首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Water-in-oil microemulsion systems have been studied in recent years for a number of applications in protein separation and enzymology. Although it is well established that reversed micelle systems provide an excellent medium for nonaqueous biocatalytic studies, there is still much speculation as to the interaction of the enzyme with the surfactant interface. Polyoxyethylene sorbitan trioleate (Tween 85) is a nonionic surfactant which has some interesting properties for microemulsion formation and protein solubilization. In conjunction with a separate article describing the structural features of Tween 85 reversed micelles in hexane with isopropanol as a cosurfactant, this work describes the activity of an enzyme, organophosphorus hydrolase, for degrading organophosphorus pesticides in this microemulsion system. Ternary phase diagrams were constructed to outline the phase boundaries at different temperatures and isopropanol concentrations, which elucidate the role of the cosurfactant alcohol, as well as some features of micelle structure. Kinetic and stability studies with organophosphorus hydrolase show the effect of enzyme partitioning between the micelle surfactant layer and aqueous core. (c) 1994 John Wiley & Sons, Inc.  相似文献   

2.
Lipases from Candida cyclindracea (L-1754) and wheat germ (L-3001) have been used to hydrolyze esters to their corresponding alcohols and acids in reverse micelles. Alcohol dehydrogenase from baker's yeast (YADH) was subsequently used to reduce the alcohol products to aldehydes. Cofactor recycling in the redox reaction was achieved using a sacrificial cosubstrate, as described previously. Four surfactants (sodium dioctylsulfosuccinate, Nonidet P-40 with Triton X-35, polyoxyethylene, 10-cetyl-ether, polyoxyethylene sorbitan trioleate) were employed to determine the effect of amphiphile on ester hydrolysis and redox reaction rates separately. The effect of type of organic solvent, W(0) [(water]/[surfactant)], and substrate concentration on separte enzyme activity were also investigated. A brief investigation of a single phase, two-step reaction catalyzed by the combination of lipase and YADH in reverse micelles is also reported. The activities of the enzymes are significantly different when used together instead of independently. (c) 1994 John Wiley & Sons, Inc.  相似文献   

3.
The performance of lipases from Candida rugosa and wheat germ have been investigated in three reaction media using three acetate hydrolyses as model reactions (ethyl acetate, allyl acetate, and prenyl acetate). The effect of substrate properties and water content were studied for each system (organic solvent, biphasic system, and reverse micelles). Not unexpectedly, the effect of water content is distinct for each system, and the optimal water content for enzyme activity is not always the same as that for productivity. A theoretical model has been used to simulate and predict enzyme performance in reverse micelles, and a proposed partitioning model for biphasic systems agrees well with experimental results. While the highest activities observed were in the micellar system, productivity in microemulsions is limited by low enzyme concentrations. Biphasic systems, however, support relatively good activity and productivity. The addition of water to dry organic solvents, combined with the dispersion of lyophilized enzyme powders in the solvent, resulted in significant enzyme aggregation, which not surprisingly limits the applicability of the "anhydrous" enzyme suspension approach. (c) 1995 John Wiley & Sons, Inc.  相似文献   

4.
The effect of a series of nonionic surfactants on the initial rate of the peroxide oxidation of 5-aminosalicylic acid in solution catalyzed by horseradish peroxidase was studied. As the surfactant concentration increases, the peroxidation rate first increases, then decreases, and the increase/decrease cycle is repeated. The primary increase may be induced by a change in properties of the medium under the action of surfactants, and the following decrease, by the enzyme inhibition. The secondary increase may be explained by to a change in the enzyme conformation and an increase in the accessibility of its active site for the substrate due to the immobilization of the protein in the surfactant aggregates, whereas the secondary decrease, by a shielding of the protein with these aggregates. For communication II, see [1].  相似文献   

5.
Lysozyme activity in the presence of nonionic detergent micelles   总被引:2,自引:0,他引:2  
The effect of a nonionic surfactant, polyoxyethylenesorbitan monolaurate (Tween 20), on the hen egg-white lysozyme catalyzed lysis of a dried cell suspension of Micrococcus lysodeikticus is analysed. A rate enhancement of up to 70% is observed in the presence of surfactant at concentrations above the critical micelle concentration. This activity increase may be explained by postulating the existence of a micelle-enzyme complex in which enzyme molecules are bound to micelles with preferential orientation of their active sites. The reaction is found to be second order with respect to substrate. A mechanism is postulated in which a substrate particle is assumed to be an energy-furnishing collision partner to the enzyme-substrate complex. This mechanism correlated data over a wide range of enzyme and substrate concentrations. Data from kinetic, ultrafiltration, ultraviolet, and fluorescence studies provide convincing evidence for the existence of a micelle-lysozyme complex. The results suggest that it is possible that immobilized enzymes mat in general be more reactive than corresponding free enzymes.  相似文献   

6.
In this article, the extraction of cytochrome c utilizing various nonionic surfactant microemulsions has been tested to determine the effect of surfactant structure on protein partitioning. Surfactants tested include a linear alcohol ethoxylate (Neodol 91-2.5), two alkyl phenol ethoxylates (lgepal CO-520, Trycol 6985), and a series of alkyl sorbitan esters that are either ethoxylated (Tweens) or un-ethoxylated (Spans). Initial attempts to extract hemoglobin into Neodol 91-2.5 Winsor II microemulsions (oil-continuous) appeared successful based on heme estimation. Careful analysis showed that the hemoglobin had dissociated prior to extraction and that only the heme was extracted with false positive results. In fact, Neodol 91-2.5 microemulsions were unable to extract a variety of proteins with differing biophysical properties. Among all the other nonionic surfactant microemulsions tested only those made using sorbitan esters extracted significant amounts of cytochrome c. The partition coefficients achieved in this study are more than an order of magnitude higher than that seen previously in the literature for comparable sorbitan systems. However, this partition coefficient is extremely sensitive to ionic strength. At an ionic strength as low as 0.001 M, the partition coefficient is reduced to that seen in previous studies. We have found that protein partitioning in sorbitan ester microemulsions is not a function of water content. In addition, extraction is not a function of either alkyl chain length, or polyethylene oxide molecular weight. Hence, the sorbitan group appears to have an important role in extraction, possibly through a weak electrostatic protein-surfactant interaction. (c) 1995 John Wiley & Sons, Inc.  相似文献   

7.
High pressure EPR studies of protein mobility in reversed micelles   总被引:1,自引:0,他引:1  
We have investigated the effect of pressure on structural properties of subtilisin solubilized in reversed micelles of Tween-85/isopropanol in hexane. Electron paramagnetic resonance (EPR) spectra of spin-labeled enzyme indicate a reduction in spin-label mobility when the enzyme is transferred from aqueous solution to the microemulsion. One explanation for the spectral broadening is a change in the protein's active-site conformation and/or dynamics. However, over a W(0) range of 80 to 180, EPR spectroscopy could detect no change in the enzyme's environment, conformation, or molecular dynamics. The EPR spectra also contained a contribution from free spin label located in an environment with a polarity roughly between that of propanol and bulk water. No changes in the polarity surrounding the free spin label nor in the enzyme's structural properties were evident at pressures up to 10,000 psi. Previous work has demonstrated that pressure can be used to manipulate the size of some reversed micelles, and the EPR data indicated that for this system such pressure tuning of micellar properties will not adversely affect the structure of solubilized enzyme. (c) 1994 John Wiley & Sons, Inc.  相似文献   

8.
Alcohol oxidase from Pichia pastoris together with catalase from bovine liver was used to oxidize n-hexanol to hexanal. For this purpose, an aqueous buffer solution was mixed with large amounts of hexanol by simple agitation, yielding a biphasic system, or by adding the nonionic surfactant Brij 35. Initial velocities and reaction yields after 24 h were measured as a function of various parameters such as the amounts of enzymes, hexanol, or surfactant. High enzymatic activity was determined for hexanol concentrations of between 20 mass% and 80 mass% without using any additional organic solvent. The homogenization of the biphasic systems with the help of Brij 35 did not yield a significant improvement of the bioconversion, which would justify the use of surfactants.  相似文献   

9.
The activities of horseradish peroxidase (HRP) and lactoperoxidase (LPO) entrapped in reverse micelles of Igepal CO-520 in cyclohexane were studied. When the molar ratio of water to surfactant, w 0 was ≥13, the activity of HRP encapsulated in the water pool of the reverse micelle was comparable with that measured in buffer. For LPO, however, lower activity was observed after its incorporation into the same system.

The activity of the investigated peroxidases was also measured in an aqueous solution of Igepal CO-720 or after incubation with this surfactant. The enzymes became inactivated in an aqueous micellar solution of Igepal CO-720, although this process was reversible.

The stability of HRP and LPO at 37 or 50°C was lower in the micellar systems than in buffer with the exception for HRP in reverse micelles at 50°C.  相似文献   

10.
The activities of horseradish peroxidase (HRP) and lactoperoxidase (LPO) entrapped in reverse micelles of Igepal CO-520 in cyclohexane were studied. When the molar ratio of water to surfactant, w0 was ≥13, the activity of HRP encapsulated in the water pool of the reverse micelle was comparable with that measured in buffer. For LPO, however, lower activity was observed after its incorporation into the same system.

The activity of the investigated peroxidases was also measured in an aqueous solution of Igepal CO-720 or after incubation with this surfactant. The enzymes became inactivated in an aqueous micellar solution of Igepal CO-720, although this process was reversible.

The stability of HRP and LPO at 37 or 50°C was lower in the micellar systems than in buffer with the exception for HRP in reverse micelles at 50°C.  相似文献   

11.
Fatty alcohols (FOHs) are important feedstocks in the chemical industry to produce detergents, cosmetics, and lubricants. Microbial production of FOHs has become an attractive alternative to production in plants and animals due to growing energy demands and environmental concerns. However, inhibition of cell growth caused by intracellular FOH accumulation is one major issue that limits FOH titers in microbial hosts. In addition, identification of FOH-specific exporters remains a challenge and previous studies towards this end are limited. To alleviate the toxicity issue, we exploited nonionic surfactants to promote the export of FOHs in Rhodosporidium toruloides, an oleaginous yeast that is considered an attractive next-generation host for the production of fatty acid-derived chemicals. Our results showed FOH export efficiency was dramatically improved and the growth inhibition was alleviated in the presence of small amounts of tergitol and other surfactants. As a result, FOH titers increase by 4.3-fold at bench scale to 352.6 mg/L. With further process optimization in a 2-L bioreactor, the titer was further increased to 1.6 g/L. The method we show here can potentially be applied to other microbial hosts and may facilitate the commercialization of microbial FOH production.  相似文献   

12.
Fusarium solani pisi recombinant cutinase solubilized in reversed micelles of a nonionic surfactant (phosphatidylcholine) in isooctane was used to catalyze the esterification of fatty acids with 2-butanol. Various parameters affecting the catalytic activity of the microencapsulated cutinase, such as pH, wo (molar ratio water/surfactant), temperature and substrate concentration were investigated. Maximal specific activity were obtained with wo=13, at pH 10.7 and 35d`C. The cutinase showed a higher specific activity for short length fatty acids, namely butyric acid. Calculation of the apparent kinetic parameters (km and Vmax) for the synthesis of butyl butyrate, showed a low apparent affinity of the cutinase in phosphatidylcholine reversed micelles for both substrates.  相似文献   

13.
An affinity-based reverse micellar system formulated with nonionic surfactant was applied to the refolding of denatured-reduced lysozyme. The nonionic surfactant of sorbitan trioleate (Span 85) was modified with Cibacron Blue F-3GA (CB) as an affinity surfactant (CB-Span 85) to form affinity-based reverse micelles in n-hexane. The water content of 15 was found optimal for lysozyme refolding in the reverse micellar system of 62.7 mmol/L Span 85 with coupled CB of 0.3 and 0.5 mmol/L. In addition, the operating conditions such as pH and the concentrations of urea and redox reagents were optimized. Under the optimized conditions, complete renaturation of lysozyme at 3-3.5 mg/mL was achieved, whereas dilution refolding in the bulk aqueous phase under the same conditions gave much lower activity recovery. Moreover, the secondary structure of the refolded lysozyme was found to be the same as the native lysozyme. Over 95% of the refolded lysozyme was recovered from CB-Span 85 reverse micelles by a stripping solution of 0.5 mol/L MgCl(2). Thus, the present system is advantageous over the conventional reverse micellar system formed with ionic surfactants in the ease of protein recovery.  相似文献   

14.
alpha-chymotrypsin is taken as a model protein to investigate three aspects of the protein extraction by reverse micelles: (1) the comparison between the two forward transfer techniques, i.e., the liquid-liquid and the solid state-liquid transfer; (2)the back-transfer, i.e., the capability of the protein to be recovered from the micellar solution; and (3) the maintainance of the enzyme activity at the end of the extraction cycle. Concerning the forward transfer from the liquid phase, we study first the effect of salt initially present in the aqueous phase on the equilibrium concentration of the extracted species; further, we study the forward protein extraction from the solid state, and the effect of pH, salt, and protein concentration on the transfer efficiency. Concerning the back transfer, we find the somewhat surprising result, that the percentage of protein back-extraction depends on the type and concentration of salt used for the forward transfer. Preliminary data concerning an alternative method for the back-transfer using silica gel to liberate the protein from the micellar environment, are presented. Finally, it is found that the enzyme activity depends again on the type and concentration of salt used for the forward transfer.  相似文献   

15.
Phase transfer studies were carried out on the solubilization of horseradish peroxidase (HRP) (E.C. 1.11.1.7) in reverse micelles formed in isooctane using the anionic surfactant, aerosol OT, at concentrations between 50 and 110mM. The selectivity of this methodology was tested, because the HRP used comprised a mixture of seven different isoenzymes with a wide range of isoelectric points. Forward and backward transfers were carried out in wellstirred vessels until equilibrium was reached. Significant protein partitioning could only be obtained by using NaCl to adjust ionic strength in pH range between 1.5 and 3.5, with a maximum at pH 3. The back transfer process was best at pH 8 with 80mM phosphate buffer and 1 M KCI. A loss of 1% to 3% of the surfactant through precipitation at the interface at pH<4 was observed, which may be due to instability in this pH region, because, even without protein, a similar precipitate was noticed. Protein partitioning was approximately constant when the ionic strength was increased up to 1 MNaCl at pH 3, but protein recovery in back transfer decreased accordingly. Hydrophobic interactions together with association between the protein and surfactant might be responsible for that behavior. Protein partitioning remained the same when the surfactant concentration was decreased to 50 mM, at the expense of higher variability. HPLC chromatograms showed no apparent damage to the protein after reverse micellar extraction. Protein partitioning is best when the temperature is kept at 25xC. The amount of protein and specific activity recovered strongly depends on the phase ratio used during forward transfer. Overall activity recovery varied from 87% to 136% when the phase ratio was increased from 1:1 to 30:1 in forward transfer. This behavior may be due to a change in the ratio of the three isoenzymes recovered after the backward transfer process, with the most active one being increasingly enriched at higher phase ratios. (c) 1994 John Wiley & Sons, Inc.  相似文献   

16.
Scott JJ 《Plant physiology》1991,95(4):1298-1301
Alkaline phytase activity, with a pH optimum of 8, was recovered from detergent extracts of dormant seeds of nine varieties of Phaseolus vulgaris L., Pisum sativum L. var. Early Alaska, and Medicago sativa L. This alkaline phytase of legume seeds was activated by calcium and differed from most seed phytases in its relative insensitivity to inhibition by fluoride.  相似文献   

17.
Structure and activity of trypsin in reverse micelles   总被引:3,自引:0,他引:3  
The kinetic properties of trypsin have been studied in reverse micelles formed by two surfactant systems, namely bis(2-ethylhexyl) sodium sulfosuccinate (AOT) in isooctane, and hexadecyltrimethyl ammonium bromide (CTAB) in chloroform/isooctane (1:1, by vol.). Three substrates have been used, namely N alpha-benzoyl-L-Arg ethyl ester, N alpha-benzoyl-L-Phe-L-Val-L-Arg p-nitroanilide (BzPheValArg-NH-Np) in AOT and N alpha-benzyloxycarbonyl-L-Lys p-nitrophenyl ester (ZLysO-Np) in CTAB. One of the main aims of the work was to compare the behaviour of trypsin in reverse micelles with that of alpha-chymotrypsin, for which an enhancement of kcat had been observed with respect to aqueous solutions. The pH profile is not significantly altered in reverse micelles with respect to water, however the kinetic parameters (kcat and Km) differ widely from one another, and are markedly affected by the micellar conditions, in particular by the water content wo (wo = [H2O]/[AOT]). Whereas in the case of BzPheValArg-NH-Np kcat is much smaller than in water, in the case of ZLysO-Np at pH 3.2 (but not at pH 6.0) a slight enhancement with respect to water is observed. On the basis of rapid kinetic spectrophotometry (stopped-flow) and solvent isotope effect studies, this enhancement is ascribed to a change in the rate-limiting step (acylation rather than hydrolysis). As in the case of alpha-chymotrypsin, the maximal activity is found for all substrates at rather small wo values (below 12), which is taken to suggest that the enzyme works better when is surrounded by only a few layers of tightly bound water. Spectroscopic studies [ultraviolet absorption, circular dichroism (CD) and fluorescence] have been carried out as a function of wo. Whereas the absorption properties are practically unchanged, the CD spectrum in AOT micelles has a lower intensity than in water, which is interpreted as a partial unfolding. The intensity is partly restored when Ca2+ ions are added, indicating that the micellar environment may cause a partial denaturation by depleting it of calcium ions. Fluorescence data show that the emission properties of the protein in reverse micelles match those in aqueous solution at around wo = 13 approx., whereas lambda max shifts towards the red by increasing wo, indicating an exposure of the tryptophan residues and probably an unfolding of the whole protein, at wo values above 15. Finally the reaction between trypsin and its specific macromolecular Kunitz inhibitor from soybeans is studied.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
Chymotrypsin is easily extracted from an aqueous solution into isooctane containing the anionic surfactant aerosol OT (AOT). The concentration of AOT needed to efficiently extract 0.5 mg/mL CMT is as low as 1 mM and as low as 0.2 mM AOT was sufficient to extract the protein into isooctane. The extraction process was unaffected by 10% (v/v) ethyl acetate in the isooctane phase. Moreover, spectroscopic analysis by electron paramagnetic resonance indicated that CMT did not exist inside a discreet water pool of a reversed micelle. Calculations of the number of AOT molecules associated per extracted CMT molecule indicate that only ca. 30 surfactant molecules interact with the protein, a value too low for reversed micellar incorporation of the protein in isooctane. These studies suggested that reversed micelles do not need to be involved in the actual transfer of the protein from the aqueous to the organic phase and protein solubilization in the organic phase is possible in the absence of reversed micelles. Based on these findings, a new mechanism has been proposed herein for protein extraction via the phase transfer method involving ionic surfactants. The central theme of this mechanism is the formation of an electrostatic complex between CMT and AOT at the aqueous/organic interface between AOT and CMT, thereby leading to the formation of a hydrophobic species that partitions into the organic phase. Consistent with this mechanism, the efficiency of extraction is dependent on the interfacial mass transfer, the concentrations of CMT and AOT in the aqueous and organic phases, respectively; the ionic strength of the aqueous phase; and the presence of various cosolvents. (c) 1994 John Wiley & Sons, Inc.  相似文献   

19.
Trypsin inhibitor was converted to hydrophobic states by covalently combining cholesteryl groups using an acylation reaction, and was immobilized in reverse micelles composed of a nonionic surfactant. Using this reverse micellar phase containing trypsin inhibitor as an affinity ligand, trypsin was selectively separated with high recoveries from a mixture of several kinds of contaminating proteins by forward and backward extraction. No loss of activity of the recovered trypsin was observed through these operations. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 53: 406-408, 1997.  相似文献   

20.
When seven different hydrolytic enzymes (four proteases and three lipases) were lyophilized from aqueous solution containing a ligand, N-Ac-L-Phe-NH(2), their catalytic activity in anhydrous solvents was far greater (one to two orders of magnitude) than that of the enzymes lyophilized without the ligand. This ligand-induced activation was expressed regardless of whether the substrate employed in organic solvents structurally resembled the ligand. Furthermore, nonligand lyoprotectants [sorbitol, other sugars, and poly(ethylene glycol)] also dramaticaliy enhanced enzymatic activity in anhydrous solvents when present in enzyme aqueous solution prior to lyophilization. The effects of the ligand and of the lyoprotectants were nonadditive, suggesting the same mechanism of action. Excipient activated and nonactivated enzymes exhibited identical activities in water. Also, addition of the excipients directly to suspensions of nonactivated enzymes in organic solvents had no appreciable effect on catalytic activity. These observations indicate that the mechanism of the excipient-induced activation is based on the ability of the excipients to alleviate reversible denaturation of enzymes upon lyophilization. Activity enhancement induced by the excipients is displayed even after their removal by washing enzymes with anhydrous solvents. Subtilisin Carlsberg, lyophilized with sorbitol, was found to be a much more efficient practical catalyst than its "regular" counterpart. (c) 1993 John Wiley & Sons, Inc.  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