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 31 毫秒
1.
The function of the mlr6791 gene from Mesorhizobium loti MAFF303099 has been identified. This gene encodes 3-hydroxy-2-methylpyridine-4,5-dicarboxylate decarboxylase (HMPDdc), an enzyme involved in the catabolism of pyridoxal 5'-phosphate (Vitamin B6). This enzyme was overexpressed in Escherichia coli and characterized. HMPDdc is a 26 kDa protein that catalyzes the decarboxylation of 3-hydroxy-2-methylpyridine-4,5-dicarboxylate to 3-hydroxy-2-methylpyridine-5-carboxylate. The KM and kcat were found to be 366 microM and 0.6 s-1, respectively. The structure of this enzyme was determined at 1.9 A resolution using SAD phasing and belongs to the class II aldolase/adducin superfamily. While the decarboxylation of hydroxy-substituted benzene rings is a common motif in biosynthesis, the mechanism of this reaction is still poorly characterized. The structural studies described here suggest that catalysis of such decarboxylations proceeds by an aldolase-like mechanism.  相似文献   

2.
Z Huang  L Dostal    J P Rosazza 《Journal of bacteriology》1994,176(19):5912-5918
A ferulic acid decarboxylase enzyme which catalyzes the decarboxylation of ferulic acid to 4-hydroxy-3-methoxystyrene was purified from Pseudomonas fluorescens UI 670. The enzyme requires no cofactors and contains no prosthetic groups. Gel filtration estimated an apparent molecular mass of 40.4 (+/- 6%) kDa, whereas sodium dodecyl sulfate-polyacrylamide gel electrophoresis showed a molecular mass of 20.4 kDa, indicating that ferulic acid decarboxylase is a homodimer in solution. The purified enzyme displayed an optimum temperature range of 27 to 30 degrees C, exhibited an optimum pH of 7.3 in potassium phosphate buffer, and had a Km of 7.9 mM for ferulic acid. This enzyme also decarboxylated 4-hydroxycinnamic acid but not 2- or 3-hydroxycinnamic acid, indicating that a hydroxy group para to the carboxylic acid-containing side chain is required for the enzymatic reaction. The enzyme was inactivated by Hg2+, Cu2+, p-chloromercuribenzoic acid, and N-ethylmaleimide, suggesting that sulfhydryl groups are necessary for enzyme activity. Diethyl pyrocarbonate, a histidine-specific inhibitor, did not affect enzyme activity.  相似文献   

3.
J G Straka  J P Kushner 《Biochemistry》1983,22(20):4664-4672
Uroporphyrinogen decarboxylase (EC 4.1.1.37) has been purified to homogeneity from bovine liver by using isoelectric and salt precipitations, followed by chromatography on DEAE-cellulose, phenyl-Sepharose, hydroxylapatite, and Sephacryl S-200. The purified enzyme is a monomer with an Mr approximately 57 000 and an isoelectric point at pH 4.6. Enzyme activity is optimal in buffers having an ionic strength of approximately 0.1 M and a pH of 6.8. The purified enzyme has a specific activity (expressed as the disappearance of uroporphyrinogen I) of 936 nmol X h-1 X (mg of protein)-1. The purified enzyme catalyzes all four decarboxylation reactions in the conversion of uroporphyrinogen I or III to the corresponding coproporphyrinogen. The rate-limiting step in the physiologically significant conversion of uroporphyrinogen III to coproporphyrinogen III is the decarboxylation of heptacarboxylate III. Kinetic data suggest that the enzyme has at least two noninteracting active sites. At least one sulfhydryl group is required for catalytic activity. The enzyme is inhibited by sulfhydryl-specific reagents and by divalent metal ions including Fe2+, Co2+, Cu2+, Zn2+, and Pb2+. The pattern of accumulation of intermediate (hepta-, hexa-, and pentacarboxylate porphyrinogens) and final (coproporphyrinogen) decarboxylation products is affected by the ratio of substrate (uroporphyrinogen I or III) concentration to enzyme concentration. Under physiologic conditions where the uroporphyrinogen to enzyme ratio is low, the substrate is nearly quantitatively decarboxylated, and the major product is coproporphyrinogen. If the ratio of uroporphyrinogen to enzyme is high, intermediates accumulate, and heptacarboxylate porphyrinogen becomes the major decarboxylation product.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Z He  J Wiegel 《Journal of bacteriology》1996,178(12):3539-3543
A 3,4-dihydroxybenzoate decarboxylase (EC 4.1.1.63) from Clostridium hydroxybenzoicum JW/Z-1T was purified and partially characterized. The estimated molecular mass of the enzyme was 270 kDa. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis gave a single band of 57 kDa, suggesting that the enzyme consists of five identical subunits. The temperature and pH optima were 50 degrees C and pH 7.0, respectively. The Arrhenius energy for decarboxylation of 3,4-dihydroxybenzoate was 32.5 kJ . mol(-1) for the temperature range from 22 to 50 degrees C. The Km and kcat for 3,4-dihydroxybenzoate were 0.6 mM and 5.4 x 10(3) min(-1), respectively, at pH 7.0 and 25 degrees C. The enzyme optimally catalyzed the reverse reaction, that is, the carboxylation of catechol to 3,4-dihydroxybenzoate, at pH 7.0. The enzyme did not decarboxylate 2-hydroxybenzoate, 3-hydroxybenzoate, 4-hydroxybenzoate, 2,3-dihydroxybenzoate, 2,4-dihydroxybenzoate, 2,5-dihydroxybenzoate, 2,3,4-trihydroxybenzoate, 3,4,5-trihydroxybenzoate, 3-F-4-hydroxybenzoate, or vanillate. The decarboxylase activity was inhibited by 25 and 20%, respectively, by 2,3,4- and 3,4,5-trihydroxybenzoate. Thiamine PPi and pyridoxal 5'-phosphate did not stimulate and hydroxylamine and sodium borohydride did not inhibit the enzyme activity, indicating that the 3,4-dihydroxybenzoate decarboxylase is not a thiamine PPi-, pyridoxal 5'-phosphate-, or pyruvoyl-dependent enzyme.  相似文献   

5.
A new enzyme, L-tartrate decarboxylase, was found in cells of Pseudomonas sp. group Ve-2. The enzyme was purified to homogeneity and characterized. The enzyme requires K+, Mg2+, and NAD+ for L-tartrate decarboxylation. The dependence of the enzymatic decarboxylation on NAD+ suggests that the decarboxylation involves redox reactions of the substrate. The enzyme catalyzes NAD(+)-linked oxidative decarboxylation of D-malate as well. The enzyme is composed of four subunits with identical molecular weight (Mr 40,000). The apparent Michaelis constants for L-tartrate and NAD+ are 1.1 mM, respectively. The cofactor requirements and the physical properties of the enzyme were similar to those of L-tartrate dehydrogenase-D-malate dehydrogenase from Rhodopseudomonas sphaeroides, and tartrate dehydrogenase from P. putida.  相似文献   

6.
Abstract A new enzyme activity, which catalyzes decarboxylation of l -2,4-diaminobutyric acid (DABA) to yield 1,3-diaminopropane (DAP), has been found in dialyzed crude extracts prepared from Vibrio alginolyticus . The pH optimum for the activity was 8.0–8.5, and the enzyme showed a pyridoxal 5'-phosphate (PLP) requirement. Mg2+ caused about 30% stimulation in activity. The enzyme was active to only l -DABA among the diamino acids examined, and the K m value for l -DABA was 0.13 mM. Ammonium sulfate fractionation of a dialyzed crude extract followed by HPLC separation allowed us to conclude that this enzyme differed from the decarboxylase which occurs in Vibrio spp. to produce norspermidine (Nspd) for carboxynorspermidine (C-Nspd) having a moiety similar in structure to DABA. The same enzyme activity was detected in several other Vibrio species.  相似文献   

7.
Phosphopantothenoylcysteine synthase catalyzes the formation of (R)-4'-phospho-N-pantothenoylcysteine from 4'-phosphopantothenate and l-cysteine: this enzyme, involved in the biosynthesis of coenzyme A (CoA), has not previously been identified. Recently it was shown that the NH(2)-terminal domain of the Dfp protein from bacteria catalyzes the next step in CoA biosynthesis, the decarboxylation of (R)-4'-phospho-N-pantothenoylcysteine to form 4'-phosphopantetheine (Kupke, T., Uebele, M., Schmid, D., Jung, G., Blaesse, M., and Steinbacher, S. (2000) J. Biol. Chem. 275, 31838-31846). We have partially purified phosphopantothenoylcysteine decarboxylase from Escherichia coli and demonstrated that the protein encoded by the dfp gene, here renamed coaBC, also has phosphopantothenoylcysteine synthetase activity, using CTP rather than ATP as the activating nucleoside 5'-triphosphate. This discovery completes the identification of all the enzymes involved in the biosynthesis of coenzyme A in bacteria.  相似文献   

8.
Carthamin, a red quinochalcone pigment in safflower (Carthamus tinctorius L.), is enzymatically converted from a yellow precursor, precarthamin. The enzyme, which catalyzes the oxidative decarboxylation of precarthamin to carthamin, was purified to apparent homogeneity from yellow petals of safflower and named precarthamin decarboxylase. The molecular mass of the denatured enzyme was estimated as 33 kDa by SDS-PAGE. The molecular mass of the native enzyme was determined by gel filtration chromatography to be 24 kDa; thus, the native enzyme is a monomer. The optimum pH of the enzyme was 5.0. The enzyme activity was inhibited by Mn2+, Fe2+, and Cu2+ and sharply decreased at temperatures higher than 50 degrees C for 10 min. The activation energy and the Arrhenius frequency factor of the enzyme reaction were 19.7 kcal mol(-1) and 9.94 x 10(11) s(-1), respectively. The saturation curve of precarthamin showed that the enzyme follows Michaelis-Menten kinetics. The Km and Vmax of the enzyme were calculated as 164 microM and 29.2 nmol/ min, respectively. The turnover number (kcat) of the enzyme was calculated as 1.42 x 10(2) s(-1). The enzyme activity was severely inhibited by reducing agents such as glutathione and DTT at pH 5.0, suggesting that a disulfide bond may play an important role in enzyme function.  相似文献   

9.
In the conversion of quinolinic acid to 6-hydroxypicolinic acid by whole cells of Alcaligenes sp. strain UK21, the enzyme reactions involved in the hydroxylation and decarboxylation of quinolinic acid were examined. Quinolinate dehydrogenase, which catalyzes the first step, the hydroxylation of quinolinic acid, was solubilized from a membrane fraction, partially purified, and characterized. The enzyme catalyzed the incorporation of oxygen atoms of H2O into the hydroxyl group. The dehydrogenase hydroxylated quinolinic acid and pyrazine-2,3-dicarboxylic acid to form 6-hydroxyquinolinic acid and 5-hydroxypyrazine-2,3-dicarboxylic acid, respectively. Phenazine methosulfate was the preferred electron acceptor for quinolinate dehydrogenase. 6-Hydroxyquinolinate decarboxylase, catalyzing the nonoxidative decarboxylation of 6-hydroxyquinolinic acid, was purified to homogeneity and characterized. The purified enzyme had a molecular mass of approximately 221 kDa and consisted of six identical subunits. The decarboxylase specifically catalyzed the decarboxylation of 6-hydroxyquinolinic acid to 6-hydroxypicolinic acid, without any co-factors. The N-terminal amino acid sequence was homologous with those of bacterial 4,5-dihydroxyphthalate decarboxylases.  相似文献   

10.
After enrichment culture with indole-3-carboxylate in static culture, a novel reversible decarboxylase, indole-3-carboxylate decarboxylase, was found in Arthrobacter nicotianae FI1612 and several molds. The enzyme reaction was examined in resting-cell reactions with A. nicotianae FI1612. The enzyme activity was induced specifically by indole-3-carboxylate, but not by indole. The indole-3-carboxylate decarboxylase of A. nicotianae FI1612 catalyzed the nonoxidative decarboxylation of indole-3-carboxylate into indole, and efficiently carboxylated indole and 2-methylindole by the reverse reaction. In the presence of 1 mM dithiothreitol, 50 mM Na2 S2O3, and 20% (v/v) glycerol, indole-3-carboxylate decarboxylase was partially purified from A. nicotianae FI1612. The purified enzyme had a molecular mass of approximately 258 kDa. The enzyme did not need any cofactor for the decarboxylating and carboxylating reactions.  相似文献   

11.
Mevalonate diphosphate decarboxylase (MVD) is an ATP-dependent enzyme that catalyzes the phosphorylation/decarboxylation of (R)-mevalonate-5-diphosphate to isopentenyl pyrophosphate in the mevalonate (MVA) pathway. MVD is a key enzyme in engineered metabolic pathways for bioproduction of isobutene, since it catalyzes the conversion of 3-hydroxyisovalerate (3-HIV) to isobutene, an important platform chemical. The putative homologue from Picrophilus torridus has been identified as a highly efficient variant in a number of patents, but its detailed characterization has not been reported. In this study, we have successfully purified and characterized the putative MVD from P. torridus. We discovered that it is not a decarboxylase per se but an ATP-dependent enzyme, mevalonate-3-kinase (M3K), which catalyzes the phosphorylation of MVA to mevalonate-3-phosphate. The enzyme''s potential in isobutene formation is due to the conversion of 3-HIV to an unstable 3-phosphate intermediate that undergoes consequent spontaneous decarboxylation to form isobutene. Isobutene production rates were as high as 507 pmol min−1 g cells−1 using Escherichia coli cells expressing the enzyme and 2,880 pmol min−1 mg protein−1 with the purified histidine-tagged enzyme, significantly higher than reported previously. M3K is a key enzyme of the novel MVA pathway discovered very recently in Thermoplasma acidophilum. We suggest that P. torridus metabolizes MVA by the same pathway.  相似文献   

12.
Phosphatidylserine decarboxylase, an intrinsic membrane protein of Escherichia coli, catalyzes the decarboxylation of phosphatidylserine, the final step in the biosynthesis of phosphatidylethanolamine, the principal membrane lipid of this organism. The purified enzyme lacks the absorption spectrum characteristic of pyridoxal-containing enzymes, and it has now been found to contain bound pyruvate, the carbonyl function of which is essential for catalytic activity. The decarboxylase is inactivated by treatment with a number of reagents that attack carbonyl groups, including sodium borohydride. Reduction with tritiated borohydride leads to the introduction of stably bound radioactivity, which, after acid hydrolysis, has been identified as tritiated lactate by several chromatographic procedures and by treatment with lactate dehydrogenase. The enzyme resists inactivation by cyanoborohydride in the absence of substrate, but is readily inactivated by this reagent in the presence of phosphatidylserine. Under the conditions of treatment of neutral pH, cyanoborohydride does not react with carbonyl residues at an appreciable rate, but reduces imino groups much more rapidly. This finding, together with demonstrated dependence of the enzyme upon the carbonyl residue of pyruvate for activity, strongly suggests that a Schiff base is formed by addition of the amino group of phosphatidylserine to the pyruvate residue of the enzyme as an essential step in the action of the decarboxylase.  相似文献   

13.
Xia C  Watton S  Nagl S  Samuel J  Lovegrove J  Cheshire J  Woo P 《FEBS letters》2004,570(1-3):217-222
The citM gene from Lactococcus lactis CRL264 was demonstrated to encode for an oxaloacetate decarboxylase. The enzyme exhibits high levels of similarity to malic enzymes (MEs) from other organisms. CitM was expressed in Escherichia coli, purified and its oxaloacetate decarboxylase activity was demonstrated by biochemical and genetic studies. The highest oxaloacetate decarboxylation activity was found at low pH in the presence of manganese, and the Km value for oxaloacetate was 0.52 ± 0.03 mM. However, no malic activity was found for this enzyme. Our studies clearly show a new group of oxaloacetate decarboxylases associated with the citrate fermentation pathway in gram-positive bacteria. Furthermore, the essential catalytic residues were found to be conserved in all members of the ME family, suggesting a common mechanism for oxaloacetate decarboxylation.  相似文献   

14.
The Bacteroides fragilis capsular polysaccharide complex is the major virulence factor for abscess formation in human hosts. Polysaccharide B of this complex contains a 2-aminoethylphosphonate functional group. This functional group is synthesized in three steps, one of which is catalyzed by phosphonopyruvate decarboxylase. In this paper, we report the cloning and overexpression of the B. fragilis phosphonopyruvate decarboxylase gene (aepY), purification of the phosphonopyruvate decarboxylase recombinant protein, and the extensive characterization of the reaction that it catalyzes. The homotrimeric (41,184-Da subunit) phosphonopyruvate decarboxylase catalyzes (kcat = 10.2 +/- 0.3 s-1) the decarboxylation of phosphonopyruvate (Km = 3.2 +/- 0.2 microm) to phosphonoacetaldehyde (Ki = 15 +/- 2 microm) and carbon dioxide at an optimal pH range of 7.0-7.5. Thiamine pyrophosphate (Km = 13 +/- 2 microm) and certain divalent metal ions (Mg(II) Km = 82 +/- 8 microm; Mn(II) Km = 13 +/- 1 microm; Ca(II) Km = 78 +/- 6 microm) serve as cofactors. Phosphonopyruvate decarboxylase is a member of the alpha-ketodecarboxylase family that includes sulfopyruvate decarboxylase, acetohydroxy acid synthase/acetolactate synthase, benzoylformate decarboxylase, glyoxylate carboligase, indole pyruvate decarboxylase, pyruvate decarboxylase, the acetyl phosphate-producing pyruvate oxidase, and the acetate-producing pyruvate oxidase. The Mg(II) binding residue Asp-260, which is located within the thiamine pyrophosphate binding motif of the alpha-ketodecarboxylase family, was shown by site-directed mutagenesis to play an important role in catalysis. Pyruvate (kcat = 0.05 s-1, Km = 25 mm) and sulfopyruvate (kcat approximately 0.05 s-1; Ki = 200 +/- 20 microm) are slow substrates for the phosphonopyruvate decarboxylase, indicating that this enzyme is promiscuous.  相似文献   

15.
Oxaloacetate decarboxylase (OXAD), the enzyme that catalyzes the decarboxylation of oxaloacetate to pyruvic acid and carbon dioxide, was purified 245-fold to homogeneity from Pseudomonas stutzeri. The three-step purification procedure comprised anion-exchange chromatography, metal-chelate affinity chromatography, and biomimetic-dye affinity chromatography. Estimates of molecular mass from sodium dodecyl sulfate-polyacrylamide gel electrophoresis and native high-performance gel-filtration liquid chromatography were, respectively, 63 and 64 kDa, suggesting a monomeric protein. OXAD required for maximum activity divalent metal cations such as Mn2+ and Mg2+ but not monovalent cations. The enzyme is not inhibited by avidin, but is competitively inhibited by adenosine 5'-diphosphate, acetic acid, phosphoenolpyruvate, malic acid, and oxalic acid. Initial velocity, product inhibition, and dead-end inhibition studies suggested a rapid-equilibrium ordered kinetic mechanism with Mn2+ being added to the enzyme first followed by oxaloacetate, and carbon dioxide is released first followed by pyruvate. Inhibition data as well as pH-dependence profiles and kinetic parameters are reported and discussed in terms of the mechanism operating for oxaloacetate decarboxylation.  相似文献   

16.
Indolepyruvate decarboxylase, a key enzyme for indole-3-acetic acid biosynthesis, was found in extracts of Enterobacter cloacae. The enzyme catalyzes the decarboxylation of indole-3-pyruvic acid to yield indole-3-acetaldehyde and carbon dioxide. The enzyme was purified to apparent homogeneity from Escherichia coli cells harboring the genetic locus for this enzyme obtained from E. cloacae. The results of gel filtration experiments showed that indolepyruvate decarboxylase is a tetramer with an M(r) of 240,000. In the absence of thiamine pyrophosphate and Mg2+, the active tetramers dissociate into inactive monomers and dimers. However, the addition of thiamine pyrophosphate and Mg2+ to the inactive monomers and dimers results in the formation of active tetramers. These results indicate that the thiamine pyrophosphate-Mg2+ complex functions in the formation of the tetramer, which is the enzymatically active holoenzyme. The enzyme exhibited decarboxylase activity with indole-3-pyruvic acid and pyruvic acid as substrates, but no decarboxylase activity was apparent with L-tryptophan, indole-3-lactic acid, beta-phenylpyruvic acid, oxalic acid, oxaloacetic acid, and acetoacetic acid. The Km values for indole-3-pyruvic acid and pyruvic acid were 15 microM and 2.5 mM, respectively. These results indicate that indole-3-acetic acid biosynthesis in E. cloacae is mediated by indolepyruvate decarboxylase, which has a high specificity and affinity for indole-3-pyruvic acid.  相似文献   

17.
Highly purified pyruvic decarboxylase (EC 4.1.1.1) from wheat germ catalyses the decarboxylation of hydroxypyruvate. A kinetic analysis of the activity of the enzyme with pyruvate and hydroxypyruvate as substrates suggests that a single enzyme is involved. The kinetics of decarboxylation are autocatalytic. The time lag before maximum activity is reached is affected by the concentration of hydroxypyruvate and the pH. The question whether or not hydroxypyruvate is a natural substrate for the enzyme remains unresolved, but it may be significant that at physiological pH (ca 7.5) the enzyme shows optimum activity with hydroxypyruvate, but negligible activity with pyruvate.  相似文献   

18.
Cations in component reactions of `malic'' enzyme catalysis   总被引:1,自引:1,他引:0       下载免费PDF全文
The ;malic' enzyme (EC 1.1.1.40) has been purified (300-fold) from wheat germ and its abilities to catalyse the decarboxylation and the hydrogenation of oxaloacetic acid and oxaloacetate esters was studied. The free 1-carboxyl group is essential for the interaction of oxaloacetates and substituted oxaloacetates with the enzyme via cations. The free 4-carboxyl group is required for the decarboxylation but is not indispensable for the hydrogenation. At high concentrations, cations inhibit the enzymic hydrogenation of oxaloacetic acid but not that of 4-ethyl oxaloacetate. A plausible inhibitory mechanism is proposed.  相似文献   

19.
S K Ng  M Wong    I R Hamilton 《Journal of bacteriology》1982,150(3):1252-1258
Oxaloacetate decarboxylase was purified to 136-fold from the oral anaerobe Veillonella parvula. The purified enzyme was substantially free of contaminating enzymes or proteins. Maximum activity of the enzyme was exhibited at pH 7.0 for both carboxylation and decarboxylation. At this pH, the Km values for oxaloacetate and Mg2+ were at 0.06 and 0.17 mM, respectively, whereas the Km values for pyruvate, CO2, and Mg2+ were 3.3, 1.74, and 1.85 mM, respectively. Hyperbolic kinetics were observed with all of the aforementioned compounds. The Keq' was 2.13 X 10(-3) mM-1 favoring the decarboxylation of oxaloacetate. In the carboxylation step, avidin, acetyl coenzyme A, biotin, and coenzyme A were not required. ADP and NADH had no effect on either the carboxylation or decarboxylation step, but ATP inhibited the carboxylation step competitively and the decarboxylation step noncompetitively. These types of inhibition fitted well with the overall lactate metabolism of the non-carbohydrate-fermenting anaerobe.  相似文献   

20.
A nonoxidative decarboxylase, 2,6-dihydroxybenzoate decarboxylase, was found in Agrobacterium tumefaciens IAM12048. The enzyme activity was induced specifically by 2,6-dihydroxybenzoate. The purified enzyme was a homotetramer of identical 38 kDa subunits. The purified decarboxylase catalyzed the nonoxidative decarboxylation of 2,6-dihydroxybenzoate and 2,3-dihydroxybenzoate without requiring any cofactors. In the presence of KHCO3, the enzyme also catalyzed the regioselective carboxylation of 1,3-dihydroxybenzene into 2,6-dihydroxybenzoate at a molar conversion ratio of 30%.  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