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 64 毫秒
1.
There is a marked hysteresis between the heating and cooling polymorphic phase transition of anhydrous cholesterol. At a scan rate of 0.05 degrees C/min the difference in transition temperatures between heating and cooling scans is approximately 10 degrees C. This phenomenon also occurs with mixtures of cholesterol with phosphatidylserine and can result in an underestimation of the amount of crystalline cholesterol in a sample that has not been cooled sufficiently. With 1-palmitoyl-2-oleoyl phosphatidylserine and 1-stearoyl-2-oleoyl phosphatidylserine the cholesterol crystallites form while the lipid remains in the L(alpha) phase. Sonication of dimyristoyl phosphatidylserine with a 0.4 mol fraction cholesterol results in the loss of cholesterol crystallite diffraction, but only a partial loss of the polymorphic transition detected by calorimetry. We therefore conclude that the thermal history of the sample can have profound effects on the appearance of the polymorphic phase transition of cholesterol by differential scanning calorimetry. Depending on the morphology of the vesicles, diffraction methods may underevaluate the amount of cholesterol crystallites present.  相似文献   

2.
Modulation by pardaxin of the phase transitions of dimyristoyl phosphatidylcholine, 1-stearoyl-2-oleoyl phosphatidylcholine or 1-stearoyl-2-oleoyl phosphatidylglycerol in the presence or absence of cholesterol was studied by differential scanning calorimetry. The transition enthalpy of each of the phospholipids was lowered by pardaxin and there was a small decrease in the transition temperature. Addition of cholesterol and pardaxin to dimyristoyl phosphatidylcholine resulted in a very marked lowering of the transition temperature. Although the peptide broadens the transition of the pure phospholipids, it sharpens the transition of mixtures of the phospholipids with cholesterol. This and the observation that pardaxin also causes the formation of crystallites of anhydrous cholesterol, suggest that the peptide promotes redistribution of cholesterol in the membrane.  相似文献   

3.
The behavior of cholesterol is different in mixtures with phosphatidylcholine as compared with phosphatidylserine. In (13)C cross polarization/magic angle spinning nuclear magnetic resonance spectra, resonance peaks of the vinylic carbons of cholesterol are a doublet in samples containing 0.3 or 0.5 mol fraction cholesterol with 1-palmitoyl-2-oleoyl phosphatidylserine (POPS) or in cholesterol monohydrate crystals, but a singlet with mixtures of cholesterol and 1-palmitoyl-2-oleoyl phosphatidylcholine (POPC). At these molar fractions of cholesterol with POPS, resonances of the C-18 of cholesterol appear at the same chemical shifts as in pure cholesterol monohydrate crystals. These resonances do not appear in samples of POPS with 0.2 mol fraction cholesterol or with POPC up to 0.5 mol fraction cholesterol. In addition, there is another resonance from the cholesterol C18 that appears in all of the mixtures of phospholipid and cholesterol but not in pure cholesterol monohydrate crystals. Using direct polarization, the fraction of cholesterol present as crystallites in POPS with 0.5 mol fraction cholesterol is found to be 80%, whereas with the same mol fraction of cholesterol and POPC none of the cholesterol is crystalline. After many hours of incubation, cholesterol monohydrate crystals in POPS undergo a change that results in an increase in the intensity of certain resonances of cholesterol monohydrate in (13)C cross polarization/magic angle spinning nuclear magnetic resonance, indicating a rigidification of the C and D rings of cholesterol but not other regions of the molecule.  相似文献   

4.
The effect of lipid composition of liposomes on peroxidation induced by ferrous ion and ascorbate was examined. Temperature affects the sensitivity of liposomes; the peroxidation rate was increased with increase of the incubation temperature. With liposomes consisting of 1-palmitoyl-2-arachidonyl phosphatidylcholine (substrate) and a peroxidation-insensitive lipid, 1-palmitoyl-2-oleoyl phosphatidylcholine, peroxidation was dependent on the density of the substrate. No appreciable peroxidation was observed with liposomes containing less than 10 mol% of the substrate at 37 degrees C. When 1 mol substrate was mixed with 9 mol dimyristoyl phosphatidylcholine, peroxidation occurred below 10 degrees C, but not above 20 degrees C. Above 20 degrees C, the substrates should be located homogeneously on the membranes, whereas they should be clustered below 10 degrees C, since the gel-liquid crystalline phase transition temperature of matrix membrane of dimyristoylphosphatidylcholine was 17-21 degrees C. Peroxidation of liposomes consisting of 1-palmitoyl-2-arachidonyl phosphatidylcholine was also suppressed by cholesterol. These findings indicate that the lateral distribution as well as the density of the substrate on membranes affects the sensitivity of the substrate to peroxidation. It was also found that alpha-tocopherol is preferentially located in the 1-palmitoyl-2-arachidonyl phosphatidylcholine-rich regions of membranes consisting of mixed phospholipids, and efficiently suppresses peroxidation of liposomal lipids.  相似文献   

5.
Model membranes composed of cholesterol plus one of two phosphatidylcholines (PC), each containing a saturated and a dienoic acyl chain, have been studied by differential scanning calorimetry. The gel to liquid-crystalline phase transition temperature of 1-palmitoyl-2-linoleoyl PC was -19.5 degrees C and that of 1-stearoyl-2-linoleoyl PC was -13.7 degrees C. The addition of cholesterol to the phosphatidylcholines in aqueous dispersion resulted in the progressive removal of the phase transition as observed by differential scanning calorimetry. Per mole of sterol in the membrane, cholesterol was more effective at reducing the enthalpy change of the phase transitions of these bilayers containing dienoic phosphatidylcholines than it is in eliminating the transition of membranes made with other phospholipids that contain more saturated chains. No transitions in membranes made with palmitoyl-linoleoyl PC or stearoyl-linoleoyl PC could be detected calorimetrically when 17 mol% cholesterol was present.  相似文献   

6.
The thermotropic phase behavior of cholesterol monohydrate in water was investigated by differential scanning calorimetry, polarizing light microscopy, and x-ray diffraction. In contrast to anhydrous cholesterol which undergoes a polymorphic crystalline transition at 39 degrees C and a crystalline to liquid transition at 151 degrees C, the closed system of cholesterol monohydrate and water exhibited three reversible endothermic transitions at 86, 123, and 157 degrees C. At 86 degrees C, cholesterol monohydrate loses its water of hydration, forming the high temperature polymorph of anhydrous cholesterol. At least 24 hours were required for re-hydration of cholesterol and the rate of hydration was dependent on the polymorphic crystalline form of anhydrous cholesterol. At 123 degrees C, anhydrous crystalline cholesterol in the presence of excess water undergoes a sharp transition to a birefringent liquid crystalline phase of smectic texture. The x-ray diffraction pattern obtained from this phase contained two sharp low-angle reflections at 37.4 and 18.7 A and a diffuse wide-angle reflection centered at 5.7 A, indicating a layered smectic type of liquid crystalline structure with each layer being two cholesterol molecules thick. The liquid crystalline phase is stable over the temperature range of 123 to 157 degrees C before melting to a liquid dispersed in water. The observation of a smectic liquid crystalline phase for hydrated cholesterol correlates with its high surface activity and helps to explain its ability to exist in high concentrations in biological membranes.  相似文献   

7.
In the present work, we demonstrate that phosphatidylcholine with (16:1)9 acyl chains undergoes polymorphic rearrangements in mixtures with 0.6-0.8 mol fraction cholesterol. Studies were performed using differential scanning calorimetry, X-ray diffraction, cryo-electron microscopy, 31P NMR static powder patterns and 13C MAS/NMR. Mixtures of phosphatidylcholine with (16:1)9 acyl chains and 0.6 mol fraction cholesterol, after being heated to 100 degrees C, can form an ordered array with periodicity 14 nm which may be indicative of a cubic phase. Our results indicate that the formation of highly curved bilayer structures, such as those required for membrane fusion, can occur in mixtures of cholesterol with certain phosphatidylcholines that do not form non-lamellar structures in the absence of cholesterol. We also determine the polymorphic behavior of mixtures of symmetric phosphatidylcholines with cholesterol. Species of phosphatidylcholine with (20:1)11, (22:1)13 or (24:1)15 acyl chains in mixtures with 0.6-0.8 mol fraction cholesterol undergo a transition to the hexagonal phase at temperatures 70-80 degrees C. This is not the case for phosphatidylcholine with (18:1)6 acyl chains which remains in the lamellar phase up to 100 degrees C when mixed with as much as 0.8 mol fraction cholesterol. Thus, the polymorphic behavior of mixtures of phosphatidylcholine and cholesterol is not uncommon and is dependent on the intrinsic curvature of the phospholipid. Crystals of cholesterol can be detected in mixtures of all of these phosphatidylcholines at sufficiently high cholesterol mole fraction. What is unusual about the formation of these crystals in several cases is that cholesterol crystals are present in the monohydrate form in preference to the anhydrous form. Furthermore, after heating to 100 degrees C and recooling, the cholesterol crystals are again observed to be in the monohydrate form, although pure cholesterol crystals require many hours to rehydrate after being heated to 100 degrees C. Both the nature of the acyl chain as well as the mole fraction cholesterol determine whether cholesterol crystals in mixtures with the phospholipids will be in the monohydrate or in the anhydrous form.  相似文献   

8.
The influence of protons or calcium ions on the miscibility of cholesterol in phosphatidylserine has been examined using differential scanning calorimetry and X-ray diffraction. At pH 2.6, where the carboxyl group of the serine moiety is protonated, two endothermic transitions are observed in cholesterol-phosphatidylserine mixtures. The midpoint of the first is at 35 degrees C in the absence of cholesterol and decreases to approx. 15 degrees C for molar fraction of cholesterol 0.5. The second transition is centered at approx. 44 degrees C, almost independent of cholesterol content. The two lower temperature phases are lamellar and the high temperature phase has hexagonal symmetry. Cholesterol is more miscible in protonated phosphatidylserine than in the sodium form: cholesterol crystals are detected at a molar ratio of phosphatidylserine to cholesterol of about 1.7:1 as compared to about 2.3:1 at neutral pH. In the presence of calcium ions (1.3 Ca2+ per phosphatidylserine), a lamellar phase is observed with layer spacing 53 A which is independent of temperature (25 degrees C-65 degrees C) and of cholesterol content. Calcium ions cause reduced cholesterol solubility: crystallites are detected already at a molar ratio of 4:1.  相似文献   

9.
Thermally induced transition between anhydrous and hydrated forms of highly crystalline beta-chitin was studied by differential thermal calorimetry (DSC) and X-ray diffraction. DSC of wet beta-chitin in a sealed pan gave two well-defined endothermic peaks at 85.2 and 104.7 degrees C on heating and one broad exothermic peak at between 60 and 0 degrees C on cooling. These peaks were highly reproducible and became more distinct after repeated heating-cooling cycles. The X-ray diffraction pattern of wet beta-chitin at elevated temperature showed corresponding changes in d-spacing between the sheets formed by stacking of chitin molecules. These phenomena clearly show that water is reversibly incorporated into the beta-chitin crystal and that the temperature change induces transitions between anhydrous, monohydrate, and dihydrate forms. The DSC behavior in heating-cooling cycles, including reversion between the two endothermic peaks, indicated that the transition between monohydrate and dihydrate was a fast and narrow-temperature process, whereas the one between the anhydrous and the monohydrate form was a slow and wide-temperature process.  相似文献   

10.
1. The specific radioactivities of individual molecular species of phosphoglycerides in the skeletal muscles of control rats and of rats deficient in essential fatty acids have been determined 3 h after intraperitoneal injection of ortho[32P] phosphate. 2. It has been demonstrated that the high average specific radioactivity of phosphoglycerides in muscles of rats deficient in essential fatty acids is due to both increased amounts and increased turnover of 1-palmitoyl-2-oleoyl phosphatidylcholine and phosphatidylethanolamine. 3. The 1-stearoyl-2-arachidonoyl phosphatidylcholine was found to turn over faster than the 1-palmitoyl-2-arachidonoyl species. In rats deficient in essential fatty acids, the 1-stearoyl-2-(5,6,11-eicosatrienoyl) phosphatidylcholine turned over more rapidly than the 1-palmitoyl-2-(5,8,11-eicosatrienoyl) species. Both findings are in constant with similar findings for liver.  相似文献   

11.
We have studied the properties of mixtures of cholesterol with dioleoylphosphatidylcholine (DOPC), and with several other phospholipids, including 1-stearoyl-2-oleoylphosphatidylcholine (SOPC) and dioleoleoylphosphatidylserine (DOPS), as a function of cholesterol molar fraction and of temperature. Mixtures of DOPC with a cholesterol molar fraction of 0.4 or greater display polymorphic behavior. This polymorphism includes the formation of structures that give rise to isotropic peaks in 31P NMR at cholesterol molar fractions between 0.4 and 0.6, dependent on the thermal history of the sample. Cryo-electron microscopy studies demonstrate the formation of small globular aggregates that would contribute to a narrowing of the 31P NMR powder pattern.At molar fraction cholesterol 0.6 and higher and at temperatures above 70 degrees C, the mixtures with DOPC convert to the hexagonal phase. Lipid polymorphism is accompanied by the phase separation of cholesterol crystals in the anhydrous form and/or the monohydrate form. The crystals that are formed have substantially altered kinetics of hydration and dehydration, compared with both pure cholesterol monohydrate crystals and with crystals formed in the presence of the other phospholipids that do not form the hexagonal phase in the presence of cholesterol. This fact demonstrates that these cholesterol crystals are in intimate contact with the DOPC phospholipid and are not present as morphologically separate structures.  相似文献   

12.
In view of the demonstrated cholesterol-binding capabilities of certain cyclodextrins, we have examined whether these agents can also catalyze efficient transfer of cholesterol between lipid vesicles. We here demonstrate that beta- and gamma-cyclodextrins can dramatically accelerate the rate of cholesterol transfer between lipid vesicles under conditions where a negligible fraction of the sterol is bound to cyclodextrin in steady state. beta- and gamma-cyclodextrin enhance the rate of transfer of cholesterol between vesicles by a larger factor than they accelerate the transfer of phospholipid, whereas, for alpha- and methyl-beta-cyclodextrin, the opposite is true. Analysis of the kinetics of cyclodextrin-mediated cholesterol transfer between large unilamellar vesicles composed mainly of 1-stearoyl-2-oleoyl phosphatidylcholine (SOPC) or SOPC/cholesterol indicates that transbilayer flip-flop of cholesterol is very rapid (halftime < 1-2 min at 37 degrees C). Using beta-cyclodextrin to accelerate cholesterol transfer, we have measured the relative affinities of cholesterol for a variety of different lipid species. Our results show strong variations in cholesterol affinity for phospholipids bearing different degrees of chain unsaturation and lesser, albeit significant, effects of phospholipid headgroup structure on cholesterol-binding affinity. Our findings also confirm previous suggestions that cholesterol interacts with markedly higher affinity with sphingolipids than with common membrane phospholipids.  相似文献   

13.
Epand RM 《Biophysical journal》2003,84(5):3102-3110
Membranes of the lens of the eye of mammals have two particular characteristics, high concentrations of sphingomyelin, and dihydrosphingomyelin and cholesterol. We have studied the miscibility of cholesterol with both egg sphingomyelin and with dihydrosphingomyelin made by hydrogenation of egg sphingomyelin. At a cholesterol mol fraction of 0.5 and lower, crystallites of cholesterol are not present with either form of sphingomyelin, as observed by differential scanning calorimetry and by (13)C CP/MAS NMR. However, in the range of 0.6 to 0.8 mol fraction of cholesterol increasing amounts of crystallites form, with the amount of anhydrous cholesterol crystals formed being somewhat greater with dihyrosphingomyelin compared with sphingomyelin. Interestingly, cholesterol monohydrate crystallites formed in these two phospholipids exhibit a temperature of dehydration higher than that of pure cholesterol monohydrate crystals. These cholesterol monohydrate crystals form more rapidly and in greater amounts with the unmodified form of sphingomyelin. This difference is likely a consequence of differences at the membrane interface. The chemical shift of the (13)C of the carbonyl group, as measured by CP/MAS NMR, shows that there are differences between the two phospholipids in both the presence and absence of cholesterol. The bilayers with dihydrosphingomyelin are more hydrogen bonded. Cholesterol crystallites are known to be present in the lens of the eye. Our studies show that the ratio of sphingomyelin to dihydrosphingomyelin can affect the rate of formation of these cholesterol crystallites and thus play a role in the membrane of cells of the lens, affecting ocular function.  相似文献   

14.
Mammalian synaptic membranes appear to contain high proportions of specific, sn-1-stearoyl-2-docosahexaenoyl- and sn-1-stearoyl-2-arachidonoyl phosphoglycerides, but the structural significance of this is unclear. Here we used a standardized approach to compare the properties of homogeneous monolayers of the corresponding phosphatidylcholines, phosphatidylethanolamines, phosphatidylserines, and phosphatidic acids with those of control monolayers of sn-1-stearoyl-2-oleoyl- and sn-1-palmitoyl-2-oleoyl phosphoglycerides. Major findings were: 1), that the presence of an sn-2-docosahexaenoyl group or an sn-2-arachidonoyl group increases the molecular areas of phosphoglycerides by 3.8 A(2) (7%) relative to the presence of an sn-2-oleoyl group; 2), that the phosphorylcholine headgroup independently increases molecular areas by a larger amount, 7.1 A(2) (13%); and 3), that the dipole moments of species having an arachidonoyl moiety or an oleoyl moiety are 83 mD (19%) higher than those of comparable docosahexaenoic acid-containing phosphoglycerides. These and other results provide new information about the molecular packing properties of polyenoic phosphoglycerides and raise important questions about the role of these phosphoglycerides in synapses.  相似文献   

15.
We have examined the effects of cholesterol on the thermotropic phase behavior and organization of aqueous dispersions of a homologous series of linear disaturated phosphatidylserines by high-sensitivity differential scanning calorimetry and Fourier transform infrared spectroscopy. We find that the incorporation of increasing quantities of cholesterol progressively reduces the temperature, enthalpy, and cooperativity of the gel-to-liquid-crystalline phase transition of the host phosphatidylserine bilayer, such that a cooperative chain-melting phase transition is completely or almost completely abolished at 50 mol % cholesterol, in contrast to the results of previous studies. We are also unable to detect the presence of a separate anhydrous cholesterol or cholesterol monohydrate phase in our binary mixtures, again in contrast to previous reports. We further show that the magnitude of the reduction in the phase transition temperature induced by cholesterol addition is independent of the hydrocarbon chain length of the phosphatidylserine studied. This result contrasts with our previous results with phosphatidylcholine bilayers, where we found that cholesterol increases or decreases the phase transition temperature in a chain length-dependent manner (1993. Biochemistry, 32:516-522), but is in agreement with our previous results for phosphatidylethanolamine bilayers, where no hydrocarbon chain length-dependent effects were observed (1999. Biochim. Biophys. Acta, 1416:119-234). However, the reduction in the phase transition temperature by cholesterol is of greater magnitude in phosphatidylethanolamine as compared to phosphatidylserine bilayers. We also show that the addition of cholesterol facilitates the formation of the lamellar crystalline phase in phosphatidylserine bilayers, as it does in phosphatidylethanolamine bilayers, whereas the formation of such phases in phosphatidylcholine bilayers is inhibited by the presence of cholesterol. We ascribe the limited miscibility of cholesterol in phosphatidylserine bilayers reported previously to a fractional crystallization of the cholesterol and phospholipid phases during the removal of organic solvent from the binary mixture before the hydration of the sample. In general, the results of our studies to date indicate that the magnitude of the effect of cholesterol on the thermotropic phase behavior of the host phospholipid bilayer, and its miscibility in phospholipid dispersions generally, depend on the strength of the attractive interactions between the polar headgroups and the hydrocarbon chains of the phospholipid molecule, and not on the charge of the polar headgroups per se.  相似文献   

16.
Monomolecular films of phospholipids in the liquid-expanded (LE) phase after supercompression to high surface pressures (pi), well above the equilibrium surface pressure (pi(e)) at which fluid films collapse from the interface to form a three-dimensional bulk phase, and in the tilted-condensed (TC) phase both replicate the resistance to collapse that is characteristic of alveolar films in the lungs. To provide the basis for determining which film is present in the alveolus, we measured the melting characteristics of monolayers containing TC dipalmitoyl phosphatidylcholine (DPPC), as well as supercompressed 1-palmitoyl-2-oleoyl phosphatidylcholine and calf lung surfactant extract (CLSE). Films generated by appropriate manipulations on a captive bubble were heated from < or =27 degrees C to > or =60 degrees C at different constant pi above pi(e). DPPC showed the abrupt expansion expected for the TC-LE phase transition, followed by the contraction produced by collapse. Supercompressed CLSE showed no evidence of the TC-LE expansion, arguing that supercompression did not simply convert the mixed lipid film to TC DPPC. For both DPPC and CLSE, the melting point, taken as the temperature at which collapse began, increased at higher pi, in contrast to 1-palmitoyl-2-oleoyl phosphatidylcholine, for which higher pi produced collapse at lower temperatures. For pi between 50 and 65 mN/m, DPPC melted at 48-55 degrees C, well above the main transition for bilayers at 41 degrees C. At each pi, CLSE melted at temperatures >10 degrees C lower. The distinct melting points for TC DPPC and supercompressed CLSE provide the basis by which the nature of the alveolar film might be determined from the temperature-dependence of pulmonary mechanics.  相似文献   

17.
The sn-1-stearoyl-2-arachidonoyl phospholipids of animal cells appear to be formed by special mechanisms. To determine whether monoacylglycerol (MG) incorporation pathways are involved we incubated quiescent Swiss 3T3 cells with [3H]glycerol-labeled sn-2-arachidonoyl MG, then analyzed the radioactive cell lipids that accumulated. We also examined cell homogenates to identify enzyme activities that might promote the incorporation of sn-2-arachidonoyl MG into other cell lipids. The cell incubation experiments demonstrated rapid labeling of several lipids, including diacylglycerol, lysophosphatidic acid, phosphatidic acid, and phosphatidylinositol. They also demonstrated selective labeling of sn-1-stearoyl-2-arachidonoyl species of phosphatidylinositol, phosphatidylethanolamine, and phosphatidylserine. The cell homogenate experiments identified an sn-2-acyl MG acyltransferase activity, an MG kinase activity that phosphorylates sn-2-arachidonoyl MG in preference to sn-2-oleoyl MG, and a stearoyl-specific acyl transferase activity that converts sn-2-arachidonoyl lysophosphatidic acid into sn-1-stearoyl-2-arachidonoyl phosphatidic acid. The results also showed that this stearoyl transferase could act with other enzymes to convert sn-2-arachidonoyl lysophosphatidic acid into sn-1-stearoyl-2-arachidonoyl phosphatidylinositol. The combined results indicate that Swiss 3T3 cells incorporate sn-2-arachidonoyl MG into phospholipids by at least two different pathways, including one that specifically forms sn-1-stearoyl-2-arachidonoyl phosphatidylinositol.  相似文献   

18.
The interaction between the cationic lipid DOTAP and cholesterol is examined in high cholesterol formulations by differential scanning calorimetry (DSC). Preparation of liposomes above 66 mol% cholesterol results in formulations that exhibit a calorimetric transition for anhydrous cholesterol at 38-40 °C. The enthalpy of this transition progressively increases at higher cholesterol contents, and is not detected below 66 mol% cholesterol. Furthermore, the enthalpy changes indicate that the composition of the non-domain forming portion containing DOTAP saturated with cholesterol is relatively constant above 66 mol% cholesterol. Greater transfection efficiency in the presence of 50% serum is observed at the formulations with high cholesterol contents where anhydrous cholesterol domains are detected by DSC. Although formulations possessing higher cholesterol exhibited a greater resistance to serum-induced aggregation, maintenance of small particle size does not appear to be responsible for the enhanced transfection efficiency. Additional studies quantifying albumin binding suggest that cholesterol domains in the lipid/DNA complex do not bind protein, and this may enable these moieties to enhance transfection by facilitating membrane fusion.  相似文献   

19.
Epand RM  Sayer BG  Epand RF 《Biochemistry》2003,42(49):14677-14689
The peptide N-acetyl-LWYIK-amide causes the reorganization of bilayers of phosphatidylcholine and cholesterol to produce domains enriched in cholesterol. At a cholesterol mol fraction of 0.5, addition of N-acetyl-LWYIK-amide results in the formation of cholesterol crystallites. Addition of this peptide to mixtures of 1-stearoyl-2-oleoylphosphatidylcholine with lower mol fractions of cholesterol results in an increase in the enthalpy of the chain melting transition of the phospholipid, indicating the depletion of cholesterol from a domain in the membrane. The peptide binds to membranes both with and without cholesterol. However, (1)H magic-angle spinning (MAS) nuclear Overhauser effect spectroscopy (NOESY) indicates that in the presence of cholesterol the peptide has greater penetration into the bilayer. (13)C MAS NMR indicates that the peptide has stronger interactions with the A ring of cholesterol than it does with the interior of the bilayer. These results are in contrast with those of another peptide, N-acetyl-KYWFYR-amide, which does not promote the formation of cholesterol crystallites and does not show preferential interaction with cholesterol by NMR. Therefore, cholesterol can promote the insertion of N-acetyl-LWYIK-amide into a membrane and this peptide will sequester cholesterol into domains. These properties help to explain the observation that this sequence is found to be important in causing the fusion protein of human immunodeficiency virus (HIV) to sequester into raft domains in biological membranes.  相似文献   

20.
Isolated human plasma low density lipoprotein (LDL) was observed to possess sphingomyelinase activity. Accordingly, the formation of ceramide was catalyzed by LDL at 37 degrees C using tertiary liposomes composed of sphingomyelin (mole fraction (x) = 0.2), 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (x = 0.7), 1, 2-dimyristoyl-sn-glycero-3-phospho-rac-glycerol (x = 0.1), and either the fluorescent sphingomyelin analog Bodipy-sphingomyelin or [(14)C]sphingomyelin as substrates. However, this activity was not present in either very low density lipoprotein or the high density lipoprotein subfractions HDL(2) and HDL(3). Oxidation of LDL abrogated its sphingomyelinase activity. Aggregation of the liposomes upon incubation with LDL was evident from the light scattering measurements. Microinjection of LDL to the surface of giant liposomes composed of 1-stearoyl-2-oleoyl-sn-glycero-3-phosphocholine (SOPC), N-palmitoyl-d-sphingomyelin (C16:0-sphingomyelin), and Bodipy-sphingomyelin as a fluorescent tracer (0.75:- 0.20:0.05, respectively) revealed the induction of vectorial budding of vesicles, resembling endocytosis.  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