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 539 毫秒
1.
Purified phosphatidylcholine exchange protein from bovine liver was used to exchange rat liver microsomal phosphatidylcholine for egg phosphatidylcholine. It was found that at 25 and 37°C rat liver microsomal phosphatidylcholine was completely and rapidly available for replacement by egg phosphatidylcholine. In contrast, phosphatidylcholine in vesicles prepared from total microsomal lipids could only be exchanged for about 60%. At 8 and 0°C complex exchange kinetics were observed for phosphatidylcholine in rat liver microsomes. The exchange process had neither effect on the permeability of the microsomal membrane to mannose 6-phosphate, nor on the permeability of the phosphatidylcholine vesicles to neodymium (III) cations.Purified phospholipase A2 from Naja naja could hydrolyze some 55–60% of microsomal phosphatidylcholine at 0°C, but 70–80% at 37°C. Microsomal phosphatidylcholine, remaining after phospholipase treatment at 37°C, could be exchanged for egg phosphatidylcholine at 37°C, but at a slower rate than with intact microsomes. Microsomal phosphatidylcholine remaining after phospholipase treatment at 0 and 37°C had a lower content of arachidonic acid than the original phosphatidylcholine.These results are discussed with respect to the localization and transmembrane movement of phosphatidylcholine in liver microsomes.  相似文献   

2.
The lipid distribution in binary mixed membranes containing charged and uncharged lipids and the effect of Ca2+ and polylysine on the lipid organization was studied by the spin label technique. Dipalmitoyl phosphatidic acid was the charged, and spin labelled dipalmitoyl lecithin was the uncharged (zwitterionic) component. The ESR spectra were analyzed in terms of the spin exchange frequency, Wex. By measuring Wex as a function of the molar percentage of labelled lecithin a distinction between a random and a heterogeneous lipid distribution could be made. It is established that mixed lecithinphosphatidic acid membranes exhibit lipid segregation (or a miscibility gap) in the fluid state. Comparative experiments with bilayer and monolayer membranes strongly suggest a lateral lipid segregation. At low lecithin concentration, aggregates containing between 25% and 40% lecithin are formed in the fluid phosphatidic acid membrane. This phase separation in membranes containing charged lipids is understandable on the basis of the Gouy-Chapman theory of electric double layers.In dipalmitoyl lecithin and in dimyristoyl phosphatidylethanolamine membranes the labelled lecithin is randomly distributed above the phase transition and has a coefficient of lateral diffusion of D = 2.8·10?8 cm2/s at 59°C.Addition of Ca2+ dramatically increases the extent of phase separation in lecithin-phosphatidic acid membranes. This chemically (and isothermally) induced phase separation is caused by the formation of crystalline patches of the Ca2+-bound phosphatidic acid. Lecithin is squeezed out from these patches of rigid lipid. The observed dependence of Wex on the Ca2+ concentration could be interpreted quantitatively on the basis of a two-cluster model. At low lecithin and Ca2+ concentration clusters containing about 30 mol% lecithin are formed. At high lecithin or Ca2+ concentrations a second type of precipitation containing 100% lecithin starts to form in addition. A one-to-one binding of divalent ions and phosphatidic acid at pH 9 was assumed. Such a one-to-one binding at pH 9 was established for the case of Mn2+ using ESR spectroscopy.Polylysine leads to the same strong increase in the lecithin segregation as Ca2+. The transition of the phosphatidic acid bound by the polypeptide is shifted from Tt = 47.5° to Tt = 62°C. This finding suggests the possibility of cooperative conformational changes in the lipid matrix and in the surface proteins in biological membranes.  相似文献   

3.
Small unilamellar vesicles of egg phosphatidylcholine (PC) or dimyristoylphosphatidylcholine, mixed with small unilamellar vesicles labelled with 2-(10-(1-pyrene)decanoyl)phosphatidylcholine, exhibit a constant average size and excimer to monomer (E/M) ratio for several hours when incubated at pH 3.6 at a temperature higher than the phase transition temperature (Tc) of the lipids. Addition of bovine serum albumin to this system produces a transient turbidity increase, a fast decrease in the E/M ratio, a partial loss of vesicle-entrapped [14C]sucrose and a measurable leak-in of externally added sucrose. Sepharose 4B filtration of the system demonstrates that the E/M ratio decrease is strictly paralleled by the formation of liposomes which exhibit a low E/M ratio and a hydrodynamic radius larger than that of small unilamellar vesicles. These data demonstrate that the E/M ratio decrease can be unequivocally ascribed to a vesicle-vesicle fusion process induced by serum albumin. The rate of serum-albumin induced fusion of small unilamellar vesicles is: (a) maximal at a stoichiometric ratio of approx. 2 albumins per vesicle: (b) sensitive to the nature of the lipid and; (c) not altered when human serum albumin replaces bovine serum albumin. The rate of albumin-induced fusion of dimyristoylphosphatidylcholine small unilamellar vesicles is higher below the Tc of the lipid and increases with temperature above the Tc. The formation of protein-bound aggregates with defined stoichiometries and a high local vesicle concentration, as well as changes in the local degree of hydration, are proposed to be the driving forces for the protein-induced vesicle fusion in this system.  相似文献   

4.
Raman spectroscopic frequency differences between selected carbon-carbon stretching modes of lipid hydrocarbon chains were determined as a function of temperature for use in monitoring lipid phase transition behavior and acyl chain disorder in both multilamellar and single-wall vesicles. Transition temperatues detected by this procedure for pure dipalmitoyl phosphatidylcholine and dimyristoyl phosphatidylcholine multilayers were observed at 39±1 °C and 23±1 °C, respectively. Although the phase transition for unilamellar vesicles of dipalmitoyl phosphatidylcholine occurred at nearly the same temperature as the multilayers, the crystal-liquid crystalline transition for the single-shell vesicles appeared to span a slightly broader temperature range, a characteristic consistent with irregularities in the packing arrangement of the hydrocarbon chains. Within the precision of the Raman spectroscopic method, however, the temperature behavior of both the multilamellar and the unilamellar dimyristoyl phosphatidylcholine assemblies appeared nearly identical. The temperature profile for the Raman frequency differences of an excess water sonicate of 25 mol percent cholesterol in dipalmitoyl phosphatidylcholine served as an example of the effect upon lipid phase transition characteristics of a bilayer component intercalated between the acyl chains. For this particular cholesterol-lipid system the phase transition was broadened over a 30 °C temperature range, in contrast to the narrow 5?4 °C range observed for pure multilayer and single-shell vesicle particles.  相似文献   

5.
We report here on a series of studies aimed at characterization of the structural and dynamical properties of the synthetic lipid diphytanoyl phosphatidylcholine, in multilamellar dispersions and vesicle suspensions.This lipid exhibits no detectable gel to liquid crystalline phase transition over a large temperature range (?120°C to +120°C).Examination of proton nuclear magnetic resonance (NMR) free induction decays obtained from multilayer dispersions of diphytanoyl phosphatidylcholine provided an estimate of the methylene proton order parameter. The estimated magnitude of 0.21 is comparable to those determined for other phospholipids.Sonication of aqueous dispersions of diphytanoyl phosphatidylcholine led to formation of bilayer vesicles as determined by the measurement of the outer/inner choline methyl proton resonances, vesicle sizes in electron micrographs, and comparison of proton NMR linewidths between multilayer and sonicated dispersions. Ultracentrifugation studies of diphytanoyl phosphatidylcholine vesicles in H2O and 2H2O media yielded a value of 1.013 ± 0.026 ml/g for the partial specific volume of this lipid.We have measured spin lattice relaxation rates for the methyl and methylenemethyne protons of the hydrocarbon chains of diphytanoyl phosphatidylcholine in bilayer vesicles over a range of temperatures and at two NMR frequencies (100 and 220 MHz). The observed relaxation rates for the methylene protons in this system were approximately twice those previously reported for dipalmitoyl phosphatidylcholine at comparable temperatures and resonance frequencies, whereas the relaxation rates measured for the methyl protons were greater than those of the straight chain lipid by an order of magnitude.Measurement of the spin lattice relaxation rates of the hydrocarbon protons of the diphytanoyl phosphatidylcholine in a 10 mol% mixture of the branched-chain lipid in a deuterated host lipid, diperdeuteropalmitoyl phosphatidylcholine, showed a discontinuity in the temperature dependence of the proton NMR longitudinal relaxation rates of the branched-chain lipid in the region of the gel to liquid crystalline phase transition temperature of the deuterated dipalmitoyl phosphatidylcholine host lipid. This result may be taken as evidence of lateral phase separation of a liquid cyrstalline phase enriched in diphytanoyl phosphatidylcholine from a gel phase enriched in diperdeuteropalmitoyl phosphatidylcholine at temperatures below the phase transition temperature of deuterated host lipid. This conclusion is supported by the observation of an abrupt change in the hydrocarbon methylene linewidth (at 100 MHz) of 10 mol% diphytanoyl phosphatidylcholine in diperdeuteropalmitoyl phosphatidylcholine over the temperature range where lateral phase separation is taking place according to differential thermograms.  相似文献   

6.
Processes occurring in dispersions of dimyristoyl phosphatidylcholine containing myristic acid have been studied by light scattering of dilute dispersions (concn. ≤ 1 mg/ml) at temperatures above and below the phase transition temperatures of these dispersions. The transition temperatures increase with increasing mol fraction of myristic acid. Above these temperatures, vesicles with different mol fractions of myristic acid exchange lipid molecules. The exchange process leads to vesicles having phase transition temperatures and radii, which are both intermediate between the initial transitions and radii, respectively. In contrast with the observations above the phase transitions, it was found that when dimyristoyl phosphatidylcholine/myristic acid vesicles were cooled to a few degrees below the phase transition, larger particles were formed. These observations are consistent with a mechanism consisting of vesicle aggregation followed by fusion of the aggregated vesicles. The aggregation process is of second order in the vesicle concentration, and its rate increases with increasing mol fraction of myristic acid.  相似文献   

7.
Lipid mixing between vesicles functionalized with SNAREs and the cytosolic C2AB domain of synaptotagmin-1 recapitulates the basic Ca2+ dependence of neuronal exocytosis. However, in the conventional ensemble lipid mixing assays it is not possible to discriminate whether Ca2+ accelerates the docking or the fusion of vesicles. Here we report a fluorescence microscopy-based assay to monitor SNARE-mediated docking and fusion of individual vesicle pairs. In situ measurement of the concentration of diffusing particles allowed us to quantify docking rates by a maximum-likelihood approach. This analysis showed that C2AB and Ca2+ accelerate vesicle-vesicle docking with more than two orders of magnitude. Comparison of the measured docking rates with ensemble lipid mixing kinetics, however, suggests that in most cases bilayer fusion remains the rate-limiting step. Our single vesicle results show that only ∼60% of the vesicles dock and only ∼6% of docked vesicles fuse. Lipid mixing on single vesicles was fast (tmix < 1 s) while an ensemble assay revealed two slow mixing processes with tmix ∼ 1 min and tmix ∼ 20 min. The presence of several distinct docking and fusion pathways cannot be rationalized at this stage but may be related to intrasample heterogeneities, presumably in the form of lipid and/or protein composition.  相似文献   

8.
The lateral diffusion of the excimer-forming probe pyrene decanoic acid has been determined in erythrocyte membranes and in vesicles of the lipid extracts. The random walk of the probe molecules is characterized by their jump frequency, vj, within the lipid matrix. At T = 35°C a value of vj = 1.6 · 103 s?1 is found in erythrocyte membranes. A somewhat slower mobility is determined in vesicles prepared from lipid extracts of the erythrocyte membrane. Depending on structure and charge of the lipids we obtain jump frequencies between 0.8 · 108 s?1 and 1.5 · 108 s?1 at T = 35°C. The results are compared with jump frequencies yielded in model membranes.The mobility of molecules perpendicular to the membrane surface (transversal diffusion) is investigated. Erythrocyte ghosts doped with pyrene phosphatidylcholine were mixed with undoped ghosts in order to study the exchange kinetics of the probe molecule. A fast transfer between the outer layers of the ghost cells (τ12 = 1.6 min at T = 37°C) is found. The exchange process between the inner and the outer layer of one erythrocyte ghost (flip-flop process) following this fast transfer occurs with a half-life time value of t12 = 100 min at T = 37°C.The application of excimer-forming probes presumes a fluid state of the membrane. Therefore we investigated the phase transition behaviour using the excimer technique. Beside a thermotropic phase transition at T = 23°C and T = 33°C we observed an additional fluidity change at T = 38°C in erythrocyte ghosts. This transition is attached to a separation of the boundary lipid layer from the intrinsic proteins. No lipid phase transition is observed in liposomes from isolated extracts of the erythrocyte membrane with our methods.  相似文献   

9.
The protein-induced lipid transfer between phosphatidylcholine vesicles was investigated. Measurements of the degree of polarization at single vesicles were made by flow cytometry using diphenylhexatriene as the optical probe. Vesicles differing in phase transition temperature could be distinguished by their degree of polarization at a temperature where one population was in the fluid (T > Tt) and the other one in the quasi-crystalline (T < Tt) state. Besides vesicles containing exchanged lipids we also observed fractions of unaffected vesicles. The lipid exchange was visualized directly by freeze-fracture electron microscopy. The characteristic ‘ripple’ structure of phosphatidylcholine vesicles disappeared upon exchange with lipid in the fluid state.  相似文献   

10.
Dimyristoyl phosphatidylcholine rapidly exchanges between vesicles at 37°C without vesicle fusion.The rate of the transbilayer movement of dimyristoyl phosphatidylcholine in sonicated vesicles has been measured employing 13C NMR using N-13CH3? labeled lipids which are introduced into the outer monolayer of non-labeled vesicles by a phosphatidylcholine exchange protein. The rate of transbilayer movement of dimyristoyl phosphatidylcholine shows a distinct maximum (halftime 4 h) in the temperature range at which the hydrocarbon phase transition occurs.The activation energy of the flip-flop rate above the phase transition is 23.7 ± 2.0 kcal/mol.  相似文献   

11.
Small unilamellar vesicles comprised of a mixture of phosphatidylethanolamine/phosphatidylcholine/cholesterol (3 : 1 : 2) fuse to form large multilamellar vesicles on increasing the temperature from 0 to 50°C. This event is associated with the appearance of lipidic particles at the fusion sites, consistent with a role as intermediary structures during the fusion process. Further, for phosphatidylcholine/cardiolipin (1 : 1) liposomes in the presence of Mn2+ a direct relationship between lipidic particles and the hexagonal (HII) phase is demonstrated which suggests that lipidic particles can also occur as intermediaries between bilayer and hexagonal (HII) structures.  相似文献   

12.
A study has been conducted of the interaction of the lytic toxin delta-haemolysin with vesicles of phospholipid, using electron microscopy, fluorescence depolarisation and excimer fluorescence. The peptide is shown to be a fusogen towards phosphatidylcholine vesicles in fluid phases. In the presence of gel phase lipid, fusion between fluid and gel phases is not seen. Fluid phase lipid vesicles are fused together to form large multilamellar structures, and initial vesicle size does not appear to be important since small unilamellar vesicles and large unilamellar vesicles are similarly affected. Fusogenic activity of delta-haemolysin is compared to that of melittin. The former is a progressive fusogen for fluid phase lipid, while the latter causes vesicle fusion in a manner related to occurrence of a lipid phase transition.  相似文献   

13.
Phospholipase A2 will act on dipalmitoyl phosphatidylcholine as substrate when the phospholipid is part of a mixed micelle with Triton X-100 at a molar ratio of Triton to phospholipid of 2:1 or greater. Kinetic studies at high molar ratios of Triton X-100 to phospholipid are reported and show that the binding of phospholipase A2 to substrate depends on the total concentration of Triton X-100 and phospholipid, but that the rate of enzymatic catalysis decreases proportionally to the Triton X-100 concentration. These results are interpreted in terms of a model involving surface dilution kinetics. The relationship of this model to that of competitive inhibition is discussed. In addition, the activity of phospholipase A2 towards dipalmitoyl phosphatidylcholine and dimyristoyl phosphatidylcholine at different temperatures is reported, and the results show a direct effect of the thermotropic phase transition of dipalmitoyl phosphatidylcholine on enzymatic activity.  相似文献   

14.
A study has been conducted of the interaction of the lytic toxin δ-haemolysin with vesicles of phospholipid, using electron microscopy, fluorescence depolarisation and excimer fluorescence. The peptide is shown to be a fusogen towards phosphatidylcholine vesicles in fluid phases. In the presence of gel phase lipid, fusion between fluid and gel phases is not seen. Fluid phase lipid vesicles are fused together to form large multilamellar structures, and initial vesicle size does not appear to be important since small unilamellar vesicles and large unilamellar vesicles are similarly affected. Fusogenic activity of δ-haemolysin is compared to that of melittin. The former is a progressive fusogen for fluid phase lipid, while the latter causes vesicle fusion in a manner related to occurrence of a lipid phase transition.  相似文献   

15.
1. A comparative study has been made of the effects of the fusogens glycerol monooleate and dimethylsulphoxide on the polymorphic phase behaviour of dipalmitoyl phosphatidylcholine and dipalmitoyl phosphatidylethanolamine by differential scanning calorimetry and 31P-NMR techniques. 2. Glycerol monooleate induces a reduction in the temperature, cooperativity and enthalpy of the gel to liquid-crystal transitions of dipalmitoyl phosphatidylcholine and dipalmitoyl phosphatidylethanolamine, whereas dimethylsulphoxide induces an increase in the temperature and enthalpy and a reduction in the cooperativity of the gel to liquid-crystal transitions for those same phospholipids. 3. Glycerol monooleate induces the formation of isotropic and hexagonal (HII) phases when mixed with either dipalmitoyl phosphatidylcholine or dipalmitoyl phosphatidylethanolamine. By contrast, in the presence of dimethylsulphoxide, those same phospholipids retain the lamellar configuration observed in the absence of fusogen. 4. These results are discussed in terms of the mechanisms of chemically induced cell fusion.  相似文献   

16.
Size enlargement of dipalmitoyl phosphatidylcholine vesicles was greatly accelerated in the range of the phase-transition temperatures, when fatty acid concentration was above a threshold level (‘critical’ concentration). This ‘critical’ concentration varied with the length of the fatty acid chain. The size enlargement process had second-order kinetics dependent on the vesicle concentration. Alkaline pH and low ionic strength inhibited the rate of size enlargement.Phospholipid exchange between dimyristoyl and dipalmitoyl phosphatidylcholine vesicles increased abruptly above a ‘critical’ fatty acid concentration. The donor vesicles were those vesicles in which fatty acids reached the ‘critical’ concentration. The phospholipid exchange occurred both in fluid- and in solid-state vesicles. The ‘critical’ fatty acid concentration accelerating the phospholipid exchange process was lower than that accelerating the size enlargement process.The phospholipid exchange process explained in terms of a diminished hydrophobic attraction among the phospholipid molecules of the bilayer occurs via a free phospholipid molecule transfer through the aqueous phase. The size enlargement process is interpreted in terms of high fatty acid concentration in the membrane fluid domains. The membrane structure is locally perturbed inducing vesicle sticking after collision.  相似文献   

17.
Transfer of phosphatidylcholine molecules between different membrane fractions of Tetrahymena pyriformis cells grown at 15, 27 and 39.5°C was studied by electron spin resonance (ESR). Microsomes were labeled densely with a phosphatidylcholine spin label and the spin-labeled microsomes were incubated with non-labeled cilia, pellicles or microsomes. The transfer of the phosphatidylcholine spin labels was measured by decrease in the exchange broadening of the electron spin resonance spectrum. In one experiment, the lipid transfer was measured between 32P-labeled microsomes and non-labeled pellicles by use of their radioactivity. The result was in good agreement with that by ESR. The fluidity of the membrane was estimated using a fatty-acid spin label incorporated into the membranes. Transfer between lipid vesicles was also studied. The results obtained were as follows: (1) The transfer between sonicated vesicles of egg- or dipalmitoyl phosphatidylcholine occurred rapidly in the liquid crystalline phase, with an activation energy of 20 kcal/mol, whereas it hardly occurred in the solid crystalline phase. (2) The transfer rate between microsomal membranes increased with temperature, and an activation energy of the reaction was 17.8 kcal/mol. (3) The transfer from the spin-labeled microsomes to subcellular membranes of the cells grown at 15°C was larger than that to the membranes of the cells grown at 39.5°C. The membrane fluidity was larger for the cells grown at lower temperature. (4) Similar tendency was observed for the transfer between microsomal lipid vesicles prepared from the cells grown at 15°C and at 39.5°C. (5) The transfer from microsomes to various membrane fractions increased in the order, cilia < pellicles < microsomes. The order of increase in the membrane fluidity was cilia < microsomes < pellicles, although the difference between microsomes and pellicles was slight. These results indicate a crucial role of the membrane fluidity in the transfer reaction. (6) Some evidence supported the idea that the lipid transfer between these organelles occurred through the lipid exchange rather than through the fusion.  相似文献   

18.
The size and size distribution of unilamellar phospholipid vesicles present in unsonicated phosphatidic acid and mixed phosphatidic acid/phosphatidylcholine dispersions were determined by gel filtration, quasi-elastic light scattering and freeze-fracture electron microscopy. The vesiculation in these dispersions was induced by a transient increase in pH as described previously (Hauser, H. and Gains, N. (1982) Proc. Natl. Acad. Sci. USA 79, 1683–1687). The resulting phospholipid dispersions are heterogeneous consisting of small unilamellar vesicles (average radius r < 50 nm) and large unilamellar vesicles (average r ranging from about 50 to 500 nm). The smallest vesicles with r = 11 ± 2 nm are observed with dispersions of pure phosphatidic acid, the population of these vesicles amounting to about 80% of the total lipid. With increasing phosphatidylcholine content the radius of the small unilamellar vesicles increases and at the same time the population of small unilamellar vesicles decreases. The average radius of small unilamellar vesicles present in phosphatidic acid/phosphatidylcholine dispersions (mole ratio, 1:1) is 17.5 ± 2 nm, the population of these vesicles amounting to about 70% of the total lipid. By a combination of gel filtration, quasi-elastic light scattering and freeze-fracture electron microscopy it was possible to characterize the large unilamellar vesicles. This population is heterogeneous with its mean radius also increasing with increasing phosphatidylcholine content. After separating the large unilamellar vesicles from small unilamellar vesicles on Sepharose 4B it can be shown by quasi-elastic light scattering that in pure phosphatidic acid dispersions 80–90% of the large unilamellar vesicle population consist of vesicles with a mean radius of 170 nm. In mixed phosphatidic acid/phosphatidylcholine dispersions this radius increases to about 265 nm as the phosphatidylcholine content is raised to 90 mol%.  相似文献   

19.
Mixtures of lipids and proteins, the ATPase from rabbit sarcoplasmic reticulum, were studied by freeze-fracture electron microscopy and by measurement of the amount of fluid lipid with the spin label 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO). In dimyristoyl phosphatidylcholine vesicles the protein molecules were randomly distributed above the transition temperature, Tt, of the lipid and aggregated below Tt. For mixtures af dimyristoyl and dipalmitoyl phosphatidylcholine the existence of fluid and solid domains was shown in the temperature interval predicted from earlier TEMPO measurements. When protein was incorporated into this lipid mixture, freeze-fracture particles were randomly distributed in fluid lipids, or aggregated when only solid lipids were present.In mixtures of dimyristoyl phosphatidylcholine with cholesterol the protein was distributed randomly above the transition temperature of the phosphatidylcholine. Below that transition temperature the protein was excluded from a banded phase of solid lipid in the case of 10 mol% cholesterol. In mixtures containing 20 mol% cholesterol, protein molecules formed linear arrays, 50–200 nm in length, around smooth patches of lipid.Phase diagrams for lipid/cholesterol and lipid/protein systems are proposed which account for many of the available data. A model for increasing solidification of lipid around protein molecules or cholesterol above the transition temperarture of the lipid is discussed.  相似文献   

20.
The interaction between amphotericin B and egg yolk phosphatidylcholine, dimyristoyl (DMPC) and dipalmitoyl phosphatidylcholine (DPPC) phospholipid bilayer vesicles has been monitored by the circular dichroism (CD) spectra of amphotericin B at a 1 · 10?5 M concentration. This method has revealed that amphotericin B may be present in a number of different forms depending on the time elapsed after the mixing, the cholesterol content of the vesicles and the vesicles' physical state. Some striking features of these CD detected species are the following: with egg yolk phosphatidylcholine and a molar cholesterol percentage lower than 25, at 25°C several forms are coexistent, their amount is time-dependent; with dipalmitoyl or dimyristoyl phosphatidylcholines without cholesterol or with a cholesterol molar percentage lower than 25, in the gel state, a form different from the former appears very rapidly; with egg yolk phosphatidylcholine, DMPC and DPPC at a molar cholesterol percentage between 25 and 50 a new form is monitored, identical in the three cases and observed in the liquid crystalline state as well as in the gel state. In the case of the three phospholipids without cholesterol a definite interaction with the antibiotic is observed but with different characteristics according to the nature of lipid.With amphotericin B ‘Fungizone’ the same species are monitored but their appearance is much slower.Two explanations are proposed for the origin of the discrepancies between CD and electronic absorption.  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