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.
1-Palmitoyllysophosphatidylcholine has been mixed in equimolar amounts with specifically deuterated palmitic acid and the structural properties of the lipid/water phase have been studied by 2H- and 31P-nuclear magnetic resonance. The order profile of the free palmitic acid is very similar to that of the parent compound 1,2-dipalmitoyl-sn-glycero-3-phosphocholine at temperatures above the gel-to-liquid crystal phase transition. The bending of the sn-2 chain which is typical for diacyl lipids is not observed for the free palmitic acid. The mixture of lysolipid and palmitic acid exhibits well-defined quadrupole splittings even at temperatures below the gel-to-liquid crystal phase transition. Hence it is possible for the first time to establish an order profile in the gel-state of the lipid bilayer phase. Between carbon atoms 5 to 12 the palmitic acid chain is found to assume the extended all-trans conformation with a very small contribution from gauche defects. Towards the methyl terminal a distinct increase in the gauche probability can be noted. The motion of the phosphocholine headgroup was also studied by 2H- and 31P-NMR using selectively deuterated 1-palmitoyllysophosphatidylcholine. The headgroup has a considerably larger motional freedom in the mixture of lysolipid and palmitic acid than in 1,2-dipalmitoyl-sn-glycero-3-phosphocholine. In addition, the average headgroup conformations are also different in the two systems.  相似文献   

2.
We have examined the thermal characteristics and barrier properties of vesicles formed from six analogues of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC). These analogues differ from DPPC in that the glycerol backbone has been replaced by each of the diastereoisomeric cyclopentane-1,2,3-triols. All of these compounds have main gel to liquid-crystal phase transition temperatures within 5 Kelvin of DPPC and four possess comparable enthalpies and entropies of transition. For two of the analogous, however, the values of the enthalpy and entropy of transition are more than double that of DPPC. The permeability characteristics and organization (as measured by diphenylhexatriene fluorescence depolarization) of vesicles formed from these two compounds suggest that their large transition enthalpy and entropy result from either a reorganization of the polar head group region during the transition or interdigitation of the acyl chains of opposing monolayers.  相似文献   

3.
Aqueous dispersions of monovalent and divalent cation salts of O-(1,2-dipalmitoyl-sn-glycero-3-phosphoryl) cholesterol form multilamellar vesicles as shown by freeze-fracture electron microscopy, by electron micrographs of the negatively stained liposomes, and by swelling curves of liposomes in hypoosmotic medium. Differential scanning calorimetry reveals that aqueous dispersions of divalent metal salts of O-(1,2-dipalmitoyl-sn-glycero-3-phosphoryl)-cholesterol undergo a characteristic thermotropic phase transition with a relatively large cooperative unit (n > 250 for the calcium salt). In contrast, monovalent cation salts of O-(1,2-dipalmitoyl-sn-glycerol-3-phosphoryl)cholesterol do not show a thermotropic phase transition under comparable conditions. The molecular area of O-(1,2-dipalmitoyl-sn-glycero-3-phosphoryl)cholesterol in a monolayer is the same in the presence and absence of Ca2+, and is virtually equal to the area of an equimolar mixture of dipalmitoyl phosphatidic acid and cholesterol. To account for the novel state induced by Ca2+ on aqueous dispersions of O-(1,2-dipalmitoyl-sn-glycero-3-phosphoryl)cholesterol (i.e., bilayer organization and highly cooperative phase transition), a linear array model is proposed in which Ca2+ bridges adjacent arrays of O-(1,2-dipalmitoyl-sn-glycero-3-phosphoryl)cholesterol molecules, thus freezing the acyl chains in their normal state. One of the main corollaries of the model is that the cooperative unit for a thermotropic phase transition is essentially one-dimensional, rather than a two-dimensional matrix. O-(1,2-Dipalmitoyl-sn-glycero-3-phosphoryl)cholesterol is proposed as an orientationally and conformationally restricted analog of glycerophospholipid plus cholesterol in bilayers.  相似文献   

4.
Incubations of rat liver inner mitochondrial membranes with liposomes prepared from 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) and cholesterol resulted in a considerable enrichment of the cholesterol composition of these membranes. This enrichment is not accompanied by an alteration in the membrane phospholipid content or fatty acid composition. The exogenous cholesterol appears to be integrated into the membrane structure because it has effects consistent with the known properties of this sterol in other natural and artificial membrane systems.Differential scanning calorimetry on both intact membranes and extracted lipids showed that as the ratio of cholesterol to phospholipid was increased, the endotherm corresponding to the lipid phase transition was reduced. Freeze-fracture electron microscopy of the native membranes showed that intramembranous particles are randomly distributed above the phase transition temperature. Below this temperature large smooth areas, believed to correspond to lipid in the gel state from which proteins have been excluded, can be observed. In the presence of high concentrations of cholesterol the fracture faces observed below the lipid transition temperature show no regions of phase segregation, an observation consistent with previous studies using pure lipids where cholesterol was observed to prevent the lipid undergoing a cooperative phase transition.The results are discussed in terms of the observed low concentrations of cholesteorl in normal liver inner mitochondrial membranes and the distribution of cholesterol within the liver cells.  相似文献   

5.
The first application of infrared difference spectroscopy to the study of a natural biological membrane is described. Perdeuterated palmitic acid was incorporated biosynthetically into the lipids of the plasma membrane of Acholeplasma laidlawii and the temperature-induced structural rearrangement of the endogenous lipids monitored via their C2H vibrational modes. Changes in infrared parameters were studied between 0 and 50°C and contrasted with those occurring in the model membrane system of 1,2-diperdeuteropalmitoyl-sn-glycero-3-phosphocholine. The phase transition of the biomembrane occurs over a 20°C range with the temperature of the maximum rate of change of absorbance coinciding with that of the sharp phase transition of the model membrane.  相似文献   

6.
The dielectric dispersion in the MHz range of the zwitterionic dipolar phosphocholine head groups has been measured from 0–70°C for various mixtures of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) and cholesterol. The abrupt change in the derived relaxation frequency f2 observed for pure DPPC at the gel-to-liquid crystalline phase transition at 42°C reduces to a more gradual increase of frequency with temperature as the cholesterol content is increased. In general the presence of cholesterol increases the DPPC head group mobility due to its spacing effect. Below 42°C no sudden changes in f2 are found at 20 or 33 mol% cholesterol, where phase boundaries have been suggested from other methods. Above 42°C, however, a decrease in f2 at cholesterol contents up to 20–30 mol% is found. This is thought to be partly due to an additional restricting effect of the cholesterol on the number of hydrocarbon chain conformations and consequently on the area occupied by the DPPC molecules.  相似文献   

7.
Some properties of monolayers of 1-palmitoyl-2-oleoyl-sn-glycero-3-phospho-rac-glycerol (POPG) alone or of POPG in mixtures with 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) have been measured near 35°C during dynamic compression and expansion at 3.6 cm2·s?1. (2) The mean values of minimum surface tension (corresponding to maximum surface pressure) which could be obtained with pure POPG monolayers at high compression ranged from 15 to 18 mN·m?1 in the presence of Na+, Ca2+ or low pH (2.0) in the subphase. (3) The presence of Ca2+ or low pH in the subphase increased the collapse plateau ratios obtained on cyclic compression. This might represent enhanced respreading into the monolayer of pure POPG from a collapsed form during reexpansion of the surface. (4) Monolayers containing 10% or 30% POPG and 90% or 70% DPPC could be compressed to surface tensions approaching zero. (5) In such mixed monolayers, 10% or 30% POPG did not appear to enhance respreading, as measured by collapse plateau ratios, in the presence of Na+ or Ca2+ in the subphase.  相似文献   

8.
A complete titration of phosphatidic acid bilayer membranes was possible for the first time by the introduction of a new anaologue, 1,2-dihexadecyl-sn-glycerol-3-phosphoric acid, which has the advantage of a high chemical stability at extreme pH values. The synthesis of this phosphatidic acid is described and the phase transition behaviour in aqueous dispersions is compared with that of three ester phosphatidic acids; 1,2-dimyristoyl-sn-glycerol-3-phosphoric acid, 1,3-dimyristoylglycerol-2-phosphoric acid and 1,2-dipalmitoyl-sn-glycerol-3-phosphoric acid.The phase transition temperatures (Tt) of aqueous phosphatidic acid dispersions at different degrees of dissociation were measured using fluorescence spectroscopy and 90° light scattering. The Tt values are comparable to the melting points of the solid phosphatidic acids in the fully protonated states, but large differences exist for the charged states.The Tt vs. pH diagrams of the four phosphatidic acids are quite similar and of a characteristic shape. Increasing ionisation results in a maximum value for the transition temperatures at pH 3.5 (pK1). The regions between the first and the second pK of the phosphatidic acids are characterised by only small variations in the transition temperatures (extended plateau) in spite of the large changes occurring in the surface charge of the membranes. The slope of the plateau is very shallow with increasing ionisation. A further decrease in the H+ concentration results in an abrupt change of the transition temperature. The slope of the Tt vs. pH diagram beyond pK2 becomes very steep. This is the  相似文献   

9.
Using differential scanning calorimetry and freeze fracture electron microscopy interactions were studied between lipids and a spectrin · action complex isolated from human erythrocyte membranes. With dispersions of 1,2-dimyristoyl-sn-glycero-3-phosphocholine, 1,2-dimyristoyl-sn-glycero-3-phosphoglycerol and mixtures of these two compounds, which for experimental reasons were chosen as the lipid counterpart, such an interaction could clearly be deduced from changes in the temperature and the enthalpy of the phase transition. Furthermore it was demonstrated that the interaction with this membrane protein protects the bilayer against the action of Ca2+ and Mg2+ and prevents fusion of lipid vesicles which easily occurs in some of the systems when divalent ions were added to the pure lipid vesicles.  相似文献   

10.
Dipalmitoyl-3-sn-phosphatidylcholine is specifically deuterated at the C-2 position of the fatty acyl chains. Using deuterium magnetic resonance it is then possible to probe the chain conformation in the vicinity the polar head groups. Three separate quadrupole splittings are observed for bilayers of 1,2[2′-2H2] palmitoyl-3-sn-phosphatidylcholine, indicating that the two chains behave differently. The synthesis of phosphatidylcholines each deuterated in only one chain allows the assignment of the three resonances. It is concluded that the beginnings of the two chains have orientations parallel and perpendicular to the bilayer normal. The data further suggest the possibility of two long-lived conformations of the glycerol constituent.  相似文献   

11.
In order to help bridge the conceptual gap between experimental data on chains of phospholipid molecules and their microscopic organization, a theoretical model has been proposed in a preceding paper. The intentions associated with the new theory were to describe a model able to reproduce accurately the experimental data. This capability is essential to monitor some of the mechanisms behind the physical data. The results presented here show first that, provided a suitable fitting of the phenomenological parameters entailed in the model, the theory indeed gives good agreement with experimental data (2H-NMR, neutron scattering, calorimetry) obtained for a dipalmitoyl-3-sn-phosphatidylcholine bilayer. This property of the model is then specifically used to describe the nature of the perturbing effects of local anaesthetics and cholesterol on the organization of the acyl chains and to correlate these effects with the experimental data. Finally the theoretical model is used to supplement experimental data by describing the acyl chain organization in terms of the most probable spectrum of chain conformations. Predictions are made about the one-, two- and three-dimensional mean spatial characteristics of the acyl chains.  相似文献   

12.
A new assay system of phospholipid exchange activities is described. The exchange activities were quantitated by measuring the stimulation of phospholipid transfer between two separate populations of liposomes, which contained, as the major constituents, phosphatidylcholine, phosphatidylethanolamine, phosphatidylinositol, sphingomyelin, and cholesterol in molar ratios of 6: 2: 1: 1: 5. One population of the liposomes was made reactive to concanavalin A by the incorporation of 1.8 mol% α-d-mannosyl-(1 → 3)-α-d- mannosyl-sn-1,2-diglyceride from Micrococcus lysodeikticus. The concanavalin A-reactive liposomes, a phospholipid donor, were doubly labelled with [6-3H]galactosylglucosyl ceramide and that class of 32P-labelled phospholipids whose exchange was being measured. The 3H-labelled glycolipid served as a non-exchangeable reference marker. The other population of the liposomes, a phospholipid acceptor, was concanavalin A nonreactive. These two populations of liposomes were incubated with the cytosol protein of rat liver in a total volume of 0.2 ml.After the incubation, two different procedures were used to separate the two liposomal populations. In one procedure concanavalin A was added to agglutinate the reactive liposomes; the flocculated lectin-liposome complex was separated from the non-reactive liposomes by brief centrifugation. In the other procedure the reactive liposomes were trapped by binding to concanavalin A covalently coupled to Sepharose 2B; the complex was separated from the nonreactive liposomes by filtration through a filter paper under suction. In both assay procedures the amount of phospholipid transferred from the donor to the acceptor liposomes was calculated from the decrease of 32P/3H ratio of the concanavalin A-reactive liposomes during the incubation. By the assay system it is possible to determine phosphatidylcholine and phosphatidylinositol exchange activities in 100 μg of rat liver cytosol protein.  相似文献   

13.
The interaction of |CnH2n+1N+(CH3)3| · I? (n = 3, 6, 9, 12, 14, 16 or 18) with egg-yolk phosphatidylcholine-water dispersions has been studied by 31P-NMR spectroscopy. It is shown that the effective anisotropy of 31P chemical shift (?Δσeff) of the lamellar phospholipid liquid-crystalline phase Lα increases with increasing concentration and alkyl chain length of the drug. Addition of |C6H13N+(CH3)3| ·I ? or |C9H19N+(CH3)3I? to the phospholipid-water dispersion at a molar ratio ammonium salt:phospholipid > 0.8 induces in the dispersion a structure with an effective isotropic phospholipid motion. This structure is unstable and slowly transforms into the hexagonal phase. These effects have not been observed in phospholipid-water dispersions mixed with the ammonium derivatives with the longer alkyl chains n  12, 14, 16 or 18. It is proposed that these results might explain the effects of the investigated drugs on the nerve, muscle and bacterial cells.  相似文献   

14.
One of the unique features of the chromaffin granule membrane is the presence of about 17 mol% lysophosphatidylcholine. Lysophosphatidylcholine isolated from the granules could be degraded by approx. 94% by lysophospholipase. This result is consistent with chemical analyses data showing that about 9% of this lysophospholipid is 1′-alkenyl glycerophosphocholine.The localization of the acylglycerophosphocholine in the chromaffin granule membrane was studied by using pure bovine liver lysophospholipases. In intact granules only about 10% of the total lysophosphatidylcholine was directly available for enzymic hydrolysis. In contrast, when granule membranes (ghosts) were treated with lysophospholipases approx. 60% of the lysophosphatidylcholine was deacylated. These values did not increase after pre-treatment of intact granules or ghosts with trypsin. Added 1-[1-14C]palmitoyl-sn-glycero-3-phosphocholine did not mix with the endogenous lysophosphatidylcholine pool(s) and remained completely accessible to added lysophospholipases.  相似文献   

15.
The unnatural amino-alcohol, N-isopropylethanolamine, is incorporated into a phospholipid by monolayers of L-M fibroblasts. This phospholipid was identified as 1,2-diacyl-sn-glycero-3-phosphoisopropylethanolamine by using chemical and enzymatic procedures combined with thin-layer and gas-liquid chromatography. Since the phospho-N-isopropylethanolamine moiety is removed by phospholipase C, the stereochemistry of the phospholipid analog is identical to naturally occurring phosphoglycerides. Incubation of cells in 10 mM N-isopropylethanolamine inhibited the incorporation of [14C]choline and [14C]ethanolamine into phospholipids and stimulated the incorporation of [1-14C]palmitic acid and [1-14C]hexadecanol into triacylglycerols and alkyldiacylglycerols. These results indicate that N-isopropylethanolamine affects glycerolipid synthesis at the diradylglycerol branch point.  相似文献   

16.
[2-14C,2-3H]1,2-Heptadecanediol administered intracerebrally to 18-day-old rats was found to be incorporated as such into phospholipid Chemical and enzymic degradation of the choline phosphatide fraction showed that an appreciable amount of radioactivity was associated with 2-O-acyl heptadecanediol phosphoryl choline.  相似文献   

17.
Sonicated emulsions of egg phosphatidylcholine containing either [14C]-dipalmitoyl phosphatidylcholine (diester-PC) or two metabolically inert analogs. [14C]-1-octadecyl-2-hexadecyl-sn-glycero-3-phosphocholine (diether-PC) and [3H]-2-tetradecyloctadecano-(1)-phosphocholine (dialkyl-PC) were injected into the vitreous of the eye of adult rabbits. After 1–40 days, radioactivities were measured in the stations of the optical pathway, and the identities of the labelled lipids arrived at the superior colliculus were ascertained by thin-layer chromatography and treatment with phospholipase A2, with the following results: (1) phosphatidylcholine and its analogs were taken up from the vitreous by the retina at similar rates: (2) all three lipids were transported in the optic nerve axons at similar rates (‘fast’). They reached maximal concentration in the superior colliculus some 20 days after injection: (3) phosphatidylcholine travelled from vitreous to superior colliculus as the intact molecule: (4) maximal accumulation of the two analogs in the superior colliculus reached only about 1 per cent of that of phosphatidylcholine. The results suggest that the vehicles of fast axonal transport can pick up intact phospholipid molecules originating in the ganglionic cell plasma membrane (and, likely, from other cellular compartments). The packaging process is promoted by the presence of carboxyl ester groups in the phospholipid; this fact suggests the involvement of ganglionic phospholipid transfer protein with specificity for these groups.  相似文献   

18.
We measured the 31P[1H] Nuclear Overhauser Effect (NOE) as a function of temperature and of 1H irradiation frequency, the linewidth Δν12 as a function of temperature and the relaxation time T1 above and below the thermal transition temperature, of the 31P-NMR signal in sonicated liposomes of 1,2-dimiristoyl-3-sn-phosphatidylcholine (DMPC), 1,2-dipalmitoyl-3-sn-phosphatidylcholine (DPPC) and 1,2-dimiristoyl-3-sn-phosphatidylcholine (DSPC). The same measurements were repeated in the presence of high molecular weight dextrans. They strongly reduce the NOE and produce longer relaxation times T1. According to the current models, we were able to evaluate, in the different situations, the correlation time of the internal motion τG and the distance r between interacting groups in the region of the polar head groups. While the first parameter changes abruptly through the phase transition and under the effect of dextrans, the latter does not appear modified in any case. These results are discussed in terms of a conformational change of the phosphocholine head groups.  相似文献   

19.
The hydration properties of Escherichia coli lipids (phosphatidylglycerol, phosphatidylethanolamine) and synthetic 1,2-dioleoyl-sn-glycero-3-phosphocholine in H2O/2H2O mixtures (9:1, v/v) were investigated with 2H-NMR. Comparison of the 2H2O spin lattice relaxation time (T1) as a function of the water content revealed a remarkable quantitative similarity of all three lipid-H2O systems. Two distinct hydration regions could be discerned in the T1 relaxation time profile. (1) A minimum of 11–16 water molecules was needed to form a primary hydration shell, characterized by an average relaxation time of T1 ≈ 90 ms. (2) Additional water was found to be in exchange with the primary hydration shell. The exchange process could be described in terms of a two-site exchange model, assuming rapid exchange between bulk water with T1 = 500 ms and hydration water with T1 = 80–120 ms. Analysis of the linewidth and the residual quadrupole splitting (at low water content) confirmed the size of the primary hydration layer. However, each lipid-water system exhibited a somewhat different linewidth behavior, and a detailed molecular interpretation appeared to be preposterous.  相似文献   

20.
The effect of the normal alcohols (up to C = 9) and three clinically used anaesthetics, on the crystalline-liquid crystalline phase transition in 1,2-dihexadecyl-sn-glycero-3-phosphorylcholine have been studied. A one-degree depression was produced by a 4.4% concentration in the membrane of n-octanol and n-nonanol agreeing well with the value calculated from the temperature and enthalpy of the transition. It is also shown that the relationship between the partition coefficient P and the water solubility S (P · S = 2), holds for the solutes investigated here. The experimental method described offers a simple way of assessing the anaesthetic potency of a wide range of compounds.  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