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 500 毫秒
1.
Osmotic regulation in the flagellate Ochromonas malhamensis Pringsheim is mainly mediated by changes in the pool size of α-galactosyl-(1 → 1)-glycerol (isofloridoside). Isofloridoside phosphate synthase, a regulated key enzyme responsible for the formation of isofloridoside phosphate, appears to exist as an inactive proenzyme which can be activated by incubation of crude cell extracts with endogenous or exogenous proteases.  相似文献   

2.
Thiolase (acetyl-coenzyme A [CoA] acetyltransferase, E.C. 2.3.1.19) from Clostridium acetobutylicum ATCC 824 has been purified 70-fold to homogeneity. Unlike the thiolase in Clostridium pasteurianum, this thiolase has high relative activity throughout the physiological range of internal pH of 5.5 to 7.0, indicating that change in internal pH during acid production is not an important factor in the regulation of this thiolase. In the condensation direction, the thiolase is inhibited by micromolar levels of CoA, and this may be an important factor in modulating the net condensation of acetyl-CoA to acetoacetyl-CoA. Other cofactors and metabolites that were tested and shown to be inhibitors are ATP and butyryl-CoA. The native enzyme consists of four 44,000-molecular-weight subunits. The kinetic binding mechanism is ping-pong. The Km value for acetyl-CoA is 0.27 mM at 30°C and pH 7.4. The Km values for sulfhydryl-CoA and acetoacetyl-CoA are, respectively, 0.0048 and 0.032 mM at 30°C and pH 8.0. The active site apparently contains a sulfhydryl group, but unlike other thiolases, this thiolase is relatively stable in the presence of 5,5′-dithiobis(2-nitrobenzoic acid). Studies of thiolase specific activity under various types of continuous fermentations show that regulation of this enzyme at both the genetic and enzyme levels is important.  相似文献   

3.
4.
α-Amylase (EC 3.2.1.1) was excreted by Calvatia gigantea in liquid growth media containing different sources of starch. Among the factors affecting enzyme production in shake flasks were the type and the concentration of starch and the nitrogen source supplied. Optimum cultural conditions for maximum enzyme production were: soluble starch concentration, 5%; inoculum size, 3.75 × 105 conidia per ml; 5-day cultivation time at 28 to 30°C. The observed maximum yield of 81.3 U of saccharifying enzyme activity per ml of growth medium was the highest ever reported in the literature for submerged cultures. Partially purified enzyme functioned optimally at pH 4.5 to 5.5 and 53 to 58°C. The activation energy of enzymic hydrolysis of starch in the range of 20 to 40°C was 8,125 cal/mol (ca. 3.41 × 104 J). The apparent Km value of the enzyme at 25°C was 7.68 × 10−4 g/ml. Some of the properties of the enzyme under investigation were similar to those of α-amylases excreted from molds producing large amounts of the enzyme.  相似文献   

5.
Turnover of galactosylglycerol and osmotic balance in ochromonas   总被引:7,自引:5,他引:2       下载免费PDF全文
Kauss H 《Plant physiology》1973,52(6):613-615
Osmotic balance in Ochromonas malhamensis is mediated directly by fluctuations in the pool size of α-galactosylglycerol (isofloridoside). Chase experiments with glucose-14C indicate that the pool is in rapid turnover even at constant size. The turnover rate is related to the pool size. Regulation of the pool size seems to occur at enzymic steps involved in the formation, as well as those involved in the degradation of isofloridoside.  相似文献   

6.
A novel thermoacidophilic pullulan-hydrolyzing enzyme (PUL) from hyperthermophilic archaeon Thermococcus kodakarensis (TK-PUL) that efficiently hydrolyzes starch under industrial conditions in the absence of any additional metal ions was cloned and characterized. TK-PUL possessed both pullulanase and α-amylase activities. The highest activities were observed at 95 to 100°C. Although the enzyme was active over a broad pH range (3.0 to 8.5), the pH optima for both activities were 3.5 in acetate buffer and 4.2 in citrate buffer. TK-PUL was stable for several hours at 90°C. Its half-life at 100°C was 45 min when incubated either at pH 6.5 or 8.5. The Km value toward pullulan was 2 mg ml−1, with a Vmax of 109 U mg−1. Metal ions were not required for the activity and stability of recombinant TK-PUL. The enzyme was able to hydrolyze both α-1,6 and α-1,4 glycosidic linkages in pullulan. The most preferred substrate, after pullulan, was γ-cyclodextrin, which is a novel feature for this type of enzyme. Additionally, the enzyme hydrolyzed a variety of polysaccharides, including starch, glycogen, dextrin, amylose, amylopectin, and cyclodextrins (α, β, and γ), mainly into maltose. A unique feature of TK-PUL was the ability to hydrolyze maltotriose into maltose and glucose.  相似文献   

7.
Solutions of the enzyme luciferase, extracted from Cypridina, were subjected at pH 6.8 to temperatures from 40–55°C. for times up to 24 hours. After the desired exposures samples were cooled rapidly to room temperature, mixed with luciferin, and the first order velocity constants (representing luciferase activity) of the resulting luminescent reactions were determined by a photo electric method. The form of the curve relating luciferase activity to time of exposure to a temperature in the above range is compound in nature. If the exposure to the high temperature is not too long, about two-thirds of the lost activity is slowly regained on standing at room temperature. The data were described by an equation which represents the following mechanism: See PDF for Equation A plot of the logarithm of the rate constant, k1, against the reciprocal of the absolute temperature yielded an experimental activation energy for this reaction of about 57,000 calories, typical of protein denaturation processes. Log k2 plotted against 1/T was described by either a curve or two straight lines, high activation energies resulting in either case, again indicating protein denaturation. The plot of log k3: vs. 1/T showed no apparent dependence upon temperature, k3 being practically constant over the range studied. This may indicate that the underlying mechanism is not actually as simple as pictured. Two other mechanisms that were also considered were discarded because of lack of experimental support. Measurements of the decrease of luciferase activity at 48°C. and at pH 6.7, pH 5.5, and pH 7.9 showed that inactivation of the enzyme at this temperature was much more rapid at pH 7.9 than at pH 6.7 and was even faster at pH 5.5. These results from the Cypridina luminescent system were compared with those of other investigators on other systems.  相似文献   

8.
Pyrophosphorylytic kinetic constants (S0.5, Vmax) of partially purified UDP-glucose- and ADP-glucose pyrophosphorylases from potato tubers were determined in the presence of various intermediary metabolites. The S0.5 of UDP-glucose pyrophosphorylase for UDP-glucose (0.17 millimolar) or pyrophosphate (0.30 millimolar) and the Vmax were not influenced by high concentrations (2 millimolar) of these substances. The most efficient activator of ADP-glucose pyrophosphorylase was 3-P-glycerate (A0.5 = 4.5 × 10−6 molar). The S0.5 for ADP-glucose and pyrophosphate was increased 3.5-fold (0.83 to 0.24 millimolar) and 1.8-fold (0.18 to 0.10 millimolar), respectively, with 0.1 millimolar 3-P-glycerate while the Vmax was increased nearly 4-fold. The magnitude of 3-P-glycerate stimulation was dependent upon the integrity of key sulfhydryl groups (−SH) and pH. Oxidation or blockage of −SH groups resulted in a marked reduction of enzyme activity. Stimulations of 3.1-, 2.9-, 4.8-, and 9.5-fold were observed at pH 7.5, 8.0, 8.5, and 9.0, respectively, in the presence of 3-P-glycerate (2 millimolar). The most potent inhibitor of ADP-glucose pyrophosphorylase was orthophosphate (I0.5 = 8.8 × 10−5. molar). This inhibition was reversed with 3-P-glycerate (1.2 × 10−4 molar), resulting in an increased I0.5 value of 1.5 × 10−3 molar. Likewise, orthophosphate (7.5 × 10−4 molar) caused a decrease in the activation efficiency of 3-P-glycerate (A0.5 from 4.5 × 10−6 molar to 6.7 × 10−5 molar). The significance of 3-P-glycerate activation and orthophosphate inhibition in the regulation of α-glucan biosynthesis in Solanum tuberosum is discussed.  相似文献   

9.
1. The cytokinase (tissue activator of plasminogen) content of several mammalian tissues was evaluated by a quantitative casein hydrolysis method. 2. An alkaline (pH10·5) extraction of cytokinase from rabbit kidney lysosome–microsome fraction, followed by chromatography on DEAE-cellulose at pH7·6 with stepwise or linear increase in concentration of phosphate buffer, gave an 86-fold purification of the enzyme. The purified material was non-proteolytic against casein and heated fibrin and was freeze-dried without significant loss of activity or solubility. 3. Cytokinase is a protein with E0·1%1cm.=0·87 at 280mμ, and does not possess sufficient hexose or sialic acid to be classified as a glycoprotein. It has S20,w 2·9–3·1s and molecular weight 50000 when measured on a calibrated Sephadex G-100 column. It has an isoelectric point between pH8 and pH9, and is maximally active and stable at pH8·5. It is inactivated by heat at 78°. 4. Cytokinase and human urokinase have the same Km value and are inhibited in a partially competitive manner by -aminohexanoic acid and aminomethylcyclohexanecarboxylic acid. They are also inhibited by cysteine and arginine, but are unaffected by iodoacetamide and p-chloromercuribenzoate. 5. On the basis of this and other evidence it is suggested that rabbit kidney cytokinase and human urokinase are similar, if not identical, enzymes.  相似文献   

10.
Regulation of 2-carboxyarabinitol 1-phosphatase   总被引:4,自引:3,他引:1       下载免费PDF全文
The regulation of 2-carboxyarabinitol 1-phosphatase (CA 1-Pase) by phosphorylated effectors was studied with enzyme purified from tobacco (Nicotiana tabacum) leaves. CA 1-Pase activity was most stimulated by fructose 1,6-bisphosphate, exhibiting an A0.5 value of 1.9 millimolar and a 10-fold enhancement of catalysis. With ribulose-1,5-bisphosphate, the A0.5 was 0.6 millimolar, and maximal stimulation of activity was 5.3-fold. Among the monophosphates, 3-phosphoglycerate and phosphoglycolate were more potent positive effectors than glyceraldehyde 3-phosphate, glucose 1-phosphate, glucose 6-phosphate, and dihydroxyacetone phosphate. Stimulation of CA 1-Pase by ribulose-1,5-bisphosphate and fructose 1,6-bisphosphate increased Vmax but did not appreciably alter Km (2-carboxyarabinitol 1-phosphate) values. Inorganic phosphate appeared to inhibit CA 1-Pase noncompetitively with respect to 2-carboxyarabinitol 1-phosphate, exhibiting a Ki of 0.3 millimolar. The results suggest that these positive and negative effectors bind to a regulatory site on CA 1-Pase and may have a physiologial role in the light regulation of this enzyme. Related experiments with CA 1-Pase inactivated by dialysis in the absence of dithiothreitol show that partial reactivation can be achieved in the presence of a range of reducing reagents, including dithiothreitol, cysteine, and reduced glutathione. This could imply an ancillary involvement of sulfhydryl reduction during light activation of CA 1-Pase in vivo. The enzyme was thermally stable up to 35°C, in contrast to ribulose-1,5-bisphosphate carboxylase/oxygenase activase which lost activity above 30°C. The activation energy for CA 1-Pase was calculated to be 56.14 kilojoules per mole.  相似文献   

11.
The Saccharomyces cerevisiae PGU1 gene was successfully expressed in Schizosaccharomyces pombe. The optimum pH and temperature for the recombinant enzyme were 5 and 40°C, respectively, these being around 0.5 U higher and 5°C lower than those shown by the native enzyme. The Km value was about fourfold higher than that of the S. cerevisiae enzyme. The recombinant endopolygalacturonase was more efficient in reducing the viscosity of polygalacturonic acid and was also more stable at different pHs and temperatures than the native enzyme.  相似文献   

12.
EstS1, a newly identified thermostable esterase from Sulfobacillus acidophilus DSM10332, was heterologously expressed in Escherichia coli and shown to enzymatically degrade phthalate esters (PAEs) to their corresponding monoalkyl PAEs. The optimal pH and temperature of the esterase were found to be 8.0 and 70°C, respectively. The half-life of EstS1 at 60°C was 15 h, indicating that the enzyme had good thermostability. The specificity constant (kcat/Km) of the enzyme for p-nitrophenyl butyrate was as high as 6,770 mM−1 s−1. The potential value of EstS1 was demonstrated by its ability to effectively hydrolyze 35 to 82% of PAEs (10 mM) within 2 min at 37°C, with all substrates being completely degraded within 24 h. At 60°C, the time required for complete hydrolysis of most PAEs was reduced by half. To our knowledge, this enzyme is a new esterase identified from thermophiles that is able to degrade various PAEs at high temperatures.  相似文献   

13.
1. β-Amylase obtained by acidic extraction of soya-bean flour was purified by ammonium sulphate precipitation, followed by chromatography on calcium phosphate, diethylaminoethylcellulose, Sephadex G-25 and carboxymethylcellulose. 2. The homogeneity of the pure enzyme was established by criteria such as ultracentrifugation and electrophoresis on paper and in polyacrylamide gel. 3. The pure enzyme had a nitrogen content of 16·3%, its extinction coefficient, E1%1cm., at 280mμ was 17·3 and its specific activity/mg. of enzyme was 880 amylase units. 4. The molecular weight of the pure enzyme was determined as 61700 and its isoelectric point was pH5·85. 5. Preliminary examinations indicated that glutamic acid formed the N-terminus and glycine the C-terminus. 6. The amino acid content of the pure enzyme was established, one molecule consisting of 617 amino acid residues. 7. The pH optimum for pure soya-bean β-amylase is in the range 5–6. Pretreatment of the enzyme at pH3–5 decreases enzyme activity, whereas at pH6–9 it is not affected.  相似文献   

14.
A manganese-containing superoxide dismutase (EC 1.15.1.1) was fully characterized from leaves of the higher plant Pisum sativum L., var. Lincoln. The amino acid composition determined for the enzyme was compared with that of a wide spectrum of superoxide dismutases and found to have a highest degree of homology with the mitochondrial manganese superoxide dismutases from rat liver and yeast. The enzyme showed an apparent pH optimum of 8.6 and at 25°C had a maximum stability at alkaline pH values. By kinetic competition experiments, the rate constant for the disproportionation of superoxide radicals by pea leaf manganese superoxide dismutase was found to be 1.61 × 109 molar−1·second−1 at pH 7.8 and 25°C. The enzyme was not sensitive to NaCN or to H2O2, but was inhibited by N3. The sulfhydryl reagent p-hydroxymercuribenzoate at 1 mm concentration produced a nearly complete inhibition of the manganese superoxide dismutase activity. The metal chelators o-phenanthroline, EDTA, and diethyldithiocarbamate all inhibited activity slightly in decreasing order of intensity. A comparative study between this higher plant manganese superoxide dismutase and other dismutases from different origins is presented.  相似文献   

15.
The β-lactamase from Klebsiella pneumoniae E70 behaved in a similar fashion to the TEM-2 plasmid mediated enzyme on reaction with clavulanic acid. Both enzymes produced two types of enzyme–clavulanate complex, a transiently stable species (t½=4min at pH7.3 and 37°C) and irreversibly inhibited enzyme. In the initial rapid reaction (2.5min) the enzymes partitioned between the transient and irreversible complexes in the ratios 3:1 for TEM-2 β-lactamase and 1:1 for Klebsiella β-lactamase. Biphasic inactivation was observed for both enzymes and the slower second phase was rate limited by the decay of the transiently stable complex. This decay released free enzyme for further reaction with fresh clavulanic acid, the products again partitioning between transiently stable and irreversibly inhibited enzyme. This cycle continued until all the enzyme had been irreversibly inhibited. A 115 molar excess of inhibitor was required to achieve complete inactivation of TEM-2 β-lactamase. Hydrolysis of clavulanic acid with product release appeared to occur with the inhibition reaction, which explained this degree of clavulanic acid turnover. The stoichiometry of the interaction with Klebsiella β-lactamase was not examined. The penicillinase from Proteus mirabilis C889 was rapidly inhibited by low concentrations of clavulanic acid. The major product was a moderately stable complex (t½=40min at pH7.3 and 37°C); the proportion of the enzyme that was irreversibly inactivated was small. The cephalosporinase from Enterobacter cloacae P99 had low affinity for the inhibitor and only reacted with high concentrations of clavulanic acid (k=4.0m−1·s−1) to produce a relatively stable complex (t½=180min at pH7.3 and 37°C). No irreversible inactivation of this enzyme was detected. The rates of decay of the clavulanate–enzyme complexes produced in reactions with Proteus and Enterobacter enzymes were markedly increased at acid pH.  相似文献   

16.
Proton (H+) conductive pathways are suggested to play roles in the regulation of intracellular pH. We characterized temperature-sensitive whole cell currents in mouse bone marrow–derived mast cells (BMMC), immature proliferating mast cells generated by in vitro culture. Heating from 24 to 36°C reversibly and repeatedly activated a voltage-dependent outward conductance with Q10 of 9.9 ± 3.1 (mean ± SD) (n = 6). Either a decrease in intracellular pH or an increase in extracellular pH enhanced the amplitude and shifted the activation voltage to more negative potentials. With acidic intracellular solutions (pH 5.5), the outward current was detected in some cells at 24°C and Q10 was 6.0 ± 2.6 (n = 9). The reversal potential was unaffected by changes in concentrations of major ionic constituents (K+, Cl, and Na+), but depended on the pH gradient, suggesting that H+ (equivalents) is a major ion species carrying the current. The H+ current was featured by slow activation kinetics upon membrane depolarization, and the activation time course was accelerated by increases in depolarization, elevating temperature and extracellular alkalization. The current was recorded even when ATP was removed from the intracellular solution, but the mean amplitude was smaller than that in the presence of ATP. The H+ current was reversibly inhibited by Zn2+ but not by bafilomycin A1, an inhibitor for a vacuolar type H+-ATPase. Macroscopic measurements of pH using a fluorescent dye (BCECF) revealed that a rapid recovery of intracellular pH from acid-load was attenuated by lowering temperature, addition of Zn2+, and depletion of extracellular K+, but not by bafilomycin A1. These results suggest that the H+ conductive pathway contributes to intracellular pH homeostasis of BMMC and that the high activation energy may be involved in enhancement of the H+ conductance.  相似文献   

17.
The β-galactosidase from the Antarctic gram-negative bacterium Pseudoalteromonas haloplanktis TAE 79 was purified to homogeneity. The nucleotide sequence and the NH2-terminal amino acid sequence of the purified enzyme indicate that the β-galactosidase subunit is composed of 1,038 amino acids with a calculated Mr of 118,068. This β-galactosidase shares structural properties with Escherichia coli β-galactosidase (comparable subunit mass, 51% amino sequence identity, conservation of amino acid residues involved in catalysis, similar optimal pH value, and requirement for divalent metal ions) but is characterized by a higher catalytic efficiency on synthetic and natural substrates and by a shift of apparent optimum activity toward low temperatures and lower thermal stability. The enzyme also differs by a higher pI (7.8) and by specific thermodynamic activation parameters. P. haloplanktis β-galactosidase was expressed in E. coli, and the recombinant enzyme displays properties identical to those of the wild-type enzyme. Heat-induced unfolding monitored by intrinsic fluorescence spectroscopy showed lower melting point values for both P. haloplanktis wild-type and recombinant β-galactosidase compared to the mesophilic enzyme. Assays of lactose hydrolysis in milk demonstrate that P. haloplanktis β-galactosidase can outperform the current commercial β-galactosidase from Kluyveromyces marxianus var. lactis, suggesting that the cold-adapted β-galactosidase could be used to hydrolyze lactose in dairy products processed in refrigerated plants.  相似文献   

18.
An alkaliphilic, thermophilic Bacillus sp. (NCIM 59) produced extracellular xylose isomerase at pH 10 and 50°C by using xylose or wheat bran as the carbon source. The distribution of xylose isomerase as a function of growth in comparison with distributions of extra- and intracellular marker enzymes such as xylanase and β-galactosidase revealed that xylose isomerase was truly secreted as an extracellular enzyme and was not released because of sporulation or lysis. The enzyme was purified to homogeneity by ammonium sulfate precipitation followed by gel filtration, preparative polyacrylamide gel electrophoresis, and ion-exchange chromatography. The molecular weight of xylose isomerase was estimated to be 160,000 by gel filtration and 50,000 by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, indicating the presence of three subunits. The enzyme is most active at pH 8.0 and with incubation at 85°C for 20 min. Divalent metal ions Mg2+, Co2+, and Mn2+ were required for maximum activity of the enzyme. The Km values for D-xylose and D-glucose at 80°C and pH 7.5 were 6.66 and 142 mM, respectively, while Kcat values were 2.3 × 102 s-1 and 0.5 × 102 s-1, respectively.  相似文献   

19.
The leucine specific serine proteinase present in the soluble fraction of leaves from Spinacia oleracea L. (called Leu-proteinase) has been purified by acetone precipitation and a combination of gel-filtration, ion exchange, and adsorption chromatography. This enzyme shows a molecular weight of 60,000 ± 3,000 daltons, an isoelectric point of 4.8 ± 0.1, and a relative electrophoretic mobility of 0.58 ± 0.03. The Leu-proteinase catalyzed hydrolysis of p-nitroanilides of N-α-substituted(-l-)amino acids as well as of chromogenic macromolecular substrates has been investigated between pH 5 and 10 at 23 ± 0.5°C and I = 0.1 molar. The enzyme activity is characterized by a bell-shaped profile with an optimum pH value around 7.5, reflecting the acid-base equilibrium of groups with pKa values of 6.8 ± 0.1 and 8.2 ± 0.1 (possibly the histidyl residue present at the active site of the enzyme and the N-terminus group). Among the substrates considered, N-α-benzoyl-l-leucine p-nitroanilide shows the most favorable catalytic parameters and allows to determine an enzyme concentration as low as 1 × 10−9 molar. In agreement with the enzyme specificity, only N-α-tosyl-l-leucine chloromethyl ketone, di-isopropyl fluorophosphate and phenylmethylsulfonyl fluoride, among compounds considered specific for serine enzymes, strongly inhibit the Leu-proteinase. Accordingly, the enzyme activity is insensitive to cations, chelating agents, sulfydryl group reagents, and activators.  相似文献   

20.
Initial reactions involved in the bacterial degradation of polycyclic aromatic hydrocarbons (PAHs) include a ring-dihydroxylation catalyzed by a dioxygenase and a subsequent oxidation of the dihydrodiol products by a dehydrogenase. In this study, the dihydrodiol dehydrogenase from the PAH-degrading Sphingomonas strain CHY-1 has been characterized. The bphB gene encoding PAH dihydrodiol dehydrogenase (PDDH) was cloned and overexpressed as a His-tagged protein. The recombinant protein was purified as a homotetramer with an apparent Mr of 110,000. PDDH oxidized the cis-dihydrodiols derived from biphenyl and eight polycyclic hydrocarbons, including chrysene, benz[a]anthracene, and benzo[a]pyene, to corresponding catechols. Remarkably, the enzyme oxidized pyrene 4,5-dihydrodiol, whereas pyrene is not metabolized by strain CHY-1. The PAH catechols produced by PDDH rapidly auto-oxidized in air but were regenerated upon reaction of the o-quinones formed with NADH. Kinetic analyses performed under anoxic conditions revealed that the enzyme efficiently utilized two- to four-ring dihydrodiols, with Km values in the range of 1.4 to 7.1 μM, and exhibited a much higher Michaelis constant for NAD+ (Km of 160 μM). At pH 7.0, the specificity constant ranged from (1.3 ± 0.1) × 106 M−1 s−1 with benz[a]anthracene 1,2-dihydrodiol to (20.0 ± 0.8) × 106 M−1 s−1 with naphthalene 1,2-dihydrodiol. The catalytic activity of the enzyme was 13-fold higher at pH 9.5. PDDH was subjected to inhibition by NADH and by 3,4-dihydroxyphenanthrene, and the inhibition patterns suggested that the mechanism of the reaction was ordered Bi Bi. The regulation of PDDH activity appears as a means to prevent the accumulation of PAH catechols in bacterial cells.  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