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 390 毫秒
1.
(1) The effect of glycophorin, a major intrinsic glycoprotein of the human erythrocyte membrane, on lipid polymorphism has been investigated by 31P-NMR (at 36.4 MHz) and by freeze-fracture electron microscopy. (2) Incorporation of glycophorin into vesicles of 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) results in the formation of unilamellar vesicles (1000–5000 Å diameter) which exhibit 31P-NMR bilayer spectra over a wide range of temperature. A reduction in the chemical shift anisotropy (Δσcsaeff) and an increase in spectral linewidth in comparison to dioleoylphosphatidylcholine liposomes may suggest a decrease in phospholipid headgroup order. (3) 1,2-Dioleoyl-sn-glycero-3-phosphoethanolamine (DOPE), in the presence of excess water, undergoes a bilayer to hexagonal (HII) phospholipid arrangement as the temperature is increased above 0°C. Incorporation of glycophorin into this system stabilizes the bilayer configuration, prohibiting the formation of the HII phase. (4) Cosonication of glycophorin with DOPE in aqueous solution (pH 7.4) produces small, stable unilamellar vesicles (300–1000 Å diameter), unlike DOPE alone which is unstable and precipitates from solution. (5) The current study demonstrates the bilayer stabilizing capacity of an intrinsic membrane protein, glycophorin, most likely by means of a strong hydrophobic interaction between the membrane spanning portion of glycophorin and the hydrophobic region of the phospholipid.  相似文献   

2.
Glycophorin was incorporated into large unilamellar dioleoylphosphatidylcholine vesicles by either a detergent dialysis method using octylglucoside or a method avoiding the use of detergents. The vesicles were characterized and the permeability properties and transbilayer movement of lipids in both vesicles were investigated as a function of the protein concentration and were compared to protein-free vesicles. An insight in the permeability properties of the vesicles was obtained by monitoring the ratio potassium (permeant): dextran (impermeant) trap immediately after separation of the vesicles from the external medium. Glycophorin incorporated without the use of detergents in 1:300 protein:lipid molar ratio induces a high potassium permeability for the majority of the vesicles as judged from the low potassium trap (K+:dextran trap = 0.21). In contrast, the vesicles in which glycophorin is incorporated via the octylglucoside method (1:500 protein:lipid molar ratio) are much less permeable to potassium (K+:dextran trap = 0.67 and t12 of potassium efflux at 22°C is 7.5 h.). The relationship between protein-induced bilayer permeability and lipid transbilayer movement in both vesicle preparations is discussed. Addition of wheat-germ agglutinin to glycophorin-containing vesicles comprised of dioleoylphosphatidylcholine and total erythrocyte lipids caused no or just a small effect (less than 20% release of potassium) on the potassium permeability of these vesicles. Also, addition of lectin to dioleoylphosphatidylethanolamine-glycophorin bilayer vesicles in a 25:1 lipid:glycophorin molar ratio had no effect on the permeability characteristics of the vesicles. In contrast, addition of wheat-germ agglutinin to bilayer vesicles made of dioleoylphosphatidylethanolamine and glycophorin in a 200:1 molar ratio resulted in a release of 74% of the enclosed potassium by triggering a bilayer to hexagonal (HII) phase transition. The role of protein aggregation and the formation of defects in the lipid bilayer on membrane permeability and lipid transbilayer movement is discussed.  相似文献   

3.
Sonicated dispersions of 1,2-dipalmitoylsn-glycero-3-phosphorylcholine and of 1,3-dipalmitoylglycero-2-phosphorylcholine were examined by proton nuclear magnetic resonance (NMR) as a function of temperature. The —(CH2)n)— peak in the spectrum of the sn-3-isomer of dipalmitoylphosphatidylcholine showed the characteristic dramatic changes in the peak intensity and width associated with the phase transition between the liquid crystalline and gel states of the phospholipid. This occurred over a 2–3°C temperature range with the midpoint of the transition at 38.5°C. With the 2-isomer the change in phase took place over a similar temperature range but the midpoint was at 33.8°C. This lower phase transition temperature is presumably the result of increased acyl chain mobility caused by the increased separation of the two acyl chains by the centre carbon of the glycerol backbone. The effect of sonication of the broadening of the range and lowering of the midpoint temperature of the phase transition from that of the corresponding unsonicated dispersions was similar with each isomer. This suggests that the overall geometry of the sonicated vesicles of the isomers is similar.  相似文献   

4.
Dimerization of the transmembrane domain of glycophorin A is mediated by a seven residue motif LIxxGVxxGVxxT through a combination of van der Waals and hydrogen bonding interactions. One of the unusual features of the motif is the large number of β-branched amino acids that may limit the entropic cost of dimerization by restricting side-chain motion in the monomeric transmembrane helix. Deuterium NMR spectroscopy is used to characterize the dynamics of fully deuterated Val80 and Val84, two essential amino acids of the dimerization motif. Deuterium spectra of the glycophorin A transmembrane dimer were obtained using synthetic peptides corresponding to the transmembrane sequence containing either perdeuterated Val80 or Val84. These data were compared with spectra of monomeric glycophorin A peptides deuterated at Val84. In all cases, the deuterium line shapes are characterized by fast methyl group rotation with virtually no motion about the Cα-Cβ bond. This is consistent with restriction of the side chain in both the monomer and dimer due to intrahelical packing interactions involving the β-methyl groups, and indicates that there is no energy cost associated with dimerization due to loss of conformational entropy. In contrast, deuterium NMR spectra of Met81 and Val82, in the lipid interface, reflected greater motional averaging and fast exchange between different side-chain conformers.  相似文献   

5.
Apolipophorin III (apoLp-III) from Locusta migratoria was employed as a model apolipoprotein to gain insight into binding interactions with lipid vesicles. Differential scanning calorimetry (DSC) was used to measure the binding interaction of apoLp-III with liposomes composed of mixtures of 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and sphingomyelin (SM). Association of apoLp-III with multilamellar liposomes occurred over a temperature range around the liquid crystalline phase transition (Lα). Qualitative and quantitative data were obtained from changes in the lipid phase transition upon addition of apoLp-III. Eleven ratios of DMPC and SM were tested from pure DMPC to pure SM. Broadness of the phase transition (T1/2), melting temperature of the phase transition (Tm) and enthalpy were used to determine the relative binding affinity to the liposomes. Multilamellar vesicles composed of 40% DMPC and 60% SM showed the greatest interaction with apoLp-III, indicated by large T1/2 values. Pure DMPC showed the weakest interaction and liposomes with lower percentage of DMPC retained domains of pure DMPC, even upon apoLp-III binding indicating demixing of liposome lipids. Addition of apoLp-III to rehydrated liposomes was compared to codissolved trials, in which lipids were rehydrated in the presence of protein, forcing the protein to interact with the lipid system. Similar trends between the codissolved and non-codissolved trials were observed, indicating a similar binding affinity except for pure DMPC. These results suggested that surface defects due to non-ideal packing that occur at the phase transition temperature of the lipid mixtures are responsible for apolipoprotein-lipid interaction in DMPC/SM liposomes.  相似文献   

6.
It is shown that sonication of phospholipid-water dispersions below the crystalline → liquid crystalline phase transition temperature (Tc) produces bilayer vesicles with structural defects within the bilayer membrane, which permit rapid permeation of ions and catalyze vesicle-vesicle fusion. These structural defects are annihilated simply by annealing the vesicle suspension above Tc. The rate of annealing was found to be slow, of the order of an hour for T = 3 °C above Tc, but annealing is complete within 10 min for T = 10 °C above Tc. It is proposed that these structural defects are fault-dislocations in the bilayer structure, which arise from a population defect in the distribution of the lipid molecules between the outer and inner monolayers, when small bilayer fragments reassemble to form the small bilayer vesicles during the sonication procedure. Such a population defect can only be remedied by lipid transport via the inside ? outside flip-flop mechanism, which would account for the slow kinetics of annealing observed even at 3 °C above the phase transition.  相似文献   

7.
Vesicles have been prepared from 18 : 1c/18 : 1c-phosphatidylcholine with or without purified glycophorin or partially purified band 3 (obtained by organomercurial gel chromatography). The vesicles have been characterized by freeze-fracture electron microscopy, binding studies to DEAE-cellulose, 31P-NMR and K+ trap measurements. Pools of phosphatidylcholine available for exchange have been investigated using phosphatidylcholine exchange protein from bovine liver. The protein-containing vesicles both exhibit exchangeable pools larger than the fraction of phosphatidylcholine in the outer monolayer, whereas in the protein-free vesicles the exchangeable pool is consistent with the outer monolayer. The results indicate that both glycophorin and the partially purified band 3 preparation enhance the transbilayer movement of phosphatidylcholine.  相似文献   

8.
A set of different biophysical approaches has been used to explore the phase behavior of palmitoylsphingomyelin (pSM)/cholesterol (Chol) model membranes in the presence and absence of palmitoylceramide (pCer). Fluorescence spectroscopy of di-4-ANEPPDHQ-stained pSM/Chol vesicles and atomic force microscopy of supported planar bilayers show gel Lβ/liquid-ordered (Lo) phase coexistence within the range XChol = 0-0.25 at 22°C. At the latter compositional point and beyond, a single Lo pSM/Chol phase is detected. In ternary pSM/Chol/pCer mixtures, differential scanning calorimetry of multilamellar vesicles and confocal fluorescence microscopy of giant unilamellar vesicles concur in showing immiscibility, but no displacement, between Lo cholesterol-enriched (pSM/Chol) and gel-like ceramide-enriched (pSM/pCer) phases at high pSM/(Chol + pCer) ratios. At higher cholesterol content, pCer is unable to displace cholesterol at any extent, even at XChol < 0.25. It is interesting that an opposite strong cholesterol-mediated pCer displacement from its tight packing with pSM is clearly detected, completely abolishing the pCer ability to generate large microdomains and giving rise instead to a single ternary phase. These observations in model membranes in the absence of the lipids commonly used to form a liquid-disordered phase support the role of cholesterol as the key determinant in controlling its own displacement from Lo domains by ceramide upon sphingomyelinase activity.  相似文献   

9.
Temperature-controlled Atomic Force Microscopy (TC-AFM) in Contact Mode is used here to directly image the mechanisms by which melting and crystallization of supported, hydrated DPPC bilayers proceed in the presence and absence of the model peptide WALP23. Melting from the gel Lβ′ to the liquid-crystalline Lα phase starts at pre-existing line-type packing defects (grain boundaries) in absence of the peptide. The exact transition temperature is shown to be influenced by the magnitude of the force exerted by the AFM probe on the bilayer, but is higher than the main transition temperature of non-supported DPPC vesicles in all cases due to bilayer–substrate interactions. Cooling of the fluid Lα bilayer shows the formation of the line-type defects at the borders between different gel-phase regions that originate from different nuclei. The number of these defects depends directly on the rate of cooling through the transition, as predicted by classical nucleation theory.The presence of the transmembrane, synthetic model peptide WALP23 is known to give rise to heterogeneity in the bilayer as microdomains with a striped appearance are formed in the DPPC bilayer. This striated phase consists of alternating lines of lipids and peptide. It is shown here that melting starts with the peptide-associated lipids in the domains, whose melting temperature is lowered by 0.8–2.0 °C compared to the remaining, peptide-free parts of the bilayer. The stabilization of the fluid phase is ascribed to adaptations of the lipids to the shorter peptide. The lipids not associated with the peptide melt at the same temperature as those in the pure DPPC supported bilayer.  相似文献   

10.
Electron transport from untreated to mersalyzed microsomal vesicles at the level of NADH-cytochrome b5 reductase or cytochrome b5 has been demonstrated in the absence of added water-soluble electron carriers. A similar effect was shown in the systems “intact mitochondria — mersalyzed microsomes” and “mersalyzed mitochondria— untreated microsomes”. No measurable electron transport between intact and mersalyzed particles of inner mitochondrial membrane was found. The obtained data suggest that the capability to carry out intermembrane electron transfer is specific for NADH-cytochrome b5 reductase and/or cytochrome b5, localized in microsomal and outer mitochondrial membranes.  相似文献   

11.
We describe the interaction of various phospholipases A2 (PLA2) from snake venoms of the family Viperidae (Macrovipera lebetina obtusa, Vipera ursinii renardi, Bothrops asper) with giant unilamellar vesicles (GUVs) composed of natural brain phospholipids mixture, visualized through fluorescence microscopy. The membrane fluorescent probes 8-anilino-1-naphthalenesulfonicacid (ANS), LAUDRAN and PRODAN were used to assess the state of the membrane and specifically mark the lipid packing and membrane fluidity. Our results have shown that the three PLA2s which contain either of aspartic acid, serine, or lysine residues at position 49 in the catalytic center, have different effects on the vesicles. The PLA2 with aspartic acid at this position causes the oval deformation of the vesicles, while serine and lysine-containing enzymes lead to an appreciable increase of fluorescence intensity in the vesicles membrane, wherein the shape and dimensions of GUVs have not changed, but in this case GUV aggregation occurs. LAURDAN and PRODAN detect the extent of water penetration into the bilayer surface. We calculated generalized polarization function (GP), showing that for all cases (D49 PLA2, S49 PLA2 and K49 PLA2) both LAUDRAN and PRODAN GP values decrease. A higher LAURDAN GP is indicative of low water penetration in the lipid bilayer in case of K49 PLA2 compared with D49 PLA2, whereas the PRODAN mainly gives information when lipid is in liquid crystalline phase.  相似文献   

12.
The impacts of pH, salt concentration (expressed as Debye length), and composition on the phase behavior of hybrid block copolymer-lipid-cholesterol bilayers incorporating carboxyl-terminated poly(butadiene)-block-poly(ethylene oxide) copolymer (PBdPEO1800(?)) or/and non-carboxyl-terminated PBdPEO (PBdPEO1800 or/and PBdPEO950), egg sphingomyelin (egg SM), and cholesterol were examined using fluorescence spectroscopy of laurdan. Laurdan emission spectra were decomposed into three lognormal curves as functions of energy. The ratio of the area of the mid-energy peak to the sum of the areas of all three peaks was evaluated as vesicles were cooled, yielding temperature breakpoint values (Tbreak) expected to be within the range of the phase transition temperature. Tbreak values displayed dependence on pH, Debye length, and vesicle composition consistent with an electrostatic repulsion contribution to vesicle phase behavior. Increased pH and Debye length, for which a greater dissociated fraction of PBdPEO1800(?) and a greater energy of electrostatic repulsion would be expected, resulted in Tbreak values as much as 10 °C less than at low pH or short Debye lengths. Additionally, at Debye lengths comparable to those at physiologically relevant ionic strength, Tbreak at pH 5.9 was observed to be slightly higher than at pH 7.0 for vesicles containing 50 mol% PBdPEO1800(?). Electrostatic effects observed for hybrid vesicles incorporating significant amounts of carboxyl-terminated polymer may have the ability to drive phase separation in response to pH drops—such as those observed after endocytosis—in physiologically relevant conditions, suggesting the utility of such materials for drug delivery.  相似文献   

13.
Glycophorin from human erythrocytes has been incorporated into liposomes of dimyristoylphosphatidylcholine (DMPC), dipalmitoylphosphatidylcholine (DPPC) and distearoylphosphatidylcholine (DSPC). The thermal properties of unsonicated liposomes with glycophorin/lipid molar ratios up to 4·10?3 have been studied by differential scanning calorimetry and the numbers of lipids withdrawn from participation in the gel-to-lamellar phase transition were found to be 42±22 (DMPC), 197±28 (DPPC) and 240±64 (DSPC). The initial rates of agglutination of sonicated liposomes with glycophorin/lipid molar ratios up to 4·10?3 by wheat germ agglutinin in the concentration range 0–7 μM have been measured over a range of temperature. Below the gel-to-lamellar phase transition (Tc) the rates of agglutination increase with acyl chain length in the sequence DMPC < DPPC < DSPC. Agglutination is found to be second order in liposome concentration and is completely reversed on saturation of the wheat germ agglutinin-binding sites by N-acetylglucosamine. Agglutination rates decrease with increasing temperature below Tc and are largely independent of temperature above Tc. The results are discussed in relation to the clustering of glycophorin in the phospholipid bilayers and its effect on binding and subsequent interliposomal bridge formation by wheat germ agglutinin.  相似文献   

14.
The phase heterogeneity of giant unilamellar dinervonoylphosphocholine (DNPC) vesicles in the course of the main phase transition was investigated by confocal fluorescence microscopy observing the fluorescence from the membrane incorporated lipid analog, 1-palmitoyl-2-(N-4-nitrobenz-2-oxa-1,3-diazol)aminocaproyl-sn-glycero-3-phosphocholine (NBDPC). These data were supplemented by differential scanning calorimetry (DSC) of DNPC large unilamellar vesicles (LUV, diameter ∼0.1 and 0.2 μm) and multilamellar vesicles (MLV). The present data collected upon cooling reveal a lack of micron-scale gel and fluid phase coexistence in DNPC GUVs above the temperature of 20.5 °C, this temperature corresponding closely to the heat capacity maxima (Tem) of DNPC MLVs and LUVs (Tem ≈21 °C), measured upon DSC cooling scans. This is in keeping with the model for phospholipid main transition inferred from our previous fluorescence spectroscopy data for DMPC, DPPC, and DNPC LUVs. More specifically, the current experiments provide further support for the phospholipid main transition involving a first-order process, with the characteristic two-phase coexistence converting into an intermediate phase in the proximity of Tem. This at least macroscopically homogenous intermediate phase would then transform into the liquid crystalline state by a second-order process, with further increase in acyl chain transgauche isomerization.  相似文献   

15.
Cationic, triple-chain amphiphiles promote vesicle fusion more than structurally related double-chain or single-chain analogues. Two types of vesicle fusion experiments were conducted, mixing of oppositely charged vesicles and acid-triggered self-fusion of vesicles composed of cationic amphiphile and anionic cholesteryl hemisuccinate (CHEMS). Vesicle fusion was monitored by standard fluorescence assays for intermembrane lipid mixing, aqueous contents mixing and leakage. Differential scanning calorimetry was used to show that triple-chain amphiphiles lower the lamellar-inverse hexagonal (Lα-HII) phase transition temperature for dipalmitoleoylphosphatidylethanolamine. The triple-chain amphiphiles may enhance vesicle fusion because they can stabilize the inversely curved membrane surfaces of the fusion intermediates, however, other factors such as extended conformation, packing defects, chain motion, or surface dehydration may also contribute. From the perspective of drug delivery, the results suggest that vesicles containing cationic, triple-chain amphiphiles (and cationic, cone-shaped amphiphiles in general) may be effective as fusogenic delivery capsules.  相似文献   

16.
Myosin light chain kinase binding to plastic   总被引:3,自引:0,他引:3  
R B Pearson  C House  B E Kemp 《FEBS letters》1982,145(2):327-331
Methionine-81 and/or -8 of the transmembrane sialoglycoprotein, glycophorin A, have been specifically alkylated with 13CH3I to produce the sulfonium ion derivatives [S-[13C]methylmethionine-8]glycophorin A and [S-[13C]methylmethionine-8 and -81]glycophorin A. 13C NMR spectra of these species show that the resonances of the methyl groups of the modified glycophorins occur at 26.1 ppm downfield from Me4Si. A spin-lattice relaxation time of 0.4 was observed for the 13C-enriched methyl resonances of the sulfonium ion derivatives of Met-8 and -81, which corresponds to an effective correlation time of < 2× 10?10 s. Demethylation of the 2 glycophorin A sulfonium ion species with 2-mercaptoethanol produces native glycophorin A which now has the ε-carbon of the methionine residue(s) 45% isotopically enriched. The ε-carbon of Met-8 was found to occur at 15.7 ppm downfield from Me4Si whereas the ε-carbon of Met-81 exhibited an unusual chemical shift of 2.0 ppm downfield from Me4Si. The spin-lattice relaxation time of both resonances was found to be ~0.3 s.  相似文献   

17.
Beef heart mitochondrial bc1 complex (ubiquinone—cytochrome c oxidoreductase) has been assayed by its ability to catalyse the reaction of duroquinol reduction of ferricytochrome c. When the isolated complex is reincorporated into lipid vesicles, the enzyme-catalysed electron transfer rate becomes uncoupler-sensitive. Initial experiments suggest that a protonmotive force is generated across the vesicles when electron transfer is initiated. Both ΔΨ and ΔpH components of this protonmotive force then influence an internal rate constant of the bc1 complex.  相似文献   

18.
Reaction progress curves for the hydrolysis of dimyristoylphosphatidylcholine by pig pancreatic phospholipase A2 exhibits a latency phase. Addition of 1-palmitoyllysophosphatidylcholine to the preformed vesicles reduces the latency phase and enhances the binding of phospholipase A2 to the vesicles. In contrast, the binary codispersions prepared from diacylphospholipids premixed with lysophosphatidylcholine do not exhibit such enhanced susceptibility to the phospholipase. This effect appears to be due to organizational defects created by asymmetrical incorporation of lysophospholipid molecules into the outer monolayer of the vesicles, and the action of phospholipase is not observed when the additive is equilibrated in both the monolayers of the vesicles.  相似文献   

19.
Composition and phase dependence of the mixing of 1,2-Dipalmitoyl-sn-glycero-3-phosphocholine (DPPC), and 1,2-Dioleoyl-sn-glycero-3-phosphocholine (DOPC), with the oxidized phospholipid, 1-palmitoyl-2-glutaryl-sn-glycero-3-phosphocholine (PGPC) were investigated by characterizing the aggregation states of DPPC/PGPC and DOPC/PGPC using a fluorescence quenching assay, dynamic light scattering, and time-resolved fluorescence quenching in the temperature range 5–60 °C. PGPC forms 3.5 nm radii micelles of aggregation number 33. In the gel phase, DPPC and PGPC fuse to form mixed vesicles for PGPC molar fraction, XPGPC  0.3 and coexisting vesicles and micelles at higher XPGPC. Data suggest that liquid phase DPPC at 50 °C forms mixed vesicles with segregated or hemi fused DPPC and PGPC for XPGPC  0.3. At 60 °C, DPPC and PGPC do not mix, but form coexisting vesicles and micelles. DOPC and PGPC do not mix in any proportion in the liquid phase. Two dissimilar aggregates of the sizes of vesicles and PGPC micelles were observed for all XPGPC for T  22 °C. DOPC–PGPC and DPPC–PGPC mixing is non-ideal for XPGPC > 0.3 in both gel and fluid phases resulting in exclusion of PGPC from the bilayer. Formation of mixed vesicles is favored in the gel phase but not in the liquid phase for XPGPC  0.3. For XPGPC  0.3, aggregation states change progressively from mixed vesicles in the gel phase to component segregated mixed vesicles in the liquid phase close to the chain melting transition temperature to separated coexisting vesicles and micelles at higher temperatures.  相似文献   

20.
Blood coagulation factors X and V, as well as prothrombin fragment 1 caused changes in the observed transition temperature (Tm) of appropriately constituted phospholipid vesicles upon binding to the membrane surface. Factor X- and prothrombin fragment 1-induced Tm shifts were calcium-dependent, while factor V changed the Tm in a calcium-independent manner. The results were consistent with clustering of the acidic phospholipid molecules due to protein binding. In all cases, protein binding to acidic phospholipid-containing vesicles caused the observed Tm to approach that for the neutral phospholipid. This resulted in a Tm increase for phospholipid mixtures containing bovine brain phosphatidylserine (PS) plus dipalmitoylphosphatidylcholine (DPPC) and a Tm decrease for mixtures of dipalmitoylphosphatidic acid (DPPA) and dimyristoylphosphatidylcholine (DMPC). Maximum Tm shifts induced in PS-DPPC (10:90) vesicles were very similar for all the prothrombinase proteins and the extent of the change was proportional to the actual amount of membrane-bound protein as determined by light-scattering techniques. For the vitamin K-dependent proteins, Tm changes were greater in the presence of protein plus calcium than in the presence of calcium alone, indicating that lateral phase separation occurs subsequent to initial protein-membrane contact. Lateral phase separation of acidic phospholipids appears to be an important process in the formation of the prothrombinase complex.  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