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 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   

2.
The ternary phase diagram of cholesteryl myristate--dimyristoyl lecithin--water has been determined by polarizing light microscopy, scanning calorimetry, and x-ray diffraction. Hydrated dimyristoyl lecithin forms a lamellar liquid--crystalline phase (L alpha) at temperatures greater than 23 degrees C into which limited amounts of cholesteryl myristate (less than 5 wt. %) can be incorporated. The amount of cholesterol ester incorporated is dependent upon the degree of hydration of the L alpha phase. Below 23 degrees C dimyristoyl lecithin forms ordered hydrocarbon chain structures (L beta' and P beta') which do not incorporate cholesterol ester. Comparison with other phospholipid--cholesterol ester--water phase diagrams suggests the following general principles: i) the incorporation of cholesterol ester occurs only into liquid crystalling phospholipid bilayers, ii) the extent of incorporation is temperature-dependent, with increasing amounts of cholesterol ester being incorporated at higher temperatures, and iii) unsaturated cholesterol esters induce increased disordering of the phospholipid bilayers.  相似文献   

3.
Physical properties of cholesteryl esters   总被引:2,自引:0,他引:2  
Cholesteryl esters, the intracellular storage form and intravascular transport form of cholesterol, can exist in crystal, liquid crystal and liquid states. The physical state of cholesteryl esters at physiologic temperatures may be a determinant of their pathogenicity. This review has surveyed saturated aliphatic cholesteryl esters of chain length 1 to 24 carbons and a series of medium-chained unsaturated cholesteryl esters from chain lengths 14 to 24 carbons. A systematic study of transition temperatures by polarizing microscopy and enthalpies by differential scanning calorimetry has provided unifying concepts concerning the phase behavior as a function of chain length and unsaturation. Neat cholesteryl esters show chain-length dependence of transition temperature and enthalpy of both the crystal and liquid crystal transitions. Double bond position along the fatty acyl chain affected stability of the liquid crystal phases; a smectic phase was not observed for any cholesteryl ester with a double bond more proximal than delta 9. 13C NMR spectroscopy in the isotropic liquid phase has provided evidence suggesting a balance of ring-ring vs. chain-chain interactions as a determinant for isotropic liquid----cholesteric vs. isotropic liquid----smectic transitions. Specifically, anisotropic molecular motions of the steroid ring are greater for cholesteryl esters forming a cholesteric phase than a smectic phase from the melt. Chain-chain interactions apparently predominate in smectic phase formation. The X-ray diffraction patterns of cholesteryl esters as a function of chain length reveal several isostructural series and known single crystal data are presented. A chain length depending on the periodicity of the smectic phase is observed which may be different for saturated vs. unsaturated esters. In summary, the phase behavior of cholesteryl ester molecules is complex and cannot be determined a priori from the phase behavior of component cholesterol and fatty acid. The data presented here should provide insight into the biological behavior of this lipid class.  相似文献   

4.
The effect of 4,4'-(isopropylidenedithio)bis(2,6-di-t-butylphenol) (probucol) on cholesteryl ester physical state was examined in dry mixtures, phospholipid-containing dispersions, and cells. Probucol has little effect on the solid to isotropic transition of cholesteryl oleate, but broadens and decreases the enthalpy of the liquid-crystalline transitions at concentrations as low as 1-2 mol %. A probucol transition is only observed at concentrations greater than 20 mol %. The mesomorphic phases of the cholesteryl oleate/probucol mixtures were identified by visual inspection and polarized light microscopy. Mixtures are liquid at probucol concentrations in excess of 5 mol % at 37 degrees C. Probucol also dramatically reduces the enthalpy of the liquid-crystalline transitions of the cholesteryl oleate core of dispersions of the ester with phospholipids at a concentration of 10 mol %, reducing the enthalpy by greater than 80% and the transition temperatures by approximately 2 degrees C. The phase state of cholesteryl esters in Fu5AH rat hepatoma cells was examined after incubation with cholesterol/phospholipid dispersions that caused the accumulation of anisotropic cholesteryl ester droplets. Differential scanning calorimetry scans of cells incubated with cholesterol-rich phospholipid dispersions indicated a phase transition near 48 degrees C, which was abolished when the cells were co-incubated with 50-100 micrograms/ml of probucol in the loading medium. Subsequent to the formation of isotropic cholesteryl ester droplets in the presence of probucol, the rate of efflux of cholesterol from the cells to phosphatidylcholine-containing acceptors in the medium was increased. These data show that probucol is relatively soluble in cholesteryl esters and that probucol changes the phase state of cholesteryl ester droplets in cells to a more fluid phase in which the cholesteryl esters are more readily mobilized.  相似文献   

5.
Human plasma low density lipoprotein displays a reversible thermal transition between 20 and 40 degrees C, due to a phase transition of its core cholesterol ester from a smectic to a more liquid-like state. To determine if the cholesterol of high density lipoprotein (HDL) displays similar thermal behavior, the human lipoprotein and its extracted lipid have been examined by differential scanning calorimetry, low angle X-ray scattering and polarizing microscopy. Neither HDL2**(d 1.063--1.125--1.21 g/ml) nor HDL3(d1.125--1.21g/ml) show thermal transitions between O and 60 degrees C. By contrast cholesterol ester isolated from HDL and mixtures of cholesterol oleate and linoleate show reversible liquid crystalline transitions between 20 and 40 degreesC. X-ray scattering studies of HDL2 and HDL3 performed at 10 degreesC show no scattering fringes attributable to a smectic phase of cholesterol ester. When HDL is heated to temperatures above 60 degreesC a broad, double-peaked endotherm is observed. The first component (peak temperature=71 degreesC) corresponds to a selective release of apoprotein A-1 from the lipoprotein, and the second component (peak temperature=90 degreesC) to a more generalized disruption of lipoprotein structure with release of cholesterol ester and apoprotein A-2. Following the thermal disruption of HDL, reversible liquid crystalline transitions of cholesterol ester can be seen by differential scanning calorimetry and polarizing microscopy, showing the presence of large domains of cholesterol ester. The absence of cholesterol ester transitions in intact HDL may indicate an interaction of cholesterol ester molecules with the protein-phospholipid surface of HDL that prevents the formation of an organized lipid phase. The high temperature behavior of HDL indicates that apoprotein A-1 is less important than apoprotein A-2 in maintaining the HDL apolar lipids in the form of a stable miroemulsion.  相似文献   

6.
D H Croll  D M Small  J A Hamilton 《Biochemistry》1985,24(27):7971-7980
The phase behavior of cholesteryl esters with triglyceride has been characterized by differential scanning calorimetry (DSC), light microscopy, and polarizing light microscopy (PLM). Temperature-dependent molecular motions determined by 13C NMR spectroscopy were correlated with thermotropic phase behavior. Two systems, cholesteryl oleate (CO) and a 3/1 w/w mixture of cholesteryl linoleate (CL) and CO, were examined in the presence of small amounts of triolein (TO). Both systems exhibited metastable cholesteric and smectic (or only smectic) phases. Increasing amounts of TO progressively lowered the liquid-crystalline phase transition temperatures and eventually abolished the cholesteric phase, but at differing amounts of TO for the two systems (between 4% and 5% with CL/CO and between 7% and 10% with CO). DSC and PLM showed a progressive broadening of the phase transitions as well as an overlapping of the temperature ranges of the cholesteric and smectic phases. At greater than or equal to 4% TO, a separate isotropic liquid phase coexisted with liquid-crystalline phases. 13C NMR spectroscopy was used to monitor the molecular motions of the cholesteryl ester steroid ring and acyl chain in liquid and liquid-crystalline phases. In the liquid phase, no significant changes in fatty acyl motions, as reflected in spin-lattice relaxation time (T1) and nuclear Overhauser enhancement (NOE) values, were found on addition of TO. The line width (v 1/2) of the steroid ring resonances increased markedly near (1-5 degrees C above) the isotropic liquid----liquid-crystal phase transition temperature (TLC). However, the C3/C6 v 1/2 ratio at 1 degree C above TLC was greater for mixtures exhibiting an isotropic----cholesteric transition than for mixtures exhibiting an isotropic----smectic transition. Rotational correlation times calculated for motions about the long molecular axis and the nonunique axis showed (i) that the ring motions became more anisotropic as TLC was approached and (ii) that the motions were more anisotropic at TLC + 1 degree C for systems exhibiting a cholesteric phase than for systems exhibiting only a smectic phase. 13C line widths in spectra of the cholesteryl ester liquid-crystalline phases suggested that TO perturbed the cholesteryl ester intermolecular interactions and increased the rates of cholesteryl ester molecular motions relative to neat esters.  相似文献   

7.
A basis for the reorganization of the bilayer structure in biological membranes is the different aggregate structures formed by lipids in water. The phase equilibria of all individual lipids and several in vivo polar lipid mixtures from acyl chain modified membranes of Acholeplasma laidlawii were investigated with different NMR techniques. All dioleoyl (DO) polar lipids, except monoglucosyldiglyceride (MGDG), form lamellar liquid crystalline (L alpha) phases only. The phase diagram of DOMGDG reveals reversed cubic (III), reversed hexagonal (HII), and L alpha phases. In mixtures of DOMGDG and dioleoyldiglycosyldiglyceride (DODGDG), the formation of an III (or HII) phase is enhanced by DOMGDG and low hydration or high temperatures. For in vivo mixtures of all polar DO lipids, a transition from an L alpha to an III phase is promoted by low hydration or high temperatures (50 degrees C). The phospholipids are incorporated in this III phase. Likewise, III and HII phases are formed at similar temperatures in a series of in vivo mixtures with different extents of acyl chain unsaturation. However, their melting temperatures (Tm) vary in an expected manner. All cubic and hexagonal phases, except the III phase with DOMGDG, exist in equilibrium with excess water. The maximum hydration of MGDG and DGDG is similar and increases with acyl chain unsaturation but is substantially lower than that for, e.g., phosphatidylcholine. The translational diffusion of the lipids in the cubic phases is rapid, implying bicontinuous structures. However, their appearances in freeze-fracture electron microscope pictures are different. The III phase of DOMGDG belongs to the Ia3d space group.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
Mechanisms and consequences of cellular cholesterol exchange and transfer   总被引:21,自引:0,他引:21  
It is apparent from consideration of the reactions involved in cellular cholesterol homeostasis that passive transfer of unesterified cholesterol molecules plays a role in cholesterol transport in vivo. Studies in model systems have established that free cholesterol molecules can transfer between membranes by diffusion through the intervening aqueous layer. Desorption of free cholesterol molecules from the donor lipid-water interface is rate-limiting for the overall transfer process and the rate of this step is influenced by interactions of free cholesterol molecules with neighboring phospholipid molecules. The influence of phospholipid unsaturation and sphingomyelin content on the rate of free cholesterol exchange are known in pure phospholipid bilayers and similar effects probably occur in cell membranes. The rate of free cholesterol clearance from cells is determined by the structure of the plasma membrane. It follows that the physical state of free cholesterol in the plasma membrane is important for the kinetics of cholesterol clearance and cell cholesterol homeostasis, as well as the structure of the plasma membrane. Bidirectional flux of free cholesterol between cells and lipoproteins occurs and rate constants characteristic of influx and efflux can be measured. The direction of any net transfer of free cholesterol is determined by the relative free cholesterol/phospholipid molar ratios of the donor and acceptor particles. Cholesterol diffuses down its gradient of chemical potential generally partitioning to the phospholipid-rich particle. Such a surface transfer process can lead to delivery of cholesterol to cells. This mechanism operates independently of any lipoprotein internalization by receptor-mediated endocytosis. The influence of enzymes such as lecithin-cholesterol acyltransferase and hepatic lipase on the direction of net transfer of free cholesterol between lipoproteins and cells can be understood in terms of their effects on the pool sizes and the rate constants for influx and efflux. Excess accumulation of free cholesterol in cells stimulates the rate of cholesteryl ester formation and induces deposition of cholesteryl ester inclusions in the cytoplasm similar to the situation in the 'foam' cells of atherosclerotic plaque. Clearance of cellular cholesteryl ester requires initial hydrolysis to free cholesterol followed by efflux of this free cholesterol. The rate of clearance of cholesteryl ester from cytoplasmic droplets is influenced by the physical state of the cholesteryl ester; liquid-crystalline cholesteryl ester is removed more slowly than cholesteryl ester in a liquid state.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
Phase behavior and structure of aqueous dispersions of sphingomyelin   总被引:6,自引:0,他引:6  
The phase behavior of bovine brain sphingomyelin in water has been determined by polarizing light microscopy, differential scanning calorimetry, and X-ray diffraction. Lamellar phases, in which water is intercalated between sheets of lipid molecules arranged in the classical bilayer fashion, are present over much of the phase diagram. An order-disorder transition separates the high temperature, liquid crystalline, lamellar phase from a more ordered lamellar phase at low temperatures. The hydration characteristics of sphingomyelin are similar to the structurally related lecithin in that only limited amounts of water are incorporated above and below the transition. Above the transition at 47 degrees C, a maximum of 35% by weight of water can be incorporated between the lipid bilayers, the total thickness at maximum hydration being 60.2 A, the lipid thickness 38 A, and the surface area per lipid molecule at the interface 60 A(2). Water in excess of 35% by weight is present as a separate phase. Below the phase transition, at 25 degrees C a maximum of 42% by weight of water may be incorporated between the lipid bilayers. On increasing the hydration, the lamellar repeat distance increases from 63.5 A to a limiting value of 76 A. Within this hydration range the calculated lipid thickness decreases from 63.5 to 42.5 A, and the surface area per lipid molecule increases from 36.1 to 53.6 A(2). Although these changes may be accounted for by a structure in which the hexagonally packed ordered hydrocarbon chains tilt progressively with respect to the normal to the bilayer plane on increasing hydration, it is possible that changes in other more complex lamellar structures may be responsible for these variations in lipid thickness and surface area.  相似文献   

10.
The phase behavior and temperature-dependent molecular motions of three cholesteryl ethers (caproyl, myristyl, oleyl) and a cholesteryl carbonate (oleyl) were characterized. The properties of each ether were qualitatively similar to, but quantitatively different from, those of the corresponding cholesteryl ester. For example, cholesteryl oleyl ether exhibited the same phase transitions as cholesteryl oleate, but at much lower temperatures (e.g., the ether isotropic liquid to cholesteric transition is at 29 degrees C). 13C NMR spectra of ethers in the isotropic liquid and liquid crystalline phases were similar to those of the ester analogue. However, near the liquid to liquid crystalline transition, the steroid ring C3 and C6 linewidths, the C3/C6 linewidth ratio, and the steroid ring rotational correlation times tau rx and tau rz calculated from the linewidths were larger for the ether than the ester analogue. The oleyl carbonate had qualitatively different properties from its analogues (e.g., stable vs. metastable cholesteric and smectic phases). Quantitative results (e.g., relatively long tau rx and tau rz in the isotropic liquid phase) for the carbonate were also distinct from those of both the ester and ether analogues. A comparison of analogues in which the polar linkage is the only structural variable yielded insights into the intermolecular interactions which influence phase behavior.  相似文献   

11.
Little or no information is available on biologically valid labeling of hypercholesterolemic plasma lipoproteins with cholesteryl ester. The esterification of labeled unesterified cholesterol in hypercholesterolemic rabbit plasma by the lecithin: cholesterol acyltransferase reaction is inefficient. The use of the d > 1.063 plasma fraction for this reaction greatly improves the efficiency, but some labeled unesterified cholesterol remains in the end products. The latter disadvantage can be avoided by the addition to whole plasma of labeled cholesteryl ester dissolved in DMSO or acetone. However, in hypercholesterolemic rabbit plasma only a small fraction of the added cholesteryl ester was associated with lipoproteins. When phosphatidylcholine/ cholesteryl ester liposomes were incubated with hypercholesterolemic rabbit plasma for 18–24 h at 37°C the labeled cholesteryl ester was quantitatively incorporated into lipoproteins. Chylomicron-like, cholesteryl ester-rich particles were removed by centrifugation (106g · min) and the subsequently isolated d < 1.019 and d = 1.019–1.063 (LDL) fractions were injected intravenously into normal and hypercholesterolemic rabbits. The disappearance of d < 1.019 and LDL cholesteryl ester and the appearance of cholesteryl ester in other lipoprotein fractions was indistinguishable from that of in vivo-labeled lipoproteins. In vivo and in vitro cholesteryl ester-labeled lipoproteins were also compared by measuring the exchangeability of their cholesteryl ester with HDL cholesteryl ester in vitro. Equal exchangeability of the two labels was observed in the d < 1.019 fraction from which the chylomicron-like particles had been removed. These findings demonstrate that when cholesteryl ester is incorporated by the liposome procedure, the distribution of labeled cholesteryl ester within the lipoprotein complex corresponds closely to that of the in vivo-incorporated labeled cholesteryl ester.  相似文献   

12.
Properties of cholesteryl esters in pure and mixed monolayers   总被引:1,自引:0,他引:1  
The surface properties of cholesteryl palmitate, stearate, linoleate, linolenate, arachidonate, and acetate were investigated. Long-chain esters were not surface-active and force-area (pi-A) isotherms were not obtained. Unsaturated cholesteryl esters were oxidized at the air-water interface and these oxidized lipids gave expanded pi-A isotherms. Cholesteryl acetate had an equilibrium spreading pressure of 14.0 dynes/cm and formed a stable monolayer indistinguishable from cholesterol below that surface pressure. Cholesteryl linoleate formed mixed monolayers with surface-active lipids, and the amount of cholesteryl linoleate in the monolayer depended both on its solubility in the other lipid and on the surface pressure. Even at moderate surface pressures cholesteryl linoleate was extruded from the monolayer into a bulk phase. Cholesteryl acetate exhibited the well-known condensing effect of cholesterol in mixed monolayers with egg lecithin.  相似文献   

13.
The action of a bacterial acyltransferase similar in overall reaction mechanism to the plasma enzyme lecithin:cholesterol acyltransferase (LCAT) has been studied using normal plasma and lipoproteins and plasma from LCAT-deficient patients. The microbial enzyme (GCAT) catalyzed acyl transfer using phosphatidylcholine and cholesterol in all of the lipoprotein fractions, presumably because it has no apolipoprotein cofactor. In addition, the enzyme was capable of hydrolyzing cholesteryl ester in lipoproteins but not in small unilamellar vesicles nor in micellar dispersions containing low amounts of Triton X-100. This suggests that cholesteryl ester is exposed on the surface of lipoprotein particles or that it may be transferred there quickly from the interior. Although considerable interconversion of radiolabeled cholesterol and cholesteryl ester could be demonstrated upon treatment of normal plasma or lipoproteins with the enzyme, there was little change in the actual amount of either steroid. This indicates that the rate of cholesteryl ester formation is very similar to the rate of hydrolysis. The relative proportions of cholesterol and cholesteryl ester in normal plasma are therefore near the equilibrium ratio for the reaction carried out by GCAT, or the ratio is controlled by the properties of the lipoproteins themselves. During reaction with the microbial acyltransferase, the ratio of cholesterol to cholesteryl ester in plasma from LCAT-deficient patients was reduced substantially, suggesting that the enzyme may have some practical applications.  相似文献   

14.
M Z Lai  N Düzgüne?  F C Szoka 《Biochemistry》1985,24(7):1646-1653
The role of the hydroxyl groups of cholesterol and tocopherol in mediating their interaction with phospholipid bilayers has been a subject of considerable interest. We have examined this question by using derivatives of cholesterol and tocopherol in which the hydroxyl group is esterified to succinate. The hemisuccinate esters of cholesterol and alpha-tocopherol can be readily incorporated into phospholipid membranes and in fact can by themselves form closed membrane vesicles as demonstrated by the encapsulation of [3H]sucrose. The thermotropic behavior of mixtures containing each succinate ester and phospholipid was studied by differential scanning calorimetry. The effect of cholesteryl hemisuccinate on the thermotropic properties of dipalmitoylphosphatidylcholine and dimyristoylphosphatidylethanolamine is very similar to that of cholesterol. This indicates that the 3 beta-OH is not required for the formation of a cholesterol-phospholipid complex. In mixtures of tocopherol acid succinate and phospholipids the peak transition temperature is progressively shifted to lower temperatures as the mole fraction of alpha-tocopherol succinate is increased, while the enthalpy of the transition is only slightly affected. At a tocopherol succinate/phospholipid molar ratio of 9/1 a phase transition is still detectable. A comparison between tocopherol succinate and tocopherol indicates that the substitution of the hydroxyl group reduces the interaction of tocopherol with phospholipids to a small but measurable extent. Thus, the hydroxyl group of tocopherol is more important than the hydroxyl group of cholesterol in influencing their interactions with phospholipids.  相似文献   

15.
The hypothesis that clearance of cellular cholesteryl ester deposits may be a function of the physical state of the stored lipid has been investigated. Cultured rat hepatoma cells were induced to store cholesteryl ester in either anisotropic inclusions by exposure to free cholesterol-rich phospholipid dispersions or isotropic inclusions by exposure to identical dispersions supplemented with oleic acid. Differential scanning calorimetry demonstrated an order/disorder transition at 43 degrees C for cholesteryl esters stored in anisotropic inclusions; the enthalpy of this transition was consistent with a smectic liquid crystalline to liquid transition. Lipids in cells with isotropic inclusions displayed no order/disorder transitions over the range 20-80 degrees C, indicating that the lipids are in a liquid state. The presence of oleic acid did not influence the mass of cholesteryl ester stored but increased the amount of stored triglyceride. Fatty acyl compositions of the cholesteryl esters were different under the two loading conditions; in particular, there was 38% cholesteryl oleate in anisotropic inclusions and 65% cholesteryl oleate in isotropic inclusions. Kinetics of cholesteryl ester clearance from cells with either anisotropic or isotropic inclusions were studied during a 12-h exposure to acceptors of free cholesterol. In both cases, cholesteryl ester clearance is essentially linear over 12 h and is directly proportional to the initial content of cholesteryl ester. However, the fraction of initial content of cholesteryl ester cleared in 12 h is 0.17 +/- 0.05 for cells with anisotropic inclusions and 0.34 +/- 0.09 for cells with isotropic inclusions. Our data demonstrate that the more rapid clearance of cholesteryl ester by cells with isotropic inclusions can be correlated with the physical state of the cholesteryl ester.  相似文献   

16.
Proton NMR spectra for gel and liquid crystalline samples, composed of dimyristoyl and/or dipalmitoyl lecithin, cholesterol and water, can be consistently interpreted in terms of mesophase symmetry and molecular diffusion according to a model proposed by Wennerstrom (Wennerstrom, H. (1973) Chem. Phys. Lett. 18, 41-44). It is shown by computer simulation that the characteristic "super-lorentzian" bandshape of the lamellar mesophase can be described by the superposition of three gaussian curves. The NMR signal of the gel phase can be simulated by the superposition of two gaussian curves with widths at half height of 2.5 kHz and 19 kHz. An upper limit of the lateral diffusion coefficient of the lecithin molecules in the gel phase is calculated to be about 5-10(-15) m-2/s. It is therefore concluded that the static intermolecular dipolar couplings average to zero in the lamellar mesophase. An estimation of the order parameter of the liquid crystalline phase is made from experimental data and a calculated "rigid lattice" linewidth. A two phase system is shown to exist in the temperature range 28-34 degrees C for a mesophase of a mixture of dimyristoyl and dipalmitoyl lecithin. The presence of cholesterol results in enhanced lateral diffusion of the lecithin molecules at temperatures below the Chapman transition point.  相似文献   

17.
(1) Parenchymal and non-parenchymal cells were isolated from rat liver. The characteristics of acid lipase activity with 4-methylumbelliferyl oleate as substrate and acid cholesteryl esterase activity with cholesteryl[1-14C]oleate as substrate were investigated. The substrates were incorporated in egg yolk lecithin vesicles and assays for total cell homogenates were developed, which were linear with the amount of protein and time. With 4-methylumbelliferyl oleate as substrate, both parenchymal and non-parechymal cells show maximal activities at acid pH and the maximal activity for non-parenchymal cells is 2.5 times higher than for parenchymal cells. It is concluded that 4-methylumbelliferyl oleate hydrolysis is catalyzed by similar enzyme(s) in both cell types. (2) With cholesteryl[1-14C]oleate as substrate both parenchymal and non-parenchymal cells show maximal activities at acid pH and the maximal activity for non-parenchymal cells is 11.4 times higher than for parenchymal cells. It is further shown that the cholesteryl ester hydrolysis in both cell types show different properties. (3) The high activity and high affinity of acid cholesteryl esterase from non-parenchymal cells for cholesterol oleate hydrolysis as compared to parenchymal cells indicate a relative specialization of non-parenchymal cells in cholesterol ester hydrolysis. It is concluded that non-parenchymal liver cells in cholesterol ester hydrolysis. It is concluded that non-parenchymal liver cells possess the enzymic equipment to hydrolyze very efficiently internalized cholesterol esters, which supports the suggestion that these cell types are an important site for lipoprotein catabolism in liver.  相似文献   

18.
Homogeneous subpopulations of human high-density lipoproteins subfraction-3 (HDL3) have been incubated at 37 degrees C with purified lecithin: cholesterol acyltransferase, human serum albumin and varying concentrations of human low-density lipoproteins (LDL). Changes in HDL particle size and composition during these incubations were monitored. Incubation of HDL3a (particle radius 4.3 nm) in the absence of LDL resulted in an esterification of more than 70% of the HDL free cholesterol after 24 h of incubation. This, however, was sufficient to increase the HDL cholesteryl ester by less than 10% and was not accompanied by any change in particle size. When this mixture was incubated in the presence of progressively increasing concentrations of LDL, which donated free cholesterol to the HDL, the molar rate of production of cholesteryl ester was much greater; at the highest LDL concentration HDL cholesteryl ester content was almost doubled after 24 h and there was an increase in the HDL particle size up to the HDL2 range. In the case of HDL3b (radius 3.9 nm), there were again only minimal changes in particle size in incubations not containing LDL. In the presence of the highest concentration of LDL tested, however, the particles were again enlarged into the HDL2 size range after 24 h incubation. These HDL2-like particles were markedly enriched with cholesteryl ester but depleted of phospholipid and free cholesterol when compared with native HDL2. Furthermore, the ratio of apolipoprotein A-I to apolipoprotein A-II resembled that in the parent-HDL3 and was very much lower than that in native HDL2. It has been concluded that purified lecithin: cholesterol acyltransferase is capable of increasing the size of HDL3 towards that of HDL2 but that other factors must operate in vivo to modulate the chemical composition of the enlarged particles.  相似文献   

19.
Cholesterol ester hydrolase activity of pig aorta has been examined under optimum experimental conditions for hydrolysis of different cholesterol esters. The enzyme specific activity values were in the numerical order of substrates hydrolyzed: cholesteryl linoleate larger than or equal to linolenate greater than palmitate larger than or equal to stearate greater than oleate. The results are discussed in relation to the arterial accumulation of cholesterol esters.  相似文献   

20.
Oriented dipalmitoyllecithin-cholesterol multibilayers with 11% water have been studied with the cholestane spin label. From the ESR spectra the order parameters and the mobility of the spin label about its long axis have been calculated. The results on pure lecithin multibilayers indicate a transition from gel to liquid crystalline phase at 52 plus or minus 2 degrees C. In the gel phase the lecithin alkyl chains are highly ordered, but tilted with respect to the normal to the bilayers by about 25 degrees. Above 52 degrees C the tilt disappears and the mobility of the cholestane spin label increases, indicating an increase of mobility of the lecithin alkyl chains. When cholesterol is added, below about 52 degrees C a decrease of order is found. Furthermore, already small cholesterol contents (smaller than or equal to 10 mole %) remove the tilt. Above about 52 degrees C cholesterol improves the order by decreasing the amplitude of the librational motions. Cholesterol lowers the transition temperature of the system and reduces the mobility of the lecithin alkyl chains in the liquid crystalline phase. However an increase in mobility is found at cholesterol contents up to 10 mole %. A very broad phase transition is observed at 50 mole % cholesterol. In all systems an increase in temperature results in a reduction of order through an increase of the amplitude of the librational motions of the molecules. The librational motions are to some extent cooperative. The asymmetry of the order matrix is found to be a measure for the lateral ordering. Cholesterol increases the lateral ordering, indicating that the flat cholesterol molecules orient parallel to each other.  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