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.
The activity and stability of yeast alcohol dehydrogenase (YADH) entrapped in aerosol OT reverse micellar droplets have been investigated spectrophotometrically. Various physical parameters, e.g., water pool size, w(0), pH, and temperature, were optimized for YADH in water/AOT/isooctane reverse micelles. It was found that the enzyme exhibits maximum activity at w(0) = 28 and pH 8.1. It was more active in reverse micelles than in aqueous buffers at a particular temperature and was denatured at about 307deg;C in both the systems. At a particular temperature YADH entrapped in reverse micelles was less stable than when it was dissolved in aqueous buffer.  相似文献   

2.
We have compared hepatic alcohol dehydrogenase activities in chick, rat and human liver with the major alcohols in commercial alcoholic beverages. 1. Chick and rat hepatic alcohol dehydrogenase was greater when assayed at a physiological pH in buffer containing chloride ions, as compared with the activity in pyrophosphate buffer at alkaline pH. 2. In contrast to reports of instability of ADH to freezing, we found the enzyme from all three species stable to freezing in 0.25 M sucrose. 3. Rat liver enzymatic activity was unstable in the presence of substrate, where as that of chick and human was not. 4. For all three species, the Km of hepatic ADH for substrate decreased with increasing chain length of alcohols. In both chick and human samples, the Vmax values for the higher chain alcohols were similar to that with ethanol, while in rat samples, ADH activity was dramatically lower with the higher chain alcohols compared to ethanol.  相似文献   

3.
Acetaldehyde and butyraldehyde are substrates for alcohol dehydrogenase in the production of ethanol and 1-butanol by solvent-producing clostridia. A coenzyme A (CoA)-acylating aldehyde dehydrogenase (ALDH), which also converts acyl-CoA to aldehyde and CoA, has been purified under anaerobic conditions from Clostridium beijerinckii NRRL B592. The ALDH showed a native molecular weight (Mr) of 100,000 and a subunit Mr of 55,000, suggesting that ALDH is dimeric. Purified ALDH contained no alcohol dehydrogenase activity. Activities measured with acetaldehyde and butyraldehyde as alternative substrates were copurified, indicating that the same ALDH can catalyze the formation of both aldehydes for ethanol and butanol production. Based on the Km and Vmax values for acetyl-CoA and butyryl-CoA, ALDH was more effective for the production of butyraldehyde than for acetaldehyde. ALDH could use either NAD(H) or NADP(H) as the coenzyme, but the Km for NAD(H) was much lower than that for NADP(H). Kinetic data suggest a ping-pong mechanism for the reaction. ALDH was more stable in Tris buffer than in phosphate buffer. The apparent optimum pH was between 6.5 and 7 for the forward reaction (the physiological direction; aldehyde forming), and it was 9.5 or higher for the reverse reaction (acyl-CoA forming). The ratio of NAD(H)/NADP(H)-linked activities increased with decreasing pH. ALDH was O2 sensitive, but it could be protected against O2 inactivation by dithiothreitol. The O2-inactivated enzyme could be reactivated by incubating the enzyme with CoA in the presence or absence of dithiothreitol prior to assay.  相似文献   

4.
Acetaldehyde and butyraldehyde are substrates for alcohol dehydrogenase in the production of ethanol and 1-butanol by solvent-producing clostridia. A coenzyme A (CoA)-acylating aldehyde dehydrogenase (ALDH), which also converts acyl-CoA to aldehyde and CoA, has been purified under anaerobic conditions from Clostridium beijerinckii NRRL B592. The ALDH showed a native molecular weight (Mr) of 100,000 and a subunit Mr of 55,000, suggesting that ALDH is dimeric. Purified ALDH contained no alcohol dehydrogenase activity. Activities measured with acetaldehyde and butyraldehyde as alternative substrates were copurified, indicating that the same ALDH can catalyze the formation of both aldehydes for ethanol and butanol production. Based on the Km and Vmax values for acetyl-CoA and butyryl-CoA, ALDH was more effective for the production of butyraldehyde than for acetaldehyde. ALDH could use either NAD(H) or NADP(H) as the coenzyme, but the Km for NAD(H) was much lower than that for NADP(H). Kinetic data suggest a ping-pong mechanism for the reaction. ALDH was more stable in Tris buffer than in phosphate buffer. The apparent optimum pH was between 6.5 and 7 for the forward reaction (the physiological direction; aldehyde forming), and it was 9.5 or higher for the reverse reaction (acyl-CoA forming). The ratio of NAD(H)/NADP(H)-linked activities increased with decreasing pH. ALDH was O2 sensitive, but it could be protected against O2 inactivation by dithiothreitol. The O2-inactivated enzyme could be reactivated by incubating the enzyme with CoA in the presence or absence of dithiothreitol prior to assay.  相似文献   

5.
Ethanol oxidation by the soluble fraction of a rat hepatoma was compared to that of the liver. Ethanol oxidation by the hepatoma was NAD+-dependent and sensitive to pyrazole, suggesting the presence of alcohol dehydrogenase. At low concentrations of ethanol (10.8 mm) the alcohol dehydrogenase activities of hepatoma and liver supernatant fractions were comparable. When the concentration of ethanol was raised to 108 mm, the activity of the liver enzyme decreased, whereas the activity in hepatoma supernatant fractions was strikingly elevated. m-Nitrobenzaldehyde-reducing activity was also conspicuously higher in hepatoma supernatant fractions. By contrast the ability to metabolize steroids and cyclohexanone was less than that in supernatant fractions of the liver.Electrophoresis of the liver supernatant fractions on ionagar at pH 7.0 revealed only one component that oxidized ethanol. On the other hand, hepatoma supernatant fractions contained two components with alcohol dehydrogenase activity; one with the same electrophoretic mobility as the liver enzyme, the other showing a slower rate of migration. The latter component, which is absent in the liver, is referred to as hepatoma alcohol dehydrogenase. By electrophoresis on starch gels at pH 8.5, it could be demonstrated that the liver and hepatoma enzymes moved in opposite directions.The liver and hepatoma enzymes differ in electrophoretic mobility, susceptibility to heat treatment, pH activity optimum and some catalytic properties. The substrate specificity of the hepatoma enzyme is narrower than that of liver alcohol dehydrogenase; cyclohexanone or 3β-hydroxysteroids of A/B cis configuration and the corresponding 3-ketones are not substrates for the hepatoma enzyme. The overall substrate specificity characteristics are, however, similar to those of the liver enzyme in that the effectiveness of substrates increases with an increase in chain length and introduction of unsaturation or an aromatic group. Both liver and hepatoma alcohol dehydrogenase cross-react with antibody to horse liver alcohol dehydrogenase EE. The Michaelis constant for ethanol with the hepatoma enzyme is 223 mm, compared to 0.3 mm for liver alcohol dehydrogenase; at 1.0 m ethanol the hepatoma enzyme is not fully saturated with substrate. The Michaelis constant for 2-hexene-1-ol is 0.3 mm, indicating that the hepatoma enzyme is better suited for dehydrogenation of longer chain alcohols. Stomach alcohol dehydrogenase has kinetic properties comparable to those of the hepatoma enzyme, as well as similar electrophoretic mobility. The hepatoma enzyme can be detected in the serum of rats bearing hepatomas.  相似文献   

6.
The enzyme activity of glutathione reductase (NAD(P)H:oxidized-glutathione oxidoreductase, EC 1.6.4.2) incorporated in CTAB/H2O/CHCl3-isooctane (1:1, v/v) reverse micelles has been investigated. Enzyme follows the Michaelis-Menten kinetics within a specified concentration range. Effects of pH, waterpool (W0), and surfactant concentration on the activity of glutathione reductase have been studied in detail. Optimum pH for the maximum enzyme activity was found to be dependent on the size of the waterpool. Further, a substrate inhibition was observed when concentration of one of the substrates was present in large excess over the other substrate. Km values for the substrate, oxidized glutathione (GSSG) and NADPH in CTAB/H2O/CHCl3-isooctane (1:1, v/v) were determined at W0 values of 14.4, 20.0, 25.5 and 29.7, at pH 8.0. These values are close to those obtained in aqueous solution, whereas the kcat values vary with W0 values of 8.8 to 32.3. Studies on the storage stability in the reverse micelle at W0 29.7 and pH 8.0 showed that glutathione reductase retained about 80% of its activity even after a month. The enzyme showed a higher stability at high waterpool. Oxidized glutathione (GSSG) provides protection to glutathione reductase against denaturation on storage in reverse micellar solution. Apparently, the enzyme is able to acquire a suitable native conformation at waterpool 29.7 and pH 8.0 and thereby exhibits an activity and stability inside the micellar cavity that are almost equivalent to that in aqueous solution.  相似文献   

7.
A simple method useful for the joint evaluation of substrate partitioning and kinetic parameters for reactions catalyzed by enzymes entrapped in reverse micelles is proposed. The method is applied to the hydrolysis of 2-naphthyl acetate (2-NA) catalyzed by lipase in sodium 1,4-bis(2-ethylhexyl) sulfosuccinate (AOT)/buffer/heptane reverse micellar solutions. In the presence of micelles, the relationship between the initial reaction rate and the analytical concentration of 2-NA was dependent on AOT concentration at a constant W ([water]/[AOT]) value. The dependence of the initial reaction rate profiles with [AOT] was analyzed according with the method proposed to obtain the partition constant of 2-NA between the micelles and the external solvent, Kp. A value of Kp = 2.7 L mol(-1) was obtained irrespective of the water content of the micelles (W from 5 to 20). The catalytic rate constant kcat in the micellar solutions was independent of [AOT] but slightly decreased with an increase in W from 2 x 10(-6) mol g(-1) s(-1) at W = 5 to 1.2 x 10(-6) mol g(-1) s(-1) at W = 20. The apparent Michaelis constant determined in terms of the analytical concentration of 2-NA increased with [AOT] at a given W and moderately decreased with W at a fixed [AOT]. The increase with [AOT] is accounted for by considering the partitioning of the substrate. After correction for the partitioning of 2-NA values of (Km)corr were obtained as 3.9 x 10(-3) mol L(-1) (W = 5), 4.6 x 10(-3) mol L(-1) (W = 10), 2.3 x 10(-3) mol L(-1) (W = 15), and 1.7 x 10(-3) mol L(-1) (W = 20). The rate parameters in the aqueous phase in the absence of micelles, were obtained as (kcat)aq = 7.9 x 10(-6) mol g(-1) s(-1) and (Km)aq = 2.5 x 10(-3) mol L(-1). In order to compare the efficiency of the enzyme in the micellar solution with that in aqueous phase, the values of (Km)corr were in turn corrected to take into account differences in the substrate activity, obtaining so a set of (Km)*corr values. The efficiency of the enzyme in the micellar solution, defined as the ratio, kcat/(Km)*corr, was found to be higher than in the aqueous phase, even at high water contents (W = 20). This higher efficiency is due to a significant decrease in (Km)*corr values.  相似文献   

8.
Catalytic and spectroscopic properties of alcohol dehydrogenase from horse liver, incorporated in reversed micellar media, have been studied. Two different reversed micellar systems have been used, one containing an anionic [sodium bis(2-ethylhexyl)sulfosuccinate, AOT], the other containing a cationic (cetyltrimethylammonium bromide, CTAB) surfactant. With 1-hexanol as substrate the turnover number of the enzyme in AOT-reversed micelles is strongly dependent on the water content of the system. At low wo ([H2O]/[surfactant]) (wo less than 20) no enzymatic activity can be detected whereas at high wo (wo = 40) the turnover is only slightly lower than in aqueous solution. In CTAB-reversed micelles the dependence of the turnover number on wo is much less. The enzymatic activity is in this case significantly lower than in aqueous solution and increases only slightly with an increasing water content of the reversed micelles. Possible interactions of the protein with the surfactant interfaces in the reversed micellar media were studied via circular dichroism and fluorescence measurements. From the circular dichroism of the protein backbone it is observed that the protein secondary structure is not significantly affected upon incorporation in the reversed micelles since the far-ultraviolet spectrum is not altered. Results from time-resolved fluorescence anisotropy experiments indicate that, especially in AOT-reversed micelles, interactions between the protein and the surfactant interface are largely electrostatic in nature, as evident from the dependence on the pH of the buffer used. In CTAB-reversed micellar solutions such interactions appear to be much less pronounced than in AOT.  相似文献   

9.
NAD-dependent 1,2-propanediol dehydrogenase (EC 1.1.1.4) activity was detected in cell-free crude extracts of various propane-grown bacteria. The enzyme activity was much lower in 1-propanol-grown cells than in propane-grown cells of Pseudomonas fluorescens NRRL B-1244, indicating that the enzyme may be inducible by metabolites of propane subterminal oxidation. 1,2-Propanediol dehydrogenase was purified from propane-grown Ps. fluorescens NRRL B-1244. The purified enzyme fraction shows a single-protein band upon acrylamide gel electrophoresis and has a molecular weight of 760,000. It consists of 10 subunits of identical molecular weight (77,600). It oxidizes diols that possess either two adjacent hydroxy groups, or a hydroxy group with an adjacent carbonyl group. Primary and secondary alcohols are not oxidized. The pH and temperature optima for 1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propan1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propan1,2-propanediol dehydrogenase are 8.5 and 20-25 degrees C, respectively. The activation energy calculated is 5.76 kcal/mol. 1,2-Propanediol dehydrogenase does not catalyze the reduction of acetol or acetoin in the presence of NADH (reverse reaction). The Km values at 25 degrees C, pH 7.0, buffer solution for 1,2-propanediol and NAD are 2 X 10(-2) and 9 X 10(-5) M, respectively. The 1,2-propanediol dehydrogenase activity was inhibited by strong thiol reagents, but not by metal-chelating agents. The amino acid composition of the purified enzyme was determined. Antisera prepared against purified 1,2-propanediol dehydrogenase from propane-grown Ps. fluorescens NRRL B-1244 formed homologous precipitin bands with isofunctional enzymes derived from propane-grown Arthrobacter sp. NRRL B-11315, Nocardia paraffinica ATCC 21198, and Mycobacterium sp. P2y, but not from propane-grown Pseudomonas multivorans ATCC 17616 and Brevibacterium sp. ATCC 14649, or 1-propanol-grown Ps. fluorescens NRRL B-1244. Isofunctional enzymes derived from methane-grown methylotrophs also showed different immunological and catalytic properties.  相似文献   

10.
Trypsin and alpha-chymotrypsin were immobilized by gelentrapment in polyacrylamide cross-linked with N,N(1)-methylenebisacrylamide. The immobilized enzymes are catalytically efficient in suspensions of reverse micelles formed in isooctane by bis(2-ethylhexyl) sodium sulfosuccinate (AOT) and water. Both entrapped enzymes are stable in reverse micellar suspension at room temperature and pH 8.2 for 3 days and lose 30-40% activity after 1 week. The enzymes obey Michaelis-Menten kinetics in the investigated concentration range with K(m) values higher than those in solution. Activity of the enzymes is independent of the water content of the micellar solution. No shift in pH optimum was observed for immobilized trypsin activity toward Nalpha-benzoyl-L-arginine ethyl ester. The utility of the procedure, which combines the advantage of enzyme immobilization and enzymology in reverse micelles, is illustrated by an example of peptide synthesis. In particular, peptide synthesis (e. g., Z--Ala--Phe--Leu--NH(2)) using water-insoluble substrate has been performed with gelentrapped alpha-chymotrypsin in reverse micellar suspension with the advantage of efficient enzyme recycling.  相似文献   

11.
Bovine liver dihydrofolate reductase has been solubilized in reverse micelles of cationic surfactant cetyltrimethylammonium bromide (CTAB) in isooctane-chloroform (1:1,V/V) mixture. Variation of waterpool (WO), pH and surfactant concentration showed that the enzyme activity was regulated by these parameters and was higher than the activity found in aqueous buffer (defined as superactivity); the maximum being at WO 13.3, pH 7.0 and CTAB concentration 75 mM. The Michaelis constants, Km for the substrate FAH2 and NADPH were found to be greater than those determined in water. Since reverse micelles have some features similar to those of biomembranes, display of super activity by dihydrofolate reductase indicates that enzymes in vivo may possess higher activity than actually observed in vitro studies in aqueous solutions.  相似文献   

12.
Bindings of the phospholipase A2 from Trimeresurus flavoviridis to the monodispersed and micellar n-alkylphosphorylcholines (n-CnPC) were studied at 25 degrees C and ionic strength 0.2 by the aromatic CD and tryptophyl fluorescence methods, respectively. The bindings to micelles of the substrate analog were analyzed by assuming that the micellar surface has multiple binding sites for the enzyme and that these sites are identical and mutually independent. The enzyme binding site was found to accommodate a constant number of the substrate (monomer) molecules, N = 9-13. The binding constant to the micelle was about 40 times greater than it was to the monodispersed substrate. The binding constant to the micellar substrate analog increased on the binding of Ca2+ to the enzyme and decreased on modification of the N-terminal alpha-NH2 group, whereas the binding to the monodispersed substrate analog was independent of pH, of the Ca2+ binding, and of the chemical modification of the alpha-NH2 group. The kinetics of the hydrolyses of monodispersed and micellar dihexanoylphosphatidylcholines (diC6PC) were studied at 25 degrees C and ionic strength 0.2 by the pH-stat method in the presence of saturating amounts of Ca2+. The catalytic center activity, kappa cat, as well as the binding constant, 1/Km, for the micellar substrate, were found to be much greater than those for the monodispersed substrate. The binding constant, 1/Km, of the monodispersed substrate was independent of pH; this was in good agreement with that of the substrate analog described above. The pH-dependence curve of kappa cat for the monodispersed substrate exhibited two transitions, one below pH 6.5 and the other above pH 9.5.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
The behavior of proteases in lecithin reverse micelles   总被引:1,自引:0,他引:1  
Reverse micelles, formed in isooctane/alcohol by phosphatidylcholines of variable chain length (i.e. 6, 7 or 8 C atoms in the fatty acid moiety) have been studied, mostly in relation to their capability of solubilizing trypsin and alpha-chymotrypsin. It has been found that the capability of the lecithin reverse micellar systems to solubilize water is strongly affected by the chain length of the alkyl group and by the alcohol used as co-surfactant. The C8-lecithin system, i.e. 1,2-dioctanoyl-sn-glycero-3-phosphocholine, in isooctane/hexanol is the system which affords the maximal solubilization of water (up to wo 60, where wo = [H2O]/[lecithin]) and of the enzymes. The water of the water pool of lecithin reverse micelles has been investigated by 1H-NMR; the proton chemical shift as a function of wo was found to be similar to the case of reverse micelles formed by the well known negatively charged surfactant sodium bis(2-ethylhexyl sulfosuccinate). 31P-NMR studies show that the ionization behavior of phosphate groups is similar to that in bulk water, suggesting no anomaly in the pH behavior of this water pool. The stability of trypsin and alpha-chymotrypsin in the various lecithin reverse micellar system is similar and occasionally better than that in aqueous solution. The same holds for the kinetic behavior (kcat and Km have been determined for a few systems). The bell-shaped curve of the pH/activity profile in lecithin reverse micelles is, for both enzymes, shifted towards more alkaline values with respect to water. Bell-shaped curves are also obtained when studying the influence of wo on the enzyme activity, with an optimal wo which is in the range 7-10, a surprisingly small value considering that we are dealing with hydrolases. Circular dichroic studies have been carried out in order to correlate the activity with the protein conformation: for both enzymes, generally no marked perturbations appear as a consequence of the solubilization in the lecithin reverse micelles, but conditions can be found under which significant alterations are present. Certain properties of the two enzymes, which in water solution are very similar, become sharply different in reverse micelles, showing that occasionally the micellization is able to enhance the relatively small structural differences between the two proteins.  相似文献   

14.
纳米磁性壳聚糖微球固定化酵母醇脱氢酶的研究   总被引:1,自引:0,他引:1  
建立了以纳米级磁性壳聚糖微球(magnetic chitosan microspheres , M-CS)为载体固定化酵母醇脱氢酶(yeast alcohol dehydrogenase,YADH)的方法,优化了YADH的固定化条件,考察了固定化酶的性质。结果表明,M-CS 呈规则的圆球形,粒径在30nm 左右,具有较好的磁响应性。酵母醇脱氢酶固定化适宜条件为:50 mg 磁性壳聚糖微球,加入20mL 0.25 mg/mL 酵母醇脱氢酶(蛋白质含量)磷酸盐缓冲液(0.05 mol/L ,pH 7.0) ,在4 ℃固定2h。M-CS 容易吸附酵母醇脱氢酶,但吸附的酶量受载体与酶的比例、溶液的离子浓度、溶液pH的影响明显,而温度对吸附的酶量的影响则相对较弱。相对于游离的酵母醇脱氢酶,固定化酶的最适温度略有升高,可明显改善其热稳定性、酸碱稳定性、操作稳定性和贮存稳定性。  相似文献   

15.
A NAD (P)-linked alcohol dehydrogenase was isolated from the soluble extract of the strictly respiratory bacterium Alcaligenes eutrophus N9A. Derepression of the formation of this enzyme occurs only in cells incubated under conditions of restricted oxygen supply for prolonged times. The purification procedure included precipitation by cetyltrimethylammonium bromide and ammonium sulfate and subsequent chromatography on DEAE-Sephacel, Cibacron blue F3G-A Sepharose and thiol-Sepharose. The procedure resulted in a 120-fold purification of a multifunctional alcohol dehydrogenase exhibiting dehydrogenase activities for 2,3-butanediol, ethanol and acetaldehyde and reductase activities for diacetyl, acetoin and acetaldehyde. During purification the ratio between 2,3-butanediol dehydrogenase and ethanol dehydrogenase activity remained nearly constant. Recovering about 20% of the initial 2,3-butanediol dehydrogenase activity, the specific activity of the final preparation was 70.0 U X mg protein-1 (2,3-butanediol oxidation) and 2.8 U X mg protein-1 (ethanol oxidation). The alcohol dehydrogenase is a tetramer of a relative molecular mass of 156000 consisting of four equal subunits. The determination of the Km values for different substrates and coenzymes as well as the determination of the pH optima for the reactions catalyzed resulted in values which were in good agreement with the fermentative function of this enzyme. The alcohol dehydrogenase catalyzed the NAD (P)-dependent dismutation of acetaldehyde to acetate and ethanol. This reaction was studied in detail, and its possible involvement in acetate formation is discussed. Among various compounds tested for affecting enzyme activity only NAD, NADP, AMP, ADP, acetate and 2-mercaptoethanol exhibited significant effects.  相似文献   

16.
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   

17.
A study was carried out to determine the Michaelian parameters relative to the action of chymosin and pepsin A on bond Phe105-Met106 of bovine kappa0-casein (carbohydrate-free fraction in micellar state). The reaction was performed in citrate buffer, pH 6.2, at 30 degrees C. The reaction mixture was analysed by reverse phase HPLC. Dosages of peptide 106-169 (caseino macropeptide) at different reaction times from recordings of its absorbance at 220 nm gave the initial rates of reaction at each substrate concentration. From these values the following parameters were determined: kcat = 68.5 s-1, Km = 0.048 mM, kcat/Km = 1,413 mM-1 s-1 for chymosin, and kcat = 45 s-1, Km = 0.018 mM, kcat/Km = 2,439 mM-1 s-1 for pepsin A. For chymosin they are similar to those obtained previously in dimethyl glutarate buffer, pH 6.6, at 30 degrees C, using fragment 98-111 of kappa-casein as substrate. It can thus be concluded that neither the micellar state nor the presence of the whole peptide chain of kappa-casein (our conditions) significantly affect the action of chymosin on fragment 98-111, which seems to contain all information that makes bond 105-106 highly sensitive to chymosin. For pepsin A, only the information contained in fragment 103-108 appears to be required.  相似文献   

18.
pH-Dependence of hydrolytic activity of trypsin has been studied in cationic reverse micellar system of cetyltrimethylammonium bromide (CTAB) in (50% v/v) chloroform/isooctane using a positively charged substrate Nα-benzoyl-L-arginine ethyl ester (BAEE). The pH of the medium was varied from 4.0 to 8.5 with addition of 0.025 M citrate-phosphate buffer containing 1 mM CaCl2. Optimum pH for maximum enzyme activity, pHopt in reverse micelles is found to be similar to that observed in bulk aqueous solution (8.0–8.5). However, changes in activity of trypsin (kcat) as a function of water content W0 (W0 = [H2O]/[CTAB]) in reverse micelles are found to be pH dependent. At low pH (4.0) and low water content (W0 = 5) the enzyme is more active in reverse micelles than in bulk aqueous solution by a factor of 2. This ‘superactivity’ is lost at higher W0 values and the kcat in reverse micelles is found to be similar to that observed in aqueous bulk. At pH 5, the enzyme activity is found to be independent of W0 while at pH 6.0–6.5 the enzyme activity is low at W0 5 and increases with water content to a constant value which is still 50% lower than that in aqueous buffer. Above pH 7, the Wo-activity profile becomes distinctly bell shaped with W0 optimum around 10–15. The enzyme activity at optimum W0 is close to that observed in aqueous bulk.  相似文献   

19.
Alcohol metabolism in vivo cannot be explained solely by the action of the classical alcohol dehydrogenase, Class I ADH (ADH1). Over the past three decades, attempts to identify the metabolizing enzymes responsible for the ADH1-independent pathway have focused on the microsomal ethanol oxidizing system (MEOS) and catalase, but have failed to clarify their roles in systemic alcohol metabolism. In this study, we used Adh3-null mutant mice to demonstrate that Class III ADH (ADH3), a ubiquitous enzyme of ancient origin, contributes to alcohol metabolism in vivo dose-dependently resulting in a diminution of acute alcohol intoxication. Although the ethanol oxidation activity of ADH3 in vitro is low due to its very high Km, it was found to exhibit a markedly enhanced catalytic efficiency (kcat/Km) toward ethanol when the solution hydrophobicity of the reaction medium was increased with a hydrophobic substance. Confocal laser scanning microscopy with Nile red as a hydrophobic probe revealed a cytoplasmic solution of mouse liver cells to be much more hydrophobic than the buffer solution used for in vitro experiments. So, the in vivo contribution of high-Km ADH3 to alcohol metabolism is likely to involve activation in a hydrophobic solution. Thus, the present study demonstrated that ADH3 plays an important role in systemic ethanol metabolism at higher levels of blood ethanol through activation by cytoplasmic solution hydrophobicity.  相似文献   

20.
Alcohol dehydrogenase [EC 1.1.1.1] was purified to homogeneity from rabbit liver by water extraction, DEAE-cellulose treatment, affinity chromatography on 5'-AMP-Sepharose and gel filtration on Sephadex G-150 using dithiothreitol as a stabilizer. The purified enzyme has an estimated molecular weight of 72,000 and consists of two subunits with a molecular weight of about 36,000 each. The enzyme contains 4 g-atoms of zinc and 18 sulfhydryl groups per mol of protein and exhibits maximal activity at pH 10.8, with a second maximum at pH 7.5. The apparent Km values for ethanol and NAD+ are 0.45 mM and 53.19 microM, respectively, at pH 10.8 and 3.33 mM and 6.94 microM, respectively, at pH 7.5. The enzyme oxidizes ethanol most readily among the aliphatic alcohols studied and has very low substrate specificity for methanol. Among steroid alcohols, 5 beta-androstan-3 beta-ol-17-one serves as a substrate for the enzyme. Pyrazole and 4-methylpyrazole (which are well known alcohol dehydrogenase inhibitors), sulfhydryl reagents, heavy metal ions and metal-chelating agents inactivate the enzyme.  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