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 0 毫秒
1.
BackgroundThe permeability of a lipid bilayer is a function of its phase state and depends non-linearly on thermodynamic variables such as temperature, pressure or pH. We investigated how shear forces influence the phase state of giant unilamellar vesicles and their membrane permeability.MethodsWe determined the permeability of giant unilamellar vesicles composed of different phospholipid species under shear flow in a tube at various temperatures around and far off the melting point by analyzing the release of fluorescently labelled dextran. Furthermore, we quantified phase state changes of these vesicles under shear forces using spectral decomposition of the membrane embedded fluorescent dye Laurdan.ResultsWe observed that the membrane permeability follows a step function with increasing permeability at the transition from the gel to the fluid phase and vice versa. Second, there was an all-or-nothing permeabilization near the main phase transition temperature and a gradual dye release far off the melting transition. Third, the Laurdan phase state analysis suggests that shear forces induce a reversible melting temperature shift in giant unilamellar vesicle membranes.Major conclusionsThe observed effects can be explained best in a scenario in which shear forces directly induce membrane pores that possess relatively long pore lifetimes in proximity to the phase transition.General significanceOur study elucidates the release mechanism of thermo-responsive drug carriers as we found that liposome permeabilization is not continuous but quantized. Furthermore, the shear force induced melting temperature shift must be taken into consideration when thermo-responsive liposomes are designed.  相似文献   

2.
Large unilamellar vesicles of two phosphatidylcholines, one saturated (DMPC) and the other unsaturated (DOPC), prepared by the reverse-phase evaporation method were studied using the quasielastic light scattering technique. The accurate sizing obtained by this technique showed an osmotic response for the two kinds of vesicles when the salinity of the external medium was diluted. The elastic moduli of lipid vesicles bilayers in the liquid phase were then estimated according to the elasticity theory of spherical shells taking into account salt leakage data known from the literature.  相似文献   

3.
The interaction of the bile salt cholate with unilamellar vesicles was studied. At low cholate content, equilibrium binding measurements with egg yolk lecithin membranes suggest that cholate binds to the outer vesicle leaflet. At increasing concentrations, further bile salt binding to the membrane is hampered. Before the onset of membrane solubilization, diphenylhexatriene fluorescence anisotropy decreases to a shallow minimum. It then increases to the initial value in the cholate concentration range of membrane solubilization. At still higher cholate concentrations, a drop in fluorescence anisotropy indicates the transformation of mixed disk micelles into spherical micelles. Perturbation of the vesicle membranes at molar ratios of bound cholate/lecithin exceeding 0.15 leads to a transient release of oligosaccharides from intravesicular space. The cholate concentrations required to induce the release depend on the size of the entrapped sugars. Cholesterol stabilizes the membrane, whereas, in spite of enhanced membrane order, sphingomyelin destabilizes the membrane against cholate. Freeze-fracture electron microscopy and phosphorus-31 nuclear magnetic resonance (31P NMR) also reflect a change in membrane structure at maximal cholate binding to the vesicles. In 31P NMR spectra, superimposed on the anisotropic line typically found in phospholipid bilayers, an isotropic peak was found. This signal is most probably due to the formation of smaller vesicles after addition of cholate. The results were discussed with respect to bile salt/membrane interactions in the liver cell. It is concluded that vesicular bile salt transport in the cytoplasm is unlikely and that cholate binding is restricted to the outer leaflet of the canalicular part of the plasma membrane.  相似文献   

4.
5.
Interaction of wheat alpha-thionin with large unilamellar vesicles.   总被引:1,自引:0,他引:1       下载免费PDF全文
The interaction of the wheat antibacterial peptide alpha-thionin with large unilamellar vesicles has been investigated by means of fluorescence spectroscopy. Binding of the peptide to the vesicles is followed by the release of vesicle contents, vesicle aggregation, and lipid mixing. Vesicle fusion, i.e., mixing of the aqueous contents, was not observed. Peptide binding is governed by electrostatic interactions and shows no cooperativity. The amphipatic nature of wheat alpha-thionin seems to destabilize the membrane bilayer and trigger the aggregation of the vesicles and lipid mixing. The presence of distearoylphosphatidylethanolamine-poly(ethylene glycol 2000) (PEG-PE) within the membrane provides a steric barrier that inhibits vesicle aggregation and lipid mixing but does not prevent leakage. Vesicle leakage through discrete membrane channels is unlikely, because the release of encapsulated large fluorescent dextrans is very similar to that of 8-aminonaphthalene-1,3,6,trisulfonic acid (ANTS). A minimum number of 700 peptide molecules must bind to each vesicle to produce complete leakage, which suggests a mechanism in which the overall destabilization of the membrane is due to the formation of transient pores rather than discrete channels.  相似文献   

6.
7.
Using the sectioning effect of the two-photon fluorescence microscope, we studied the behavior of phospholipid giant unilamellar vesicles (GUVs) composed of pure diacylphosphatidylcholine phospholipids during the gel-to-liquid crystalline phase transition. We used the well-characterized excitation generalized polarization function (GP(ex)) of 6-dodecanoyl-2-dimethylamine-naphthalene (LAURDAN), which is sensitive to the changes in water content in the lipid vesicles, to monitor the phase transition in the GUVs. Even though the vesicles do not show temperature hysteresis at the main phase transition, we observed different behaviors of the vesicle shape, depending on how the GUV sample reaches the main phase transition. During the cooling cycles, we observed an increase in the vesicle diameter at the phase transition ( approximately 0.5-1%), followed by a decrease in the diameter when the vesicle reached the gel phase. During the heating cycles and close to the phase transition temperature, a surprising behavior is observed, showing a sequence of different vesicle shapes as follows: spherical-polygonal-ellipsoidal. We attribute these changes to the effect of lipid domain coexistence on the macroscopic structure of the GUVs. The "shape hysteresis" in the GUVs is reversible and largely independent of the temperature scan rate. In the presence of 30 mol% of cholesterol the events observed at the phase transition in the GUVs formed by pure phospholipids were absent.  相似文献   

8.
Transbilayer movement of phospholipids in biological membranes is mediated by energy-dependent and energy-independent flippases. Available methods for detection of flippase mediated transversal flip-flop are essentially based on spin-labeled or fluorescent lipid analogues. Here we demonstrate that shape change of giant unilamellar vesicles (GUVs) can be used as a new tool to study the occurrence and time scale of flippase-mediated transbilayer movement of unlabeled phospholipids. Insertion of lipids into the external leaflet created an area difference between the two leaflets that caused the formation of a bud-like structure. Under conditions of negligible flip-flop, the bud was stable. Upon reconstitution of the energy-independent flippase activity of the yeast endoplasmic reticulum into GUVs, the initial bud formation was reversible, and the shapes were recovered. This can be ascribed to a rapid flip-flop leading to relaxation of the monolayer area difference. Theoretical analysis of kinetics of shape changes provides self-consistent determination of the flip-flop rate and further kinetic parameters. Based on that analysis, the half-time of phospholipid flip-flop in the presence of endoplasmic reticulum proteins was found to be on the order of few minutes. In contrast, GUVs reconstituted with influenza virus protein formed stable buds. The results argue for the presence of specific membrane proteins mediating rapid flip-flop.  相似文献   

9.
Static and dynamic light scattering measurements have been used to characterize the size, size distribution, and shape of extruded vesicles under isotonic conditions. Dynamic light scattering was then used to characterize osmotically induced shape changes by monitoring changes in the hydrodynamic radius (R(h)) of large unilamellar vesicles (LUVs). These changes are compared to those predicted for several shapes that appear in trajectories through the phase diagram of the area difference elasticity (ADE) model (. Phys. Rev. E. 52:6623-6634). Measurements were performed on dioleoylphosphatidylcholine (DOPC) vesicles using two membrane-impermeant osmolytes (NaCl and sucrose) and a membrane-permeant osmolyte (urea). For all conditions, we were able to produce low-polydispersity, nearly spherical vesicles, which are essential for resolving well-defined volume changes and consequent shape changes. Hyper-osmotic dilutions of DOPC vesicles in urea produced no change in R(h), whereas similar dilutions in NaCl or sucrose caused reductions in vesicle volume resulting in observable changes to R(h). Under conditions similar to those of this study, the ADE model predicts an evolution from spherical to prolate then oblate shapes on increasing volume reduction of LUVs. However, we found that DOPC vesicles became oblate at all applied volume reductions.  相似文献   

10.
The exchange of cholesterol between two populations of small unilamellar vesicles has been investigated using a new system. Uniformly sized egg lecithin-cholesterol vesicles containing [3H]cholesterol and the glycolipid N-palmitoyl-DL-dihydrolactocerebroside were used as donors, whereas similar vesicles containing unlabelled cholesterol and no glycolipid were used as cholesterol acceptors. The two populations of vesicles were separated with the castor bean lectin Ricinus communis. It was found that greater than 90% of the cholesterol in the donor vesicle could be exchanged with a single time constant, the half-time for the completion of this exchange process being 1.5 h at 37 degrees C. Therefore, the rate of transmembrane movement or flip-flop of cholesterol in these vesicles must be at least as fast as the intermembrane exchange process. Similar results were obtained using hemoglobin-free human erythrocyte ghosts as the acceptor membrane. If the molecular-sieve chromatography step used to fractionate the vesicles was omitted, a non-exchangeable pool of cholesterol was detected which was shown not to be due to the presence of multilamellar vesicles.  相似文献   

11.
Osmotic properties of large unilamellar vesicles prepared by extrusion.   总被引:3,自引:5,他引:3  
We have examined the morphology and osmotic properties of large unilamellar vesicles (LUVs) prepared by extrusion. Contrary to expectations, we observe by cryo-electron microscopy that such vesicles, under isoosmotic conditions, are non-spherical. This morphology appears to be a consequence of vesicle passage through the filter pores during preparation. As a result when such LUVs are placed in a hypoosmotic medium they are able to compensate, at least partially, for the resulting influx of water by "rounding up" and thereby increasing their volume with no change in surface area. The increase in vesicle trapped volume associated with these morphological changes was determined using the slowly membrane-permeable solute [3H]-glucose. This allowed calculation of the actual osmotic gradient experienced by the vesicle membrane for a given applied differential. When LUVs were exposed to osmotic differentials of sufficient magnitude lysis occurred with the extent of solute release being dependent on the size of the osmotic gradient. Surprisingly, lysis was not an all-or-nothing event, but instead a residual osmotic differential remained after lysis. This differential value was comparable in magnitude to the minimum osmotic differential required to trigger lysis. Further, by comparing the release of solutes of differing molecular weights (glucose and dextran) a lower limit of about 12 nm diameter can be set for the bilayer defect created during lysis. Finally, the maximum residual osmotic differentials were compared for LUVs varying in mean diameter from 90 to 340 nm. This comparison confirmed that these systems obey Laplace's Law relating vesicle diameter and lysis pressure. This analysis also yielded a value for the membrane tension at lysis of 40 dyn cm-1 at 23 degrees C, which is in reasonable agreement with previously published values for giant unilamellar vesicles.  相似文献   

12.
Small, unilamellar phospholipid vesicles have been prepared using a new, high-powdered cup sonifier that avoids contact of the sample with a titanium probe. These vesicles have been characterized by gel filtration chromatography both before and after fractionation by high-speed centrifugation. Plots of the turbidity of centrifuged vesicles between 300 and 650 nm against the reciprocal fourth power of the scattering wavelength were linear with zero intercepts (extrapolated to infinite wavelength). In the presence of minute quantities of large, multilamellar vesicles, these plots remained linear but had intercepts quantitatively proportional to the amount of contaminating large vesicles. Since this measurement requires only a standard spectrophotometer and very small quantities of lipid, this method is suggested as a useful assay for determining contamination of small vesicle preparations by large vesicles. Two applications of this method as well as a practical limitation are discussed.  相似文献   

13.
Multilamellar and unilamellar vesicles can be generated by a variety of techniques which lead to systems with differing lamellarity, size, trapped volume and solute distribution. The straight-forward hydration of lipid to produce multilamellar vesicles (MLVs) results in systems which exhibit low trapped volumes and where solutes contained in the aqueous buffer are partially excluded from the MLV interior. Large trapped volumes and equilibrium solute distributions can be achieved by freeze-thawing or by ‘reverse phase’ procedures where the lipid is hydrated after being solubilized in organic solvent. Unilamellar vesicles can be produced directly from MLVs by extrusion or sonication or, alternatively, can be obtained by reverse phase or detergent removal procedures. The advantages and limitations of these techniques are discussed.  相似文献   

14.
Melittin, the soluble lipophilic peptide of bee venom, causes fusion of phospholipid vesicles when vesicle suspensions are heated or cooled through their thermal phase transition. Fusion was detected using a new photochemical method (Morgan, C.G., Hudson, B. and Wolber, P. (1980) Proc. Natl. Acad. Sci. U.S.A. 77, 26–30) which monitors lipid mixing. Electron microscopy and gel filtration confirmed that most of the lipid formed large vesicular structures. Fluorescence experiments with a water-soluble, membrane-impermeable complex of terbium (Wilschut, J. and Papahadjopoulos, D. (1979) Nature 281, 690–692) demonstrate that these ionic contents are released during fusion. The large structures formed by melittin-induced fusion are impermeable to these ions and are resistant to further fusion. This is in contrast to the behavior observed for the cationic detergent cetyltrimethylammonium bromide (CETAB). The large size of the vesicles formed, the extreme speed of the fusion event and the appearance of electron microscope images of the vesicles prior to fusion suggest that the mechanism of the fusion process includes a preaggregation step.  相似文献   

15.
The effects of poly(ethylene glycol) (PEG) on the phase transition of phospholipid multilamellar vesicles (MLVs) were investigated by using differential scanning calorimetry (DSC). Main transition temperature (Tm) and the pre-transition temperature (Tp) of neutral phospholipid-, DMPC-1, DPPC- and DSPC-MLVs increased with an increase in PEG concentration. The subtransition temperature of DPPC-MLV also increased with an increase in PEG concentration. These results could be qualitatively explained by enhancement of the lateral packing on the basis of the osmoelastic coupling theory. The pretransition temperature increased faster than the main transition temperature did with an increase in PEG concentration. The increment of Tm depended on the hydrocarbon chain length, the shorter the hydrocarbon chain length was, the larger the increment was. The transition width in the DSC peak was broadened with an increase in PEG concentration. These three above-mentioned effects are the main differences between the effects of the osmotic stress on the phase transition of MLVs and those of hydrostatic pressure. On the other hand, ethylene glycol (EG), which is the monomer of PEG, had a biphasic effect on transition temperature of DPPC-, DSPC-, and DMPC-MLV, reducing Tm and Tp at low concentrations, but increasing Tm and extinguishing pretransition at high concentrations. This is explained by the induction of an interdigitated gel phase at high concentrations of EG, which indicates that EG can easily penetrate into the head group region of the lipid, in contrast with PEG 6K, because EG is small. Temperature-EG concentration phase diagrams for the various PC-MLVs were determined.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The ability of astaxanthin and canthaxanthin as chain-breaking antioxidants was studied in Cu(2+)-initiated peroxidation of phosphatidylcholine large unilamellar vesicles (LUVs). Both carotenoids increased the lag period that precedes the maximum rate of lipid peroxidation, though astaxanthin showed stronger activity. For these experiments, different amounts of xanthophylls were exogenously added to previously made LUVs, non-incorporated pigment being afterwards removed. Differential scanning calorimetry assays with L-beta,gamma-dimyristoyl-alpha-phosphatidylcholine LUVs demonstrated that xanthophylls incorporated as described interact with the lipid matrix becoming interspersed among the phospholipid molecules.  相似文献   

17.
There is an active interest in peptides that readily cross cell membranes without the assistance of cell membrane receptors(1). Many of these are referred to as cell-penetrating peptides, which are frequently noted for their potential as drug delivery vectors(1-3). Moreover, there is increasing interest in antimicrobial peptides that operate via non-membrane lytic mechanisms(4,5), particularly those that cross bacterial membranes without causing cell lysis and kill cells by interfering with intracellular processes(6,7). In fact, authors have increasingly pointed out the relationship between cell-penetrating and antimicrobial peptides(1,8). A firm understanding of the process of membrane translocation and the relationship between peptide structure and its ability to translocate requires effective, reproducible assays for translocation. Several groups have proposed methods to measure translocation into large unilamellar lipid vesicles (LUVs)(9-13). LUVs serve as useful models for bacterial and eukaryotic cell membranes and are frequently used in peptide fluorescent studies(14,15). Here, we describe our application of the method first developed by Matsuzaki and co-workers to consider antimicrobial peptides, such as magainin and buforin II(16,17). In addition to providing our protocol for this method, we also present a straightforward approach to data analysis that quantifies translocation ability using this assay. The advantages of this translocation assay compared to others are that it has the potential to provide information about the rate of membrane translocation and does not require the addition of a fluorescent label, which can alter peptide properties(18), to tryptophan-containing peptides. Briefly, translocation ability into lipid vesicles is measured as a function of the Foster Resonance Energy Transfer (FRET) between native tryptophan residues and dansyl phosphatidylethanolamine when proteins are associated with the external LUV membrane (Figure 1). Cell-penetrating peptides are cleaved as they encounter uninhibited trypsin encapsulated with the LUVs, leading to disassociation from the LUV membrane and a drop in FRET signal. The drop in FRET signal observed for a translocating peptide is significantly greater than that observed for the same peptide when the LUVs contain both trypsin and trypsin inhibitor, or when a peptide that does not spontaneously cross lipid membranes is exposed to trypsin-containing LUVs. This change in fluorescence provides a direct quantification of peptide translocation over time.  相似文献   

18.
Preparation and characterization of unilamellar myelin vesicles   总被引:1,自引:0,他引:1  
Myelin vesicles have been obtained from isolated rat brain myelin and shown by electron microscopy to consist of single bilayer membranes. The yield of the preparation is approximately 25% of the myelin proteins. The vesicles show a typical myelin protein pattern on sodium dodecyl sulfate-polyacrylamide gel electrophoresis and contain activity for the myelin marker enzyme, 2',3'-cyclic nucleotide-3'-phosphohydrolase (CNPase). The preparation consists of both inside-out and right-side-out vesicles, and the proportion in each orientation varies from one preparation to another. The occurrence of two populations is demonstrated by the observation that hypotonically lysed vesicles compete to a greater extent than intact vesicles in a competitive enzyme-linked immunosorbent assay with myelin basic protein antiserum. In addition, only a portion of the CNPase activity of the vesicles is trypsin-sensitive and detectable in the absence of detergent; the remaining, trypsin-insensitive activity is present in detergent-disrupted membranes. Thus, there are vesicle populations in which myelin basic protein and CNPase are accessible and others in which they are inaccessible. A population of uniformly oriented right-side-out vesicles has been obtained by ConA-Agarose affinity column chromatography and elution of the bound fraction with methyl-alpha-D-mannopyranoside. In the absence of detergent, less than 10% of the total CNPase activity of these vesicles can be demonstrated, suggesting that the active site of CNPase is opposite to that of the ConA binding site and, therefore, appears to be on the cytoplasmic face of the myelin membrane.  相似文献   

19.
We have investigated the stability of giant unilamellar vesicles (GUVs) and large unilamellar vesicles (LUVs) of lipid membranes in the liquid-ordered phase (lo phase) against a detergent, Triton X-100. We found that in the presence of high concentrations of Triton X-100, the structure of GUVs and LUVs of dipalmitoyl-PC (DPPC)/cholesterol (chol) and sphingomyelin (SM)/chol membranes in the lo phase was stable and no leakage of fluorescent probes from the vesicles occurred. We also found that ether-linked dihexadecylphosphatidylcholine (DHPC) membranes containing more than 20 mol% cholesterol were in the lo phase, and that DHPC/chol-GUV and DHPC/chol-LUV in the lo phase were stable and no leakage of internal contents occurred in the presence of Triton X-100. In contrast, octylglucoside solution could easily break these GUVs and LUVs of the lo phase membranes and induced internal contents leakage. These data indicate that GUVs and LUVs of the lo phase membranes are very valuable for practical use.  相似文献   

20.
The ability of lipid asymmetry to regulate Ca(2+)-stimulated fusion between large unilamellar vesicles has been investigated. It is shown that for 100-nm-diameter LUVs composed of dioleoylphosphatidylcholine, dioleoylphosphatidylethanolamine, phosphatidylinositol, and dioleoylphosphatidic acid (DOPC/DOPE/PI/DOPA; 25:60:5:10) rapid and essentially complete fusion is observed by fluorescent resonance energy transfer techniques when Ca2+ (8 mM) is added. Alternatively, for LUVs with the same lipid composition but when DOPA was sequestered to the inner monolayer by incubation in the presence of a pH gradient (interior basic), little or no fusion is observed on addition of Ca2+. It is shown that the extent of Ca(2+)-induced fusion correlates with the amount of exterior DOPA. Further, it is shown that LUVs containing only 2.5 mol % DOPA, but where all the DOPA is in the outer monolayer, can be induced to fuse to the same extent and with the same rate as LUVs containing 5 mol % DOPA. These results strongly support a regulatory role for lipid asymmetry in membrane fusion and indicate that the fusogenic tendencies of lipid bilayers are largely determined by the properties of the monolayers proximate to the fusion interface.  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