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 15 毫秒
1.
A "lactase solution" was prepared from Escherichia coli. The mechanism of its action has been studied and changes in the rate of hydrolysis under various conditions investigated. The hydrolysis of lactose by the enzyme approximates the course of reaction of the integrated Michaelis-Menten equation. One molecule of enzyme combines with one molecule of substrate. E. coli lactase is readily inactivated at pH 5.0, and its optimal activity at 36°C. is reached between pH 7.0 and pH 7.5. The optimal temperature for its action was found to be 46°C. when determinations were carried out after an incubation period of 30 minutes. Its inactivation by heat follows the course of a first order reaction, and the critical thermal increment between the temperatures of 45°C. and 53°C. was calculated to be 56,400 calories per mol. The enzyme is activated by potassium cyanide, sodium sulfide, and cysteine, and irreversibly inactivated by mercuric chloride, silver nitrate, and iodine. After inactivation with copper sulfate partial reactivation is possible, while the slight inhibition brought about by hydrogen peroxide is completely reversible. The possible structure of the active groups of E. coli lactase as compared with other enzymes has been discussed.  相似文献   

2.
Inactivation of the alcohol oxidase enzyme system of Pichia pastoris, during the whole-cell bioconversion of ethanol to acetaldehyde, was due to catabolite inactivation. Electron microscopy showed that methanol-grown cells contained peroxisomes but were devoid of these microbodies after the bioconversion. Acetaldehyde in the presence of O2 was the effector of catabolite inactivation. The process was initiated by the appearance of free acetaldehyde, and was characterized by an increase in the level of cyclic AMP, that coincided with a rapid 55% drop in alcohol oxidase activity. Further enzyme inactivation, believed to be due to proteolytic degradation, then proceeded at a constant but slower rate and was complete 21 h after acetaldehyde appearance. The rate of catabolite inactivation was dependent on acetaldehyde concentration up to 0.14 mM. It was temperature dependent and occurred within 24 h at 37°C and by 6 days at 15°C but not at 3°C. Alcohol oxidase activity was psychrotolerant, with only a 17% decrease in initial specific activity over a temperature drop from 37 to 3°C. In contrast, protease activity was inhibited at temperatures below 15°C. When the bioconversion was run at 3°C, catabolite inactivation was prevented. In the presence of 3 M Tris hydrochloride buffer, 123 g of acetaldehyde per liter was produced at 3°C, compared with 58 g/liter at 30°C. By using 0.5 M Tris in a cyclic-batch procedure, 140.6 g of acetaldehyde was produced.  相似文献   

3.
1. Alkaline phosphatase of Escherichia coli undergoes below pH 6·0 a reversible acid inactivation that has been studied and related to the extent of uptake of inorganic phosphate occurring below pH 6·0. 2. The rate of inactivation is rapid in the first few minutes but later it decreases markedly. Temperature, pH, composition of buffer and other factors have an important effect on the inactivation. 3. About 60% of the activity lost at pH values above 3·5 is rapidly recovered when the enzyme is taken back to pH 8·0, independently (within certain limits) of the extent of the inactivation. 4. Phosphate and Zn2+, although very good protectors of the inactivation by acid, are not by themselves able to reverse the acid inactivation. 5. Inorganic phosphate seems not to be incorporated into the acid-inactivated enzyme. 6. Incorporation of more than one mole of phosphate/mole of enzyme has been obtained, but the phosphate residues seem to be incorporated to serine residues with a common sequence, suggesting two identical active serine residues/molecule of active enzyme.  相似文献   

4.
Weis E 《Plant physiology》1982,70(5):1530-1534
The most heat-sensitive functions of chloroplasts in Spinacia oleracea L. including the stromal carboxylation reaction, the light-induced electrical field gradient across the thylakoid membrane, as well as the overall photosynthetic CO2 fixation were less affected by heat if chloroplasts were heated in the light: 50% inactivation occurred around 35°C in the dark and around 40°C in the light. Relative low light intensities were sufficient to obtain optimal protection against heat. In contrast, the light-induced ΔpH across the thylakoid membrane, the photophosphorylation, and the photochemical activity of photosystem II which were less sensitive to heat in the dark (50% inactivation above 40°C) were not protected by light. Photosystem II even was destabilized somewhat by light.

The effect of light on the heat sensitivity of the water-splitting reaction was dependent on the pH in the medium. Protection by light only occurred at alkaline pH, in which case heat sensitivity was high (50% inactivation at 33°C in the dark and at 38°C in the light). Protection was prevented by uncouplers. At pH 6.8 when the heat sensitivity was low in any case (50% inactivation at 41°C in the dark), light had no further protecting effect.

Protection by light has been discussed in terms of light-induced transport of protons from the stroma to the thylakoid space and related ion fluxes.

  相似文献   

5.
Thermal stability of antioxidant defense enzymes was investigated in leaf and inflorescence of heat adaptive weed Chenopodium album. Leaf samples were taken at early and late seedling stage in December (LD, 20 °C/4 °C) and March (LM, 31 °C/14 °C). Young inflorescence (INF) was sampled at flowering in April (40 °C/21 °C). LD, LM and INF crude protein extracts were subjected to elevated temperatures (5 to 100 °C) for 30′. Superoxide dismutase (SOD) was the most heat stable enzyme followed by Ascorbate peroxidase (APX). Two heat stable SOD isozymes were visible on native-PAGE at 100 °C in both leaf and INF. Some heat stable APX isozymes were more abundant in INF than leaf. Thermostability of catalase (CAT) increased with age and increasing ambient temperatures in leaves. CAT activity was observed up to 60 °C in leaves and INF while peroxidase (POX) retained activity up to 100 °C in INF due to one thermostable isozyme. Glutathione reductase (GR), dehydroascorbate reductase (DHAR, EC 1.8.5.1) and monodehydroascorbate reductase (MDHAR) showed activity up to 70 °C in both leaves and INF. DHAR activity was stable up to 60 °C while GR and MDHAR declined sharply after 40 °C. Constitutive heat stable isozymes of SOD and APX in leaves and INF may contribute towards heat tolerance in C. album.  相似文献   

6.
Phleinase induced in stem base of orchardgrass (Dactylis glomerata L.) after defoliation was partially purified with ammonium sulfate precipitation, DEAE-Sephadex chromatography, gel filtration, and preparative polyacrylamide gel electrophoresis. The molecular weight of phleinase was 57,000 as determined by gel chromatography. The enzyme showed normal Michaelis-Menten kinetics and its Km value was 91 millimolar for phlein of mean degree of polymerization 60 as substrate. Reaction velocity of the enzyme was proportional to molarity of phlein irrespective of its chain length (mean degree of polymerization, 30 to 314). Phleinase attacked terminal fructosyl linkage of phlein by multi-chain mechanism. Phleinase cleaved β-2,6 linkage, β-2,6 linkage branched with β-2,1 linkage, and β-2,1 linkage of fructan in order of affinity, but not sucrose. Phleinase exhibited an optimum activity at pH 5.5 at 40°C. Its complete inactivation occurred at 60 and 70°C without and with phlein, respectively. Heat inactivation of the enzyme was enhanced by p-chloromercuribenzoate and protected partially by l-cysteine. The enzyme was inhibited by sulfhydryl reagents such as p-chloromercuribenzoate and Hg2+. The modes of action of phleinase were compared with those of the related enzymes.  相似文献   

7.
The gene encoding subtilisin-like protease T. kodakaraensis subtilisin was cloned from a hyperthermophilic archaeon Thermococcus kodakaraensis KOD1. T. kodakaraensis subtilisin is a member of the subtilisin family and composed of 422 amino acid residues with a molecular weight of 43,783. It consists of a putative presequence, prosequence, and catalytic domain. Like bacterial subtilisins, T. kodakaraensis subtilisin was overproduced in Escherichia coli in a form with a putative prosequence in inclusion bodies, solubilized in the presence of 8 M urea, and refolded and converted to an active molecule. However, unlike bacterial subtilisins, in which the prosequence was removed from the catalytic domain by autoprocessing upon refolding, T. kodakaraensis subtilisin was refolded in a form with a putative prosequence. This refolded protein of recombinant T. kodakaraensis subtilisin which is composed of 398 amino acid residues (Gly−82 to Gly316), was purified to give a single band on a sodium dodecyl sulfate (SDS)-polyacrylamide gel and characterized for biochemical and enzymatic properties. The good agreement of the molecular weights estimated by SDS-polyacrylamide gel electrophoresis (44,000) and gel filtration (40,000) suggests that T. kodakaraensis subtilisin exists in a monomeric form. T. kodakaraensis subtilisin hydrolyzed the synthetic substrate N-succinyl-Ala-Ala-Pro-Phe-p-nitroanilide only in the presence of the Ca2+ ion with an optimal pH and temperature of pH 9.5 and 80°C. Like bacterial subtilisins, it showed a broad substrate specificity, with a preference for aromatic or large nonpolar P1 substrate residues. However, it was much more stable than bacterial subtilisins against heat inactivation and lost activity with half-lives of >60 min at 80°C, 20 min at 90°C, and 7 min at 100°C.  相似文献   

8.
9.
Agro-industrial residues and cow dung were used as the substrate for the production of alkaline protease by Bacillus cereus strain AT. The bacterial strain Bacillus cereus strain AT produced a high level of protease using cow dung substrate (4813 ± 62 U g−1). Physiological fermentation factors such as the incubation time (72 h), the pH (9), the moisture content (120%), and the inoculum level (6%) played a vital role in the enzyme bioprocess. The enzyme production improved with the supplementation of maltose and yeast extract as carbon and nitrogen sources, respectively. Sodium dodecyl sulfate–polyacrylamide gel electrophoresis and zymogram analysis of the purified protease indicated an estimated molecular mass of 46 kDa. The protease enzyme was stable over a temperature range of 40–50 °C and pH 6–9, with maximum activity at 50 °C and pH 8. Among the divalent ions tested, Ca2+, Na+ and Mg2+ showed activities of 107 ± 0.7%, 103.5 ± 1.3%, and 104.6 ± 0.9, respectively. The enzyme showed stability in the presence of surfactants such as sodium dodecyl sulfate and on various commercially available detergents. The crude enzyme effectively de-haired goat hides within 18 h of incubation at 30 °C. The enzymatic properties of this protease suggest its suitable application as an additive in detergent formulation and also in leather processing. Based on the laboratory results, the use of cow dung for producing and extracting enzyme is not cumbersome and is easy to scale up. Considering its cheap cost and availability, cow dung is an ideal substrate for enzyme bioprocess in an industrial point of view.  相似文献   

10.
Developing soybean (Glycine max) seeds respond to a change in growth temperature by changing the level of stearoyl acyl carrier protein desaturase activity in the tissue. After 20 hours in liquid culture, seeds grown at 20°C show an increase in activity while seeds grown at 35°C show a decrease in activity, relative to their preculture levels. Analysis of the enzyme from both growth conditions shows the change not to be due to induction of kinetically distinct iosenzymes; desaturase activities from both 20°C and 35°C have identical behavior with regard to pH, temperature optimum, substrate concentration and cofactor requirements. Experiments with boiled extracts indicate that the modulation is not caused by induction of metabolic effectors. From these data, it appears that stearoyl-acyl carrier protein desaturase responds to changes in growth temperature by altering the level of active enzyme present in the tissue. The magnitude of this response is a function of the developmental stage of the seed and not a function of the growth conditions of the parent plant. Changing the age of the seeds from early late R5 changed the ratio of 20:35°C activity from 3.8:1 to 1.2:1. Changing the temperature at which the parent plants were grown over a range from 20/12°C to 34/28°C (day/night) produced only minor, and inconsistent, changes in the ratio of 20:35°C activities.  相似文献   

11.
The thermal inactivation kinetics of Salmonella enteritidis PT4 between 49 and 60°C were investigated. Using procedures designed to eliminate methodological artifacts, we found that the death kinetics deviated from the accepted model of first-order inactivation. When we used high-density stationary-phase populations and sensitive enumeration, the survivor curves at 60°C were reproducibly biphasic. The decimal reduction time at 60°C (D60°C) of the tail subpopulation was more than four times that of the majority population. This difference decreased with decreasing temperature; i.e., the survivor curves became more linear, but the proportion of tail cells remained a constant proportion of the initial population, about 1 in 104 to 105. Z plots (log D versus temperature) for the two populations showed that the D values coincided at 51°C, indicating that the survivor curves should be linear at this temperature, and this was confirmed experimentally. Investigations into the nature of the tails ruled out genotypic differences between the populations and protection due to leakage from early heat casualties. Heating of cells at 59°C in the presence of 5 or 100 μg of chloramphenicol per ml resulted in reductions in the levels of tailing. These reductions were greatest at the higher chloramphenicol concentration. Our results indicate that de novo protein synthesis of heat shock proteins is responsible for the observed tailing. Chemostat-cultured cells heated at 60°C also produced biphasic survivor curves in all but one instance. Cells with higher growth rates were more heat sensitive, but tailing was comparable with batch cultures. Starved cells (no dilution input) displayed linear inactivation kinetics, suggesting that during starvation a rapid heat shock response cannot be initiated.  相似文献   

12.
The heat resistance of foot-and-mouth disease virus (FMDV) strains isolated from outbreaks in Thailand was investigated in phosphate-buffered saline (PBS) at 50, 60, 70, 80, 90, and 100°C. The first-order kinetic model fitted most of the observed linear inactivation curves. The ranges of decimal-reduction time (D value) of FMDV strains at 50, 60, 70, 80, 90, and 100°C were 732 to 1,275 s, 16.37 to 42.00 s, 6.06 to 10.87 s, 2.84 to 5.99 s, 1.65 to 3.18 s, and 1.90 to 2.94 s, respectively. The heat resistances of FMDV strains at lower temperature (50°C) were not serotype specific. The effective inactivating temperature is approximately 60°C. Heat resistances of FMDV strains at 90 and 100°C were not statistically different (P > 0.05), while the FMDV serotype O (OPN) appeared to be the most heat resistant at 60 to 80°C. The other observed inactivation curves were linear with shoulder or tailing (biphasic curves). The shoulder effect was mostly observed at 90 and 100°C, while the tailing effect was mostly observed at 50 to 80°C. The adjusted D values in the case of shoulder and tailing effects did not affect the overall estimated heat resistance of these FMDV strains, so even unadjusted D values of deviant inactivation curves were legitimate. The z values of FMDV serotypes O, A, and Asia 1 were 21.78 to 23.26, 20.75 to 22.79, and 19.87°C, respectively. The z values of FMDV strains studied were not statistically significantly different (P > 0.05). The results of this study indicated that the heat resistance in PBS of FMDV strains from Thailand was much less than had been reported for foreign epidemic FMDV strains.  相似文献   

13.
1. We confirm the observation of Bürk (1965) that Neurospora crassa NADP-linked glutamate dehydrogenase normally exists in an inactive form below pH7·0 and in a fully active form above pH8·0 in either tris or orthophosphate buffer. At pH7·4 the enzyme is about half activated at 25°. 2. The variety of the enzyme produced by the mutant am2l shows a similar behaviour except that the transition is shifted about one pH unit in the alkaline direction. 3. The am2l enzyme has previously been reported to be activated by brief warming to 30° in phosphate buffer at pH8·0. The wild-type enzyme shows a similar effect at pH7·0. In tris buffer this effect is much less pronounced. 4. The am2l enzyme is extremely unstable at 47° at pH7·0; its stability is somewhat greater at lower pH, and is markedly increased by increasing the pH in the range 7·0–8·7. The wild-type enzyme also shows an indication of a stability minimum at pH7·0, but a temperature of 60° is needed for a measurable rate of inactivation. 5. The inactive form of the enzyme is much more subject to thermal irreversible denaturation than is the active form.  相似文献   

14.
Peroxidase from date palm (Phoenix dactylifera L.) leaves was purified to homogeneity and characterized biochemically. The enzyme purification included homogenization, extraction of pigments followed by consecutive chromatographies on DEAE-Sepharose and Superdex 200. The purification factor for purified date palm peroxidase was 17 with 5.8% yield. The purity was checked by SDS and native PAGE, which showed a single prominent band. The molecular weight of the enzyme was approximately 55 kDa as estimated by SDS–PAGE. The enzyme was characterized for thermal and pH stability, and kinetic parameters were determined using guaiacol as substrate. The optimum activity was between pH 5–6. The enzyme showed maximum activity at 55 °C and was fairly stable up to 75 °C, with 42% loss of activity. Date palm leaves peroxidase showed Km values of 0.77 and 0.045 mM for guaiacol and H2O2, respectively. These properties suggest that this enzyme could be a promising tool for applications in different analytical determinations as well as for treatment of industrial effluents at low cost.  相似文献   

15.
1. Pancreatic islets from several mammalian species were investigated for hydrolytic activity towards glucose 6-phosphate. Both the total phosphatase activity towards this substrate and the proportion cleaving glucose 6-phosphate in preference to β-glycerophosphate varied widely between species. In pancreatic-islet homogenates prepared from mice and guinea pigs there was a higher rate of liberation of Pi at pH6·7 from glucose 6-phosphate than from β-glycerophosphate. In these two species cortisone treatment enhanced the enzyme activity towards glucose 6-phosphate but not that towards β-glycerophosphate. Simultaneous injections of ethionine or puromycin blocked this stimulating effect of cortisone. 2. With whole homogenates of mouse pancreatic islets, inverse plots of the relationship between glucose 6-phosphate concentration and enzyme activity suggested the simultaneous action of two enzymes with different Km values. After fractionation of islets from obese–hyperglycaemic mice by differential centrifugation, one of these enzymes could be shown to be localized in the microsome fraction. It had Km for glucose 6-phosphate about 0·5mm and optimum pH6·7. It split glucose 6-phosphate in preference to β-glycerophosphate, glucose 1-phosphate, fructose 6-phosphate and fructose 1,6-diphosphate. Incubation of the microsomes at pH5·0 and 37° for 15min. decreased the enzyme activity by about 80%. Glucose was a potent inhibitor, the type of inhibition being neither strictly competitive nor non-competitive. It is suggested that the results indicate the presence of glucose 6-phosphatase in mammalian endocrine pancreas, and that this enzyme may play a role in the metabolic regulation of release of insulin.  相似文献   

16.
The activity of cytidine 5′-diphosphate (CDP) choline: 1,2-diacylglycerol cholinephosphotransferase (EC 2.7.8.2) in developing soybean (Glycine max L. var Williams 82) seeds was 3 to 5 times higher in cotyledons grown at 20°C than in those grown at 35°C. Some characteristics of the enzyme from cotyledons cultured at 20 and 35°C were compared. In preparations from both growth temperatures, the enzyme showed a pH optimum of 7, Km of 7.0 micromolar for CDP-choline, and an optimum assay temperature of 45°C. Both enzyme preparations were stimulated by increasing concentrations of Mg2+ or Mn2+, up to 10 millimolar and 50 micromolar, respectively, though Mn2+ produced lower activities than Mg2+. Enzymes from both 20 and 35°C show the same specificity for exogenous diacylglycerol. No metabolic effectors were detected by addition of heat treated extracts to the assay mixture. The above findings suggest that the higher enzyme activity at 20°C can be attributed to a higher level of the enzyme rather than to the involvement of isozymes or metabolic effectors. Enzyme activity decreased rapidly during culture at 35°C, indicating a rapid turnover of the enzyme. The level of temperature modulation was found to be a function of seed developmental stage.  相似文献   

17.
1. Pyruvate kinase of Alaskan king-crab leg muscle exists in two kinetically distinct forms, each of which displays a different temperature-dependence in the Km for phosphoenolpyruvate. 2. A `cold' variant of the enzyme has hyperbolic kinetics and exhibits a minimal Km for substrate at 5°. At physiological concentrations of phosphoenolpyruvate the `cold' enzyme is active only below 10°. A `warm' pyruvate kinase has a minimal Km for substrate at about 12°. This enzyme displays sigmoidal kinetics and is likely to be inactive, at physiological substrate concentrations, at temperatures below 9°. 3. The combined activities of these two pyruvate kinases yield highly temperature-independent rates of catalysis, at physiological substrate concentrations, over the range of habitat temperatures encountered by the organism, namely 4–12°. 4. The two variants of pyruvate kinase do not appear to be isoenzymes in the conventional sense. Electrophoretic and electrofocus analyses revealed only single peaks of activity. 5. The results suggest that the `warm' pyruvate kinase and the `cold' pyruvate kinase are formed by a temperature-dependent interconversion of one protein species. This interconversion has major adaptive significance: as the temperature is lowered the `warm' enzyme is converted into the `cold' enzyme; the opposite situation obtains when the temperature is raised. Temperature changes thus mimic the effects noted for fructose 1,6-diphosphate on certain mammalian pyruvate kinases.  相似文献   

18.
The mevalonate-based isoprenoid biosynthetic pathway is responsible for producing cholesterol in humans and is used commercially to produce drugs, chemicals, and fuels. Heterologous expression of this pathway in Escherichia coli has enabled high-level production of the antimalarial drug artemisinin and the proposed biofuel bisabolane. Understanding the kinetics of the enzymes in the biosynthetic pathway is critical to optimize the pathway for high flux. We have characterized the kinetic parameters of phosphomevalonate kinase (PMK, EC 2.7.4.2) from Saccharomyces cerevisiae, a previously unstudied enzyme. An E. coli codon-optimized version of the S. cerevisiae gene was cloned into pET-52b+, then the C-terminal 6X His-tagged protein was expressed in E. coli BL21(DE3) and purified on a Ni2+ column. The KM of the ATP binding site was determined to be 98.3 µM at 30°C, the optimal growth temperature for S. cerevisiae, and 74.3 µM at 37°C, the optimal growth temperature for E. coli. The KM of the mevalonate-5-phosphate binding site was determined to be 885 µM at 30°C and 880 µM at 37°C. The Vmax was determined to be 4.51 µmol/min/mg enzyme at 30°C and 5.33 µmol/min/mg enzyme at 37°C. PMK is Mg2+ dependent, with maximal activity achieved at concentrations of 10 mM or greater. Maximum activity was observed at pH = 7.2. PMK was not found to be substrate inhibited, nor feedback inhibited by FPP at concentrations up to 10 µM FPP.  相似文献   

19.
Kasamo K 《Plant physiology》1988,87(1):126-129
Proton-translocating ATPase (H+-ATPase) was purified from mung bean (Vigna radiata L.) roots. Treatment of this enzyme with the arginine-specific reagent 2,3-butanedione in the presence of borate at 37°C (pH 7.0), caused a marked decrease in its activity. Under this condition, half-maximal inhibition was brought about by 20 millimolar 2,3-butanedione at 12 minutes. MgATP and MgADP, the physiological substrate and competitive inhibitor of the ATPase, respectively, provided partial protection against inactivation. Loss of activity followed pseudo-first order kinetics with respect to 2,3-butanedione concentration, and double log plots of pseudo-first order rate constants versus reagent concentration gave a curve with a slope of 0.984. Thus, inactivation may possibly result from reaction of one arginine residue at each active site of the enzyme. The results obtained from the present study indicate that at least one arginyl residue performs an essential function in the plasma membrane H+-ATPase, probably at the catalytic site.  相似文献   

20.
Translational thermotolerance in Saccharomyces cerevisiae   总被引:1,自引:0,他引:1       下载免费PDF全文
While protein synthesis is rapidly inactivated in Saccharomyces cerevisiae, cells shifted from log growth at 30°C to 43°C, a 1-h 37°C treatment given to cells just prior to the shift to 43°C partially blocks this inactivation. By contrast, such a pre-heat shock treament has no protective effect on translational inactivation at 45°C or higher. Cells allowed to approach stationary phase not only develop an enhanced thermotolerance relative to log cells but also exhibit a pronounced resistance to inactivation of protein synthesis at 43°C as well as at 45°C. We have found that this ‘translational thermotolerance’ can also be induced in S. cerevisiae by briefly treating log phase cells at 30°C with cycloheximide. Using such a procedure to induce stabilization of protein synthesis at 43°C, we have been able to show that heat shock-induced proteins are not responsible for the establishment of this protective effect. This work shows that enhanced thermotolerance can be induced in log cells even after a shift to 43°C, as long as a prior translational thermotolerance has been established. Futhermore, we show that the capacity of plateau cells to maintain translation at 43°C contributes significantly to their state of enhanced thermotolerance.  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