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 15 毫秒
1.
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   

2.
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   

3.
4.
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   

5.
31P-NMR is used to characterize the phase behavior of phosphonolipids in both model and biological membranes. (1′,2′-Dipalmitoyl-sn-glyceryl)-2-aminoethylphosphonate gives rise to static chemical shift tensor elements (?87, 5 and 63 ppm) which differ considerably from those reported for the analogous phospholipid, 1,2-dipalmitoyl-sn-glycero-3-phosphoethanolamine (?81, ?20 and 105 ppm). Phosphonolipid, as well as a mixture of phosphonolipid and 1,2-dipalmitoyl-sn-glycero-3-phosphocholine, in aqueous dispersion gives rise to 31P spectra which may be interpreted in terms of lamellar structures. A mixture of phosphonolipid and egg phosphatidylethanolamine exhibits a bilayer-to-hexagonal phase transition with a concomitant decrease by one-half in the value of the 31P chemical shift anisotropies of both the phosphonate and phosphate resonances. The chemical shift anisotropy associated with phosphonolipid has been found to be consistently smaller than that observed for the analogous phospholipid. 31P-NMR spectra of total lipid extracts of Tetrahymena sp. indicate that both phospho- and phosphonolipids have a bilayer organization between ?20 and 20°C.  相似文献   

6.
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   

7.
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   

8.
Raman spectroscopic techniques have been used to construct phase diagrams for the binary phospholipid systems, DPPC-d62/DPPE and DPPC/DPPE (DPPC, dipalmitoyl phosphatidylcholine; DPPE, dipalmitoyl phosphatidylethanolamine). For the former, the half-width of the C-2H stretching modes of the deuterated component near 2100 cm?1 serves as an indicator of phospholipid fluidity. The phase behavior is described semi-quantitatively using regular solution theory with the following non-ideality parameters:
ρ0(1)=0.75kcal/mol and ρ0(s)=1.05 kcal/mol
The use of deuterated phospholipids as one component of a binary mixture permits direct evaluation of the conformation of a particular component in the mixture throughout the phase separation region. The approach is demonstrated with the help of a simple model correlating the half-width of the symmetric C-2H stretching mode with the fraction of DPPC-d62 hydrocarbon chains in the liquid crystalline state.The effect of chain perdeuteration on the phase behavior of DPPC with DPPE is evaluated by comparison of the phase diagram of the DPPC-d62/DPPE system with that of DPPC-DPPE. The latter has been constructed previously from both probe and calorimetric techniques, and is created from the Raman spectroscopic data using the I(11301100) ratio to characterize the transgauche population ratio in non-deuterated hydrocarbon chains. A reasonable fit to the phase behavior is obtained using:
ρ0(1)=0.85kcal/mol and ρ0(s)=0.90 kcal/mol
The similarities of the non-ideality parameters in the two phase diagrams indicate that the effect of perdeuteration on the phase behavior of DPPC is not extensive. The use of deuterated phospholipids as essentially unperturbed components of a model membrane system is justified.  相似文献   

9.
Synthesis and phase transition characteristics of aqueous dispersions of the homologous (12 : 0, 14 : 0, 16 : 0) diphosphatidylglycerols (cardiolipins) and phosphatidyldiacylglycerols are reported. Electron microscopy of the negatively stained aqueous dispersions reveals a characteristic lamellar structure suggesting that these phospholipid molecules are organized as bilayers in the aqueous dispersions. The phase transition temperature (Tm) and the enthalpy of transition (ΔH) increase monotonically with chain length in the cardiolipin and phosphatidyldiacylglycerol series; Tm for phosphatidyldiacylglycerol is higher than that for cardiolipin of the same chain-length. The transition temperatures for the enantiomeric sn-3,3- and sn-1,1-phosphatidyldiacylglycerol and for the diastereomeric, meso-sn-1,3-phosphatidyldiacylglycerol are approximately the same. The molar enthalpy for the transition of cardiolipin-NH4+ bilayers is approximately twice the value for the phosphatidylcholines of the same chain length, i.e., the molar enthalpy per acyl chain is approximately the same in the two systems. The transition temperatures for metal ion salts of C1 6-cardiolipin exhibit a biphasic dependence upon the unhydrated ionic radii, i.e. the highest Tm is observed for Ca2+- cardiolipin and decreases for the salts of ions with smaller and larger ionic radii than that of Ca2+. The lowest Tm is observed for Rb+-cardiolipin. Monovalent metal salts of cardiolipin exhibit two phase transitions. This effect may result from different conformational packing of the four acyl chains due to differences in metal-phosphate binding.  相似文献   

10.
The in vivo chain elongation of rat liver DNA following partial hepatectomy was studied using alkaline sucrose gradients. DNA made in 5 min was less than 4 × 107 daltons and that made in 30 min was heterodisperse and by 4 hr 75% of the DNA became larger than 1 × 109 daltons. Administration of 1-β-D-arabinofuranosyl cytosine (ara-C) 5 min after thymidine-3H injection inhibited the chain elongation, whereas if given 30 minutes after thymidine-3H pulse did not inhibit the chain elongation. Thus the in vivo chain elongation of rat liver DNA consists of at least two steps 1) a step sensitive to ara-C involving nucleotides addition and 2) the other insensitive to ara-C and probably involving ligation of polynucleotide chains.  相似文献   

11.
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   

12.
The crystal structures of a triclinic form (HPA1) and a monoclinic form (HPA2) of hexadecyl-2-deoxyglycerophosphoric acid monohydrate were determined by single crystal analysis. The unit cell dimensions for HPA1 are a = 4.75, b = 5.72, c = 44.36 A? and α = 91.0, β = 101.5, γ = 100.5° (P1) and for HPA2, a = 4.75, b = 5.72, c = 88.72 A? and γ = 100.8° (P21). In both structures the molecules are fully extended and pack tail-to-tail in bilayers with tilting (47°) hydrocarbon chains. In HPA2, however, the chain tilt alternatingly changes direction in adjacent bilayers, giving rise to a doubled unit cell which spans two bilayers. The dihydrogen phosphate groups interact by hydrogen bonds and are arranged in rows. Laterally between these phosphate rows the water molecules are accommodated producing a compact two-dimensional network of hydrogen bonds. The packing cross-section in the layer plane of the dihydrogen phosphate monohydrate group is 26.7 Å2 in both structures. The hydrocarbon chains pack according to the triclinic (T|) chain packing mode. In HPA2, however, the chain packing is somewhat less compact with accounts for a 2% increase in the molecular volume. In both structures the ether oxygen is accommodated into the hydrocarbon matrix without distortion of the chain packing.  相似文献   

13.
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   

14.
Moving the phosphatidylcholine group from the 3-to the 2-position in monolayers of distearoyl-sn-glycero-3-phosphocholine at the oil/water interface expands the surface pressure-area isotherm and markedly increases the surface pressure at which phase separation occurs with only a slight change in the monolayer surface density at the onset of the transition. This is interpreted in terms of a change in an ordering parameter in the solid-condensed state.  相似文献   

15.
Perfused rat hearts were treated with isoprenaline (10?6M) or ouabain (5.5 × 10?6M). The phosphate contents of troponin-I and myosin P light chains were established by radiolabelling with 32P; in the case of the light chains, direct chemical analysis of total and of specifically alkali-labile phosphate was also performed. Addition of isoprenaline caused phosphorylation of both troponin-I and myosin P light chains, reaching a maximum increment, after several minutes, of 1 mol/mol and 0.30 mol/mol, respectively. The Mg2+-ATPase activities, at saturating Ca2+ concentrations, of natural actomyosin isolated from treated hearts were significantly depressed, and an inverse correlation was established between the phosphate content of troponin-I and the Vmax[Ca2+] of this ATPase activity. The Ca2+ sensitivity of the Ca2+Mg2+-ATPase was also decreased. These changes were all reversed by an incubation permitting dephosphorylation of proteins by endogenous phosphatases.Treatment of hearts with ouabain caused no increment in troponin-I phosphorylation, but increased the P light chain phosphate content to a maximum of 0.30 mol/mol after some minutes. A positive correlation was evident between phosphate content of the light chains (in all experiments) and the maximum myosin Ca2+-ATPase activities. In addition, the Vmax[ATP] of the Ca2+Mg2+-ATPase of natural actomyosin was increased when light chain phosphorylation had occurred in the absence of troponin-I phosphorylation. P-light chain phosphorylation did not affect the Ca2+ sensitivity of Ca2+Mg2+-ATPase activity.We suggest that the effects of phosphorylation of troponin-I are to diminish thin filament sensitivity to Ca2+, and to decrease the efficiency of the transduction process along neighbouring actin monomers, such that the number of actin-myosin crossbridge interactions is decreased even in the presence of Ca2+ excess. Phosphorylation of P light chains of myosin has an activating effect on myosin Ca2+-ATPase activity, as well as on the rate of cross-bridge formation.  相似文献   

16.
Dispersions (50 wt% water) containing 95 mol% dipalmitoyl phosphatidylcholine/5 mol% deuterated cholesteryl palmitate (or stearate) were studied using 2H-NMR. Incorporation of ester into the phospholipid bilayer was found to be 0.5 mol% at 50°C. From the profile of 2H quadrupolar splitting vs. chain position, support for an average conformation resembling a ‘horseshoe’ within the bilayer is obtained. Quadrupolar relaxation times T2e of approx. 250 μs and approx. 850 μs are measured for cholesteryl palmitate-2,2-d2 and cholesteryl palmitate-16,16,16-d3, respectively, which are less than one-half those obtained for the corresponding positions in dipalmitoyl-d62 phosphatidylcholine. This is ascribed to a slower rate of motion of the ester chain and/or an extra, slow motion of the molecule.  相似文献   

17.
The antibiotic A23187 carries Ca2+ across Müller-Rudin membranes made from 1,2-dierucoyl-sn-glycero-3-phosphocholine and n-decane. The conductance of the membranes is not increased by the Ca2+-transport. The flux depends linearly on Ca2+ concentration and ionophore concentration (above pH 6). It increases with increasing pH, approximately by a factor of 4–5 between pH 6 and pH 8. Maximal Ca2+-fluxes of about 10?10mol · cm?2 · s?1 were found. A counter transport of H+ could not be detected.The complex formation between A23187 and Ca2+ in egg phosphatidylcholine vesicles was studied spectroscopically. The results are consistent with the formation of a 2 : 1 complex. Optical absorption measurements on single phosphatidylcholine membranes were used to calculate the concentration of membrane-bound ionophore A23187.  相似文献   

18.
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   

19.
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   

20.
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   

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