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.
Organic solvents strip water off enzymes   总被引:12,自引:0,他引:12  
Exchange of enzyme-bound H(2)O with T(2)O in aqueous solution followed by freeze drying provided tritiated water bound to chymotrypsin, subtilisin Carlsberg, and horseradish peroxidase. The desorption of T(2)O from these enzymes suspended in various organic solvents showed that all three enzymes lost enzyme-bound water with peroxidase losing the most T(2)O of the three in solvents of moderate to high polarity. Polar solvent resulted in the highest degree of T(2)O desorption (e.g., methanol desorbed from 56%-62% of the bound T(2)O), while nonpolar solvents resulted in the lowest degree of desorption (e.g., hexane desorbed from 0.4%-2% of the bound T(2)O). Desorption is nearly immediate with most of the desorbable T(2)O being released from the enzymes within the first 5 min. Both solvent dielectric and a measure of the saturated molar solubility of water in a given solvent provide accurate correlations between the properties of the organic solvents and the extent of T(2)O desorption. This investigation shows that water stripping from an enzyme into a nonaqueous medium does occur and can be significant in polar solvents.  相似文献   

2.
A new technique with controlled interface generation allows separation and quantitation of enzyme inactivation by both solvent/aqueous interface and dissolved solvent. This has now been used in n-butanol, isopropylether, 2-octanone, n-hexane, n-butylbenzene, and n-tridecane. Ribonuclease was stable with all the solvent/aqueous interfaces studied. Chymotrypsin was mainly inactivated by the more hydrophobic solvent/aqueous interfaces, whereas lipase was only inactivated by the less hydrophobic solvent/aqueous interfaces. Urease was inactivated by some interfaces, but not all, without an obvious trend. Thus, the commonly expected simple relationship with solvent polarity (e.g., log P) does not apply when interfacial inactivation is determined specifically. Greater dissolved solvent inactivation occurred with the more polar solvents, though only a general trend was apparent with log P. A better correlation was noted with the Hilde-brand solubility parameter. Interfacial effects are discussed with reference to enzyme molecular weight, denaturation temperature, hydrophobicity, and adiabatic compressibility. (c) 1994 John Wiley & Sons, Inc.  相似文献   

3.
Lanne于1987年提出了生物催化剂工程(Biocatalyst engimeering)和介质工程(Medium enineering)的概念[1].有机相生物催化中溶剂的选择也是介质工程的内容之一。纯酶在有机相中的催化作用已有大量报道[2],但对完整细胞研究甚少。本文以甲基单胞菌(Methylomonos)Z201完整细胞为生物催化剂,丙烯环氧化为指标反应,研究有机溶剂对活性的影响并对催化活性-溶剂疏水性进行了相关性分析。研究了水-十六烷两相体系中十六烷含量和搅拦速度对丙烯环氧化速度的影响和细胞的操作稳定性。  相似文献   

4.
Approaches for increasing the solution stability of proteins   总被引:1,自引:0,他引:1  
Stabilization of proteins through proper formulation is an important challenge for the pharmaceutical industry. Two approaches for stabilization of proteins in solution are discussed. First, work describing the effect of additives on the thermally induced denaturation and aggregation of low molecular weight urokinase is presented. The effects of these additives can be explained by preferential exclusion of the solute from the protein, leading to increased thermal stability with respect to denaturation. Diminished denaturation leads to reduced levels of aggregation. The second approach involves stoichiometric replacement of polar counter ions (e.g., chloride, acetate, etc.) with anionic detergents, in a process termed hydrophobic ion pairing (HIP). The HIP complexes of proteins have increased solubility in organic solvents. In these organic solvents, where the water content is limited, the thermal denautration temperatures greatly exceed those observed in aqueous solution. In addition, it is possible to use HIP to selectively precipitate basic proteins from formulations that contain large amounts of stabilizers, such as human serum albumin (HSA), with a selectivity greater than 2000-fold. This has been demonstrated for various mixtures of HSA and interleukin-4. (c) 1995 John Wiley & Sons, Inc.  相似文献   

5.
A solvent engineering strategy was applied to the lipase-catalyzed synthesis of xylitol-oleic acid monoesters. The different esterification degrees for this polyhydroxylated molecule were examined in different organic solvent mixtures. In this context, conditions for high selectivity towards monooleoyl xylitol synthesis were enhanced from 6 mol% in pure n-hexane to 73 mol% in 2-methyl-2-propanol/dimethylsulfoxide (DMSO) 80:20 (v/v). On the contrary, the highest production of di- and trioleoyl xylitol, corresponding to 94 mol%, was achieved in n-hexane. Changes in polarity of the reaction medium and in the molecular interactions between solvents and reactants were correlated with the activity coefficients of products. Based on experimental results and calculated thermodynamic activities, the effect of different binary mixtures of solvents on the selective production of xylitol esters is reported. From this analysis, it is concluded that in the more polar conditions (100% dimethylsulfoxide (DMSO)), the synthesis of xylitol monoesters is favored. However, these conditions are unfavorable in terms of enzyme stability. As an alternative, binary mixtures of solvents were proposed. Each mixture of solvents was characterized in terms of the quantitative polarity parameter E(T)(30) and related with the activity coefficients of xylitol esters. To our knowledge, the characterization of solvent mixtures in terms of this polarity parameter and its relationship with the selectivity of the process has not been previously reported.  相似文献   

6.
The title complex cation, [Sb(tbpc)(OH)(2)](+) (where tbpc denotes tetra(tert-butyl)phthalocyaninate, C(48)H(48)N(8)(2-)), has been prepared by oxidizing [Sb(tbpc)]I(3) with tert-butyl perbenzoate in CH(2)Cl(2), CHCl(3), o-dichlorobenzene and also without solvent in the range of 20-80 degrees C. This species has been isolated as I(3)(-) salt and characterized by elemental analysis, ESI-MS, FT-IR, optical absorption and emission, and magnetic circular dichroism spectroscopy. This compound is quite well soluble in common polar organic solvents (e.g., CH(2)Cl(2), acetonitrile, acetone) without detectable aggregation at least up to ca. 10(-4)M while much less soluble (e.g., benzene, chloronaphthalene) or insoluble (hexane) in non-polar solvents. Although this compound is insoluble in water, it makes hydrophilic colloids in acetone-water mixtures. The most intense absorption band (Q-band) in a specific solvent red-shifts with an increase in the refractive index of the solvent. However, considerable deviation of the Q-band positions in donor-solvents from linear correlation between the positions and Onsager's solvent polarity function suggests that there are significant specific chemical interactions between the axial hydroxyl groups and the surrounding donor molecules. The low fluorescence quantum yield (ca. 0.01) for [Sb(tbpc)(OH)(2)](+) suggests that the singlet excited state of this species is considerably quenched by the presence of antimony ion in the chromophore.  相似文献   

7.
Laabe于1987年提出了生物催化剂工程(Biocatalyst engineering)和介质工程(Medium engineering)的概念[1]。有机相生物催化中溶剂的选择也是介质工程的内容之一。纯酶在有机相中的催化作用已有大量报道[2],但对完整细胞研究甚步。本文以甲基单胞菌(Methylomonas Z201)完整细胞为生物催化剂.丙烯环氧化为指标反应.研究有机溶剂对活性的影响并对催化活性——溶剂疏水性进行了相关性分析。研究了水一十六烷两相体系中十六烷含量和搅拌速度对丙烯环氧化速度的影响和细胞的操作稳定性。  相似文献   

8.
The solvation properties of ubiquinone-10 and ubiquinol-10 in a wide variety of solvents of polarity varying from alkanes to water are reported. Greatest solubility is observed in solvents of intermediate polarity and particularly where low polarity is combined with a pronounced tendency to interact with the benzoquinone substituent of the ubiquinone molecule. This includes solvents like chloroform and benzene. Ubiquinone-10 is somewhat less polar than ubiquinol-10 as judged by comparative solubilities of the two molecules. Proton-NMR chemical shift measurements and aggregation studies in selected solvents indicate that in ubiquinone-10 in the liquid phase and in solution in hydrocarbons like dodecane the molecules have a preferred association possibly involving stacking of the benzoquinone rings. Surface balance studies indicated that the surface-active character of ubiquinone-10 is relatively weak and only in a comparatively polar and highly structured solvent, formamide, was there evidence of an effect on surface tension of the solvent. The critical micelle concentratiom in this solvent was estimated to be about 5 M on the basis of surface tension measurements. Ubiquinone-10 is well known to form virtually insoluble monolayers at the air/water interface. Studies of the partition of ubiquinone-10 in binary mixtures of solvents suggest that the interaction of the benzoquinone ring substituent with structured polar solvents is considerably weaker than the internal cohesion between molecules of the solvent. No evidence on the basis of wide-angle X-ray diffraction measurements was obtained to indicate that solvent molecules were a component of the crystal lattice of ubiquinone-10 that had precipitated from solvent mixtures.  相似文献   

9.
Steroids are generally sparingly soluble in water. Thus, for in vitro studies of steroid metabolism or enzymology it is common practice to solubilize steroids by the addition of a small amount (2–10%, v/v) of an organic cosolvent. Methanol, ethanol, and 1,2-propanediol, singly or in combination, have been widely used (1). Effects of organic solvents on the kinetic parameters, Km and Vmax, of steroid-metabolizing enzymes with various substrates have been demonstrated (2,3), and the results are consistent with the conclusion that organic solvent influences on catalytic activity reflect, in part, effects on the aggregation state and solubility of steroid substrates.Light-scattering measurements have been applied extensively in studies of macromolecular structure (4) and micelle formation by a large variety of amphiphilic substances [reviewed in Ref. (5)]. Jones and Gordon (6) used a commercial instrument, designed specifically for light-scattering measurements, to characterize micelle formation in aqueous solutions by Δ5-3-ketosteroids containing various substituents at the 17β position. They showed that turbidity versus concentration plots were of the form seen in studies of micelle formation (5) and that steroids can exist in solution in monomeric or micellar forms, their aggregation state being a function of the polarity of the steroid solute and the composition of the solvent.To estimate solubility quantitatively 3H- or 14C-labeled steroids have been used in conjunction with centrifugation (3), dialysis (7), or filtration (8). These techniques allow for accurate estimates of solubility, but one may encounter problems due to nonspecific absorption on membranes or the unavailability of the labeled steroid of interest.We have observed that steroid aggregation and solubility can be estimated easily and with high sensitivity with a commercially available fluorometer. In this report the method is described and examples demonstrating the reproducibility and sensitivity of the technique are presented.  相似文献   

10.
Solvent selection tests were carried out for the Delta(1) dehydrogenation of 6-alpha-methylhydrocortisone-21-acetate by Arthrobacter simplex cells in the presence of organic solvents. Solubility limits were determined for substrate and product in dry and water-saturated solvents and solvent mixtures. Molecular toxicity levels were estimated by measuring the dehydrogenation activity decay of freely suspended cells incubated in solvent-saturated aqueous media. Chloroform and n-decan-1-ol were the best choice of solvent, for both solubility and catalytic stability. High concentrations of water-soluble additives, such as monosodium glutamate, were found to greatly improve the retention of activity in chloroform-saturated media.  相似文献   

11.
The process of reversible denaturation of several proteins (alpha-chymotrypsin, trypsin, laccase, chymotrypsinogen, cytochrome c and myoglobin) by a broad series of organic solvents of different nature was investigated using both our own and literature data, based on the results of kinetic and spectroscopic measurements. In all systems studied, the denaturation proceeded in a threshold manner, i.e. an abrupt change in catalytic and/or spectroscopic properties of dissolved proteins was observed after a certain threshold concentration of the organic solvent had been reached. To account for the observed features of the denaturation process, a thermodynamic model of the reversible protein denaturation by organic solvents was developed, based on the widely accepted notion that an undisturbed water shell around the protein globule is a prerequisite for the retention of the native state of the protein. The quantitative treatment led to the equation relating the threshold concentration of the organic solvent with its physicochemical characteristics, such as hydrophobicity, solvating ability and molecular geometry. This equation described well the experimental data for all proteins tested. Based on the thermodynamic model of protein denaturation, a novel quantitative parameter characterizing the denaturing strength of organic solvents, called the denaturation capacity (DC), was suggested. Different organic solvents, arranged according to their DC values, form the DC scale of organic solvents which permits theoretical prediction of the threshold concentration of any organic solvent for a given protein. The validity of the DC scale for this kind of prediction was verified for all proteins tested and a large number of organic solvents. The experimental data for a few organic solvents, such as formamide and N-methylformamide, did not comply with equations describing the denaturation model. Such solvents form the group of so-called 'bad' solvents; reasons for the occurrence of 'bad' solvents are not yet clear. The DC scale was further extended to include also highly nonpolar solvents, in order to explain the well-known ability of enzymes to retain catalytic activity and stability in biphasic systems of the type water/water-immiscible organic solvent. It was quantitatively demonstrated that this ability is accounted for by the simple fact that nonpolar solvents are not sufficiently soluble in water to reach the inactivation threshold concentration.  相似文献   

12.
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   

13.
Flavonoids are polyphenolic secondary plant metabolites which possess antioxidant and anti-inflammatory properties. Besides, they have been shown to exhibit increased antioxidant properties in their polymerized form. Catechins are one of the attractive class of flavonoids which belong to the group of flavan-3-ols. Polymerization of catechins have been investigated in numerous studies indicating the requirement of certain amount of organic solvent to provide the solubility of the monomer. However, many research projects have been conducted recently to replace toxic organic contaminants of the processes with environmentally friendly solvents. In this aspect, deep eutectic solvents (DESs) that are regarded as “green solvents” have been studied extensively in various enzyme catalyzed reactions. In the present study, we focused on establishing a green pathway for laccase catalyzed polycatechin synthesis by replacing organic solvent content with DESs as green solvents. For this aim, various parameters were investigated, such as DES types and concentrations laccase amount and reaction time. Consequently, the highest molecular weight polycatechin was obtained using 5% (v/v) B–M, 125?U laccase in 1?hr of reaction time, at 30°C, as 4,354?±?678?g?mol?1. Corresponding X/XO inhibitory activity and superoxide radical scavenging activities were achieved as, 59 and 50%, respectively.  相似文献   

14.
The stability of biocatalysis in systems containing organic solvents is reviewed. Among the examples presented are homogeneous mixtures of water and water-miscible organic solvents, aqueous/organic two-phase systems, solid biocatalysts suspended in organic solvents, enzymes in reverse micelles and modified enzymes soluble in water immiscible solvents. The stability of biocatalysts in organic solvents depends very much on the conditions. The hydrophobicity or the polarity of the solvent is clearly of great importance. More hydrophobic solvents (higher log P values) are less harmful to enzymes than less hydrophobic solvents. The water content of the system is a very important parameter. Some water is essential for enzymatic activity; however, the stability of enzymes decreases with increasing water content. Mechanisms of enzyme inactivation are discussed.  相似文献   

15.
Why do crown ethers activate enzymes in organic solvents?   总被引:2,自引:0,他引:2  
One of the major drawbacks of enzymes in nonaqueous solvents is that their activity is often dramatically low compared to that in water. This limitation can be largely overcome by crown ether treatment of enzymes. In this paper, we describe a number of carefully designed new experiments that have improved the insights into the mechanisms that are operative in the crown ether activation of enzymes in organic solvents. The enhancement of enzyme activity upon addition of 18-crown-6 to the organic solvent can be reconciled with a mechanism in which macrocyclic interactions of 18-crown-6 with the enzyme play an important role. Macrocyclic interactions (e.g., complexation with lysine ammonium groups of the enzyme) can lead to a reduced formation of inter- and intramolecular salt bridges and, consequently, to lowering of the kinetic conformational barriers, enabling the enzyme to refold into thermodynamically stable, catalytically (more) active conformations. This assumption is supported by the observation that the crown-ether-enhanced enzyme activity is retained after removal of the crown by washing with a dry organic solvent. A much stronger crown ether activation is observed when 18-crown-6 is added prior to lyophilization, and this can be explained by a combination of two effects: the before-mentioned macrocyclic complexation effect, and a less specific, nonmacrocyclic, lyoprotecting effect. The magnitude of the total crown ether effect depends on the polarity and thermodynamic water activity of the solvent, the activation being highest in dry and apolar media, where kinetic conformational barriers are highest. By determination of the specific activity of crown-ether-lyophilized enzyme as a function of the enzyme concentration, the macrocyclic crown ether (linearly dependent on the enzyme concentration) and the nonmacrocyclic lyoprotection effect (not dependent on the enzyme concentration) could be separated. These measurements reveal that the contribution of the nonmacrocyclic effect is significantly larger than the macrocyclic refolding effect.  相似文献   

16.
The effects of various organic solvents on penicillin acylase-catalyzed synthesis of β-lactam antibiotics (pivampicillin and ampicillin) have been investigated in water-solvent mixtures. The rates of penicillin acylase-catalyzed reactions were found to be significantly reduced by the presence of a small amount of organic solvent. In particular, the rate of enzyme catalysis was extremely low in the presence of ring-structured solvents and acids while enzyme activities were fully restored after removing the solvents. This indicates that interactions between the solvents and the enzyme are specific and reversible. To correlate the inhibitory effects of organic solvents with solvent properties the influence of solvent hydrophobicities and solvent activity on the rate of pivampicillin synthesis was examined. The reaction rate was found to decrease with increasing solvent hydrophobicities, and a better correlation was observed between the reaction rate and solvent activity. The effects of ionic strength on the synthesis of pivampicillin and ampicillin were also examined. The ionic strength dependence indicates that electrostatic interactions are involved in the binding of ionic compounds to the enzyme. On the basis of the active site structure of penicillin acylase, a possible mechanism for molecular interactions between the enzyme and organic solvents is suggested.  相似文献   

17.
Using the solubility parameter mapping technique (Eisenbach, M., Caplan, S.R. and Tanny, G (1979) Biochim. Biophys. Acta 554, 269-280) we studied spectroscopically the mode of interaction between the purple membrane of Halobacterium halobium and pure organic solvents or solvent mixtures. Although the interacting solvents formed a well-defined closed region in the interaction maps, mapping the modes of interaction did not reveal a closed region for each spectrally classifiable type. A suggested interpretation for this is that interaction with the purple membrane chromophore requires that a solvent (or solvent mixture) possess apolar groups in order to obtain access to the chromophore, together with a polar character and hydrogen-bonding capacity. The mode of interaction, however, is dependent on the specificity of the reactive group of the solvent for retinal, and this has nothing to do with membrane properties. We also examined the influence of the duration of the interaction and of illumination. Some solvents appeared to react more sluggishly than others, but no generalization in terms of the solubility parameter mapping was found, probably because the map describes thermodynamic rather than kinetic phenomena. The only effect of illumination was to enhance the reaction of some of these solvents. It did not change the solubility parameters of purple membrane.  相似文献   

18.
Solid-phase synthesis of dipeptides in low-water media was achieved using AOT ion-paired alpha-chymotrypsin solubilized in organic solvents. Multiple solvents and systematic variation of water activity, a(w), were used to examine the rate of coupling between N-alpha-benzyloxycarbonyl-L-phenylalanine methyl ester (Z-Phe-OMe) and leucine as a function of the reaction medium for both solid-phase and solution-phase reactions. In solution, the observed maximum reaction rate in a given solvent generally correlated with measures of hydrophobicity such as the log of the 1-octanol/water partitioning coefficient (log P) and the Hildebrand solubility parameter. The maximum rate for solution-phase synthesis (13 mmol/h g-enzyme) was obtained in a 90/10 (v/v) isooctane/tetrahydrofuran solvent mixture at an a(w) of 0.30. For the synthesis of dipeptides from solid-phase leucine residues, the highest synthetic rates (0.14-1.3 mmol/h g-enzyme) were confined to solvent environments that fell inside abruptly defined regions of solvent parameter space (e.g., log P > 2.3 and normalized electron acceptance index <0.13). The maximum rate for solid-phase synthesis was obtained in a 90/10 (v/v) isooctane/tetrahydrofuran solvent mixture at an a(w) of 0.14. In 90/10 and 70/30 (v/v) isooctane/tetrahydrofuran environments with a(w) set to 0.14, seven different N-protected dipeptides were synthesized on commercially available Tentagel support with yields of 74-98% in 24 h.  相似文献   

19.
Summary Lipase fromCandida rugosa was immobilized by adsorption on three supports which could contain water available for the hydrolysis of olive oil in a reverse phase system. To select the most suitable solvent for this system, the effect of organic solvents on the stability and catalytic activity of immobilized lipase for the hydrolysis reaction has been examined. The results revealed that isooctane was superior to any other solvents tested in this study for enzymatic fat splitting in a reverse phase system. Also the effect of the solvent polarity on the hydrolysis of olive oil has been examined in detail using various organic solvents mixed with an equivolume of isooctane. It was found that the hydrolysis of olive oil by immobilized lipase was markedly affected by the polarity of reaction solvents.  相似文献   

20.
Recombinant proteins are often expressed in the form of insoluble inclusion bodies in bacteria. To facilitate refolding of recombinant proteins obtained from inclusion bodies, 0.1 to 1 M arginine is customarily included in solvents used for refolding the proteins by dialysis or dilution. In addition, arginine at higher concentrations, e.g., 0.5-2 M, can be used to extract active, folded proteins from insoluble pellets obtained after lysing Escherichia coli cells. Moreover, arginine increases the yield of proteins secreted to the periplasm, enhances elution of antibodies from Protein-A columns, and stabilizes proteins during storage. All these arginine effects are apparently due to suppression of protein aggregation. Little is known, however, about the mechanism. Various effects of solvent additives on proteins have been attributed to their preferential interaction with the protein, effects on surface tension, or effects on amino acid solubility. The suppression of protein aggregation by arginine cannot be readily explained by either surface tension effects or preferential interactions. In this review we show that interactions between the guanidinium group of arginine and tryptophan side chains may be responsible for suppression of protein aggregation by arginine.  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