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.
Tryptophan octyl ester (TOE) represents an important model for membrane-bound tryptophan residues. In this article, we have explored the effect of sphere-to-rod transition of sodium dodecyl sulfate micelles on the dynamics of the membrane-bound tryptophan analogue, TOE, utilizing a combination of fluorescence spectroscopic approaches which include red edge excitation shift (REES). Our results show that REES and fluorescence spectroscopic parameters such as lifetime, anisotropy and acrylamide quenching of micelle-bound TOE are sensitive to the change in micellar organization accompanied by the sphere-to-rod transition.  相似文献   

2.
The distribution of indole and tryptophan derivatives between sodium dodecyl sulfate (SDS) micellar and aqueous phases was analyzed using conventional methods of ultraviolet (UV) absorption spectroscopy and measurement of fluorescence quenching by succinimide. On the assumption of a simple pseudo-phase equilibrium between both phases the distribution coefficient was easily obtained by the measurement of the ratioR pv of the absorbance intensity in the peak to that in the valley of the UV spectra or the fluorescence quenching constant Ksv. The possibilities and limitations of utilizing the ratio of the collisional quenching constant estimating from theK sv value in the micellar phase to that in the aqueous phase for a measure of the polarity of the microenvironment around the tryptophan derivatives in the SDS micelle is discussed in comparison with theR pv values for the UV spectra. The indole ring in the derivatives in the SDS micelle is localized near or on the micelle-water interface with its imino group directed toward the aqueous phase. Thus it can serve as a feasible model for interpreting the distribution coefficients andR pv values obtained for the various indole and tryptophan derivatives.Abbreviations UV ultraviolet - SDS sodium dodecyl sulfate - ATEE N-acetyl-l-tryptophan ethyl ester - ATA N-acetyl-l-tryptophan-amide - CMC critical micelle concentration  相似文献   

3.
Luminescent heteroleptic ruthenium(II) complexes of type RuLnX3–n [L = 1,10‐phenanthroline (phen), X = 4,7 diphenyl phenanthroline disulfonate, (dpsphen) n = 0,1,2,3] were synthesized and their photophysical properties investigated in homogeneous and cationic (CTAB), anionic (SDS) and nonionic (Triton X‐100) micelles. The luminescent quantum yield and lifetime of the complexes were found to increase in the presence of micellar media and on the introduction of a disulfonate ligand into the coordination sphere. Both electrostatic and hydrophobic interactions play an important role in the micellar media. Thus, by changing the nature of the ligands and the medium, we were able to tune the photophysical properties of Ru(II) complexes. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

4.
Antioxidants solubilized in micellar solutions can change micellar properties like the size and shape of micelles, critical micellar concentration (cmc) and viscosity. Interactions arising between antioxidants and the surfactant determine the locations of antioxidants and vice versa. The location and interaction are dependent on the type of both the antioxidant and surfactant. Influences of various antioxidants on the physical and structural properties were tested in micellar systems of cationic CTAB, non-ionic Brij 58 and anionic SDS. The antioxidants used to investigate the effects of gradually increasing lipophilicity were gallic acid (GA) and the gallate esters from methyl to octyl gallate (MG-OG). Hydroxy cinnamic acids (HCAs) like -coumaric acid (pC), caffeic acid (CA), ferulic acid (FA) and sinapic acid (SA) were employed to observe effects of functional groups like hydroxyl and methoxy groups. Micellar size and shape determined by small angle neutron scattering (SANS), viscosity and cmc were chosen to characterize the antioxidant influence. In Brij 58 systems propyl gallate (PG) did not affect the cmc or aggregation number but decreased micellar size slightly due to an intercalation of PG into the region of the polyoxyethylene chain and the first adjacent alkyl chain methylene groups. In SDS systems the micellar size and cmc decreased in the presence of PG. This was attributed to PG residing in the Stern layer. However, in CTAB systems micelles swelled at low PG concentration and in the presence of GA, while higher PG concentrations and more lipophilic antioxidants led to a sphere-to-rod transition with a simultaneous increase in viscosity and decrease in cmc. This revealed the intercalation of antioxidants in the palisade layer of CTAB micelles entering into strong interactions of electrostatic and hydrophobic origin. It could be demonstrated that the interactions became stronger the more lipophil an antioxidant is and the more hydroxyl groups are attached to the aromatic ring. Differences in the location and interaction of antioxidant and micelles are proposed as being responsible for the effectiveness of antioxidants.  相似文献   

5.
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   

6.
The effects of pressure and temperature on the critical micelle concentration (CMC) of sodium dodecylsulfate (SDS) were measured in the presence of various concentrations of an inhalation anesthetic, methoxyflurane. The change in the partial molal volume of SDS on micellization, ΔVm, increased with the increase in the concentration of methoxyflurane. The CMC-decreasing power, which is defined as the slope of the linear plot between ln(CMC) vs. mole fraction of anesthetic, was determined as a function of pressure and temperature. Since the CMC-decreasing power is correlated to the micelle/water partition coefficient of anesthetic, the volume change of the transfer (ΔVpo) of methoxyflurane from water to the micelle can be determined from the pressure dependence of the CMC-decreasing power. The value of ΔVpo amounts 6.5±1.8 cm3·mol?1, which is in reasonable agreement with the volume change determined directly from the density data, 5.5±0.6 cm3 · mol?1. Under the convention of thermodynamics, this indicates that the application of pressure squeezes out anesthetic molecules from the micelle. The transfer enthalpy of anesthetic from water to the micelle is slightly endothermic. The partial molal volume of methoxyflurane in the micelle (112.0 cm3·mol?1) is smaller than that in decane (120.5 cm3·mol?1) and is larger than that in water (108.0 cm3·mol?1). This indicates that the anesthetic molecules are incorporated into the micellar surface region, i.e., the palisade layer of the micelle in contact with water molecules, rather than into the micelle core.  相似文献   

7.
Structural transition can be induced in charged micelles by increasing the ionic strength of the medium. We have monitored the organization and dynamics of the functionally important tryptophan residues of gramicidin in spherical and rod-shaped sodium dodecyl sulfate micelles utilizing a combination of wavelength-selective fluorescence and related fluorescence approaches. Our results show that tryptophans in gramicidin, present in the single-stranded beta(6.3) conformation, experience slow solvent relaxation giving rise to red edge excitation shift in spherical and rod-shaped micelles. In addition, changes in fluorescence polarization with increasing excitation or emission wavelength reinforce that the gramicidin tryptophans are localized in motionally restricted regions of these micelles. Fluorescence quenching experiments using acrylamide as a quencher of tryptophan fluorescence show that there is reduced water penetration in rod-shaped micelles. Taken together, we show that gramicidin conformation and dynamics is sensitive to the salt-induced structural transition in charged micelles. In addition, these results demonstrate that deformation of the host assembly could modulate protein conformation and dynamics.  相似文献   

8.
Chromatography of glutaraldehyde-fixed skim-milk on controlled-pore glass (CPG-10, 300 nm) gave three micellar fractions whose averaged diameters, measured by electron microscopy, decreased progressively with increasing elution volume. Casein micelles with diameters up to 680 nm were detected. The casein composition of the same fractions from unfixed skim-milk was determined. As the fraction elution volume increased, κ-casein varied from 7.7 to 11.4% of total casein, giving αs/κ ratios of 6.1, 4.7 and 3.3.A plot of κ-casein content versus micelle surface-to-volume ratio for skim-milk and the column fractions approximated to a straight line. Re-calculation of the published results from two other studies also gave linear relationships between κ-casein content and surface area for artificial micelles. The three regression lines thus obtained had small intercepts. It was concluded that the data indicated the same fundamental structure for casein micelles, with a pre-dominant surface location for κ-casein, whether the micelles are natural or artificial and whether they are aggregated or by Ca2+ alone oy Ca2+ together with calcium phosphate-citrate complex.  相似文献   

9.
Structural transitions involving shape changes play an important role in cellular physiology. Such transition can be conveniently induced in charged micelles by increasing ionic strength of the medium. Shape changes have recently been shown to result in altered packing and lowering of micellar polarity. As a consequence of reduced polarity, the ionization states of micelle-bound molecules vary in micelles of different shape. The changes in micellar organization and dynamics due to structural transition can be effectively monitored utilizing the red edge excitation shift (REES). These changes are influenced by the position (location) of the probe in the micelle, i.e., the region of the micelle being monitored. Changes in organization and dynamics of probes and peptides upon structural transition are discussed with representative examples. We envisage that the reduction in micellar polarity and tighter packing upon structural transition represent important factors in the incorporation of drugs in micelles (nano-carriers), since micellar polarity plays a crucial role in the incorporation of drugs.  相似文献   

10.
An absorption and fluorescence spectral and temporal studies on the solubilzation properties of adrenaline in micellar environment in sodium dodecyl sulfate (SDS) and in tetradodecyltrimethyl ammonium bromide (TTABr) has been carried out. Observed Stokes shifts have been correlated with polarity parameters which allowed an estimate of the dielectric constant of the adrenaline environment in SDS and TTABr micelles at 44 and 58, respectively. Experiments with methanol-water mixtures indicate that the hydrogen bonding formation with solvent and the hydrophilic nature of adrenaline influence its solubilization in micelles. Fluorescence and anisotropy decay analysis has shown that neutral adrenaline in SDS micelle is partitioned between aqueous phase (70%) and less polar, micellar phase (30%) and the interactions are limited to the Guy-Chapman layer without deeper penetration into micellar hydrophobic core.  相似文献   

11.
The critical micellar concentration (CMC) of stearoylcarnitine was determined at different pH values at room temperature by fluorescence spectroscopy, monitoring the spectral changes of 8-anilinonaphthalene-1-sulfonate (ANS). The CMC was found to vary with pH, increasing from about 10 μM at pH 3.0 to ca. 25 μM at pH 7.0, but decreasing slightly with further increase in pH to approximately 19 μM at pH 10.0. Differential scanning calorimetry (DSC) shows that stearoylcarnitine dispersed in water at low concentration undergoes a broad thermotropic phase transition at 44.5°C, with a transition enthalpy of 15.0 kcal/mol. The transition temperature (T t) shifts to ca. 50.5°C in the presence of 1 mM EDTA or when the concentration is increased significantly. The turbidity of aqueous dispersions of stearoylcarnitine was found to be considerably high at low temperatures, which decreases quite abruptly over a short temperature range, indicating that a transition occurs from a phase of large aggregates to one of much smaller aggregates, most likely micelles. The phase transition temperature was determined as 29.1°C at pH 3.0, which increased with increasing pH up to a value of 55.3°C at pH 8.6 and remains nearly constant thereafter up to pH 11.2. The pH dependence of CMC and T t suggest that the pK a of the carboxyl group of long chain acylcarnitines shifts to higher temperatures upon aggregation (micelles or bilayer membranes).  相似文献   

12.
The effect of sodium salicylate (NaSal) on the spherical-to-threadlike micelle shape transition in 3-hexadecyloxy-2-hydroxy-propyl trimethyl ammonium bromide (R16HTAB) solution was studied using molecular dynamics simulation. The simulations were started from a preassembled infinitely long threadlike micelle of R16HTAB. By analyzing the aggregation morphologies and structural details, we find that the preassembled threadlike micelle in the absence of NaSal was unstable and assembled into a spherical micelle. While in the presence of NaSal, the threadlike micelle exhibited fluctuations but remained the threadlike shape during the long simulation run. The Sal? ions were found to penetrate inside the micelle, which promoted the junction between the surfactant and salicylate counterion. The aromatic Sal? ions located in the surfactant headgroup region with their phenyl groups pointing toward the interior core region of the micelle. From another simulation started with two individual spherical micelles, we found that the Sal? ions can link the two spherical micelles into a long threadlike micelle, in accordance with a mode proposed by experimental studies. Our studies showed that the H-bonds and electrostatic interactions between the Sal? ions and the surfactants played an important role in micellar growth and stabilising the threadlike micelle.  相似文献   

13.
The correlation between membrane protein solubilisation and detergent aggregation in aqueous solution is studied for a series of n-alkyl-β-d-maltosides (CxG2 with x = 10, 11, 12 being the number of carbon atoms in the alkyl chain) using the trimeric photosystem I core complex (PSIcc) of oxygenic photosynthesis from Thermosynechococcus elongatus as model protein. While protein solubilisation is monitored via the turbidity of the solution, the aggregation behavior of the detergent is probed via the fluorescence spectrum of the polycyclic aromatic hydrocarbon pyrene. In addition, changes of the fluorescence spectrum of PSIcc in response to formation of the detergent belt surrounding its hydrophobic surface are investigated. Solubilisation of PSIcc and aggregation of detergent into micelles or belts are found to be strictly correlated. Both processes are complete at the critical solubilisation concentration (CSC) of the detergent, at which the belts are formed. The CSC depends on the concentration of the membrane protein, [prot], and is related to the critical micelle concentration (CMC) by the empirical law ln(CSC/CMC) = 0 [prot], where the constant 0 = (2.0 ± 0.3) μM−1 is independent of the alkyl chain length x. Formation of protein-free micelles below the CSC is not observed even for x = 10, where a significant excess of detergent is present at the CSC. This finding indicates an influence of PSIcc on micelle formation that is independent of the binding of detergent to the hydrophobic protein surface. The role of the CSC in the optimisation of membrane protein crystallisation is discussed.  相似文献   

14.
Oil extraction from microalgae for biodiesel production   总被引:3,自引:0,他引:3  
This study examines the performance of supercritical carbon dioxide (SCCO2) extraction and hexane extraction of lipids from marine Chlorococcum sp. for lab-scale biodiesel production. Even though the strain of Chlorococcum sp. used in this study had a low maximum lipid yield (7.1 wt% to dry biomass), the extracted lipid displayed a suitable fatty acid profile for biodiesel [C18:1 (∼63 wt%), C16:0 (∼19 wt%), C18:2 (∼4 wt%), C16:1 (∼4 wt%), and C18:0 (∼3 wt%)]. For SCCO2 extraction, decreasing temperature and increasing pressure resulted in increased lipid yields. The mass transfer coefficient (k) for lipid extraction under supercritical conditions was found to increase with fluid dielectric constant as well as fluid density. For hexane extraction, continuous operation with a Soxhlet apparatus and inclusion of isopropanol as a co-solvent enhanced lipid yields. Hexane extraction from either dried microalgal powder or wet microalgal paste obtained comparable lipid yields.  相似文献   

15.
The amount of casein found in the milk of the tammar wallaby increases as lactation progresses. The increase is due to increasing amounts of β-casein; the α-casein remains largely constant. The α-casein is the more highly phosphorylated; the most abundant form is the 10-P, throughout lactation. The level of phosphorylation of β-casein shifts to lower average values in late lactation, possibly indicating the enzymatic reaction is overloaded by the increasing amounts of β-casein. Unlike bovine casein micelles, the wallaby micelles are not completely disrupted at pH 7.0 by sequestration of their calcium content with ethylene diamine tetraacetic acid (EDTA). Complete disruption only follows the addition of sodium dodecyl sulphate, indicating considerably greater importance for hydrophobic bonds in maintaining their integrity. This micellar behaviour indicates that, despite the evolutionary divergence of marsupials millennia ago, the caseins of wallaby milk assemble into micelles in much the same fashion as in bovine milk.  相似文献   

16.
17.
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   

18.
We have used cryo-transmission electron microscopy (cryo-TEM) for inspection of aggregates formed by dimyristoylphosphatidylcholine (DMPC) and dihexanoylphosphatidylcholine (DHPC) in aqueous solution at total phospholipid concentrations cL≤5% and DMPC/DHPC ratios q≤4.0. In combination with ocular inspections, we are able to sketch out this part of phase-diagram at T=14-80 °C. The temperature and the ratio q are the dominating variables for changing sample morphology, while cL to a lesser extent affects the aggregate structure. At q=0.5, small, possibly disc-shaped, aggregates with a diameter of ∼6 nm are formed. At higher q-values, distorted discoidal micelles that tend to short cylindrical micelles are observed. The more well-shaped discs have a diameter of around 20 nm. Upon increasing q or the temperature, long slightly flattened cylindrical micelles that eventually branch are formed. A holey lamellar phase finally appears upon further elevation of q or temperature. The implications for biological NMR work are two. First, discs prepared as membrane mimics are frequently much smaller than predicted by current “ideal bicelle” models. Second, the q≈3 preparations used for aligning water-soluble biomolecules in magnetic fields consist of perforated lamellar sheets. Furthermore, the discovered sequence of morphological transitions may have important implications for the development of bicelle-based membrane protein crystallization methods.  相似文献   

19.
《Luminescence》2003,18(2):97-102
Benzo[a]pyrene causes cancer at cellular level and is widely present in the environment. Conventional spectroscopic methods for analysis of this compound need a pre‐separation procedure due to severe spectral overlap from other polycyclic aromatic hydrocarbons. We report a simple method that avoids spectral overlap of benzo[a]pyrene from other impurities or polycyclic aromatic hydrocarbons (PAHs), thus it can easily identify benzo[a]pyrene in a complex PAH mixture. The method could easily identify benzo[a]pyrene in an 18‐component PAH mixture. Calibration plots in methanol solution and in micellar media show a good linearity (R > 0.9997) in the benzo[a]pyrene concentration range generally found in the environment. The method gives a detection limit of 1.52 × 10?9 mol/L in CTAB micellar medium and 2.55 × 10?9 mol/L in methanol solution. The proposed method is selective, sensitive and fast. The fluorescence response of benzo[a]pyrene is found to be a potential candidate to sense the critical micellar concentration (CMC) of CTAB micelles. Copyright © 2003 John Wiley & Sons, Ltd.  相似文献   

20.
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   

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