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.
Abstract Crystalline 3-methylaspartase (EC 4.3.1.2) from Escherichia coli strain YG1002 that had been isolated from soil was characterized. The enzyme activity was induced when the organism was grown statically on medium containing ( S )-glutamic acid. Its molecular mass is about 84 kDa, and it may be composed of two identical subunits of 42 kDa. The enzyme requires both divalent and monovalent cations such as Mg2+ and K+, respectively. The enzyme catalyzes reversible amination-deamination between mesaconic acid and (2 S ,3 S )-methylaspartic acid, which is the best substrate.  相似文献   

2.
N P Botting  M A Cohen  M Akhtar  D Gani 《Biochemistry》1988,27(8):2956-2959
3-Methylaspartate ammonia-lyase catalyzes the deamination of (2S)-aspartic acid 137 times more slowly than the deamination of (2S,3S)-3-methylaspartic acid but catalyzes the amination of fumaric acid 1.8 times faster than the amination of mesaconic acid [Botting, N.P., Akhtar, M., Cohen, M. A., & Gani, D. (1988) Biochemistry (preceding paper in this issue)]. In order to understand the mechanistic basis for these observations, the deamination reaction was examined kinetically with (2S)-aspartic acid, (2S,3S)-3-methylaspartic acid, (2S,3S)-3-ethylaspartic acid, and the corresponding C-3-deuteriated isotopomers. Comparison of the double-reciprocal plots of the initial reaction velocities for each of the three pairs of substrates revealed that the magnitude of the primary isotope effect on both Vmax and V/K varied with the substituent at C-3 of the substrate. 3-Methylaspartic acid showed the largest isotope effect (1.7 on Vmax and V/K), 3-ethylaspartic acid showed a smaller isotope effect (1.2 on Vmax and V/K), and aspartic acid showed no primary isotope effect at all. These results, which are inconsistent with earlier reports that there is no primary isotope effect for 3-methylaspartic acid [Bright, H. J. (1964) J. Biol. Chem. 239, 2307], suggest that for both 3-methylaspartic acid and 3-ethylaspartic acid elimination occurs via a predominantly concerted mechanism whereas for aspartic acid an E1cb mechanism prevails.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

3.
Methylaspartate ammonia lyase (MAL; EC 4.3.1.2) catalyzes the reversible addition of ammonia to mesaconate to give (2S,3S)-3-methylaspartate and (2S,3R)-3-methylaspartate as products. MAL is of considerable biocatalytic interest because of its potential use for the asymmetric synthesis of substituted aspartic acids, which are important building blocks for synthetic enzymes, peptides, chemicals, and pharmaceuticals. Here, we have cloned the gene encoding MAL from the thermophilic bacterium Carboxydothermus hydrogenoformans Z-2901. The enzyme (named Ch-MAL) was overproduced in Escherichia coli and purified to homogeneity by immobilized metal affinity chromatography. Ch-MAL is a dimer in solution, consisting of two identical subunits (∼49 kDa each), and requires Mg2+ and K+ ions for maximum activity. The optimum pH and temperature for the deamination of (2S,3S)-3-methylaspartic acid are 9.0 and 70°C (k cat = 78 s−1 and K m = 16 mM). Heat inactivation assays showed that Ch-MAL is stable at 50°C for >4 h, which is the highest thermal stability observed among known MALs. Ch-MAL accepts fumarate, mesaconate, ethylfumarate, and propylfumarate as substrates in the ammonia addition reaction. The enzyme also processes methylamine, ethylamine, hydrazine, hydroxylamine, and methoxylamine as nucleophiles that can replace ammonia in the addition to mesaconate, resulting in the corresponding N-substituted methylaspartic acids with excellent diastereomeric excess (>98% de). This newly identified thermostable MAL appears to be a potentially attractive biocatalyst for the stereoselective synthesis of aspartic acid derivatives on large (industrial) scale.  相似文献   

4.
Structure and Function of Amino Acid Ammonia-lyases   总被引:1,自引:0,他引:1  
Histidine ammonia-lyase (HAL) and methylaspartate ammonia-lyase (MAL) belong to the family of carbon-nitrogen lyases (EC 4.3.1). The enzymes catalyze the α,β-elimination of ammonia from (S)-His to yield urocanic acid, and (S)-threo-(2S,3S)-3-methylaspartic acid to mesaconic acid, respectively. Based on structural analyses, the peptide at the active center of HAL from Pseudomonas putida is considered to be post-translationally dehydrated to form an electrophilic 4-methylidene-imidazole-one (MIO) group. A reaction mechanism was proposed with the structure. On the other hand, the structure of MAL from Citrobacter amalonaticus was found to be a typical TIM barrel structure with Mg2+ coordinated to the 4-carbonyl of the substrate methylaspartate. Unlike HAL, MIO was not observed in MAL, and the reaction of MAL appears to be completely different from phenylalanine ammonia-lyase (PAL), HAL, and other amino acid ammonia-lyases. A reaction mechanism is proposed in which the hydrogen at the β to the amino group of the substrate is abstracted forming an enolate type intermediate and then ammonia is released.  相似文献   

5.
The nicotinamide adenine dinucleotide phosphate (NADP)-dependent glutamate dehydrogenase (NADP-GDH) of Agaricus bisporus, a key enzyme in ammonia assimilation, was purified to apparent electrophoretic homogeneity with 27% recovery of the initial activity. The molecular weight of the native enzyme was 330 kDa. The enzyme is probably a hexamer, composed of identical subunits of 48 kDa. The isoelectric point of the enzyme was found at pH 4.8. The N-terminus appeared to be blocked. The enzyme was specific for NADP(H). The Km-values were 2.1, 3.2, 0.074, 27.0, and 0.117mM for ammonia, 2-oxoglutarate, NADPH, L-glutamate, and NADP respectively. The pH optima for the amination and deamination reactions were found to be 7.6 and 9.0, respectively. The temperature optimum was 33°C. The effect of several metabolites on the enzyme's activity was tested. Pyruvate, oxaloacetate, ADP, and ATP showed some inhibitory effect. Divalent cations slightly stimulated the aminating reaction. Antibodies raised against the purified enzyme were able to precipitate NADP-GDH activity from a cell-free extract in an anticatalytic immunoprecipitation test. Analysis of a Western blot showed the antibodies to be specific for NADP-GDH.  相似文献   

6.
We have partially purified S-adenosylmethionine decarboxylase (EC 4.1.1.50, SAMDC) from carnation (Dianthus caryophyllus L.) petals and generated polyclonal antibodies against CSDC 16 protein (Leeet al., 1996) overexpressed inE. coli. The protein has been purified approximately 126.8 fold through the steps involving ammonium sulfate fractionation, DEAE-Sepharose column chromatography and Sephacryl S-300 gel filtration. Its molecular mass was 42 kDa in native form and we could also detect a band of 32 kDa molecular mass on SDS-PAGE in western blot analysis using the polyclonal antibodies. The Km value of this enzyme forS-adenosylmethionine was 26.3 μM. The optimum temperature and pH forS-adenosylmethionine decarboxylase activity were 35°C and pH 8.0, respectively. Putrescine and Mg2+ had no effects on the activation of the enzyme activity. Mg2+ did not have any significant effects on the enzyme activity. SAMDC activity was inhibited by putrescine, spermidine and spermine. Methylglyoxal bis-(guanylhydrazone) (MGBG), carbonyl reagents such as hydroxylamine and phenylhydrazine, and sulfhydryl reagent such as 5,5′dithio-bis (2-nitrobenzoic acid) (DTNB) were effective inhibitors of the enzyme. However, isonicotinic acid hydrazide known as an inhibitor of 5′-pyridoxal phosphate (PLP) dependent enzyme activity had no significant effect on the enzyme activity. These results and our previously reported results (Leeet al., 1997b) suggest thatS-adenosylmethionine decarboxylase is a heterodimer, αβ, and some carbonyl group and sulfhydryl group are involved in the catalytic activity.  相似文献   

7.
β-Galactosidase was isolated from the cell-free extracts ofLactobacillus crispatus strain ATCC 33820 and the effects of temperature, pH, sugars and monovalent and divalent cations on the activity of the enzyme were examined.L. crispatus produced the maximum amount of enzyme when grown in MRS medium containing galactose (as carbon source) at 37°C and pH 6.5 for 2 d, addition of glucose repressing enzyme production. Addition of lactose to the growth medium containing galactose inhibited the enzyme synthesis. The enzyme was active between 20 and 60°C and in the pH range of 4–9. However, the optimum enzyme activity was at 45°C and pH 6.5. The enzyme was stable up to 45°C when incubated at various temperatures for 15 min at pH 6.5. When the enzyme was exposed to various pH values at 45°C for 1 h, it retained the original activity over the pH range of 6.0–7.0. Presence of divalent cations, such as Fe2+ and Mn2+, in the reaction mixture increased enzyme activity, whereas Zn2+ was inhibitory. TheK m was 1.16 mmol/L for 2-nitrophenyl-β-d-galactopyranose and 14.2 mmol/L for lactose.  相似文献   

8.
1,4-Benzoquinone reductase was purified to electrophoretic homogeneity from bovine liver, and the purified enzyme found to have a molecular mass of 29 kDa, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme exhibited pH optimum between 8.0 and 8.5. The apparentK m for 1,4-benzoquinone was 1.643 mM, and the apparent Km for NADH was 1.837 mM. Various divalent cations, such as Hg2+, Cu2+, and Zn2+, exhibited strong inhibitory effects. The enzyme activity was also strongly inhibited by quercetin, dicumarol, and benzoic acid. Incubation of the enzyme withN-bromosuccinimide and pyridoxal 5′-phosphate led to inhibitions of 100% and 99%, respectively. Accordingly, these results suggest that tryptophan and lysine residues are involved at or near the active sites of the enzyme.  相似文献   

9.
Purified preparations of NAD(H)‐glutamate dehydrogenase (GDH, EC 1.4.1.2.) were assayed to determine the effects of mono‐ and divalent cations, nucleotides and select carbon compounds on NAD(H)‐dependent GDH activity. The amination reaction was stimulated 2‐ to 17‐fold by divalent cations (Ca2+ > Cd2+ > Co2+ > Mg2+ > Mn2+ > Zn2+ between 1 and 1000 µ M ), but the reaction was unaffected by monovalent cations (Na + and K +). The amination reaction was most responsive to changes in Ca2+ in a NADH‐dependent manner. The addition of EDTA or EGTA nullified the stimulatory effects of Ca2+. Calmodulin alone or in combination with calmodulin antagonists did not affect the amination reaction. Divalent cations (at 1 m M ) inhibited the rate of the deamination reaction by 15 to 25%, while monovalent cations had no effect. ATP inhibited the amination reaction by 10 to 60%, while ADP had little or no effect. ATP or ADP decreased the rate of the deamination reaction 23 to 60 or 20 to 38%, respectively. Many tricarboxylic acid cycle intermediates inhibited the amination reaction, 20 to 50% of the inhibition could be attributed to the chelating capacity of intermediates. Conversely, most of the carbon sources tested did not affect the deamination reaction, the only appreciable differences were increases in activity with sucrose (21%) and glucose (41%) and a decrease in activity with pyruvate (34%). Inhibitors of sulfhydryl groups were used to examine the importance of reduced thiol groups in the amination or deamination reactions. The amination was not dependent on reduced thiol groups, whereas the deamination reaction was dependent on reduced thiol groups.  相似文献   

10.
S K Goda  N P Minton  N P Botting  D Gani 《Biochemistry》1992,31(44):10747-10756
The gene encoding methylaspartase (EC 4.3.1.2) from Clostridium tetranomorphum has been cloned, sequenced, and expressed in Escherichia coli. The open reading frame (ORF) codes for a polypeptide of 413 amino acid residues (M(r) 45,539) of which seven are cysteine residues. The size of the ORF indicates that methylaspartase is a homodimer rather than an (AB)2 tetramer. The deduced primary structure of the protein shows no homology to enzymes that catalyze similar reactions or, indeed, any convincing homology with any other characterized protein. The recombinant protein is identical to the enzyme isolated directly from C. tetanomorphum as determined by several criteria. The enzyme is obtained in a highly active form (approximately 70% of the activity of the natural enzyme) and migrates as a single band (M(r) 49,000) in SDS-polyacrylamide gels. The kinetic parameters for the deamination of (2S,3S)-3-methylaspartic acid by the natural and recombinant proteins are very similar, and the proteins display identical potassium ion-dependent primary deuterium isotope effects for V and V/K when (2S,3S)-3-methylaspartic acid is employed as the substrate. In accord with the activity of the natural enzyme, the recombinant protein is able to catalyze the slow formation of (2S,3R)-3-methylaspartic acid, the L-erythro-epimer of the natural substrate, from mesaconic acid and ammonia. Earlier work in which the cysteine residues in the protein were labeled with N-ethylmaleimide had indicated that there were eight cysteine residues per protein monomer. One cysteine residue was protected by substrate. Here evidence is forwarded to suggest that the residue that was protected by the substrate is not a cysteine residue but the translation product of a serine codon. Kinetic data indicate that this serine residue may be modified in the active enzyme. The implications of these findings on the mechanism of catalysis are discussed within the context of a few emerging mode of action for methylaspartate ammonia-lyase.  相似文献   

11.
N-acetyltransferase from Aeromonas hydrophilia was purified by ultrafiltration, DEAE-Sephacel, gel filtration chromatography on Sephadex G-100, and DEAE-5pw on high performance liquid chromatography, as judged by sodium dodecyl sulfate-polyacrylamine gel electrophoresis (SDS-PAGE) on a 12.% (wt/vol) slab gel. The enzyme had a molecular mass 44.9 kDa. The purified enzyme was thermostable at 37°C for 1 h with a half-life 28 min at 37°C, and displayed optimum activity at 37°C and pH 7.0. The K m and V max values for 2-aminofluorene were determined to be 0.896 mM and 2.456 nmol/min/mg protein, respectively. Among a series of divalent cations and salts, Zn2+, Ca2+, and Fe2+ were demonstrated to be the most potent inhibitors. Received: 10 November 1997 / Accepted: 17 February 1998  相似文献   

12.
Glutamate dehydrogenase [L-glutamate : NAD(P) oxidoreductase(deaminating) EC 1.4.1.3 [EC] .] has been purified from the mitochondrialfraction of green tobacco callus tissue. The enzyme was stableat –20?C for several months. The pH optimum for the aminationreaction was 7.8. But the optimum for the deamination reactionwas indistinct because it was in an extremely alkaline domain.Relative activities of the enzyme for amination were 50 withNADH and 10 with NADPH, and those for deamination were 5 withNAD and 1 with NADP at pH 7.9. The enzyme was inactivated by EDTA, but its activity partiallyrestored by the addition of divalent cations such as Ca2+, Mn2+,Zn2+, Cu2+ and Mg2+. Ca2+, Mn2+ and Zn2+ activated the reductiveamination 141, 122 and 39% respectively, but these divalentcations scarcely affected the oxidative deamination. Citrate and fumarate acted as inhibitors for reductive amination,and oxaloacetate for oxidative deamination of the enzyme reaction.These inhibitions were counteracted by the addition of Ca2+.ATP and ADP exerted an inhibitory effect on both directionsof the enzyme reaction. The inhibitory effect was hardly preventedby the addition of AMP. Ca2+ caused considerable recovery fromthe inhibition of ATP and ADP. Amino acids scarcely affectedthe enzyme activity. Michaelis constants were 0.28 mM for NAD, 0.065 mM for NADH,2.19 mM for a-ketoglutarate, 43.6 mM for ammonium chloride and4.24 mM for L-glutamate. 1To whom requests for reprints should be addressed. (Received June 25, 1980; )  相似文献   

13.
Rhodococcus globerulus K1/1 was found to express an inducible (S)-specific N-acetyl-2-amino-1-phenyl-4-pentene amidohydrolase. Optimal bacterial growth and amidohydrolase expression were both observed at about pH 6.5. Purification of the enzyme to a single band in a Coomassie blue-stained SDS-PAGE gel was achieved by nucleic acid and ammonium sulfate precipitation of Rhodococcus globerulus K1/1 crude extract and column chromatography on TSK Butyl-650(S) Fractogel and Superose 12HR. The amidohydrolase was purified to a homogeneity leading to a tenfold increase of the specific activity with a recovery rate of 65%. At pH 7.0 and 23 °C the enzyme showed no loss of activity after 30 days incubation. The amidohydrolase was stable up to 55 °C. The enzyme was inhibited strongly only by 10 mM Zn2+ among the tested metal cations and was inhibited 100% by 0.01 mM phenylmethanesulfonyl fluoride. The molecular weight of the native enzyme was estimated to be 92 kDa by gel filtration and 55 kDa by SDS-PAGE, suggesting a homodimeric structure. Received: 8 February 1999 / Received revision: 3 May 1999 / Accepted: 7 May 1999  相似文献   

14.
An NAD-dependent, morpholine-stimulated l-alanine dehydrogenase activity was detected in crude extracts from morpholine-, pyrrolidine-, and piperidine-grown cells of Mycobacterium strain HE5. Addition of morpholine to the assay mixture resulted in an up to 4.6-fold increase of l-alanine dehydrogenase activity when l-alanine was supplied at suboptimal concentration. l-Alanine dehydrogenase was purified to near homogeneity using a four-step purification procedure. The native enzyme had a molecular mass of 160 kDa and contained one type of subunit with a molecular mass of 41 kDa, indicating a tetrameric structure. The sequence of 30 N-terminal amino acids was determined and showed a similarity of up to 81% to that of various alanine dehydrogenases. The pH optimum for the oxidative deamination of l-alanine, the only amino acid converted by the enzyme, was determined to be pH 10.1, and apparent K m values for l-alanine and NAD were 1.0 and 0.2 mM, respectively. K m values of 0.6, 0.02, and 72 mM for pyruvate, NADH, and NH4 +, respectively, were estimated at pH 8.7 for the reductive amination reaction. Received: 25 September 1998 / Accepted: 11 March 1999  相似文献   

15.
Alanine dehydrogenase [L-alanine:NAD+ oxidoreductase (deaminating), EC 1.4.1.4.] catalyses the reversible oxidative deamination of L-alanine to pyruvate and, in the anaerobic bacterium Bilophila wadsworthia RZATAU, it is involved in the degradation of taurine (2-aminoethanesulfonate). The enzyme regenerates the amino-group acceptor pyruvate, which is consumed during the transamination of taurine and liberates ammonia, which is one of the degradation end products. Alanine dehydrogenase seems to be induced during growth with taurine. The enzyme was purified about 24-fold to apparent homogeneity in a three-step purification. SDS-PAGE revealed a single protein band with a molecular mass of 42 kDa. The apparent molecular mass of the native enzyme was 273 kDa, as determined by gel filtration chromatography, suggesting a homo-hexameric structure. The N-terminal amino acid sequence was determined. The pH optimum was pH 9.0 for reductive amination of pyruvate and pH 9.0-11.5 for oxidative deamination of alanine. The apparent Km values for alanine, NAD+, pyruvate, ammonia and NADH were 1.6, 0.15, 1.1, 31 and 0.04 mM, respectively. The alanine dehydrogenase gene was sequenced. The deduced amino acid sequence corresponded to a size of 39.9 kDa and was very similar to that of the alanine dehydrogenase from Bacillus subtilis.  相似文献   

16.
The bacterium Variovorax paradoxus, grown in a minimal medium in which silk fibroin represents the sole source of carbon and nitrogen, produces an extracellular protease that hydrolyzes fibroin as well as casein and, to a smaller extent, collagen and albumin. The optimal pH for activity was found to be in the acid range (optimum pH 5.8–6.4) and the enzyme activity was stimulated by the addition of divalent cations, either manganese or magnesium. Gel permeation chromatography and SDS-PAGE provided evidence that the native enzyme is a monomer with a Mr of ca. 21 kDa.  相似文献   

17.
Cloning and expression of the L-phenylalanine dehydrogenase (PheDH) gene from Bacillus sphaericus in B. subtilis was performed. It was ligated into the pHY300PLK shuttle vector and the resulting plasmid, pHYDH encoding polypeptide with molecular weight of 340 kDa, then transformed in B. subtilis ISW1214 and Escherichia coli JM109 competent cells for expression. Bacillus subtilis ISW1214/pHYDH only produced PheDH enzyme (4700 U/l). The recombinant PheDH was purified to near homogeneity as judged by SDS–polyacrylamide gel electrophoresis (M r 41000 Da) and the result was 40-fold with a yield of about 54%. Apparent K m values for L-phenylalanine (Phe), L-tyrosine and NAD+ were 0.24, 0.48 and 0.19 mM respectively. The optimum pH of the recombinant enzyme was 11 for the oxidative deamination, 10.2 for the reductive amination. The features of recombinant PheDH enzyme were comparable with the wild type PheDH protein.  相似文献   

18.
NADP+-specific glutamate dehydrogenase (EC 1.4.1.4) was purified to homogeneity from the extremely thermophilic, strictly anaerobic, sulfate-reducing archaeon Archaeoglobus fulgidus strain 7324. The native enzyme (263 kDa) is composed of subunits of mol. mass 46 kDa, suggesting a hexameric structure. The temperature optimum for enzyme activity was > 95° C. The enzyme was highly thermostable, having a half-life of 140 min at 100° C. Potassium phosphate, KCl, and NaCl enhanced the thermal stability and increased the rate of activity three- to fourfold. The N-terminal 26-amino-acid sequence showed a high degree of similarity to glutamate dehydrogenases from Pyrococcus spp. and Thermococcus spp. Received: 25 March 1997 / Accepted: 11 July 1997  相似文献   

19.
An isozyme of acid phosphatase-1, acid phosphatase-11, was purified from the leaves of tomato (Lycopersicon esculentum) to homogeneity and characterized. The purified enzyme was homogeneous on polyacrylamide gel electrophoresis with or without sodium dodecyl sulfate. The gel filtration analysis showed that the native molecule had a relative molecular mass of about 61 kilodaltons (kDa). The relative molecular mass of the subunit on gel electrophoresis with sodium dodecyl sulfate was about 32 kDa, indicating that the native form of the enzyme was a homodimer. It was suggested by periodic acid-Schiff staining on the gel that the enzyme was a glycoprotein. The Km for p-nitrophenylphosphate was 2.9 × 10?3 m. The enzyme had a pH optimum of 4.5 in 0.15 m potassium acetate buffer with p-nitrophenylphosphate as a substrate. This enzyme was activated by divalent metal ions, such as Zn2+, Mg2+, and Mn2+. The N-terminal amino acids were sequenced after the purified enzyme was treated with pyroglutamylpeptidase. It was suggested that the N-terminal amino acid was pyroglutamate.  相似文献   

20.
The enzyme uridine diphosphate N-acetylglucosamine pyrophosphorylase was purified about 330-fold from an extract of baker’s yeast by the treatment with protamine sulfate and column chromatographies on DEAE-cellulose, hydroxylapatite and Sephadex G–150. The purified enzyme was proved to be homogeneous by disc gel electrophoresis. The molecular weight was determined to be approximately 37,000 by gel filtration. The enzyme had an optimum reactivity in the pH range of 7.5-8.5 and was stable at 4°C in potassium phosphate buffer, pH 7.5, containing 0.1 mm dithiothreitol, but was unstable when stored at ?20°C. The addition of dithiothreitol also increased the thermal stability of enzyme. The enzyme was specific for UDP-N-acetylglucosamine as substrate, and none of the other sugar nucleotides could serve as nucleotide substrate. The estimated values of Km were 6.1 × 10?3 m for UDP-N-acetylglucosamine and 5.0 × 10?3 m for inorganic pyrophosphate. The enzyme required some divalent cations for activity. Magnesium ion was the most effective among the cations tested. The enzyme activity was highly stimulated by the addition of dithiothreitol or dithioerythritol.  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