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.
The head-group orientations and molecular dynamics of three glyceroglycolipids in aqueous dispersions, as determined by 2H-NMR, are compared. 1,2-Di-O-tetradecyl-3-O-(alpha-D-glucopyranosyl)-sn-glycerol (alpha-DTGL) and 1,2-di-O-tetradecyl-3-O-(alpha-D-mannopyranosyl)-sn-glycerol (alpha-DTML), selectively 2H-labelled on the pyranose ring, at the exocyclic hydroxymethyl group, and at C3 of glycerol, have been studied by 2H-NMR and the results compared with those reported earlier for 1,2-di-O-tetradecyl-3-O-(beta-D-glucopyranosyl)-sn-glycerol (beta-DTGL). The alpha-glucolipid exhibits a gel-to-liquid crystal phase transition and a lamellar to hexagonal mesophase transition at temperatures which are similar to those of the beta-anomer, beta-DTGL. However, alpha-DTGL exhibits head group orientations and molecular ordering in the lamellar and hexagonal phases which differ strikingly with those reported for the corresponding beta-glucolipid. Whereas the head group of beta-DTGL is extended away from the bilayer surface into the aqueous phase, that of alpha-DTGL is almost parallel to the bilayer surface. alpha-DTGL exhibits a molecular order parameter of 0.56 which is substantially greater than that of its anomer, beta-DTGL, 0.45. The latter indicates that the head group region of the alpha-glyceroglucolipid is characterized by smaller angular fluctuations than that of beta-DTGL. On entering the hexagonal mesophase the pyranose ring of the beta-glucolipid undergoes a large reorientation relative to the motional axis of the head group, whereas the alpha-anomer exhibits only a small orientational change. 1,2-Di-O-tetradecyl-3-O-(alpha-D-mannopyranosyl)-sn-glycerol (alpha-DTML) undergoes a phase transition at 47 degrees C, attributed to the unusual lamellar gel to hexagonal phase transition. The pyranose ring of alpha-DTML, in a mixture with dimyristoylphosphatidylcholine (1:9 mol ratio) to give a lamellar liquid crystalline phase, is oriented away from the bilayer surface into the aqueous environment and has an Smol of 0.75. The results for alpha-DTML, 2H-labelled at the C3 position of glycerol, suggest that this segment also has high molecular ordering. alpha-DTML in a lamellar environment has the least flexible membrane surface of the glyceroglycolipids investigated to date. 2H-NMR spin lattice relaxation times have been used to probe the head group motions of the glycolipids. The results indicate that the rate of head group motion increases in the order alpha-DTML less than alpha-DTGL less than beta-DTGL.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
The polar headgroup contribution to monolayer behavior of dipalmitoylglucosylglycerol has been examined through studies of 1,2-di-O-palmitoyl-3-O-(alpha-D-glucopyranosyl)-sn-glycerol (di-16:0-alpha GlcDG) and 1,2-di-O-palmitoyl-3-O-(beta-D-glucopyranosyl)-sn-glycerol (di-16:0-beta GlcDG) in which the sugar headgroup is linked via an alpha or beta linkage to the diacylglycerol moiety. The results indicate that the limiting areas per molecule of the resultant condensed states are smaller than those of the corresponding phosphatidylcholine (DPPC) but larger than those of dipalmitoylphosphatidylethanolmine (DPPE). In the expanded state, while the areas per molecule are similar to those of DPPC at low pressures, both glycolipids occupy smaller areas at higher pressures. The expanded-state areas of the glucolipids are also slightly greater than those of DPPE. The initial compressional phase transition pressure of the glucolipid liquid-expanded/liquid-condensed transition (pi t) is, however, less sensitive to temperature than are the pi t values of phospholipids. Both of these effects must relate to strong headgroup/water interactions, which, in turn, result in a stabilization of the liquid-expanded states. In the expanded states the alpha anomers are slightly less tightly packed than the beta anomers, as is indicated by the somewhat higher areas per molecule of the expanded states and the lower transition temperatures. These differences in chain-melting temperatures are slightly smaller than those observed in bilayers. While the areas per molecule of the dipalmitoyl glucolipids are greater than those of dipalmitoylphosphatidylethanolamine, they nevertheless exhibit a greater tendency to form nonbilayer structures. Such observations indicate that other factors besides geometric shape play a role in bilayer/nonbilayer transitions.  相似文献   

3.
Niu SL  Litman BJ 《Biophysical journal》2002,83(6):3408-3415
Lateral domain or raft formation in biological membranes is often discussed in terms of cholesterol-lipid interactions. Preferential interactions of cholesterol with lipids, varying in headgroup and acyl chain unsaturation, were studied by measuring the partition coefficient for cholesterol in unilamellar vesicles. A novel vesicle-cyclodextrin system was used, which precludes the possibility of cross-contamination between donor-acceptor vesicles or the need to modify one of the vesicle populations. Variation in phospholipid headgroup resulted in cholesterol partitioning in the order of sphingomyelin (SM) > phosphatidylserine > phosphatidylcholine (PC) > phosphatidylenthanolamine (PE), spanning a range of partition DeltaG of -1181 cal/mol to +683 cal/mol for SM and PE, respectively. Among the acyl chains examined, the order of cholesterol partitioning was 18:0(stearic acid),18:1n-9(oleic acid) PC > di18:1n-9PC > di18:1n-12(petroselenic acid) PC > di18:2n-6(linoleic acid) PC > 16:0(palmitic acid),22:6n-3(DHA) PC > di18:3n-3(alpha-linolenic acid) PC > di22:6n-3PC with a range in partition DeltaG of 913 cal/mol. Our results suggest that the large differences observed in cholesterol-lipid interactions contribute to the forces responsible for lateral domain formation in plasma membranes. These differences may also be responsible for the heterogeneous cholesterol distribution in cellular membranes, where cholesterol is highly enriched in plasma membranes and relatively depleted in intracellular membranes.  相似文献   

4.
The mechanism whereby bacteriorhodopsin (BR), the light driven proton pump from the purple membrane of Halobacterium halobium, arranges in a 2D-hexagonal array, has been studied in bilayers containing the protein, 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and various fractions of H. halobium membrane lipids, by freeze fracture electron microscopy and examination of optical diffractograms of the micrographs obtained. Electron micrographs of BR/DMPC complexes containing the entire polar lipid component of H. halobium cell membranes or the total lipid component of the purple membrane, with a protein-to-total lipid molar ratio of less than 1:50 and to which 4 M NaCl had been added, revealed that trimers of BR formed into an hexagonal 2D-array similar to that found in the native purple membrane, suggesting that one or more types of the purple membrane polar lipids are required for array formation. To support this suggestion, bacteriorhodopsin was purified free of endogenous purple membrane lipids and reconstituted into lipid bilayer complexes by detergent dialysis. The lipids used to form these complexes are 1,2-dimyristoyl-sn-glycerol-phosphocholine (DMPC) as the major lipid and, separately, each of the individual lipid types from the H. halobium cell membranes, namely 2,3-di-O-phytanyl-sn-glycero-1-phosphoryl-3'-sn-glycerol 1'-phosphate (DPhPGP), 2,3-di-O-phytanyl-sn-glycero-1-phosphoryl-3'-sn-glycerol 1'-sulphate (DPhPGS), 2,3-di-O-phytanyl-sn-glycero-1-phosphoryl-3'-sn-glycerol (DPhPG) and 2,3-di-O-phytanyl-1-O-[beta-D-Galp-3-sulphate-(1----6)-alpha-D- Manp-(1----2)-alpha-D-Glcp]-sn-glycerol (DPhGLS). When examined by freeze-fracture electron microscopy, only the complexes containing 2,3-di-O-phytanyl-sn-glycero-1-phosphoryl-3'-sn-glycerol- 1'-phosphate or 2,3-di-O-phytanyl-sn-glycero-1-phosphoryl-3'-sn-glycerol-1'-sulphate, at high protein density (less than 1:50, bacteriorhodopsin/phospholipid, molar ratio) and to which 4 M NaCl had been added, showed well defined 2D hexagonal arrays of bacteriorhodopsin trimers similar to those observed in the purple membrane of H. halobium.  相似文献   

5.
Cholesterol was found to inhibit full fusion of oppositely charged phospholipid bilayer vesicles by stabilizing the contacting membranes at the stage of the hemifused intermediate. Vesicles of opposite charge containing different amounts of cholesterol were prepared using cationic (1,2-dioleoyl-sn-glycero-3-ethylphosphocholine) and anionic (dioleoylphosphatidylglycerol) phospholipids. Pairwise interactions between such vesicles were observed by fluorescence video microscopy in real time after electrophoretically maneuvering the vesicles into contact. Hemifusion accounted for more than 80% of the observed events when the vesicles contained 33-50 mole% cholesterol. In contrast, vesicles containing only a small proportion of cholesterol (相似文献   

6.
The formability, current-voltage characteristics and stability of the planar lipid bilayer membranes from the synthetic phytanyl-chained glycolipids, 1, 3-di-O-phytanyl-2-O-(beta-glycosyl)glycerols (Glc(Phyt)(2), Mal(N)(Phyt)(2)) were studied. The single bilayer membranes were successfully formed from the glycolipid bearing a maltotriosyl group (Mal(3)(Phyt)(2)) by the folding method among the synthetic glycolipids examined. The membrane conductance of Mal(3)(Phyt)(2) bilayers in 100 mM KCl solution was significantly lower than that of natural phospholipid, soybean phospholipids (SBPL) bilayers, and comparable to that of 1,2-diphytanoyl-sn-glycero-3-phosphocholine (DPhPC) bilayers. From the permeation measurements of lipophilic ions through Mal(3)(Phyt)(2) and DPhPC bilayers, it could be presumed that the carbonyl groups in glycerol backbone of the lipid molecule are not necessarily required for the total dipole potential barrier against cations in Mal(3)(Phyt)(2) bilayer. The stability of Mal(3)(Phyt)(2) bilayers against long-term standing and external electric field change was rather high, compared with SBPL bilayers. Furthermore, a preliminary experiment over the functional incorporation of membrane proteins was demonstrated employing the channel proteins derived from octopus retina microvilli vesicles. The channel proteins were functionally incorporated into Mal(3)(Phyt)(2) bilayers in the presence of a negatively charged glycolipid. From these observations, synthetic phytanyl-chained glycolipid bilayers are promising materials for reconstitution and transport studies of membrane proteins.  相似文献   

7.
Cholesterol and saturated lipid species preferentially partition into liquid ordered microdomains, such as lipid rafts, away from unsaturated lipid species for which the sterol has less affinity in the surrounding liquid-disordered membrane. To observe how cholesterol interacts with unsaturated phospholipids, we have determined, from one-dimensional neutron scattering length density profiles, the depth of cholesterol in phosphatidylcholine (PC) bilayers with varying amounts of acyl chain unsaturation. Through the use of [2,2,3,4,4,6-(2)H(6)]-labeled cholesterol, we show that in 1-palmitoyl-2-oleoylphosphatidylcholine (16:0-18:1 PC), 1,2-dioleoylphosphatidylcholine (18:1-18:1 PC), and 1-stearoyl-2-arachidonylphosphatidylcholine (18:0-20:4 PC) bilayers the center of mass of the deuterated sites is approximately 16 A from the bilayer center. This location places the hydroxyl group of the sterol moiety at the hydrophobic/hydrophilic bilayer interface, which is the generally accepted position. In dramatic contrast, for 20:4-20:4 PC membranes the hydroxyl group is found, unequivocally, sequestered in the bilayer center. We attribute the change in location to the high disorder of polyunsaturated fatty acids (PUFA) that is incompatible with close proximity to the steroid moiety in its usual "upright" orientation.  相似文献   

8.
Massey JB  Pownall HJ 《Biochemistry》2005,44(30):10423-10433
7-Ketocholesterol is an oxidized derivative of cholesterol with numerous physiological effects. In model membranes, 7-ketocholesterol and cholesterol were compared by physical measures of bilayer order and polarity, formation of detergent resistant domains (DRM), phase separation, and membrane microsolubilization by apolipoprotein A-I. In binary mixtures of a saturated phosphatidylcholine (PC), dipalmitoyl-PC (DPPC), and cholesterol or 7-ketocholesterol, the sterols modulate bilayer order and polarity and induce DRM formation to a similar extent. Cholesterol induces formation of ordered lipid domains (rafts) in tertiary mixtures with dioleoyl-PC (DOPC) and DPPC, or DOPC and sphingomyelin (SM). In tertiary mixtures, cholesterol increased lipid order and reduces bilayer polarity more than 7-ketocholesterol. This effect was more pronounced when the mixtures were in a miscible liquid-disordered (L(d)) phase. Substitution of 7-ketocholesterol for cholesterol dramatically reduced the extent of DRM formation in DOPC/DPPC and DOPC/SM bilayers and ordered lipid phase separation in mixtures of a spin-labeled PC with DPPC and with SM. Compared to cholesterol, 7-ketocholesterol decreased the rate for the microsolubilization of dimyristoyl-PC multilamellar vesicles by apolipoprotein A-I. The membrane effects of 7-ketocholesterol were dependent on the phospholipid matrix. In L(d) phase phospholipids, a model for 7-ketocholesterol indicates that the proximity of the 7-keto and 3beta-OH groups puts both polar moieties at the lipid-water interface to tilt the sterol nucleus to the plane of the bilayer. 7-Ketocholesterol was less effective in forming ordered lipid domains, in decreasing the level of bilayer hydration, and in forming phase boundary bilayer defects. Compared to cholesterol, 7-ketocholesterol can differentially modulate membrane properties involved in protein-membrane association and function.  相似文献   

9.
l-1-Deoxy-1-fluoro-6-O-methyl-myo-inositol was epimerized by chloral/DCC in boiling 1,2-dichloroethane yielding D-1-O-cyclohexylcarbamoyl-2-deoxy-2-fluoro-3-O-methyl-5,6-O-[(R/S)-2,2,2-trichloroethylidene]-chiro-inositol. The latter and l-4-O-benzyl-3-O-cyclohexylcarbamoyl-5-O-methyl-1,2-O-(2,2,2-trichloroethylidene)-muco-inositol, l-4-O-benzyl-3-O-cyclohexylcarbamoyl-1,2-O-ethylidene-5-O-methyl-muco-inositol, d-1-O-cyclohexylcarbamoyl-2-deoxy-5,6-O-ethylidene-2-fluoro-3-O-methyl-chiro-inositol, as well as D-5-O-benzyl-4-O-cyclohexylcarbamoyl-3-deoxy-3-(N,N'-dicyclohexylureido)-6-O-methyl-1,2-O-(2,2,2-trichloroethylidene)-chiro-inositol were deprotected with boiling 57% aq hydrogen iodide. Ether, urethane and ethylidene acetal functions were simultaneously cleaved by the reagent, whereas the trichloroethylidene groups were still intact or were only removed in small quantities. Especially, the urea function of D-5-O-benzyl-4-O-cyclohexylcarbamoyl-3-deoxy-3-(N,N'-dicyclohexylureido)-6-O-methyl-1,2-O-(2,2,2-trichloroethylidene)-chiro-inositol was decomposed to a cyclohexylamino group. The hydrodechlorination of D-1-O-cyclohexylcarbamoyl-2-deoxy-2-fluoro-3-O-methyl-5,6-O-[(R/S)-2,2,2-trichloroethylidene]-chiro-inositol using Raney-Nickel yielded a mixture of the corresponding 5,6-O-ethylidene- and 5,6-O-chloroethylidene derivatives. The three synthetic steps-hydrodehalogenation, HI-deprotection and peracylation- were combined without purification of the intermediates.  相似文献   

10.
1. Phase transitions in sonicated (vesicles) and unsonicated liposomes composed of various synthetic phosphatidylcholines are monitored using differential scanning calorimetry and 31P NMR. 2. The temperature (Tc), heat content and width of the phase transition are comparable in both vesicles and liposomes prepared from 1,2-dipalmitoyl phosphatidylcholine and 1,2-dimyristoyl phosphatidylcholine. In vesicles composed of a (1 : 1) mixture of 1,2-dipalmitoyl phosphatidylcholine and 1,2-dioleoyl phosphatidylcholine phase separation occurs as in the bilayers of the unsonicated liposomes. 3. The linewidth of the 31P resonances in vesicles is not greatly dependent upon the fatty acid composition when the lipids are in the disordered liquid crystalline state (above Tc). When the lipids are in the gel state (below Tc), however, there is a marked increase in linewidth, demonstrating a reduction in motion of the phosphate group. 4. The ratio of the amounts of phosphatidylcholine present in the outside and inside monolayter of the vesicle membrane was determined with 31P NMR using Nd3+ as a non-permeating shift reagent. 5. The outside/inside ratio is dependent upon the hydrocarbon chain length. Increasing chain length gives a lower outside/inside ratio and a larger vesicle. Introduction of cis or trans double bonds in the chain influences the outside/inside ratio slightly. 6. The incorporation of cholesterol decreases the outside/inside ratio and increases the size of 1,2-dimyristoyl phosphatidylcholine vesicles. The cholesterol concentration in the outside and inside monolayer is approximately the same. The size of the 1,2-dioleoyl phosphatidylcholine vesicles is also increased by cholesterol incorporation but the outside/inside distribution is also increased, especially between 30 and 50 mol% cholesterol. In these vesicles cholesterol is asymmetrically distributed and strongly prefers the inside monolayer of the vesicle.  相似文献   

11.
Treatment of methyl 3,4,6-tri-O-benzyl-2-O-(2,3,4-tri-O-acetyl-alpha-D-mannopyranosyl)-alpha -D- mannopyranoside with N,N-diethylaminosulfur trifluoride (Et2NSF3), followed by O-deacetylation and catalytic hydrogenolysis, afforded methyl 2-O-(6-deoxy-6-fluoro-alpha-D-mannopyranosyl)-alpha-D-mannopyranoside (8). Methyl 6-deoxy-6-fluoro-2-O-alpha-D-mannopyranosyl-alpha-D-mannopyranoside (11) was similarly obtained from methyl 3-O-benzyl-2-O-(2,3,4,6-tetra-O-acetyl-alpha-D-mannopyranosyl-alpha-D- mannopyranoside. 1,2,3,4-Tetra-O-acetyl-6-deoxy-6-fluoro-beta-D-mannopyranose (13), used for the synthesis of the 4-nitrophenyl analogs of 8 and 11, as well as their 3-O-linked isomers, was obtained by treatment of 1,2,3,4-tetra-O-acetyl-beta-D-mannopyranose with Et2NSF3. Treatment of 13 with 4-nitrophenol in the presence of tin(IV) chloride, followed by sequential O-deacetylation, isopropylidenation, acetylation, and cleavage of the acetal group, afforded 4-nitrophenyl 4-O-acetyl-6-deoxy-6-fluoro-alpha-D-mannopyranoside (18). Treatment of 13 with HBr in glacial acetic acid furnished the 6-deoxy-6-fluoro bromide 19. Glycosylation of diol 18 with 20 gave 4-nitrophenyl 4-O-acetyl-6-deoxy-6-fluoro-3-O- (21) and -2-O-(2,3,4,6-tetra-O-acetyl-alpha-D-mannopyranosyl)-alpha-D- mannopyranoside (23) in the ratio of approximately 2:1, together with a small proportion of a branched trisaccharide. 4-Nitrophenyl 4,6-di-O-acetyl-alpha-D-mannopyranoside was similarly glycosylated with bromide 19 to give 4-nitrophenyl 4,6-di-O-acetyl-3-O- and -2-O-(2,3,4-tri- O-acetyl-6-deoxy-6-fluoro-alpha-D-mannopyranosyl)-alpha-D-mannopyranosid e. The various di- and tri-saccharides were O-deacetylated by Zemplén transesterification.  相似文献   

12.
C C Kan  Z S Ruan  R Bittman 《Biochemistry》1991,30(31):7759-7766
Cholesterol undergoes exchange between membranes containing sphingomyelin at a much slower rate than between membranes lacking sphingomyelin. To investigate the role of the hydroxy group at the 3-position of sphingomyelin in the interaction between sphingomyelin and cholesterol, we have measured the rates of [4-14C]cholesterol exchange between unilamellar vesicles prepared with N-stearoylsphingomyelin or with synthetic analogues in which the hydroxy group is replaced with an O-alkyl group or with hydrogen. Vesicles prepared from 3-deoxy- and 3-O-methyl-N-stearoylsphingomyelin had the same rate of [14C]-cholesterol desorption. The half-times for exchange from vesicles prepared with 3-O-methyl- and 3-deoxy-N-stearoylsphingomyelins and 10 mol % of cholesterol were only slightly faster (a factor of only 1.5) than that found from vesicles prepared from N-stearoylsphingomyelin and 10 mol % cholesterol. The rate of cholesterol desorption from vesicles could be accelerated by preparing vesicles from bulky 3-O-alkyl analogues of sphingomyelin. Vesicles containing 3-O-ethyl-N-stearoylsphingomyelin and 3-O-tetrahydropyranyl egg sphingomyelin gave rate enhancements of approximately 14 and 35, compared with the rates observed in vesicles made from N-stearoyl- and egg sphingomyelin, respectively. These data indicate that insertion of sterically bulky groups at the 3-position of sphingomyelin (such as ethoxy and tetrahydropyranyloxy) in place of hydroxy interferes markedly with the molecular packing of cholesterol and sphingomyelin in bilayer membranes; however, the hydroxy group of sphingomyelin is not critical for the strong interaction of cholesterol with sphingomyelin.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
In this study, we have examined the membrane properties and sterol interactions of phosphatidyl alcohols varying in the size of the alcohol head group coupled to the sn-3-linked phosphate. Phosphatidyl alcohols of interest were dipalmitoyl derivatives with methanol (DPPMe), ethanol (DPPEt), propanol (DPPPr), or butanol (DPPBu) head groups. The Phosphatidyl alcohols are biologically relevant, because they can be formed in membranes by the phospholipase D reaction in the presence of alcohol. The melting behavior of pure phosphatidyl alcohols and mixtures with 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) or cholesterol was assessed using high sensitivity differential scanning calorimetry (DSC). DPPMe had the highest melting temperature (∼ 49 °C), whereas the other phosphatidyl alcohols had similar melting temperatures as DPPC (∼ 40-41 °C). All phosphatidyl alcohols, except DPPMe, also showed good miscibility with DPPC. The effects of cholesterol on the melting behavior and membrane order in multilamellar bilayer vesicles were assessed using steady-state anisotropy of 1,6-diphenyl-1,3,5-hexatriene (DPH) and DSC. The ordering effect of cholesterol in the fluid phase was lower for all phosphatidyl alcohols as compared to DPPC and decreased with increasing head group size. The formation of ordered domains containing the phosphatidyl alcohols in complex bilayer membranes was determined using fluorescence quenching of DPH or the sterol analogue cholesta-5,7,(11)-trien-3-beta-ol (CTL). The phosphatidyl alcohols did not appear to form sterol-enriched ordered domains, whereas DPPMe, DPPEt appeared to form ordered domains in the temperature window examined (10-50 °C). The partitioning of CTL into bilayer membranes containing phosphatidyl alcohols was to a small extent increased for DPPMe and DPPEt, but in general, sterol interactions were weak or unfavorable for the phosphatidyl alcohols. Our results show that the biophysical and sterol interacting properties of phosphatidyl alcohols, having identical acyl chain structures, are markedly dependent on the size of the head group.  相似文献   

14.
In this study stereochemically pure d-erythro-sphingomyelins (SMs) with either 16:0 or 18:1(cisDelta9) as the N-linked acyl-chain were synthesized. Our purpose was to examine the properties of these sphingomyelins and acyl-chain matched racemic (d-erythro/l-threo) sphingomyelins in model membranes. Liquid-expanded d-erythro-N-16:0-SM in monolayers was observed to pack more densely than the corresponding racemic sphingomyelin. Cholesterol desorption to beta-cyclodextrin was significantly slower from d-erythro-N-16:0-SM monolayers than from racemic N-16:0-SM monolayers. Significantly more condensed domains were seen in cholesterol/d-erythro-N-16:0-SM monolayers than in the corresponding racemic mixed monolayers, when [7-nitrobenz-2-oxa-1, 3-diazol-4-yl]phosphatidylcholine was used as a probe in monolayer fluorescence microscopy. With monolayers of N-18:1-SMs, both the lateral packing densities (sphingomyelin monolayers) and the rates of cholesterol desorption (mixed cholesterol/sphingomyelin monolayers) was found to be similar for d-erythro and racemic sphingomyelins. The phase transition temperature and enthalpy of d-erythro-N-16:0-SM in bilayer membranes were slightly higher compared with the corresponding racemic sphingomyelin (41.1 degrees C and 8.4 +/- 0.4 kJ/mol, and 39.9 degrees C and 7.2 +/- 0.2 kJ/mol, respectively). Finally, d-erythro-sphingomyelins in monolayers (both N-16:0 and N-18:1 species) were not as easily degraded at 37 degrees C by sphingomyelinase (Staphylococcus aureus) as the corresponding racemic sphingomyelins. We conclude that racemic sphingomyelins differ significantly in their biophysical properties from the physiologically relevant d-erythro sphingomyelins.  相似文献   

15.
It is known that ceramides can influence the lateral organization in biological membranes. In particular ceramides have been shown to alter the composition of cholesterol and sphingolipid enriched nanoscopic domains, by displacing cholesterol, and forming gel phase domains with sphingomyelin. Here we have investigated how the bilayer content of ceramides and their chain length influence sterol partitioning into the membranes. The effect of ceramides with saturated chains ranging from 4 to 24 carbons in length was investigated. In addition, unsaturated 18:1- and 24:1-ceramides were also examined. The sterol partitioning into bilayer membranes was studied by measuring the distribution of cholestatrienol, a fluorescent cholesterol analogue, between methyl-β-cyclodextrin and large unilamellar vesicle with defined lipid composition. Up to 15 mol% ceramide was added to bilayers composed of DOPC:PSM:cholesterol (3:1:1), and the effect on sterol partitioning was measured. Both at 23 and 37 °C addition of ceramide affected the sterol partitioning in a chain length dependent manner, so that the ceramides with intermediate chain lengths were the most effective in reducing sterol partitioning into the membranes. At 23 °C the 18:1-ceramide was not as effective at inhibiting sterol partitioning into the vesicles as its saturated equivalent, but at 37 °C the additional double bond had no effect. The longer 24:1-ceramide behaved as 24:0-ceramide at both temperatures. In conclusion, this work shows how the distribution of sterols within sphingomyelin-containing membranes is affected by the acyl chain composition in ceramides. The overall membrane partitioning measured in this study reflects the differential partitioning of sterol into ordered domains where ceramides compete with the sterol for association with sphingomyelin.  相似文献   

16.
Rats were maintained on nutritionally complete diets enriched in unsaturated (menhaden fish oil) or saturated (butter fat) triacylglycerols. After 4 weeks, the animals were killed, proximal small intestinal brush-border membranes were prepared, and examined and compared with respect to their lipid composition, molecular species of phosphatidylcholine, lipid fluidity and sodium-dependent D-glucose transport. Membranes prepared from the two dietary groups were found to possess similar ratios of cholesterol/phospholipid (mol/mol), sphingomyelin/phosphatidylcholine (mol/mol), and protein/lipid (w/w). In contrast to these findings, however, striking differences were noted in the total fatty acid compositions of these membranes. Plasma membranes prepared from animals fed the fish oil diet possessed higher percentages of saturated fatty acids as well as (n - 3) unsaturated fatty acids and lower percentages of monounsaturated and (n - 6) unsaturated fatty acids than those prepared from animals fed the butter fat diet. Analysis of the molecular species of phosphatidylcholine by HPLC, moreover, revealed that membranes from rats fed fish oil had higher levels of 16:0-20:5, 16:0-22:6 and 18:0-20:5 and lower levels of 18:0-18:2 and 16:0-18:1 than their butter fat counterparts. As assessed by steady-state fluorescence polarization, differential polarized phase fluorometric and excimer/monomer fluorescence intensity techniques using various fluorophores, the lipid fluidity of membranes from rats fed fish oil was also found to be significantly lower compared to membranes from rats fed butter fat. Finally, comparison of the kinetic parameters of Na+-dependent D-glucose transport revealed that fish oil-membrane vesicles had a higher maximum velocity (Vmax) than butter fat membrane vesicles but a similar Km for glucose.  相似文献   

17.
We have prepared acyl chain-defined D-erythro-sphingomyelins and D-erythro-dihydrosphingomyelins and compared their properties in monolayer and bilayer membranes. Surface pressure/molecular area isotherms of D-erythro-N-16:0-sphingomyelin (16:0-SM) and D-erythro-N-16:0-dihydrosphingomyelin (16:0-DHSM) show very similar packing properties, except that the expanded-to-condensed phase transition (crystallization) occurs at a lower surface pressure for 16:0-DHSM. The measured surface potential was generally about 100 mV less for 16:0-DHSM monolayers compared to 16:0-SM monolayers. The condensed domains (crystals) that formed in 16:0-SM monolayers as a function of compression displayed star-shaped morphology when viewed under an epifluorescence microscope. 16:0-DHSM monolayers did not form similar crystals upon compression. 16:0-DHSM was degraded much faster by sphingomyelinase from Staphylococcus aureus than 16:0-SM (10-fold difference in enzyme activity needed for comparable hydrolytic rate). Cholesterol desorption from 16:0-DHSM to cyclodextrin was slightly slower (approximately 20%) than the rate measured from 16:0-SM monolayers (at 60 mol % cholesterol). The bilayer melting temperature of 16:0-DHSM was 47.7 degrees C (DeltaH 8.3 kcal/mol) whereas it was 41.2 degrees C for 16:0-SM (DeltaH 8.1 kcal/mol). Cholesterol/16:0-DHSM bilayers (15 mol % sterol) had more condensed domains than comparable 16:0-SM bilayers, as evidenced from the quenching resistance of DPH in DHSM membranes. We conclude that cholesterol interacts more favorably with 16:0-DHSM and that the membranes are more condensed than comparable 16:0-SM-containing membranes.  相似文献   

18.
Cationic amphiphiles used for transfection can be incorporated into biological membranes. By differential scanning calorimetry (DSC), cholesterol solubilization in phospholipid membranes, in the absence and presence of cationic amphiphiles, was determined. Two different systems were studied: 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC)+cholesterol (1:3, POPC:Chol, molar ratio) and 1-palmitoyl-2-oleoyl-sn-glycero-3-[phospho-l-serine] (POPS)+cholesterol (3:2, POPS:Chol, molar ratio), which contain cholesterol in crystallite form. For the zwitterionic lipid POPC, cationic amphiphiles were tested, up to 7 mol%, while for anionic POPS bilayers, which possibly incorporate more positive amphiphiles, the fractions used were higher, up to 23 mol%. 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP) and DOTAP in methyl sulfate salt form (DOTAPmss) were found to cause a small decrease on the enthalpy of the cholesterol transition of pure cholesterol aggregates, possibly indicating a slight increase on the cholesterol solubilization in POPC vesicles. With the anionic system POPS:Chol, the cationic amphiphiles dramatically change the cholesterol crystal thermal transition, indicating significant changes in the cholesterol aggregates. For structural studies, phospholipids spin labeled at the 5th or 16th carbon atoms were incorporated. In POPC, at the bilayer core, the cationic amphiphiles significantly increase the bilayer packing, decreasing the membrane polarity, with the cholesterol derivative 3 beta-[N-(N',N'-dimethylaminoethane)-carbamoyl]-cholesterol (DC-chol) displaying a stronger effect. In POPS and POPS:Chol, DC-chol was also found to considerably increase the bilayer packing. Hence, exogenous cationic amphiphiles used to deliver nucleic acids to cells can change the bilayer packing of biological membranes and alter the structure of cholesterol crystals, which are believed to be the precursors to atherosclerotic lesions.  相似文献   

19.
Cationic amphiphiles used for transfection can be incorporated into biological membranes. By differential scanning calorimetry (DSC), cholesterol solubilization in phospholipid membranes, in the absence and presence of cationic amphiphiles, was determined. Two different systems were studied: 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) + cholesterol (1:3, POPC:Chol, molar ratio) and 1-palmitoyl-2-oleoyl-sn-glycero-3-[phospho-l-serine] (POPS) + cholesterol (3:2, POPS:Chol, molar ratio), which contain cholesterol in crystallite form. For the zwitterionic lipid POPC, cationic amphiphiles were tested, up to 7 mol%, while for anionic POPS bilayers, which possibly incorporate more positive amphiphiles, the fractions used were higher, up to 23 mol%. 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP) and DOTAP in methyl sulfate salt form (DOTAPmss) were found to cause a small decrease on the enthalpy of the cholesterol transition of pure cholesterol aggregates, possibly indicating a slight increase on the cholesterol solubilization in POPC vesicles. With the anionic system POPS:Chol, the cationic amphiphiles dramatically change the cholesterol crystal thermal transition, indicating significant changes in the cholesterol aggregates. For structural studies, phospholipids spin labeled at the 5th or 16th carbon atoms were incorporated. In POPC, at the bilayer core, the cationic amphiphiles significantly increase the bilayer packing, decreasing the membrane polarity, with the cholesterol derivative 3β-[N-(N′,N′-dimethylaminoethane)-carbamoyl]-cholesterol (DC-chol) displaying a stronger effect. In POPS and POPS:Chol, DC-chol was also found to considerably increase the bilayer packing. Hence, exogenous cationic amphiphiles used to deliver nucleic acids to cells can change the bilayer packing of biological membranes and alter the structure of cholesterol crystals, which are believed to be the precursors to atherosclerotic lesions.  相似文献   

20.
Molecular dynamics (MD) computer simulations of five different hydrated unsaturated phosphatidylcholine lipid bilayers built up by 18:0/18:1(n-9)cis PC, 18:0/18:2(n-6)cis PC, 18:0/18:3(n-3)cis PC, 18:0/20:4(n-6)cis PC, and 18:0/22:6(n-3)cis PC molecules with 40 mol% cholesterol, and the same five pure phosphatidylcholine bilayers have been performed at 303 K. The simulation box of a lipid bilayer contained 96 phosphatidylcholines, 64 cholesterols, and 3840 water molecules (48 phosphatidylcholine molecules and 32 cholesterols per layer and 24 water molecules per phospholipid or cholesterol in each case). The lateral self-diffusion coefficients of the lipids in these systems and mass density profiles with respect to the bilayer normal have been analyzed. It has been found that the lateral diffusion coefficients of phosphatidylcholine molecules increase with increasing number of double bonds in one of the lipid chains, both in pure bilayers and in bilayers with cholesterol. It has been found as well that the lateral diffusion coefficient of phosphatidylcholine molecules of a lipid bilayer with 40 mol% cholesterol is smaller than that for the corresponding pure phosphatidylcholine bilayer.  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