首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The enantioselectivities of subtilisin Carlsberg and Rhizomucor miehei lipase in organic solvents are found to strongly depend on the method by which the enzymes are prepared. For the transesterification between sec-phenethyl alcohol and vinyl butyrate in dioxane at 7 degrees C, the enantioselectivity of subtilisin precipitated with isopropanol is more than twice that of the enzyme prepared by lyophilization from aqueous buffer. Furthermore, the temperature dependence of the enantioselectivity is influenced by the mode of enzyme preparation. For example, in the aforementioned process the enantioselectivities of subtilisin lyophilized from aqueous buffer and crosslinked subtilisin crystals increase when the temperature is raised from 7 to 45 degrees C. In contrast, the enantioselectivities decrease with temperature for the enzyme precipitated from aqueous solution with acetone or isopropanol and for the enzymatic hydrolysis in water. The temperature dependence of the enantioselectivity of subtilisin lyophilized from buffer is markedly affected by the solvent: In acetonitrile and nitromethane the enzyme is more enantioselective at higher temperatures, while negligible temperature effects have been found in tetrahydrofuran and pyridine. Lyophilized lipase exhibits striking temperature dependencies of its enantioselectivity in dioxane, acetonitrile, and nitromethane, while showing almost none in pyridine, triethylamine, and tetrahydrofuran. The results underscore the importance of the mode of enzyme recovery on enantioselectivity and its temperature dependence in enzymatic reactions in organic solvents (in contrast to those in water). (c) 1996 John Wiley & Sons, Inc.  相似文献   

2.
The addition of simple inorganic salts to aqueous enzyme solutions prior to lyophilization results in a dramatic activation of the dried powder in organic media relative to enzyme with no added salt. Activation of both the serine protease subtilisin Carlsberg and lipase from Mucor javanicus resulting from lyophilization in the presence of KCl was highly sensitive to the lyophilization time and water content of the sample. Specifically, for a preparation containing 98% (w/w) KCl, 1% (w/w) phosphate buffer, and 1% (w/w) enzyme, varying the lyophilization time showed a direct correlation between water content and activity up to an optimum, beyond which the activity decreased with increasing lyophilization time. The catalytic efficiency in hexane varied as much as 13-fold for subtilisin Carlsberg and 11-fold for lipase depending on the lyophilization time. This dependence was apparently a consequence of including the salt, as a similar result was not observed for the enzyme freeze-dried without KCl. In the case of subtilisin Carlsberg, the salt-induced optimum value of kcat/Km for transesterification in hexane was over 20,000-fold higher than that for salt-free enzyme, a substantial improvement over the previously reported enhancement of 3750-fold (Khmelnitsky, 1994). As was found previously for pure enzyme, the salt-activated enzyme exhibited greatest activity when lyophilized from a solution of pH equal to the pH for optimal activity in water. The active-site content of the lyophilized enzyme samples also depended upon lyophilization time and inclusion of salt, with opposite trends in this dependence observed for the solvents hexane and tetrahydrofuran. Finally, substrate selectivity experiments suggested that mechanism(s) other than selective partitioning of substrate into the enzyme-salt matrix are responsible for salt-induced activation of enzymes in organic solvents.  相似文献   

3.
A method to improve the enantioselectivity of lipase-catalyzed kinetic resolution (KR) of trans-2-phenyl-cyclopropane-1-carboxylic acid derivatives in water–acetone solution is presented. Two different approaches were compared: enzyme-catalyzed esterification and enzymatic hydrolysis of the target ester. A substantial influence of enzyme type, ethoxy group donor, and solvent on conversion and enantioselectivity of the enzymatic esterification was noted. While enzymatic esterification proceeds with poor enantioselectivity, the hydrolysis of target ester proceeds efficiently. Studies on the influence of cosolvent used for the enzymatic hydrolysis reaction showed that kinetic resolution can be performed in acetone and water buffer mixture predominantly containing organic solvent. Any change in organic solvent content resulted in a substantial decrease in enantioselectivity from almost E = 150 to less than 5.  相似文献   

4.
Lyophilized enzyme powder is often used in organic solvents. However, the enzymatic activity decreases during the reaction process. In the present study, the relation between structural stability and enzymatic activity in an organic solvent was investigated. 13C cross-polarization magic angle spinning NMR spectroscopy was used to determine the secondary structure of lyophilized papain in the solid-state. Deconvolution of the peaks of the backbone carbonyl carbons suggested that the proportion of beta-sheet conformation increased after lyophilization from a phosphate buffer solution. The esterification of N-benzyloxycarbonyl phenylalanylalanine amide was attempted using the lyophilized papain as a catalyst in anhydrous 1-propanol. The yield of ester was 46.1% after 48 h at 50 degrees C, but this reaction slowed remarkably after 48 h. When the lyophilized papain was suspended in anhydrous 1-propanol for 7 days without the substrate, the proportion of beta-sheet conformation was further increased and the suspended papain had no activity. These results suggest that the increase in beta-sheet conformation caused inactivation of papain. The increase in beta-sheet conformation caused by both lyophilization and suspension in propanol was found, which was related to a decrease in enzymatic activity.  相似文献   

5.
The kinetic parameters, kcat and KM, for the hydrolysis of N-α-tosyl-L-arginine methyl ester (1, TAME) by the wild-type subtilisins Carlsberg and BPN′ as well as the BPN′ mutants Glyl66Ser, GLyl66Asn, and Met222Phe, were determined in the presence of 5 and 15% (v/v) of a selection of water-soluble organic solvents. The goals were to compare and evaluate the solvent effects with a view to expanding their use in organic synthetic applications of the WT and mutant subtilisins. The results showed that subtilisin BPN′ and its mutants were much less affected by organic solvents than subtilisin Carlsberg. The BPN′ mutant Met222Phe demonstrated the greatest resistance to cosolvent inactivation, making it a particularly attractive mutant for peptide synthesis. Dimethyl sulfoxide, acetone, and branched alcohols were found to be the most benign solvents, whereas dioxane, THF, and N-methyl-2-pyrrolidinone seriously reduced catalytic activities, even at low concentrations. The results parallel the solvent-effect data available for other proteinases, including α-chymotrypsin.  相似文献   

6.
Summary Subtilisin from Bacillus subtilis was modified with polyethylene glycol (PEG), or adsorbed either on celite or porous glass, or directly used as a suspended powder to catalyse peptide synthesis and transesterification reactions in organic solvents. The rather low yield of peptide synthesis probably resulted from the enzyme tendency to catalyse hydrolysis and transesterification side reactions. The kinetics of transesterification catalysed by PEG-subtilisin was consistent with a ping-pong mechanism modified by a hydrolytic branch. Initial rates of transesterification were found to be dependent on alcohol and organic base concentrations in the reaction mixture. The high affinity of benzyloxycarbonyl-l-serine-methyl ester for the enzyme indicated that a change in substrate specificity of subtilisin occurred in organic phase. The 50-fold increase in the rate of synthesis of benzyloxycarbonyl-l-serine-l-phenylalanine amide which was observed when PEG-subtilisin was used instead of immobilized or powdered enzyme, suggested that a higher flexibility of the polypeptide chain modified by the covalent attachment of a number of soluble PEG moieties occurred in organic solvents. This also resulted in a lower stability of PEG-subtilisin at high temperature.Offprint requests to: A. Puigserver  相似文献   

7.
We developed an FTIR (Fourier transform infrared) methodology for quantitatively assessing the secondary structure of proteins suspended in nonaqueous media. This methodology was used to measure the percentages of alpha-helices and beta-sheets of subtilisin Carlsberg, prepared under different conditions, placed in various organic solvents. The title question was addressed with respect to some instances of markedly influencing the subtilisin activity in organic solvents reported in the literature. It is concluded that the mechanism of subtilisin activation by KCl and N-Ac-L-Phe-NH(2) present in the aqueous solution of the enzyme prior to lyophilization may be due to their preservation of the secondary structure, otherwise altered by the dehydration. Likewise, subtilisin inactivation in the protein-dissolving solvent DMSO (dimethyl sulfoxide) is likely caused by enzyme denaturation (the loss of both alpha-helices and beta-sheets). On the other hand, some other ligands, as well as protein nondissolving organic solvents, while greatly affecting the subtilisin activity, have little effect on its secondary structure, thus ruling out the causal relationship between the two. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 53: 351-362, 1997.  相似文献   

8.
Enzymatic synthesis of propyl gallate in organic solvent was studied using cell-associated tannase (EC 3.1.1.20) of Bacillus massiliensis. Lyophilized biomass showing tannase activity was used as the biocatalyst. The effects of solvent, surfactant treatment, and bioimprinting on the propyl gallate synthesis were studied and subsequently optimized. Among various solvents, benzene followed by hexane was found to be the most favorable. Treatment of the biocatalyst with Triton X-100 at a lower concentration (0.2% w/v), before lyophilization, increased the propyl gallate yield by 24.5% compared to the untreated biocatalyst. The biocatalyst was imprinted with various concentrations of gallic acid and tannic acid. Biocatalyst imprinted with tannic acid showed 50% enhancement in the propyl gallate yield compared to the non-imprinted biocatalyst.  相似文献   

9.
The catalytic activity of α-chymotrypsin in the enzymatic peptide synthesis of N-acetyl-l-tryptophan ethyl ester with glycyl glycinamide was examined in ionic liquids and organic solvents. The water content in 1-ethyl-3-methylimidazolium bis(fluorosulfonyl)imide ([emim][FSI]) affected the initial rates of peptide synthesis and hydrolysis. The activity of α-chymotrypsin was influenced by a kind of anions consisting of the same cation, [emim], when an ionic liquid was used as a solvent. The initial rate of peptide synthesis was improved 16-fold by changing from an organic solvent, acetonitrile, to an ionic liquid, [emim][FSI], at 25 °C. The activity of α-chymotrypsin in the peptide synthesis in [emim][FSI] was 17 times greater than that in acetonitrile at 60 °C, although the activity of α-chymotrypsin in the peptide synthesis gradually decreased with an increase in reaction temperature in [emim][FSI], similar to organic solvents. Moreover, α-chymotrypsin exhibited activity in [emim][FSI] and [emim][PF6] at 80 °C.  相似文献   

10.
The serine protease subtilisin BPN' is a useful catalyst for peptide synthesis when dissolved in high concentrations of a water-miscible organic co-solvent such as N,N-dimethylformamide (DMF). However, in 50% DMF, the k(cat) for amide hydrolysis is two orders of magnitude lower than in aqueous solution. Surprisingly, the k(cat) for ester hydrolysis is unchanged in 50% DMF. To explain this alteration in activity, the structure of subtilisin 8397+1 was determined in 20, 35, and 50% (v/v) DMF to 1.8 A resolution. In 50% DMF, the imidazole ring of His64, the central residue of the catalytic triad, has rotated approximately 180 degrees around the Cbeta-Cgamma bond. Two new water molecules in the active site stabilize the rotated conformation. This rotation places His64 in an unfavorable geometry to interact with the other members of the catalytic triad, Ser221 and Asp32. NMR experiments confirm that the characteristic resonance due to the low barrier hydrogen bond between the His64 and Asp32 is absent in 50% DMF. These experiments provide a clear structural basis for the change in activity of serine proteases in organic co-solvents.  相似文献   

11.
The hydrolysis of p-nitrophenyl acetate is catalyzed by imidazole, free in solution or as the side chain in poly(His-Ala-Glu). This is based on the observations that the reaction is first order in ester and first order in nonprotonated imidazole. Catalysis of p-nitrophenyl acetate hydrolysis is dependent on solvent conditions. The effect of low concentrations of ethanol, dioxane, and trifluoroethanol were investigated. As the concentration of organic solvent is increased, the second-order rate constant for imidazole catalysis decreases. The decrease, however, is greater for imidazole than for poly(His-Ala-Glu). In 2% trifluoroethanol/water solution, free imidazole has twice the catalytic activity of polymeric imidazole, while in 40% trifluoroethanol/water they have equal activity. Since under the latter solvent conditions poly(His-Ala-Glu) is partially α-helical, the relative improvement in polymeric–imidazole catalysis may be attributed to imidazole hydrogen-bonded to a carboxylate ion. With this assumption the carboxylate–imidazole hydrogen-bonded system has been calculated to have three times the base catalytic activity of imidazole.  相似文献   

12.
Bacillus subtilis IBTC-3 subtilisin was purified by gel filtration on Sephadex G 75 and affinity chromatography on bacitracin-CNBr-Sepharose 4B and characterized. Its molecular mass of 27 kDa was determined by SDS-PAGE, and isoelectric pH of 8.4 by chromatofocusing. FT-Raman and FT-IR spectroscopy studies revealed fragments with alpha-helix and irregular secondary structures within the polypeptide chain. The beta-sheet conformation was observed only in second-derivatives of FT-RS and FT-IR spectra, in the range of the amide II, III, and I bands. Tyr residues were shown to be hydrogen bonded and CSCH(3) groups adopted two conformations (P(H)-T and P(C)-G conformers). Kinetic properties of B. subtilis IBTC-3 subtilisin in hydrolysis of ethyl esters of amino acid derivatives were compared with that of alkaline peptidase from Bacillus alcalophilus PB92. The first enzyme displayed the highest affinity for NAc-Phe-OEt, both in hydrolysis (K(m) of 0.22 mM) and in synthesis (K(m) of 0.85 mM), whereas PB92 peptidase preferred Tyr derivatives (NAc-Tyr-OEt, K(m) of 0.043 and 0.75 mM, respectively). In contrast to the latter enzyme, B. subtilis IBTC-3 subtilisin catalyzed hydrolysis and synthesis of Bz-Arg-OEt.  相似文献   

13.
The release rate of vanillylamine from its hydrochloride salt was the limiting step in the lipase-catalysed synthesis of olvanil, a capsaicin analogue amide, in organic solvents. When the tertiary amine base concentration (N,N-diisopropylethylamine) was increased from 20 mM to 360 mM, the initial rate of amide synthesis increased proportionally. At a 12 molar excess of N,N-diisopropylethylamine and 30 min of preincubation, both the initial rate and total conversion were the same as those with free vanillylamine (80% conversion in 20 h). This result was independent of the organic solvent used. It is also shown that N,N-diisopropylethylamine does not enhance lipase activity.  相似文献   

14.
The activity and stability of native subtilisin Karlsberg and subtilisin 72 and their complexes with sodium dodecyl sulfate (SDS) in organic solvents were studied. The kinetic constants of the hydrolysis of specific chromogenic peptide substrates Z-Ala-Ala-Leu-pNA and Glp-Ala-Ala-Leu-pNA by the subtilisins were determined. It was found that the subtilisin Karlsberg complex with SDS in anhydrous organic solvents is an effective catalyst of peptide synthesis with multifunctional amino acids in positions P 1 and P 1 (Glu, Arg, and Asp) containing unprotected side ionogenic groups.  相似文献   

15.
Thermolysin-catalyzed peptide synthesis using N-benzyloxycarbonyl)-l-phenylalanine (Z-Phe) and l-phenylalanine methyl ester (Phe-OMe) as substrates was done mainly in a water-organic one phase solvent system. The organic solvent content used was less than the saturation concentration in buffer. With organic solvents with high log P values, the enzymatic activity increased as the organic solvent content increased; but further increases in the organic solvent content decreased the enzymatic activity, showing an “organic activity” profile. On the other hand, with organic solvents of low log P values, the enzymatic reaction was inhibited even by the initial addition of organic solvents. When a correlation between maximum activities and logP values or Hildebrand solubility parameters was investigated, a linear correlation was obtained among the same category of organic solvents, but not between categories. This suggests that the direct effect of organic solvents on the microenvironment of the enzyme largely depends on the molecular structure of the solvents.  相似文献   

16.
The employment of enzymes as catalysts within organic media has traditionally been hampered by the reduced enzymatic activities when compared to catalysis in aqueous solution. Although several complementary hypotheses have provided mechanistic insights into the causes of diminished activity, further development of biocatalysts would greatly benefit from effective chemical strategies (e.g., PEGylation) to ameliorate this event. Herein we explore the effects of altering the solvent composition from aqueous buffer to 1,4-dioxane on structural, dynamical, and catalytic properties of the model enzyme subtilisin Carlsberg (SBc). Furthermore, we also investigate the effects of dissolving the enzyme in 1,4-dioxane through chemical modification with poly(ethylene)-glycol (PEG, M(W) = 20 kDa) on these enzyme properties. In 1,4-dioxane a 10(4)-fold decrease in the enzyme's catalytic activity was observed for the hydrolysis reaction of vinyl butyrate with D(2)O and a 50% decrease in enzyme structural dynamics as evidenced by reduced amide H/D exchange kinetics occurred. Attaching increasing amounts of PEG to the enzyme reversed some of the activity loss. Evaluation of the structural dynamic behavior of the PEGylated enzyme within the organic solvent revealed an increase in structural dynamics at increased PEGylation. Correlation analysis between the catalytic and structural dynamic parameters revealed that the enzyme's catalytic activity and enantioselectivity depended on the changes in protein structural dynamics within 1,4-dioxane. These results demonstrate the importance of protein structural dynamics towards regulating the catalytic behavior of enzymes within organic media.  相似文献   

17.
Summary An enzymatic method for synthesis of monoglyceride from 1,2-isopropylidene glycerol and n-3 polyunsaturated fatty acid concentrate was investigated in organic solvent. Optimal reaction conditions for monoglyceride synthesis by lipase were established. Lipase IM-60 fromMucor miehei produced yields of monoglyceride of up to 80% in this system. The resultant monoglyceride contained 76.2% n-3 polyunsaturated fatty acid (eicosapentaenoic acid 43.3%; docosahexaenoic acid, 32.7%). Isooctane and hexane were suitable organic solvents for monoglyceride synthesis and optimal initial water content was 2.5%. Lipase IM-60 was relatively stable in organic solvent and is easily recovered for reuse.  相似文献   

18.
The kinetics of the -chymotrypsin catalysed transesterification of N-acetyl-l-phenylalanine ethyl ester with 1-butanol and the competing hydrolysis were evaluated at fixed water activity in two ionic liquids and two non-ionic organic solvents. In most respects the four solvents behaved similarly. However, at a water activity of 0.33, higher catalytic activity was observed in the ionic liquid, 1-butyl-3-methylimidazolium bis[(trifluoromethyl)sulfonyl]amide, than in the other solvents, and at aw=0.11 catalysis was only observed in this solvent.  相似文献   

19.
The aim of this work was to purify and characterize the extra-cellular leucine amino peptidase (LAP) from Streptomyces gedanensis and also study its applications for protein hydrolysis. The enzyme was purified to homogeneity by ammonium sulfate fractionation and sequential chromatography steps. LAP appeared to be a monomeric enzyme with a molecular weight of ~75 kDa determined by sodium dodecyl sulfate poly acryl amide gel electrophoresis (SDS-PAGE). The enzyme preferentially hydrolyzed leucine p-nitroanilide followed by Met, Phe, Lys and Arg derivatives. Kinetic studies on the purified enzyme confirmed that it can hydrolyze peptide as well as ester substrates at comparable rates. This amino peptidase was highly resistant to different concentrations of various organic solvents. The characteristics of this amino peptidase, including thermo stability, organic solvent resistance, its activity against various substrates, and also it showed esterase and peptidase activity at comparable rates; identified this amino peptidase as a novel one. The specificity towards aromatic and hydrophobic amino acid residues, the solvent-resistance and thermo stability make this amino peptidase could offer interesting possibilities for various industrial applications including debittering of protein hydrolysates, peptide and ester synthesis.  相似文献   

20.
Summary Release of autolysin during the late exponential growth phase of Clostridium acetobutylicum resulted in early lysis of the culture and reduction of solvent formation. A simple and effective way of reducing autolysin activity and increasing solvent production is partial inhibition of protein synthesis with chloramphenicol (CAP). The extracellular autolytic activity in the culture, determined by following loss of turbidity of washed clostridial cells in 0.04m sodium phosphate buffer at 37° C, was decreased by 40% after CAP treatment. This caused an extension of cell viability by 12 h and an increase in butanol production by 30%. The optimal time of CAP addition was 12 h of incubation, and the optimal antibiotic concentration was 120 g/ml. The effects of CAP on the fermentation are due to the inhibition of protein synthesis leading to a decrease in autolysin level in the culture. The results obtained provide economic advantages for industrial production of solvents by minimizing autolysin activity and maximizing solvent yield during the critical solvent-producing phase. Correspondence to: R. W. Traxler  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