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.
According to the liquid hydrocarbon model, the lipid bilayer is viewed simply as a thin slice of bulk hydrocarbon liquid. This allows the water permeability of the bilayer to be calculated from bulk properties. In this paper the prediction of the liquid hydrocarbon model is compared with the known water permeability coefficient of the glycerol monoolein/n-hexadecane bilayer (Fettiplace, R. (1978) Biochim. Biophys. Acta 513, 1–10). As the alkyl chain of glycerol monoolein is equivalent to 8-heptadecene, the water permeability coefficient of 8-heptadecene/n-hexadecane mixtures was measured for temperatures between 20 and 35°C. The mole fraction of n-hexadecane in the bulk liquid was chosen at each temperature to match the known mole fraction of n-hexadecane in the bilayer (White, S. (1976) Nature 262, 421–422). The predicted water permeability coefficient agrees with the measured value at 32°C but is 40% above the measured value at 20°C. The apparent activation energy predicted by the liquid hydrocarbon model is 9.0 ± 0.3 kcal/mol, while the measured value is 14.2 ± 1.0 kcal/mol. The failure of the liquid hydrocarbon model probably results from a different molecular organization of the hydrocarbon chains in the bilayer and in the bulk liquid.  相似文献   

2.
T1 nuclear relaxation measurements of 1H and 17O of water have been applied to study the kinetics of the diffusional transport of water across the cytoplasmic cell membrane of Dunaliella salina and Dunaliella bardawil. The water permeability coefficients at 25°C were found to be 1.5·10−3 cm/s and 1.8·10−3 cm/s, respectively, with an activation energy of 3.7 kcal/mol. The results indicate that the cell membrane of Dunaliella exhibits high diffusional permeability to water, similar in magnitude to that found for other cells and model membranes, and a relatively low activation energy. This regularity is in contrast to the exceptionally low glycerol permeability of the membrane (Brown, F.F., Sussman, I., Avron, M. and Degani, H. (1982) Biochim. Biophys. Acta 690, 165–173).  相似文献   

3.
In this paper we present the results of a Monte Carlo study of the effects of protein, cholesterol, bilayer curvature, and mobility on the chain order parameters of a lipid layer. The Monte Carlo method used is identical to the version developed earlier (Scott, Jr., H.L. (1977) Biochim. Biophys. Acta 469, 264–271). Simulations of protein and cholesterol effects are accomplished by insertion of a rigid stationary cylinder into the lipid matrix. The protein studies show the presence of boundary lipid (Jost, P., Griffith, O.H., Capaldi, R.H. and Vanderkooi, G. (1973) Biochim. Biophys. Acta 311, 141–152). The effect of cholesterol is dependent upon the length of the lipid hydrocarbon chains relative to the cholesterol depth of penetration. Our computer studies of bilayer curvature show the manner in which this curvature disrupts chain packing and are consistent with experimental results (Chrzeszczyk, A., Wishnia, A. and Springer, C.S. (1977) Biochim. Biophys. Acta, 470, 161–171). We also find that restricting lateral motion in chains, the simplest manner in which head group interactions can affect hydrocarbon chain order, does not measurably alter the order parameters. We argue that this provides some support for an earlier hypothesis by Scott (Scott, Jr., H.L. (1975) Biochim. Biophys. Acta 406, 329–346) regarding head group-chain interaction in monolayer experiments.  相似文献   

4.
Solid core liposomes with encapsulated colloidal gold particles were prepared through four major steps: Preparation of prevesicles with encapsulated solid cores of agarose-gelatin by emulsification of agarose-gelatin sol in organic solvent containing emulsifiers followed by cooling. Extraction of lipophilic components from prevesicles to obtain microspherules of agarose-gelatin. Introducing colloidal gold particles into microspherules and coating with protein molecules. Encapsulation of colloidal gold-bearing microspherules with the modified organic solvent spherule evaporation method for preparation of liposomes (Kim et al. (1983) Biochim. Biophys. Acta 728, 339-348 and Kim et al. (1984) Biochim. Biophys. Acta 812, 793-801). Electron micrographs showed that if liposomes were prepared by using a lipid mixture containing dioleoylphosphatidylcholine/cholesterol/dioleoylphosphatidylglycerol/tri olein (molar ratio 4.5:4.5:1:1), there was only a single continuous bilayer membrane for each solid core liposome. However, if no triolein was added to the lipid mixture, it would cause the formation of multilamellar liposomes. In both cases, there were hundreds to thousands of colloidal gold particles within each solid core liposome.  相似文献   

5.
The interaction of surfactants with the vesicle membrane of the negatively charged lipid, dilauroylphosphatidic acid, was investigated through their effect on the gel-to-liquid-crystalline phase transition of the lipid bilayer. Three types of surfactants (anionic, cationic and non-ionic) with different hydrocarbon chain length were examined. (i) Anionic sodium alkylsulfates affected the phase transition temperature, Tm, only weakly. (ii) Non-ionic alkanoyl-N-methylglucamides decreased Tm monotonously with increasing concentration. The depression of Tm induced by these surfactants was analyzed by applying the van't Hoff model for the freezing-point depression, and the partition coefficients of the surfactants between bulk water and lipid membrane were estimated. (iii) Cationic alkyltrimethylammonium bromides affected Tm in a complex manner depending on the hydrocarbon chain length of the surfactants. Octyl-/tetradecyl-trimethylammonium bromide depressed/elevated Tm monotonously with increasing concentration, whereas the change in Tm induced by decyl- and dodecyltrimethylammonium bromides was not monotonous but biphasic. This complex behavior of the phase transition temperature was well explained, based on the statistical mechanical theory presented by Suezaki et al. (Biochim. Biophys. Acta, 818 (1985) 31-37), which takes into account the interaction between surfactant molecules incorporated in the lipid membrane.  相似文献   

6.
Spectrin-depleted inside-out vesicles (IOV's) prepared from human erythrocyte membranes were characterized in terms of size, ground permeability to hydrophilic nonelectrolytes and their sensitivity to modification by SH reagents, DIDS and trypsin. IOV's proved to have the same permeability of their lipid domain to erythritol as native erythrocytes, in contrast to resealed ghosts (Klonk, S. and Deuticke, B. (1992) Biochim. Biophys. Acta 1106, 126-136 (Part I in this series)), which have a residual leak. On the other hand, IOV's have a slightly elevated permeability for mannitol and sucrose, nonelectrolytes which are almost (mannitol) or fully (sucrose) impermeant in the native membrane. These increased fluxes, which have a high activation energy and can be stimulated by phloretin, are, however, also much smaller than the corresponding leak fluxes observed in resealed ghosts. In view of these differences, formation of IOV's can be concluded to go along with partial annealing of barrier defects persisting in the erythrocyte membrane after preparation of resealed ghosts. Oxidation of SH groups of the IOV membrane by diamide produces an enhancement of permeability for hydrophilic nonelectrolytes which is much less pronounced than that induced by a similar treatment of erythrocytes or ghosts (Klonk, S. and Deuticke, B. (1992) Biochim. Biophys. Acta 1106, 126-136 (Part I in this series)). Moreover, proteolytic treatment of the vesicle membrane, although leading to a marked digestion of integral membrane proteins, only induces a minor, saturating increase of permeability, much lower than that in trypsinized resealed ghosts (Klonk, S. and Deuticke, B. (1992) Biochim. Biophys. Acta 1106, 137-142 (Part II of this series)). Since absence of the cytoskeletal proteins, spectrin and actin, is the major difference between IOV's and resealed ghosts, these results may be taken as further evidence for a dependence of the barrier properties of the erythrocyte membrane bilayer domain on its interaction with cytoskeletal elements. In contrast, these barrier properties seem to be rather insensitive to perturbations of integral proteins.  相似文献   

7.
The passive leakage of glucose across bilayers of dimyristoylphosphatidylcholine (DMPC), cholesterol (variable), and dicetyl phosphate (constant 5.9 mol%) has been measured as efflux over 30 min from multilamellar vesicles. Bilayer cholesterol was varied from 20 mol% to 40 mol%. Glucose permeation rates were measured from 10 degrees C to 36 degrees C, and showed a maximum in permeability at 24 degrees C, the DMPC phase transition temperature. Increasing the bilayer cholesterol content above 20 mol% reduced that permeability peak. These results are quite consistent with a large number of similar bilayer permeability studies over the past 25 years. However, they are not consistent with a previous study of these same systems, which reported increased glucose permeability with temperature, without any maximum at or near the lipid chain melting temperature (K. Inoue, Biochim. Biophys. Acta 339 (1974) 390-402).  相似文献   

8.
H Schindler  J Seelig 《Biochemistry》1975,14(11):2283-2287
The physical properties of bilayers of dipalmitoyl-3-sn-phosphatidylcholine are analyzed in terms of a statistical model proposed by Marcelja (S. Marcelja (1974), Biochim. Biophys. Acta 367, 165). The model is used to calculate the segmental order parameters of the hydrocarbon chains, the transition temperature of the crystalline leads to liquid crystalline phase transition, the entropy change of the transition, the bilayer thickness, and the linear thermal expansion coefficient. The theoretical predictions are in excellent agreement with experimental results obtained by deuterium magnetic resonance, differential scanning calorimetry, and X-ray diffraction. The model yields the probabilities of trans and gauche conformations and also those of more specific conformational defects like kinks or jogs.  相似文献   

9.
P J Spooner  D M Small 《Biochemistry》1987,26(18):5820-5825
Triacylglycerols are the major substrates for lipolytic enzymes that act at the surface of emulsion-like particles such as triglyceride-rich lipoproteins, chylomicrons, and intracellular lipid droplets. This study examines the effect of cholesterol on the solubility of a triacylglycerol, triolein, in phospholipid surfaces. Solubilities of [carbonyl-13C]triolein in phospholipid bilayer vesicles containing between 0 and 50 mol % free cholesterol, prepared by cosonication, were measured by 13C NMR. The carbonyl resonances from bilayer-incorporated triglyceride were shifted downfield in the 13C NMR spectra from those corresponding to excess, nonincorporated material. This enabled solubilities to be determined directly from carbonyl peak intensities at most cholesterol concentrations. The bilayer solubility of triolein was inversely proportional to the cholesterol/phospholipid mole ratio. In pure phospholipid vesicles the triolein solubility was 2.2 mol %. The triglyceride incorporation decreased to 1.1 mol % at a cholesterol/phospholipid mole ratio of 0.5, and at a mole ratio of 1.0 for the bilayer lipids, the triolein solubility was reduced to just 0.15 mol %. The effects of free cholesterol were more pronounced and progressive than observed previously on the bilayer solubility of cholesteryl oleate (Spooner, P. J. R., Hamilton, J. A., Gantz, D. L., & Small, D. M. (1986) Biochim. Biophys. Acta 860, 345-353]. As with cholesteryl oleate, we suggest that cholesterol also displaces solubilized triglyceride to deeper regions of the bilayer.  相似文献   

10.
Oxidation of erythrocyte membrane SH-groups by diamide and tetrathionate induces cross-linking of spectrin (Haest, C.W.M., Kamp, D., Plasa, G. and Deuticke, B. (1977) Biochim. Biophys. Acta 469, 226–230). This cross-linking was now shown to go along with a concentration- and time-dependent enhancement of membrane permeability for hydrophilic nonelectrolytes and ions. The enhancement is specific for oxidative SH-group modifications, is reversible by reduction of the induced disulfides, can be suppressed by a very brief pre-treatment of the cells with low concentrations of N-ethylmaleimide and is strongly temperature-dependent. The pathway of the induced permeability discriminates nonelectrolytes on the basis of molecular size and exhibits a very low activation energy (Ea 3–8 kcal/mol). These findings are reconcilable with the formation of a somewhat inhomogeneous population of aqueous pores with radii probably ? 0.65 nm. Estimated pore numbers vary with the size of the probe molecule. Assuming a diffusion coefficient as in bulk water within the pore, at least 20 pores per cell have to be postulated; more realistic lower diffusion coefficients increase that number. Alterations of the lipid domain by changes of cholesterol contents and insertion of hexanol or nonionic detergents alter the number or size of the pores. Since aggregation of skeletal and intrinsic membrane proteins also occurs after the SH-oxidation, in parallel to the formation of membrane leaks, one may consider (a) defects in the disturbed bilayer interface, (b) a mismatch between lipid and intrinsic proteins or (c) channels inbetween aggregated intrinsic proteins as structures forming the pores induced by diamide treatment.  相似文献   

11.
Errata     
Human erythrocytes were exposed to oxidative stress by iodate and periodate. Oxidation causes a time- and concentration-dependent increase in membrane permeability for hydrophilic molecules and ions. The induced leak discriminates nonelectrolytes on the basis of molecular size and exhibits a very low activation energy (Ea = 1–4 kcal · mol?1). These results are reconcilable with the formation of aqeous pores. The pore size was approximated to be between 0.45 and 0.6 nm. This increase in permeability is reversible upon treatment with dithioerythritol. Blocking of membrane thiol groups with N-ethylmaleimide protects the membranes against leak formation. The oxidation causes dithioerythritol-reversible modification of membrane proteins as indicated by the gel electrophoretic behavior. These modifications can also be suppressed by blocking the membrane thiol groups with N-ethylmaleimide. About half of the membrane methionine is oxidized to acid hydrolysis-stable derivatives. A fast saturating increase in diene conjugation was observed in whole cells but not in isolated membranes, with only minor degradation of fatty acid chains. The oxidation of cell membrane lipids as well as oxidation of cell surface carbohydrates are not involved in leak formation. Taken together with earlier data (Deuticke, B., Poster, B., Lütkemeier, P., and Haest, C.W.M. (1983) Biochim. Biophys. Acta 731, 196–210), these findings indicate that formation of disulfide bonds by different oxidative mechanisms results in leaks with similar properties.  相似文献   

12.
The binding of cytochrome b5 to single-walled liposomes of egg phosphatidylcholine was inhibited by the presence of cholesterol in the lipid bilayer under conditions where a limited amount of liposomes was incubated with the cytochrome. Since similar conditions seem to apply for the binding of cytochrome b5 to erythrocyte ghosts, this observation supports the conclusion of Enomoto and Sato (Enomoto, K. and Sato, R. (1977) Biochim. Biophys. Acta 466, 136–147) that the localization of cholesterol on the outer surface of the ghost membrane prevents the binding of cytochrome b5 to this surface. The finding reported by Roseman et al. (Roseman, M.A., Holloway, P.W. and Calabro, M.A. (1978) Biochim. Biophys. Acta 507, 552–556) that cholesterol did not prevent the cytochrome binding to phosphatidylcholine liposomes in the presence of a large excess of liposomes could be confirmed in the present study, but this does not contradict the abovementioned conclusion.  相似文献   

13.
The model of Cruzeiro-Hansson et al. (Biochim. Biophys. Acta (1989) 979, 166-1176) for lipid-cholesterol bilayers at low cholesterol concentrations is used to predict the thermodynamic properties and the passive ion permeability of lipid bilayers as a function of acyl-chain length and cholesterol concentration. Numerical simulations based on the Monte Carlo method are used to determine the equilibrium state of the system near the main gel-fluid phase transition. The permeability is calculated using an ansatz which relates the passive permeability to the amount of interfaces formed in the bilayer when cholesterol is present. The model predicts at low cholesterol contents an increase in the membrane permeability in the transition region both for increasing cholesterol concentration and for decreasing chain length at a given value of the reduced temperature. This is in contrast to the case of lipid bilayers containing high cholesterol concentrations where the cholesterol strongly suppresses the permeability. Experimental results for the Na+ permeability of C15PC and DPPC (C16PC) bilayers containing cholesterol are presented which confirm the theoretical predictions at low cholesterol concentrations.  相似文献   

14.
In 1972, Lieb, W. R.; Stein, W. D. (Biochim. Biophys. Acta 265, 187–207) in their review of sugar transport in human erythrocytes concluded that the conventional two-state carrier model was inconsistent with the experimental data available at that time. Since then, other papers have appeared which question the validity of the model. In this paper, we give a brief derivation of the equations describing the two-state carrier model, and analyze the predictions of the model in the classical experiments, i.e. zero-trans, infinite-cis, and equilibrium exchange. We show that the estimate of the half saturatiion constant of 2.8 mM for glucose at the inner face of the human red cell membrane for the infinite-cis procedure reported by Hankin, B. L., Lieb, W. R. and Stein, W. D ((1972) Biochim. Biophys. Acta 288, 114–126) is unreliable. We note that all of the other experimental findings are consistent with the asymmetric carrier model.  相似文献   

15.
Membrane cholesterol in porcine and bovine erythrocytes was elevated up to 165% of its normal value by incubation of the cells in cholesterol/phosphatidylcholine dispersions with or without serum. This alteration of membrane lipid composition brought about only a minor (10–40%) decrease of the permeability to glycerol, erythritol and to organic acids penetrating by non-ionic diffusion, although additional cholesterol had actually been incorporated into the lipid bilayer, as indicated by determinations of cell surface area from the critical hemolytic volume, in combination with quantitative evaluation of freeze-etch electron micrographs.On the basis of this finding and of the previously demonstrated (Grunze, M. and Deuticke, B. (1974) Biochim. Biophys. Acta 356, 125–130) considerable increase of permeability in cholesterol-depleted cells, it is proposed that in the erythrocyte membrane a pronounced “specific” reduction of permeability by cholesterol occurs only up to a molar ratio cholesterol/polar lipid of 0.6. At higher ratios cholesterol affects permeability only slightly, owing to an “unspecific” rigidifying effect on the membrane lipid phase.  相似文献   

16.
Membrane cholesterol in porcine and bovine erythrocytes was elevated up to 165% of its normal value by incubation of the cells in cholesterol/phosphatidylcholine dispersions with or without serum. This alteration of membrane lipid composition brought about only a minor (10-40%) decrease of the permeability to glycerol, erythritol and to organic acids penetrating by non-ionic diffusion, although additional cholesterol had actually been incorporated into the lipid bilayer, as indicated by determinations of cell surface area from the critical hemolytic volume, in combination with quantitative evaluation of freeze-etch electron micrographs. On the basis of this finding and of the previously demonstrated (Grunze, M. and Deuticke, B. (1974) Biochim. Biophys. Acta 356, 125-130) considerable increase of permeability in cholesterol-depleted cells, it is proposed that in the erythrocyte membrane a pronounced "specific" reduction of permeability by cholesterol occurs only up to a molar ratio cholesterol/polar lipid of 0.6. At higher ratios cholesterol affects permeability only slightly, owing to an "unspecific" rigidifying effect on the membrane lipid phase.  相似文献   

17.
The oxygen permeability coefficient across the membrane made of the total lipid extract from the plasma membrane of calf lens was estimated from the profile of the oxygen transport parameter (local oxygen diffusion-concentration product) and compared with those estimated for membranes made of an equimolar 1-palmitoyl-2-oleoylphosphatidylcholine/cholesterol (POPC/Chol) mixture and of pure POPC. Profiles of the oxygen transport parameter were obtained by observing the collision of molecular oxygen with nitroxide radical spin labels placed at different depths in the membrane using the saturation-recovery EPR technique and were published by us earlier (J. Widomska, M. Raguz, J. Dillon, E. R. Gaillard, W. K. Subczynski, Biochim. Biophys. Acta. 1768 (2007) 1454-1465). At 35 °C, the estimated oxygen permeability coefficients were 51.3, 49.7, and 157.4 cm/s for lens lipid, POPC/Chol, and POPC membranes, respectively (compared with 53.3 cm/s for a water layer with the same thickness as a membrane). Membrane permeability significantly decreases at lower temperatures. In the lens lipid membrane, resistance to the oxygen transport is located in and near the polar headgroup region of the membrane to the depth of the ninth carbon, which is approximately where the steroid-ring structure of cholesterol reaches into the membrane. In the central region of the membrane, oxygen transport is enhanced, significantly exceeding that in bulk water. It is concluded that the high level of cholesterol in lens lipids is responsible for these unique membrane properties.  相似文献   

18.
The action of gramicidin S and melittin on human erythrocytes, Staphylococcus aureus and Escherichia coli was studied as an extension of the previous study (Katsu, T., Ninomiya, C., Kuroko, M., Kobayashi, H., Hirota, T. and Fujita, Y. (1988) Biochim. Biophys. Acta 939, 57-63). These amphipathic peptides stimulated the release of membrane phospholipids outside cells in a concentration range causing permeability change. The shape change of erythrocytes from normal discoid to spiculate form was observed just prior to the release of membrane components. We have proposed the following action mechanism of gramicidin S and melittin. The peptide molecules were predominantly accumulated in the outer half of the bilayer, deforming the erythrocyte cell into crenature. A large accumulation made the membrane structure unstable, resulting in the release of membrane fragments and the simultaneous enhancement of permeability. The action mechanism of these peptides was compared with that of simple surfactants.  相似文献   

19.
A L MacDonald  D A Pink 《Biochemistry》1987,26(7):1909-1917
We have developed a model of glycophorin in a phospholipid bilayer membrane in order to study the thermodynamics of this system and to understand the detailed behavior of recent calorimetric data. We assume that the larger glycophorin polar group can be considered as either adopting a pancakelike conformation at the bilayer interface (D state) or be directed generally away from the interface (U state) [Ruppel, D., Kapitza, H.G., Galla, H.J., Sixl, F., & Sackmann, E. (1982) Biochim. Biophys. Acta 692, 1-17]. Lipid hydrocarbon chains are described either as excited (e state) with high energy and relatively many gauche conformers or as generally extended (g state) with low energy. We performed a Monte-Carlo simulation using the Glauber and Kawasaki procedures on a triangular lattice which represents the plane of half of the bilayer. Lattice sites can be occupied either by lipid hydrocarbon chains or by model glycophorin alpha-helical hydrophobic cores. The states D and U are represented by hexagons of different sizes in the plane of the lattice, and the hard core repulsion between two such polar groups is accounted for by forbidding hexagon-hexagon overlap. We have studied the effect of having the glycophorin polar group interact in various ways with the lipid bilayer. We find that the protein polar group in its D state interacts, either directly or indirectly, with the lipid bilayer so as to reduce the effective lateral pressure acting on the lipid hydrocarbon chains by about 3 dyn/cm. Polar groups in their U states do not reduce this lateral pressure.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
The nicotinic acetylcholine receptor (nAcChoR) has an absolute requirement for cholesterol if agonist-stimulated channel opening is to occur [Biochemistry 25 (1986) 830]. Certain non-polar analogs could replace cholesterol in vectorial vesicle permeability assays. Using a stopped-flow fluorescence assay to avoid the limitations of permeability assays imposed by vesicle morphology, it was shown that polar conjugates of cholesterol could also satisfy the sterol requirement [Biochim. Biophys. Acta 1370 (1998) 299]. Here this assay is used to explore the chemical specificity of sterols. Affinity-purified nAcChoRs from Torpedo were reconstituted into bilayers at mole ratios of 58:12:30 [1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC)/1,2-dioleoyl-sn-glycero-3-phosphate (DOPA)/steroid]. When the enantiomer of cholesterol was used, or when the stereochemistry at the 3-hydroxy group was changed from β to α by substituting epicholesterol for cholesterol, activation was still supported. The importance of cholesterol's planar ring structure was tested by comparing planar cholestanol (5α-cholestan-3β-ol) with nonplanar coprostanol (5β-cholestan-3β-ol). Both supported activation. Thus, these steroids support activation independent of structural features known to be important for modulation of lipid bilayer properties. This provides indirect support for a steroid binding site possessing very lax structural requirements.  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