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.
In this paper, we describe the derivation and the validation of original RESP atomic partial charges for the N, N-dimethyl-dodecylamine oxide (LDAO) surfactant. These charges, designed to be fully compatible with all the AMBER force fields, are at first tested against molecular dynamics simulations of pure LDAO micelles and with a fragment of the lipid kinase PIK4A (DI) modeled with the QUARK molecular modeling server. To model the micelle, we used two distinct AMBER force fields (i.e. Amber99SB and Lipid14) and a variety of starting conditions. We find that the micelle structural properties (such as the shape, size, the LDAO headgroup hydration, and alkyl chain conformation) slightly depend on the force field but not on the starting conditions and more importantly are in good agreement with experiments and previous simulations. We also show that the Lipid14 force field should be used instead of the Amber99SB one to better reproduce the C(sp3)C(sp3)C(sp3)C(sp3) conformation in the surfactant alkyl chain. Concerning the simulations with LDAO-DI protein, we carried out different runs at two NaCl concentrations (i.e. 0 and 300 mM) to mimic, in the latter case, the experimental conditions. We notice a small dependence of the simulation results with the LDAO parameters and the salt concentration. However, we find that in the simulations, three out of four tryptophans of the DI protein are not accessible to water in agreement with our fluorescence spectroscopy experiments reported in the paper.  相似文献   

2.
Solubilization of large unilamellar 1,2-dioleoylphosphatidylcholine (DOPC) vesicles by N-dodecyl-N,N-dimethylamine-N-oxide (LDAO) was studied using turbidimetry. From turbidity data, the LDAO partition coefficient between the aqueous phase and DOPC bilayers was obtained. Using this partition coefficient, the LDAO:DOPC molar ratio in the bilayer was calculated and effects of LDAO on the bilayer stability, bilayer thickness and on the phosphohydrolase activity of sarcoplasmic reticulum Ca(2+) transporting ATPase (SERCA) reconstituted into DOPC were compared at the same LDAO:DOPC molar ratios in the bilayer. The sequence "bilayers in vesicles - bilayer fragments (flat mixed micelles) - tubular mixed micelles - globular mixed micelles" was suggested for the solubilization mechanism of DOPC vesicles from the combined turbidimetric and small-angle neutron scattering (SANS) results. The effective molecular packing parameter delta = 0.5, corresponding to the mixed bilayer - mixed tubular micelle transition, was calculated from fragmental DOPC and LDAO volumes at the molar ratio LDAO:DOPC = 2.00 in bilayers, in the middle of transition region observed earlier experimentally by small-angle neutron scattering (SANS). The bilayer thickness decrease induced by LDAO in DOPC observed by SANS did not result in the SERCA phosphohydrolase activity decrease and this indicates that some other factors compensated this bilayer effect of LDAO. The ATPase activity decrease at higher LDAO concentrations was caused by the bilayer deformation. This deformation resulted in the formation of non-bilayer aggregates in LDAO+DOPC system.  相似文献   

3.
The synthesis and high-pressure liquid chromatographic purification of the homogeneous nonionic surfactant p-(1,1,3,3-tetramethylbutyl)phenoxynonaoxyethylene glycol (OPE-9) in quantities suitable for membrane solubilization studies is reported. Micelles of OPE-9 and mixed micelles of OPE-9 with dimyristoyl and dipalmitoyl phosphatidylcholine as well as phosphatidylserine, phosphatidylethanolamine, lysophosphatidylcholine, sphingomyelin, and palmitic acid were characterized by column chromatography on 6% agarose. It was found that at 28°C OPE-9 micelles have a Stokes' radius of 32 Å, giving a molecular weight for a spherical micelle of about half that of micelles of the polydisperse nonionic surfactant Triton X-100 under the same conditions. The micelle size is temperature dependent: at 40°C the OPE-9 micelles have a Stokes' radius of 44 Å, giving a molecular weight for a spherical micelle of about twice that of the OPE-9 micelles at 28°C. The size of the mixed micelles varies linearly (as measured by Kav) with the mole fraction of phospholipid. The mixed micelle size was found to be relatively independent of the absolute concentration of surfactant over a four-fold range if the mole fraction of phospholipid is kept constant. The usefulness of the OPE-9/phospholipid mixed micelle system for lipolytic enzyme substrates and membrane-related studies is considered.  相似文献   

4.
Mixed micelle formation of the polydisperse nonionic surfactant Triton X-100 as well as its homogeneous analogue, p-(1,1,3,3-tetramethylbutyl)-phenoxynonaoxyethylene glycol (OPE-9), with bovine brain sphingomyelin or dipalmitoyl phosphatidylcholine has been characterized by column chromatography on 6% agarose. At 40 degrees C, mixtures of OPE-9 and either sphingomyelin or dipalmitoyl phosphatidylcholine give a narrow size distribution for mixed micelles. A this temperature the size distribution of Triton X-100-containing mixed micelles is complicated because of the polydispersity of the oxyethylene chains. At 20 degrees C narrow size distributions are observed for mixed micelles of sphingomyelin/Triton X-100 and sphingomyelin/OPE-9 up to at least 0.06 mol fraction of lipid. For dipalmitoyl phosphatidylcholine this is observed only with OPE-9. At intermediate mol fractions of lipid (around 0.25), two populations of mixed micelles exist for sphingomyelin/Trition X-100, sphingomyelin/OPE-9, and dipalmitoyl phosphatidylcholine/OPE-9. At high mol fractions of lipid only one population of mixed micelles again exists. At 20 degrees C, sphingoymelin forms a clear solution with Triton X-100 and OPE-9 to a lipid mol fraction of at least 0.46 and 0.67, respectively. Dipalmitoyl phosphatidylcholine forms a clear solution with OPE-9 to a lipid mol fraction of at least 0.57 at the same temperature. Triton X-100 and dipalmitoyl phosphatidylcholine do not form stable, clear solutions at 20 degrees C unless the lipid mol fraction is extremely low. These results show that surfactant polydispersity and temperature are important determinants in the solubilization of lipids by nonionic surfactants. It is also shown that pure surfactant micelles and lipid/surfactant mixed micelles do not co-exist in the same solution.  相似文献   

5.
The extraction of a relatively large molecular weight protein, bovine serum albumin (BSA), using nano-sized reverse micelles of nonionic surfactant polyoxyethylene p-t-octylphenol (Triton-X-100) is attempted for the first time. Suitability of reverse micelles of anionic surfactant sodium bis (2-ethyl hexyl) sulfosuccinate (AOT) and Triton-X-100/AOT mixture in organic solvent toluene for BSA extraction is also investigated. Although, the size of the Triton-X-100 reverse micelle in toluene is large enough to host BSA molecule in the hydraulic core, the overall extraction efficiency is found to be low, which may be due to lack of strong driving force. AOT/toluene system resulted in complete forward extraction at aqueous pH 5.5 and a surfactant concentration of 160 mM. The back extraction with aqueous phase (pH 5.5) resulted in 100% extraction of BSA from the organic phase. The addition of Triton-X-100 to AOT reduced the extraction efficiency of AOT reverse micelles, which may be attributed to reduced hydrophobic interaction. The circular dichroism (CD) spectrum of BSA extracted using AOT/toluene reverse micelles indicated the structural stability of the protein extracted.  相似文献   

6.
Using ultracentrifugation, the systems of reversed micelles of aerosol OT in octane containing solubilized protein (alpha-chymotrypsin, lysozyme, trypsin, egg albumin, alcohol dehydrogenase from horse liver and gamma-globulin) were studied. The changes in the sedimentation coefficients of reversed micelles during incorporation of the protein are correlated (within a wide range of experimental conditions, e. g. degree of surfactant hydration or protein concentration) exclusively with the molecular weight of the solubilized protein. The simplest solubilization model, according to which the protein molecule is incorporated into the inner cavity of the reversed micelle at the stoichiometric ratio of 1 : 1, which does not affect the external sizes of the reversed micelle, has been proposed. Using alpha-chymotrypsin as an example, the conditions, under which the sedimentation properties of the systems deviate from this model, have been found. These deviations occurred at sufficiently low degrees of the surfactant hydration, when the inner cavity of the reversed micelle is smaller than the effective size of the solubilized protein molecule. In the latter case the protein forms a new micelle of necessary (i. e. larger) size. Since the hydrated micelle can be regarded as an elementary (30-100 A) fragment of biomembranes, the results obtained should be taken into consideration when analyzing the structural organization and functioning of the latter.  相似文献   

7.
Although reconstitution of membrane proteins within protein detergent complexes is often used to enable their structural or biophysical characterization, it is unclear how one should rationally choose the appropriate micellar environment to preserve native protein folding. Here, we investigated model mixed micelles consisting of a nonionic glucosylated alkane surfactant from the maltoside and thiomaltoside families, bile salt surfactant, and the steryl derivative cholesteryl hemisuccinate. We correlated several key attributes of these micelles with the in vitro ligand-binding activity of hA2aR in these systems. Through small-angle neutron scattering and radioligand-binding analysis, we found several key aspects of mixed micellar systems that preserve the activity of hA2aR, including a critical amount of cholesteryl hemisuccinate per micelle, and an optimal hydrophobic thickness of the micelle that is analogous to the thickness of native mammalian bilayers. These features are closely linked to the headgroup chemistry of the surfactant and the hydrocarbon chain length, which influence both the morphology and composition of resulting micelles. This study should serve as a general guide for selecting the appropriate mixed surfactant systems to stabilize membrane proteins for biophysical analysis.  相似文献   

8.
The scientific study of protein surfactant interactions goes back more than a century, and has been put to practical uses in everything from the estimation of protein molecular weights to efficient washing powder enzymes and products for personal hygiene. After a burst of activity in the late 1960s and early 1970s that established the general principles of how charged surfactants bind to and denature proteins, the field has kept a relatively low profile until the last decade. Within this period there has been a maturation of techniques for more accurate and sophisticated analyses of protein-surfactant complexes such as calorimetry and small angle scattering techniques. In this review I provide an overview of different useful approaches to study these complexes and identify eight different issues which define central concepts in the field. (1) Are proteins denatured by monomeric surfactant molecules, micelles or both? (2) How does unfolding of proteins in surfactant compare with "proper" unfolding in chemical denaturants? Recent work has highlighted the role of shared micelles, rather than monomers, below the critical micelle concentration (cmc) in promoting both protein denaturation and formation of higher order structures. Kinetic studies have extended the experimentally accessible range of surfactant concentrations to far above the cmc, revealing numerous different modes of denaturation by ionic surfactants below and above the cmc which reflect micellar properties as much as protein unfolding pathways. Uncharged surfactants follow a completely different denaturation strategy involving synergy between monomers and micelles. The high affinity of charged surfactants for proteins means that unfolding pathways are generally different in surfactants versus chemical denaturants, although there are common traits. Other issues are as follows: (3) Are there non-denaturing roles for SDS? (4) How reversible is unfolding in SDS? (5) How do solvent conditions affect the way in which surfactants denature proteins? The last three issues compare SDS with "proper" membranes. (6) Do anionic surfactants such as SDS mimic biological membranes? (7) How do mixed micelles interact with globular proteins? (8) How can mixed micelles be used to measure the stability of membrane proteins? The growing efforts to understand the unique features of membrane proteins have encouraged the development of mixed micelles to study the equilibria and kinetics of this class of proteins, and traits which unite globular and membrane proteins have also emerged. These issues emphasise the amazing power of surfactants to both extend the protein conformational landscape and at the same time provide convenient and reversible short-cuts between the native and denatured state for otherwise obdurate membrane proteins.  相似文献   

9.
Micelles of lysomyristoylphosphatidylcholine (LMPC) and mixed micelles of LMPC with anionic detergent sodium dodecyl sulfate (SDS) have been characterized by spin-probe-partitioning electron paramagnetic resonance (SPPEPR) and time-resolved fluorescence quenching (TRFQ) experiments. SPPEPR is a novel new method to study structure and dynamics in lipid assemblies successfully applied here for the first time to micelles. Several improvements to the computer program used to analyze SPPEPR spectra have been incorporated that increase the precision in the extracted parameters considerably from which micelle properties such as effective water concentration and microviscosity may be estimated. In addition, with this increased precision, it is shown that it is feasible to study the rate of transfer of a small spin probe between micelles and the surrounding aqueous phase by SPPEPR. The rate of transfer of the spin probe di-tert-butyl nitroxide (DTBN) and the activation energy of the transfer process in LMPC and LMPC-SDS micelles have been determined with high precision. The rate of transfer increases with temperature and SDS molar fraction in mixed micelles, while it remains constant with LMPC concentration in pure LMPC micelles. The activation energy of DTBN transfer in pure lysophospholipid micelles does not change with LMPC concentration while it decreases with the increasing molar fraction of SDS in mixed LMPC-SDS micelles. Both this decrease in activation energy and the increase in the rate of transfer are rationalized in terms of an increasing micelle surface area per molecule (decreasing compactness) as SDS molecules are added. This decreasing compactness as a function of SDS content is confirmed by TRFQ measurements showing an aggregation number that decreases from 122 molecules for pure LMPC micelles to 80 molecules for pure SDS micelles. The same increase in surface area per molecule is predicted to increase the effective water concentration in the polar shell of the micelles. This increase in hydration with SDS molar fraction is confirmed by measuring the effective water concentration in the polar shell of the micelles from the hyperfine spacing of DTBN. This work demonstrates the potential to design mixed lysophospholipid surfactant micelles with variable physicochemical properties. Well-defined micellar substrates, in terms of their physicochemical properties, may improve the studies of protein structure and enzyme kinetics.  相似文献   

10.
Sarker M  Jackman D  Booth V 《Biochemistry》2011,50(22):4867-4876
Surfactant protein A (SP-A) is the most abundant protein component of lung surfactant, a complex mixture of proteins and lipids. SP-A performs host defense activities and modulates the biophysical properties of surfactant in concerted action with surfactant protein B (SP-B). Current models of lung surfactant mechanism generally assume SP-A functions in its octadecameric form. However, one of the findings of this study is that when SP-A is bound to detergent and lipid micelles that mimic lung surfactant phospholipids, it exists predominantly as smaller oligomers, in sharp contrast to the much larger forms observed when alone in water. These investigations were carried out in sodium dodecyl sulfate (SDS), dodecylphosphocholine (DPC), lysomyristoylphosphatidylcholine (LMPC), lysomyristoylphosphatidylglycerol (LMPG), and mixed LMPC + LMPG micelles, using solution and diffusion nuclear magnetic resonance (NMR) spectroscopy. We have also probed SP-A's interaction with Mini-B, a biologically active synthetic fragment of SP-B, in the presence of micelles. Despite variations in Mini-B's own interactions with micelles of different compositions, SP-A is found to interact with Mini-B in all micelle systems and perhaps to undergo a further structural rearrangement upon interacting with Mini-B. The degree of SP-A-Mini-B interaction appears to be dependent on the type of lipid headgroup and is likely mediated through the micelles, rather than direct binding.  相似文献   

11.
《Biophysical journal》2021,120(18):4115-4128
Empirically, α-helical membrane protein folding stability in surfactant micelles can be tuned by varying the mole fraction MFSDS of anionic (sodium dodecyl sulfate (SDS)) relative to nonionic (e.g., dodecyl maltoside (DDM)) surfactant, but we lack a satisfying physical explanation of this phenomenon. Cysteine labeling (CL) has thus far only been used to study the topology of membrane proteins, not their stability or folding behavior. Here, we use CL to investigate membrane protein folding in mixed DDM-SDS micelles. Labeling kinetics of the intramembrane protease GlpG are consistent with simple two-state unfolding-and-exchange rates for seven single-Cys GlpG variants over most of the explored MFSDS range, along with exchange from the native state at low MFSDS (which inconveniently precludes measurement of unfolding kinetics under native conditions). However, for two mutants, labeling rates decline with MFSDS at 0–0.2 MFSDS (i.e., native conditions). Thus, an increase in MFSDS seems to be a protective factor for these two positions, but not for the five others. We propose different scenarios to explain this and find the most plausible ones to involve preferential binding of SDS monomers to the site of CL (based on computational simulations) along with changes in size and shape of the mixed micelle with changing MFSDS (based on SAXS studies). These nonlinear impacts on protein stability highlights a multifaceted role for SDS in membrane protein denaturation, involving both direct interactions of monomeric SDS and changes in micelle size and shape along with the general effects on protein stability of changes in micelle composition.  相似文献   

12.
Mechanisms of protein solubilization in reverse micelles   总被引:4,自引:0,他引:4  
Solubilization properties of alpha-chymotrypsin and alcohol dehydrogenase (LADH) in reverse micelles are reported for three different solubilization techniques. The solubilization properties for these two proteins depend on the method used for protein addition. The addition of a dry protein powder to a reverse-micelle-containing organic phase does not appreciably solubilize the protein until the diameter of the reverse micelle is similar to that of the protein. However, when an aqueous protein solution is injected an organic phase, protein solubilization is not strongly dependent on micelle size. For chymotrypsin, multiple protein occupancy occurs at large micelle size, with as many as 11 chymotrypsin molecules solubilized in one reverse micelle. The solubilization of chymotrypsin using a phase-transter technique with a positively charged surfactant follows the expected traned based on protein-surfactant electrostatic interactions. When a negatively charged sufactants is used for phase transfer, at low pH the solubilization data do not fit this electrostatic interaction mechanism. In this case, proteinsurfactant aggregation may be occurring at the aqueousorganic interface.  相似文献   

13.
Characterization of the solubilization of lipid bilayers by surfactants   总被引:11,自引:0,他引:11  
This communication addresses the state of aggregation of lipid-detergent mixed dispersions. Analysis of recently published data suggest that for any given detergent-lipid mixture the most important factor in determining the type of aggregates (mixed vesicles or mixed micelles) and the size of the aggregate is the detergent to lipid molar ratio in these aggregates, herein denoted the effective ratio, Re. For mixed bilayers this effective ratio has been previously shown to be a function of the lipid and detergent concentrations and of an equilibrium partition coefficient, K, which describes the distribution of the detergent between the bilayers and the aqueous phase. We show that, similar to mixed bilayers, the size of mixed micelles is also a function of the effective ratio, but for these dispersions the distribution of detergent between the mixed micelles and the aqueous medium obeys a much higher partition coefficient. In practical terms, the detergent concentration in the mixed micelles is equal to the difference between the total detergent concentration and the critical micelle concentration (cmc). Thus, the effective ratio is equal to this difference divided by the lipid concentration. Transformation of mixed bilayers to mixed micelles, commonly denoted solubilization, occurs when the surfactant to lipid effective ratio reaches a critical value. Experimental evaluation of this critical ratio can be based on the linear dependence of detergent concentration, required for solubilization, on the lipid concentration. According to the 'equilibrium partition model', the dependence of the 'solubilizing detergent concentration' on the lipid concentration intersects with the lipid axis at -1/K, while the slope of this dependence is the critical effective ratio. On the other hand, assuming that when solubilization occurs the detergent concentration in the aqueous phase is approximately equal to the critical micelle concentration, implies that the above dependence intersects with the detergent axis at the critical micelle concentration, while its slope, again, is equal to the critical effective ratio. Analysis of existing data suggests that within experimental error both these distinctively different approaches are valid, indicating that the critical effective ratio at which solubilization occurs is approximately equal to the product of the critical micelle concentration and the distribution coefficient K. Since the nature of detergent affects K and the critical micelle concentration in opposite directions, the critical ('solubilizing') effective ratio depends upon the nature of detergent less than any of these two factors.  相似文献   

14.
The effect of water-insoluble compounds on enzyme catalytic properties was studied using a colloidal solution of water in organic solvent as reaction medium. In this microheterogeneous medium enzyme is entrapped into hydrated reversed micelles of a surfactant, the dimensions of the internal hole of the micelles being dependent on the ratio of water to surfactant. At sufficiently low values of this ratio the molecule of entrapped enzyme has limited mobility in the micelle. Because of this the interaction of the enzyme with water-insoluble compound which is added in assay solution and intercalated in the surface layer of the micelle may be manifested. The suggested method was used to study the inhibitory action of dihydroriboflavin esters on D-amino acid oxidase from pig kidney and soybean lipoxygenase. The reaction medium was hydrated reversed micelles of Aerosol OT in octane. The method of sedimentation in an analytical ultracentrifuge has shown the dihydroriboflavin esters to be completely included into reversed micelles.  相似文献   

15.
Phytophenols were solubilized in nonionic surfactant micelles to form antimicrobially active and thermodynamically stable microemulsions. Formulation of phytophenols in microemulsions has previously been shown to improve their antimicrobial activity in model microbiological and food systems. Carvacrol and eugenol were incorporated in micellar solutions of two nonionic surfactants (Surfynol® 485W and Surfynol® 465) by mixing at room temperature. Particle size of formed microemulsions was determined by dynamic light scattering, and structural information about the mixed micellar system was obtained by nuclear magnetic resonance spectroscopy (NMR). Uptake of carvacrol and eugenol in surfactant micelles as determined by ultrasonic velocity measurements was very rapid, e.g., below the maximum additive concentration, the phytophenols were completely solubilized in the micelles in less than 30 min. Depending on the surfactant–phytophenol combination, the self-assembled surfactant–phytophenol aggregates had mean particle diameters between 3 and 17 nm. Elucidation of the structure of aggregates by 1H NMR studies indicated that micelles had a “bracket-like” structure with phytophenols being located inside the palisade layer of the micelle in direct contact with adjacent surfactant monomers. Encapsulation of phytophenols in surfactant micelles enables the incorporation of large amounts of hydrophobic antimicrobials in aqueous phases. Formulation of antimicrobial microemulsions may thus offer a means to deliver high concentrations of phytophenols to the bacterial surfaces of foodborne pathogens to affect kill.  相似文献   

16.
Psachoulia E  Bond PJ  Sansom MS 《Biochemistry》2006,45(30):9053-9058
Mistic is an unusual membrane protein from Bacillus subtilis. It appears to fold and insert autonomously into a lipid bilayer and has been suggested as a tool that aids the targeting of eukaryotic membrane proteins to bacterial membranes. The NMR structure of Mistic in detergent (LDAO) micelles has revealed it to be a four alpha-helix bundle. From a structural perspective, Mistic does not resemble other membrane proteins. Its external surface is not very hydrophobic, and standard methods do not predict any of its helices to be in the transmembrane orientation. Molecular dynamics simulations (simulation times approximately 30 ns) in water and in detergent micelles have been used to explore the conformational stability of Mistic as a function of its environment. In water, the protein is stable, exhibiting no significant change in fold on a 30 ns time scale. In contrast, in three simulations in detergent micelles, the partial unfolding of Mistic occurred, whereby the H4 helix drifted away from the H1-H3 core. This was due to the penetration of detergent molecules between H4 and the remainder of the protein. This is unlike the behavior of several other membrane proteins, both alpha-helix bundles and beta-barrels, in comparable detergent micelle simulations. The unfolding of H4 from the H1-H3 core of Mistic could be partially reversed by a simulation in which the detergent molecules were removed, and the unfolded protein was simulated in water. These results suggest that Mistic may not be a stable integrated membrane protein but rather that it may undergo a conformational change upon interaction with a membrane or membrane-like environment.  相似文献   

17.
Bovine-heart mitochondrial complex III was purified in the presence of Triton X-100, and the size and shape of the resulting protein-surfactant-phospholipid mixed micelles were investigated by laser light-scattering. The protein appears to be present in the form of a dimer, irrespective of temperature (between 25 and 40 degrees C) and protein concentration (between 0.5 and 5 mg/ml). The molecular weight of the micelle increases with temperature from 600 000 (25 degrees C) to 692 000 (40 degrees C). The variation of the solvent second virial coefficient in this temperature range suggests that, with increasing temperature, some of the free surfactant molecules become integrated in the mixed micelles. The average quadratic radius of gyration of these is of 42 +/- 5 nm, corresponding in our case to an ellipsoidal shape.  相似文献   

18.
The interactions between the diblock copolymer S(15)E(63) and the surfactants sodium dodecyl sulfate (SDS), sodium decyl sulfate (SDeS), and sodium octyl sulfate (SOS) have been investigated by dynamic light scattering (DLS), transmission electron microscopy (TEM), and isothermal titration calorimetry (ITC). The surfactants with the same headgroup differentiate in their chain length. At 20 degrees C, the block copolymer is associated into micelles with a hydrodynamic radius of 11.6 nm, which is composed of a hydrophobic styrene oxide (S) core and a water-swollen oxypolyethylene (PEO or E) corona. The different copolymer/surfactant systems have been studied at a constant copolymer concentration of 2.5 g dm(-3) and in a vast range of surfactant concentrations, from 7.5 x 10(-6) up to 0.75 M. When SDS and SDeS are added to the block copolymer solution, different regions are observed in the DLS data: at low surfactant concentrations (c < 1.0 x 10(-4) M), single surfactant molecules associate with the copolymer micelle, probably the former being solubilized in the micelle core, leading to a certain disruption of the mixed micelle due to repulsive electrostatic interactions between surfactant headgroups followed by a stabilization of the mixed micelle. At higher concentrations (1.0 x 10(-4) < c < 0.1 M), two types of copolymer-surfactant complexes coexist: one large copolymer-rich/surfactant complex and one small complex consisting of one or a few copolymer chains and rich in surfactants. At higher SDS and SDeS concentrations, complete disintegration of mixed micelles takes place. In contrast, SOS-S(15)E(63) interactions are less important up to surfactant concentrations of 0.05 M due to its higher hydrophilicity, reducing the hydrophobic interactions between surfactant alkyl chains and copolymer micelles. At concentration larger than the critical aggregation concentration (cac) of the system, 0.05 M, disruption of copolymer micelles occurs. These regions have been confirmed by transmission electron microscopy. On the other hand, the titration calorimetric data for SDS and SDeS present an endothermic increase indicating the formation of mixed copolymer-rich-surfactant micelles. From that point, important differences in the ITC plot for both surfactants are present. However, the ITC curve obtained after titration of a SOS solution in the copolymer solution is quite similar to that of its titration in water.  相似文献   

19.
One of the several uses of sucrose detergents, as well as other micelle forming detergents, is the solubilization of different membrane proteins. Accurate knowledge of the micelle properties, including size and shape, are needed to optimize the surfactant conditions for protein purification and membrane characterization. We synthesized sucrose esters having different numbers of methylene subunits on the substituent to correlate the number of methylene groups with the size of the corresponding micelles. We used Fluorescence Correlation Spectroscopy (FCS) and two photon excitation to determine the translational D of the micelles and calculate their corresponding hydrodynamic radius, R(h). As a fluorescent probe we used LAURDAN (6-dodecanoyl-2-dimethylaminonaphthalene), a dye highly fluorescent when integrated in the micelle and non-fluorescent in aqueous media. We found a linear correlation between the size of the tail and the hydrodynamic radius of the micelle for the series of detergents measured.  相似文献   

20.
This communication addresses the state of aggregation of lipid-detergent mixed dispersions. Analysis of recently published data suggest that for any given detergent-lipid mixture the most important factor in determining the type of aggregates (mixed vesicles or mixed micelles) and the size of the aggregate is the detergent to lipid molar ratio in these aggregates, herein denoted the effective ratio, Re. For mixed bilayers this effective ratio has been previously shown to be a function of the lipid and detergent concentrations and of an equilibrium partition coefficient, K, which describes the distribution of the detergent between the bilayers and the aqueous phase. We show that, similar to mixed bilayers, the size of mixed micelles is also a function of the effective ratio, but for these dispersions the distribution of detergent between the mixed micelles and the aqueous medium obeys a much higher partition coefficient. In practical terms, the detergent concentration in the mixed micelles is equal to the difference between the total detergent concentration and the critical micelle concentration (cmc). Thus, the effective ratio is equal to this difference divided by the lipid concentration. Transformation of mixed bilayers to mixed micelles, commonly denoted solubilization, occurs when the surfactant to lipid effective ratio reaches a critical value. Experimental evaluation of this critical ratio can be based on the linear dependence of detergent concentration, required for solubilization, on the lipid concentration. According to the ‘equilibrium partition model’, the dependence of the ‘solubilizing detergent concentration’ on the lipid concentration intersects with the lipid axis at −1/K, while the slope of this dependence is the critical effective ratio. On the other hand, assuming that when solubilization occurs the detergent concentration in the aqueous phase is approximately equal to the critical micelle concentration, implies that the above dependence intersects with the detergent axis at the critical micelle concentration, while its slope, again, is equal to the critical effective ratio. Analysis of existing data suggests that within experimental error both these distinctively different approaches are valid, indicating that the critical effective ratio at which solubilization occurs is approximately equal to the product of the critical micelle concentration and the distribution coefficient K. Since the nature of detergent affects K and the critical micelle concentration in opposite directions, the critical (‘solubilizing’) effective ratio depends upon the nature of detergent less than any of these two factors.  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