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.
Kinetic analysis of the activating effect of substrate on the cholinesterase catalysis is performed. There are determined values of coefficient of activation A in the pH zone 5.0-7.5 for the process of hydrolysis of acetylcholine, indophenylacetate (IPA), and 2,6-dichlorophenolindophenylacetate (DIPA) by cholinesterase (ChE) of horse blood serum, as well as of IPA and DIPA by ChE of optical ganglia of the Pacific squid Todarodes pacificus. The phenomenon of activation has not been revealed at hydrolysis of phenylacetate by the horse blood serum ChE. The conclusion is made that the cause of the activating effect of substrate on the process of enzymatic hydrolysis by ChEs of different origin is the presence of the onium grouping in the structure of substrates.  相似文献   

2.
The ability of a previously enriched microbial population to utilize isopropanol (IPA) as the sole carbon source within a minimal salts medium is studied. The advantage of prior enrichment procedures for the improvement of IPA biodegradation performance is demonstrated for an IPA concentration of up to 24 g L(-1). Results showing the interrelationship between temperature and substrate utilization and inhibition levels at temperatures of between 2 degrees C and 45 degrees C are examined. Models of inhibition based on enzyme kinetics are assessed via nonlinear analysis, in order to accurately represent the growth kinetics of this solvent-tolerant mixed culture. The model that best describes the data is the Levenspiel substrate inhibition model, which can predict the maximum substrate level above which growth is completely limited. This is the first report of IPA treatment of up to 24 g L(-1) by an aerobic solvent-tolerant population.  相似文献   

3.
The sensitivity to the inhibitor of two forms of reindeer liver carboxylesterases differing in electrophoretic mobility and conventionally termed as "slow" and "fast" forms were investigated. The rate constants for the interaction of organophosphorous irreversible inhibitors--diisopropylfluorophosphate (DPP) and two methylthiophosphonic acid thioesters--C5H11O(CH3)P(O)S(CH2)SCH2C(O)OCH3 (Sh-205) and, C8H17O(CH3)P(O)S(CH2)SCH2C(O)OCH2 (Sh-207)--with the "fast" form are hundreds of times as high as those with the "slow" one. The rate constants for irreversible carbamate inhibitor interaction byehone with both carboxylesterase forms were approximately equal to 1.2 X 10(3) M-1 X min-1 and 2.0 X 10(3) M-1 X min-1, respectively. The reversible inhibitors potassium benzylate and kathapin also inhibited the "fast" carboxylesterase form in the indophenylacetate (IPA) hydrolysis reaction (770 and 1700-fold, respectively). On the contrary, N-methylpiperidinyl ester of benzyl acid inhibited the "slow" form three times stronger. Carbophos reversibly inhibited IPA hydrolysis in the presence of both enzyme forms, but the carboxyester carbophos group was hydrolyzed at a measurable speed only by the "slow" form.  相似文献   

4.
We re-examined the kinetics of the bisphosphatase reaction of rat hepatic 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase after depleting the enzyme of bound fructose 6-phosphate and found a hyperbolic dependence on fructose 2,6-bisphosphate at concentrations below 100 nM. The Michaelis constant was 4 nM, the Vmax was about 12 nmol X mg-1 X min-1 at 22 degrees C but the substrate inhibited at concentrations above 100 nM. Both phosphate and alpha-glycerol phosphate strongly inhibited phosphoenzyme formation and hydrolytic rate below 100 nM, but relieved the inhibition by substrate at higher concentrations probably by antagonizing substrate binding. A number of observations support the proposition that the phosphoenzyme is a necessary participant in catalysis. 1) The amount of phosphoenzyme measured during steady-state hydrolysis as a function of substrate concentration correlated with the velocity profile. 2) Rapid mixing experiments demonstrated that over a broad range of substrate concentrations phosphoenzyme formation was faster than the net rate of hydrolysis. 3) Both phosphate and alpha-glycerol phosphate inhibited the rate of phosphoenzyme formation and, at low substrate concentrations, reduced the steady-state phosphoenzyme levels. The latter correlated with inhibition of substrate hydrolysis. 4) Both phosphate and alpha-glycerol phosphate stimulate the rate of phosphoenzyme breakdown, consistent with their stimulation of substrate hydrolysis at high substrate concentrations. 5) The fractional rate of phosphoenzyme breakdown, which was pH and substrate dependent, multiplied by the amount of phosphoenzyme obtained in the steady state at that pH and substrate concentration approximated the observed rate of hydrolysis. We conclude that the phosphoenzyme is a reaction intermediate in the hepatic fructose-2,6-bisphosphatase reaction.  相似文献   

5.
Quinine activates the hydrolysis of phosphatidyl choline suspensions by phospholipase C (E.C. 3.1.4.3) obtained from Clostridium welchii. Low levels of calcium are an absolute requirement for this activation: Mg2+, Ba2+, Sr2+, and Zn2+ are ineffective. The induction period, or lag phase for this enzyme is dependent upon both calcium concentration and substrate interfacial surface area. At low concentrations (less then 50 muM) calcium ions affect the induction period but not the maximal rate of hydrolysis, whereas guinine predominantly affects the rate of hydrolysis by alterations in the surface charge carried by the substrate.  相似文献   

6.
All the equilibrium conformations of 34 analogues of acetylcholine (ACh) with the general formula R-C(O)O-Alk-N+(CH3)3 are calculated by the method of molecular mechanics. In the series R-C(O)O-(CH2)2-N+(CH3)3, a reliable correlation is found between the molecular volume of the substrate and the rate of its hydrolysis by acetylcholinesterase (AChE); the absence of such a correlation is demonstrated for butyrylcholinesterase (BChE). Theoretical conformational analysis confirms that the completely extended tt conformation of ACh is productive for the hydrolysis by AChE, which agrees with the results of X-ray analysis of AChE. AChE is shown to hydrolyze only those substrates that form equilibrium conformers compatible in the mutual arrangement of trimethylammonium group, carbonyl carbon, and carbonyl oxygen with the tt conformation of ACh; in this case, the rate of substrate hydrolysis depends on the total population of these conformers. A reliable correlation was found between the population of the semifolded (tg-) conformation of the choline moiety of substrate molecules and the rate of their BChE hydrolysis. In a series of CH3-C(O)O-Alk-N+(CH3)3, the rate of BChE hydrolysis is demonstrated to depend on the total population of conformations compatible in the mutual arrangement of functionally important atoms with the tg- conformation of ACh. The tg- conformation of ACh is concluded to be productive for BChE hydrolysis. Similar orientations of the substrate molecules relative to the catalytic triads of both AChE and BChE are proven to coincide upon the substrate productive sorption in their active sites. It is hypothesized that the sorption stage is rate-limiting in cholinesterase hydrolysis and the enzyme hydrolyzes the ACh molecule in its energetically favorable conformation.  相似文献   

7.
When concentrated dispersions of tripalmitin in Triton X-100 are added to reaction mixtures containing soluble beef liver lipase, the rate of hydrolysis of tripalmitin increases with incubation time. When the diluted substrate is aged at 37 degrees C for 3 hr before the addition of enzyme, the rate of hydrolysis is greater than the rate with freshly diluted dispersions and is constant for at least 2 hr. The reciprocal of the rate of hydrolysis is a complex function of the reciprocal of the substrate concentration when measured with freshly diluted substrate dispersions. A linear relationship between these reciprocals is obtained when measured with aged preparations of substrate. The rate and extent of increase of the velocity of hydrolysis of the aged substrate in relation to the velocity of hydrolysis of freshly diluted substrate are directly proportional to the substrate concentration and inversely proportional to the Triton X-100 concentration. The apparent V(max) of beef liver lipase for tripalmitin in diluted and aged dispersions is independent of the Triton X-100 concentration, while the apparent K(m) is inversely proportional to the Triton X-100 concentration. The apparent K(m) for tripalmitin complexes at zero Triton X-100 concentration was judged to be 7.5 x 10(-5) m. The molecular size of dispersion complexes does not change significantly as dispersions are aged. The spherical diameter of the complexes assessed by gel filtration techniques is in the order of 100 A.  相似文献   

8.
Mycobacterium convolutum strain NPA-1 can utilize n-propylamine (NPA), isopropylamine (IPA), and 1,3-propane diamine (PD) as sole source of carbon, nitrogen, and energy. Enzyme assays, fatty acid profiles, and 14CO2 incorporation experiments indicate that NPA is deaminated to propionate and further metabolized via the methylmalonyl succinate pathway, and IPA and PD were metabolized (after deamination) through a C2 + C1 cleavage. An inducible amine dehydrogenase was present in cell extracts after growth on the three amines. Polyacrylamide gel electrophoresis of cell extracts from NPA- and IPA-grown cells yielded one major band of amine dehydrogenase activity. When extracts of NPA-grown cells were assayed with NPA, IPA, or PD as substrate, the relative position of the major band on gel electrophoresis was equivalent. Similar results were obtained with extracts prepared from IPA-grown cells. Sephadex G-100 chromatography also indicated one major peak of activity. This suggests that one enzyme of broad specificity is involved in deamination of IPA, NPA, and PD. IPA-grown cells utilized NPA readily, whereas NPA-grown cells could not utilize IPA without lag. Since amine dehydrogenase activity was present in extracts of cells after growth on either substrate, this lag was probably due to the inability to transport IPA without an induction period. The molecular weight of the amine dehydrogenase was approximately 38,500 as determined by gel filtration.  相似文献   

9.
B Z Yu  O G Berg  M K Jain 《Biochemistry》1999,38(32):10449-10456
Hydrolysis of monodisperse short chain phosphatidylcholines, far below their critical micelle concentration, by phospholipase A2 (PLA2) and other interfacial enzymes is characterized. Results show that virtually all the observed hydrolysis by pancreatic and human inflammatory PLA2 occurs on surfaces of the reaction vessel or air bubbles. Conditions to eliminate such extraneous contributions at low substrate concentrations are established. Premicellar aggregates are apparently formed near the critical micelle concentration. The observation window at low substrate concentrations is used to obtain an upper limit estimate of the rate of hydrolysis through the monodisperse Michaelis complex. A limit estimate of <0.1 s-1 is obtained for the hydrolysis of monodisperse substrates by pig pancreatic phospholipase A2. These results show that the observed rate of hydrolysis of dihexanoyl- and diheptanoylphosphatidylcholines with pig pancreatic phospholipase A(2) through the monomer path is insignificant compared to the rate of >1000 s-1 seen at the saturating levels of the micellar substrate. These protocols should be useful for evaluating reactions catalyzed at vessel walls. Implications of these results for assays and models of interfacial activation of pancreatic PLA2 are discussed.  相似文献   

10.
Dihexanoylphosphatidylethanolamine (DiC6-PE) was prepared by phospholipase D catalyzed transphosphatidylation of dihexanoylphosphatidylcholine (DiC6-PC). Below the critical micellar concentration the pKa of the amino group is 9.4 +/- 0.05. The critical micellar concentration of the zwitterionic species is 5.3 +/- 0.2 mM, while that of the anionic species is 11.0 +/- 0.05 mM. Based on the pH dependence of the rate of hydroxide ion catalyzed hydrolysis, the second order rate constant for hydrolysis of the zwitterionic species is 0.70 +/- 0.021 s-1 M-1, while that for the anionic species is 0.040 +/- 0.011 s-1 M-1. The pH-dependence of phospholipase A2 catalyzed hydrolysis at substrate concentrations below the critical micellar concentration shows that the zwitterionic species is the preferred substrate, and the anionic species is either a competitive inhibitor of the hydrolysis of the zwitterionic species or poor substrate. DiC6-PE is hydrolyzed by C. adamanteus at about 1% the rate of DiC6-PC.  相似文献   

11.
1. N-Benzoyl-l-serine methyl ester was synthesized and evaluated as a substrate for bromelain (EC 3.4.22.4) and for papain (EC 3.4.22.2). 2. For the bromelain-catalysed hydrolysis at pH7.0, plots of [S(0)]/v(i) (initial substrate concn./initial velocity) versus [S(0)] are markedly curved, concave downwards. 3. Analysis by lattice nomography of a modifier kinetic mechanism in which the modifier is substrate reveals that concave-down [S(0)]/v(i) versus [S(0)] plots can arise when the ratio of the rate constants that characterize the breakdown of the binary (ES) and ternary (SES) complexes is either less than or greater than 1. In the latter case, there are severe restrictions on the values that may be taken by the ratio of the dissociation constants of the productive and non-productive binary complexes. 4. Concave-down [S(0)]/v(i) versus [S(0)] plots cannot arise from compulsory substrate activation. 5. Computational methods, based on function minimization, for determination of the apparent parameters that characterize a non-compulsory substrate-activated catalysis are described. 6. In an attempt to interpret the catalysis by bromelain of the hydrolysis of N-benzoyl-l-serine methyl ester in terms of substrate activation, the general substrate-activation model was simplified to one in which only one binary ES complex (that which gives rise directly to products) can form. 7. In terms of this model, the bromelain-catalysed hydrolysis of N-benzoyl-l-serine methyl ester at pH7.0, I=0.1 and 25 degrees C is characterized by K(m) (1) (the dissociation constant of ES)=1.22+/-0.73mm, k (the rate constant for the breakdown of ES to E+products, P)=1.57x10(-2)+/-0.32x10(-2)s(-1), K(a) (2) (the dissociation constant that characterizes the breakdown of SES to ES and S)=0.38+/-0.06m, and k' (the rate constant for the breakdown of SES to E+P+S)=0.45+/-0.04s(-1). 8. These parameters are compared with those in the literature that characterize the bromelain-catalysed hydrolysis of alpha-N-benzoyl-l-arginine ethyl ester and of alpha-N-benzoyl-l-arginine amide; K(m) (1) and k for the serine ester hydrolysis are somewhat similar to K(m) and k(cat.) for the arginine amide hydrolysis and K(as) and k' for the serine ester hydrolysis are somewhat similar to K(m) and k(cat.) for the arginine ester hydrolysis. 9. A previous interpretation of the inter-relationships of the values of k(cat.) and K(m) for the bromelain-catalysed hydrolysis of the arginine ester and amide substrates is discussed critically and an alternative interpretation involving substantial non-productive binding of the arginine amide substrate to bromelain is suggested. 10. The parameters for the bromelain-catalysed hydrolysis of the serine ester substrate are tentatively interpreted in terms of non-productive binding in the binary complex and a decrease of this type of binding by ternary complex-formation. 11. The Michaelis parameters for the papain-catalysed hydrolysis of the serine ester substrate (K(m)=52+/-4mm, k(cat.)=2.80+/-0.1s(-1) at pH7.0, I=0.1, 25.0 degrees C) are similar to those for the papain-catalysed hydrolysis of methyl hippurate. 12. Urea and guanidine hydrochloride at concentrations of 1m have only small effects on the kinetic parameters for the hydrolysis of the serine ester substrate catalysed by bromelain and by papain.  相似文献   

12.
Sodium chloride, phosphate buffer and ethanol were studied for their effect on butyryl cholinesterase hydrolysis rate of acetylcholine, acetylthiocholine, butyrylthiocholine and nonion substrate of indophenylacetate. The concentrations of 1.10(-2) = 1.10(-1) M of sodium chloride activated enzymatic hydrolysis of ion substrates at the concentrations lower than 1.10(-4) M but sodium chloride is a competitive inhibitor at higher concentrations. Phosphate buffer also activates substrates enzyme hydrolysis at the concentrations of 2.10(-4) M and lower, but it inhibits incompetitively the nonion substrate indophenylacetate hydrolysis. Ethanol activates butyrylthiocholine hydrolysis and is a competitive inhibitor in acetylthiocholine and indophenylacetate hydrolysis. The observed effects are discussed on the assumption of two forms of butyrylcholinesterase E' and E" existence. These two forms are determined by different kinetic parameters and are in equilibrium.  相似文献   

13.
Homogenates of rat neostriatum hydrolysed cGMP faster than cAMP at both high (100 microM) and low (1 microM) substrate concentrations, although the hydrolysis of both nucleotides exhibited similar kinetic properties. Kinetic analysis of the effect of substrate concentration on the rate of cAMP and cGMP hydrolysis gave results characteristic of a negatively cooperative enzyme species, with two apparent Km's for each nucleotide. The ratio between the Vmax of the high Km form and the Vmax of the low Km form was similar in various subcellular fractions of neostriatal tissue, in a preparation of synaptic membranes from whole brain, and in homogenates of other brain regions, including both neural-rich and glial-rich tissues. In homogenates of neostriatum cAMP could almost completely block cGMP hydrolysis and vice versa. The kinetics of this inhibition were competitive at low (1 microM) substrate concentrations, and non-competitive at high (100 microM) substrate concentrations. Various phosphodiesterase inhibitors failed to preferentially inhibit the hydrolysis of either nucleotide at high or low nucleotide concentrations. Preliminary studies of the effect of a Ca(2+)-dependent endogenous activator preparation on the hydrolysis of cyclic nucleotides in homogenates of rat neostriatum showed a specific activation of cGMP hydrolysis at low nucleotide concentrations. The rate of cGMP hydrolysis at 1 microM substrate concentration was doubled in the presence of the activator preparation and 100 microM-CaCl2, while cGMP hydrolysis at 100 microM or cAMP hydrolysis at both 1 microM and 100 microM remained unaffected. These observations raise the possibility that cAMP and cGMP may be hydrolysed by the same enzyme in rat neostriatum, and that an endogenous activating factor may determine the relative affinities of the enzyme for the two nucleotides.  相似文献   

14.
采用三因素二次通用旋转设计和体外检测法,对胰蛋白酶水解β-乳球蛋白获得ACE抑制肽的条件进行优化。结果表明,底物浓度(X1)、温度(X2)、酶与底物的质量比(X3)对ACE抑制率的影响回归方程为:Y=50.62-2.33X1-1.97X2+5.81 X3-3.36X2X3-6.56X22-1.96X32,胰蛋白酶水解β-乳球蛋白获得ACE抑制肽的最优水解条件为:底物质量浓度为60 g/L,水解温度30℃,酶与底物的质量比为5.5%,水解时间6 h,水解产物对ACE抑制活性最大抑制率为53.86%。  相似文献   

15.
A new vanadyl complex, bis(5-iodopicolinato)oxovanadium(IV), VO(IPA)2, with a VO(N2O2) coordination mode, was prepared by mixing 5-iodopicolinic acid and VOSO4 at pH 5, with the structure characterized by electronic absorption, IR, and EPR spectra. Introduction of the halogen atom on to the ligand enhanced the in vitro insulinomimetic activity (IC50 = 0.45 mM) compared with that of bis(picolinato)oxovanadium(IV) (IC50 = 0.59 mM). The hyperglycemia of streptozotocin-induced insulin-dependent diabetic rats was normalized when VO(IPA)2 was given by daily intraperitoneal injection. The normoglycemic effect continued for more than 14 days after the end of treatment. To understand the insulinomimetic action of VO(IPA)2, the organ distribution of vanadium and the blood disposition of vanadyl species were investigated. In diabetic rats treated with VO(IPA)2, vanadium was distributed in almost all tissues examined, especially in bone, indicating that the action of vanadium is not peripheral. Vanadyl concentrations in the blood of normal rats given VO(IPA)2 remain significantly higher and longer than those given other complexes because of its slower clearance rate. VO(IPA)2 binds with the membrane of erythrocytes, probably owing to its high hydrophobicity in addition to its binding with serum albumin. The longer residence of vanadyl species shows the higher normoglyceric effects of VO(IPA)2 among three complexes with the VO(N2O2) coordination mode. On the basis of these results, VO(IPA)2 is indicated to be a preferred agent to treat insulin-dependent diabetes mellitus in experimental animals.  相似文献   

16.
1. The velocity of hydrolysis of gelatin by trypsin increases more slowly than the gelatin concentration and finally becomes nearly independent of the gelatin concentration. The relative velocity of hydrolysis of any two substrate concentrations is independent of the quantity of enzyme used to make the comparison. 2. The rate of hydrolysis is independent of the viscosity of the solution. 3. The percentage retardation of the rate of hydrolysis by inhibiting substances, is independent of the substrate concentration. 4. There is experimental evidence that the enzyme and inhibiting substance are combined to form a widely dissociated compound. 5. If the substrate were also combined with the enzyme, an increase in the substrate concentration should affect the equilibrium between the enzyme and the inhibiting substance. This is not the case. 6. The rate of digestion of a mixture of casein and gelatin is equal to the sum of the rates of hydrolysis of the two substances alone, as it should be if the rate is proportional to the concentration of free enzyme. This contradicts the saturation hypothesis. 7. If the reaction is followed by determining directly the change in the substrate concentration, it is found that this change agrees with the law of mass action; i.e., the rate of digestion is proportional to the substrate concentration.  相似文献   

17.
The hydrolysis of small unilamellar vesicles made of dipalmitoylphosphatidylcoline by pancreatic phospholipase A2 has been studied under various conditions of temperature and enzyme and substrate concentration using the following three different experimental protocols. When the enzyme was added to the substrate vesicles after being separately adjusted to the temperature of the experiments hydrolysis occurred instantaneously only in the temperature range where the lipid is known to exist in its gel phase, while above the transition range no hydrolysis occurred. Within the transition range, the time course of hydrolysis was characterized by initial very slow rate of hydrolysis (latency phase) followed by an abrupt increase in the rate after a time tau, which is a complex function of temperature and enzyme to substrate ratio. When an enzyme-substrate mixture was first preincubated below Tm and then temperature jumped to a temperature above or within the transition range, the latency phase was markedly shortened. When the temperature jump was to the transition range, this effect is observed even if Ca2+ is absent in the preincubation mixture. However, instantaneous hydrolysis was observed upon temperature jumping the mixture to a temperature high above Tm only if Ca2+ was present in the preincubation medium. In temperature-scanning experiments, hydrolysis was followed while changing the temperature of the enzyme-substrate mixture continuously. Heating an enzyme-substrate mixture from room temperature resulted in an abrupt onset of hydrolysis when the transition range was approached. These results lead us to conclude that two distinctly different steps precede rapid hydrolysis of dipalmitoylphosphatidylcholine small unilamellar vesicles by pancreatic phospholipase A2: a Ca2+-independent binding of the enzyme to the substrate vesicles, which for chemically pure bilayers occurs best in the gel phase. This step is followed by a Ca2+-dependent activation of the initially formed enzyme-substrate complex. The latter step only occurs under conditions where the bilayer possesses packing irregularities and probably involves a reorganization of the enzyme-substrate complex. At least one of these two steps appears to involve enzyme-enzyme interaction.  相似文献   

18.
The substrate specificities of the phospholipase and triglyceridase activities of purified rat liver hepatic lipase were compared using lipid monolayers so that the substrates were presented to the enzyme in a controlled physical state. The rate of hydrolysis of 14C-labeled lipid at constant surface pressure in the presence of hepatic lipase and fatty acid-free bovine serum albumin at 33 degrees C was determined by monitoring the decrease of surface radioactivity. In monolayers of sphingomyelin/cholesterol (2:1, mol/mol) containing either 1 mol% triacylglycerol, 1 mol% phosphatidylethanolamine, or 10 and 20 mol% phosphatidylcholine, hepatic lipase clearly showed a preference for unsaturated over saturated lipids. In addition, with a sphingomyelin/cholesterol (2:1) monolayer containing 1 mol% of lipid substrate, hepatic lipase showed the following preference: triolein = dioleoylphosphatidylethanolamine much greater than dioleoylphosphatidylcholine; the respective rates of hydrolysis were 15.3 +/- 1.2, 14.9 +/- 0.8, and 0.5 +/- 0.1 mumol fatty acid produced/h per mg hepatic lipase. Overall, it appears that when comparing rates of hydrolysis of molecules within a given lipid class, hydrocarbon chain interactions are important. However, when comparing different lipid classes such as phosphatidylcholines and phosphatidylethanolamines, it is apparent that the polar group has a significant influence on the rate of hydrolysis. The rate of [14C]triolein hydrolysis, when mixed at surface concentrations of up to 2 mol% in a sphingomyelin/cholesterol (2:1) monolayer, was significantly faster than when triolein was present in a 1-oleyl-2-palmitylphosphatidylcholine monolayer; the rates of hydrolysis were 47.7 +/- 5.4 and 8.9 +/- 0.8 mumol fatty acid produced/h per mg hepatic lipase, respectively. The monolayer physical state and the miscibility of the substrate in the inert matrix influence the presentation of the substrate to the enzyme, thereby affecting the hydrolysis rate.  相似文献   

19.
In an effort to better understand the role of the substrate in the rapid fall off in the rate of enzymatic hydrolysis of cellulose with conversion, substrate reactivity was measured as a function of conversion. These measurements were made by interrupting the hydrolysis of pretreated wood at various degrees of conversion; and, after boiling and washing, restarting the hydrolysis in fresh buffer with fresh enzyme. The comparison of the restart rate per enzyme adsorbed with the initial rate per enzyme adsorbed, both extrapolated back to zero conversion, provides a measurement of the substrate reactivity without the complications of product inhibition or cellulase inactivation. The results indicate that the substrate reactivity falls only modestly as conversion increases. However, the restart rate is still higher than the rate of the uninterrupted hydrolysis, particularly at high conversion. Hence we conclude that the loss of substrate reactivity is not the principal cause for the long residence time required for complete conversion. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 56: 650-655, 1997.  相似文献   

20.
The extent of substrate enantioselectivity and regioselectivity of a series of aliphatic 2,3-dialkyl- and trialkylsubstituted oxiranes in their in vitro epoxide-hydrolase-catalyzed hydrolysis depends on the size of the alkyl residues and on the substitution pattern of the oxirane ring. The enzyme-catalyzed hydrolysis of cis-oxiranes, containing at least one methyl substituent, shows complete or nearly complete substrate enantioselectivity and regioselectivity with nucleophilic attack by water occurring with inversion of configuration at the methylsubstituted ring carbon atom of (S)-configuration. In the hydrolysis of the isomeric trans-oxiranes, both enantiomers are metabolized with a higher rate for the (2S;3S)-enantiomer. The conversion of trimethyloxirane occurs with high substrate enantioselectivity in favor of the (S)-enantiomer and with complete regioselectivity at the monomethylsubstituted ring carbon atom. The differentiation of the enantiotopic ring carbon atoms (product enantioselectivity) in the smallest aliphatic meso-oxirane, cis-2,3-dimethyloxirane, leads to (2R;3R)-butane-2,3-diol with ee = 86%. cis-2-Ethyl-3-propyloxirane, possessing alkyl residues larger than methyl, represents an extremely poor substrate in the epoxide-hydrolase-catalyzed hydrolysis process.  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