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 453 毫秒
1.
The influence of microheterogeneity on enzyme inactivation kinetics is presented. Examples of different enzymes are given where microheterogeneity has been detected by different techniques. The different statistical models are presented which include the influence of microheterogeneity on enzyme inactivation kinetics and stability. As the microheterogeneity of the enzyme increases, there is a sharper decline in the normalized activity during the initial stages of the deactivation but a greater stability and activity, compared to similar homogeneous enzyme, as the deactivation proceeds. Microheterogeneity makes the deactivation reaction have a higher apparent order of reaction. The implications of microheterogeneity on enzyme inactivations are high lighted by different examples. The analysis provides fresh physical insights into the chemistry, subpopulations, structure, and function of enzymes.  相似文献   

2.
Irreversible thermal inactivation of the tetrameric form of human plasma butyrylcholinesterase (cholinesterase; EC 3.1.1.8) was studied in water and in deuterium oxide at pH 7 in the temperature range 53-65 degrees C. The enzyme inactivation follows a complex kinetics that may be described by the sum of two apparent first-order processes. The Eyring plot for enzyme inactivation exhibits a wavelike discontinuity over a span of 2 C degrees around 58 degrees C. This transition was interpreted in terms of equilibrium between two temperature-dependent conformational states. Though 2H2O does not alter the overall multistep inactivation process, a slight solvent isotope effect was observed: a stabilizing effect and a shift in the transition temperature. A comparison between several enzyme preparations revealed differences in thermodynamic activation parameters of inactivation suggesting microheterogeneity in enzyme structures. Kinetics of inactivation of usual (E1uE1u) and atypical (E1aEa1a++) enzymes were compared. The atypical enzyme was found to be more stable than the usual phenotype.  相似文献   

3.
The process of thermal inactivation of triosephosphate isomerase covalently attached to a silica-based support activated with p-benzoquinone was found to be a complex one. At 50 degrees C, a characteristic activation preceding the thermal inactivation was observed. Following the intramolecular changes caused by heat, the values of K(M) and V(max) were determined during the activation. It was presumed that the complex thermal inactivation kinetics reflects the microheterogeneity of the immobilized enzyme molecules. The phosphate ion proved to be a better stabilizer than the substrate. (c) 1992 John Wiley & Sons, Inc.  相似文献   

4.
5.
The thermal stability of a highly purified preparation of D-amino acid oxidase from Trigonopsis variabilis (TvDAO), which does not show microheterogeneity due to the partial oxidation of Cys-108, was studied based on dependence of temperature (20-60°C) and protein concentration (5-100 µmol L-1). The time courses of loss of enzyme activity in 100 mmol L-1 potassium phosphate buffer, pH 8.0, are well described by a formal kinetic mechanism in which two parallel denaturation processes, partial thermal unfolding and dissociation of the FAD cofactor, combine to yield the overall inactivation rate. Estimates from global fitting of the data revealed that the first-order rate constant of the unfolding reaction (k a) increased 104-fold in response to an increase in temperature from 20 to 60°C. The rate constants of FAD release (k b) and binding (k -b) as well as the irreversible aggregation of the apo-enzyme (k agg) were less sensitive to changes in temperature, their activation energy (E a) being about 52 kJ mol-1 in comparison with an E a value of 185 kJ mol-1 for k a. The rate-determining step of TvDAO inactivation switched from FAD dissociation to unfolding at high temperatures. The model adequately described the effect of protein concentration on inactivation kinetics. Its predictions regarding the extent of FAD release and aggregation during thermal denaturation were confirmed by experiments. TvDAO is shown to contain two highly reactive cysteines per protein subunit whose modification with 5,5'-dithio-bis (2-nitrobenzoic acid) was accompanied by inactivation. Dithiothreitol (1 mmol L-1) enhanced up to 10-fold the recovery of enzyme activity during ion exchange chromatography of technical-grade TvDAO. However, it did not stabilize TvDAO at all temperatures and protein concentrations, suggesting that deactivation of cysteines was not responsible for thermal denaturation.  相似文献   

6.
L S Cook  H Im    F R Tabita 《Journal of bacteriology》1988,170(12):5473-5478
Ribulose 1,5-bisphosphate (RuBP) carboxylase/oxygenase (RuBPC/O) was inactivated in crude extracts of Rhodospirillum rubrum under atmospheric levels of oxygen; no inactivation occurred under an atmosphere of argon. RuBP carboxylase activity did not decrease in dialyzed extracts, indicating that a dialyzable factor was required for inactivation. The inactivation was inhibited by catalase. Purified RuBPC/O is relatively oxygen stable, as no loss of activity was observed after 4 h under an oxygen atmosphere. The aerobic inactivation catalyzed by endogenous factors in crude extracts was mimicked by using a model system containing purified enzyme, ascorbate, and FeSO4 or FeCl3. Dithiothreitol was found to substitute for ascorbate in the model system. Preincubation of the purified enzyme with RuBP led to enhanced inactivation, whereas Mg2+ and HCO3- significantly protected against inactivation. Unlike the inactivation catalyzed by endogenous factors from extracts of R. rubrum, inactivation in the model system was not inhibited by catalase. It is proposed that ascorbate and iron, in the presence of oxygen, generate a reactive oxygen species which reacts with a residue at the activation site, rendering the enzyme inactive.  相似文献   

7.
8.
A cytosolic glutathione S-transferase from pig lung was purified 210-fold to apparent homogeneity. The enzyme was classified as a class pi isoenzyme on the basis of its physical and chemical properties. It is homodimeric with a subunit Mr of 23,500, has a pI of 7.2, and shows a high specific activity towards ethacrynic acid. The glutathione analogues, S-hexylglutathione and glutathione sulfonate, were strong reversible inhibitors. The enzyme's primary structure, established entirely by protein chemical methods, consists of 203 amino acids and is highly similar (82-84% residue identity) to the rat and human class pi isoenzymes. Furthermore, there was no evidence of microheterogeneity or post-translational modifications. Each subunit contains a highly reactive cysteine residue, the modification of which leads to enzyme inactivation. None of the cysteine residues in the pig enzyme appear to form intramolecular disulfide bonds. Singel crystals of the glutathione-S-transferase-glutathione-sulfonate complex were obtained by the hanging-drop method of vapour diffusion from poly(ethylene glycol) 4000 solutions. The crystals belong to the orthorhombic space group P212121 with unit cell dimensions of a = 10.125 nm, b = 8.253 nm and c = 5.428 nm and diffract to better than 0.22 nm.  相似文献   

9.
The enzymatic oxidation of D-glucose to 2-keto-D-glucose (D-arabino-hexos-2-ulose, D-glucosone) is of prospective industrial interest. Pyranose oxidase (POx) from Peniphora gigantea is deactivated during the reaction. To develop a kinetic model including the main reaction and the enzyme inactivation, possible side-reactions of the non-stabilised enzyme with D-glucosone, hydrogen peroxide, and peroxide radicals were considered. A developed step-by-step combined experimental and computational procedure allowed to discriminate among alternative inactivation mechanisms and provides an increased model reliability. The most probable scheme is the enzyme inactivation by hydroxyl radicals formed from produced H2O2 in the presence of Fe2+ ions. This .OH reaction is supported by matrix assisted laser desorption ionisation-mass spectrometry (MALDI-MS) measurement. The estimated kinetic parameter values for the main reaction are of the same order of magnitude as those reported in the literature. The identified model allows a satisfactory process simulation and highlights measures to prevent the enzyme activity loss.  相似文献   

10.
The half-time method for the determination of Michaelis parameters from enzyme progress-curve data (Wharton, C.W. and Szawelski, R.J. (1982) Biochem. J. 203, 351-360) has been adapted for analysis of the kinetics of irreversible enzyme inhibition by an unstable site-specific inhibitor. The method is applicable to a model in which a product (R) of the decomposition of the site-specific reagent, retaining the chemical moiety responsible for inhibitor specificity, binds reversibly to the enzyme with dissociation constant Kr: (formula; see text). Half-time plots of simulated enzyme inactivation time-course data are shown to be unbiased, and excellent estimates of the apparent second-order rate constant for inactivation (k +2/Ki) and Kr can be obtained from a series of experiments with varying initial concentrations of inhibitor. Reliable estimates of k +2 and Ki individually are dependent upon the relative magnitudes of the kinetic parameters describing inactivation. The special case, Kr = Ki, is considered in some detail, and the integrated rate equation describing enzyme inactivation shown to be analogous to that for a simple bimolecular reaction between enzyme and an unstable irreversible inhibitor without the formation of a reversible enzyme-inhibitor complex. The half-time method can be directly extended to the kinetics of enzyme inactivation by an unstable mechanism-based (suicide) inhibitor, provided that the inhibitor is not also a substrate for the enzyme.  相似文献   

11.
Enoate reductase (EC 1.3.1.31) can stereospecifically reduce a variety of alpha,beta-unsaturated carboxylates. Its use was extended to apolar media by incorporating the enzyme into a reversed micellar medium. The kinetics of the enzyme in such a medium have been investigated using 2-methylbutenoic acid as substrate and NADH as a cofactor and compared with the reaction rates in aqueous solution. In aqueous solution the enzyme obeys a ping pong mechanism [Bühler et al. (1982) Hoppe-Seyler's Z. Physiol. Chem 363, 609-625]. In 50 mM Hepes pH = 7.0 with ionic strength of 0.05 M the Michaelis constants for NADH and 2-methylbutenoic acid are 20 microM and 6.0 mM respectively. In reversed micelles the kinetics of the reaction (Michaelis constant, maximum velocity as well as inhibitory effects) were markedly different. The rate of the enzymatic reaction of enoate reductase was studied using various concentrations of 2-methylbutenoic acid and various NADH concentrations. In reversed micelles composed of the anionic detergent sodium di(ethylhexyl)sulphosuccinate, the enzymatic reaction deviates substantially from the values in aqueous solution. Using our model (see preceding paper in this issue of the journal), all kinetics could be explained as evolving from enclosure in reversed micelles without any change in the intrinsic rate parameters of the enzyme. So the enzyme itself is unaffected by incorporation in reversed micelles, but the rate of intermicellar exchange as well as the microheterogeneity of the medium, resulting in very high local concentrations of the substrate, are the most important factors altering the reaction pattern. The effect of the composition of the reversed micellar medium was also investigated using either a nonionic or a cationic surfactant. In these solutions too, exchange and microheterogeneity of the medium proved to be the most important parameters influencing the enzymatic reaction. In all reversed micellar solutions inhibition by the enoate was observed at an overall concentration of 0.5-5 mM, implying that a concentration of substrate equal to the Km value in aqueous solution may already cause inhibition in reversed micelles. At this level no inhibition by NADH was observed. The microheterogeneity of the medium also explains this inhibition of the enzyme at relatively low 2-methylbutenoic acid concentrations.  相似文献   

12.
Thermal inactivation of jack bean urease (EC 3.5.1.5) was investigated in a 0.1 M phosphate buffer with pH 7. An injection flow calorimetry method was adapted for the measurement of the enzyme activity. The inactivation curves were measured in the temperature range of 55 to 87.5 degrees C. The curves exhibited a biphasic pattern in the whole temperature range and they were well fitted with a biexponential model. A simultaneous fit of all inactivation data was based on kinetic models that were derived from different inactivation mechanisms and comprised the material balances of several enzyme forms and the enthalpy balance characterizing the initial heating period of enzyme solution. The multitemperature evaluation revealed that an adequate model had to incorporate at least three reaction steps. It was concluded that the key reaction steps at urease thermal inactivation were the reversible dissociation/denaturation of native form into an inactive denatured form, and irreversible association reactions of both the denatured and native forms.  相似文献   

13.
Hydrogenases catalyze the reversible activation of dihydrogen. The hydrogenases from the aerobic, N2-fixing microorganisms Azotobacter vinelandii and Rhizobium japonicum are nickel- and iron-containing dimers that belong to the group of O2-labile enzymes. Exposure of these hydrogenases to O2 results in an irreversible inactivation; therefore, these enzymes are purified anaerobically in a fully active state. We describe in this paper an electron acceptor-requiring and pH-dependent, reversible inactivation of these hydrogenases. These results are the first example of an anaerobic, reversible inactivation of the O2-labile hydrogenases. The reversible inactivation required the presence of an electron acceptor. The rate of inactivation was first-order, with similar rates observed for methylene blue, benzyl viologen, and phenazine-methosulfate. The rate of inactivation was also dependent on the pH. However, increasing the pH of the enzyme in the absence of an electron acceptor did not result in inactivation. Thus, the reversible inactivation was not a result of high pH alone. The inactive enzyme could not be reactivated by H2 or other reductants at high pH. Titration of enzyme inactivated at high pH back to low pH was also ineffective at reactivating the enzyme. However, if reductants were present during this titration, the enzyme could be fully reactivated. The temperature dependence of inactivation yielded an activation energy of 44 kJ X mol-1. Gel filtration chromatography of active and inactive hydrogenase indicated that neither dissociation nor aggregation of the dimer hydrogenase was responsible for this reversible inactivation. We propose a four-state model to describe this reversible inactivation.  相似文献   

14.
The bi-exponential time-course of detergent inactivation at 37 degrees C of C12E8-solubilized (Na+ + K+)-ATPase from shark rectal glands and ox kidney was investigated. The data for shark enzyme, obtained at detergent/protein weight ratios between 2 and 16, are interpreted in terms of a simple model where the membrane bound enzyme is solubilized predominantly as (alpha-beta)2 diprotomers at low detergent concentrations and as alpha-beta protomers at high C12E8 (octaethyleneglycoldodecylmonoether) concentrations. It is observed that the protomers are inactivated 15-fold more rapidly than the diprotomers, and that the rate of inactivation of both oligomers is proportional to the detergent/protein ratio. Inactivation of kidney enzyme was biexponential with a very rapid inactivation of up to 40% of the enzyme activity. The observed rate of inactivation of the slower phase varied with the detergent/protein ratio, but the inactivation pattern for the kidney enzyme could not readily be accommodated within the model for inactivation of the shark enzyme. The rates of inactivation at 37 degrees C were about the same in KCl and NaCl, i.e., in the E2(K) and E1 X Na forms, for both enzymes.  相似文献   

15.
A number of years ago we reported a two‐step inactivation mechanism for α‐amylase (enzyme) on the basis of theoretical and experimental studies in aqueous solutions. In the first step the metal (Ca2+) ion dissociates reversibly from the enzyme followed by an irreversible thermal inactivation of the apoenzyme. In this study we report inactivation of the enzyme in the presence of ethanol–water solutions. We noticed that as the concentration of ethanol in the aqueous solution is increased, the thermal inactivation of the enzyme is suppressed with almost no inactivation (in 1 h, 30°C) when 50% alcohol is present in the solution. These results are explained by the two‐step inactivation model. © 2016 American Institute of Chemical Engineers Biotechnol. Prog., 32:1271–1275, 2016  相似文献   

16.
It was shown that the blockage of epsilon-amino group of Lis-126 residue by 2,2,6,6-tetramethyl-4-oxo-piperidine-1-oxyl (TMPO) leads to the cooperative inactivation of glutamate dehydrogenase (L-glutamate-NAD(P)-oxidoreductase, EC 1.4.1.3). The data concerning cooperative inactivation of the enzyme are interpreted by the model of hexamer with identical orientation of subunits. It was shown that the modification of any of enzyme subunits is accompanied by an inactivation of the hexamer's fragment which is a dimer, with subunits interacting reciprocally by means of isological contacts.  相似文献   

17.
Lysosomal α-glucosidase (EC 3.2.1.20) was purified from a lysosome-enriched fraction of rat liver using an improved procedure. A purification factor of 2900-fold was reached, with a yield of 35%. Polyacrylamide gel electrophoresis of the purified enzyme in nondenaturing conditions, or in the presence of SDS, showed only one band. However, a microheterogeneity among enzyme subunits was detected by high-resolution two-dimensional electrophoresis.  相似文献   

18.
A series-type enzyme deactivation model is used to model and to quantitate some more complex enzyme deacti-vations. The influence of temperature, pH, immobilization, chemical modifier (inhibitor or protector), substrate, and metal ion on the inactivation kinetics and on the parameter values is examined. In some cases the influence of two parameters on enzyme inactivations is presented. This provides further physical insights into enzyme inactivation and stabilization processes.  相似文献   

19.
The molecular and kinetic properties of cytidine deaminase from E. coli and chicken liver show several interesting differences and similarities: 1. Both enzymes possess an oligomeric structure, and linear kinetics. 2. The chicken liver enzyme is strictly dependent on the presence of reducing agents and presents a microheterogeneity in the pure preparation. 3. Both enzymes display identical specificity and share a rapid-equilibrium random Uni-Bi mechanism of catalysis. 4. The chicken liver enzyme is inhibited competitively by dTTP, CMP and dCMP.  相似文献   

20.
A series-type model is utilized to show the influence of pH on enzyme inactivation kinetics and stability. Examples of enzyme inactivations involving both single-step and series-type mechanisms are presented. Empirical relations for the inactivation rate constant for the first step and the residual activity as a function of pH are presented. This provides physical insights into the enzyme inactivation processes. The analysis forms the beginning of a framework within which one could quantitatively manipulate the inactivation rate constants and the residual activity for enzymes in desired directions as a function of pH.  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