochondrial decarboxylases as measured by the ferricyanide assay showed that pyruvate and α-ketoglutarate were decarboxylated twice as rapidly as α-ketoisovalerate and four to ten times as fast as α-keto-β-methylvalerate and α-ketoisocaproate. The three branched-chain α-keto acids individually inhibit pyruvate and α-ketoglutarate decarboxylases. Inactivation of mitochondrial branched-chain α-keto acid decarboxylase activity by freezing and thawing and by prolonged storage resulted in a proportional decrease in decarboxylase activity toward each of the three branched-chain α-keto acids. However, hypophysectomy was found to increase decarboxylase activity with α-keto-β-methylvalerate to four times normal and with α-ketoisovalerate to three times normal, but the activity with α-ketoisocaproate was not changed. Hypophysectomy did not alter mitochondrial decarboxylase activity with pyruvate, α-ketoglutarate, or α-ketovalerate. The finding that hypophysectomy differentially alters the mitochondrial decarboxylase activity with the three branched-chain α-keto acids suggests either that there is more than one substrate-specific enzyme with branched-chain α-keto acid decarboxylase activity or that there is a modification of one enzyme such that the catalytic activity is selectively altered toward the three substrates.  相似文献   

6.
gamma-Carboxymuconolactone decarboxylase (EC 4.1.1.44) from Azotobacter vinelandii resembled the isofunctional enzymes from Acinetobacter calcoaceticus and Pseudomonas putida. All three decarboxylases appeared to be hexamers formed by association of identical subunits of about 13,300 daltons. The A. vinelandii and P. putida decarboxylases cross-reacted immunologically with each other, and the NH2-terminal amino acid sequences of the enzymes differed in no more than 7 of the first 36 residues. In contrast, the A. calcoaceticus decarboxylase did not cross-react with the decarboxylase from A. vinelandii or P. putida; the NH2-terminal amino acid sequences of these enzymes diverged about 50% from the NH2-terminal amino acid sequence of the A. calcoaceticus decarboxylase.  相似文献   

7.
8.
9.
A sensitive method for the determination of arginine decarboxylase from oat seedlings has been developed, which is based upon the estimation of agmatine, the decarboxylation product of arginine. In the presence of pea seedling amine oxidase, the agmatine is oxidised and the hydrogen peroxide generated is estimated as a red-brown chromogen formed on the peroxidative oxidation of guaiacol. The method may be applied to the estimation of other amino acid decarboxylases.  相似文献   

10.
Animal aspartate decarboxylase (ADC), glutamate decarboxylase (GDC) and cysteine sulfinic acid decarboxylase (CSADC) catalyze the decarboxylation of aspartate, glutamate and cysteine sulfinic acid to β-alanine, γ-aminobutyric acid and hypotaurine, respectively. Each enzymatic product has been implicated in different physiological functions. These decarboxylases use pyridoxal 5-phosphate (PLP) as cofactor and share high sequence homology. Analysis of the activity of ADC in the presence of different amino determined that beta-alanine production from aspartate was diminished in the presence of cysteine. Comparative analysis established that cysteine also inhibited GDC and CSADC in a concentration-dependent manner. Spectral comparisons of free PLP and cysteine, together with ADC and cysteine, result in comparable spectral shifts. Such spectral shifts indicate that cysteine is able to enter the active site of the enzyme, interact with the PLP-lysine internal aldimine, form a cysteine-PLP aldimine and undergo intramolecular nucleophilic cyclization through its sulfhydryl group, leading to irreversible ADC inactivation. Cysteine is the building block for protein synthesis and a precursor of cysteine sulfinic acid that is the substrate of CSADC and therefore is present in many cells, but the presence of cysteine (at comparable concentrations to their natural substrates) apparently could severely inhibit ADC, CSADC and GDC activity. This raises an essential question as to how animal species prevent these enzymes from cysteine-mediated inactivation. Disorders of cysteine metabolism have been implicated in several neurodegenerative diseases. The results of our study should promote research in terms of mechanism by which animals maintain their cysteine homeostasis and possible relationship of cysteine-mediated GDC and CSADC inhibition in neurodegenerative disease development.  相似文献   

11.
Control of Acid Resistance in Escherichia coli   总被引:4,自引:0,他引:4       下载免费PDF全文
Acid resistance (AR) in Escherichia coli is defined as the ability to withstand an acid challenge of pH 2.5 or less and is a trait generally restricted to stationary-phase cells. Earlier reports described three AR systems in E. coli. In the present study, the genetics and control of these three systems have been more clearly defined. Expression of the first AR system (designated the oxidative or glucose-repressed AR system) was previously shown to require the alternative sigma factor RpoS. Consistent with glucose repression, this system also proved to be dependent in many situations on the cyclic AMP receptor protein. The second AR system required the addition of arginine during pH 2.5 acid challenge, the structural gene for arginine decarboxylase (adiA), and the regulator cysB, confirming earlier reports. The third AR system required glutamate for protection at pH 2.5, one of two genes encoding glutamate decarboxylase (gadA or gadB), and the gene encoding the putative glutamate:gamma-aminobutyric acid antiporter (gadC). Only one of the two glutamate decarboxylases was needed for protection at pH 2.5. However, survival at pH 2 required both glutamate decarboxylase isozymes. Stationary phase and acid pH regulation of the gad genes proved separable. Stationary-phase induction of gadA and gadB required the alternative sigma factor sigmaS encoded by rpoS. However, acid induction of these enzymes, which was demonstrated to occur in exponential- and stationary-phase cells, proved to be sigmaS independent. Neither gad gene required the presence of volatile fatty acids for induction. The data also indicate that AR via the amino acid decarboxylase systems requires more than an inducible decarboxylase and antiporter. Another surprising finding was that the sigmaS-dependent oxidative system, originally thought to be acid induced, actually proved to be induced following entry into stationary phase regardless of the pH. However, an inhibitor produced at pH 8 somehow interferes with the activity of this system, giving the illusion of acid induction. The results also revealed that the AR system affording the most effective protection at pH 2 in complex medium (either Luria-Bertani broth or brain heart infusion broth plus 0.4% glucose) is the glutamate-dependent GAD system. Thus, E. coli possesses three overlapping acid survival systems whose various levels of control and differing requirements for activity ensure that at least one system will be available to protect the stationary-phase cell under naturally occurring acidic environments.  相似文献   

12.
Some lactic acid bacteria contain a tyrosine decarboxylase (TDC) which converts tyrosine to tyramine, a biogenic amine frequently encountered in fermented food and wine. Purification and microsequencing of the TDC of Lactobacillus brevis IOEB 9809 allowed us to determine a partial sequence of the TDC gene encoding 264 amino acids of the enzyme. Analysis of this protein sequence revealed typical features of pyridoxal phosphate-dependent amino acid decarboxylases while not any known decarboxylase was closely related to the TDC of L. brevis IOEB 9809. In addition, we could detect other L. brevis strains carrying a TDC gene in a rapid assay based on the polymerase chain reaction.  相似文献   

13.
The induction of several amino acid decarboxylases under anaerobic conditions at low pH has been known for many years, but the mechanism associated with this type of regulation has not been elucidated. To study the regulation of the biodegradative arginine and lysine decarboxylases of Escherichia coli K12, Mudlac fusions to these genes were isolated. Mudlac fusion strains deficient for lysine decarboxylase or arginine decarboxylase were identified using decarboxylase indicator media and analysed for their regulation of beta-galactosidase expression. The position of the Mudlac fusion in lysine decarboxylase-deficient strains has been mapped to the cadA gene at 93.7 minutes, while the Mudlac fusions exhibiting a deficiency in the inducible arginine decarboxylase have been mapped to 93.4 minutes.  相似文献   

14.
During a screening program intended to identify genes encoding enzymes typical for secondary metabolism in Sorangium cellulosum So ce90, an aromatic amino acid decarboxylase gene (ddc) was detected. Expression of ddc in Escherichia coli and subsequent enzyme assays with cell-free extracts confirmed the proposed function derived from amino acid sequence comparisons. In contrast to other aromatic amino acid decarboxylases of eukaryotic origin, the S. cellulosum Ddc converted only L-dihydroxy phenylalanine. This is the first report of a gene encoding an L-dihydroxy phenylalanine decarboxylase in bacteria.  相似文献   

15.
The lysA gene of Bacillus methanolicus MGA3 was cloned by complementation of an auxotrophic Escherichia coli lysA22 mutant with a genomic library of B. methanolicus MGA3 chromosomal DNA. Subcloning localized the B. methanolicus MGA3 lysA gene into a 2.3-kb SmaI-SstI fragment. Sequence analysis of the 2.3-kb fragment indicated an open reading frame encoding a protein of 48,223 Da, which was similar to the meso-diaminopimelate (DAP) decarboxylase amino acid sequences of Bacillus subtilis (62%) and Corynebacterium glutamicum (40%). Amino acid sequence analysis indicated several regions of conservation among bacterial DAP decarboxylases, eukaryotic ornithine decarboxylases, and arginine decarboxylases, suggesting a common structural arrangement for positioning of substrate and the cofactor pyridoxal 5'-phosphate. The B. methanolicus MGA3 DAP decarboxylase was shown to be a dimer (M(r) 86,000) with a subunit molecular mass of approximately 50,000 Da. This decarboxylase is inhibited by lysine (Ki = 0.93 mM) with a Km of 0.8 mM for DAP. The inhibition pattern suggests that the activity of this enzyme in lysine-overproducing strains of B. methanolicus MGA3 may limit lysine synthesis.  相似文献   

16.
The product of the ARO10 gene from Saccharomyces cerevisiae was initially identified as a thiamine diphosphate-dependent phenylpyruvate decarboxylase with a broad substrate specificity. It was suggested that the enzyme could be responsible for the catabolism of aromatic and branched-chain amino acids, as well as methionine. In the present study, we report the overexpression of the ARO10 gene product in Escherichia coli and the first detailed in vitro characterization of this enzyme. The enzyme is shown to be an efficient aromatic 2-keto acid decarboxylase, consistent with it playing a major in vivo role in phenylalanine, tryptophan and possibly also tyrosine catabolism. However, its substrate spectrum suggests that it is unlikely to play any significant role in the catabolism of the branched-chain amino acids or of methionine. A homology model was used to identify residues likely to be involved in substrate specificity. Site-directed mutagenesis on those residues confirmed previous studies indicating that mutation of single residues is unlikely to produce the immediate conversion of an aromatic into an aliphatic 2-keto acid decarboxylase. In addition, the enzyme was compared with the phenylpyruvate decarboxylase from Azospirillum brasilense and the indolepyruvate decarboxylase from Enterobacter cloacae. We show that the properties of the two phenylpyruvate decarboxylases are similar in some respects yet quite different in others, and that the properties of both are distinct from those of the indolepyruvate decarboxylase. Finally, we demonstrate that it is unlikely that replacement of a glutamic acid by leucine leads to discrimination between phenylpyruvate and indolepyruvate, although, in this case, it did lead to unexpected allosteric activation.  相似文献   

17.
草酸脱羧酶及其应用   总被引:2,自引:0,他引:2  
草酸脱羧酶是一种含锰的酶,在白腐菌中广泛存在,少数低等真菌和细菌中也能产生。目前,至少10多种草酸脱羧酶得到了分离和纯化。该酶是一种氨基酸残基在379个左右的单体酶,一般都为酸性糖蛋白,含有2个锰离子,形成2个活性区域;表面一些氨基酸被不同程度地糖基化。晶体结构和其它一些波谱学研究解释了其空间结构和可能的电子传递机制。运用PCR技术和cDNA文库技术,越来越多的草酸脱羧酶基因被克隆。已研究的该酶基因中都含有17个左右的内含子,这些内含子在活性域位置上有比较高的保守性。一些特殊氨基酸序列的存在决定了该酶的表达形式为诱导型,菌株的基因调控序列中含有一段受草酸化合物作用的序列。该酶在一些酵母和植物等异源表达系统中有成功表达的报道。该酶的应用主要集中在以下几方面:造纸废水中的草酸盐降解;食品中的草酸降解;草酸生物检测(如,临床诊断)等。  相似文献   

18.
Lactic acid bacteria play a pivotal role in many food fermentations and sometimes represent a health threat due to the ability of some strains to produce biogenic amines that accumulate in foods and cause trouble following ingestion. These strains carry specific enzymatic systems catalyzing the uptake of amino acid precursors (e.g., ornithine and lysine), the decarboxylation inside the cell, and the release of the resulting biogenic amines (e.g., putrescine and cadaverine). This study aimed to identify the system involved in production of cadaverine from lysine, which has not been described to date for lactic acid bacteria. Strain Lactobacillus saerimneri 30a (formerly called Lactobacillus sp. 30a) produces both putrescine and cadaverine. The sequencing of its genome showed that the previously described ornithine decarboxylase gene was not associated with the gene encoding an ornithine/putrescine exchanger as in other bacteria. A new hypothetical decarboxylation system was detected in the proximity of the ornithine decarboxylase gene. It consisted of two genes encoding a putative decarboxylase sharing sequence similarities with ornithine decarboxylases and a putative amino acid transporter resembling the ornithine/putrescine exchangers. The two decarboxylases were produced in Escherichia coli, purified, and characterized in vitro, whereas the transporter was heterologously expressed in Lactococcus lactis and functionally characterized in vivo. The overall data led to the conclusion that the two decarboxylases and the transporter form a three-component decarboxylation system, with the new decarboxylase being a specific lysine decarboxylase and the transporter catalyzing both lysine/cadaverine and ornithine/putrescine exchange. To our knowledge, this is an unprecedented observation of a bacterial three-component decarboxylation system.  相似文献   

19.
Biosynthetic ornithine decarboxylase was purified 4300-fold from Escherichia coli to a purity of approximately 85% as judged by polyacrylamide gel electrophoresis. The enzyme showed hyperbolic kinetics with a Km of 5.6 mM for ornithine and 1.0 micronM for pyridoxal phosphate and it was competitively inhibited by putrescine and spermidine. The biosynthetic decarboxylase was compared with the biodegradative ornithine decarboxylase [Applebaum, D., et al. (1975), Biochemistry 14, 3675]. Both enzymes were dimers of 80 000-82 000 molecular weight and exhibited similar kinetic properties. However, they differed significantly in other respects. The pH optimum of the biosynthetic enzyme was 8.1, compared with 6.9 for the biodegradative. Both enzymes were activated by nucleotides, but with different specificity. Antibody to the purified biodegradative ornithine decarboxylase did not cross-react with the biosynthetic enzyme. The evolutionary relationship of these two decarboxylases to the other amino acid decarboxylases of E. coli is discussed.  相似文献   

20.
Histidine decarboxylases from Klebsiella planticola and Enterobacter aerogenes were purified to homogeneity and compared with the histidine decarboxylase from Morganella morganii. All three enzymes required pyridoxal 5'-phosphate as a coenzyme, showed optimal activity at pH 6.5, decarboxylated only histidine among the amino acids derived from protein, and were tetramers or dimers of identical subunits. Amino-terminal sequences of the three enzymes showed up to 81% homology through residue 33, but the enzymes differed sufficiently in amino acid composition and sequence so that no cross-reaction occurred between the K. planticola or E. aerogenes enzymes and antibodies to the decarboxylase from M. morganii. All three enzymes were inhibited by carbonyl reagents; by amino-, carboxyl-, and some methyl-substituted histidines; and by alpha-fluoromethylhistidine. These decarboxylases, all from gram-negative organisms, differed greatly in subunit structure, biogenesis, and other properties from the pyruvoyl-dependent histidine decarboxylases from gram-positive organisms described previously.  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