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 properties of horseradish peroxidase in sodium dodecyl sulfate (DDS) reversed micelles in benzene-pentanol-water solutions are studied. The potential of the analytical application of direct and reversed DDS micelles is demonstrated using newly developed methods for the determination of peroxidase substrates (hydrogen peroxide and cystein), inhibitor (sulfanylamide), and activator (imidazole) via the oxidation of o-dianisidine (o-D) with hydrogen peroxide.  相似文献   

2.
《Electroanalysis》2005,17(19):1780-1788
The amperometric biosensing of aromatic amines using a composite glucose oxidase (GOD)‐peroxidase (HRP) biosensor in reversed micelles is reported. Rigid composite pellets of graphite and Teflon, in which GOD and HRP were coimmobilized by simple physical inclusion, were employed for the biosensor design. This design allows the in situ generation of the H2O2 needed for the enzyme reaction with the aromatic amines, thus preventing the negative effect that the presence of a high H2O2 concentration in solution has on HRP activity. The H2O2 in situ generation is performed by oxidation of glucose catalyzed by GOD. The effect of the composition of the reversed micelles, i.e., the nature of the organic solvent used as the continuous phase, the nature and concentration of the surfactant used as emulsifying agent, the aqueous 0.05 mol L?1 phosphate buffer percentage used as the dispersed phase, and the glucose concentration in the aqueous phase, on the biosensor response was evaluated. Reversed micelles formed with ethyl acetate, a 5% of phosphate buffer (pH 7.0) containing 3.0×10?3 mol L?1 glucose, and 0.1 mol L?1 AOT (sodium dioctylsulfosuccinate), were selected as working medium. Well‐defined and reproducible amperometric signals at 0.00 V were obtained for p‐phenylenediamine, 2‐aminophenol, o‐phenylenediamine, m‐phenylenediamine, 1‐naphthylamine, o‐toluidine and aniline. The useful lifetime of one single biosensor was of 60 days. The trend in sensitivity observed for the aromatic amines is discussed considering the effect of their structure on the stabilization of the radicals formed in the enzyme reaction which are electrochemically reduced. The behavior of the composite bienzyme electrode was also evaluated in a FI (flow injection) system using reversed micelles as the carrier. The suitability of the composite bienzyme electrode for the analysis of real samples was demonstrated by determining aniline in spiked carrots.  相似文献   

3.
We have investigated the mixing behavoir of a pH‐mediated ternary surfactant mixture at constant ratio of dodecyldimethylamine oxide (DDAO) and Triton X‐100 (9:1). From the equilibrium surface tension measurements at different pHs, the critical micelle concentration (cmc) data were obtained as functions of the pH. Values of the cmc and composition of the micelles were predicted using the regular solution approximation. To some extent, the experimental cmc values agree with the predicted cmc. The average degree of ionization of dodecyldimethylamine oxide in the mixed surfactant systems was estimated using potentiometric titrations. The surface electric potential of the micelles (Ψo) was determined using two methods, one by hydrogen ion titration and the other by the dissociation constants of an acid‐base indicator. In a high degree of ionization of DDAO in the micelles phase (am), Ψo estimated from acid‐base indicator is much higher than that from hydrogen ion titration. In the protonated dodecyldimethylamine oxide/TX‐100 binary surfactant system, Ψo estimated from hydrogen ion titration was as high as 89 mV. The micellar aggregation numbers evaluated by the steady‐state fluorescence probe method increase with pH except at pH=5.03. At pH=5.03, the maximum micelle aggregation number was observed.  相似文献   

4.
《Colloids and Surfaces》1988,29(4):403-410
The equilibrium distribution of a hydrophilic solute (Mz+ between an aqueous phase and a reversed micellar organic phase (consisting of a surfactant HA with aggregation number x, and dissolved in a hydrocarbon diluent) is analyzed quantitatively by treating the reversed micelles as a pseudophase. It is shown that when the M—A complex is strongly solubilized by the micellar pseudophase, the distribution coefficient (D) has a first-order dependence on the concentration of micellized surfactant (Cs). On the other hand, when the M—A complex is not solubilized by the reversed micelles, a plot of log D versus log Cs has a slope of (z/x); in this case the monomeric species HA is the active extractant and any effect that decreases surfactant aggregation (e.g. low aggregation number, small aggregation equilibrium constant) leads to an increase in the distribution coefficient.  相似文献   

5.
A sodium 1,4-bis[(2-ethylhexyl)oxy]-1,4-dioxybutane-2-sulfonate (NaАОТ)–water–isooctane three-component system is calculated by the molecular-dynamics method. In a wide range of relative water contents w 0, reverse micelles are obtained with different morphologies: single spherical and cylindrical micelles and their spatial networks. It is shown that w 0 and surfactant concentration are the main shape-generating factors. The data obtained are in good agreement with previous results of simulations and experimental data.  相似文献   

6.
The conformation of various basic poly (-amino acid)s was investigated by CD measurements in aqueous solutions containing bis (2-ethylhexyl)sodium sulfosuccinate (AOT) as well as in the AOT reversed micelles. The addition of AOT into an aqueous solution of poly(L-lysine) induces the conformational transition from coil to ordered structure, followed by aggregation. On the other hand, poly(L-lysine) assumes-structure in the reversed micelles at low wovalue (wo=[H2O]/[AOT]). Similarly to poly(L-lysine), poly(L-ornithine) takes an ordered structure in the aqueous solution containing AOT and-structure in the reversed micelles. In this case, however, these ordered structures are not so stable, compared with that of poly(L-lysine). Poly(L-arginine) undergoes the conformational transition from coil to helix by addition of AOT into the aqueous solution. Further addition of AOT allows transformation into-structure. Copoly(L-lysyl-L-leucine) with 63% leucine residue was shown to take a stable helical conformation even in pure water. In the reversed micelles, however, this ordered structure is significantly changed probably because the hydrophobic interaction among the leucyl residues is lowered in the reversed micelles.  相似文献   

7.
Nucleotide coupling was investigated in reverse micelles formed by (cetyl)trimethylammonium bromide (CTAB), in hexane/pentan-1-o1. In particular, the coupling of 2′ -deoxy-5′-O-methylcytidine 3′ O-phosphate, prepared by phosphoramidite chemistry, with 5′-amino-5-deoxythymidine was studied in the presence of a H2O-soluble carbodiimide at (wo) = 11 and 22 (wo=[H2O]/[CTAB]). The effect of wo on the reaction rate was investigated. A solid-phase strategy was developed for the synthesis of 2′-deoxy-5′O-methyl-cytidyl-(3′-5′)-5′-amino-5′deoxythymidine. The nucleotide coupling yieldig the expected product occurred readily in reverse micelles. Nucleotide coupling is thus possible in reverse micelles, and this is discussed in connection with the micellar self-replication program.  相似文献   

8.
Being solubilized in the systems of the surfactant reversed micelles (RM), the macromolecules incorporate into the inner cavity of RM, whose size can be changed by varying the surfactant hydration degree Wo, i.e. [H2O]/[Surfactant] molar ratio. The conjugates of macromolecules (protein-protein, protein-linear polyelectrolyte) are synthesized in RM of AOT (Aerosol OT, sodium bis[2-ethylhexyl]sulfosuccinate) in octane. The yield critically depends on the hydration degree: the reaction does not proceed at low Wo, but at Wo exceeding threshold value (which differs for various macromolecules) the yield increases sharply and reaches 100%. Using the ultracentrifugation it was demonstrated that at Wo lower than threshold the polyelectrolyte represents a compact globule compressed by the micellar matrix in the RM inner cavity. Under these conditions there is no space for the protein in the RM, containing polyelectrolyte, and therefore reaction does not proceed. At Wo higher than threshold RM become large enough to entrap the conjugated macromolecules simultaneously. The possibility of regulation of the conjugate composition by variation of Wo (size of micellar matrix) was demonstrated. The RM are applied as universal matrix microreactors, for modification of macromolecules with water insoluble reagents and for regulation of supramacromolecular composition of oligomeric enzymes.  相似文献   

9.
The enzymatic synthesis of amide bonds catalyzed by penicillin acylase is investigated both in H2O solution and in organic solvents containing reverse micelles. The specificity of the reaction is rather high on the side of the acyl component, practically only phenylacetic acid gives sizeable yields. On the contrary, a variety of amino-acid esters, dipeptides, and tripeptides can be used as amino component, e.g., serine methyl ester, methionine ethyl ester, tyrosine ethyl ester, Gly–Asp, Ala–Tyr, Gly–Tyr–Gly etc. However, Many other amino-acid residues do not react, and the possible reasons for this are discussed. Yields varyin rang the 10–80%. A. systematic study to optimize yields by varying the solvent composition is presented for one model reaction. The enzyme is also able to couple certain D-amino-acid residues (e.g. D-methionine ethyl ester or Gly-D -Asp) though at lower rate. Reverse micelles formed by the cationic surfactant cetyltrimethylammonium bromide (CTAB) in CHCl3/isooctane are used to host penicillin acylase and to perform amide synthesis in which the product is perferentially soluble in the organic solvent mixture. The reaction is studied as a function of pH and certain micellar parameters, e. g. wo (wo = [H2O]/[CTAB]). A new membrane enzyme reactor is utilized to separate the product from the enzyme-containing micelles. The adavantges and the limits of this approach are discussed.  相似文献   

10.
A common cationic surfactant,n-hexadecylammonium hydrogensulphate, dissolved in concentrated sulphuric acid, has been studied by static and dynamic light scattering. Micelle formation has been observed even in this unusual solvent. An apparent molar mass of 45 500±4.5% was found for the aggregates. A translational diffusion coefficientD 0=5.5×10–9 cm2/s was measured which gave a hydrodynamically effective radius ofR h=17.7 nm. The geometric radius of gyration wasR g=76.2 nm. The ratioR g/R h=4.33 is indicative for rodlike structures. Assuming a polydispersity ofL w/L n=2 this corresponds to a cylinder ofL w=152 nm. An axial ratiop w=(L w/d)=60.4 nm was estimated which leads to a cylinder diameter of 2.53 nm. At surfactant concentrations higher than 5% (w/vol) the rod-like micelles aggregate to form more globular structures. The time correlation function, recorded by dynamic light scattering, exhibited a two-step decay which indicates a bimodal distribution of particle sizes. The fast motion coincides with that of the micelles at low concentrations while the other is slower than the fast one by three orders of magnitude and corresponds to the translational motion of large clusters.  相似文献   

11.
The interaction of nonionic triblock copolymers of poly(ethyleneoxide) (PEO) and poly(propyleneoxide) (PPO) (PEOnPPOmPEOn) with a series of cationic surface-active ionic liquids in aqueous solutions have been investigated. The cationic surface-active ionic liquids include 1-alkyl-3-methylimidazolium bromide (CnmimBr, n?=?8, 10, 12, 14, 16) and N-alkyl-N-methylpyrrolidinium bromide (CnMPB, n?=?12, 14, 16). For different polymer-surfactant systems, the critical aggregation surfactant concentration (cac), the surfactant concentration to form free micelles (C m), and the saturation concentration of surfactant on the polymer chains (C 2) were determined using isothermal titration microcalorimetry (ITC) and conductivity measurements. The structure of the formed aggregates depended strongly on the hydrophobicity of the surfactant and the ratio of polymer/surfactant concentration. For C8mimBr, there were not any micelle-like surfactant?Cpolymer clusters detected in the solution, and only micelles appeared. For other surfactants, the polymer?Csurfactant aggregates were formed in the solution, which was verified by the appearance of a broad endothermic peak in the ITC thermograms. The intensity of polymer?Csurfactant interaction increased with the hydrophobicity of the surfactants and the polymers but was not affected by the surfactant headgroups.  相似文献   

12.
The wetting properties of pulmonary surfactant aqueous solutions with respect to solid surfaces with different degree of hydrophobicity have been studied. The contact angles θ of drops from a pulmonary surfactant solution onto SiO2-glass surfaces have been measured as a function of their degree of hydrophobicity θ w. The completely hydrophilic SiO2-glass surface is essentially hydrophobized by the animal-derived pulmonary surfactant Curosurf. The hydrophobization depends on the surfactant concentration—the contact angles increase with increasing the Curosurf concentration C s in the low concentration range, but they remain almost constant in a wide range of C s >90 μg/ml. Additions like NaCl and bovine serum albumin influence the θ-values. The contact angles θ naturally increase with increasing θ w but this dependence is not linear—the curve steepens at larger θ w values. The thickness h of the wetting thin liquid films from Curosurf aqueous solutions depends on the hydrophobicity θ w of the solid surface and the h(θ w) curves always pass a minimum. The h-values, as well as the h(θ w) curve, are mainly determined by the steric and hydrophobic disjoining pressures, which depend on the orientations and conformations of the molecules adsorbed on the solid surface from the very complicated multi-component aqueous solutions.  相似文献   

13.
ApparentpK a values of thymol blue solubilized in Aerosol-OT reversed micelles in carbon tetrachloride at 25 °C were determined spectrophotometrically. The effects of the Aerosol-OT concentration and the (water)/(surfactant)R ratio were investigated. The apparentpK a values increase as a function of increasing (R). All the measuredpK a values are less than that in water. The decrease ranges from 1.23 units (detergent=0.4 M,R=1.39) to 0.42 units (Aerosol-OT=0.6 M,R=9,25). These results are rationalized in terms of decreased hydronium ion activity in the micellar core due to its binding to the detergent SO3 ? groups.  相似文献   

14.
Three proteins (horse liver alcohol dehydrogenase, ribonuclease, lysozyme) were solubilized in hydrocarbon with the help of reverse micelles formed by aqueous di(2-ethyl-hexyl) sodium sulfosuccinate (AOT). Sedimentation and diffusion coefficients of the micellar aggregates were measured with an analytical ultracentrifuge. Partial specific volumes were also evaluated from density measurements. The molecular weight of the protein-containing reverse micelles (M t ) could thus be determined for each protein system at various w0 values (w0 - [H2O]/[AOT]). For horse liver alcohol dehydrogenase at w0 = 46.4, for example,M t is ca. 2,670,000 Daltons; for lysozyme at wo = 22.5,M t is ca. 323,000 Daltons and increases by increasing w0. On the basis of these experimentally determined molecular weights, a structural model for the protein-containing reverse micelle is proposed. The model is based upon the assumption that the protein is confined in the water pool of a spherical micelle, and that the inner core volume is the sum of the protein volume and the volume of all water molecules present in a micelle. It is possible then to calculate the micellar structure at each w0 value. For example, in the case of ribonuclease at w0 = 20, the inner core radius is ca. 37.5 A, and a layer of water of ca. 22 A separates the protein surface from the surfactant layer. The possible implications of this model for the reactivity of enzymes solubilized in hydrocarbons by reverse micelles are discussed.  相似文献   

15.
Aqueous solutions of anionic surfactant, sodium oleate (NaOA), have been studied by means of steady-state shear rheology and dynamic oscillatory technique. The micellar structure can be changed upon the addition of NaCl, Na2CO3 and NaCl/NaOH while NaOA concentration is maintained at 0.060 M. These systems except NaOA/NaCl show high viscosity and strong viscoelasticity. The hydroxide ion is very important for the formation of wormlike micelles. The anions of salts also have effect on the rheological properties of wormlike micelles. Three parameters: intersection frequency ωi, plateau modulus G0 and relaxation time τ are also discussed. The Maxwell model and Cole-Cole plot are applied to investigate the dynamic viscoelasticity of wormlike micelles. Variation in surfactant packing parameter RP can be used to explain the change of rheology and microstructure of the micelles.  相似文献   

16.
A lipase from a wild strain ofPenicillium citrinum was encapsulated in AOT/isooctane-reversed micelles, and the kinetic parameters were studied relative to triolein hydrolysis. Lipolytic activity was strongly dependent on the water amount in the system (Wo) and presented a bell-shaped curve for this parameter, with a maximum in the range of Wo 10–15. Optimum conditions for enzyme activity were pH 8.0 and 45?C. The influence of substrate concentration was also studied. The enzyme showed a Michaelis-Menten behavior and the apparent kinetics constants were calculated as beingV max.app. - 120 U/mg and Kmapp = 49.2 mM.  相似文献   

17.
The micellar properties of tetradecyltrimethylammonium nitrate (C14TANO3) in aqueous solutions in the temperature range of 10 to 35 °C and in aqueous solutions of benzyl alcohol (BzOH) at 25 °C were studied conductometrically. The specific conductivity data served for the evaluation of critical micelle concentration, cmc, and the degree of ionization of the micelles, , of the surfactant. From the temperature dependence of the cmc the thermodynamic parameters for micellization of C14TANO3 were calculated by applying Mullers modified equation. BzOH was found to affect strongly the cmc and values of the surfactant. The plot of the cmc/cmco ratio (where cmco is for pure water) as a function of BzOH molality, exhibits a characteristic break, which was attributed to the commencement of self-association of BzOH in aqueous solution at a molality of ca. 0.05. By applying the theoretical treatment suggested by Motomura for binary surfactant systems, the molar fraction of BzOH in the micelles at cmc, was estimated as a function of molality of the alcohol. C14TANO3 appears to be slightly more hydrophobic compared to the corresponding bromide.  相似文献   

18.
The enthalpies of mixing of some n-nitriles (from acetonitrile to valeronitrile) aqueous solutions with dodecyltzimethylammonium bromide, sodium dodecylsulfate and dodecyltzimethylammonium oxide micellar solutions were determined. The measurements were performed by systematically changing the surfactant concentration at a given solute concentration. The experimental enthalpies were rationalized in terms of the standard enthalpy of transfer of solute from the aqueous to the micellar phase and of the distribution constant between the two phase. Information on the effect of the nature of the surfactant on the standard thermodynamic quantities of transfer(G t o , H t o , TS t o ) is reported. The present data are compared to those previously reported for primary alcohols and the solubilizing properties shown by the different types of micelles are discussed.  相似文献   

19.
The association of many classes of surface active molecules into micellar aggregates is a well-known phenomenon. Micelles are often drawn as static structures of spherical aggregates of oriented molecules. However, micelles are in dynamic equilibrium with surfactant monomers in the bulk solution constantly being exchanged with the surfactant molecules in the micelles. Additionally, the micelles themselves are continuously disintegrating and reforming. The first process is a fast relaxation process typically referred to as τ1. The latter is a slow relaxation process with relaxation time τ2. Thus, τ2 represents the entire process of the formation or disintegration of a micelle. The slow relaxation time is directly correlated with the average lifetime of a micelle, and hence the molecular packing in the micelle, which in turn relates to the stability of a micelle. It was shown earlier by Shah and coworkers that the stability of sodium dodecyl sulfate (SDS) micelles plays an important role in various technological processes involving an increase in interfacial area, such as foaming, wetting, emulsification, solubilization and detergency. The slow relaxation time of SDS micelles, as measured by pressure-jump and temperature-jump techniques was in the range of 10−4–101 s depending on the surfactant concentration. A maximum relaxation time and thus a maximum micellar stability was found at 200 mM SDS, corresponding to the least foaming, largest bubble size, longest wetting time of textile, largest emulsion droplet size and the most rapid solubilization of oil. These results are explained in terms of the flux of surfactant monomers from the bulk to the interface, which determines the dynamic surface tension. The more stable micelles lead to less monomer flux and hence to a higher dynamic surface tension. As the SDS concentration increases, the micelles become more rigid and stable as a result of the decrease in intermicellar distance. The smaller the intermicellar distance, the larger the Coulombic repulsive forces between the micelles leading to enhanced stability of micelles (presumably by increased counterion binding to the micelles). The Center for Surface Science & Engineering at the University of Florida has developed methods using stopped-flow and pressure-jump with optical detection to determine the slow relaxation time of micelles of nonionic surfactants. The results show relaxation times τ2 in the range of seconds for Triton X-100 to minutes for polyoxyethylene alkyl ethers. The slow relaxation times are much longer for nonionic surfactants than for ionic surfactants, because of the absence of ionic repulsion between the head groups. The observed relaxation time τ2 was related to dynamic surface tension and foaming experiments. A slow break-up of micelles, (i.e. a long relaxation time τ2) corresponds to a high dynamic surface tension and low foamability, whereas a fast break-up of micelles, leads to a lower dynamic surface tension and higher foamability. In conclusion, micellar stability and thus the micellar break-up time is a key factor in controlling technological processes involving a rapid increase in interfacial area, such as foaming, wetting, emulsification and oil solubilization. First, the available monomers adsorb onto the freshly created interface. Then, additional monomers must be provided by the break-up of micelles. Especially when the free monomer concentration is low, as indicated by a low CMC, the micellar break-up time is a rate limiting step in the supply of monomers, which is the case for many nonionic surfactant solutions. Therefore, relaxation time data of surfactant solutions enables us to predict the performance of a given surfactant solution. Moreover, the results suggest that one can design appropriate micelles with specific stability or τ2 by controlling the surfactant structure, concentration and physico-chemical conditions, as well as by mixing anionic/cationic or ionic/nonionic surfactants for a desired technological application.  相似文献   

20.
The effect of concentration on the behavior of high-molecular-mass sodium polymethacrylate (M w = 3.9 × 105) in a 0.1 M NaCl aqueous solution was studied by the methods of dynamic and static light scattering, and capillary and rotational viscometry. It was shown that the concentration corresponding to the formation of fluctuation network of polyelectrolyte entanglements (6.4%) is substantially higher than the crossover concentration (0.25%). This fact indicates the existence of a wide concentration interval for a semidilute unentangled polyanion solution. The introduction of minor amounts of micelles of cationic amphiphilic polystyrene-poly(N-ethyl-4-vinylpyridinium bromide) diblock copolymer is accompanied by the development of the network of an interpolyelectrolyte complex. The junctions of this network are diblock copolymer micelles that are linked to polyanion macromolecules via salt bonds. The formation of interpolyelectrolyte network leads to the additional structuring of the sodium polymethacrylate solution and an increase in its viscosity. The growth in viscosity is most pronounced in the concentration region above the concentration corresponding to the formation of entanglement network of the free polyanion.  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