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.
Alcaligenes xylosoxydans subsp. xylosoxydans A-6 (Alcaligenes A-6) produced N-acyl-D-aspartate amidohydrolase (D-AAase) in the presence of N-acetyl-D-aspartate as an inducer. The enzyme was purified to homogeneity. The enzyme had a molecular mass of 56 kDa and was shown by sodium dodecyl sulfate (SDS)–polyacrylamide gel electrophoresis (PAGE) to be a monomer. The isoelectric point was 4.8. The enzyme had maximal activity at pH 7.5 to 8.0 and 50°C, and was stable at pH 8.0 and up to 45°C. N-Formyl (Km=12.5 mM), N-acetyl (Km=2.52 mM), N-propionyl (Km=0.194 mM), N-butyryl (Km=0.033 mM), and N-glycyl (Km =1.11 mM) derivatives of D-aspartate were hydrolyzed, but N-carbobenzoyl-D-aspartate, N-acetyl-L-aspartate, and N-acetyl-D-glutamate were not substrates. The enzyme was inhibited by both divalent cations (Hg2+, Ni2+, Cu2+) and thiol reagents (N-ethylmaleimide, iodoacetic acid, dithiothreitol, and p-chloromercuribenzoic acid). The N-terminal amino acid sequence and amino acid composition were analyzed.  相似文献   

2.
An intermediate radical, ?H2OH, was produced in aqueous methanol solution containing nitrous oxide by γ-irradiation. Yields of ethylene glycol and formaldehyde, the major and the minor product from ?H2OH, respectively, changed on the addition of some solutes. Cysteine lowered the both product yields to zero even at a low concentration of 5 × 10?5m. Oxygen of low concentrations (2.5~7.5 × 10?5 m) changed effectively the major product from ethylene glycol to formaldehyde. k (CySH+?H2OH)/k(O2+?H2OH) was calculated as 0.5.

Ascorbic acid (5 × 10?5 m) lowered ethylene glycol yield to 48%, cystine (10?3m) to 15%, methionine (10?3m) to 31%, histidine (10?3m) to 42%, tryptophan (10?3m) 46%, tyrosine (10?3m) to 77%, phenylalanine (10?3m) to 73%, hypoxanthine (10?3m) to 37%, adenine (10?3m) to 52%, uracil (10?3m) to 20%, thymine (10?3m) to 10%, cytosine (10?3 m) to 49%, rutin (10?3m) to 23%, pyrogallol (10?3m) to 41%, and gallic acid (10?3m) to 78% of the control. These results suggest that the reactions of the secondary radicals such as ?H2OH perform an important role in material change of foods irradiated with γ rays.  相似文献   

3.
Abstract

We present procedures for nucleoside and oligonucleotide synthesis, binding affinity (T m) and structural analysis (CD spectra) of 2′-deoxy-2′,2″-difluoro-α-D-ribofuranosyl and 2′-deoxy-2′,2″-difluoro-β-D-ribofuranosyl oligothymidylates. Possible reasons for the thermal instability of duplexes formed between these compounds and RNA or DNA targets are discussed.  相似文献   

4.
A newly found methanol-using bacterium, Mycobacterium gastri MB19, is a facultative methylotroph which assimilates methanol via the ribulose monophosphate pathway. 3-Hexulose phosphate synthase was purified from the organism and characterized. This enzyme was found to use glycolaldehyde (Km = 4.3 mm) and methylglyoxal (Km = 5.7 mm) as well as formaldehyde (Km = 1.4 mm) in the presence of d-ribulose 5-phosphate as an acceptor. The product of the condensation of glycolaldehyde with d-ribulose 5-phosphate was isolated by ion-exchange chromatography. The dephosphorylated product was tentatively identified as a heptulose with the molecular formula C7H14O7 from its spectrophotometric properties and GC-MS results.  相似文献   

5.
Various conditions for obtaining hybrids of the auxotrophic mutants SH1509 and SH1512 of Saccharomyces cerevisiae by electrofusion were investigated. An AC field of 400 Vp/cm and a DC field of 2 square pulses (7 kV/cm; 60/βsec each) at an interval of 0.5 sec were effective. Treatment with 0.2 (SH1509) or l.0 mg/ml (SH1512) Zymolyase for 1 or 1.5 hr was essential. As to the molarity of the osmotic stabilizer (sorbitol), the hybrid yield peaked at 0.6 m. The presence of CaCl2 (up to 0.4 mm) or 0.1 mm CaCl2 with 0.1 mm MgCl2 enhanced the yield. The temperature of the spheroplast suspension during pulsations also affected the yield, the most suitable temperature being 28°C.  相似文献   

6.
Uridine diphosphate (UDP)-glucose 4-epimerase (EC 5.1.3.2) has been purified over 1000-fold from extracts of wheat germ by MnCl2 treatment, (NH4)2SO4 fractionation, Sephadex column chromatography, and adsorption onto and elution from calcium phosphate gel. The enzyme has a pH optimum of 9.0. Km values are 0.1 mm for UDP-d-galactose and 0.2 mm for UDP-d-glucose. NAD is required for activity; Ka = 0.04 mm. NADH is an inhibitor strictly competitive with NAD; Ki = 2 μm. Wheat germ also contains UDP-l-arabinose 4-epimerase (EC 5.1.3.5) and thymidine diphosphate (TDP)-glucose 4-epimerase which are distinct from UDP-glucose 4-epimerase.  相似文献   

7.
Abstract

(1′R, 3′S and R, 5′S)-4′-Oxo-2′-oxabicyclo[3.1.0]hexan-3′-yl pyrimidines and purines were synthesized from ribonucleosides in 2-5 steps. The configurations of the base moieties in the cyclopropano keto-nucleosides were determined by NOE difference spectroscopy.  相似文献   

8.
The intra- and intermolecular hydrogen bonding (ΔGº298K ≈ 2, kcal mol?1) of 2′-OH in nucleos(t)ides has been reported by the temperature- and concentration-dependent NMR study in conjunction with dihedral dependence of the NMR derived both endo (3 J H,H)- and exocyclic (3 JH,OH) coupling constants, nOe contacts and lineshape analyses of hydroxyl protons for EtpA (1), 3′-dA (2), rA (3), 2′-dA (4) [Fig. 1] in DMSO-d 6 at 500 MHz.

Figure 1. The schematic representation of the bias of the dymanic two-state pseudorotational equilibrium between the North-type (N, C2′-exo -C3′-endo) and the South-type (S, C3′-exo-C2′-endo) [3a] Thibaudeau, C. and Chattopadhyaya, J. 1999. Stereoelectronic Effects in Nucleosides and Nucleotides and their Structural Implications Sweden: (ISBN 91-506-1351-0), Department of Bioorganic Chemistry, Uppsala University Press (fax: +4618554495). For review see: and references therein [Google Scholar] pseudorotamers of the sugar moeity for EtpA (1), 3′-dA (2), rA (3), 2′-dA (4) and torsion (Φ) around C2′/3′-O bond viz. Φ1 = ΦH2′?C2′?O?H and Φ2 = ΦH3′?C3′?O?H except in 1 where the torsion across C3′-O3′ bond is actually ?? [C4′-C3′-O3′-P].  相似文献   

9.
D-Galacturonic acid reductase, a key enzyme in ascorbate biosynthesis, was purified to homogeneity from Euglena gracilis. The enzyme was a monomer with a molecular mass of 38–39 kDa, as judged by SDS–PAGE and gel filtration. Apparently it utilized NADPH with a Km value of 62.5±4.5 μM and uronic acids, such as D-galacturonic acid (Km=3.79±0.5 mM) and D-glucuronic acid (Km=4.67±0.6 mM). It failed to catalyze the reverse reaction with L-galactonic acid and NADP+. The optimal pH for the reduction of D-galacturonic acid was 7.2. The enzyme was activated 45.6% by 0.1 mM H2O2, suggesting that enzyme activity is regulated by cellular redox status. No feedback regulation of the enzyme activity by L-galactono-1,4-lactone or ascorbate was observed. N-terminal amino acid sequence analysis revealed that the enzyme is closely related to the malate dehydrogenase families.  相似文献   

10.
1. Phosphomevalonate kinase and 5-pyrophosphomevalonate decarboxylase have been purified from the freeze-dried latex serum of the commercial rubber tree Hevea brasiliensis. 2. The phosphomevalonate kinase was acid- and heat-labile and required the presence of a thiol to maintain activity. 3. The 5-pyrophosphomevalonate decarboxylase was relatively acid-stable and more heat-stable than the phosphokinase. 4. Maximum activity of the phosphokinase was achieved at pH 7.2 with 0.2mm-5-phosphomevalonate (Km 0.042mm), 2.0mm-ATP (Km 0.19mm) and 8mm-Mg2+ at 40°C. The apparent activation energy was 14.8kcal/mol. 5. Maximum activity of 5-pyrophosphomevalonate decarboxylase was achieved at pH5.5–6.5 with 0.1mm-5-pyrophosphomevalonate (Km 0.004mm), 1.5mm-ATP (Km 0.12mm) and 2mm-Mg2+. The apparent activation energy was 13.7kcal/mol. The enzyme was somewhat sensitive to inhibition by its products, isopentenyl pyrophosphate and ADP.  相似文献   

11.
The protease from Streptomyces cellulosae preferentially catalyzed the condensation reaction producing tripeptide amides in highly concentrated mixture solutions of various dipeptides and amino acid amides, although it weakly hydrolyzed the substrates at the same time. The tripeptide amides formed were l-Leu-Gly-Gly-NH2 (PLGGN) from l-Leu-Gly and Gly-NH2 and l-Leu-Gly-l-Leu-NH2 (PLGLN) from l-Leu-Gly and l-Leu-NH2. Moreover, the ratio of the rate of PLGLN formation per the proteolytic activity of this enzyme was much larger than those of the other proteases tested.

The formation of PLGLN was studied at various concentrations of the substrates (l-Leu-Gly and. l-Leu-NH2). The dependences of the initial velocities of PLGLN formation on the substrates concentrations could be explained by a two-substrate, one-product reaction mechanism involving a single active center forming the peptide bonds and two substrate-binding sites. The values of the substrate dissociation constants for enzyme-substrate complexes were about 0.6 m for l-Leu-Gly and 0.008 m for l-Leu-NH2.  相似文献   

12.
1. The influence of cations on the active transport into cells of rat-brain-cortex slices of l-histidine, an amino acid that is not metabolized by this tissue, has been studied. 2. Like other amino acids, l-histidine accumulated in the cells in the presence of glucose in concentrations up to over double that in the incubation medium. 3. The active transport of l-histidine was highest in a medium containing Ca2+ (3mm). The addition of K+ (27mm) led to a marked decrease in the intracellular concentration of l-histidine, though the oxygen uptake of the slices was higher. 4. The active l-histidine transport was inhibited by NH4+. The inhibitory effect increased with the NH4+ concentration, being about 25% at 8mm, 65% at 20mm, and 90% at 27 and 50mm. The oxygen uptake of the brain slices was depressed by only 25% by the highest NH4+ concentration used, and less by lower concentrations.  相似文献   

13.
ABSTRACT

The nematocidal activities of the fatty acid esters of d-allose were examined using the larvae of C. elegans. Among the fatty acid esters, 6-O-octanoyl-d-allose (3) showed significant activity. 6-O-octanoyl-d-glucose (5) showed no activity, indicating that the D-allose moiety is essential for the nematocidal activity of 3. A nonhydrolyzable alkoxy analog 6-O-octyl-d-allose (6) also showed activity equivalent to that of 3.  相似文献   

14.
The cepA putative gene encoding a cellobiose phosphorylase of Thermotoga maritima MSB8 was cloned, expressed in Escherichia coli BL21-codonplus-RIL and characterized in detail. The maximal enzyme activity was observed at pH 6.2 and 80°C. The energy of activation was 74 kJ/mol. The enzyme was stable for 30 min at 70°C in the pH range of 6-8. The enzyme phosphorolyzed cellobiose in an random-ordered bi bi mechanism with the random binding of cellobiose and phosphate followed by the ordered release of D-glucose and α-D-glucose-1-phosphate. The K m for cellobiose and phosphate were 0.29 and 0.15 mM respectively, and the k cat was 5.4 s-1. In the synthetic reaction, D-glucose, D-mannose, 2-deoxy-D-glucose, D-glucosamine, D-xylose, and 6-deoxy-D-glucose were found to act as glucosyl acceptors. Methyl-β-D-glucoside also acted as a substrate for the enzyme and is reported here for the first time as a substrate for cellobiose phosphorylases. D-Xylose had the highest (40 s-1) k cat followed by 6-deoxy-D-glucose (17 s-1) and 2-deoxy-D-glucose (16 s-1). The natural substrate, D-glucose with the k cat of 8.0 s-1 had the highest (1.1×104 M-1 s-1) k cat/K m compared with other glucosyl acceptors. D-Glucose, a substrate of cellobiose phosphorylase, acted as a competitive inhibitor of the other substrate, α-D-glucose-1-phosphate, at higher concentrations.  相似文献   

15.
The substrate specificity of rice α-glucosidase II was studied. The enzyme was active especially on nigerose, phenyl-α-maltoside and maltooligosaccharides. The actions on isomaltose and phenyl-α-glucoside were weak, and on sucrose and methyl-α-glucoside, negligible. The α-glucans, such as soluble starch, amylopectin, β-limit dextrin, glycogen and amylose, were also hydrolyzed.

The ratio of the maximum velocities for hydrolyses of maltose (G2), nigerose (N), kojibiose (K), isomaltose (I), phenyl-α-maltoside (?M) and soluble starch (SS) was estimated to be 100: 94.4: 14.2: 7.1: 89.5: 103.1 in this order, and that for hydrolyses of malto-triose (G3), -tetraose (G4), -pentaose (G5), -hexaose (G6), -heptaose (G7), -octaose (G8), and amyloses ( and ), 113: 113: 113: 106: 113: 100: 106: 106. The Km values for N, K, I, ?M and SS were 2.4 mm, 0.58 mm, 20 mm, 1.6 mm and 5.0 mg/ml, respectively; those for G2, G3, G4, G5, G6, G7, G8, and , 2.4 mm, 2.2 mm, 2.1 mm, 1.5 mm, 1.0 mm, 1.1 mm, 0.95 mm, 1.5 mm and 1.1 mm.

Rice α-glucosidase II is considered an enzyme with a preferential activity on maltooligosaccharides.  相似文献   

16.
Polyamine oxidase from Penicillium chrysogenum oxidized spermine rapidly and spermidine slightly at pH 7.5. The apparent Km values for spermine and spermidine were calculated to be 2.25 × 10?5 m and 9.54 × 10?6 m, respectively. The relative maximum velocities for spermine and spermidine were 3.37 × 10?3 m (H2O2) per min per mg of protein and 2.08 × 10?4 m (H2O2) per min per mg of protein, respectively. Spermine oxidation of the enzyme was competitively inhibited by spermidine and putrescine. The apparent Ki values by spermidine and putrescine were calculated to be 3.00 × 10?5 m and 1.80 × 10?8 m, respectively. On the other hand, polyamine oxidase from Aspergillus terreus rapidly oxidized both spermidine and spermine at pH 6.5. The apparent Km values for spermidine and spermine were 1.20 × 10?8 m and 5.37 × 10?7 m, respectively. The relative maximum velocities for spermidine and spermine were 1.55 × 10?2 m (H2O2) per min per mg of protein and 6.20 × 10?3 m (H2O2) per min per mg of protein, respectively.

Differential determination of spermine and spermidine was carried out using the two enzymes. The initial rate was assayed with Penicillium enzyme and the end point was measured afte addition of Aspergillus enzyme. Small amounts of polyamines (25 to 200 nmol of spermine and 25 to 250 nmol of spermidine) were assayed by solving two simultaneous equations obtained from the rate assay method and the end point assay method. The calculated values were in close agreement with those obtained by an amino-acid analyzer.  相似文献   

17.
A restriction endonuclease, designated as DmaI, was purified from cell-free extracts of Deleya marina IAM 14114 by streptomycin treatment, ammonium sulfate fractionation and two steps of chromatographies on heparin-Sepharose CL-6B and Mono Q (HR 5/5, FPLC). The purified enzyme was homogeneous on SDS-polyacrylamide gel disk electrophoresis and a ligation-recutting test. The relative molecular mass measurements of the purified enzyme gave 28,000 daltons by SDS-polyacrylamide gel disk electrophoresis and 56,000 daltons by gel filtration. These data indicated that the purified enzyme (56,000 daltons) has a dimeric structure composed of two 28,000-dalton subunits. The isoelectric point was 5.5. The purified enzyme worked best at 37°C in a reaction mixture (50 μl) containing 1.0 μg λDNA, 10 mm Tris–HCl, 7 mm 2-mercaptoethanol, 7 mm MgCl2 and 100 mm NaCl (pH 7.5). The enzyme was stable up to 55°C and between pH 7.0 and 9.0. The purified enzyme recognizes the palindromic hexanucleotide DNA sequence 5′-CAGCTG-3′, cuts between G and C and produces a flush end (isoschizomer of PvuII).  相似文献   

18.
The effects of cytokinins, a class of plant hormones, on cell proliferation and protein synthesis were studied in rat-derived L6 myoblasts cultured in a serum-free medium. Of the three cytokinins tested, isopentenyladenine, zeatin and ribosylzeatin, isopentenyladenine (5 μm) most stimulated the growth and DNA synthesis of the myoblasts, and it dose-dependently (0 ~ 10μm) enhanced the proliferation and DNA synthesis of the cells. Isopentenyladenine (5 and 10 μm) increased protein synthesis to twice that of control (0μm). These results suggest that isopentenyladenine, a trace component in plant food and a plant hormone, can affect the metabolism of an animal cell line of myoblasts.  相似文献   

19.
meso-Diaminopimelate dehydrogenase (EC 1.4.1.16) was purified to homogeneity from Corynebacterium glutamicum ATCC 13032. The enzyme had a molecular weight of about 70,000 and consisted of two subunits identical in molecular weight. The enzyme was highly specific for meso-2,6-diaminopimelate. The pH optima for deamination and amination were about 9.8 and 7.9, respectively. The Michaelis constants were 3.1mm for meso-2,6-diaminopimelate, 0.12mm for NADP+, 0.28 mm for l-2-amino-6-ketopimelate, 36 mm for ammonia, and 0.13 mm for NADPH. d and l isomers of 2,6-diaminopimelate competitively inhibited the oxidative deamination of meso-2,6-diaminopimelate. The enzyme was distributed in a wider range of bacterial species than reported previously [Misono et al., J. Bacteriol., 137, 22 (1979)] when assayed by a sensitive formazan formation method.  相似文献   

20.
In a continuing effort to further explore the use of the average local ionization energy $ \overline{\mathrm{I}}\left( \mathbf{r} \right) $ as a computational tool, we have investigated how well $ \overline{\mathrm{I}}\left( \mathbf{r} \right) $ computed on molecular surfaces serves as a predictive tool for identifying the sites of the more reactive electrons in several nonplanar defect-containing model graphene systems, each containing one or more pentagons. They include corannulene (C20H10), two inverse Stone-Thrower-Wales defect-containing structures C26H12 and C42H16, and a nanotube cap model C22H6, whose end is formed by three fused pentagons. Coronene (C24H12) has been included as a reference planar defect-free graphene model. We have optimized the structures of these systems as well as several monohydrogenated derivatives at the B3PW91/6-31G* level, and have computed their $ \overline{\mathrm{I}}\left( \mathbf{r} \right) $ on molecular surfaces corresponding to the 0.001 au, 0.003 au and 0.005 au contours of the electronic density. We find that (1) the convex sides of the interior carbons of the nonplanar models are more reactive than the concave sides, and (2) the magnitudes of the lowest $ \overline{\mathrm{I}}\left( \mathbf{r} \right) $ surface minima (the $ {{\overline{\mathrm{I}}}_{{\mathrm{S}\text{,}\min }}} $ ) correlate well with the interaction energies for hydrogenation at these sites. These $ {{\overline{\mathrm{I}}}_{{\mathrm{S}\text{,}\min }}} $ values decrease in magnitude as the nonplanarity of the site increases, consistent with earlier studies. A practical benefit of the use of $ \overline{\mathrm{I}}\left( \mathbf{r} \right) $ is that a single calculation suffices to characterize the numerous sites on a large molecular system, such as graphene and defect-containing graphene models.
Figure
Convex 0.001 au molecular surface of hydrogenated inverse Stone-Thrower-Wales defect-containing model 4H, with the hydrogen attached to one of the central carbons fusing the two pentagons  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