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.
Measurements of the reflectance of monoolein n-alkane and monoolein/squalene lipid bilayers have been made. The total thickness of the bilayer was calculated from the dependence of reflectance on the refractive index of the aqueous salt or sucrose solution surrounding the bilayer. The total thickness was then compared to the thickness of the hydrocarbon chain region as determined from capacitance measurements. From this comparison, we found that the thickness of each polar region of the bilayers in salt solutions was 0.5 ± 0.1 nm, independent of the hydrocarbon solvent used. When the aqueous solutions contained sucrose, each polar region was approx. 0.9 nm thick. When n-tetradecane and n-hexadecane were used as solvents, microlenses of solvent trapped in the monoolein bilayer increased the reflectance. After about one hour, the coalescence of microlenses into larger lenses allowed the reflectance of the bilayer alone to be measured. The use of reflectance to measure the thickness of monoolein bilayers appears to be consistent with other methods and to give useful information about the structure of lipid bilayers.  相似文献   

2.
Two-component lipid bilayers of dipalmitoylphosphatidylcholine and dimyristoylphosphatidylcholine were studied by measuring, ultrasonic velocity and absorption at 3 MHz. The phase diagram of the two-component lipid bilayers is discussed based upon the transition anomalies of the ultrasonic velocity as well as absorption, and it is suggested that this binary system has two critical points. The bulk modulus of lipid bilayers was determined from the ultrasonic velocity to be (2.2–3.0) × 1010dynecm2, whereas the bulk viscosity calculated from the absorption was 10–20 P except for the transition regions.  相似文献   

3.
The changes in leaflet thickness and surface area per lipid molecule as a function of added amount of n-alkane solvents have been studied in a number of lipid water systems by low angle X-ray diffraction techniques. The most probable site of accumulation of the n-alkane is identified as the middle of the leaflet as opposed to intercalation with lipid hydrocarbon chains. Adsorption of n-alkane depends on alkyl chain length and the organisation of the lipid hydrocarbon chains in the lipid water phases. Attention is drawn to the possible relationship between these results, the effect of n-alkanes on isolated lipid bilayers and their effects as anaesthetics.  相似文献   

4.
5.
Glycerol substitutes for water in multilamellar phosphatidylcholine liposomes in that the fluid spaces between bilayers, as well as their main transition temperatures, heat capacities, and ethalpies are very similar in water and in pure glycerol. One major difference is that the gel state phase of dipalmitoylphosphatidylcholine (DPPC) in glycerol consists of bilayers with fully interdigitated hydrocarbon chains. Interdigitated DPPC phases are also formed in ethylene glycol or in methanol (at low methanol content). In solutions of glycerol and water, the fluid spacing between bilayers is a function of mole fraction of glycerol Xg, reaching maximum values at Xg ≌ 0.1 for lipid in the liquid crystalline phase and at Xg ≌ 0.3 for the gel phase. These changes are explained in terms of a modification of the long-range Van der Waals attractive forces by glycerol.  相似文献   

6.
7.
Differential polarized phase fluorometry was used to quantify the rotational rate (R) and limiting anisotropy (r) of the membrane probe diphenylhexatriene (DPH) in solvents and lipid vesicles exposed to hydrostatic pressures ranging from 1 bar to 2 kbar. These measurements reveal the effect of pressure on the phase-transition temperatures of the phosphatidylcholine vesicles, and the effects of pressure on order parameter of the acyl side-chain region of the membranes, the latter as indicated by r. In addition to the well-known elevation of the transition temperature (Tc) with pressure, our results demonstrate that increased pressure restores the order of the bilayers to that representative of temperatures below the transition temperature. We also found that solvents which allowed free isotropic rotation of DPH at 1 bar no longer allowed free rotation when sufficiently compressed; moreover, the apparent DPH rotational rate increased with r. Pressure studies using both DPH and the charged DPH analogue, trimethylammonium DPH (TMA-DPH) indicated that the Tc of dipalmitoylphosphatidylcholine vesicles increased 23 K/kbar and an apparent volume change of 0.036 ml/mol lipid at the phase transition. Assuming, as has been proposed, that TMA-DPH is localized near the glycerol backbone region of the bilayers, these results indicate a similar temperature- and pressure-dependent phase transition in this region and the acyl side-chain region of the membrane.  相似文献   

8.
The effect of n-alkanes on the ultrastructure of lipid bilayers has been investigated using freeze-fracture and negative stain electron microscopy. It has been found that the morphology of bilayers containing the long alkane tetradecane is quite different from bilayers containing the short alkane hexane. The smooth fracture faces of gel and liquid crystalline state bilayers are unmodified by tetradecane. However, hexane dramatically alters the hydrophobic bilayer interior, producing large (20 to 50 nm) mounds and depressions in the fracture faces. The fracture steps in these multilayer preparations containing hexane are variable in thickness and often considerably wider than the corresponding fracture steps in multilayers which contain tetradecane or are solvent-free. Alkanes also modify the structure of the Pβ′ or ‘banded’ phase of phosphatidylcholine bilayers. The incorporation of tetradecane removes the banded structure from both the bilayer's hydrophilic surface, as viewed by negative staining, and the bilayer's hydrophobic interior, as viewed by the freeze-fracture technique. These results are consistent with X-ray diffraction data which imply that long alkanes are primarily located between adjacent lipid hydrocarbon chains in each monolayer of the bilayer, while short alkanes can partition into the geometric center of the bilayer between apposing monolayers.  相似文献   

9.
The structure and dynamics of phosphatidylcholine bilayers containing chlorophyll were studied by X-ray diffraction and absorption polarization spectroscopy in the form of hydrated orientated multilayers below the thermal phase transition of the lipid chains and by nuclear magnetic resonance in the form of single-wall vesicles above the thermal transition. Our results show that (a) chlorophyll is incorporated into the phosphatidylcholine bilayers with its porphyrin ring located anisotropically in the polar headgroup layer of the membrane and with its phytol chain penetrating in a relatively extended form between the phosphatidylcholine fatty acid chains in the hydrocarbon core of the mixed bilayer membrane and (b) the intramolecular anisotropic rotational dynamics of the host phosphatidylcholine molecules are significantly perturbed upon chlorophyll incorporation into the bilayer at all levels of the phosphatidylcholine structure. These dynamics for the host phosphatidtlcholine fatty acid chains are qualitatively different from that of the incorporated chlorophyll phytol chains on a 10?9 ? 10?10s time scale in the ideally mixed two-component bilayer.  相似文献   

10.
The major capsid protein of M13 bacteriophage is incorporated at each stage of infection into the host plasma membrane with its amino terminus exposed on the outer surface. Purified M13 coat protein is incorporated with the same asymmetry into synthetic phosphatidylcholine vesicles formed near the Tm of the lipid by a cholate dilution technique. We now report that the lipid in the pre-dilution mixture exists as mixed micelles of uniform size. Prior to dilution, the coat protein is present in at least two states of aggregation, both of which behave similarly in the model membrane assembly reaction. No detectable lipid-protein interaction occurs prior to dilution. Upon dilution there is rapid production of small closed vesicles and coat protein is converted to a chymotrypsin-resistant form, presumably reflecting its incorporation into these vesicle bilayers. Formation of large (>6000 A? diameter) vesicles occurs slowly with preservation of coat protein asymmetry and internal volume. A model for this assembly reaction is proposed.  相似文献   

11.
When human red cells are treated with the mercurial sulfhydryl reagent, p-chloromercuribenzene sulfonate, osmotic water permeability is suppressed and only diffusional water permeability remains (Macey, R.I. and Farmer, R.E.L. (1970) Biochim. Biophys. Acta 211, 104–106). It has been suggested that the route for the remaining water permeation is by diffusion through the membrane lipids. However, after making allowance for the relative lipid area of the membrane, the water diffusion coefficient through lipid bilayers which contain cholesterol is too small by a factor of two or more. We have measured the permeability coefficient of normal human red cells by proton T1 NMR and obtained a value of 4.0 · 10?3 cm · s?1, in good agreement with published values. In order to study permeation-through red cell lipids we have perturbed extracted red cell lipids with the lipophilic anesthetic, halothane, and found that halothane increases water permeability. The same concentration of halothane has no effect on the water permeability of human red cells, after maximal pCMBS inhibition. In order to compare halothane mobility in extracted red cell membrane lipids with that in red cell ghost membranes, we have studied halothane quenching of N-phenyl-1-naphthylamine by equilibrium fluorescence and fluorescence lifetime methods. Since halothane mobility is similar in these two preparations, we have concluded that the primary route of water diffusion in pCMBS-treated red cells is not through membrane lipids, but rather through a membrane protein channel.  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.
Studies of the influence of fatty acids, which were the component of intestinal mucosal lipids, on the permeability of several drugs across bilayer lipid membranes generated from egg phosphatidylcholine and intestinal lipid have been pursued. The permeability coefficients of p-aminobenzoic acid, salicylic acid and p-aminosalicylic acid (anionic-charged drug) increased when fatty acids such as lauric, stearic, oleic, linoleic and linolenic acid were incorporated into the bilayer lipid membranes generated from phosphatidylcholine. In the presence of methyl linoleate and oleyl alcohol, no enhancing effect on p-aminobenzoic acid transfer was obtained. The effect of fatty acids was more marked at pH 6.5 than at pH 4.5. In contrast, upon the addition of fatty acids to intestinal lipid membranes which originally contained fatty acids, the permeability coefficient of p-aminobenzoic acid tended to decrease, though the permeability through intestinal lipid membranes was larger than that of phosphatidylcholine membranes. The permeability of p-aminobenzoic acid across bilayer lipid membranes from intestinal phospholipids was significantly decreased to about equal that of phosphatidylcholine membranes, and reverted to the value of intestinal lipid membranes when fatty acids were added to intestinal phospholipids. It seemed reasonable to assume that free fatty acids in the intestinal neutral lipid fraction could contribute to the increase in the permeability of p-aminobenzoic acid. On the basis of above results, possible mechanisms for good absorbability of weakly acidic drugs from the intestine are discussed.  相似文献   

14.
Planar bilayer membranes were formed from monolayers of a series of monounsaturated monoglycerides and lecithins. The hydrocarbon thickness of these membranes, as calculated from the electrical capacity, increases with the length of the fatty acid chain. The specific capacity of monoolein bilayers was found to be 0.745 μF/cm2 which is nearly twice that of a monoolein black film made in the presence of decane, but is close to that obtained after freezing out the solvent from the black film. The hydrocarbon thickness of the bilayer, as calculated with a dielectric constant of 2.1, is considerably less than twice the length of the extended hydrocarbon chain of the monoglyceride.The specific capacity (Cm) of bilayers made from monoolein monolayers showed a negligible voltage dependence, whereas the Cm increased significantly at a voltage of 150 mV in the case of Mueller-Rudin-type monoolein films with n-decane as a solvent.  相似文献   

15.
16.
The ionization of fatty acids, fatty amines and N-acylamino acids incorporated in phosphatidylcholine single-walled vesicles has been measured. The guest molecules have been specifically enriched with 13C and titrated by using NMR spectroscopy. The apparent pKa of fatty acids in phosphatidylcholine bilayers is 7.2–7.4 and those of fatty amines are approx. 9.5. These pKa values depend on many different parameters related to the structure of the lipid/ solution interface, to the composition of the aqueous medium and to the localization of the ionizable groups. A special sensitivity to the ionic strength and to the surface charge has been found. A positive surface charge decreases the pKa value whereas a negative one increases it, the total range of variation being 2.5–3 units. In a qualitative macroscopic interpretation, it is proposed that pKa is essentially determined by the low polarity of the lipidic matrix.  相似文献   

17.
Sarcoplasmic reticulum membranes from skeletal muscle contain a proteolipid (Mr ≈ 12 000) which reduces both the nonspecific ion and water permeabilities of artificial planar phospholipid bilayers. The proteolipid does not show any ionophoric effect or specific pore formation for Ca2+. The a.c. capacitance of the bilayers is unaffected whereas the refractive index is increased by the presence of proteolipid. The results support the view that the proteolipid interacts with the phospholipids in the bilayer interior and causes a condensation in the packing of the alkyl chains.  相似文献   

18.
Neutron diffraction experiments on selectively deuterated lipids provide a new method of determining to a segmental resolution the mean conformation of a lipid molecule as projected along the bilayer normal, despite the high amount of disorder that exists in these bilayers. In addition, a time-averaged picture of the extent of the positional fluctuations of the individual segments in this direction can be given. This is demonstrated for a multilamellar system of bilayers of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine. In this paper the head group region of the molecule is examined and this carries the zwitterionic phosphocholine group that determines the electrostatic interaction in the bilayer. Samples deuterated at four different positions in the head group region were measured as oriented samples at 6% (ww) water content at 20 °C (Lβ′ phase) and at 10% (ww) at 70 °C (Lα phase) and as unsonicated dispersions with 25% (ww) water at 28 °C (Lβ′ phase) and 50 °C (Lα phase). From the oriented samples, reflections up to ten orders, and from the powder type samples only four orders, were collected. The derived structure factors for the deuterated segments were fitted assuming a Gaussian distribution of the segments along the bilayer normal. The mean label position was determined for each label under different conditions of water content and temperature with a precision of better than ± 1 ångström in most cases. The data clearly show that the average orientation of the zwitterionic phosphocholine group is almost parallel to the membrane surface in the gel state (Lβ′) as well as in the liquid crystalline state (Lα). It is interesting to note that in a recent dielectric investigation on this multilamellar system at 25% (ww) water content the same mean orientation of the dipole was found (Shepherd &; Büldt, 1978).  相似文献   

19.
20.
Dipalmitoylphosphatidylcholine (DPPC) dispersed in perdeuterated glycerol was investigated in order to determine the effects on the Raman spectra of hydrocarbon chain interdigitation in gel-phase lipid bilayers. Interdigitated DPPC bilayers formed from glycerol dispersions in the gel phase showed a decrease in the peak height intensity I2850/I2880 ratio, for the symmetric and asymmetric methylene CH stretching modes, respectively, as compared to non-interdigitated DPPC/water gel-phase dispersions. The decrease in this spectral ratio is interpreted as an increase in chain-chain lateral interactions. Spectra recorded in the 700–740 cm?1 CN stretching mode region, the 1000–1200 cm?1 CC stretching mode region and the 1700–1800 cm? CO stretching mode region were identical for both the interdigitated and non-interdigitated hydrocarbon chain systems. At low temperatures the Raman peak height intensity ratios I2935/I2880 were identical for the DPPC/glycerol and DPPC/water dispersions, indicating that this specific index for monitoring bilayer behavior is insensitive to acyl chain interdigitation. The increase, however, in the change of this index at the gel-liquid crystalline phase transition temperature for the DPPC/glycerol dispersions implies a larger entropy of transition in comparison to the non-interdigitated DPPC/water bilayer system.  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