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.
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.
The influence of sucrose, between 0 and 70% in the aqueous phase, upon multilamellar liposomes of dimyristoylphosphatidylcholine was examined by differential scanning calorimetry and X-ray diffraction analysis. Increasing concentrations of sucrose increase the temperatures of both the main transition and the pretransition of the lipid. The effect is greater on the pretransition than on the main transition. At 35 degrees C the interlamellar spacing in the multilamellar liposomes is reduced by increasing sucrose concentration in the aqueous phase and no significant effects are seen in the chain lattice of the bilayers. This result is interpreted as a dehydrating effect of sucrose upon the bilayer-water system at 35 degrees C. At 5 degrees C the interlamellar spacing is increased and this increase is, at high (70%) sucrose concentrations, attributable to an untilting of the lipid acyl chains with no change in the thickness of the aqueous layers in the multilamellae.  相似文献   

3.
The purpose of this study was to correlate the effectiveness of the lysoPC to disrupt bilayers with the effects of trehalose and sucrose on the hydration sites of a lipid bilayer. The vibration frequencies of carbonyls and phosphates was measured at 18 degrees C for different ratios of monomyristoylphosphatidylcholine and dimyristoylphosphatidylcholine vesicles prepared in water, sucrose and trehalose. The disruption point of the bilayer, evaluated by following the changes in the turbidity of the suspension of unilamellar vesicles, was decreased when the vesicles were prepared in 100 mM sucrose. The increase of the lytic action is directly related to the extent of hydration of the carbonyl populations. It is interpreted that the insertion of the sucrose molecule in the interface causes local changes in interfacial structure, such as the dehydration of the second population of the carbonyls that may be identified as defects of packing. In contrast, the insertion of trehalose by replacing water simultaneously at the carbonyls and the phosphates does not cause defects of packing. For this reason, the lytic action is produced at a concentration very similar to that found in water.  相似文献   

4.
Molecular simulations of hydrated dipalmitoylphosphatidylcholine lipid bilayers have been performed for temperatures in the range of 250-450 K. The area per headgroup increases with temperature from 58 to 77 A(2). Other properties such as hydration number, alkyl tail order parameter, diffusion coefficients, and radial distribution functions exhibit a clear dependence on temperature. Simulations of bilayers have also been performed in the presence of two disaccharides, namely trehalose and sucrose, at concentrations of up to 18 wt % (lipid-free basis). The simulated area per headgroup of the bilayer is not affected by the presence of the disaccharides, suggesting that the overall structure of the bilayer remains undisturbed. The results of simulations reveal that the interaction of disaccharide molecules with the bilayer occurs at the surface of the bilayer, and it is governed by the formation of multiple hydrogen bonds to specific groups of the lipid. Disaccharide molecules are observed to adopt specific conformations to fit onto the surface topology of the bilayer, often interacting with up to three different lipids simultaneously. At high disaccharide concentrations, the results of simulations indicate that disaccharides can serve as an effective replacement for water under anhydrous conditions, which helps explain their effectiveness as lyophilization agents for liposomes and cells.  相似文献   

5.
The fusion of small unilamellar vesicles of phosphatidylcholines during freeze-thawing and freeze-drying/rehydration, and the suppression of fusion under these conditions by various saccharides, was investigated by gel filtration on Sepharose 4B, quasielastic light scattering, high-resolution 1H-NMR, ESR spin labeling, and differential scanning calorimetry. Freeze-thawing and freeze-drying of aqueous small unilamellar vesicle suspensions in the presence of sufficient sucrose had no significant effect on the average size and size distribution of small unilamellar vesicles. In the presence of sucrose the structural integrity and the permeability properties of the phosphatidylcholine bilayers were retained during freeze-thawing and freeze-drying. A comparison of the stabilizing effect of sucrose with that of trehalose and glucose showed that the stabilization is not sugar-specific but is a general property of saccharides. The fraction of small unilamellar vesicles recovered after freeze-thawing depended on the saccharide/phosphatidylcholine molar ratio. The mechanism of the cryoprotective effect involves binding of the sugar to the phospholipid polar group, probably through hydrogen bonding.  相似文献   

6.
A new pair of fluorescence-energy-transferring probes, dansylphosphatidylethanolamine and dioctadecylindocarbocyanine, were incorporated separately into phospholipid vesicles to monitor intervesicle lipid mixing under various conditions. The transfer efficiencies of mixtures of sonicated vesicles labeled with 2 wt% donor dansylphosphatidylethanolamine (DnsPE) or with 1 wt% acceptor dioctadecylindocarbocyanine (DiI-C18) were negligible, but increased to about 25% after the vesicles had been frozen in a solid CO2/ethanol bath, thawed and diluted. The freeze-thaw-induced mixing of lipids between vesicles, signified by energy transfer, was dependent on lipid concentration and was promoted by 0.5-1.5 M KCl, 0.5 M potassium trichloroacetate and 5 mM sodium acetate (pH 4) and inhibited by 0.5 M LiCl, 0.5 M glycerol, 0.5 M sucrose, 0.15 M KCl and 0.15-1.5 M NaCl. These results support and complement previously reported measurements of the trapped volumes, turbidities and population size distributions of similarly treated liposomes. Comparison of the responses of paucilamellar vesicles with those of multilamellar vesicles suggests that lipid mixing during freeze-thawing can occur either during interaction of the outermost bilayers of vesicles or during interaction of all bilayers, possibly as a result of breakdown and reformation of bilayer structure.  相似文献   

7.
The vesicle-to-micelle transition of egg phosphatidylcholine LUVs induced by octylglucoside was studied in buffers with 0-4 M sodium chloride, sucrose or urea. We used both light scattering and fluorescent probes to follow the lipid-detergent complexes in these buffers. The vesicle-to-micelle transition process was fundamentally the same in each solute. However, the detergent-to-lipid ratio required for micelle formation shifted in ways that depended on the aqueous solute. The partitioning of octylglucoside between the vesicles and the aqueous phase was primarily determined by the change in its critical micelle concentration (cmc) induced by each solute. Specifically, the cmc decreased in high salt and sucrose buffers but increased in high concentrations of urea. Cmc for two additional nonionic detergents, decyl- and dodecyl-maltoside, and three zwittergents (3-12, 3-14 and 3-16) were determined as a function of concentration for each of the solutes. In all cases NaCl and sucrose decreased the solubility of the detergents, whereas urea increased their solubilities. The effects clearly depended on acyl chain length in urea-containing solutions, but this dependence was less clear with increasing NaCl and sucrose concentrations. The contributions of these solutes to solubility and to interfacial interactions in the bilayers, pure and mixed micelles are considered.  相似文献   

8.
In this communication it is demonstrated that the cell wall of the gram-positive bacterium Micromonospora purpurea contains a cell wall channel for the passage of hydrophilic solutes. The channel-forming protein was identified in sucrose step-density-gradient fractions of the cell envelope and in whole cell extracts using either organic solvent or detergent and the lipid bilayer technique. The fractions of the sucrose step-density centrifugation were assayed for NADH-oxidase activity and for the formation of ion-permeable channels in lipid bilayers. The highest NADH-oxidase activity and the highest channel-forming ability were found in different fractions. The cell wall fraction was identified by the presence of meso-diaminopimelic acid and contained an ion-permeable channel with the extremely high single-channel conductance of about 14 nS in 1 M KCl. The channel-forming unit was purified to homogeneity by FPLC on a HiTrap-Q column. It was identified as a heat- and SDS-resistant 200-kDa band on SDS-PAGE and formed the same general diffusion pores in lipid bilayer membranes as those formed by detergent extracts of the cell wall fraction of the sucrose step-density centrifugation. The channels were slightly selective for potassium ions over chloride, possibly caused by an excess of negative charges in or near the channel.  相似文献   

9.
The disaccharides trehalose and sucrose have small effects on temperature and enthalpy of the pre- and main phase transition in hydrated DMPC bilayers. In contrast, these sugars cause a considerable retention of carboxyfluorescein when large unilamellar vesicles of DMPC are heated through the main transition. This effect is sugar specific, as the monosaccharides glucose and fructose are less effective and ethyleneglycol has no effect at all.  相似文献   

10.
The effect of the small anesthetic molecule, benzyl alcohol, on the structure of various bilayer system has been studied by optical, electrical, and x-ray diffraction techniques. We find that the modifications in bilayer thickness caused by benzyl alcohol differ dramatically for planar (or black lipid) bilayers containing solvent, planar bilayers containing little or no solvent, and vesicular bilayers. Benzyl alcohol increases the thickness of planar bilayers containing n-alkane solvents, yet decreases the thickness of "solvent-free" planar bilayers. The effect of benzyl alcohol on vesicular bilayers below the phase transition temperature also depends on whether solvent is present in the bilayers. Without solvent, gel-state bilayers are reduced in thickness by benzyl alcohol, whereas in the presence of solvent, the thickness is unchanged. Above the phase transition temperature, benzyl alcohol has no measurable effect on vesicular bilayer thickness, whether solvent is present or not. These results indicate that different model membrane systems respond quite differently to a particular anesthetic.  相似文献   

11.
CFTR is a monomer: biochemical and functional evidence   总被引:3,自引:0,他引:3  
Although the CFTR protein alone is sufficient to generate a regulated chloride channel, it is unknown how many of the polypeptides form the channel. Using biochemical and functional assays, we demonstrate that the CFTR polypeptide is a monomer. CFTR sediments as a monomer in a linear, continuous sucrose gradient. Cells co-expressing different epitope-tagged CFTR provide no evidence of co-assembly in immunoprecipitation and nickel affinity binding experiments. Co-expressed wild-type and DF508 CFTR are without influence on each other in their ability to progress through the secretory pathway, suggesting they do not associate in the endoplasmic reticulum. No hybrid conducting single channels are seen in planar lipid bilayers with which membrane vesicles from cells co-expressing similar amounts of two different CFTR conduction species have been fused.  相似文献   

12.
Proton translocating ATPase (F0F1) from bovine heart mitochondria was reconstituted into planar phospholipid bilayers, and its electrogenicity was directly demonstrated. The F0F1 ATPase was solubilized using 3-[(3-cholamidopropyl)-dimethylammonio]-1-propanesulfonic acid (CHAPS) as a detergent followed by sucrose density gradient centrifugation according to the method originally described by McEnery et al. for rat liver mitochondria (McEnery et al. (1986) J. Biol. Chem. 259, 4642-4651), with minor modifications. The purified ATPase was reconstituted into proteoliposomes and then reconstituted into planar phospholipid bilayers by the modified fusion method (Hirata et al. (1986) J. Biol. Chem. 261, 9839-9843). A short-circuit current of up to 0.4 pA was induced by adding ATP, and this current was suppressed by the F1 ATPase inhibitor NaN3 or by a specific mitochondrial F0 inhibitor, oligomycin. The direction of the current corresponded to the flow of positive charges from the F1 side to the F0 side. All these facts clearly demonstrate that the mitochondrial F0F1 ATPase was successfully reconstituted into planar phospholipid bilayers, and the current was generated by the ATPase.  相似文献   

13.
Li L  Cheng JX 《Biochemistry》2006,45(39):11819-11826
We report a new type of gel-liquid phase segregation in giant unilamellar vesicles (GUVs) of mixed lipids. Coexisting patch- and stripe-shaped gel domains in GUV bilayers composed of DOPC/DPPC or DLPC/DPPC are observed by confocal fluorescence microscopy. The lipids in stripe domains are shown to be tilted according to the DiIC18 fluorescence intensity dependence on the excitation polarization. The patch domains are found to be mainly composed of DPPC-d62 according to the coherent anti-Stokes Raman scattering (CARS) images of DOPC/DPPC-d62 bilayers. When cooling GUVs from above the miscibility temperature, the patch domains start to appear between the chain melting and the pretransition temperature of DPPC. In GUVs containing a high molar percentage of DPPC, the stripe domains form below the pretransition temperature. Our observations suggest that the patch and stripe domains are in the Pbeta' and Lbeta' gel phases, respectively. According to the thermoelastic properties of GUVs described by Needham and Evans [(1988) Biochemistry 27, 8261-8269], the Pbeta' and Lbeta' phases are formed at relatively low and high membrane tensions, respectively. GUVs with high DPPC percentage have high membrane surface tension and thus mainly exhibit Lbeta' domains, while GUVs with low DPPC percentage have low membrane surface tension and form Pbeta' domains accordingly. Adding negatively charged lipid to the lipid mixtures or applying an osmotic pressure to GUVs using sucrose solutions releases the surface tension and leads to the disappearance of the Lbeta' gel phase. The relationship between the observed domains in free-standing GUV bilayers and those in supported bilayers is discussed.  相似文献   

14.
Hydration force and bilayer deformation: a reevaluation   总被引:20,自引:0,他引:20  
T J McIntosh  S A Simon 《Biochemistry》1986,25(14):4058-4066
The hydration repulsive force between lipid bilayers and the deformability of both gel and liquid-crystalline bilayers have been quantitated by an X-ray diffraction analysis of osmotically stressed liposomes. Both sampling theorem reconstructions and electron density distributions were calculated from diffraction data obtained from multilayers with applied osmotic pressures of 0-50 atm. The bilayer thickness and area per lipid molecule remain nearly constant (to within about 4%) in this pressure range, as adjacent bilayers move from their equilibrium separation in excess water to within 2-4 A of each other. This analysis indicates that the bilayers are relatively incompressible. This results differs from previously published X-ray diffraction studies of bilayer compressibility but agrees with direct mechanical measurements of the bilayer compressibility modulus. It is also found that the hydration repulsive force decays exponentially with separation between bilayers with a decay constant of 1.4 A for gel-state dipalmitoylphosphatidylcholine and 1.7 A for liquid-crystalline egg phosphatidylcholine bilayers. This implies that the exponential decay constant is not necessarily equal to the diameter of a water molecule, as has been previously suggested on experimental and theoretical grounds.  相似文献   

15.
Green tea contains a high concentration of such catechins as (-)-epicatechin (EC), (-)-epigallocatechin (EGC), (-)-epicatechin gallate (ECg), and (-)-epigallocatechin gallate (EGCg). Their biological activities have been evaluated by in vitro experiments using cultured cells or bacteria, but the order of activity of the various catechins differed with the study. We have been studying the interaction of tea catechins with lipid bilayers, and clarified that the number of hydroxyl groups on the B-ring, the presence of the galloyl moiety, and the stereochemical structure of each catechin govern their affinity for lipid bilayers. We investigated in this present study the effects of various external factors on the affinity of tea catechins for lipid bilayers by using liposomes as model membranes. The amount of tea catechins incorporated into the lipid bilayers increased with increasing salt concentration in an aqueous medium and decreased with increasing negative electric charge of the lipid bilayers. Furthermore, the amount of EGCg or ECg incorporated into the lipid bilayers increased with increasing EC concentration. These results reveal that the salt concentration in an aqueous medium, the electric charge of the membrane, and the presence of other catechins governed the affinity of tea catechins for the lipid bilayers.  相似文献   

16.
It has been suggested that the polyunsaturated omega-3 fatty acid, docosahexaenoic acid (DHA), can adopt unique closely packed arrays in lipid bilayers (Glomset and Applegate. (1986) J. Lipid Res. 27, 658-680). These conformations are predicted on the basis of molecular dynamics calculations and are in contrast to the expanded conformations characteristic of omega-6 unsaturated fatty acids. It has also been suggested that close packing of omega-3 acyl chains could have a substantial affect on the physical properties of lipid bilayers (e.g. permeability). We report here some experimental tests of these predictions. Surface pressure-area experiments have been carried out on DHA and its mixtures with stearic and oleic acids. At low surface pressures DHA is more expanded than oleic acid. Extrapolation to the high surface pressures characteristic of lipid bilayers indicates that the area per molecule of DHA is only marginally less than that for oleic acid. Thus there is no compelling evidence to suggest that the average area per molecule of the omega-3 fatty acid is substantially different from the omega-6 fatty acid at high surface pressures. Experiments also show that the permeability of bilayers to glucose and the rates of dissociation of pyrenyl cholesterol from bilayers were similar for bilayers containing DHA compared to bilayers containing oleic acid or linoleic acid.  相似文献   

17.
Potassium conductance of single muscle fibres from Rana esculenta was studied in isotonic K2SO4 solution under constant current conditions using double sucrose gap method. At room temperature the channel formation by gramicidin was much faster than that of the synthetic head to head covalently linked gramicidin dimer. The increase of temperature by 8-10 degrees C resulted in a considerable rise of both dimer- and gramicidin-induced conductances. The effect was much greater than in the case of bilayers indicating a remarkable entropy change in the muscle fibre membrane. The temperature dependence of adsorption was more pronounced than that of desorption: there was no effect on desorption of dimer and only 20% of the temperature-activated desorption of gramicidin irreversibly bound at room temperature.  相似文献   

18.
The effects of three so-called kosmotropic solutes, namely, betaine, sucrose, and choline chloride on 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine large unilamellar vesicles, were studied by measuring the generalized polarization (GP) for the fluorescence emission of the membrane partitioning probe Laurdan. The latter has been shown to be sensitive to the depth of water penetration into phospholipid bilayers. At equal osmotic pressures the three solutes produced different increments in GP, with a qualitative positive correlation. However, the increments in GP correlated also quantitatively with the increase of air-water surface tension caused by the three kosmotropes. Our findings suggest surface tension to determine the impact of these solutes on the lateral packing of the lipid bilayer. Based on the changes in area/lipid at different surface tensions, the equilibrium lateral pressure for a 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine bilayer at 25 degrees C was estimated to be approximately 34 mN/m.  相似文献   

19.
Reactive disulfide reagents (RDSs) with a biotin moiety have been synthesized and found to cause Ca2+ release from sarcoplasmic reticulum (SR) vesicles. The RDSs oxidize SH sites on SR proteins via a thiol-disulfide exchange, with the formation of mixed disulfide bonds between SR proteins and biotin. Biotinylated RDSs identified a 106-kDa protein which was purified by biotin-avidin chromatography. Disulfide reducing agents, like dithiothreitol, reverse the effect of RDSs and thus promoted active re-uptake of Ca2+ and dissociated biotin from the labeled protein indicating that biotin was covalently linked to the 106-kDa protein via a disulfide bond. Several lines of evidence indicate that this protein is not Ca2+, Mg2+-ATPase and is not a proteolytic fragment or a subunit of the 400-kDa Ca2+-ryanodine receptor complex (RRC). Monoclonal antibodies against the ATPase did not cross-react with the 106-kDa protein, and polyclonal antibodies against the 106-kDa did not cross-react with either the ATPase or the 400-kDa RRC. RDSs did not label the 400-kDa RRC with biotin. Linear sucrose gradients used to purify the RRC show that the 106-kDa protein migrated throughout 5-20% linear sucrose gradients, including the high sucrose density protein fractions containing 400-kDa RRC. Protease inhibitors diisopropylfluorophosphate used to prevent proteolysis of 400-kDa proteins did not alter the migration of 106-kDa in sucrose gradients nor the patterns of biotin labeling of the 106-kDa protein. Incorporation of highly purified 106-kDa protein (free of RRC) in planar bilayers revealed cationic channels with large Na+ (gNa+ = 375 +/- 15 pS) and Ca2+ (gCa2+ = 107.7 +/- 12 pS) conductances which were activated by micromolar [Ca2+]free or millimolar [ATP] and blocked by micromolar ruthenium red or millimolar [Mg2+]. Thus, the SR contains a sulfhydryl-activated 106-kDa Ca2+ channel with apparently similar characteristics to the 400-kDa "feet" proteins.  相似文献   

20.
Measurement and modification of forces between lecithin bilayers.   总被引:15,自引:8,他引:7       下载免费PDF全文
We probe in two different ways the competing attractive and repulsive forces that create lamellar arrays of the phospholipid lecithin when in equilibrium with pure water. The first probe involves the addition of low molecular weight solutes, glucose and sucrose, to a system where the phospholipid is immersed in a large excess of water. Small solutes can enter the aqueous region between bilayers. Their effect is first to increase and then to decrease the separation between bilayers as sugar concentration increases. We interpret this waxing and waning of the lattice spacing in terms of the successive weakening and strengthening of the attractive van der Waals forces originally responsible for creation of a stable lattice. The second probe is an "osmotic stress method," in which very high molecular weight neutral polymer is added to the pure water phase but is unable to enter the multilayers. The polymer competes for water with the lamellar lattice, and thereby compresses it. From the resulting spacing (determined by X-ray diffraction) and the directly measured osmotic pressure, we find a force vs. distance curve for compressing the lattice (or, equivalently, the free energy of transfer to bulk water of water between bilayers. This method reveals a very strong, exponentially varying "hydration force" with a decay distance of about 2 A.  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