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 the search for the mechanisms whereby water is transported across biological membranes, we hypothesized that in the airways, the hydration of the periciliary fluid layer is regulated by luminal-to-basolateral water transport coupled to active transepithelial sodium transport. The luminal-to-basolateral (JWL-->B) and the basolateral-to-luminal (JWB-->L) transepithelial water fluxes across ovine tracheal epithelia were measured simultaneously. The JWL-->B (6.1 microliter/min/cm2) was larger than JWB-->L (4.5 microliter/min/cm2, p < 0.05, n = 30). The corresponding water diffusional permeabilities were PdL-->B = 1.0 x 10(-4) cm/s and PdB-->L = 7.5 x 10(-5) cm/s. The activation energy (Ea) of JWL-->B (11.6 kcal/mol) was larger than the Ea of JWB-->L (6.5 kcal/mol, p < 0.05, n = 5). Acetylstrophanthidin (100 microM basolateral) reduced JWL-->B from 6.1 to 4.4 microliter/min/cm2 (p < 0. 05, n = 5) and abolished the PD. Amiloride (10 microM luminal) reduced JWL-->B from 5.7 to 3.7 microliter/min/cm2 (p < 0.05, n = 5) and reduced PD by 44%. Neither of these agents significantly changed JWB-->L. These data indicate that in tracheal epithelia under homeostatic conditions, JWB-->L was dominated by diffusion (Ea = 4.6 kcal/mol), whereas approximately 30% of JWL-->B was coupled to the active Na+,K+-ATPase pump (Ea = 27 kcal/mol).  相似文献   

2.
Mechanism of electroporative dye uptake by mouse B cells.   总被引:3,自引:0,他引:3       下载免费PDF全文
The color change of electroporated intact immunoglobulin G receptor (Fc gammaR-) mouse B cells (line IIA1.6) after direct electroporative transfer of the dye SERVA blue G (Mr 854) into the cell interior is shown to be dominantly due to diffusion of the dye after the electric field pulse. Hence the dye transport is described by Fick's first law, where, as a novelty, time-integrated flow coefficients are introduced. The chemical-kinetic analysis uses three different pore states (P) in the reaction cascade (C <==> P1 <==> P2 <==> P3), to model the sigmoid kinetics of pore formation as well as the biphasic pore resealing. The rate coefficient for pore formation k(p) is dependent on the external electric field strength E and pulse duration tE. At E = 2.1 kV cm(-1) and tE = 200 micros, k(p) = (2.4 +/- 0.2) x 10(3) s(-1) at T = 293 K; the respective (field-dependent) flow coefficient and permeability coefficient are k(f)0 = (1.0 +/- 0.1) x 10(-2) s(-1) and P0 = 2 cm s(-1), respectively. The maximum value of the fractional surface area of the dye-conductive pores is 0.035 +/- 0.003%, and the maximum pore number is Np = (1.5 +/- 0.1) x 10(5) per average cell. The diffusion coefficient for SERVA blue G, D = 10(-6) cm2 s(-1), is slightly smaller than that of free dye diffusion, indicating transient interaction of the dye with the pore lipids during translocation. The mean radii of the three pore states are r(P1) = 0.7 +/- 0.1 nm, r(P2) = 1.0 +/- 0.1 nm, and r(P3) = 1.2 +/- 0.1 nm, respectively. The resealing rate coefficients are k(-2) = (4.0 +/- 0.5) x 10(-2) s(-1) and k(-3) = (4.5 +/- 0.5) x 10)(-3) s(-1), independent of E. At zero field, the equilibrium constant of the pore states (P) relative to closed membrane states (C) is K(p)0 = [(P)]/[C] = 0.02 +/- 0.002, indicating 2.0 +/- 0.2% water associated with the lipid membrane. Finally, the results of SERVA blue G cell coloring and the new analytical framework may also serve as a guideline for the optimization of the electroporative delivery of drugs that are similar in structure to SERVA blue G, for instance, bleomycin, which has been used successfully in the new discipline of electrochemotherapy.  相似文献   

3.
Lateral diffusion measurements of PEG-lipid incorporated into magnetically aligned bicelles are demonstrated using stimulated echo (STE) pulsed field gradient (PFG) proton (1H) nuclear magnetic resonance (NMR) spectroscopy. Bicelles were composed of dimyristoyl phosphatidylcholine (DMPC) plus dihexanoyl phosphatidylcholine (DHPC) (q = DMPC/DHPC molar ratio = 4.5) plus 1 mol % (relative to DMPC) dimyristoyl phosphatidylethanolamine-N-[methoxy(polyethylene glycol)-2000] (DMPE-PEG 2000) at 25 wt % lipid. 1H NMR STE spectra of perpendicular aligned bicelles contained only resonances assigned to residual HDO and to overlapping contributions from a DMPE-PEG 2000 ethoxy headgroup plus DHPC choline methyl protons. Decay of the latter's STE intensity in the STE PFG 1H NMR experiment (g(z) = 244 G cm(-1)) yielded a DMPE-PEG 2000 (1 mol %, 35 degrees C) lateral diffusion coefficient D = 1.35 x 10(-11) m2 s(-1). Hence, below the "mushroom-to-brush" transition, DMPE-PEG 2000 lateral diffusion is dictated by its DMPE hydrophobic anchor. D was independent of the diffusion time, indicating unrestricted lateral diffusion over root mean-square diffusion distances of microns, supporting the "perforated lamellae" model of bicelle structure under these conditions. Overall, the results demonstrate the feasibility of lateral diffusion measurements in magnetically aligned bicelles using the STE PFG NMR technique.  相似文献   

4.
Fluorescence depolarization techniques are used to determine the molecular order and reorientational dynamics of the probe molecule TMA-DPH embedded in the lamellar L alpha and the hexagonal HII phases of lipid/water mixtures. The thermotropically induced L alpha----HII phase transition of the lipid DOPE is used to obtain macroscopically aligned samples in the hexagonal HII phase at 45 degrees C from samples prepared in the lamellar L alpha phase at 7 degrees C. The interpretation of angle-resolved fluorescence depolarization experiments on these phases, within the framework of the rotational diffusion model, yields the order parameters (P2) and (P4), and the diffusion constants for the reorientational motions. The reorientational motion rates of the TMA-DPH molecules in the hexagonal HII phase are comparable with those in the lamellar L alpha phase. Furthermore, the lateral diffusion of the probe molecule on the surface of the lipid/water cylinder in the hexagonal phase is found to be considerably slower than the reorientational motion.  相似文献   

5.
The chloroplast galactolipids monogalactosyldiacylglycerol (MGDG) and digalactosyldiacylglycerol (DGDG) were isolated from wheat leaves. The phase equilibria of galactolipid-water systems with MGDG / DGDG molar ratios equal to 0:1, 1:2, 1.2:1, 2:1 and 1:0 were investigated, using nuclear magnetic resonance (NMR) methods. MGDG and DGDG form reversed hexagonal and lamellar phases, respectively, at temperatures between 10 and 40°C at all water contents studied (up to about 14 mol 2H2O per mol lipid). The galactolipid mixtures show a complex phase forming reversed hexagonal, lamellar and reversed cubic phases, depending on water content and temperature. It was found that the water hydration is similar for the lamellar and hexagonal phases formed by DGDG and MGDG, respectively. The non-lamellar phase areas increase with increasing content of MGDG. Small-angle X-ray measurements show that the cubic phase belongs to the Ia3d space group. From translational diffusion studies by NMR it is concluded that the structure of this cubic phase is bicontinuous.  相似文献   

6.
The ternary phase diagram of cholesteryl linolenate-egg lecithin-water has been determined by polarizing light microscopy, calorimetry and X-ray diffraction at 23 °C. Hydrated lecithin forms a lamellar liquid-crystalline structure into which small amounts of cholesteryl linolenate are incorporated. The maximum incorporation of cholesterol ester into this lamellar structure varies with the degree of hydration. Increasing the water concentration from 10 to 15% (w/w) increased the limiting molar ratio of cholesteryl linolenate to lecithin in the lamellar phase from 1:50 to 1:22. At intermediate concentrations (15 to 30% water) the cholesteryl linolenate:lecithin ratio remains constant at 1:22. When water is increased to 42.5%, the maximum water content in the lamellar phase, the molar ratio decreased to 1:32. At low water concentrations the cholesterol ester appears to be entirely in the apolar region of the lecithin bilayer, while at higher water concentrations the ester groups of cholesteryl linolenate may be located at the lipid-water interface. At high water concentrations the ester appears to disorder the alkyl chains of the lecithin, giving rise to a thinner lipid layer and an increased surface area per lipid molecule when compared to the lecithin-water system in the absence of cholesteryl linolenate.The lamellar phase is the only phase (except at water concentrations less than 5%) in which all three components mutually interact. All mixtures of the three components having compositions outside the one-phase (lamellar) zone produce additional phases of cholesteryl linolenate or water, or both. Between 23 °C and 60 °C only minor changes in the phase diagram are observed.  相似文献   

7.
The mechanism of water permeation across the sarcolemma was characterized by examining the kinetics and temperature dependence of osmotic swelling and shrinkage of rabbit ventricular myocytes. The magnitude of swelling and the kinetics of swelling and shrinkage were temperature dependent, but the magnitude of shrinkage was very similar at 6 degrees, 22 degrees, and 37 degrees C. Membrane hydraulic conductivity, Lp, was approximately 1.2 x 10(-10) liter.N-1.s-1 at 22 degrees C, corresponding to an osmotic permeability coefficient, Pf, of 16 microns.s-1, and was independent of the direction of water flux, the magnitude of the imposed osmotic gradient (35-165 mosm/liter), and the initial cell volume. This value of Lp represents an upper limit because the membrane was assumed to be a smooth surface. Based on capacitive membrane area, Lp was 0.7 to 0.9 x 10(-10) liter.N-1.s-1. Nevertheless, estimates of Lp in ventricle are 15 to 25 times lower than those in human erythrocytes and are in the range of values reported for protein- free lipid bilayers and biological membranes without functioning water channels (aquaporin). Evaluation of the effect of unstirred layers showed that in the worst case they decrease Lp by < or = 2.3%. Analysis of the temperature dependence of Lp indicated that its apparent Arrhenius activation energy, Ea', was 11.7 +/- 0.9 kcal/mol between 6 degrees and 22 degrees C and 9.2 +/- 0.9 kcal/mol between 22 degrees and 37 degrees C. These values are significantly greater than that typically found for water flow through water-filled pores, approximately 4 kcal/mol, and are in the range reported for artificial and natural membranes without functioning water channels. Taken together, these data strongly argue that the vast majority of osmotic water flux in ventricular myocytes penetrates the lipid bilayer itself rather than passing through water-filled pores.  相似文献   

8.
Lateral diffusion measurements of polyethylene glycol(PEG)-lipid incorporated into magnetically aligned lipid bilayers, composed of dimyristoyl phosphatidylcholine (DMPC) plus dihexanoyl phosphatidylcholine (DHPC) plus 1 mol % (relative to DMPC) dimyristoyl phosphatidylethanolamine-n-[methoxy(polyethylene glycol)-2000] (DMPE-PEG 2000), were performed using stimulated-echo pulsed-field-gradient proton (1H) nuclear magnetic resonance spectroscopy. The DMPE-PEG 2000 (1 mol %, 35°C) lateral diffusion coefficient D varied directly with the mole fraction of DMPC, XDMPC = q/(1+q) where q = DMPC/DHPC molar ratio, decreasing progressively from D = 1.65 × 10−11 m2 s−1 at q ≈ 4.7 to D = 0.65 × 10−11 m2 s−1 at q ≈ 2.5. Possible sources of this dependence, including orientational disorder, obstruction, and PEG-lipid sequestration, were simulated using, respectively, a diffusion-in-a-cone model, percolation theory, and a two-phase PEG distribution model. Orientational disorder alone was not capable of reproducing the observations, but in combination with either obstruction or PEG-lipid two-phase distribution models did so satisfactorily. A combination of all three models yielded the most reasonable fit to the observed dependence of lateral diffusion on q. These same effects would be expected to influence lateral diffusion of any bilayer-associating species in such systems.  相似文献   

9.
The lateral diffusion constants of 1-palmitoyl-2-oleoyl-sn-glycero-3 phosphocholine (POPC), water, and ibuprofen were measured in multilamellar liposomes using pulsed field gradient magic-angle spinning (PFG-MAS) (1)H NMR. The analysis of diffusion data obtained in powder samples and a method for liposome curvature correction are presented. At 322 K POPC has a diffusion constant of (8.6 +/- 0.2) x 10(-12) m(2)/s when dehydrated (8.2 waters/lipid) and (1.9 +/- 0.1) x 10(-11) m(2)/s in excess water. The diffusion constant of water in dehydrated POPC was found to be (4.7 +/- 0.1) x 10(-10) m(2)/s. The radius of curvature is 21 +/- 2 microm for the dehydrated sample and 4.5 +/- 0.5 microm for POPC sample containing excess water. The activation energies of diffusion are 40.6 +/- 0.4 kJ/mole for dehydrated POPC, 30.7 +/- 0.9 kJ/mole for POPC with excess water, and 28.6 +/- 1.5 kJ/mole for water in dehydrated POPC. The diffusion constants and activation energies for a sample of POPC/ibuprofen/water (1:0.56:15) were also measured. The ibuprofen, which locates in the lipid-water interface, diffuses faster than POPC but has a slightly higher activation energy of lateral diffusion. Within certain restrictions, PFG-MAS NMR provides a useful method for characterizing membrane organization and mobility.  相似文献   

10.
A Derksen  D M Small 《Biochemistry》1989,28(2):900-906
The cholesterol content of triglyceride-rich lipoproteins increases during their catabolism in circulation. We therefore studied the binding of the exchangeable apoprotein apoA-1 and apoE-3 to triolein-rich emulsions with increasing cholesterol content. Five emulsion systems containing 83.1-88.8% (w/w) triolein, 9.3-10.1% egg yolk phosphatidylcholine, and 1.1-7.3% cholesterol were isolated from sonicated lipid mixtures by flotation. Negative stain EM of emulsions containing 1.1 and 7.3% cholesterol showed polydisperse populations of large spherical particles with diameters of 106 +/- 39 and 108 +/- 57 nm. These values are similar to particle diameters calculated from the lipid composition data. No lamellar structures were observed by EM, even after addition of apoA-1 at a molar ratio to lecithin of 10(-2). Apolipoproteins apoA-1 and apoE-3 bound to the particles in a saturable manner without altering particle morphology. We found a dissociation constant Kd = 7.4 x 10(-7) M and a binding capacity N = 3.9 x 10(-3) proteins/lecithin for apoA-1 with particles containing 1.1% cholesterol; the Kd and N values for apoE-3 were very similar. When the emulsion particles were saturated with cholesterol at 7.3%, the protein binding capacity N sharply decreased to 0.6 x 10(-3) (apoA-1) and 0.7 x 10(-3) proteins/lecithin (apoE-3), but the Kd values were virtually unchanged. No change in N occurred when the particle cholesterol content was increased from 1.1 to 3.7%, which spans the normal physiological range. These results suggest that increases in lipoprotein cholesterol content above 3.7% may be responsible for impaired apoprotein redistribution and altered metabolism of remnants such as beta-VLDL.  相似文献   

11.
Solid state deuterium (2H) NMR inversion-recovery and Jeener-Broekaert relaxation experiments were performed on oriented multilamellar dispersions consisting of 1,2-dilauroyl-sn-glycero-3-phosphatidylcholine and 2H exchange-labeled gramicidin D, at a lipid to protein molar ratio (L/P) of 15:1, in order to study the dynamics of the channel conformation of the peptide in a liquid crystalline phase. Our dynamic model for the whole body motions of the peptide includes diffusion of the peptide around its helix axis and a wobbling diffusion around a second axis perpendicular to the local bilayer normal in a simple Maier-Saupe mean field potential. This anisotropic diffusion is characterized by the correlation times, tau R parallel and tau R perpendicular. Aligning the bilayer normal perpendicular to the magnetic field and graphing the relaxation rate, 1/T1Z, as a function of (1-S2N-2H), where S2N-2H represents the orientational order parameter, wer were able to estimate the correlation time, tau R parallel, for rotational diffusion. Although in the quadrupolar splitting, which varies as (3 cos2 theta D-1), has in general two possible solutions to theta D in the range 0 < or = theta D < or = 90 degrees, the 1/T1Z vs. (1-S2N-2H) curve can be used to determine a single value of theta D in this range. Thus, the 1/T1Z vs. (1-S2N-2H) profile can be used both to define the axial diffusion rate and to remove potential structural ambiguities in the splittings. The T1Z anisotropy permits us to solve for the two correlation times (tau R parallel = 6.8 x 10(-9) s and tau R perpendicular = 6 x 10(-6) s). The simulated parameters were corroborated by a Jeener-Broekaert experiment where the bilayer normal was parallel to the principal magnetic field. At this orientation the ratio, J2(2 omega 0)/J1(omega 0) was obtained in order to estimate the strength of the restoring potential in a model-independent fashion. This measurement yields the rms angle, <theta 2>1/2 (= 16 +/- 2 degrees at 34 degrees C), formed by the peptide helix axis and the average bilayer normal.  相似文献   

12.
The longitudinal diffusion of a homologous series of monoamides through lecithin-water lamellar phases with aqueous channel widths of 16–27 Å has been studied. The diffusion coefficients relative to water of the hydrophilic amides, formamide and acetamide, depend logarithmically on solute molar volume, as previously demonstrated in human red cells. Aqueous diffusion of amides in red-cell membranes is similar to that in a lecithin-water phase of aqueous channel width less than 16 Å, the smallest channel width used. Partition coefficients of the lipophilic amides, valeramide and isovaleramide, between lecithin vesicles and water are 1.64 and 1.15 at 20 °C. These data enabled us to compute a valeramide diffusion coefficient of 6.5 · 10−7cm2 · s−1 at 20 °C in the lipid region of a lamellar phase containing 30% water about one order of magnitude greater than the diffusion coefficient of spin-labelled analogs of phosphatidylcholine. The discrimination between the permeability coefficients of valeramide and isovaleramide is more than twice as great in the human red cell as between lipid diffusion coefficients in a phase containing 8% water. This suggests that the lipid region of the human red cell is more highly organized than lipid in the lecithin-water lamellar phase.  相似文献   

13.
We have studied the flickering of erythrocytes at wavelengths comparable to the cell dimension. To do this we have analyzed the edge fluctuations of the cell to a resolution of 5 nm by combining phase contrast microscopy with fast image processing. By measuring the edge excitations simultaneously at four orthogonal positions around the cell, the eigenmodes of equal azimuthal mode numbers m = 0,1,2 could be separated. From a continuous time sequence of 100 s of video frames taken at 40 ms time intervals, we determined the time-auto correlation function for the modes m = 0,1,2 and calculated their mean square amplitudes <delta n2m> as well as their decay times tau m. To explain the results we also present the theoretically calculated energy eigenmodes of an erythrocyte, accounting for the constraint that the cell is in contact with the substrate along an annular ring, which agreed well with the experimental findings. We found that the softest mode is a "hindered translational" mode with m = 1 of the adhered cell, which is almost insensitive to the shear elastic modulus. Comparison of the calculated and measured amplitudes yielded an average value for the bending stiffness of kc = 4 x 10(-19) J, which is much larger than the value obtained by flicker analysis at short wavelengths (kc = 2.3 x 10(-20) J). It would, however, agree well with the value expected from the red cell membrane area compressibility modulus of K = 4.5 x 10(-1)N/m, which corresponds to a lipid bilayer containing approximately 50 mol % of cholesterol. In contradiction to our theoretical expectations we found that the flicker eigenmodes seemed not to be influenced by the membrane shear elasticity, which will be discussed in terms of an unusual coupling between the lipid bilayer and the cytoskeleton.  相似文献   

14.
J Shah  R I Duclos  Jr    G G Shipley 《Biophysical journal》1994,66(5):1469-1478
The structural and thermotropic properties of 1-stearoyl-2-acetyl-phosphatidylcholine (C(18):C(2)-PC) were studied as a function of hydration. A combination of differential scanning calorimetry and x-ray diffraction techniques have been used to investigate the phase behavior of C(18):C(2)-PC. At low hydration (e.g., 20% H2O), the differential scanning calorimetry heating curve shows a single reversible endothermic transition at 44.6 degrees C with transition enthalpy delta H = 6.4 kcal/mol. The x-ray diffraction pattern at -8 degrees C shows a lamellar structure with a small bilayer periodicity d = 46.3 A and two wide angle reflections at 4.3 and 3.95 A, characteristic of a tilted chain, L beta' bilayer gel structure. Above the main transition temperature, a liquid crystalline L alpha phase is observed with d = 53.3 A. Electron density profiles at 20% hydration suggest that C(18):C(2)-PC forms a fully interdigitated bilayer at -8 degrees C and a noninterdigitated, liquid crystalline phase above its transition temperature (T > Tm). Between 30 and 50% hydration, on heating C(18):C(2)-PC converts from a highly ordered, fully interdigitated gel phase (L beta') to a less ordered, interdigitated gel phase (L beta), which on further heating converts to a noninterdigitated liquid crystalline L alpha phase. However, the fully hydrated (> 60% H2O) C(18):C(2)-PC, after incubation at 0 degrees C, displays three endothermic transitions at 8.9 degrees C (transition I, delta H = 1.6 kcal/mol), 18.0 degrees C (transition II), and 20.1 degrees C (transition III, delta HII+III = 4.8 kcal/mol). X-ray diffraction at -8 degrees C again showed a lamellar gel phase (L beta') with a small periodicity d = 52.3 A. At 14 degrees C a less ordered, lamellar gel phase (L beta) is observed with d = 60.5 A. However, above the transition III, a broad, diffuse reflection is observed at approximately 39 A, consistent with the presence of a micellar phase. The following scheme is proposed for structural changes of fully hydrated C(18):C(2)-PC, occurring with temperature: L beta' (interdigitated)-->L beta (interdigitated)-->L alpha(noninterdigitated)-->Micelles. Thus, at low temperature C(18):C(2)-PC forms a bilayer gel phase (L beta') at all hydrations, whereas above the main transition temperature it forms a bilayer liquid crystalline phase L alpha at low hydrations and a micellar phase at high hydrations (> 60 wt% water).  相似文献   

15.
We have measured the lateral diffusion coefficient (D), of active dansyl-labeled gramicidin C (DGC), using the technique of fluorescence photobleaching recovery, under conditions in which the cylindrical dimer channel of DGC predominates. In pure, hydrated, dimyristoylphosphatidylcholine (DMPC) multibilayers (MBL), D decreases from 6 X 10(-8) cm2/s at 40 degrees C to 3 X 10(-8) cm2/s at 25 degrees C, and drops 100-fold at 23 degrees C, the phase transition temperature (Tm) of DMPC. Above Tm, addition of cholesterol decreases D; a threefold stepwise drop occurs between 10 and 20 mol %. Below Tm, increasing cholesterol increases D; a 10-fold increase occurs between 10 and 20 mol % at 21 degrees C, between 20 and 25 mol % at 15 degrees C, and between 25 and 30 mol % at 5 degrees C. In egg phosphatidylcholine (EPC) MBL, D decreases linearly from 5 X 10(-8) cm2/s at 35 degrees C to 2 X 10(-8) cm2/s at 5 degrees C; addition of equimolar cholesterol reduces D by a factor of 2. Thus this transmembrane polypeptide at low membrane concentrations diffuses quite like a lipid molecule. Its diffusivity in lipid mixtures appears to reflect predicted changes of lateral composition. Increasing gramicidin C (GC) in DMPC/GC MBL broadened the phase transition, and the diffusion coefficient of the lipid probe N-4-nitrobenzo-2-diazole phosphatidylethanolamine (NBD-PE) at 30 degrees C decreases from 8 X 10(-8) cm2/s below 5 mol % GC to 2 X 10(-8) cm2/s at 14 mol % GC; D for DGC similarly decreases from 4 X 10(-8) cm2/s at 2 mol % GC to 1.4 X 10(-8) cm2/s at 14 mol % GC. Hence, above Tm, high concentrations of this polypeptide restrict the lateral mobility of membrane components.  相似文献   

16.
The effect of increasing concentrations of lipid X (2,3-bis(3-hydroxymyristoyl)-alpha-D-glucosamine 1-phosphate) on the phase behaviour of EPC (egg phosphatidylcholine) and EPE (egg phosphatidylethanolamine) is studied at a pH greater than or equal to 7 where lipid X carries one to two negative charges. Small amounts of lipid X (molar ratio approximately 0.01) induce continuous swelling of EPC and EPE bilayers and consequently the formation of large unilamellar vesicles in excess water. In many respects, the effect of lipid X on EPC and EPE bilayers is similar to that of phosphatidic acid. However, lipid X/EPC mixtures form micelles in excess lipid X whereas mixtures of phosphatidic acid/EPC vesiculate at all ratios. The same is true for lipid X/EPE mixtures. Small unilamellar vesicles of an average diameter of 40 nm form spontaneously upon dispersion of a dry lipid X/EPE film (molar ratio = 10). Unsonicated dispersions of lipid X/EPC (molar ratio = 1) are subjected to pH-jump treatment which involves raising of the pH to 11-12 and subsequent lowering of the pH to between 7.5 and 8.5. Such a treatment has little effect on the vesicle size and size distribution as compared to a control dispersion at pH 8.2. The mean size is determined to be 92 +/- 60 nm. Electron micrographs of freeze-fractured samples of lipid X/EPC (molar ratio = 1) reveal the presence of mainly micelles at pH 12. Upon lowering the pH to neutrality these micelles become unstable and aggregate/fuse rapidly to unilamellar vesicles (average diameter 95 +/- 40 nm). Sonication of equimolar mixtures of lipid X and EPC at pH 7 yields small unilamellar vesicles of a diameter of 20-25 nm as well as mixed micelles of a size between 15 and 17 nm. This behaviour is again different from that of mixed EPC/phosphatidic acid dispersions which form small unilamellar vesicles. The presence of lipid X in such mixtures does not prevent the aggregation/fusion to larger vesicles during freezing of the dispersion. As with pure EPC bilayers, stabilization is, however, achieved in the presence of 10% sucrose. This indicates that the covalently bonded glucosamine group of lipid X cannot substitute water of hydration in neighbouring EPC molecules.  相似文献   

17.
Cold acclimation of Arabidopsis thaliana includes the expression of cold-regulated (COR) genes and the accumulation of COR polypeptides. The hydration characteristics of two COR polypeptides, COR6.6 and COR15am, have been determined and their effects on the dehydration-induced liquid crystalline-to-gel and lamellar-to-hexagonal II phase transitions in phospholipid mixtures have been examined. After dehydration at osmotic pressures between 8 and 150 MPa, the water content of the COR polypeptides was less than that of bovine serum albumin, with COr15am the least hydrated: bovine serum albumin > COR6.6 > COR15am. Neither COR6.6 nor COR15am altered the dehydration-induced gel lamellar --> fluid lamellar phase transition temperature of either dipalmitoylphosphatidylcholine or dioleoylphosphatidylcholine (DOPC). In multilamellar vesicles of dioleoylphosphatidylethanolamine:DOPC (1:1, mol:mol) prepared by either freeze-thaw or reverse-phase evaporation methods, neither COR6.6, COR15am, nor bovine serum albumin altered the incidence of the dehydration-induced formation of the inverted hexagonal phase as a function of osmotic pressure. However, a specific ultrastructural alteration--the formation of a striated surface morphology in the lamellar domains--was observed in mixtures of dioleoylphosphatidylethanolamine:DOPC that were dehydrated in the presence of COR15am. Nevertheless, neither COR6.6 nor COR15am appears to participate in a specific protein-phospholipid interaction that alters the dehydration-induced phase behavior of phospholipid vesicles.  相似文献   

18.
Three bipolar archaeal-type diglycerophosphocholine tetraether lipids (also known as bolalipids) have been prepared to determine (1) the influence of molecular structure on the physical properties of bolalipid membranes and (2) their impact on the functional reconstitution of Ste14p, a membrane-associated isoprenylcysteine carboxyl methyltransferase from Saccharomyces cerevisiae. Three bolalipids were synthesized: C20BAS, C32BAS, and C32phytBAS. These bolalipid structures differ in that the C20BAS derivative has a short sn-1 glyceryl diether C20H40 transmembrane alkyl chain and two ether-linked sn-2 n-decyl chains, whereas the C32BAS and C32phytBAS derivatives have a longer sn-1 diether C32H64 membrane-spanning chain and two ether-linked sn-2 n-hexadecyl or phytanyl chains, respectively. Differential scanning calorimetry and temperature-dependent 31P NMR was used to determine the gel-to-liquid crystalline phase transition temperatures of the bolalipids (C32BAS Tm > 85 degrees C; C32phytBAS Tm = 14 degrees C; and C20BAS Tm = 17 degrees C). The bolalipid lateral diffusion coefficients, determined by fluorescence recovery after photobleaching at 25 degrees C, were 1.5 x 10(-8) and 1.8 x 10(-9) cm2/s for C20BAS and C32phytBAS, respectively. The mobility of C32BAS could not be measured at this temperature. Ste14p activity was monitored by an in vitro methyltransferase assay in reconstituted vesicle dispersions composed of DMPC, C20BAS/E. coli polar lipid, C20BAS/POPC, C32phytBAS/E. coli polar lipid, and C32phytBAS/POPC. Ste14p activity was lost in vesicles composed of 75-100 mol % C20BAS and 0-100 mol % C32BAS but retained in vesicles with 0-50 mol % C20BAS and 0-100 mol % C32phytBAS. Confocal immunofluorescence microscopy confirmed the presence of Ste14p in 100 mol % C20BAS and 100 mol % C32phytBAS vesicle dispersions, even though the lamellar liquid crystalline phase thickness of C20BAS is only 32 A. Because Ste14p activity was not affected by either the gel-to-liquid-crystal phase transition temperature of the lipid or the temperature of the assay, the low activity observed in 75-100 mol % C20BAS membranes can be attributed to hydrophobic mismatch between this bolalipid and the hydrophobic surface of Ste14p.  相似文献   

19.
The egg of Manduca sexta contains a very high density lipophorin (VHDLp-E; Mr approximately equal to 4.14 x 10(5), rho = 1.238 g/ml) that is derived from the high density lipophorin (HDLp-A; Mr approximately equal to 7.63 x 10(5), rho = 1.076 g/ml) of the hemolymph. The selective uptake of HDLp-A into the egg and its subsequent conversion to VHDLp-E was studied both in vivo and in vitro. Upon entering the egg, an estimated 530 mol of lipid were stripped from each mol of HDLp-A, and 68% of the diacylglycerol fraction was converted to triacylglycerol. In addition, the two molecules of the low molecular weight apolipoprotein, apolipophorin-III, of HDLp-A were dissociated from the lipophorin particle. The VHDLp-E thus formed consisted of 80% protein and 20% lipid, 75% of which was phospholipid. HDLp-A labeled in vivo with [35S]methionine in its apoprotein moiety was injected into females at the onset of egg development, and its incorporation in a series of follicles at different stages of growth was measured. There was increased accumulation of [35S]HDLp-A in the follicles as they matured. The apoproteins of [35S]HDLp-A were not hydrolyzed when the particle was internalized by the follicle. In the accompanying paper we have presented the evidence that the apoproteins of HDLp-A are retained in the follicles (Kawooya, J.K., and Law, J.H. (1988) J. Biol. Chem. 263, 8748-8753).  相似文献   

20.
Differential scanning calorimetry and x-ray diffraction techniques have been used to investigate the structure and phase behavior of hydrated dimyristoyl lecithin (DML) in the hydration range 7.5 to 60 weight % water and the temperature range -10 to +60 degrees C. Four different calorimetric transitions have been observed: T1, a low enthalpy transition (deltaH approximately equal to 1 kcal/mol of DML) at 0 degrees C between lamellar phases (L leads to Lbeta); T2, the low enthalpy "pretransition" at water contents greater than 20 weight % corresponding to the transition Lbeta leads to Pbeta; T3, the hydrocarbon chain order-disorder transition (deltaH = 6 to 7 kcal/mol of DML) representing the transition of the more ordered low temperature phases (Lbeta, Pbeta, or crystal C, depending on the water content) to the lamellar Lalpha phase; T4, a transition occurring at 25--27 degrees C at low water contents representing the transition from the lamellar Lbeta phase to a hydrated crystalline phase C. The structures of the Lbeta, Pbeta, C, and Lalpha phases have been examined as a function of temperature and water content. The Lbeta structure has a lamellar bilayer organization with the hydrocarbon chains fully extended and tilted with respect to the normal to the bilayer plane, but packed in a distorted quasihexagonal lattice. The Pbeta structure consists of lipid bilayer lamellae distorted by a periodic "ripple" in the plane of the lamellae; the hydrocarbon chains are tilted but appear to be packed in a regular hexagonal lattice. The diffraction pattern from the crystalline phase C indexes according to an orthorhombic cell with a = 53.8 A, b = 9.33 A, c = 8.82 A. In the lamellae bilayer Lalpha strucure, the hydrocarbon chains adopt a liquid-like conformation. Analysis of the hydration characteristics and bilayer parameters (lipid thickness, surface area/molecule) of synthetic lecithins permits an evaluation of the generalized hydration and structural behavior of this class of lipids.  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