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.
2.
The hydrogenation of various nitrogen-, oxygen- or sulfur-heterocyclic aromatic compounds by various surfactant-stabilized aqueous rhodium(0) colloidal suspensions was investigated. The nanocatalysts in the size range of 2.1-2.4 nm have been synthesized by reducing RhCl3 · 3H2O with sodium borohydride and were stabilized by highly water soluble N,N-dimethyl-N-cetyl-N-(2-hydroxyethyl)ammonium bromide or chloride salts. The catalytic reactions were performed under mild reaction conditions, namely room temperature and under atmospheric hydrogen pressure. The influence of the bromide or chloride nature of the surfactant counter-ion on the recycling of the aqueous phase containing the Rh(0) particles was studied.  相似文献   

3.
The structure and physical properties of solloids (surfactant aggregates adsorbed on surfaces) adsorbed on particles are of general interest. The relationship between solloid structure and properties of hexadecyltrimethylammonium bromide (HTAB), cetylpyridinium chloride (CPC) and cetylpyridinium salicylate (CPS) adsorbed on silica particles was studied by electron paramagnetic resonance (EPR) spectroscopy using the spin-probes peroxylaminedisulfonate (PADS) and 4-[N,N-dimethyl-N-(n-hexadecyl)ammonium]-2,2,6,6-tetramethylpiperidinyl-N-oxy bromide (HTAB*). Using HTAB* incorporated in HTAB, CPC and CPC solloids and comparing the results to those in micelles, it was determined that for silica around pH 4 the solloids are very similar in properties to the micelles. This is consistent with a linear solvation–energy relationship (LSER) analysis of solute equilibration data which indicates that at pH 5 HTAB solloids have similar properties to HTAB micelles. The PADS spin-probe appears to be more sensitive to changes in the properties of the double layer, and substantial differences were observed between HTAB, CPC and CPS and as a function of HTAB concentration for HTAB solloids on silica.  相似文献   

4.
The effects of surfactants on the human liver hexosaminidase A-catalysed hydrolysis of Gm2 ganglioside were assessed. Some non-ionic surfactants, including Triton X-100 and Cutscum, and some anionic surfactants, including sodium taurocholate, sodium dodecyl sulphate, phosphatidylinositol and N-dodecylsarcosinate, were able to replace the hexosaminidase A-activator protein [Hechtman (1977) Can. J. Biochem. 55, 315–324; Hechtman & Leblanc (1977) Biochem. J. 167, 693–701) and also stimulated the enzymic hydrolysis of substrate in the presence of saturating concentrations of activator. Other non-ionic surfactants, such as Tween 80, Brij 35 and Nonidet P40, and anionic surfactants, such as phosphatidylethanolamine, did not enhance enzymic hydrolysis of Gm2 ganglioside and inhibited hydrolysis in the presence of activator. The concentration of surfactants at which micelles form was determined by measurements of the minimum surface-tension values of reaction mixtures containing a series of concentrations of surfactant. In the case of Triton X-100, Cutscum, sodium taurocholate, N-dodecylsarcosinate and other surfactants the concentration range at which stimulation of enzymic activity occurs correlates well with the critical micellar concentration. None of the surfactants tested affected the rate of hexosaminidase A-catalysed hydrolysis of 4-methylumbelliferyl N-acetyl-β-d-glucopyranoside. Both activator and surfactants that stimulate hydrolysis of Gm2 ganglioside decrease the Km for Gm2 ganglioside. Inhibitory surfactants are competitive with the activator protein. Evidence for a direct interaction between surfactants and Gm2 ganglioside was obtained by comparing gel-filtration profiles of 3H-labelled GM2 ganglioside in the presence and absence of surfactants. The results are discussed in terms of a model wherein a mixed micelle of surfactant or activator and GM2 ganglioside is the preferred substrate for enzymic hydrolysis.  相似文献   

5.
The stopped-flow kinetics of the reaction between oligopeptides containing tryptophan residues andN-bromosuccinimide (NBS) were studied in 50 mM sodium phosphate buffer (pH 7.0) containing sodium dodecyl sulfate (SDS). Decreases in the reaction rates attributable to the interaction between oligopeptides and SDS were observed, and oligopeptides studied were classified into types I and II on the basis of the interaction modes. Type I oligopeptides were dissolved in SDS micelles; type II oligopeptides interacted cooperatively with SDS monomers. The manner of interaction between SDS and oligopeptides of type II could be interpreted by a simple equilibrium relation: oligopeptide+n·(SDS)=oligopeptide·(SDS) n .  相似文献   

6.
Reverse non-equilibrium (RNE) method is applied to investigate the rheological properties of surfactant solutions by using dissipative particle dynamics algorithm. Results show that the surfactant solutions exhibit a shear-thinning behaviour. With the strengthening of shear, the worm-like micelles are gradually oriented in the x direction and then broken up into small spherical micelles. The process is also shown by the decrease of viscosity, which starts quickly, levels off for moderate shear rates and speeds up again over a critical value. We find that this critical value is independent of surfactant concentration. Compared with the shear flow method, the RNE method is credible for moderate shear rate and has more computational efficiency.  相似文献   

7.
Four chiral analogues of the surfactant Aerosol-OT (AOT) have been synthesized and characterized. All of them form reverse micelles in apolar solvents in the w0 range 0–30 (w0 = [water]/[tenside]). Reverse micellar solutions have been investigated by UV absorption and circular dichroism spectroscopies with the aim of clarifying whether the formation of the macromolecular micellar structure induces the appearance of new chromophoric bands or perturbs the existing ones. Methanolic solutions of the surfactants, in which no micellar aggregates are formed, were taken as references. One of the products 1(S),1′(S)-dimethylbisheptylsulphosuccinate sodium salt (MH-AOT) was capable of forming reverse micelles of relatively high water content (w0 up to 40) and this process was accompanied by a specific increase in the intensity of the circular dichroism band associated with the ester absorbance of the molecule. As no concomitant changes were seen in the UV absorbance spectrum, it was concluded that this observation reflected conformational events occurring within the surfactant rather than chromophoric perturbation. These results are qualitatively similar to those found recently for lecithin reverse micelles which, however, form gels at sufficiently high water contents. The chiroptical properties of these supramolecular aggregates are compared with those of covalent macromolecular systems such as polypeptides.  相似文献   

8.
The objectives of this study were to determine the concentrations of free benzalkonium chloride (BAC) and apparent partitions coefficients (K m) in micelle solutions and to explore its application in formulation development. Ultrafiltration (UF) was carried out using 10K Nanosep® devices and centrifugation at 5,000 rpm for 5 min. The separation of free BAC from micellar solutions was also conducted using ultracentrifugation (UC) method for the comparison with UF method. Capillary electrophoresis method was used for the identification of micelles. Results showed that a UF method was applicable for quantitatively evaluating BAC–micelle interaction in micellar solutions. Unlike UF, UC could not completely separate free BAC from the micelles. The free BAC concentrations in the micelle solutions decreased with increasing surfactant concentrations. Among polysorbate 80, cremophor EL, and tyloxapol, BAC had the highest K m in polysorbate 80 solutions. The K m was significantly lower in non-buffered aqueous solutions than that in citric buffers. Moreover, increasing surfactant concentrations led to reducing antimicrobial activity. The UF is a rapid and accurate method that minimally alters the micellar equilibrium for the determination of free BAC and K m in micellar solutions. In conclusion, free BAC concentration, which is a function of surfactant type, surfactant concentration, and ion strength of solution, is likely associated with the antimicrobial activity.  相似文献   

9.
Abstract

The effect of C12E8 (polyoxyethylene 8 lauryl ether) and sodium laurate on the structural properties of soybean lecithin vesicles was studied in different concentrations of surfactants, using the fluorescent probes NBD-PE (N-7-nitro-2,1,3-benzodiazoyl phosphatidylethanolamine) and N-Rh-PE (N-lissamine rhodamine B sulfonyl phosphatidylethanolamine). Direct energy transfer studies were carried out in labelled vesicles with addition of surfactants. Rhodamine emission with maximum at β585 nm was detected by excitation of NBD at λ = 473 nm. This fact is caused by direct energy transfer process from NBD to rhodamine. The yield of this process decreases with increasing amounts of surfactants, indicating that the average spatial separation of the lipid probes increases, as a result of the enlargement of the vesicle size and due to alteration of its structure to mixed micelles.  相似文献   

10.
A new nonionic reverse micellar system is developed by blending two nonionic surfactants, Triton X‐45 and Span 80. At total surfactant concentrations lower than 60 mmol/L and molar fractions of Triton X‐45 less than 0.6, thermodynamically stable reverse micelles of water content (W0) up to 30 are formed. Di(2‐ethylhexyl) phosphoric acid (HDEHP; 1–2 mmol/L) is introduced into the system for chelating transition metal ions that have binding affinity for histidine‐rich proteins. HDEHP exists in a dimeric form in organic solvents and a dimer associated with one transition metal ion, including copper, zinc, and nickel. The copper‐chelate reverse micelles (Cu‐RM) are characterized for their W0, hydrodynamic radius (Rh), and aggregation number (Nag). Similar with reverse micelles of bis‐2‐ethylhexyl sodium sulfosuccinate (AOT), Rh of the Cu‐RM is also linearly related to W0. However, Nag is determined to be 30–90 at W0 of 5–30, only quarter to half of the AOT reverse micelles. Then, selective metal‐chelate extraction of histidine‐rich protein (myoglobin) by the Cu‐RM is successfully performed with pure and mixed protein systems (myoglobin and lysozyme). The solubilized protein can be recovered by stripping with imidazole or ethylinediaminetetraacetic acid (EDTA) solution. Because various transition metal ions can be chelated to the reverse micelles, it is convinced that the system would be useful for application in protein purification as well as simultaneous isolation and refolding of recombinant histidine‐tagged proteins expressed as inclusion bodies. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2010  相似文献   

11.

Background

Respiratory distress syndrome in preterm babies is caused by a pulmonary surfactant deficiency, but also by its inactivation due to various conditions, including plasma protein leakage. Surfactant replacement therapy is well established, but clinical observations and in vitro experiments suggested that its efficacy may be impaired by inactivation. A new synthetic surfactant (CHF 5633), containing synthetic surfactant protein B and C analogs, has shown comparable effects on oxygenation in ventilated preterm rabbits versus Poractant alfa, but superior resistance against inactivation in vitro. We hypothesized that CHF 5633 is also resistant to inactivation by serum albumin in vivo.

Methodology/Principal Findings

Nineteen preterm lambs of 127 days gestational age (term = 150 days) received CHF 5633 or Poractant alfa and were ventilated for 48 hours. Ninety minutes after birth, the animals received albumin with CHF 5633 or Poractant alfa. Animals received additional surfactant if PaO2 dropped below 100 mmHg. A pressure volume curve was done post mortem and markers of pulmonary inflammation, surfactant content and biophysiology, and lung histology were assessed. CHF 5633 treatment resulted in improved arterial pH, oxygenation and ventilation efficiency index. The survival rate was significantly higher after CHF 5633 treatment (5/7) than after Poractant alfa (1/8) after 48 hours of ventilation. Biophysical examination of the surfactant recovered from bronchoalveolar lavages revealed that films formed by CHF 5633-treated animals reached low surface tensions in a wider range of compression rates than films from Poractant alfa-treated animals.

Conclusions

For the first time a synthetic surfactant containing both surfactant protein B and C analogs showed significant benefit over animal derived surfactant in an in vivo model of surfactant inactivation in premature lambs.  相似文献   

12.
Measurements of time-resolved fluorescence quenching have been performed in the binary lauroyllysophosphatidylcholine (LaLPC)/water system. The aggregation numbers, N, are determined for the micellar solution phase (Nmicelle ≈ 80) and the cubic liquid crystalline I1 phase (Ncub ≈ 90) at 298-303 K. When a quencher is present, the fluorescence decays for the hexagonal phase of the LaLPC/water system and for the bicontinuous cubic phase of monooleoylglycerol/water system are nonexponential, as expected for phase structures having long-range continuous apolar regions. Nuclear magnetic resonance (NMR) measurements of the lipid translational diffusion conclusively show that the cubic I1 phase consists of closed micelles. NMR spectra of 31P obtained at 202.4 MHz of this cubic phase exhibit a characteristic line shape, which is compatible with a phase structure containing short nonspherical micelles. A comparison between electron spin resonance (ESR) spin-label spectra recorded for a micellar solution and the cubic phases of the LaLPC and monooleoylglycerol systems are also shown to support a structure of closed micelles in the cubic I1 phase of the lysolecithin system.  相似文献   

13.
The hydrolysis of olive oil catalyzed by Candida rugosa lipase in sodium bis(2-ethylhexyl)sulfosuccinate (AOT)/isooctane and the synthetic sodium bis(2-ethylhexyl polyoxyethylene)sulfosuccinate (MAOT)/isooctane reverse micellar systems was investigated in a polysulfone hollow fiber membrane reactor with recycle of the reaction mixture. Lipase was completely retained by the membrane while olive oil and oleic acid freely passed through. The retention of reverse micelles depended on W 0 (molar ratio of water to surfactant). At an olive oil concentration of 0.23 mol l–1 the final substrate conversion in the MAOT micellar system was about 1.4 times of that in the AOT micellar system.  相似文献   

14.
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   

15.
Catalysis by laccase from Coriolus uersicolor solubilized in the ternary systems of surfactant/water/organic solvent type, namely, Aerosol OT/water/octane, Brij 56/water/cyclohexane and egg lecithin/water/octane + pentanol + methanol mixture, has been studied. The laccase activity is found to depend, in principle, not only on the water/surfactant molar ratio, but on the surfactant concentration (with its hydration degree being constant) as well. The following inferences should be emphasized. Firstly, in all the systems under study, the catalytic activity (kcat) of laccase entrapped into surfactant reversed micelles increases more than 50 times (when the surfactant concentration is extrapolated to zero) compared with the kcat value in aqueous solution. Secondly, the catalytic activity (kcat) of laccase entrapped in hydrated Aerosol OT aggregates, having lamellar, reversed cylindrical (hexagonal) and reversed micellar structure, depends greatly on the aggregate type. In other words, the phase transitions, i.e. an alteration in the packing of hydrated Aerosol OT molecules, evokes a sharp reversible change in the enzymatic activity. Thirdly, in the same phase, the catalytic activity of the solubilized enzyme depends on the linear dimensions of water cavities inside the surfactant aggregates (i.e. on the water content in the system under study). All these effects, regulating enzymatic activity, are probably caused by an alteration of the conformational mobility of laccase molecules incorporated into the inner polar cavities inside the surfactant aggregates.  相似文献   

16.
An enhanced thiosemicarbazide(TSC)–H2O2 chemiluminescence (CL) system was established and proposed as a new analytical method for determination of β‐lactam antibiotics, ampicillin sodium and amoxicillin at microgram levels. The method is based on the inhibition of CL emission accompanying oxidation of TSC by H2O2 in alkaline medium. The effect of anionic, cationic, and non‐ionic surfactants on the CL emission of the system was studied. Both N‐cetyl‐N,N,N‐trimethylammonium bromide (CTMAB) and Triton X‐100, unlike sodium dodecyl sulfate (SDS), reinforced the CL intensity and were efficient to approximately the same level. The effect of the presence of eight non‐aqueous solvents on the CL system was also investigated. Upon addition of both of the non‐ionic surfactant, Triton X‐100, and the non‐aqueous solvent, N,N‐dimethyl formamide (DMF), the intensity of the CL reaction was increased 100‐fold. This method allows the measurement of 25–545 µg amoxicillin, and 35–350 µg ampicillin sodium. The detection limits are 8 µg for amoxicillin and 9 µg for ampicillin sodium. The relative standard deviations of six replicate measurements of 200 µg amoxicillin and 200 µg ampicillin sodium were 1.9 and 2.1%, respectively. The effect of foreign species on the determination of amoxicillin and ampicillin sodium was also examined. The proposed method was successfully applied to the determination of ampicillin sodium and amoxicillin in some pharmaceutical dosage forms. Copyright © 2008 John Wiley & Sons, Ltd.  相似文献   

17.
The present investigation aims at studying the effect of mixed surfactant system of sodium lauryl sulphate (SLS) and alkyl polyglucosides (C10APG, C12APG and C12/14APG) on dissolution rate enhancement of poorly water soluble drug. Aceclofenac—a non-steroidal anti-inflammatory agent was used as a model drug as it has limited water solubility. The influence of the surfactant concentration in various blends on dissolution rate of Solid Dispersion (SD), prepared using solution method with ethanol as the solvent was studied and the advantage of mixed surfactant systems over the individual surfactants was illustrated by differences in the in-vitro dissolution profiles of SD. Physico chemical evaluation (critical micellar concentration, zeta potential and β-parameter calculations) was carried out to study the mixed surfactant systems. Solid mixtures were characterized by Infrared spectroscopy (FT-IR); X-ray diffraction studies (XRD) and scanning electron microscopy (SEM). It was seen that the dissolution rate of aceclofenac from SD increased with the increase in the APG proportion relative to SLS with the optimum ratio of 0.2 SLS:0.8 APG showing the best effect in all cases. Results obtained from physico-chemical evaluation (the decrease in the value of critical micelle concentration and higher negative value of β-parameters) suggested the existence of synergism between surfactants blends. The observed results in the dissolution rate enhancement could be attributed to the drug—surfactant interactions as evident from FT-IR, SEM and XRD results.  相似文献   

18.
α-Chymotrypsin (α-CT) activity was tested in aqueous media with the following cetyltrialkylammonium bromide surfactants in the series methyl, ethyl, propyl and butyl, different in the head group size, and for the sake of comparison also with the anionic sodium n-dodecyl sulfate and the zwitterionic myristyldimethylammonium propanesulfonate. N-glutaryl-l-phenylalanine p-nitroanilide hydrolysis rate was monitored at surfactant concentration above the critical micellar one. Only some cationic surfactants gave superactivity and the head group size had a major weight. The highest superactivity was measured in the presence of cetyltributylammonium bromide. The effect of both nature and concentration of three different buffers was also investigated. There is a dependence of enzyme superactivity on buffer type. Michaelis–Menten kinetics were found. The binding constants of substrate with micellar aggregates were determined in the used buffers and the effective improvement of reaction rate (at the same free substrate concentration in the medium) was calculated. kcat significantly increased while Km was little changed after correction to free substrate concentration. The ratio of kcat to Km was between 12 and 35 times higher than in pure buffer, depending on buffer and surfactant concentrations. The increase of α-CT activity (30%) was less important in the presence of 1×10−2 M tetrabutylammonium bromide, a very hydrophobic salt, unable to micellise. Fluorescence spectra showed differences of enzyme conformation in the presence of various surfactants.  相似文献   

19.
Summary A novel aqueous two-phase system containing hydrophobically modified ethylene oxide (HM-EO) and sodium dodecyl sulphate (SDS) was developed to enhance the selectivity of protein partitioning in two phases. Phase diagrams of HM-EO/H2O and HM-EO/SDS/H2O were measured, and the mechanism of interaction between HM-EO polymer and the anionic surfactant sodium dodecyl sulphate (SDS) was also discussed. It was found that the improvement of selectivity of protein partitioning was related to the increase of electrostatic potential difference between the two phases because of the charged network formed by mixed micelles of HM-EO and SDS in the bottom phase. With bovine serum albumin (BSA) and lysozyme as model proteins, some factors, such as pH, SDS concentration, conductivity and temperature of the system, were investigated for the influences of protein partition in HM-EO/SDS/H2O systems. The results showed that the addition of SDS not only changed the phase behaviour, but also played an important role in protein partitioning.  相似文献   

20.
The interaction of the surfactant octyl glucoside (OG) with dimyristoylphosphatidylcholine (DMPC), dipalmitoylphosphatidylcholine (DPPC), distearoylphosphatidylcholine (DSPC), and soy bean phosphatidylcholine (soy bean PC) was studied using high-sensitivity titration calorimetry. We determined the partition coefficient of OG between water and lipid bilayers and the transfer enthalpy of the surfactant by addition of lipid vesicles to OG monomers or vice versa. Comparison with the micellization enthalpy of the surfactant gives information on differences in the hydrophobic environment of OG in a liquid-crystalline bilayer or a micelle. The average partition coefficient P in mole fraction units for xe≈0.12–0.2 decreases slightly from 4152 at 27°C to 3479 at 70°C for DMPC and from 4260 to 3879 for soy bean PC, respectively. The transfer enthalpy ΔHT of OG into lipid vesicles is positive at 27°C and negative at 70°C. Its temperature dependence is larger for the incorporation of OG into DMPC than into soy bean PC vesicles. It is concluded that OG in DMPC vesicles is better shielded from water than in soy bean PC vesicles or in micelles. Titration calorimetry was also used to determine the phase boundaries of the coexistence region of mixed vesicles and mixed micelles in the systems OG/DMPC, OG/DPPC, OG/DSPC, and OG/soy bean PC vesicles at 70°C in the liquid-crystalline phase. DMPC and soy bean PC solubilization was also studied at 27°C to investigate the effect of temperature. The effective surfactant to lipid ratios at saturation, Resat, for all PCs studied are in the range between 1.33–1.72 and the ratios at complete solubilization, Resol, are between 1.79–3.06. At 70°C, the Resat values decrease with increasing chain length of the saturated PC. The ratios depend also slightly on temperature and the degree of unsaturation of the fatty acyl chains. For the OG/soy bean PC system, the coexistence range for mixed vesicles and mixed micelles is larger than for the corresponding PCs with saturated chains.  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