首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
R E Brown  K J Hyland 《Biochemistry》1992,31(43):10602-10609
The spontaneous incorporation of II3-N-acetylneuraminosylgangliotetraosylceramide (GM1) from its micelles into phospholipid bilayer vesicles has been investigated to determine whether curvature-induced changes in membrane lipid packing influence ganglioside uptake. Use of conventional liquid chromatography in conjunction with technically-improved molecular sieve gels permits ganglioside micelles to be separated from phospholipid vesicles of different average size including vesicles with diameters smaller than 40 nm and, thus, allows detailed study of native ganglioside GM1 incorporation into model membranes under conditions where complicating processes like fusion are readily detected if present. At 45 degrees C, the spontaneous transfer rate of GM1 from its micelles to small unilamellar vesicles (SUVs) comprised of 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) is at least 3-fold faster than that to similar composition large unilamellar vesicles (LUVs) prepared by octyl glucoside dialysis. Careful analysis of ganglioside GM1 distribution among vesicle populations of differing average size reveals that GM1 preferentially incorporates into the smaller vesicles of certain populations. This behavior is observed in SUVs as well as in LUV-SUV mixtures and actually serves as a sensitive indicator for the presence of trace quantities of SUVs in various LUV preparations. Analysis of the results shows that both differences in the diffusional collision frequency between GM1 monomers and either SUVs or LUVs and curvature-induced changes in the interfacial lipid packing in either SUVs or LUVs can dramatically influence spontaneous ganglioside uptake.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Thrombin-activated factor Va and factor Va subunit binding to large-volume vesicles was investigated by a technique based on the separation by centrifugation of phospholipid-bound protein from the bulk solution. This technique allows the direct measurement of free-protein concentration. It is concluded that the phospholipid binding site on factor Va is located on a basic factor Va subunit with Mr 80 000 (factor Va-LC). The effects of phospholipid vesicle composition, calcium concentration, pH, and ionic strength on the equilibrium constants of factor Va- and factor Va-LC-phospholipid interaction were studied. Factor Va and factor Va-LC binding to phospholipid requires the presence of negatively charged phospholipids. It is further demonstrated that the following occur: (a) Calcium ions compete with factor Va and factor Va-LC for phospholipid-binding sites. (b) The dissociation constant of protein-phospholipid interaction increases with the ionic strength, whereas the maximum protein-binding capacity of the phospholipid vesicle was not affected by ionic strength. (c) The dissociation constant for factor Va-phospholipid interaction depends on pH when the vesicle consists of phosphatidic acid. It is concluded that factor Va-phospholipid interaction is primarily electrostatic in nature, where positively charged groups on the protein directly interact with the phosphate group of net negatively charged phospholipids. The results suggest that factor Va, like factor Xa and prothrombin, has the characteristics of an extrinsic membrane protein.  相似文献   

3.
Sedimentation field flow fractionation (SF3), a relatively new instrumentation methodology for separating particles according to size, has been used to monitor changes in vesicle size during the formation of fused unilamellar vesicles of dipalmitoylphosphatidylcholine. The fusion of 500-A small unilamellar vesicles to 700-A large unilamellar vesicles (LUVs) and the slower conversion to 950-A LUVs at 4 degrees C was readily monitored by SF3 over the time course of a month. Changes in the physical state of these vesicles as a result of structural reorganizations were characterized by high sensitivity differential scanning calorimetry. The advantages and limitations of SF3 are discussed and correlated with electron microscopy and gel filtration.  相似文献   

4.
The ability of high pressure to dissociate several peripheral protein-membrane complexes was investigated. Three vitamin K-dependent proteins (factor X, protein Z, and prothrombin) dissociated from small unilamellar vesicles (SUVs, 30 nm diameter) composed of 25% phosphatidylserine (PS) and 75% phosphatidylcholine (PC) at comparable pressures (midpoints of 0.3-0.6 kbar). The pressure-induced dissociation curves for the factor X-SUV interaction followed the expected behavior for an interaction with an apparent dissociation equilibrium constant at atmospheric pressure, KD(atm), of 9 x 10(-7) M and a change in volume of association, delta Va, of 88 mL/mol. Factor X also dissociated from large unilamellar vesicles (LUVs, 100 nm diameter, 25% PS:75% PC) with a midpoint of 0.5 kbar. A second group of calcium-dependent membrane-binding proteins included protein kinase C (PKC), a 64-kDa protein, and a 32-kDa protein. The 32-kDa protein dissociated from SUVs (midpoint of 0.8 kbar), whereas PKC and the 64-kDa protein did not dissociate to a significant degree. The differences in dissociability of these proteins appeared to be a result of the differences in their KD(atm)'s (decreased dissociability with decreased KD(atm)). This pattern was further demonstrated by the relatively high midpoint of dissociation (1.1-1.4 kbar) of serum amyloid P component (SAP; KD(atm) ca. 10(-11)) and the limited dissociation of factor Va light chain (KD(atm) ca. 10(-11)). Changing the vesicle composition to phosphatidylethanolamine in place of PC gave higher affinity and decreased dissociation of the 32-kDa protein and SAP.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
The transfer kinetics of the neutral glycosphingolipid gangliotetraosylceramide (asialo-GM1) were investigated by monitoring tritiated asialo-GM1 movement from donor to acceptor vesicles. Two different methods were employed to separate donor and acceptor vesicles at desired time intervals. In one method, a negative charge was imparted to dipalmitoylphosphatidylcholine donor vesicles by including 10 mol% dipalmitoylphosphatidic acid. Donors were separated from neutral dipalmitoylphosphatidylcholine acceptor vesicles by ion-exchange chromatography. In the other method, small, unilamellar donor vesicles (20-nm diameter) and large, unilamellar acceptor vesicles (70-nm diameter) were coincubated at 45 degrees C and then separated at desired time intervals by molecular sieve chromatography. The majority of asialo-GM1 transfer to acceptor vesicles occurred as a slow first-order process with a half-time of about 24 days assuming that the relative concentration of asialo-GM1 in the phospholipid matrix was identical in each half of the donor bilayer and that no glycolipid flip-flop occurred. Asialo-GM1 net transfer was calculated relative to that of [14C]cholesteryl oleate, which served as a nontransferable marker in the donor vesicles. A nearly identical transfer half-time was obtained when the phospholipid matrix was changed from dipalmitoylphosphatidylcholine to palmitoyloleoylphosphatidylcholine. Varying the acceptor vesicle concentration did not significantly alter the asialo-GM1 transfer half-time. This result is consistent with a transfer mechanism involving diffusion of glycolipid through the aqueous phase rather than movement of glycolipid following formation of collisional complexes between donor and acceptor vesicles. When viewed within the context of other recent studies involving neutral glycosphingolipids, these findings provide additional evidence for the existence of microscopic, glycosphingolipid-enriched domains within the phospholipid bilayer.  相似文献   

6.
The thermodynamics of binding of the antibacterial peptide magainin 2 amide (M2a) to negatively charged small (SUVs) and large (LUVs) unilamellar vesicles has been studied with isothermal titration calorimetry (ITC) and CD spectroscopy at 45 degrees C. The binding isotherms as well as the ability of the peptide to permeabilize membranes were found to be qualitatively and quantitatively similar for both model membranes. The binding isotherms could be described with a surface partition equilibrium where the surface concentration of the peptide immediately above the plane of binding was calculated with the Gouy-Chapman theory. The standard free energy of binding was deltaG0 approximately -22 kJ/mol and was almost identical for LUVs and SUVs. However, the standard enthalpy and entropy of binding were distinctly higher for LUVs (deltaH0 = -15.1 kJ/mol, deltaS0 = 24.7 J/molK) than for SUVs (deltaH0 = -38.5 kJ/mol, deltaS0 = -55.3 J/molK). This enthalpy-entropy compensation mechanism is explained by differences in the lipid packing. The cohesive forces between lipid molecules are larger in well-packed LUVs and incorporation of M2a leads to a stronger disruption of cohesive forces and to a larger increase in the lipid flexibility than peptide incorporation into the more disordered SUVs. At 45 degrees C the peptide easily translocates from the outer to the inner monolayer as judged from the simulation of the ITC curves.  相似文献   

7.
The binding of peptides or proteins to a bilayer membrane is often coupled with a random coil-->alpha-helix transition. Knowledge of the energetics of this membrane-induced folding event is essential for the understanding of the mechanism of membrane activity. In a recent study [Wieprecht et al., J. Mol. Biol. 294 (1999) 785-794], we have developed an approach which allows an analysis of the energetics of membrane-induced folding. We have systematically varied the helix content of the amphipathic peptide magainin-2-amide by synthesizing analogs where two adjacent amino acid residues were substituted by their corresponding D-enantiomers and have measured their binding to small unilamellar vesicles (SUVs). Correlation of the binding parameters with the helicities allowed the evaluation of the thermodynamic parameters of helix formation. Since SUVs (30 nm in diameter) are characterized by a non-ideal lipid packing due to their high membrane curvature, we have now extended our studies to large unilamellar vesicles (LUVs) (100 nm in diameter) with a lipid packing close to planar membranes. While the free energy of binding was similar for SUVs and LUVs, the binding enthalpies and entropies were distinctly different for the two membrane systems. The thermodynamic parameters of the coil-helix transition were nevertheless not affected by the vesicle size. Helix formation at the membrane surface of LUVs (SUVs) was characterized by an enthalpy change of -0.8 (-0.7) kcal/mol per residue, an entropy change of-2.3 (-1.9) cal/mol K per residue, and a free energy change of -0.12 (-0.14) kcal/mol per residue. Helix formation accounted for approximately 50% of the free energy of binding underlining its major role as a driving force for membrane-binding.  相似文献   

8.
Interactions of apolipoprotein A-I (apoA-I) with cell membranes appear to be important in the initial steps of reverse cholesterol transport. The objective of this work was to examine the effect of three distinct conformations of apoA-I (lipid-free and in 78 A or 96 A reconstituted high density lipoproteins, rHDL) on its ability to bind to, and abstract lipids from, palmitoyl oleoyl phosphatidylcholine membrane vesicles (small unilamellar vesicles, SUV, and giant unilamellar vesicles, GUV). The molecular interactions were observed by two-photon fluorescence microscopy, and the binding parameters were quantified by gel-permeation chromatography or isothermal titration microcalorimetry. Rearrangement of apoA-I-containing particles after exposure to SUVs was examined by native gel electrophoresis. The results indicate that lipid-free apoA-I binds reversibly, with high affinity, to the vesicles but does not abstract a significant amount of lipid nor perturb the vesicle structure. The 96 A rHDL, where all the amphipathic helices of apoA-I are saturated with lipid within the particles, do not bind to vesicles or perturb their structure. In contrast, the 78 A rHDL have a region of apoA-I, corresponding to a few amphipathic helical segments, which is available for external or internal phospholipid binding. These particles bind to vesicles with measurable affinity (lower than lipid-free apoA-I), abstract lipids from the membranes, and form particles of larger diameters, including 96 A rHDL. We conclude that the conformation of apoA-I regulates its binding affinity for phospholipid membranes and its ability to abstract lipids from the membranes.  相似文献   

9.
Association of protein kinase C with phospholipid vesicles   总被引:7,自引:0,他引:7  
The Ca2+- and phospholipid-dependent protein kinase, protein kinase C (PKC), was purified from bovine brain by a modified procedure that provided sufficient quantities of stable protein for analysis of physical properties of protein-membrane binding. The binding of PKC to phospholipid vesicles of various compositions was investigated by light-scattering and fluorescence energy transfer measurements. The binding properties for membranes of low phosphatidylserine (PS) content were consistent with a peripheral membrane association; PKC showed Ca2+ -dependent binding to phospholipid vesicles containing phosphatidylserine, phosphatidylinositol, or phosphatidylglycerol. Membranes containing 0-20% PS (the remainder of the phospholipid was phosphatidylcholine) bound less protein than membranes containing greater than 20% PS; the factor limiting protein binding to membranes containing low PS appeared to be the availability of acidic phospholipids. Increasing the PS content above 20% did not increase the amount of membrane-bound protein at saturation, and the limiting factor was probably steric packing of protein on the membrane surface. The membranes bound about 1 g of protein/g of phospholipid at steric saturation. Binding was of relatively high affinity (Kd less than 5 nM), and the association rate was rapid on the time scale of the experiments. Addition of ethylene glycol bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid to phospholipid-bound PKC caused dissociation of the complex, and the properties of this dissociation indicated an equilibrium binding of protein to membrane. However, only partial dissociation of PKC was achieved when the PS content of the vesicles exceeded 20%. A number of comparisons revealed that binding of protein to the membrane, even in the presence of phorbol esters, was insufficient for development of enzyme activity.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Osmotic properties of large unilamellar vesicles prepared by extrusion.   总被引:8,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   

11.
We have studied the heme oxidation kinetics of purified human hemoglobin (Hb) in the presence of lipid vesicles of dipalmitoyl phosphatidylcholine and bovine brain phosphatidylserine that exhibited minimal lipid peroxidation. We showed that the lipid vesicles enhanced Hb oxidation and that small unilamellar vesicles (SUVs) exerted a larger effect than large unilamellar vesicles (LUVs). We have determined pseudo first-order rate constants for the initial disappearance of oxygenated ferrous Hb (k0) and for the initial formation of several ferric Hb species (methemoglobin, hemichrome, and choleglobin) in the presence of SUVs and LUVs. k0 and other rate constants depended linearly on lipid-to-hemoglobin molar ratio (lipid/Hb), with k0SUV (h-1) = k0auto (h-1) + 3.7 x 10(-3) x lipid/Hb, and k0LUV (h-1) = k0auto (h-1) + 0.2 x 10(-3) x lipid/hb, where k0auto is the rate constant for Hb autoxidation in the absence of vesicles. Thus, in the absence of lipid peroxidation products, lipid vesicles themselves promote Hb oxidation by enhancing the rate of Hb oxidation. The enhanced oxidation was inhibited by catalase, but not by butylated hydroxytoluene. The rate constants were independent of Hb concentration, in the range of about 3.1 to 100 microM. We suggest that the lipid surface properties, including surface curvature, surface energy, and hydrophobicity, promote hemoglobin oxidation.  相似文献   

12.
Transbilayer lipid distribution of small unilamellar vesicles (SUVs) and large unilamellar vesicles (LUVs) was measured using 31P-nuclear magnetic resonance (NMR) spectroscopy, chemical modification with 2,4,6-trinitrobenzene sulfonic acid (TNBS) and dithionite reduction of N-(7-nitrobenz-2-oxa-1,3-diazole-4-yl)-labeled lipid (NBD-lipid). The dithionite assay was the most reproducible of the three assays, with 1.2% error for SUVs and 3.9% error for LUVs. The dithionite assay also agreed best with theoretical inner:outer leaflet ratios, based on vesicle diameters determined by electron microscopy (Thomas et al. (1989) Biochem. Biophys. Acta 978, 85–90). Dithionite assay measurements were within 2.7% of theoretical ratios for SUVs and 2.3% for LUVs, while the NMR assay for SUVs was 14% lower than theoretical ratios and 23% lower for LUVs. The accuracy of NBD-lipids as markers for total transbilayer lipid was investigated. NBD-labeled phosphatidylserine, phosphatidylcholine and phosphatidylglycerol were accurate markers for total transbilayer lipid distribution, as their distributions were in close agreement with theoretical ratios. However, NBD-labeled phosphatidylethanolamine displayed a slight preference for the inner leaflet at low mole fractions of phosphatidylethanolamine, while native phosphatidylethanolamine showed a preference for the outer leaflet at the same concentration. NBD-labeled phosphatidic acid also showed a slight preference for the inner leaflet. We conclude that although dithionite-based assessment of NBD-labeled lipids across membrane bilayers can be a powerful analytical tool, caution must be used in the interpretation of results.  相似文献   

13.
Complement components C5b-6 and C7 assemble to form C5b-7, which then interacts with membranes and commits the membrane attack complex to a target site. This protein-membrane association event was investigated to determine possible structural features that could contribute to a selective membrane attack. This system may also suggest general properties of protein-membrane insertion events. Initial binding of C5b-6 to membranes could potentially determine the site of assembly. However, binding of C5b-6 to membranes required phosphatidylglycerol or phosphatidic acid produced from egg phosphatidylcholine while binding of C5b-6 to phosphatidylcholine, phosphatidylserine, or phosphatidylinositol was undetectable. Binding to phosphatidic acid was irreversible, and the bound C5b-6 could no longer interact with C7. In contrast, C5b-7 interacted with all phospholipids tested. The rate-limiting process was the interaction of C5b-6 and C7, which displayed bimolecular properties and an activation energy of 37 kcal/mol. The C5b-7 complex showed 20-fold selectivity for small unilamellar phospholipid vesicles over large unilamellar vesicles. Vesicles carrying high negative charge densities were selected over neutral vesicles by a factor of about 5. Vesicles formed from phospholipids with short, saturated hydrocarbon side chains (dimyristoylphosphatidylcholine and dipalmitoylphosphatidylcholine) were about 5-fold less effective than those formed from phospholipids with natural fatty acid distributions. The gel vs. fluid state had little influence on C5b-7 insertion.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
N E Gabriel  M F Roberts 《Biochemistry》1986,25(10):2812-2821
Stable unilamellar vesicles formed spontaneously upon mixing aqueous suspensions of long-chain phospholipid (synthetic, saturated, and naturally occurring phosphatidylcholine, phosphatidylethanolamine, and sphingomyelin) with small amounts of short-chain lecithin (fatty acid chain lengths of 6-8 carbons) have been characterized by using NMR spectroscopy, negative staining electron microscopy, differential scanning calorimetry, and Fourier transform infrared (FTIR) spectroscopy. This method of vesicle preparation can produce bilayer vesicles spanning the size range 100 to greater than 1000 A. The combination of short-chain lecithin and long-chain lecithin in its gel state at room temperature produces relatively small unilamellar vesicles, while using long-chain lecithin in its liquid-crystalline state produces large unilamellar vesicles. The length of the short-chain lecithin does not affect the size distribution of the vesicles as much as the ratio of short-chain to long-chain components. In general, additional short-chain decreases the average vesicle size. Incorporation of cholesterol can affect vesicle size, with the solubility limit of cholesterol in short-chain lecithin micelles governing any size change. If the amount of cholesterol is below the solubility limit of micellar short-chain lecithin, then the addition of cholesterol to the vesicle bilayer has no effect on the vesicle size; if more cholesterol is added, particle growth is observed. Vesicles formed with a saturated long-chain lecithin and short-chain species exhibit similar phase transition behavior and enthalpy values to small unilamellar vesicles of the pure long-chain lecithin prepared by sonication. As the size of the short-chain/long-chain vesicles decreases, the phase transition temperature decreases to temperatures observed for sonicated unilamellar vesicles. FTIR spectroscopy confirms that the incorporation of the short-chain lipid in the vesicle bilayer does not drastically alter the gauche bond conformation of the long-chain lipids (i.e., their transness in the gel state and the presence of multiple gauche bonds in the liquid-crystalline state).  相似文献   

15.
B Mütsch  N Gains  H Hauser 《Biochemistry》1986,25(8):2134-2140
The kinetics of lipid transfer from small unilamellar vesicles as the donor to brush border vesicles as the acceptor have been investigated by following the transfer of radiolabeled or spin-labeled lipid molecules in the absence of exchange protein. The labeled lipid molecules studied were various radiolabeled and spin-labeled phosphatidylcholines, radiolabeled cholesteryl oleate, and a spin-labeled cholestane. At a given temperature and brush border vesicle concentration similar pseudo-first-order rate constants (half-lifetimes) were observed for different lipid labels used. The lipid transfer is shown to be an exchange reaction leading to an equal distribution of label in donor and acceptor vesicles at equilibrium (time t----infinity). The lipid exchange is a second-order reaction with rate constants being directly proportional to the brush border vesicle concentration. The results are only consistent with a collision-induced exchange of lipid molecules between small unilamellar phospholipid vesicles and brush border vesicles. Other mechanisms such as collision-induced fusion or diffusion of lipid monomers through the aqueous phase are negligible at least under our experimental conditions.  相似文献   

16.
K M Eum  G Riedy  K H Langley  M F Roberts 《Biochemistry》1989,28(20):8206-8213
Small unilamellar vesicles which form when gel-state long-chain phosphatidylcholines are mixed with micellar short-chain lecithins undergo an increase in size as the long-chain species melts to its liquid-crystalline form. Analysis of the vesicle population with quasi-elastic light scattering shows that the particle size increases from 90-A radius to greater than 5000-A radius. Resonance energy transfer experiments show total mixing of lipid probes with unlabeled vesicles only when the Tm of the long-chain phosphatidylcholine is exceeded. This implies that the large size change represents a fusion process. Aqueous compartments are also mixed during this transition. 31P NMR analysis of the vesicle mixtures above the phase transition shows a great degree of heterogeneity with large unilamellar particles coexisting with oligo- and multilamellar structures. Upon cooling the vesicles below the Tm, the original size distribution (e.g., small unilamellar vesicles) is obtained, as monitored by both quasi-elastic light scattering and 31P NMR spectroscopy. This temperature-induced fusion of unilamellar vesicles is concentration dependent and can be abolished at lower total phospholipid concentrations. It occurs over a wide range of long-chain to short-chain ratios and occurs with 1-palmitoyl-2-stearoylphosphatidylcholine and dimyristoylphosphatidylcholine as well. Characterization of this fusion event is used to understand the anomalous kinetics of water-soluble phospholipases toward these unusual vesicles.  相似文献   

17.
The size and size distribution of unilamellar phospholipid vesicles present in unsonicated phosphatidic acid and mixed phosphatidic acid/phosphatidylcholine dispersions were determined by gel filtration, quasi-elastic light scattering and freeze-fracture electron microscopy. The vesiculation in these dispersions was induced by a transient increase in pH as described previously (Hauser, H. and Gains, N. (1982) Proc. Natl. Acad. Sci. USA 79, 1683–1687). The resulting phospholipid dispersions are heterogeneous consisting of small unilamellar vesicles (average radius r < 50 nm) and large unilamellar vesicles (average r ranging from about 50 to 500 nm). The smallest vesicles with r = 11 ± 2 nm are observed with dispersions of pure phosphatidic acid, the population of these vesicles amounting to about 80% of the total lipid. With increasing phosphatidylcholine content the radius of the small unilamellar vesicles increases and at the same time the population of small unilamellar vesicles decreases. The average radius of small unilamellar vesicles present in phosphatidic acid/phosphatidylcholine dispersions (mole ratio, 1:1) is 17.5 ± 2 nm, the population of these vesicles amounting to about 70% of the total lipid. By a combination of gel filtration, quasi-elastic light scattering and freeze-fracture electron microscopy it was possible to characterize the large unilamellar vesicles. This population is heterogeneous with its mean radius also increasing with increasing phosphatidylcholine content. After separating the large unilamellar vesicles from small unilamellar vesicles on Sepharose 4B it can be shown by quasi-elastic light scattering that in pure phosphatidic acid dispersions 80–90% of the large unilamellar vesicle population consist of vesicles with a mean radius of 170 nm. In mixed phosphatidic acid/phosphatidylcholine dispersions this radius increases to about 265 nm as the phosphatidylcholine content is raised to 90 mol%.  相似文献   

18.
G Ramsay  R Prabhu  E Freire 《Biochemistry》1986,25(8):2265-2270
A newly designed high-sensitivity isothermal reaction calorimetry system has been used to investigate the thermodynamics of the association between myelin basic protein and phosphatidylserine vesicles. This instrument has allowed us to measure directly the energetics of the protein-lipid interaction under various conditions. Above the phospholipid phase transition temperature the enthalpy of association is highly exothermic amounting to -160 kcal/mol of protein. Below the phospholipid phase transition temperature the enthalpy of association is exothermic at protein/lipid ratios smaller than 1/50 and endothermic at higher protein/lipid ratios. These studies indicate that the association of myelin basic protein to phosphatidylserine vesicles consists of at least two stages involving different types of binding. The first stage, at low protein/lipid ratios, involves a strong exothermic association of the protein to the membrane and the second, at high protein/lipid ratios, a weaker association probably involving attachment of the protein to the membrane surface only. In the gel phase the second binding stage is endothermic and appears to be correlated with the formation of large vesicle aggregates. This vesicle aggregation is a reversible process dependent upon the physical state of the membrane. The isothermal titration studies have been complemented with high-sensitivity differential scanning calorimetry experiments. It is shown that the dependence of the phospholipid transition enthalpy on the protein/lipid molar ratio can be expressed in terms of the different protein-membrane association enthalpies in the gel and fluid phases of the membrane.  相似文献   

19.
Nano-differential scanning calorimetry (nano-DSC) is a powerful tool in the investigation of unilamellar (small unilamellar, SUVs, or large unilamellar, LUVs) vesicles, as well as lipids on supported bilayers, since it measures the main gel-to-liquid phase transition temperature (Tm), enthalpies and entropies. In order to assign these transitions in single component systems, where Tm often occurred as a doublet, nano-DSC, dynamic light scattering and cryo-transmission electron microscopy (cryo-TEM) data were compared. The two Tms were not attributable to decoupled phase transitions between the two leaflets of the bilayer, i.e. nano-DSC measurements were not able to distinguish between the outer and inner leaflets of the vesicle bilayers. Instead, the two Tms were attributed to mixtures of oligolamellar and unilamellar vesicles, as confirmed by cryo-TEM images. Tm for the oligolamellar vesicles was assigned to the peak closest to that of the parent multilamellar vesicle (MLV) peak. The other transition was higher than that of the parent MLVs for 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC), and increased in temperature as the vesicle size decreased, while it was lower in temperature than that of the parent MLVs for 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC), and decreased as the vesicle size decreased. These subtle shifts arose due to small differences in the values of ΔH and ΔS, since Tm is determined by their ratio (ΔH/ΔS). It was not possible to completely eliminate oligolamellar structures for MLVs extruded with the 200 nm pore size filter, even after 120 passes, while these structures were eliminated for MLVs extruded through the 50 nm pore size filter.  相似文献   

20.
Rao CS  Lin X  Pike HM  Molotkovsky JG  Brown RE 《Biochemistry》2004,43(43):13805-13815
Glycolipid transfer protein (GLTP) catalyzes the intermembrane transfer of lipids that have sugars beta-linked to either diacylglycerol or ceramide backbones, including simple glycosphingolipids (GSLs) and gangliosides. The present study provides a quantitative understanding of GLTP action involving bilayer vesicles that have high and low curvature stress, i.e., small and large unilamellar vesicles (SUVs and LUVs). When the GSL intervesicular transfer was monitored in real time using an established fluorescence resonance energy approach, the initial GSL transfer rates (v(0)) and net transfer equilibrium (K(eq)) were determined for GLTP-mediated transfer from SUVs and LUVs over the temperature range of 30-44 degrees C. v(0) exhibited a linear dependence with respect to varying GLTP concentrations (0-143 nM range) in SUVs and LUVs, suggesting a first order dependence on the GLTP bulk concentration. Thermodynamic parameters associated with the GLTP-GSL transition-state complex and GSL net transfer were determined from linear Arrhenius and van't Hoff plots, respectively. Although initial transfer rates were lower for LUVs than for SUVs, the activation energy barriers were higher for LUVs, while the Gibbs's free energy of the transition states were similar. The formation of a transition-state complex was predominantly enthalpy driven, whereas the net transfer of GSLs was mainly entropy driven. The rate-limiting step for GLTP action appeared to be the surface processes leading to the GLTP-GSL complex formation and release associated with a shuttle/carrier mode of action. Because surface processes leading to the GLTP-GSL complex formation were limiting for GLTP action with SUVs and LUVs, it was concluded that GLTP is likely to be a valuable tool to probe and manipulate GSL environments in membranes.  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