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 medicinal mushroom Agaricus blazei produced high amounts of laccase (up to 5,000 units l–1) in a complex, agitated liquid medium based on tomato juice, while only traces of the enzyme (<100 units l–1) were detected in synthetic glucose-based medium. Purification of the enzyme required three chromatographic steps, including anion and cation exchanging. A. blazei laccase was expressed as a single protein with a molecular mass of 66 kDa and an isoelectric point of 4.0. Spectroscopic analysis of the purified enzyme confirmed that it belongs to the blue copper oxidases. The enzymes pH optimum for 2,6-dimethoxyphenol (DMP) and syringaldazine was pH 5.5; but for 2,2-azino-bis(3-ethylthiazoline-6-sulfonate) (ABTS) no distinct pH optimum was observed (highest activity at the lowest pH tested). Purified laccase was stable at 20°C, pH 7.0 and pH 3.0, but rapidly lost its activity at 40°C or pH 10. Sodium chloride strongly inhibited the enzyme activity, although the inhibition was completely reversible. The following kinetic constants were determined (Km, kcat): 63 M, 21 s–1 for ABTS, 4 M, 5 s–1 for syringaldazine, 1,026 M, 15 s–1 for DMP and 4307 M, 159 s–1 for guaiacol. The results show that—in addition to the wood-colonizing white-rot fungi—the typical litter-decomposing basidiomycetes can also produce high titers of laccase in suitable liquid media.  相似文献   

2.
Summary Soft water of low buffer capacity was drawn from near the branchial surface of rainbow trout (Salmo gairdneri) at 15°C, using opercular catheters, to determine pH changes in water passing over the gills. Latex masks allowed measurement of ventilation volume, and concentrations of carbon dioxide, oxygen, ammonia, and titratable base in expired water were compared to concentrations in inspired water. Water passing over the gills was more basic than inspired water if the inspired water was pH 4–6 (maximum increase: +0.7 pH units near pH 5). Expired water was more acidic than inspired water if the inspired water was pH 6–10 (maximum decrease: –1.7 pH units near pH 9). Ventilation volume (0.37 l·kg–1·min–1) and oxygen consumption (1.7 mmol·kg–1·h–1) were constant in the pH range 4.6–10.1, but both increased by 1.6–2.4× near pH 4. Carbon dioxide transfer near the gills was about 100 M, ammonia transfer about 15 M, and titratable base added at the gills was about 30 M. A theoretical model using CO2, titratable base, and ammonia added at the gills, the titration characteristics of the defined soft water medium, and aquatic equilibria for CO2 and ammonia, adequately explained the experimentally observed changes in pH near trout gills. Our observations and predictive model indicate that any gill contaminant whose toxicity varies with pH may be more or less toxic at the gills than predicted from bulk water chemistry alone.Abbreviations pH ex expired pH - pH in inspired pH  相似文献   

3.
Summary In order to obtain a better understanding of the behaviour ofPediococcus pentosaceus in food products as well to facilitate the designing of industrial production processes for the organism, the growth and lactic acid production ofPediococcus pentosaceus in a complex glucose medium was followed in batch cultures at different gas environments (CO2, air, N2 and static cultures without gasflow), temperatures (10–50°C), pH (4.3–7.3) and nitrite concentrations (0–700 ppm). Optimal growth was obtained in CO2 at 40°C and pH 6.3 and resulted in a maximum specific growth rate ( max) of 1.27 h–1. In static culture at 40°C and pH 6.3 the max was 1.21 h–1. The max was, compared with static culture, reduced in air (12%) and nitrogen (26%). At 10°C the max was reduced by 99% and at 50°C by 88%. The reduction at pH 4.3 and 7.3 was 65% and 57%, respectively. Nitrite did not affect the max at any pH but increased the lag phase at pH 4.3 by a factor of 12. The lactic acid production was linked to the growth. The total amount of lactic acid produced was the same in all the tested gases and nitrite concentrations and also within the wide temperature range (15–45°C) and pH range (5.3–7.3). Mainly L(+)-lactic acid was produced during the exponential growth phase, but after this growth declined about 30% of the L(+)-lactic acid was converted to D(–)-lactic acid. The lactic acid product yield and the cellmass yied were both affected by the temperature but not by the pH.  相似文献   

4.
Using primary cultures of gill pavement cells from freshwater rainbow trout, a method is described for achieving confluent monolayers of the cells on glass coverslips. A continuous record of intracellular pH was obtained by loading the cells with the pH-sensitive flourescent dye 2,7-bis(2-carboxyethyl)-5(6)-carboxyfluorescein and mounting the coverslips in the flowthrough cuvette of a spectrofluorimeter. Experiments were performed in HEPES-buffered media nominally free of HCO3. Resting intracellular pH (7.43 at extracellular pH=7.70) was insensitive to the removal of Cl or the application of 4-acetamido-4-isothiocyanatostilbene-2,2-disulfonic acid (0.1 mmol·l–1), but fell by about 0.3 units when Na+ was removed or in the presence of amiloride (0.2 mmol·l–1). Exposure to elevated ammonia (ammonia prepulse; 30 mmol·l–1 as NH4Cl for 6–9 min) produced an increase in intracellular pH (to about 8.1) followed by a slow decay, and washout of the pulse caused intracellular pH to fall to about 6.5. Intracellular non-HCO 3 buffer capacity was about 13.4 slykes. Rapid recovery of intracellular pH from intracellular acidosis induced by ammonia prepulse was inhibited more than 80% in Na+-free conditions or in the presence of amiloride (0.2 mmol·l–1). Neither bafilomycin A1 (3 mol·l–1) nor Cl removal altered the intracellular pH recovery rate. The K m for Na+ of the intracellular pH recovery mechanism was 8.3 mmol·l–1, and the rate constant at V max was 0.008·s–1 (equivalent to 5.60 mmol H+·l–1 cell water·min–1), which was achieved at external Na+ levels from 25 to 140 mmol·l–1. We conclude that intracellular pH in cultured gill pavement cells in HEPES-buffered, HCO 3 -free media, both at rest and during acidosis, is regulated by a Na+/H+ antiport and not by anion-dependent mechanisms or a vacuolar H+-ATPase.Abbreviations BCECF 2,7-bis(2-carboxyethyl)-5(6)-carboxy-fluorescein - BCECF/AM 2,7-bis(2-carboxyethyl)-5(6)-carboxy-fluorescein, acetoxymethylester - Cholin-Cl choline chloride - DMSO dimethyl sulfoxide - EDTA ethylene diamine tetra-acetic acid - FBS foetal bovine serum - H + -ATPase Proton-dependent adenosine triphosphatase - HEPES N-[2-hydroxyethyl]piperazine-N[2-ethanesulfonic acid] - pH i intracellular pH - pH e extracellular pH - PBS phosphate-buffered saline - SITS 4-acetamido-4-isothiocyanatostilbene-2,2-disulfonic acid  相似文献   

5.
The effect of pH and transmembrane pH on the efficiency of the proton pump of the mitochondrialbc 1 complex bothin situ and in the reconstituted state was studied. In both cases the H+/e ratio for vectorial proton translocation by thebc 1 complex respiring at the steady state, under conditions in which the transmembrane pH difference (pH) represents the only component of the proton motive force (p), was significantly lower than that measured under level flow conditions. The latter amounts, at neutral pH, to 1 (2 including the scalar H+ release). In the reconstituted system steady-state pH was modulated by changing the intravesicular buffer as well as the intra/extra-liposomal pH. Under these conditions the H+/e ratio varied inversely with the pH. The data presented show that pH exerts a critical control on the proton pump of thebc 1 complex. Increasing the external pH above neutrality caused a decrease of the level flowH +/e ratio. This effect is explained in terms of proton/electron linkage inb cytochromes.  相似文献   

6.
Cell extracts of Agrobacterium tumefaciens, immobilised in calcium alginate beads, had a 7-fold increase in N-carbamoylase (N-carbamylamino acid amidohydrolase E.C. 3.5.1) activity on reaction with N-carbamylglycine. The hydantoinase (dihydropyrimidinase E.C. 3.5.2.2) and N-carbamoylase activities remained stable over 4 weeks storage at 4°C relative to the non-immobilised enzymes, with the hydantoinase activity showing a 5-fold increase in activity relative to the non-immobilised hydantoinase. The pH optima of the immobilised hydantoinase and N-carbamoylase enzymes decreased to pH 7 and pH 8, respectively. The temperature optimum remained at 40°C for the N-carbamoylase enzyme while the hydantoinase activity was optimal at 50°C.  相似文献   

7.
Pythium fluminum produced zoospores most abundantly at 15°C, whereas the optima forPythium group F andP. marsipium were 20 and 25°C, respectively. Increasing the incubation temperature above the optimum resulted in the decrease of the duration of zoospore production. InPythium group F the ability to produce zoospores was not lost even after incubation at 40°C for 24 h. On the other hand,P. marsipium andP. fluminum lost the ability under these conditions. Zoospore production was inhibited at pH 4.5 and 10.5 in all the species tested.Pythium fluminum andP. marsipium were found to have two pH optima for zoospore production (7.5 and 9.5 for the former and 5.5 and 8.5 for the latter). The optimum pH for zoospore production byPythium group F was 6.5–7.5. Moderate osmotic potentials (–0.27–0.47 MPa) appeared to favor zoospore production by the pythia tested. The effect of temperature, pH and osmotic potential on zoospore production was discussed in relation to pollution of pond water.  相似文献   

8.
Diatom communities of acidic mountain streams in Poland   总被引:1,自引:0,他引:1  
Kwandrans  Janina 《Hydrobiologia》1993,269(1):335-342
A comparison has been made of the species composition of diatom communities developing in acidic Polish mountain streams which flow over calcium-poor substrates: sandstones in the Silesian Beskid (section of the Western Carpathians), the witokrzyskie Mts, and over granite in the Karkonosze range (in the Sudetic Mts). The number of taxa and diversity of the diatom assemblages decreased along a decreasing pH gradient. The correlation between pH and the number of taxa was positive and significant (r 2 = 0.69, p < 0.005). A small number of species (< 20) and low diversity were found in the communities developing in strongly acidic streams such as in the witokrzyskie Mts with pH 4.1–5.2, and in the Silesian Beskid with pH 3.5–4.0. In the stream of the Karkonosze Mts, with pH 5.2–6.0, the communities were characterized by their greater number of species and higher diversity.Acidobiontic and acidophilous diatoms were generally dominant. The pH-indiferent forms were less abundant, and their proportion increased above pH 5.0. Eunotia exigua, E. paludosa var. trinacria, E. tenella and Pinnularia subcapitata dominated in streams with the lowest pH, while E. exigua, E. sudetica and Achnanthes kryophila predominated in a stream with water pH above 5.2. Eunotia exigua, a common acidobiontic species was present in all the examined communities, and was a strong dominant in waters of pH 5.0. A corresponding decrease in abundance of E. exigua was observed with an increase in pH.  相似文献   

9.
In this study, the behaviour of Lactobacillus fermentum CRL 722 and CRL 251 were evaluated under different pH conditions (pH 6.0, 5.5, 5.0, 4.5) and without pH control. Growth was similar under all conditions assayed except at pH 4.5. These microorganisms were able to eliminate raffinose, a nondigestible -oligosaccharide (NDO) found in soy products, showing a consumption rate of 0.25 g l–1 h–1 (pH 6.0–5.0). The removal of raffinose was due to the high -galactosidase (-gal) activities of these lactic acid bacteria, which was highest at pH 5.5 (5.0 U/ml). The yield of organic acids produced during raffinose consumption was also highest at this pH. The results of this study will allow selection of the optimum growth conditions of L. fermentum with elevated levels of -gal to be used in the reduction of NDO in soy products when used as starter cultures.  相似文献   

10.
Summary In jejunal brush-border membrane vesicles, an out-wardly directed OH gradient (in>out) stimulates DIDS-sensitive, saturable folate (F) uptake (Schron, C.M., 1985).J. Clin. Invest. 76:2030–2033), suggesting carrier-mediated folate: OH exchange (or phenomenologically indistiguishable H+: folate cotransport). In the present study, the precise role of pH in the transport process was elucidated by examinin F uptake at varying pH. For pH gradients of identical magnitude, F uptake (0.1 M) was geater at lower (pHint/pHext:5.5/4.5) compared with higher (6.5/5.5) pH ranges. In the absence of a pH gradient, internal Ftrans stimulated DIDS-sensitive3H-folate uptake only at pH6.0. Since setepwise increments ininternal pH (4.57.5; pHext=4.5) stimulated F uptake, an inhibitory effect of higherinternal pH was excluded. In contrast, with increasing external pH(4.356.5; pHint=7.8), a 50-fold decrement in F uptake was observed (H+ K m =12.8±1.2m). Hill plots of these data suggest involvement of at least one H+ (OH) at high pH (divalent F–2 predominates). Since an inside-negative electrical potential did not affect F uptake at either pHext 4.55 or 5.8, transport of F and F–2 is electroneutral. Kinetic parameters for F and F–2 were calculated from uptake data at pHext 4.55 and 5.0. Comparision of predictedvs. experimentally determined kinetic parameters at pHext 5.8 (K m =1.33vs. 1.70 m;V max=12.8vs. 58.0 pmol/mg prot min) suggest that increasing external pH lowers theV max, but does not affect thatK m, for carrier-mediated F transport. These data are consistent with similarK i's for sulfasalazine (competitive inhibitor) at pHext 5.35 and 5.8 (64.7 and 58.5 m, respectively). In summary, the jejunal F carrier mediates electroneutral transport of mono- and divalen F and is sensitive to extermal pH with a H+ K m (or OH IC50) corresponding to pH 4.89. External pH affects theV max, but not theK m for carriermediated F uptake suggesting a reaction mechanism involving a ternary complex between the outward-facing conformation of the carrier and the transported ions (F and either OH or H+) rather than competitive binding that is mutually exclusive.  相似文献   

11.
An extracellular protease was produced by Arthrobacter ramosus isolated from the alkaline lake of Lonar, Buldhana District of Maharashtra, India when grown on a synthetic medium of pH 10 containing casein. The optimum conditions for production were 3.0% initial casein concentration, 2% inoculum of 1 × 108 cells/ml, pH 9.0, temperature 30 °C and shaken culture conditions. The protease was purified by ammonium sulphate precipitation followed by Sephadex G-100 chromatography. Two proteases viz. Arthro I and Arthro II, having molecular weights 21 and 11.4 kDa respectively were isolated. The Arthro II fraction had K m 395 g/ml and V max 10.55 g/min for azocasein. The maximum activity of enzyme was at 55 °C and pH 8. It was thermostable (up to 80 °C), alkali stable (pH 12) and stable in commercial detergent. The enzyme may contain a thiol group at the active site.  相似文献   

12.
A nonsporulating fungus isolated from dioxin-containing tropical soils forms cellobiose dehydrogenase when grown in media supplemented by a source of cellulose. The enzyme purified to homogeneity by SDS-PAGE (yield, 43%) had an Mr of 95 kDa; its pH optimum was in the range 5.5–7.0; more than 50% activity was retained at pH 4.0–8.0 (citrate–phosphate buffer). The absorption spectrum of the enzyme in the visible range had the characteristic appearance of flavocytochrome proteins. Cellobiose dehydrogenase oxidized cellobiose and lactose (the respective K M values at pH 6.0 equaled 4.5 ± 1.5 and 56 M) in the presence of dichlorophenolindophenol (K M,app = 15 ± 3 M at pH 6.0) taken as an electron acceptor. Other sugars were barely if at all oxidized by the enzyme. Neither ethyl--D-cellobioside, heptobiose, nor chitotriose inhibited the enzymatic oxidation of lactose, even under the conditions of 100-fold molar excess. The enzyme was weakly inhibited by sodium azide dichlorophenolindophenol reduction and exhibited an affinity for amorphous cellulose. At 55°C and pH 6.0 (optimum stability), time to half-maximum inactivation equaled 99 min. The enzyme reduced by cellobiose was more stable than the nonreduced form. Conversely, the presence of an oxidizer (dichlorophenolindophenol) decreased the stability eight times at pH 6.0. In addition, the enzyme acted as a potent reducer of the one-electron acceptor cytochrome c 3+ (K M app = 15 M at pH 6.0).  相似文献   

13.
α-Amylase and glucoamylase production by Schwanniomyces castellii   总被引:1,自引:0,他引:1  
A chromogenic substrate (Cibachron blue-amylose), and soluble starch and maltose were used to characterize the amylolytic system from Schwanniomyces castellii 3754. The strain was able to produce inducible -amylase (EC 3.2.1.1) and glucoamylase (EC 3.2.1.3) when grown on different C sources. The effect of the C source was slightly different for -amylase and glucoamylase production. Melezitose, maltose and soluble starch enhanced both -amylase and glucoamylase synthesis to nearly the same extent; amylose, trehalose and cellobiose particularly induced -amylase synthesis. The optimal pH for the release of both amylases was 5.5–7.0; maximal -amylase synthesis, on the other hand, was observed in the medium buffered at pH 6.0. The optimal pH for -amylase and glucoamylase activity was in the range of 4.5–7.2 and 4.2–5.5, respectively. Temperatures allowing maximal activity were 45°C for -amylase and 45–52°C for glucoamylase; a rapid decline of both activities was observed just above these temperatures.The species Schwanniomyces castellii (together with Schw. alluvius) is now considered to be synonymous with Schw. occidentalis var. occidentalis (Kreger-Van Rij 1984).  相似文献   

14.
Phenanthrene solubilization and biodegradation with a biosurfactant (rhamnolipid) solution were investigated as a function of pH. Batch phenanthrene solubilization experiments were performed in the pH range 4–8 and the highest solubilities with the biosurfactant were detected around a pH of 4.5–5.5. The apparent solubility at pH 5.5 was 3.8 times greater than at pH 7 in the presence of 240 ppm rhamnolipid, probably due to the rhamnolipid—an anionic surfactant—forming different pH-dependent structures. Biodegradation experiments using Pseudomonas putida CRE 7 were performed in the absence and the presence of the rhamnolipid solution. Without the biosurfactant, the specific growth rate () at pH 6 was higher than at other pH values, and analysis for the total phenanthrene loss confirmed the trends in , with the greatest phenanthrene removal at pH 6. In presence of the rhamnolipid, the maximum value shifted to around pH 5, which showed maximum enhancement of solubility in the abiotic experiment. Although there was an increase in the observed specific growth rate with the biosurfactant, this increase was not as great as the increase in solubilization. For example, the 1.44 times increase in the value at pH 5 was lower than the 3.8 times enhancement in the solubility at the same pH. Thus, as observed by others, not all of the solubilized phenanthrene was bioavailable to the microorganisms. Interestingly, the results of a size distribution experiment showed that a large portion of the phenanthrene-rhamnolipid aggregates existed at a molecular weight of >300,000. Furthermore, this fraction appeared to be the most available for biodegradation, although not all the phenanthrene was bioavailable.  相似文献   

15.
Summary In jejunal brush-border membrane vesicles, an outwardly directed OH gradient (in>out) stimulates DIDS-sensitive, saturable folate (F) uptake (Schron, C.M. 1985.J. Clin. Invest. 76:2030–2033), suggesting carrier-mediated folate: OH exchange (or phenomenologically indistinguishable H+: folate cotransport). In the present study, the precise role of pH in the transport process was elucidated by examining F uptake at varying pH. For pH gradients of identical magnitude, F uptake (0.1 M) was greater at lower (pHint/pHext: 5.5/4.5) compared with higher (6.5/5.5) pH ranges. In the absence of a pH gradient, internal Ftrans stimulated DIDS-sensitive3H-folate uptake only at pH6.0. Since stepwise increments ininternal pH (4.57.5; pHext=4.5) stimulated F uptake, an inhibitory effect of higherinternal pH was excluded. In contrast, with increasing external pH (4.356.5; pHint=7.8), a 50-fold decrement in F uptake was observed (H+ K m =12.8±1.2 M). Hill plots of these data suggest involvement of at least one H+ (OH) at low pH (monovalent F predominates) and at least 2 H+ (OH) at high pH (divalent F–2 predominates). Since an inside-negative electrical potential did not affect F uptake at either pHext 4.55 or 5.8, transport of F and F–2 is electroneutral. Kinetic parameters for F and F–2 were calculated from uptake data at pHext 4.55 and 5.0. Comparison of predictedvs. experimentally determined kinetic parameters at pHext5.8 (K m =1.33vs. 1.70 M;V max=123.8vs. 58.0 pmol/mg prot min) suggest that increasing external pH lowers theV max, but does not affect theK m for carrier-mediated F transport. These data are consistent with similarK i ' s for sulfasalazine (competitive inhibitor) at pHext 5.35 and 5.8 (64.7 and 58.5 M, respectively). In summary, the jejunal F carrier mediates electroneutral transport of mono- and divalent F and is sensitive to external pH with a H+ K m (or OH lC50) corresponding to pH 4.89. External pH effects theV max, but not theK m for carriermediated F uptake suggesting a reaction mechanism involving a ternary complex between the outward-facing conformation of the carrier and the transported ions (F and either OH or H+),rather than competitive binding that is mutually exclusive.  相似文献   

16.
Summary The time-course of changes in skeletal muscle pH during arousal from hibernation in the Columbian ground squirrel was studied using31P NMR spectroscopy. In hibernation (T re 7–9°C), shoulder/neck muscle pH was 7.45±0.03 and Im was 0.60. In euthermia (T re 37°C), muscle pH was 7.24±0.05 and Im was 0.75. Thus the overall pH-temperature coefficient was-0.009 pH units/°C, indicating acidification of the muscle in hibernation. During the transition from hibernation to euthermia, however, the muscle shows a nonlinear pattern of pH change. In early arousal (T sh<20–25°C,T re<15°C) muscle pH does not change and muscle Im increases to 0.72. In later arousal (T sh>20–25°C,T re>15°C) muscle pH decreases gradually toward the euthermic value and muscle Im increases only slightly from 0.72 to 0.75. These results support the hypothesis that intracellular acidification of the muscle, present during hibernation, is reversed in early arousal. This may facilitate an increase in muscle metabolism and the contribution of maximal shivering thermogenesis to rewarming of the animal.Abbreviations Im dissociation ratio of protein imidazole buffergroups - NST non-shivering thermogenesis - BAT brown adipose tissue - dp H/dT temperature coefficient of pH - pH i intracellular pH - 31 P NMR 31Phosphorus nuclear magnetic resonance - P i chemical shift of inorganic phosphate relative to PCr - PCr phosphocreatine - T b body temperature - T re rectal temperature - T sh subcutaneous shoulder temperature - T a ambient temperature  相似文献   

17.
Competitive inhibition of soybean urease was studied at 36°C in aqueous solution (pH 4.95) in the presence of polycarbonyl compounds (PCCs): oxalyldihydrazide (ODH), its polydisulfide (poly(DSODH)), three cyclic -triketones (CTKs), and seven cyclic PCC species of differing structure. The inhibition constants of ureolysis (Ki) varied in the range 8.5–3800 µM depending on the structure of organic chelators for the nickel atom in urease. It was shown that pH variation within the range from 3.85 to 7.40 exerted a strong effect on the values of Ki of three CTKs and hydroxyurea, which was used as a reference: pH dependences of log Ki were linear in all cases and displayed a break at pH 6.0–6.5. The most effective inhibitor of ureolysis was poly(DSODH), which contained 28 carbonyl groups in the polymer molecule. The role of such factors as the number of carbonyl groups per PCC molecule, mutual arrangement, and reaction medium pH in the efficiency of the process of urease inhibition is discussed.Translated from Prikladnaya Biokhimiya i Mikrobiologiya, Vol. 41, No. 1, 2005, pp. 17–22.Original Russian Text Copyright © 2005 by Tarun, Rubinov, Metelitza.  相似文献   

18.
Summary Of 373 anaerobic microbial isolates screened for the enzymatic conversion of dihydrouracil to N-carbamyl--alanine, several strains of Clostridium spp., C. glycolicum, C. subterminale and Peptococcus anaerobius were positive. These Clostridium and Peptococcus strains produced also N-carbamyl-d-amino acids from the respective 5-monosubstituted hydantoins. The d-hydantoinase activity from whole cell suspensions of P. anaerobius strain CRDA 303 was characterized with regard to pH and temperature stability and activity by using dihydrouracil (DHU) and isopropylhydantoin (IPH) as substrates. The d-hydantoinase from P. anaerobius was optimal at 60°C and at pH 6.5–9.5 for the substrate DHU. It was stable up to 55°C and at pH 5.0–9.5 and could be stored at 4°C under an aerobic atmosphere for at least 14 days. Offprint requests to: A. Morin  相似文献   

19.
A method for isolating extracellular glucose oxidase from the fungus Penicillium funiculosum 46.1 using ultrafiltration membranes was developed. Two samples of the enzyme with a specific activity of 914–956 IU were obtained. The enzyme exhibited a high catalytic activity at pH above 6.0. The effective rate constant of glucose oxidase inactivation at pH 2.6 and 16°C was 2.74 × 10–6 s–1. This constant decreased significantly as the pH of the medium increased (4.0–10.0). The temperature optimum for glucose oxidase–catalyzed -D-glucose oxidation was in the range 30–65°C. At temperatures below 30°C, the activation energy for -D-glucose oxidation was 6.42 kcal/mol; at higher temperatures, this parameter was equal to 0.61 kcal/mol. Kinetic parameters of glucose oxidase–catalyzed -D-glucose oxidation depended on the initial concentration of the enzyme in the solution. Glucose oxidase also catalyzed the oxidation of 2-deoxy-D-glucose, maltose, and galactose.  相似文献   

20.
Mid-log-phase cell suspensions of Corydalis sempervirens Pers., when incubated in micromolar or submicromolar concentrations of fusicoccin, strongly acidified the culture medium. High-affinity fusicoccin-binding sites were found in microsomes prepared from these cells using the radioligand [3H]-9-norfusicoccin-8-alcohol. Binding was saturable with an apparent dissociation constant (K d) of 2.8 nM, a pH optimum of 6.0, a temperature optimum of 35° C and was rapid (t1/2 = 8 min). The site abundance was 0.76±0.17 pmol · (mg of protein)–1. In the same membrane preparations, the K+, Mg2+-ATPase (EC 3.6.1.3) was characterized. The enzyme was highly vanadate-sensitive (IC50=6.5 M) and nucleotide-specific (ATPNTP), had a pH optimum of 6.2, an apparent K m for ATP of 0.23±0.12 mM, and V max of 10.6±1.8 nkat (mg of protein)–1. Fusicoccin doubled V max and lowered, by a factor of 2, the apparent K m for ATP of the enzyme when the cells were incubated with the toxin for 30 min prior to homogenization of the cells. The stimulation of the enzyme was also pronounced when fusicoccin was added to the homogenization medium just prior to homogenization of the cells, but was slight to zero when the toxin was added at the microsomal stage. The pronounced stimulatory effect of fusicoccin on the ATPase was seen at pH 7.1, i.e. at a pH typical for the cytoplasmic compartment, but was not detectable at pH 6.2, the pH optimum of the enzyme. The implications of these findings for an understanding of fusicoccin action are discussed.Abbreviations [3H]ABE-FC 9-nor-8-(4-azido-3,5-[3H]-benzoyl-diaminoethyl)-fusicoccin - FC fusicoccin - FCol 9-norfusicoc-cin-8-alcohol - Mes 2(N-morpholino)ethanesulfonic acid This work was supported by the Deutsche Forschungsgemeinschaft, Bonn, FRG and the Fonds der Chemischen Industrie, Frankfurt, FRG (literature provision).  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