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.
The transfer of 14C-labeled, reduced and carboxymethylated human apolipoprotein A-II (RCM-AII) between small unilamellar vesicles (SUV) has been investigated. Ion-exchange chromatography was used for rapid separation of negatively charged egg phosphatidylcholine (PC)/dicetyl phosphate donor SUV containing bound 14C-labeled RCM-AII from neutral egg PC acceptor SUV present in 10-fold molar excess. The kinetics of 14C-labeled RCM-AII transfer in incubations of up to 12 h at 37 degrees C are consistent with the existence of fast, slow, and apparently "nontransferrable" pools of SUV-associated apolipoprotein; the transfers from these pools occur on the time scales of seconds or less, hours, and days/weeks, respectively. For donor SUV (0.15 mg of phospholipid/mL reaction mixture) containing about 15 RCM-AII molecules per vesicle, the sizes of the fast, slow, and nontransferrable pools are 13, 69, and 18%, respectively. The transfer of RCM-AII from the slow kinetic pool follows first-order kinetics, and the half-time (t 1/2) is about 3 h. The different kinetic pools of SUV-associated RCM-AII probably reflect apoprotein in different conformations of the SUV surface. Increasing the number of RCM-AII per donor SUV enlarges the size of the fast pool and increases the t 1/2 of transfer from the slow pool. In contrast, raising the incubation temperature reduces the t 1/2 of slow transfer. The t 1/2 of RCM-AII transfer from the slow kinetic pool is inversely proportional to the acceptor/donor SUV ratio which suggests that the transfer of apoprotein molecules in this kinetic pool is mediated by SUV collisions.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The kinetics and mechanism of transfer of 14C-labeled human apolipoproteins A-I, A-II and C-III1 between small unilamellar vesicles (SUV) have been investigated. Ion exchange chromatography was used for rapid separation of negatively charged egg phosphatidylcholine (PC)/dicetyl phosphate donor SUV containing bound 14C-labeled apoprotein from neutral egg PC acceptor SUV present in 10-fold molar excess. The transfer kinetics of these apolipoproteins at 37 degrees C are consistent with the existence of fast, slow and apparently 'nontransferrable' pools of SUV-associated lipoprotein: the transfers from these pools occur on timescales of seconds (or less), minutes/hours and days/weeks, respectively. For donor SUV containing about 15 apoprotein molecules per vesicle and at a donor SUV concentration of 0.15 mg phospholipid/ml incubation mixture, the sizes of the fast kinetic pools for apolipoproteins A-I, A-II and C-III1 associated with donor SUV are 2, 10 and 11%, respectively. The sizes of the slow kinetic pools for these apolipoproteins are 16, 71 and 50%, respectively. The transfer of the various apolipoproteins from the slow kinetic pool follows first order kinetics and the half-time (t1/2) values are in the order: apo C-III1 less than apo A-I. Increasing the number of apoprotein molecules per donor SUV enlarges the size of the fast pool and increases the t1/2 of slow transfer. The differences in the kinetics of apolipoprotein transfer between SUV are consequences of the variations in the primary and secondary structures of the apolipoprotein molecules. The slow transfer of apoprotein molecules is mediated by collisions between donor and acceptor SUV; the rate is dependent on the apoprotein molecular weight with larger molecules transferring more slowly from donor SUV containing the same lipid/protein molar ratio. The hydrophobicity of the apoprotein molecule is also significant with less hydrophobic molecules transferring more rapidly. Further understanding of the differences in the kinetics of transfer of these apolipoproteins will require more knowledge of their secondary and tertiary structures.  相似文献   

3.
Effect of membrane additives on vesicle fusion   总被引:1,自引:0,他引:1  
A large variety of alkyl derivatives were found to either slow or block the low-temperature induced fusion of dipalmitoylphosphatidylcholine small unilamellar vesicles (DPPC SUV) when incorporated into the SUV bilayer at five mol%. Only corn oil was fusogenic.  相似文献   

4.
We have investigated the fusion of phospholipid vesicles induced by lysozyme and glyceraldehyde-3-phosphate dehydrogenase (GAPDH). Vesicles were composed of dimyristoylphosphatidylcholine/dioleoylphosphatidylethanolamine/ cholesterol (DMPC:DOPE:Chol, 2:1:1). Small unilamellar vesicles (SUV, diameter ca. 30 nm) obtained by extensive sonication or large unilamellar vesicles (LUV, diameters ranged from 100 to 400 nm) obtained by extrusion methods were used. Fusion of LUV induced by lysozyme and GAPDH was drastically decreased when the diameter of the vesicles increased over a value of 100 nm. Lysozyme effect was stopped at the aggregation step while GAPDH effect was stopped at the fusion (lipid mixing) step. Fusion of heterogeneous vesicle populations (SUV with LUV) was observed only with GAPDH and this happened only when the lipids were in the liquid-crystalline state.  相似文献   

5.
We have investigated the behavior of two populations of doxorubicin (DXR)-containing phospholipid vesicles with regard to various physical and pharmacological parameters. DXR-containing liposomes were prepared by ultrasonic irradiation, the lipid composition being phosphatidylglycerol (or phosphatidylserine), phosphatidylcholine and cholesterol. The vesicles were fractionated into oligolamellar vesicles (OLV) and small unilamellar vesicles (SUV) by preparative differential ultracentrifugation (150,000 x g for 1 h). Unentrapped DXR was removed by gel exclusion chromatography. OLV and SUV liposomes differed in size (mean diameters, 247 +/- 113 nm and 61 +/- 16 nm, respectively) and number of lamellae (two for OLV, one for SUV). Drug entrapment per unit of lipid was three to 5-fold higher in OLV than in SUV. In both liposome populations more than 95% of the entrapped drug was membrane-associated. Physical studies on these two vesicle populations revealed higher motional restriction and greater susceptibility to iodide-mediated fluorescence collisional quenching of DXR in the small vesicles. OLV showed superior stability in the presence of plasma as determined by the fraction of DXR retained by the vesicles. It was also found that the tissue distribution of DXR in SUV follows a pattern different from that of DXR in OLV and resembling that of soluble DXR. In accordance with these differences in patterns of tissue distribution, animal studies demonstrated that DXR in OLV is significantly less toxic than DXR in SUV and more effective in a tumor model with predominant involvement of the liver. These results indicate that vesicle size and/or number of lamellae play an important role in optimizing liposome-mediated delivery of DXR, and that oligolamellar liposomes are distinctively superior to small unilamellar liposomes when fluid phase formulations (Tm less than 37 degrees C) with bilayer-associated DXR are considered.  相似文献   

6.
The specific ultrasonic absorption coefficient per wavelength as a function of temperature in the vicinity of the phase transition of liposomes, composed of a 4:1 mixture of dipalmitoylphosphatidylcholine (DPPC) and dipalmitoylphosphatidylglycerol (DPPG), of different sizes was determined using an acoustic interferometer. Small unilamellar vesicles (SUV) and multilamellar vesicles (MLV) yielded results similar to those in the literature, viz., an absorption maximum at the transition temperature. Seven intermediate sizes including several size distributions of large unilamellar vesicles (LUV) were studied, yielding information on size dependencies of the temperatures at which the peaks occur, the widths at half peak amplitude, and the peak amplitudes. All liposome sizes except the SUV exhibited approximately the same transition temperature as did the largest MLV. The widths of the peaks were inversely related to liposome size, with a strong dependence for the smallest vesicles and an approach to independence for the largest vesicles. The amplitudes of the peaks exhibited a general increase with size with two exceptions, viz., the SUV and the vesicles with average diameters of 90-100 nm. It was also found that the membrane permeability increased near the transition temperature. The temperature dependencies of ultrasonic absorption and membrane permeability are compared.  相似文献   

7.
To investigate the role of membrane proteins in the fusion process, linear hydrophobic polypeptide gramicidin was used as fusogenic agent in small unilamellar vesicles (SUV) constituted of saturated lecithins. It was found that gramicidin, externally added to a suspension of vesicles, induces a reversible vesicles aggregation. When incorporated into the bilayer, gramicidin induces increase in vesicle size. The vesicle size increase was monitored by column chromatography and transmission electron microscopy. The process of vesicle size increase occurs only when the lipid membrane is in the gel state. A maximum is observed in the kinetics at a temperature of approx. 25 degrees C lower than the phase transition temperature of lipids. Higher rates of vesicle size increase are obtained as the lipid chain length increases. The process is accompanied by a release of internal vesicle content and by membrane lipid mixing.  相似文献   

8.
M E Haque  A J McCoy  J Glenn  J Lee  B R Lentz 《Biochemistry》2001,40(47):14243-14251
The effects of hemagglutinin (HA) fusion peptide (X-31) on poly(ethylene glycol)- (PEG-) mediated vesicle fusion in three different vesicle systems have been compared: dioleoylphosphatidylcholine (DOPC) small unilamellar vesicles (SUV) and large unilamellar vesicles (LUV) and palmitoyloleoylphosphatidylcholine (POPC) large unilamellar perturbed vesicles (pert. LUV). POPC LUVs were asymmetrically perturbed by hydrolyzing 2.5% of the outer leaflet lipid with phospholipase A(2) and removing hydrolysis products with BSA. The mixing of vesicle contents showed that these perturbed vesicles fused in the presence of PEG as did DOPC SUV, but unperturbed LUV did not. Fusion peptide had different effects on the fusion of these different types of vesicles: fusion was not induced in the absence of PEG or in unperturbed DOPC LUV even in the presence of PEG. Fusion was enhanced in DOPC SUV at low peptide surface occupancy but hindered at high surface occupancy. Finally, fusion was hindered in proportion to peptide concentration in perturbed POPC LUV. Contents leakage assays demonstrated that the peptide enhanced leakage in all vesicles. The peptide enhanced lipid transfer between both fusogenic and nonfusogenic vesicles. Peptide binding was detected in terms of enhanced tryptophan fluorescence or through transfer of tryptophan excited-state energy to membrane-bound diphenylhexatriene (DPH). The peptide had a higher affinity for vesicles with packing defects (SUV and perturbed LUV). Quasi-elastic light scattering (QELS) indicated that the peptide caused vesicles to aggregate. We conclude that binding of the fusion peptide to vesicle membranes has a significant effect on membrane properties but does not induce fusion. Indeed, the fusion peptide inhibited fusion of perturbed LUV. It can, however, enhance fusion between highly curved membranes that normally fuse when brought into close contact by PEG.  相似文献   

9.
Unilamellar vesicles of varying and reasonably uniform size were prepared from 1,2-dipalmitoyl-3-sn-phosphatidylcholine (DPPC) by the extrusion procedure and sonication. Quasi-elastic light scattering was used to show that different vesicle preparations had mean (Z-averaged) diameters of 1340, 900, 770, 630, and 358 A (sonicated). Bilayer-phase behavior as detected by differential scanning calorimetry was consistent with the existence of essentially uniform vesicle populations of different sizes. The response of these different vesicles to treatment with poly(ethylene glycol) (PEG) was monitored using fluorescence assays for lipid transfer, contents leakage, and contents mixing, as well as quasi-elastic light scattering. No fusion, as judged by vesicle contents mixing and change in vesicle size, was detected for vesicles of diameter greater than 770 A. The diameters of smaller vesicles increased dramatically when treated with high concentrations of PEG, although mixing of their contents could not be detected both because of their small trapped volumes and because of the extensive leakage induced in small vesicles by high concentrations of PEG. Lipid transfer was detected between vesicles of all sizes. We conclude the high bilayer curvature does encourage fusion of closely juxtaposed membrane bilayers but that highly curved vesicles appear also to rupture and form larger structures when diluted from high PEG concentration, a process that can be confused with fusion. Despite the failure of PEG to induce fusion of large, uncurved vesicles composed of a single phosphatidylcholine, these vesicles can be induced to fuse when they contain small amounts of certain amphiphathic compounds thought to play a role in cellular fusion processes. Thus, vesicles which contained 0.5 mol % L-alpha-lysopalmitoylphosphatidylcholine, 5 mol % platelet activating factor, or 0.5 mol % palmitic acid fused in the presence of 30%, 25%, and 20% (w/w) PEG, respectively. However, vesicles containing 1,2-dipalmitoyl-sn-glycerol, 1,2-dioleoyl-sn-glycerol, 1-oleoyl-2-acetyl-sn-glycerol, or monooleoyl-rac-glycerol at surface concentrations up to 5 mol % did not fuse in the presence or absence of PEG. There was no correlation between the abilities of these amphipaths to induce phase separation or nonlamellar phases and their abilities to support fusion of pure DPPC unilamellar vesicles in the presence of high concentrations of PEG. The results are discussed in terms of the type of disrupted lipid packing that could be expected to favor PEG-mediated fusion.  相似文献   

10.
We have employed both small unilamellar vesicles (SUV) and large unilamellar vesicles formed by the reverse phase evaporation technique (REV) to study the initial kinetics of membrane aggregation and fusion. Stopped flow measurements of the calcium-induced changes in the turbidity of SUV and REV, formed from 1:1 (mol/mol) mixtures of bovine phosphatidylserine (PS) and Escherichia coli phosphatidylethanolamine (PE), were used to follow particle aggregation. Simultaneous measurements of the fluorescence resonance energy transfer from N-(7-nitro2,1,3-benzoxadiazol-4-yl) (NBD)-PE to rhodamine (Rho)-PE incorporated into the vesicle bilayers established that 1) both initial aggregation and fusion can be described as a bimolecular process and 2) the rate-limiting step of membrane fusion is aggregation. Thus fusion takes place in the microsecond time domain. Parallel experiments, which simultaneously measured aggregation and the dequenching of encapsulated carboxyfluorescein (CF) in the presence and absence of antifluorescein antibodies in the suspension medium, established that the small unilamellar vesicles rapidly lose their contents of CF as they fuse. On the other hand, the first few cycles of fusion of the large unilamellar vesicles are nonleaky, but leakage develops within 1-2 s as the particles grow in size. Thus the results demonstrate that the SUV are poor models for the study of nonleaky fusion, while the REV must be carefully tested before unambiguous interpretation of fusion assays involving the formation of tight complexes (such as the terbium-dipicolinic assay) can be made. NBD-PE undergoes very rapid, Ca2+-promoted changes in quantum yield which can obscure the resonance energy transfer signals. Thus data from the NBD-PE/Rho-PE energy transfer pair must be carefully scrutinized for artifacts.  相似文献   

11.
We have measured the phase behavior of mixed dipentadecanoylphosphatidylglycerol (DC15PG)/dimyristoylphosphatidylcholine (DMPC) small unilamellar vesicles (SUV) in the presence of saturating (greater than 98% occupancy of binding sites) concentrations of bovine prothrombin fragment 1 and 5 mM Ca2+. Binding of fragment 1 in the presence of Ca2+ was verified by an increase in 90 degrees light scattering. Only in the cases of DC15PG/DMPC SUV below their phase transition and of pure DMPC SUV were such light scattering measurements not reversible upon addition of ethylenediaminetetraacetic acid to complex Ca2+. Phase-behavior changes of DC15PG/DMPC SUV as monitored by diphenylhexatriene fluorescence anisotropy occurred in concert with the binding of fragment 1. The major effects of peptide binding on SUV phase behavior were to raise the phase-transition temperature by 2-15 degrees C, depending on vesicle composition, and, in general, to make the phase diagram for these small vesicles closely resemble that of large vesicles. No evidence was obtained for the existence of lateral membrane domains with distinct compositions induced by the binding of prothrombin fragment 1 plus Ca2+. Surprisingly, fragment 1 without Ca2+ also altered the phase behavior of DC15PG/DMPC SUV. Most striking was the effect of fragment 1 (with or without Ca2+) on DMPC SUV phase behavior. Freeze-fracture electron microscopy demonstrated that pure DMPC vesicles were induced to fuse in the presence of fragment 1, while vesicles containing DC15PG remained intact. The rate of DMPC SUV fusion (followed by 90 degrees light scattering) increased with increasing fragment 1 concentration but was not saturable up to 40 microM fragment 1, suggesting a weak, nonspecific interaction between fragment 1 and the neutral phospholipid vesicle.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Cupp D  Kampf JP  Kleinfeld AM 《Biochemistry》2004,43(15):4473-4481
Understanding the mechanism that governs the transport of long chain free fatty acids (FFA) across lipid bilayers is critical for understanding transport across cell membranes. Conflicting results have been reported for lipid vesicles; most investigators report that flip-flop occurs within the resolution time of the method (<5 ms) and that dissociation from the membrane is rate limiting, while other studies find that flip-flop is rate limiting and on the order of seconds. We have reinvestigated this problem and find that the methods used in studies reporting rapid flip-flop have not been interpreted correctly. We find that accurate information about transport of FFA across lipid vesicles requires that FFA be delivered to the vesicles as complexes with albumin (BSA). For example, we find that stopped-flow mixing of uncomplexed FFA with small unilamellar vesicles (SUV) containing pyranine yields the very fast influx rates reported previously (>100 s(-1)). However, these influx rates increase linearly with lipid vesicle concentration and can therefore not, as previously interpreted, represent flip-flop. In contrast, measurements of influx rates in SUV and giant unilamellar vesicles performed with oleate-BSA complexes reveal no dependence on vesicle concentration and yield influx rate constants of approximately 4 and approximately 0.5 s(-1), respectively. Rate constants for efflux and dissociation were determined from the transfer of oleate from vesicles to BSA and reveal similar influx and efflux but dissociation rate constants that are approximately 5-10-fold greater. We conclude that flip-flop is rate limiting for transport of FFA across lipid vesicles and slows with an increasing radius of curvature. These results, in contrast to those reporting that flip-flop is extremely fast, indicate that the lipid bilayer portion of biological membranes may present a significant barrier to transport of FFA across cell membranes.  相似文献   

13.
Monovalent cation-induced fusion of acidic phospholipid vesicles   总被引:1,自引:0,他引:1  
Fusion of small unilamellar vesicles (SUV) consisting of dimyristoylphosphatidylcholine (DMPC), dimyristoylphosphatidylglycerol (DMPG) and phosphatidylglycerol (PG) from egg yolk, dipalmitoylphosphatidylserine (DPPS) and phosphatidylserine (PS) from bovine brain was studied as a function of monovalent cation concentration. Fusion was detected by measuring the changes in the excimer to monomer fluorescence intensity ratio (IE/M) of pyrene-labeled phospholipid analogues upon fusion of the pyrene-labeled and unlabeled vesicles. No fusion was observed from vesicles consisting of DMPC, PS from bovine brain or PG from egg yolk upon addition of NaCl (up to 1 M). However, considerable fusion was evident for vesicles consisting of DMPG or DPPS upon addition of monovalent cations (300 mM to 1 M). Fusion kinetics were fast reaching a plateau after 5 min of addition of cations. The order of efficiency of different monovalent cations to induce the fusion of DMPG vesicles as judged by the changes of the IE/M ratio was Li+ greater than Na+ greater than K+ greater than Cs+. DSC-scan of sonicated DMPG vesicles showed, in the absence of salt, a phase transition at 19.2 degrees C with enthalpy of 1.1 kcal.mol-1. After incubation in the presence of 600 mM NaCl the DSC scan showed a narrow phase transition at 24.1 degrees C with enthalpy of 6.9 kcal.mol-1 and a pronounced pretransition, both supporting that the fusion of the vesicles had occurred in the presence of NaCl. The results indicate that sonicated vesicles consisting of acidic phospholipids with fully saturated fatty acids fuse in the presence of monovalent cations, whereas those containing unsaturated fatty acids do not.  相似文献   

14.
We have investigated the temperature dependence of the fusion of phospholipid vesicles composed of pure bovine brain phosphatidylserine (PS) induced by Ca2+ or Mg2+. Aggregation of the vesicles was monitored by 90 degrees light-scattering measurements, fusion by the terbium/dipicolinic acid assay for mixing of internal aqueous volumes, and release of vesicle contents by carboxyfluorescein fluorescence. Membrane fluidity was determined by diphenylhexatriene fluorescence polarization measurements. Small unilamellar vesicles (SUV, diameter 250 A) or large unilamellar vesicles (LUV, diameter 1000 A) were used, and the measurements were done in 0.1 M NaCl at pH 7.4. The following results were obtained: (1) At temperatures (0-5 degrees C) below the phase transition temperature (Tc) of the lipid, LUV (PS) show very little fusion in the presence of Ca2+, although vesicle aggregation is rapid and extensive. With increasing temperature, the initial rate of fusion increases dramatically. Leakage of contents at the higher temperatures remains limited initially, but subsequently complete release occurs as a result of collapse of the internal aqueous space of the fusion products. (2) SUV (PS) are still in the fluid state down to 0 degree C, due to the effect of bilayer curvature, and fuse rapidly in the entire temperature range from 0 to 35 degrees C in the presence of Ca2+. The initial rate of leakage is low relative to the rate of fusion. At higher temperatures (15 degrees C and above), subsequent collapse of the vesicles' internal space causes complete release.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
We have studied the formation of a supported bilayer containing both cationic and zwitterionic lipids by fusion of small unilamellar vesicles (SUV) onto the solid surface at low salt conditions using a combination of attenuated total reflection infrared (ATR-IR) and deuterium NMR spectroscopy with microcalorimetry. The data suggest that a significant cationic lipid asymmetry between the outer (distal) and the inner (proximal) monolayer of a supported bilayer results under conditions of prolonged incubation times of the solid support with the SUV coating solution. For a SUV composition of DPPC/DHDAB (4:1) we observed an enrichment of the cationic component in the proximal monolayer of up to 200% compared to the distal monolayer after 12 h incubation. It is suggested that the electrostatic potential arising from the solid surface is the driving force for the creation of this asymmetry by means of directed flip-flop between the monolayers and/or by temporary fusion between SUV from the bulk with the supported bilayer.  相似文献   

16.
Addition of bovine adrenal cytochrome P-450scc to small unilamellar dioleoylphosphatidylcholine vesicles (DOPC-SUV) produces a complex sequence of interactions, indicating exceptional cytochrome mobility. First, cholesterol transfer from cytochrome to vesicles indicated rapid dissociation of P-450scc oligomers and integration of monomers into the membrane (delta A 390-420 nm; t1/2 = 2 s). After 10-15 s, P-450scc-induced aggregation of the vesicles starts, as indicated by increased turbidity (delta A 448 or 520 nm; complete in 6-8 min). Fluorescence quenching experiments indicate that this aggregation does not lead to measurable vesicle fusion during this period. Aggregation is prevented by mild heat denaturation of P-450scc, by addition of anti-P-450scc IgG, and also by 1:1 complex formation with the electron donor adrenodoxin (ADX). P-450scc, therefore, links two vesicles through two separate domains involved in, respectively, membrane integration (lipophilic) and ADX binding (charged). Although completely bound by DOPC-SUV, as evidenced by Sephadex elution, P-450scc has access within 1 min to cholesterol in secondary SUV. This is indicated by spectral changes (cholesterol complex formation) and by metabolism of secondary vesicle cholesterol. Since cholesterol equilibrates slowly between vesicles (t1/2 = 1-2 h), these changes arise from P-450scc transfer. This transfer was maximally slowed after a 5-min preincubation with primary vesicles, reflecting more extensive integration into the membrane than is necessary for the rapid initial cholesterol transfer to P-450scc. P-450scc transfer probably results from simultaneous interaction of P-450scc with two vesicles that may also initiate aggregation. Weaker integration into primary dimyristoylphosphatidylcholine vesicles facilitates exchange but prevents aggregation. Integration and aggregation are both enhanced by incorporation of 10% phosphatidylinositol into SUV, while exchange is slowed. This mobility of P-450scc is most probably a consequence of the absence of amino-terminal anchoring. P-450scc-induced association of inner mitochondrial membrane segments may contribute to the exceptionally vesiculated structure of adrenal and ovarian mitochondria that parallels increased P-450scc content.  相似文献   

17.
Summary Fusion between unilamellar vesicles of both egg phosphatidylcholine and bovine phosphatidylserine was induced by polyethylene glycol. Aggregation and fusion events were monitored by electron microscopy and turbidity measurements. The threshold concentration of polyethylene glycol for aggregation and fusion is found to be independent of lipid concentration. Typically, aggregation of phosphatidylcholine vesicles starts at 2.5% (wt/wt) polyethylene glycol, but fusion is not significant until the polyethylene glycol concentration reaches 35%. Multilamellar vesicles were formed as a result of fusion.Abbreviations PEG Polyethylene glycol - IMP Intramembranous particle - PC Phosphatidylcholine - PS Phosphatidylserine - SUV Small unilamellar vesicles - MLV Multilamellar vesicles - DPPC Dipalmitoyl phosphatidylcholine - DSC Differential scanning calorimetry  相似文献   

18.
G Nemecz  F Schroeder 《Biochemistry》1988,27(20):7740-7749
The fluorescent sterol delta 5,7,9(11),22-ergostatetraen-3 beta-ol (dehydroergosterol) was investigated as a cholesterol analogue to examine sterol domains in and spontaneous exchange of sterol between 1-palmitoyl-2-oleoylphosphatidylcholine (POPC) small unilamellar vesicles (SUV). Fluorescence lifetime, acrylamide quenching analyses, and intermembrane exchange kinetics were consistent with the presence of at least two sterol domains in POPC. Fluorescence lifetime was determined by phase and modulation fluorescence spectroscopy and analyzed by nonlinear least-squares as well as continuous distributional analyses. Both methods demonstrated that pure dehydroergosterol in POPC SUV had two lifetime components (C) and fractional intensities (F) near C1 = 0.851 ns (F1 0.96) and C2 = 2.668 ns (F2 0.004). In contrast to component C1, the center of lifetime distribution, fractional intensity, and peak width of dehydroergosterol lifetime component C2 was dependent on the polarity of the medium and vesicle curvature. The sterol domain corresponding to dehydroergosterol component C2 was preferentially quenched by acrylamide. Acrylamide quenching of dehydroergosterol fluorescence demonstrated that the two lifetime components of dehydroergosterol were not due to transbilayer sterol domains with different lifetimes. In a spontaneous exchange assay not requiring separation of donor and acceptor SUV, the lifetime component C2, but not C1, shifted to a shorter lifetime with altered distributional width. The kinetics of these lifetime and distributional width changes best fitted a two-exponential function, with a fast exchange rate constant K1 = 0.0325 min-1, t1/2 = 21.3 min, and a slow rate constant k2 = 0.00275 min-1, t1/2 = 261 min. The fast exchanging pool correlates with the longer lifetime component C2. These kinetics were confirmed both by dehydroergosterol exchange measured with fluorescence intensity and by [3H]cholesterol exchange. In summary, lifetime, distributional width, acrylamide quenching, and classical exchange assay data are consistent with the presence of at least two pools of sterol in POPC SUV.  相似文献   

19.
Fusion of exocytotic vesicles with the plasma membrane gives rise to an increase in membrane surface area, whereas the surface area is decreased when vesicles are internalized during endocytosis. Changes in membrane surface area, resulting from fusion and fission of membrane vesicles, can be followed by monitoring the corresponding proportional changes in membrane capacitance. Using the cell‐attached configuration of the patch‐clamp techniques we were able to resolve the elementary processes of endo‐ and exocytosis in yeast protoplasts at high temporal and spatial resolution. Spontaneous capacitance changes were predominantly in the range of 0.2–1 fF which translates to vesicle diameters of 90–200 nm. The size distribution revealed that endocytotic vesicles with a median at about 132 nm were smaller than exocytotic vesicles with a median at 155 nm. In energized and metabolizing protoplasts, endo‐ and exocytotic events occurred at frequencies of 1.6 and 2.7 events per minute, respectively. Even though these numbers appear very low, they are in good agreement with the observed growth rate of yeast cells and protoplasts.  相似文献   

20.
This study aimed to investigate the influence of the preparation conditions on the performance of an ethosomal formulation for topical delivery of the local anesthetic agent, benzocaine (BZC). Ethosomes were prepared with different techniques, such as thin-layer evaporation, freezing and thawing, reverse-phase evaporation, extrusion and sonication, obtaining, respectively, multilayer vesicles (MLVs), frozen and thawed MLV (FATMLV), large unilamellar vesicles (LUVs), and small unilamellar vesicles (SUVs). The obtained vesicles were characterized for morphology, size, zeta potential, and entrapment efficiency (EE%), and their stability was monitored during storage at 4°C. In vitro permeation properties from gels incorporating drug ethosomal dispersions were evaluated in vitro by using artificial lipophilic membranes, while their anesthetic effect was determined in vivo on rabbits. The results suggested that the vesicle preparation method plays an important role in affecting the properties and effectiveness of ethosomal formulations. MLVs and LUVs exhibited higher drug EE% and better stability than FATMLV and SUV vesicles. The In vitro drug permeation rate was directly related to the vesicle EE% and varied in the order MLV>LUV≈FATMLV>SUV. The therapeutic efficacy of BZC ethosomal formulations was significantly improved with respect to the corresponding BZC solution. The best results, in terms of enhanced intensity of anesthetic effect, were given by formulations containing MLVs and LUVs, and the order of effectiveness was MLV≈LUV>FATMLV≈SUV, rather similar to that found in permeation studies. On the contrary, unexpectedly, the effectiveness order in increasing the duration of drug action was SUV≥MLV>LUV≈FATMLV. The highest efficacy of SUVs was probably due to the more intimate contact with the epithelium due to their greatest surface area, which allowed the longest extension of drug therapeutic action. The overall results suggest that a suitably developed ethosomal formulation of BZC can be of actual value for improving its clinical effectiveness in topical anesthesia.  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