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 331 毫秒
1.
Zeolite molecular sieves very commonly are used as in situ drying agents in reaction mixtures of enzymes in nonaqueous media. They often affect enzyme behavior, and this has been interpreted in terms of altered hydration. Here, we show that zeolites can also have dramatic acid-base effects on enzymes in low water media, resulting from their cation-exchange ability. Initial rates of transesterification catalyzed by cross-linked crystals of subtilisin were compared in supercritical ethane, hexane, and acetonitrile with water activity fixed by pre-equilibration. Addition of zeolite NaA (4 A powder) still caused remarkable rate enhancements (up to 20-fold), despite the separate control of hydration. In the presence of excess of an alternative solid-state acid-base buffer, however, zeolite addition had no effect. The more commonly used Merck molecular sieves (type 3 A beads) had similar but somewhat smaller effects. All zeolites have ion-exchange ability and can exchange H+ for cations such as Na+ and K+. These exchanges will tend to affect the protonation state of acidic groups in the protein and, hence, enzymatic activity. Zeolites pre-equilibrated in aqueous suspensions of varying pH-pNa gave very different enzyme activities. Their differing basicities were demonstrated directly by equilibration with an indicator dissolved in toluene. The potential of zeolites as acid-base buffers for low-water media is discussed, and their ability to overcome pH memory is demonstrated.  相似文献   

2.
Enzyme performance is often impaired in supercritical carbon dioxide. We were able to enhance enzyme activity in this medium via changes in acid-base conditions by using ion-exchange materials (solid H(+)/Na(+) buffer pairs and a zeolite), which were selected on the basis of the response of an organosoluble acid-base indicator. The concentration of ion-exchange materials had an important effect on the catalytic activity of subtilisin Carlsberg cross-linked enzyme crystals (CLECs), and this was related to the protonation and hydration states of the enzyme. The buffer Na(2)CO(3)/NaHCO(3) gave the highest enhancement in enzyme activity (by a factor of 54), probably as a result of its high basicity and capacity to counteract the deleterious effect of carbonic acid to a greater extent than the other materials tested.  相似文献   

3.
We have previously demonstrated that a solid-state buffer could be successfully used to control the ionization state of subtilisin Carlsberg cross-linked microcrystals (CLECs) suspended in supercritical ethane (sc-ethane) in the presence of acid–base active species such as salt hydrates and zeolite molecular sieves. Here we studied the effect of six zwitterionic proton/sodium (pH–pNa) solid-state acid–base buffers on the catalytic activity of subtilisin CLECs in sc-ethane at high and low water activity (aW). CLECs were strongly activated by increasing aW. At high aW, and despite the high hydrolysis rates, transesterification activities were still about one order of magnitude higher than those observed at lower aW. This is in contradiction with what was previously reported in the absence of acid–base control and supports the hypothesis that the poor catalytic performance of subtilisin CLECs at high aW observed in those studies was due to the inhibitory effect of the hydrolytic by-product, rather than to the competition of water with propanol for the acyl-enzyme intermediate. Although the catalytic activity of subtilisin showed a general positive correlation with the aqueous pKa of the acid–base buffers tested here, our results also show that as expected, the acid–base behavior of the buffers in nonaqueous media is more complex than what can be predicted from aqueous-based parameters alone. This work further confirms the usefulness of solid-state acid–base buffers in supercritical biocatalysis but highlights the need for further research on the topic.  相似文献   

4.
The stability of the serine proteases from Bacillus amyloliquefaciens (subtillisin BPN') and Bacillus licheniformis (subtilisin Carlsberg) was investigated in various anhydrous solvents at 45 degrees C. The half-life of subtilisin BPN' in dimethyl-formamide dramatically depends on the pH of the aqueous solutions from which the enzyme was lyophilized, increasing from 48 min to 20 h when the pH is raised from 6.0 to 7.9. Both subtilisins exhibited substantial inactivation during multihour incubations in tert-amyl alcohol and acetonitrile when enzymatic activities were also measured in these solvents; however, when the enzymes were assayed in water instead, hardly any loss of activity was detected. This surprising difference appears to stem from the partitioning of the bound water essential for catalytic activity from the enzymes into the solvents. When assayed in organic solvents, this time-dependent stripping of water results in decay of enzymatic activity; however, when assayed in water, where the dehydrated subtilisins can undergo rehydration thereby recovering catalytic activity, little inactivation is observed. In agreement with this hypothesis, the addition of small quantities of water tert-amyl alcohol stabilized the subtilisins in it even when enzymatic activity was measured in the nonaqueous solvent. Ester substrates (vinyl butyrate and trichloroethyl butyrate) greatly enhanced the stability of both subtilisins in organic solvents possibly because of the formation of the acyl-enzymes.  相似文献   

5.
Fourier-transform infrared (FTIR) spectroscopy has been used to quantify the alpha-helix and beta-sheet contents of subtilisin Carlsberg dissolved in several nonaqueous, as well as aqueous, solvents. Independently, the catalytic activity of the enzyme has been measured in the same solvents. While our previous FTIR studies revealed no connection between the secondary structure and enzymatic activity for subtilisin suspended in various organic solvents, a very different situation is observed herein for the dissolved enzyme. Specifically, if either the alpha-helix or beta-sheet content in a given solvent is higher or lower than in water, no appreciable enzymatic catalysis is observed. Conversely, when the secondary structure of subtilisin dissolved in a given nonaqueous solvent is similar to that in water, so is the enzymatic activity. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 56: 485-491, 1997.  相似文献   

6.
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   

7.
Pressure affects enzyme function in nonaqueous media. Activation volumes have been determined and provide evidence that the primary effect of pressure is to enhance the stripping of water off an enzyme in polar organic solvents and leads to decreased enzymatic activity. Activation volumes of subtilisin Carlsberg in organic solvents, particularly with the enzyme hydrated, have a larger magnitude than activation volumes determined in aqueous solutions. This study provides further evidence that enzymatic activity in polar organic solvents is dominated by the interaction of enzyme-bound water with the solvent. From a practical standpoint, however, the results of this study suggest that enzymatic catalysis in organic solvents may be controlled by the combined effects of pressure and enzyme hydration. (c) 1993 John Wiley & Sons, Inc.  相似文献   

8.
Chymotrypsin (EC 3.4.21.1) powder suspended in hexane in the presence of Na2CO3.10H2O is a good catalyst for peptide synthesis. The salt hydrate releases water to fix the thermodynamic water activity of the system in accord with its dissociation pressure. Salt hydrates can be useful to buffer water activity in mainly organic enzyme reaction mixtures at a value permitting activity of the catalyst while minimising hydrolytic side reactions.  相似文献   

9.
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   

10.
We report on the performance of cross-linked enzyme microcrystals (CLECs) of subtilisin Carlsberg in supercritical fluids (SC-fluids). The catalytic activity of CLECs in SC-ethane was found to be 2- to 10-fold greater than in hexane under the same conditions, using CLECs dried by propanol washing. Air-dried CLECs and lyophilized powders showed much lower activities, reflecting the same hydration hysteresis effects as in organic solvents. Reaction rates were much lower in SC-CO(2), especially at higher water activity, probably as a result of acid-base effects of carbonic acid on the enzyme.  相似文献   

11.
Single amino acid substitutions increase the activity and stability of subtilisin E in mixtures of organic solvents and water, and the effects of these mutations are additive. A variant of subtilisin E that exhibits higher activity in mixtures of dimethylformamide (DMF) and water (Q103R) was created by random mutagenesis combined with screening for improved activity (K. Chen and F. H. Arnold, in preparation). Another mutation, N218S, known to improve both the activity and stability of subtilisin BPN', also improves the activity and stability of subtilisin E in the presence of DMF. The effects of the two substitutions on transition-state stabilization are additive. Furthermore, the Q103R mutation that improves activity has no deleterious effect on subtilisin stability. The double mutant Q103R+N218S is 10 times more active than the wild-type enzyme in 20% (v/v) DMF and twice as stable in 40% DMF. Although the effects of single mutations can be impressive, a practical strategy for engineering enzymes that function in nonaqueous solvents will most likely require multiple changes in the amino acid sequence. These results demonstrate the excellent potential for engineering nonaqueous-solvent-compatible enzymes.  相似文献   

12.
Enzymatic catalysis in nonaqueous solvents   总被引:39,自引:0,他引:39  
Subtilisin and alpha-chymotrypsin vigorously act as catalysts in a variety of dry organic solvents. Enzymatic transesterifications in organic solvents follow Michaelis-Menten kinetics, and the values of V/Km roughly correlate with solvent's hydrophobicity. The amount of water required by chymotrypsin and subtilisin for catalysis in organic solvents is much less than needed to form a monolayer on its surface. The vastly different catalytic activities of chymotrypsin in various organic solvents are partly due to stripping of the essential water from the enzyme by more hydrophilic solvents and partly due to the solvent directly affecting the enzymatic process. The rate enhancements afforded by chymotrypsin and subtilisin in the transesterification reaction in octane are of the order of 100 billion-fold; covalent modification of the active center of the enzymes by a site-specific reagent renders them catalytically inactive in organic solvents. Upon replacement of water with octane as the reaction medium, the specificity of chymotrypsin toward competitive inhibitors reverses. Both thermal and storage stabilities of chymotrypsin are greatly enhanced in nonaqueous solvents compared to water. The phenomenon of enzymatic catalysis in organic solvents appears to be due to the structural rigidity of proteins in organic solvents resulting in high kinetic barriers that prevent the native-like conformation from unfolding.  相似文献   

13.
The water activities (a(w)) of 13 salt hydrate pairs were determined from vapor pressure measurements; a(w) values for a subset were also estimated from a study of water transfer to isopropylether. The application of salt hydrates as water buffers was investigated in two models: (i) effect of hydration on the initial rate of subtilisincatalyzed transesterification of the nitrophenol ester of CBZ-alanine with butanol; and (ii) effect of hydrates on the equilibrium concentrations of reactants in the esterification of dodecanol and decanoic acid, catalyzed by lipase. Transfer of ions from salt to enzyme particles was also demonstrated. The implications of the results for the successful use of salt hydrates as water buffers are discussed. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 55: 367-374, 1997.  相似文献   

14.
The rates of transesterification reactions catalyzed by the protease subtilisin Carlsberg suspended in various anhydrous solvents at 30 degrees C can be increased more than 100-fold by the addition of denaturing organic cosolvents (dimethyl sulfoxide or formamide); in water, the same cosolvents exert no enzyme activation. At 4 degrees C, the activation effect on the lyophilized protease is even higher, reaching 1000-fold. Marked enhancement of enzymatic activity in anhydrous solvents by formamide is also observed for two other enzymes, alpha-chymotrypsin and Rhizomucor miehei lipase, and is manifested in two transesterification reactions. In addition to lyophilized subtilisin, crosslinked crystals of subtilisin are also amenable to the dramatic activation by the denaturing cosolvents. In contrast, subtilisin solubilized in anhydrous media by covalent modification with poly(ethylene glycol) exhibits only modest activation. These observations are rationalized in terms of a mechanistic hypothesis based on an enhanced protein flexibility in anhydrous millieu brought about by the denaturing organic cosolvents. The latter exert their lubricating effect largely at the interfaces between enzyme molecules in a solid preparation, thus easing the flexibility constraints imposed by protein-protein contacts. (c) 1996 John Wiley & Sons, Inc.  相似文献   

15.
The alkaline proteases subtilisin Carlsberg and alcalase possess substantial enzymatic activity even when dissolved in ethanol. The crude enzymes were purified by gel filtration and the main fractions suspended in ethanol to give a translucent suspension. Both the supernatant and the resuspended precipitate after high-speed centrifugation were found to have enzymatic activities. The solubility of subtilisin Carlsberg in anhydrous ethanol was found to be 45.1g/ml and that of alcalase was 48.1g/ml by Coomassie blue dye-binding method using bovine serum albumin as a standard. In the presence of water, the solubility of both enzymes increased with water content. The stability of enzymes incubated in ethanol was assayed by their amidase and transesterase activities using Ala-Ala-Pro-Phe-pNA as substrate in phosphate buffer (pH8.2) and Moz-Leu-OBzl as substrate in anhydrous ethanol, respectively. The soluble enzymes have a half-life of about 36 hr and that of suspended enzymes about 50 hr in the amidase activity assay, whereas the same soluble enzymes have a half-life of about several hours and that of suspended enzymes 1 h by the transesterase activity assay. The stability of both enzymes decreased as water concentration increased. The diastereoselectivity of the enzyme-catalyzed hydrolysis of diastereo pairs of tetrapeptide esters,l-Ala-l-Ala-(d-orl-)Pro-l-Phe-OMe andl-Ala-l-Ala-(d-orl-)Ala-l-Phe-OMe, in phosphate is as high as that of the transesterification of these substrates in ethanol. It is concluded that active sites and selectivity of alkaline serine proteases in anhydrous alcohol are probably very similar to those in aqueous solution in spite of the fact that a lower reactivity is usually associated with the enzymes in nonaqueous solvents.  相似文献   

16.
Salt hydrate pairs were used to control water activity in the ionic liquid 1-butyl-3-methylimidazolium hexafluorophosphate. It was shown that salt hydrate pairs behave essentially the same in ionic liquids as they do in organic solvents as long as they do not dissolve. Initial rate-water activity profiles were prepared for the immobilized Candida antarctica lipase catalyzed synthesis of 2-ethylhexyl methacrylate. The ability to use salt hydrate pairs for the control of water activity in ionic liquids should allow for improved comparison of enzyme activity and specificity in ionic liquids and conventional solvents.  相似文献   

17.
A R Rendina  W W Cleland 《Biochemistry》1984,23(22):5157-5168
Aldehyde analogues of the normal alcohol substrates induce ATPase activities by glycerokinase (D-glyceraldehyde), fructose-6-phosphate kinase (2,5-anhydromannose 6-phosphate), fructokinase (2,5-anhydromannose or 2,5-anhydrotalose), hexokinase (D-gluco-hexodialdose), choline kinase (betaine aldehyde), and pyruvate kinase (glyoxylate). Since purified deuterated aldehydes give V and V/K isotope effects near 1.0 for glycerokinase, fructokinase with 2,5-anhydro[1-2H]talose, hexokinase, choline kinase, and pyruvate kinase, the hydrates of these almost fully hydrated aldehydes are the activators of the ATPase reactions. Fructose-6-phosphate kinase and fructokinase with 2,5-anhydro[1-2H]mannose show V/K deuterium isotope effects of 1.10 and 1.22, respectively, suggesting either that both hydrate and free aldehyde may be activators (predicted values are 1.37 if only the free aldehyde activates the ATPase) or, more likely, that the phosphorylated hydrate breaks down in a rate-limiting step on the enzyme while MgADP is still present and the back-reaction to yield free hydrate in solution is still possible. 18O was transferred from the aldehyde hydrate to phosphate during the ATPase reactions of glycerokinase, fructose-6-phosphate kinase, fructokinase, and hexokinase but not with choline kinase or pyruvate kinase. Thus, direct phosphorylation of the hydrates by the first four enzymes gives the phosphate adduct of the aldehyde, which decomposes nonenzymatically, while with choline kinase and pyruvate kinase the hydrates induce transfer to water (metal-bound hydroxide or water with pyruvate kinase on the basis of pH profiles). Observation of a lag in the release of phosphate from the glycerokinase ATPase reaction at 15 degrees C supports the existence of a phosphorylated hydrate intermediate with a rate constant for breakdown of 0.035-0.043 s-1 at this temperature. Kinases that phosphorylate creatine, 3-phosphoglycerate, and acetate did not exhibit ATPase activities in the presence of keto or aldehyde analogues (N-methylhydantoic acid, D-glyceraldehyde 3-phosphate, and acetaldehyde, respectively), possibly because of the absence of an acid-base catalytic group in the latter two cases. These analogues were competitive inhibitors vs. the normal substrates, and in the latter case, the hydrate of acetaldehyde was shown to be the inhibitory species on the basis of the deuterium isotope effect on the inhibition constant.  相似文献   

18.
Salt-activation of nonhydrolase enzymes for use in organic solvents   总被引:1,自引:0,他引:1  
Enzymatic reactions are important for the synthesis of chiral molecules. One factor limiting synthetic applications of enzymes is the poor aqueous solubility of numerous substrates. To overcome this limitation, enzymes can be used directly in organic solvents; however, in nonaqueous media enzymes usually exhibit only a fraction of their aqueous-level activity. Salt-activation, a technique previously demonstrated to substantially increase the transesterification activity of hydrolytic enzymes in organic solvents, was applied to horse liver alcohol dehydrogenase, soybean peroxidase, galactose oxidase, and xanthine oxidase, which are oxidoreductase and oxygenase enzymes. Assays of the lyophilized enzyme preparations demonstrated that the presence of salt protected enzymes from irreversible inactivation. In organic solvents, there were significant increases in activity for the salt-activated enzymes compared to nonsalt-activated controls for every enzyme tested. The increased enzymatic activity in organic solvents was shown to result from a combination of protection against inactivation during the freeze-drying process and other as-yet undetermined factors.  相似文献   

19.
The dramatic activation of serine proteases in nonaqueous media resulting from lyophilization in the presence of KCl is shown to be unrelated to relaxation of potential substrate diffusional limitations. Specifically, lyophilizing subtilisin Carlsberg in the presence of KCl and phosphate buffer in different proportions, ranging from 99% (w/w) enzyme to 1% (w/w) enzyme in the final lyophilized solids, resulted in biocatalyst preparations that were not influenced by substrate diffusion. This result was made evident through use of a classical analysis whereby initial catalytic rates, normalized per weight of total enzyme in the catalyst material, were measured as a function of active enzyme for biocatalyst preparations containing different ratios of active to inactive enzyme. The active enzyme content of a given biocatalyst preparation was controlled by mixing native subtilisin with subtilisin preinactivated with PMSF, a serine protease inhibitor, and lyophilizing the enzyme mixture in the presence of different fractions of KCl and phosphate buffer. Plots of initial reaction rates as a function of percent active subtilisin in the biocatalyst were linear for all biocatalyst preparations. Thus, enzyme activation (reported elsewhere to be as high as 3750-fold in hexane for the transesterification of N-Ac-L-Phe-OEt with n-PrOH) is a manifestation of intrinsic enzyme activation and not relaxation of diffusional limitations resulting from diluted enzyme preparations. Similar activation is reported herein for thermolysin, a nonserine protease, thereby demonstrating that enzyme activation due to lyophilization in the presence of KCl may be a general phenomenon for proteolytic enzymes.  相似文献   

20.
A soluble enzyme preparation from the leaves of sweet marjoram (Majorana hortensis Moench) catalyzes the divalent cation-dependent cyclization of [1-3H]geranyl pyrophosphate to the bicyclic monoterpene alcohols (+)-[6-3H]cis- and (+)-[6-3H]-transsabinene hydrate, providing labeling patterns consistent with current mechanistic considerations. No free intermediates were detectable in the conversion of geranyl pyrophosphate to the sabinene hydrates as determined by isotopic dilution experiments. Label from H2(18)O water was quantitatively incorporated into the products, indicating that the hydroxyl oxygen atoms of both cis- and trans-sabinene hydrate are derived from water and not from the pyrophosphate ester moiety of the substrate. The two enzymatic activities were inseparable by several chromatographic procedures, and differential inactivation studies suggested that the two activities reside with the same enzyme. The sabinene hydrate cyclase (synthase) has an apparent molecular weight of 56,000, shows a pH optimum near 7.0, and requires a divalent metal ion (either Mn2+ or Mg2+) for activity. The enzyme preparation is also capable of cyclizing neryl pyrophosphate, the cis-isomer of geranyl pyrophosphate, and analysis of mixed substrate incubations indicated that the two precursors are mutually competitive. Kinetic analysis and comparison of Vrel/Km values revealed that geranyl pyrophosphate is the more efficient substrate. This is the first report on an enzyme preparation capable of cyclizing geranyl pyrophosphate and neryl pyrophosphate to the isomeric sabinene hydrates.  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