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 46 毫秒
1.
Transmembrane pH gradients have previously been shown to induce an asymmetric transmembrane distribution of simple lipids that exhibit weak acid or basic characteristics (Hope, M.J. and Cullis, P.R. (1987) J. Biol. Chem. 262, 4360-4366). In the present study we have examined the influence of proton gradients on the inter-vesicular exchange of stearylamine and oleic acid. We show that vesicles containing stearylamine immediately aggregate with vesicles containing phosphatidylserine and that disaggregation occurs subsequently as stearylamine equilibrates between the two vesicle populations. Despite visible flocculation during the aggregation phase, vesicle integrity is maintained. Stearylamine is the only lipid to exchange, fusion does not occur and vesicles are able to maintain a proton gradient. When stearylamine is sequestered to the inner monolayer in response to a transmembrane pH gradient (inside acidic) aggregation is not observed and diffusion of stearylamine to acceptor vesicles is greatly reduced. The ability of delta pH-dependent lipid asymmetry to modulate lipid exchange is also demonstrated for fatty acids. Oleic acid can be induced to transfer from one population of vesicles to another by maintaining a basic interior pH in the acceptor vesicles. Moreover, it is shown that the same acceptor vesicles are capable of depleting serum albumin of bound fatty acid. These results are discussed with respect to the mechanism and modulation of lipid flow between membranes both in vitro and in vivo.  相似文献   

2.
The interaction of two types of vesicle systems was investigated: micrometer-sized, giant unilamellar vesicles (GUVs) formed from 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) and submicrometer-sized, large unilamellar vesicles (LUVs) formed from oleic acid and oleate, both in a buffered aqueous solution (pH 8.8). Individual POPC GUVs were transferred with a micropipette into a suspension of oleic acid/oleate LUVs, and the shape changes of the GUVs were monitored using optical microscopy. The behavior of POPC GUVs upon transfer into a 0.8 mM suspension of oleic acid, in which oleic acid/oleate forms vesicular bilayer structures, was qualitatively different from the behavior upon transfer into a 0.3 mM suspension of oleic acid/oleate, in which oleic acid/oleate is predominantly present in the form of monomers and possibly non-vesicular aggregates. In both cases, changes in vesicle morphology were observed within tens of seconds after the transfer. After an initial increase of the vesicle cross-section, the vesicle started to evaginate, spawning dozens of satellite vesicles connected to the mother vesicle with narrow necks or tethers. In 60% of the cases of transfer into a 0.8 mM oleic acid suspension, the evagination process reversed and proceeded to the point where the membrane formed invaginations. In some of these cases, several consecutive transitions between invaginated and evaginated shapes were observed. In the remaining 40% of the cases of transfer into the 0.8 mM oleic acid suspension and in all cases of vesicle transfer into the 0.3 mM oleic acid suspension, no invaginations nor subsequent evaginations were observed. An interpretation of the observed vesicle shape transformation on the basis of the bilayer-couple model is proposed, which takes into account uptake of oleic acid/oleate molecules by the POPC vesicles, oleic acid flip-flop processes and transient pore formation.  相似文献   

3.
Proton-induced fusion of oleic acid-phosphatidylethanolamine liposomes   总被引:2,自引:0,他引:2  
Liposomes composed of oleic acid and phosphatidylethanolamine (3:7 mole ratio) aggregate, become destabilized, and fuse below pH 6.5 in 150 mM NaCl. Fusion is monitored by (i) the intermixing of internal aqueous contents of liposomes, utilizing the quenching of aminonaphthalene-3,6,8-trisulfonic acid (ANTS) by N,N'-p-xylylenebis(pyridinium bromide) (DPX) encapsulated in two separate populations of vesicles, (ii) a resonance energy transfer assay for the dilution of fluorescent phospholipids from labeled to unlabeled liposomes, (iii) irreversible changes in turbidity, and (iv) quick-freezing freeze-fracture electron microscopy. Destabilization is followed by the fluorescence increase caused by the leakage of coencapsulated ANTS/DPX or of calcein. Ca2+ and Mg2+ also induce fusion of these vesicles at 3 and 4 mM, respectively. The threshold for fusion is at a higher pH in the presence of low (subfusogenic) concentrations of these divalent cations. Vesicles composed of phosphatidylserine/phosphatidylethanolamine or of oleic acid/phosphatidylcholine (3:7 mole ratio) do not aggregate, destabilize, or fuse in the pH range 7-4, indicating that phosphatidylserine and phosphatidylcholine cannot be substituted for oleic acid and phosphatidylethanolamine, respectively, for proton-induced membrane fusion. Freeze-fracture replicas of oleic acid/phosphatidylethanolamine liposomes frozen within 1 s of stimulation with pH 5.3 display larger vesicles and vesicles undergoing fusion, with membrane ridges and areas of bilayer continuity between them. The construction of pH-sensitive liposomes is useful as a model for studying the molecular requirements for proton-induced membrane fusion in biological systems and for the cytoplasmic delivery of macromolecules.  相似文献   

4.
Studies on membrane fusion. III. The role of calcium-induced phase changes.   总被引:12,自引:0,他引:12  
The interaction of phosphatidylserine vesicles with Ca2+ and Mg2+ has been examined by several techniques to study the mechanism of membrane fusion. Data are presented on the effects of Ca2+ and Mg2+ on vesicle permeability, thermotropic phase transitions and morphology determined by differential scanning calorimetry, X-ray diffraction, and freeze-fracture electron microscopy. These data are discussed in relation to information concerning Ca2+ binding, charge neutralization, molecular packing, vesicle aggregation, phase transitions, phase separations and vesicle fusion. The results indicate that at Ca2+ concentrations of 1.0-2.0 mM, a highly cooperative phenomenon occurs which results in increased vesicle permeability, aggregation and fusion of the vesicles. Under these conditions the hydrocarbon chains of the lipid bilayers undergo a phase change from a fluid to a crystalline state. The aggregation of vesicles that is observed during fusion is not sufficient range of 2.0-5.0 mM induces aggregation of phosphatidylserine vesicles but no significant fusion nor a phase change. From the effect of variations in pH, temperature, Ca2+ and Mg2+ concentration on the fusion of vesicles, it is concluded that the key event leading to vesicle membrane fusion is the isothermic phase change induced by the bivalent metals. It is proposed that this phase change induces a transient destabilization of the bilayer membranes that become susceptible to fusion at domain boundaries.  相似文献   

5.
We have investigated the initial kinetics of Ca2+-induced aggregation and fusion of phosphatidylserine large unilamellar vesicles at 3, 5 and 10 mM Ca2+ and 15, 25 and 35 degrees C, utilizing the Tb/dipicolinate (Tb/DPA) assay for mixing of aqueous vesicle contents and a resonance energy transfer (RET) assay for mixing of bilayer lipids. Separate rate constants for vesicle aggregation as well as deaggregation and for the fusion reaction itself were determined by analysis of the data in terms of a mass action kinetic model. At 15 degrees C the aggregation rate constants for either assay are the same, indicating that at this temperature all vesicle aggregation events that result in lipid mixing lead to mixing of aqueous contents as well. By contrast, at 35 degrees C the RET aggregation rate constants are higher than the Tb/DPA aggregation rate constants, indicating a significant frequency of reversible vesicle aggregation events that do result in mixing of bilayer lipids, but not in mixing of aqueous vesicle contents. In any conditions, the RET fusion rate constants are considerably higher than the Tb/DPA fusion rate constants, demonstrating the higher tendency of the vesicles, once aggregated, to mix lipids than to mix aqueous contents. This possibly reflects the formation of an intermediate fusion structure. With increasing Ca2+ concentrations the RET and the Tb/DPA fusion rate constants increase in parallel with the respective aggregation rate constants. This suggests that fusion susceptibility is conferred on the vesicles during the process of vesicle aggregation and not solely as a result of the interaction of Ca2+ with isolated vesicles. Aggregation of the vesicles in the presence of Mg2+ produces neither mixing of aqueous vesicle contents nor mixing of bilayer lipids.  相似文献   

6.
Unilamellar vesicles composed of phosphatidylcholine (PC) and either phosphatidic acid (PA) or phosphatidylglycerol (PG) partition to the upper poly(ethylene glycol) (PEG)-rich phase of a charge-sensitive 5%:5% (w/w) PEG 8000/Dextran T-500 phase system containing 10 mM sodium phosphate at pH 7, consistent with the vesicles bearing a net negative charge. When prepared in the presence of a pH gradient (interior acidic), PC/PA vesicles exhibit an increased partition to the top PEG-rich phase, consistent with a redistribution of the PA from the inner to the outer monolayer of the vesicle bilayer. Conversely, when prepared in the presence of a pH gradient (interior basic), PC/PG vesicles exhibit a decreased top-phase partition, consistent with a redistribution of the PG from the outer to the inner monolayer of the vesicle bilayer. Unilamellar vesicles composed of PC and stearylamine partition to the lower dextran-rich phase of a 5%:5% (w/w) PEG 8000/Dextran T-500 phase system containing 10 mM sodium phosphate at pH 8.5, consistent with the vesicles bearing a net positive charge. When prepared in the presence of a pH gradient (interior acidic), conditions under which the stearylamine is trapped on the inner monolayer of the bilayer, the vesicles now partition predominantly to the interface in a manner similar to vesicles composed of PC alone. These results demonstrate that partitioning in aqueous two-phase polymer systems is a sensitive method for monitoring the asymmetry of charged lipids in model membrane systems and also suggests that partitioning in charge-sensitive systems depends only on the physical nature of the exterior surface of the membrane.  相似文献   

7.
In the process of lipoprotein lipolysis, masses of fatty acid are generated at the surface of the lipoprotein. The newly generated fatty acid may at least partly redistribute from the site of lipolysis to phospholipid-rich membranes and to albumin. We have studied the distribution of [1-13C]oleic acid in model systems consisting of chylomicron-like triacylglycerol-rich emulsions, unilamellar phosphatidylcholine vesicles, and bovine serum albumin. By using high resolution 13C NMR spectroscopy it was possible to distinguish fatty acid in each compartment (emulsion, vesicle, albumin) and quantitate the fatty acid distribution under various conditions of lipid compartment concentration and aqueous pH. When emulsions and vesicles were present in equivalent mass amounts, fatty acid exhibited a profound preference for the lipid bilayers. The release of oleic acid to phospholipid bilayers was presumably also a function of its high molar stoichiometry (5:1) with the albumin present. More equitable distributions of fatty acid between vesicles and emulsions were seen when higher concentrations of emulsion were used. The distribution of fatty acid between compartments was in good agreement with predictions made using the apparent ionization constant, expressed as pKapp, of 7.5 and the surface to core (phospholipid to triacylglycerol) distribution coefficient of 7.0, measured for unionized oleic acid in chylomicron particles (Spooner, P. J. R., Bennett Clark, S., Gantz, D. L., Hamilton, J. A., and Small, D. M. (1988) J. Biol. Chem. 263, 1444-1455). These results indicate that the affinities of fatty acid for phospholipid bilayer and chylomicron-like emulsion surfaces are equivalent. Redistribution of lipolytically generated fatty acid from chylomicron surface to cell membrane may simply be driven by the predominant quantity of the cell membrane surfaces.  相似文献   

8.
Small unilamellar phosphatidylserine/phosphatidylcholine liposomes incubated on one side of planar phosphatidylserine bilayer membranes induced fluctuations and a sharp increase in the membrane conductance when the Ca2+ concentration was increased to a threshold of 3--5 mM in 100 mM NaCl, pH 7.4. Under the same ionic conditions, these liposomes fused with large (0.2 micrometer diameter) single-bilayer phosphatidylserine vesicles, as shown by a fluorescence assay for the mixing of internal aqueous contents of the two vesicle populations. The conductance behavior of the planar membranes was interpreted to be a consequence of the structural rearrangement of phospholipids during individual fusion events and the incorporation of domains of phosphatidylcholine into the Ca2+-complexed phosphatidylserine membrane. The small vesicles did not aggregate or fuse with one another at these Ca2+ concentrations, but fused preferentially with the phosphatidylserine membrane, analogous to simple exocytosis in biological membranes. Phosphatidylserine vesicles containing gramicidin A as a probe interacted with the planar membranes upon raising the Ca2+ concentration from 0.9 to 1.2 mM, as detected by an abrupt increase in the membrane conductance. In parallel experiments, these vesicles were shown to fuse with the large phosphatidylserine liposomes at the same Ca2+ concentration.  相似文献   

9.
Fusion of multilamellar phospholipid vesicles with planar phospholipid bilayer membranes was monitored by the rate of appearance in the planar membrane of an intrinsic membrane protein present in the vesicle membranes. An essential requirement for fusion is an osmotic gradient across the planar membrane, with the cis side (the side containing the vesicles) hyperosmotic to the opposite (trans) side; for substantial fusion rates, divalent cation must also be present on the cis side. Thus, the low fusion rates obtained with 100 mM excess glucose in the cis compartment are enhanced orders of magnitude by the addition of 5-10 mM CaCl2 to the cis compartment. Conversely, the rapid fusion rates induced by 40 mM CaCl2 in the cis compartment are completely suppressed when the osmotic gradient (created by the 40 mM CaCl2) is abolished by addition of an equivalent amount of either CaCl2, NaCl, urea, or glucose to the trans compartment. We propose that fusion occurs by the osmotic swelling of vesicles in contact with the planar membrane, with subsequent rupture of the vesicular and planar membranes in the region of contact. Divalent cations catalyze this process by increasing the frequency and duration of vesicle-planar membrane contact. We argue that essentially this same osmotic mechanism drives biological fusion processes, such as exocytosis. Our fusion procedure provides a general method for incorporating and reconstituting transport proteins into planar phospholipid bilayer membranes.  相似文献   

10.
P Meers  K Hong  D Papahadjopoulos 《Biochemistry》1988,27(18):6784-6794
The effect of free fatty acids on the cation-induced fusion of large unilamellar vesicles (liposomes) was investigated by using fluorescent assays which monitor the mixing of aqueous contents of liposomes. Overall fusion was modeled as a two-step process involving aggregation of vesicles followed by actual fusion. Different experimental conditions were used which favored either aggregation or fusion as the rate-limiting step in the overall process. When phosphatidylserine liposomes were induced to fuse by 4 mM Ca2+ plus 5 mM Mg2+, preincubation with arachidonic acid showed a dramatically increased overall rate of fusion compared to the same liposomes not treated with fatty acid. When fusion was induced by 3 mM Ca2+, arachidonic acid had little effect. These results were interpreted in terms of the action of arachidonic acid only at the fusion step per se and not the aggregation step. Therefore, the enhancement of the overall fusion rate would be observed solely under conditions where the actual fusion of liposomes was rate limiting (Ca/Mg) rather than the aggregation of liposomes (Ca alone). When other liposome systems were tested, the effect of arachidonic acid was observed only under fusion rate-limiting conditions. Arachidonic acid was found to act synergistically with promoters of liposomal aggregation, such as Mg2+, spermine, and synexin, to enhance the overall rate of liposome fusion, as would be expected from action at separate kinetic steps. The dependence of the fusion rates on arachidonic acid concentration demonstrated an apparently cooperative effect. The structure of the fatty acid is of critical importance in determining its effects, as shown by the fact that 16-doxylstearic acid always increased the rate of fusion while 5-doxylstearic acid always decreased the rate of fusion under all conditions tested. A number of different fatty acids, including oleic acid, elaidic acid, 16-doxylstearic acid, myristic acid, and stearic acid, were effective at increasing the fusion rate to varying extents. In general, unsaturated fatty acids were more effective than saturated ones, either due to partitioning into the membrane or because of structural requirements for promotion of fusion.  相似文献   

11.
Shingles R  Roh MH  McCarty RE 《Plant physiology》1996,112(3):1375-1381
Chloroplast inner envelope membrane vesicles that are loaded with the pH-sensitive fluorophore, pyranine, show rapid internal acidification when nitrite is added. Acidification is dependent upon [delta]pH, with the inside of vesicles being alkaline with respect to the outside. The rate of vesicle acidification was directly proportional to the concentration of nitrite that was added and the imposed pH difference across the membrane. In contrast, added nitrate had no effect on vesicle acidification. Nitrite also caused acidification of asolectin vesicles. The extent of vesicle acidification is dependent on the internal volume of vesicles. Inner envelope and asolectin vesicles that were prepared by extrusion were approximately the same size, allowing them to be compared when the final extent of acidification, measured after the pH gradient had collapsed, was similar. The rate of nitrite-dependent acidification was similar in these two preparations at any single nitrite concentration. These results indicate that nitrite movement occurs by rapid diffusion across membranes as nitrous acid, and this movement is dependent on a proton gradient across the lipid bilayer. Under conditions approximating those in vivo, the rate of diffusion of nitrous acid far exceeds that of nitrite reduction within chloroplasts.  相似文献   

12.
The ability of oligo- and polymers of the basic amino acids L-lysine, L-arginine, L-histidine and L-ornithine to induce lipid intermixing and membrane fusion among vesicles containing various anionic phospholipids has been investigated. Among vesicle consisting of either phosphatidylinositol or mixtures of phosphatidic acid and phosphatidylethanolamine rapid and extensive lipid intermixing, but not complete fusion, was induced at neutral pH by poly-L-ornithine or L-lysine peptides of five or more residues. When phosphatidylcholine was included in the vesicles, the lipid intermixing was severely inhibited. Such lipid intermixing was also much less pronounced among phosphatidylserine vesicles. Poly-L-arginine provoked considerable leakage from the various anionic vesicles and caused significantly less lipid intermixing than L-lysine peptides at neutral pH. When the addition of basic amino acid polymer was followed by acidification to pH 5-6, vesicle fusion was induced. Fusion was more pronounced among vesicles containing phosphatidylserine or phosphatidic acid than among those containing phosphatidylinositol, and occurred also with vesicles whose composition resembles that of cellular membranes (i.e., phosphatidylcholine/phosphatidylethanolamine/phosphatidylserine, 50:30:20, by mol). Liposomes with this composition are resistant to fusion by Ca2+ or by acidification after lectin-mediated contact. The tight interaction among vesicles at neutral pH, resulting in lipid intermixing, does not seem to be necessary for the fusion occurring after acidification, but the basic peptides nevertheless appear to play a more active role in the fusion process than simply bringing the vesicles in contact. However, protonation of the polymer side chains and transformation of the polymer into a polycation does not explain the need for acidification, since the pH-dependence was quite similar for poly(L-histidine)- and poly(L-lysine)-mediated fusion.  相似文献   

13.
J Wilschut  S Nir  J Scholma  D Hoekstra 《Biochemistry》1985,24(17):4630-4636
We have investigated the kinetics of Ca2+-induced aggregation and fusion of large unilamellar vesicles composed of an equimolar mixture of bovine heart cardiolipin and dioleoylphosphatidylcholine. Mixing of bilayer lipids was monitored with an assay based on resonance energy transfer (RET) and mixing of aqueous vesicle contents with the Tb/dipicolinate assay. The results obtained with either assay were analyzed in terms of a mass action kinetic model, providing separate rate constants for vesicle aggregation and for the fusion reaction proper. At different Ca2+ concentrations, either at 25 degrees C or at 37 degrees C, aggregation rate constants derived from the data obtained with the RET assay were the same as those derived from the Tb/dipicolinate data, indicating that mixing of bilayer lipids occurred only during vesicle aggregation events that resulted in mixing of aqueous contents as well. At 25 degrees C, identical fusion rate constants were obtained with either assay, indicating that at this temperature the probability of lipid mixing and that of aqueous contents mixing, occurring after vesicle aggregation, were the same. The fusion rate constants for the RET assay increased more steeply with increasing temperature than the fusion rate constants derived from the Tb/dipicolinate data. As a result, at 37 degrees C the tendency of the vesicles, after aggregation, to mix lipids was slightly higher than their tendency to mix aqueous contents. The aggregation rate constants increased steeply with Ca2+ concentrations increasing in a narrow range (9.5-11 mM), indicating that, in addition to a Ca2+-dependent charge neutralization on the vesicle surface, structural changes in the lipid bilayer are involved in the aggregation process.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Membrane vesicles composed of fatty acids can be made to grow and divide under laboratory conditions, and thus provide a model system relevant to the emergence of cellular life. Fatty acid vesicles grow spontaneously when alkaline micelles are added to buffered vesicles. To investigate the mechanism of this process, we used stopped-flow kinetics to analyze the dilution of non-exchanging FRET probes incorporated into preformed vesicles during growth. Oleate vesicle growth occurs in two phases (fast and slow), indicating two pathways for the incorporation of fatty acid into preformed vesicles. We propose that the fast phase, which is stoichiometrically limited by the preformed vesicles, results from the formation of a "shell" of fatty acid around a vesicle, followed by rapid transfer of this fatty acid into the preformed vesicle. The slower phase may result from incorporation of fatty acid which had been trapped in an intermediate state. We provide independent evidence for the rapid transformation of micelles into an aggregated intermediate form after transfer from high to low pH. Our results show that the most efficient incorporation of added oleate into oleic acid/oleate vesicles occurs under conditions that avoid a large transient increase in the micelle/vesicle ratio.  相似文献   

15.
Annexin A4 belongs to a class of Ca(2+)-binding proteins for which different functions in the cell have proposed, e.g. involvement in exocytosis and in the coagulation process. All these functions are related to the ability of the annexins to bind to acidic phospholipids. In this study the interaction of annexin A4 with large unilamellar vesicles (LUV) prepared from phosphatidylserine (PS) or from phosphatidic acid (PA) is investigated at neutral and acidic pH. Annexin A4 strongly binds to either lipid at acidic pH, whereas at neutral pH only weak binding to PA and no binding to PS occurs. Addition of 40 microM Ca(2+) leads to a strong binding to the lipids also at neutral pH. This is caused by the different electric charge of the protein below and above its isoelectric point. Binding of annexin A4 induces dehydration of the vesicle surface. The strength of the effects is much greater at pH 4 than at pH 7.4. At pH 7.4 annexin A4 reduces the Ca(2+)-threshold concentration necessary to induce fusion of PA LUV. The Ca(2+) induced fusion of PS LUV is not affected by annexin A4 at pH 7.4. At pH 4 annexin A4 induces fusion of either vesicles without Ca(2+). Despite the low binding extents at neutral pH annexin A4 induces a Ca(2+) independent leakage of PS- or PA-LUV. The leakage extent is increased at acidic pH. From the data two suggestions are made: (1) At pH 4 annexin A4 (at least partially) penetrates into the bilayer in contrast to the preferred location at the vesicle surface at neutral pH. The conformation of annexin A4 seems to be different at the two conditions. (2) At neutral pH, Annexin A4 seems to be able to bind two PA vesicles simultaneously; however, only one PS vesicle at the same time. This behavior might be related to a recently described double Ca(2+) binding site, which appears to be uniquely suited for PS.  相似文献   

16.
Calcium uptake was examined in sealed plasma membrane vesicles isolated from red beet (Beta vulgaris L.) storage tissue using (45)Ca(2+). Uptake of (45)Ca(2+) by the vesicles was ATP-dependent and radiotracer accumulated by the vesicles could be released by the addition of the calcium ionophore A23187. The uptake was stimulated by gramicidin D but slightly inhibited by carbonylcyanide m-chlorophenylhydrazone. Although the latter result might suggest some degree of indirect coupling of (45)Ca(2+) uptake to ATP utilization via deltamuH(+), no evidence for a secondary H(+)/Ca(2+) antiport in this vesicle system could be found. Following the imposition of an acid-interior pH gradient, proton efflux from the vesicle was not enhanced by the addition of Ca(2+) and an imposed pH gradient could not drive (45)Ca(2+) uptake. Optimal uptake of (45)Ca(2+) occurred broadly between pH 7.0 and 7.5 and the transport was inhibited by orthovanadate, N,N'-dicyclohexylcarbodiimide, and diethylstilbestrol but insensitive to nitrate and azide. The dependence of (45)Ca(2+) uptake on both calcium and Mg:ATP concentration demonstrated saturation kinetics with K(m) values of 6 micromolar and 0.37 millimolar, respectively. While ATP was the preferred substrate for driving (45)Ca(2+) uptake, GTP could drive transport at about 50% of the level observed for ATP. The results of this study demonstrate the presence of a unique primary calcium transport system associated with the plasma membrane which could drive calcium efflux from the plant cell.  相似文献   

17.
Two major types of planar lipid bilayers, painted and folded, were compared with respect to vesicle fusion using one chamber for the preparation of both bilayers. Liposomes containing ion channels composed of nystatin and ergosterol were used as the vesicle sample. Fusion of the liposome to either bilayer elicited a spike-like current change, which corresponds to a fusion event. The lag time between the first fusion event and the addition of the vesicles is an index of the ease with which the vesicles fuse with the bilayers. The lag time in the painted bilayer at a KCl concentration (cis) of 450 mM was 1.58+/-1.18 min, similar to that in the folded bilayer (1.65+/-0.64 min). The lag time decreased with increase of the osmotic difference across the painted bilayer, whereas this change was small in the folded bilayer. The fusion of the liposomes to the painted bilayer was markedly reduced by stopping the stirring in the cis compartment, whereas the fusion to the folded bilayer was not affected significantly. These results imply that no practical difference exists in the ability of vesicles to fuse with the painted and folded bilayers. For the study of single channel behavior, the painted bilayer could have an advantage because simply stopping the stirring prevents excess fusion.  相似文献   

18.
Role of channels in the fusion of vesicles with a planar bilayer.   总被引:7,自引:3,他引:7       下载免费PDF全文
Fluorescence microscopy combined with electrical conductance measurements were used to assess fusion of phospholipid vesicles with a planar bilayer. Large unilamellar vesicles (0.5-3 microns diam.) filled with the fluorescent dye, calcein, were made both with or without porin channels. Vesicle-bilayer fusion was induced by increasing the osmolarity of the solution on the side of the bilayer to which the vesicles were added. Fusion was detected optically by the fluorescent flash due to release of vesicular contents. Although both porin-containing and porin-free vesicles give the same kind of flash upon content release, the conditions necessary to induce release are very different. Only 4% of the porin-free vesicles fuse (release their contents) when subjected to 3 M urea. However, the same conditions induce 53% of the porin-containing vesicles to fuse and most of these fusions occur at a lower osmolarity ([urea] less than 400 mM). Thus channels greatly enhance fusion in this model system. A physical model based on the postulate that fusion is induced by an increase in surface tension, predicts that three conditions are necessary for fusion in this system: (a) an open channel in the vesicle membrane, (b) an osmotic gradient across the bilayer, and (c) the vesicle in contact with the planar membrane. These are the conditions that experimentally produce fusion in the model system.  相似文献   

19.
The effect of incorporation of glycophorin, the major integral sialoglycoprotein of the erythrocyte membrane, into bovine brain phosphatidylserine (PS) vesicles on the Ca2+-induced fusion of these vesicles has been investigated. Fusion was monitored by the terbium-dipicolinic acid fluorescence assay for the mixing of aqueous contents of the vesicles and by a resonance energy transfer assay that follows the intermixing of membrane lipids. The Ca2+-induced fusion of PS vesicles is completely prevented by incorporation of glycophorin (molar ratio of PS/glycophorin = 400-500:1) for Ca2+ concentrations up to 50 mM. The ability to fuse is partially restored after treating the glycophorin-containing vesicles with neuraminidase, which removes the negatively charged sialic acid residues of glycophorin. Fusion is further facilitated by trypsin treatment, removing the entire extravesicular glycosylated head group of glycophorin. However, Ca2+-induced fusion of enzyme-treated glycophorin-PS vesicles proceeds at a slower rate and to a smaller extent than fusion of protein-free PS vesicles. The influence of the aggregation state of the glycophorin molecules on fusion has been investigated in experiments using wheat germ agglutinin (WGA). Addition of WGA to the glycophorin-PS vesicles does not induce fusion. However, upon subsequent addition of Ca2+, distinct fusion occurs concomitantly with release of vesicle contents. The inhibition of Ca2+-induced fusion of PS vesicles by incorporation of glycophorin is explained by a combination of steric hindrance and electrostatic repulsion between the vesicles by the glycosylated head group of glycophorin and a direct bilayer stabilization by the intramembranous hydrophobic part of the glycophorin molecule.  相似文献   

20.
Ca2+-induced fusion of phospholipid vesicles (phosphatidylcholine/phosphatidic acid, 9:1 mol/mol) prepared by ethanolic injection was followed by five different procedures: resonance energy transfer, light scattering, electron microscopy, intermixing of aqueous content, and gel filtration through Sepharose 4-B. The five methods gave concordant results, showing that vesicles containing only 10% phosphatidic acid can be induced to fuse by millimolar concentrations of Ca2+. When the fusing capability of several soluble proteins was assayed, it was found that concanavalin A, bovine serum albumin, ribonuclease, and protease were inactive. On the other hand, lysozyme, L-lactic dehydrogenase, and muscle and yeast glyceraldehyde-3-phosphate dehydrogenase were capable of inducing vesicle fusion. Glyceraldehyde-3-phosphate dehydrogenase from rabbit muscle, the most extensively studied protein, proved to be very effective: 0.1 microM was enough to induce complete intermixing of bilayer phospholipid vesicles. Under conditions used in this work, fusion was accompanied by leakage of internal contents. The fusing capability of glyceraldehyde-3-phosphate dehydrogenase was not affected by 5 mM ethylenediaminetetraacetic acid. The Ca2+ concentration in the medium, as determined by atomic absorption spectroscopy, was 5 ppm. Heat-denatured enzyme was incapable of inducing fusion. We conclude that glyceraldehyde-3-phosphate dehydrogenase is a soluble protein inherently endowed with the capability of fusing phospholipid vesicles.  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