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.
When mouse myeloid leukemia M1 cells were induced to differentiate into macrophages by bacterial lipopolysaccharide (LPS), phospholipids and gangliosides of the cells changed markedly. The amounts of phosphatidylcholine, phosphatidylethanolamine and phosphatidylinositol per mg protein increased 30%, 20% and 30%, respectively, during differentiation, but the others, phosphatidylserine and sphingomyelin, did not increase significantly. Three species of gangliosides constituted of major portions of gangliosides in M1 cells. Several-fold increase in monosialoganglioside GM1 was observed in the LPS-treated cells with a concomitant decrease in disialogangliosides. Based upon the treatment with sialidase, this GM1 was identified to be GM1b, which was recently found in rat ascites hepatoma cells and human erythrocyte membranes.  相似文献   

2.
The interaction of myelin basic protein with monosialoganglioside GM1 was investigated. It was found that the emission maximum of the tryptophan of the protein is blue-shifted due to the interaction. In mixtures of the monosialoganglioside with phosphatidylcholine, the myelin basic protein induces phase separation of the lipids as inferred from differential scanning calorimetry experiments.  相似文献   

3.
K Hong  V D Vacquier 《Biochemistry》1986,25(3):543-549
Lysin, a protein of Mr 16 000 from the acrosome granule of the abalone, is responsible for the dissolution of the egg vitelline layer. The primary structure of this cationic protein projects some hydrophobic domains in the secondary structure. Lysin was found to associate nonselectively with phospholipid bilayers and cause a spontaneous release of encapsulated carboxyfluorescein in liposomes. The association of lysin with phosphatidylcholine liposomes suggests that there is a hydrophobic interaction between lysin and lipid bilayers. Binding of lysin to phospholipid resulted in the aggregation of phosphatidylserine-containing liposomes, but aggregation was not observed in neutral phosphatidylcholine liposomes. Resonance energy transfer and dequenching of fluorescent 1-palmitoyl-2-cis-parinaroylphosphatidylcholine were both used to determine the fusogenic activity of lysin in aggregated liposomes. Results from both assays are consistent. Lysin-induced fusion was observed in all the phosphatidylserine-containing liposomes, and the general trend of fusion susceptibility was phosphatidylserine/phosphatidylcholine (1:2) approximately equal to phosphatidylserine/phosphatidylcholine/phosphatidylethanolamine (1:1:1) greater than phosphatidylserine/phosphatidylethanolamine (1:2). Cholesterol up to 30% did not affect the intrinsic fusion susceptibility. A hydrophobic penetration by protein molecules and the packing of phospholipid bilayers are used to interpret the fusion susceptibility. Lysin-induced liposome aggregation was highly independent of the state of self-association of lysin in ionic medium. However, the fusogenic activity of self-associated lysin was found to be much less than the monodispersed one. Liposomes preincubated with Ca2+ did not fuse initially as readily as those without Ca2+ treatment.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Hydrolysis of lipid mixtures by rat hepatic lipase   总被引:1,自引:0,他引:1  
The hydrolysis of phospholipid mixtures by purified rat hepatic lipase, also known as hepatic triglyceride lipase, was studied in a Triton X-100/lipid mixed micellar system. Column chromatography of the mixed micelles showed elution of Triton X-100 and binary lipid mixtures of phosphatidic acid, phosphatidylcholine, phosphatidylethanolamine and phosphatidylserine as a single peak. This indicated that the mixed micelles were homogenous and contained all components in the designated molar ratios. The molar ratio of Triton X-100 to lipid was kept constant at 4 to 1. Labeling one lipid with 3H and the other lipid with 14C enabled us to determine the hydrolysis of both components of these binary lipid mixed micelles. We found that the hydrolysis of phosphatidylcholine was activated by the inclusion of small amounts of phosphatidic acid (2.5-fold), phosphatidylethanolamine (1.5-fold) or phosphatidylserine (1.4-fold). The maximal activation of phosphatidylcholine hydrolysis was observed when 5 mol% of phosphatidylethanolamine, 7.5 mol% phosphatidic acid or 5 mol% phosphatidylserine was added to Triton X-100 mixed micelles. The hydrolysis of phosphatidic acid was activated 30%, and that of phosphatidylserine was inhibited 30% when the molar proportion of phosphatidylcholine was less than 50 mol%. The hydrolysis of phosphatidylethanolamine was slightly activated when the mol% of phosphatidylcholine was below 5. The hydrolysis of phosphatidylserine was inhibited by phosphatidylethanolamine when the mol% of the latter was 50 or less whereas phosphatidylethanolamine hydrolysis was not affected by phosphatidylserine. Under the conditions used sphingomyelin and cholesterol did not have a significant effect on the hydrolysis of the phospholipids studied. In agreement with our previous study (Kucera et al. (1988) J. Biol. Chem. 263, 1920-1928) these studies show that the phospholipid polar head group is an important factor which influences the action of hepatic lipase and that the interfacial properties of the substrate play a role in the expression of the activity of this enzyme. The molar ratios of phosphatidic acid, phosphatidylethanolamine and phosphatidylserine which activated phosphatidylcholine hydrolysis correspond closely to the molar ratios of these lipids found in the surface lipid film of lipoproteins e.g., high density lipoproteins.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

5.
On the substrate specificity of rat liver phospholipase A1   总被引:1,自引:0,他引:1  
The substrate specificity of purified phospholipase A1 was studied using mixed micelles of phospholipid and Triton X-100. The kinetic analysis employed determined Vmax, Ks (a dissociation constant for the phospholipase A1-mixed micelle complex), and Km (the Michaelis constant for the catalytic step which reflects the binding of the enzyme to the substrate in the interface). The order of Vmax values was phosphatidic acid greater than phosphatidylethanolamine greater than phosphatidylcholine greater than phosphatidylserine. The order of Ks values was phosphatidylcholine greater than phosphatidylethanolamine greater than phosphatidic acid greater than phosphatidylserine; the order of Km values was phosphatidic acid greater than phosphatidylethanolamine = phosphatidylserine greater than phosphatidylcholine. When present together, phosphatidylcholine inhibited the hydrolysis of phosphatidylethanolamine but phosphatidylethanolamine did not affect the hydrolysis of phosphatidylcholine. Sphingomyelin, phosphatidylcholine plasmalogen, and phosphatidylethanolamine plasmalogen had no effect on the hydrolysis of phosphatidylethanolamine. The effects of the reaction products, lysolipids and/or fatty acids, were also considered for their influence on phosphatidylethanolamine hydrolysis catalyzed by phospholipase A1. Free fatty acid was found to inhibit, whereas lysophospholipids stimulated hydrolysis of phosphatidylethanolamine. In a mixture of 1,2- and 1,3-diacylglycerides in mixed micelles, only the acyl chain at the sn-1 position of the 1,2 compound was hydrolyzed. Surface charge did not modulate the hydrolysis of phosphatidylcholine vesicles or mixed micelles. In conclusion, it is hypothesized that steric hindrance at position 3 of the glycerol regulates substrate binding in the active site and that an acyl group in position 1 is favored over a vinyl ether linkage for binding.  相似文献   

6.
Ca2+-induced phase separation in phosphatidylserine/phosphatidylethanolamine and phosphatidylserine/phosphatidylethanolamine/phosphatidylcholine model membranes was studied using spin-labeled phosphatidylethanolamine and phosphatidylcholine and compared with that in phosphatidylserine/phosphatidylcholine model membranes studied previously. The phosphatidylethanolamine-containing membranes behaved in qualitatively the same way as did phosphatidylserine/phosphatidylcholine model membranes. There were some quantitative differences between them. The degree of phase separation was higher in the phosphatidylethanolamine-containing membranes. For example, the degree of phase separation in phosphatidylserine/phosphatidylethanolamine membranes containing various mole fractions of phosphatidylserine was 94--100% at 23 degrees C and 84--88% at 40 degrees C, while the corresponding value for phosphatidylserine/phosphatidylcholine membranes was 74--85% at 23 degrees C and 61--79% at 40 degrees C. Ca2+ concentration required for the phase separation was lower for phosphatidylserine/phosphatidylethanolamine than that for phosphatidylserine/phosphatidylcholine membranes; concentration to cause a half-maximal phase separation was 1.4 . 10(-7) M for phosphatidylserine-phosphatidylethanolamine and 1.2 . 10(-6) M for phosphatidylserine/phosphatidylcholine membranes. The phase diagram of phosphatidylserine/phosphatidylethanolamine membranes in the presence of Ca2+ was also qualitatively the same as that of phosphatidylserine/phosphatidylcholine except for the different phase transition temperatures of phosphatidylethanolamine (17 degrees C) and phosphatidylcholine (-15 degrees C). These differences were explained in terms of a greater tendency for phosphatidylethanolamine, compared to phosphatidylcholine, to form its own fluid phase separated from the Ca2+-chelated solid-phase phosphatidylserine domain.  相似文献   

7.
The effect of phospholipid suspensions on the proteolysis of isolated spectrin was examined by SDS-polyacrylamide gradient gel electrophoresis. Proteolysis of spectrin in the membranes by trypsin and pronase was also studied. It was found that electrophoretic patterns of spectrin fragments were influenced by the presence of the suspension prepared from phosphatidylethanolamine:phosphatidylserine (60:40) mixture and of phosphatidylcholine. Qualitative changes in the proteolytic patterns obtained after proteolysis of spectrin by pronase in the presence of phosphatidylcholine suspension were observed. The changes in the sensitivity of spectrin towards proteases result probably from changes in the accessibility of some peptide bonds upon the interaction of this extrinsic protein with phospholipids.  相似文献   

8.
Phospholipids and regulation of protein kinase reaction   总被引:1,自引:0,他引:1  
Naturally occurring phospholipids such as phosphatidylinositol and phosphatidylserine inhibited cAMP-dependent protein kinase by interacting with the substrate protein (phosphate acceptor). This inhibition was observed both in the presence and absence of cAMP when histone H2B, protamine, and myelin basic protein were used, but was not detected when casein was used as the substrate. Other phospholipids such as phosphatidylethanolamine and phosphatidylcholine did not inhibit the kinase but did stimulate the kinase when protamine served as the substrate. Both cAMP-dependent and cGMP-dependent protein kinases were inhibited by phosphatidylserine when histone H2B was used as substrate. The substrate protein binding to phosphatidylinositol and phosphatidylserine was observed when these phospholipids were added to the incubation mixture, suggesting that direct interaction between the substrate protein and the phospholipids resulted in inhibition of cAMP and cGMP-dependent protein kinase. Thus the substrate protein for protein kinase probably plays an important role in regulating the kinase activity related to various phospholipids.  相似文献   

9.
The role of methionine residues in the interaction of the phosphatidylcholine transfer protein from bovine liver with phospholipid vesicles was investigated by specific modification of these residues with iodoacetamide. The modified protein was digested with cyanogen bromide in order to determine which methionine residues had become resistant to this cleavage. Automated Edman degradation on the digest indicated that after 72 h of reaction, Met-1 was modified for 80%, Met-73 for 50%, Met-109 for 20%, whilst Met-173 and Met-203 were found to be unmodified. This distinct modification did not result in any loss of phosphatidylcholine transfer activity. The interaction of the phosphatidylcholine transfer protein with phospholipid vesicles was investigated by making use of electron spin resonance spectroscopy. The interaction of unmodified protein with vesicles composed of phosphatidylcholine/phosphatidic acid/spin-labeled phosphatidylethanolamine (79:16:5, mol%) or composed of phosphatidylserine/spin-labeled phosphatidylethanolamine (95:5, mol%), gave an increase of about 50% in the rotation correlation time. A similar increase was observed with the modified protein. This interaction was further investigated by labeling Met-1 and Met-73 in the transfer protein with iodoacetamidoproxyl spin-label. Spin-labeling did not inactivate the transfer protein. In addition, the electron spin resonance spectra of the spin-labeled protein were not affected upon addition of vesicles composed of phosphatidylcholine/phosphatidic acid (80:20, mol%). These experiments strongly suggest that Met-1 and Met-73 are not part of the site that interacts with the membrane.  相似文献   

10.
The phospholipid composition and phospholipid fatty acid composition of purified Rickettsia prowazeki were determined. The lipid phosphorous content was 6.8 +/- 1.3 microgram/mg of total rickettsial protein. The major phospholipid was phosphatidylethanolamine (60 to 70%); phosphatidylglycerol constituted 20%, and phosphatidylcholine constituted 15%. Small amounts of phosphatidylserine and cardiolipin were detected. The principal fatty acids were 18:1, 16:1, and 16:0. The fatty acid composition of the phosphatidylcholine in the rickettsial extracts was very different than that of the other rickettsial phosphatides and very similar to that of normal yolk sac phosphatidylcholine. The specific of the phosphatidylcholine of rickettsiae grown in the presence of 32P was markedly lower than that of phosphatidylethanolamine and phosphatidylglycerol. It is suggested that the phosphatidylcholine in the rickettsial extract is yolk sac derived and either tightly absorbed or exchanged into the rickettsial membrane.  相似文献   

11.
Synapsin I, a major neuron-specific phosphoprotein, is localized on the cytoplasmic surface of small synaptic vesicles to which it binds with high affinity. It contains a collagenase-resistant head domain and a collagenase-sensitive elongated tail domain. In the present study, the interaction between synapsin I and phospholipid vesicles has been characterized, and the protein domains involved in these interactions have been identified. When lipid vesicles were prepared from cholesterol and phospholipids using a lipid composition similar to that found in native synaptic vesicle membranes (40% phosphatidylcholine, 32% phosphatidylethanolamine, 12% phosphatidylserine, 5% phosphatidylinositol, 10% cholesterol, wt/wt), synapsin I bound with a dissociation constant of 14 nM and a maximal binding capacity of about 160 fmol of synapsin I/microgram of phospholipid. Increasing the ionic strength decreased the affinity without greatly affecting the maximal amount of synapsin I bound. When vesicles containing cholesterol and either phosphatidylcholine or phosphatidylcholine/phosphatidylethanolamine were tested, no significant binding was detected under any conditions examined. On the other hand, phosphatidylcholine vesicles containing either phosphatidylserine or phosphatidylinositol strongly interacted with synapsin I. The amount of synapsin I maximally bound was directly proportional to the percentage of acidic phospholipids present in the lipid bilayer, whereas the Kd value was not affected by varying the phospholipid composition. A study of synapsin I fragments obtained by cysteine-specific cleavage showed that the collagenase-resistant head domain actively bound to phospholipid vesicles; in contrast, the collagenase-sensitive tail domain, though strongly basic, did not significantly interact. Photolabeling of synapsin I was performed with the phosphatidylcholine analogue 1-palmitoyl-2-[11-[4-[3-(trifluoromethyl)diazirinyl]phenyl] [2-3H]undecanoyl]-sn-glycero-3-phosphocholine; this compound generates a highly reactive carbene that selectively interacts with membrane-embedded domains of membrane proteins. Synapsin I was significantly labeled upon photolysis when incubated with lipid vesicles containing acidic phospholipids and trace amounts of the photoactivatable phospholipid. Proteolytic cleavage of photolabeled synapsin I localized the label to the head domain of the molecule.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

12.
Membrane proteins interact with phospholipids either via an annular layer surrounding the transmembrane segments or by specific lipid-protein interactions. Although specifically bound phospholipids are observed in many crystal structures of membrane proteins, their roles are not well understood. Na,K-ATPase is highly dependent on acid phospholipids, especially phosphatidylserine, and previous work on purified detergent-soluble recombinant Na,K-ATPase showed that phosphatidylserine stabilizes and specifically interacts with the protein. Most recently the phosphatidylserine binding site has been located between transmembrane segments of αTM8–10 and the FXYD protein. This paper describes stimulation of Na,K-ATPase activity of the purified human α1β1 or α1β1FXYD1 complexes by neutral phospholipids, phosphatidylcholine, or phosphatidylethanolamine. In the presence of phosphatidylserine, soy phosphatidylcholine increases the Na,K-ATPase turnover rate from 5483 ± 144 to 7552 ± 105 (p < 0.0001). Analysis of α1β1FXYD1 complexes prepared with native or synthetic phospholipids shows that the stimulatory effect is structurally selective for neutral phospholipids with polyunsaturated fatty acyl chains, especially dilinoleoyl phosphatidylcholine or phosphatidylethanolamine. By contrast to phosphatidylserine, phosphatidylcholine or phosphatidylethanolamine destabilizes the Na,K-ATPase. Structural selectivity for stimulation of Na,K-ATPase activity and destabilization by neutral phospholipids distinguish these effects from the stabilizing effects of phosphatidylserine and imply that the phospholipids bind at distinct sites. A re-examination of electron densities of shark Na,K-ATPase is consistent with two bound phospholipids located between transmembrane segments αTM8–10 and TMFXYD (site A) and between TM2, -4, -6, -and 9 (site B). Comparison of the phospholipid binding pockets in E2 and E1 conformations suggests a possible mechanism of stimulation of Na,K-ATPase activity by the neutral phospholipid.  相似文献   

13.
When human erythrocytes are incubated with spin-labeled analogues of sphingomyelin, phosphatidylcholine, phosphatidylserine, or phosphatidylethanolamine, with a short beta chain (C5) bearing a doxyl group at the fourth carbon position, the labeled lipids incorporate readily in the outer monolayer. The incorporation is followed in fresh erythrocytes by a selective inward diffusion of the amino derivatives. This observation led us to postulate the existence of a selective ATP-dependent system that would flip aminophospholipids from the outer to the inner monolayer [Seigneuret, M., & Devaux, P. F. (1984) Proc. Natl. Acad. Sci. U.S.A. 81, 3751-3755]. This study further examines the nature of this selective transport and demonstrates that it is mediated by a specific membrane protein. By measurement of the initial rate of transverse diffusion of spin-labeled lipids incorporated at various concentrations in the membrane outer leaflet of packed erythrocytes, apparent Km values were determined for the phosphatidylserine and phosphatidylethanolamine analogues. A ratio of approximately equal to 1/9.4 [corrected] was obtained (KmPS/KmPE). Using spin-labels bearing either a 14N or a 15N isotope, we have carried out competition experiments allowing us to measure simultaneously the transport of two different phospholipids. By this procedure, we show that phosphatidylserine and phosphatidylethanolamine compete for the same transport site but that phosphatidylserine has a higher affinity, in agreement with a lower apparent Km. On the other hand, the slow diffusion of the phosphatidylcholine or sphingomyelin analogues has no influence on the transport of phosphatidylserine or phosphatidylethanolamine. Experiments carried out in ghosts loaded with ATP enabled us to determine the activation energies for phosphatidylserine and phosphatidylcholine transverse diffusion.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Ca2+-induced phase separation in phosphatidylserine/phosphatidylethanolamine and phosphatidylserine/phosphatidylethanolamine/phosphatidylcholine model membranes was studied using spin-labeled phosphatidylethanolamine and phosphatidylcholine and compared with that in phosphatidylserine/phosphatidylcholine model membranes studied previously. The phosphatidyl-ethanolamine-containing membranes behaved in qualitatively the same way as did phosphatidylserine/phosphatidylcholine model membranes. There were some quantitative differences between them. The degree of phase separation was higher in the phosphatidylethanolamine-containing membranes. For example, the degree of phase separation in phosphatidylserine/phosphatidylethanolamine membranes containing various mole fractions of phosphatidylserine was 94–100% at 23°C and 84–88% at 40°C, while the corresponding value for phosphatidylserine/phosphatidylcholine membranes was 74–85% at 23°C and 61–79% at 40°C. Ca2+ concentration required for the phase separation was lower for phosphatidylserine/phosphatidylethanolamine than that for phosphatidylserine/phosphatidylcholine membranes; concentration to cause a half-maximal phase separation was 1.4 · 10?7 M for phosphatidylserine-phosphatidylethanolamine and 1.2 · 10?6 M for phosphatidylserine/phosphatidylcholine membranes. The phase diagram of phosphatidylserine/phosphatidylethanolamine membranes in the presence of Ca2+ was also qualitatively the same as that of phosphatidylserine/phosphatidylcholine except for the different phase transition temperatures of phosphatidylethanolamine (17°C) and phosphatidylcholine (?15°C). These differences were explained in terms of a greater tendency for phosphatidylethanolamine, compared to phosphatidylcholine, to form its own fluid phase separated from the Ca2+-chelated solid-phase phosphatidylserine domain.  相似文献   

15.
The effects of low concentrations of cholesterol in mixtures of a negatively charged phospholipid (phosphatidylserine or phosphatidylglycerol) and another phospholipid (phosphatidylcholine, sphingomyelin or phosphatidylethanolamine) have been studied by differential scanning calorimetry. Only mixtures which showed a gel phase miscibility gap have been employed. It was demonstrated that in mixtures with phosphatidylethanolamine, cholesterol was preferentially associated with the negatively charged phospholipid, regardless whether this species represented the component with the high or with the low transition temperature in the mixture. In mixtures of a negatively charged phospholipid and phosphatidylcholine, cholesterol associated with the negatively charged phospholipid; when the phosphatidylcholine was the species with the low transition temperature, cholesterol had an affinity for the phosphatidylcholine and for the negatively charged phospholipid as well. Cholesterol, in a mixture of sphingomyelin with a high and phosphatidylserine with a low transition temperature, was preferentially associated with sphingomyelin.From these experiments it is concluded that phospholipids show a decrease in affinity for cholesterol in the following order: sphingomyelin ? phosphatidylserine, phosphatidylglycerol > phosphatidylcholine ? phosphatidylethanolamine.  相似文献   

16.
The effect of associating acidic and basic polypeptides with dilute suspensions of vesicles composed of various unsaturated phospholipids was assessed with regard to optical density and ultraviolet absorption. Associating basic polypeptides with phosphatidylserine, phosphatidylethanolamine, and phosphatidylglycerol vesicles, or acidic polypeptide with phosphatidylcholine vesicles, caused an increase in the optical density of the preparations, with no measurable effect on the intensity of the ultraviolet spectrum of the olefinic bonds of the lipid. Associating basic polypeptides with phosphatidylcholine vesicles, in addition to causing similar increases in optical density, resulted in a large decrease in the intensity of ultraviolet absorption by the olefinic bonds. This implies that the interaction between the basic polypeptides and phosphatidylcholine vesicles results in major alterations in the microenvironment of the olefinic bonds, which would require intimate association of the polypeptide with the ninth carbon region of the acyl chains. These observations support the conclusion, drawn from our earlier studies, that the association of basic polypeptides and liquid crystalline phase phosphatidylcholine vesicles is peculiar in that it involves a major hydrophobic component.  相似文献   

17.
Effect of membrane phospholipids on the activity of cytosolic protein-tyrosine kinase from porcine spleen (CPTK-40) has been studied. Using poly(Glu Na, Tyr)4:1 as a substrate, phosphatidylethanolamine, phosphatidylcholine and phosphatidylserine had stimulatory effects on that phosphorylation activity, however phosphatidic acid had inhibitory and phosphatidylinositol had no effects. Similar results were obtained using[Val5]angiotensin II as a substrate. On the other hand using basic protein (H2B histone and myelin basic protein) as substrates, phosphatidic acid stimulated the activity of CPTK-40, while phosphatidylinositol inhibited the activity. Phosphatidylethanolamine, phosphatidylcholine and phosphatidylserine caused different effect on the activity of CPTK-40 depending on the substrate employed. However using acidic protein (tubulin and casein) as substrates, the activity of CPTK-40 was neither stimulated nor inhibited by any phospholipids. These results suggest that phospholipids may modulate the activity of CPTK-40.  相似文献   

18.
Exosomes are small vesicles released from cells into extracellular space. We have isolated exosomes from neuroblastoma cells and investigated their influence on the aggregation of α-synuclein, a protein associated with Parkinson disease pathology. Using cryo-transmission electron microscopy of exosomes, we found spherical unilamellar vesicles with a significant protein content, and Western blot analysis revealed that they contain, as expected, the proteins Flotillin-1 and Alix. Using thioflavin T fluorescence to monitor aggregation kinetics, we found that exosomes catalyze the process in a similar manner as a low concentration of preformed α-synuclein fibrils. The exosomes reduce the lag time indicating that they provide catalytic environments for nucleation. The catalytic effects of exosomes derived from naive cells and cells that overexpress α-synuclein do not differ. Vesicles prepared from extracted exosome lipids accelerate aggregation, suggesting that the lipids in exosomes are sufficient for the catalytic effect to arise. Using mass spectrometry, we found several phospholipid classes in the exosomes, including phosphatidylcholine, phosphatidylserine, phosphatidylethanolamine, phosphatidylinositol, and the gangliosides GM2 and GM3. Within each class, several species with different acyl chains were identified. We then prepared vesicles from corresponding pure lipids or defined mixtures, most of which were found to retard α-synuclein aggregation. As a striking exception, vesicles containing ganglioside lipids GM1 or GM3 accelerate the process. Understanding how α-synuclein interacts with biological membranes to promote neurological disease might lead to the identification of novel therapeutic targets.  相似文献   

19.
To identify the specific component(s) in the target membrane involved in fusion of vesicular stomatitis virus (VSV), we examined the interaction of the virus with human erythrocyte membranes with asymmetric and symmetric bilayer distributions of phospholipids. Fusion was monitored spectrofluorometrically by the octadecylrhodamine dequenching assay. Fusion of VSV with lipid-symmetric erythrocyte ghosts was rapid at 37 degrees C and low pH, whereas little or no fusion was observed with lipid-asymmetric ghosts. Conversion of phosphatidylserine in the lipid-symmetric ghost membrane to phosphatidylethanolamine by means of the enzyme phosphatidylserine decarboxylase did not alter the target membrane's susceptibility to VSV fusion. Spin-labeled phospholipid analogues with phosphatidylserine, phosphatidylethanolamine, and phosphatidylcholine headgroups incorporated into the outer leaflet of lipid-asymmetric erythrocytes did not render those membranes fusogenic. Electron spin resonance spectra showed an increased mobility of a phosphatidylcholine spin-label incorporated into the outer leaflet of lipid-symmetric erythrocyte ghosts as compared to that of lipid-asymmetric ghosts. These results indicate that the susceptibility to VSV fusion is not dependent on any particular phospholipid but rather is related to packing characteristics of the target membrane.  相似文献   

20.
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   

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