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.
Partially reassembled high density lipoproteins (R-HDL) composed of apolipoprotein A-I and nonhydrolyzable analogues of phosphatidylcholine have been prepared, and their physical properties and reactivities as substrates for lecithin: cholesterol acyltransferase and three phospholipases were tested. The stereo-chemical pairs L-DMPC-ether (1,2-O-ditetradecyl-sn-glycero-3-phosphorylcholine) and D-DMPC-ether (2,3-O-ditetradecyl-sn-glycero-1-phosphoryline) or L-DMPC (1,2-dimyristoyl-sn-glycero-3-phosphoryl-choline) and D-DMPC (2,3-dimyristoyl-sn-glycero-1-phosphorylcholine) have similar thermal properties. R-HDL composed of these four lipids also have similar thermal properties as well as lipid/protein ratios, molecular weights, and protein conformations. Vmax and apparent Km values for lecithin: cholesterol acyltransferase on R-HDL consisting of linear combinations of L-DMPC and D-DMPC, L-DMPC-ether, or D-DMPC-ether plus 6 mol % cholesterol were measured. For the ether lecithins, there was a linear increase in Vmax with percentage of the acyl donor, L-DMPC, in R-HDL; over the same range, there was no change in Km. A comparison with bee venom and Naja melanoleuca phospholipase A2 demonstrated that the venom enzymes have turnover numbers almost 3 orders of magnitude greater than has lecithin:cholesterol acyltransferase; the activity of the phospholipases was profoundly affected by the physical state of the lipid, whereas lecithin: cholesterol acyltransferase activity was not. The differences between these two types of enzymes, which cleave the same bonds of a phosphatidylcholine, are assigned to different catalytic mechanisms. These studies show that R-HDL containing sn-glycero-3-phosphorylcholines and sn-glycero-3-phosphorylcholine ethers have similar structure, properties, and affinities for phospholipolytic enzymes.  相似文献   

2.
Rat plasma lecithin: cholesterol acyltransferase, a 68 kDa glycoprotein, has been purified 14 000-fold by a modification of a procedure used for the human enzyme. The activity of lecithin: cholesteryl acyltransferase in human and rat plasma are the same, although activation of both enzymes by human apolipoprotein A-I is greater than that produced by rat apolipoprotein A-I. Using reassembled high-density lipoproteins composed of human apolipoprotein A-I, phosphatidylcholine ethers and a series of different phosphatidylcholines, the separate effects of molecular species specificity and microenvironment on the rate of cholesteryl ester formation was determined. Substitution of a fluid lipid, 1-palmityl-2-oleyl-sn-glycero-3-phosphorylcholine, for a solid lipid, 1,2-dipalmityl-sn-glycero-3-phosphorylcholine, produced an 8-fold increase in the activity of all molecular species of phosphatidylcholine. With either solid or fluid lipid environments, the activity decreased as a function of increasing chain length of saturated acyl groups. Addition of one or more double bonds greatly increased the activity of a given saturated homologue. One major difference between the molecular specificity of rat and human lecithin: cholesteryl acyltransferase was that the latter had a two-fold preference for phosphatidylcholines containing arachidonate at the sn-2-position.  相似文献   

3.
In a patient with lecithin: cholesterol acyltransferase deficiency, free cholesterol was markedly increased, and esterified cholesterol was diminished. In the patient's plasma, an increase in phosphatidylcholine (PC) and a decrease in sphingomyelin were observed. Concomitantly, an increase in a shorter acyl chain 16:0 was noted in PC, sphingomyelin and phosphatidylethanolamine (PE). In contrast to these results, longer chains such as 22:0 and 24:0 were decreased, especially in sphingomyelin. Unsaturated double bonds such as 18:1 was also increased in PC and PE. In the red-cell membrane lipids, the increase in free cholesterol was counteracted by an increase in PC and by a decrease in sphingomyelin and PE, reflecting changes in the patient's plasma lipids. Increased 16:0 (in PC) and decreased 18:0 and 24:0 were observed. The increased plasma free cholesterol due to metabolic defect (lecithin:cholesterol acyltransferase deficiency) led to decreased red-cell membrane fluidity. This effect appeared to be counteracted by changing phospholipid composition (increased PC and decreased sphingomyelin and PE), by increasing shorter chains (16:0), by decreasing longer chains (18:0 and 24:0) and by increasing unsaturated double bonds (18:2). These results can be interpreted as a self-adaptive modification of lecithin:cholesterol acyltransferase deficiency-induced red-cell membrane abnormalities, to maintain normal membrane fluidity. This speculation was supported by the ESR spin-label studies on the patient's membrane lipids. The normal order parameters in intact red cells and in total lipid liposomes were decreased if cholesterol-depleted membrane liposomes were prepared. Thus, the hardening effect of cholesterol appeared to be counteracted by the softening effects described above. Overall membrane fluidity in intact red cells of the lecithin:cholesterol acyltransferase-deficient patient was maintained normally, judged by order parameters in ESR spin-label studies.  相似文献   

4.
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   

5.
Model high density lipoproteins containing human apolipoprotein A-I, cholesterol, and a variety of phosphatidylcholines (PCs) have been prepared and tested. The PCs included 1-palmitoyl-2-oleoyl PC (POPC) and its diether analog 1-O-hexadecyl-2-oleyl PC (POPC ether), 1,2-diphytanoyl PC (DPhPC), 1-palmitoyl-2-phytanoyl PC, and 1-phytanoyl-2-palmitoyl PC. All ester PCs were good acyl donors for the transesterification of cholesterol catalyzed by human lecithin-cholesterol acyltransferase except DPhPC, which showed no reactivity. The PCs containing one phytanoyl chain donated an acyl chain to cholesterol as fast as non-branched fatty acyl chains. However, the competitive inhibition of lecithin-cholesterol acyltransferase by POPC ether and DPhPC was similar, and both lipids formed a macromolecular matrix that supported the reactivity of other ester PC substrates. The bulk of physicochemical properties of model high density lipoproteins composed of DPhPC were indistinguishable from those of POPC ether. These properties included 1) alpha-helical content of the apoprotein as assessed by circular dichroism, 2) microviscosity as determined from the fluorescence polarization and lifetime of the probe 1,6-diphenyl-1,3,5-hexatriene, 3) macromolecular weight based upon analytical gel filtration chromatography, and 4) surface polarity revealed by the fluorescence of 6-propionyl-2(dimethylamino)naphthalene. The only major difference in a physicochemical property was that the molecular surface area of DPhPC (area = 69 A2 at collapse pressure) determined by monolayer methods was 17 A2 greater than that of POPC (area = 53 A2 at collapse pressure) at all surface pressures measured. We suggest that the properties of DPhPC in being enzymatically nonreactive but a competitive inhibitor are due to its much larger size and that the active site of lecithin-cholesterol acyltransferase cannot bind phospholipid substrates in a catalytically productive way if they have surface areas of 70 A2 or more.  相似文献   

6.
By use of carboxyfluorescein-loaded multilamellar liposomes prepared from synthetic phosphatidylcholine (PC) or sphingomyelin and cholesterol in a molar ratio of 1:1, we studied whether or not fatty acyl domain of the phospholipids affects the membrane-damaging action (or channel formation) of Staphylococcus aureus alpha-toxin on the phospholipid-cholesterol membranes. Our data indicated: (1) that toxin-induced carboxyfluorescein-leakage from the liposomes composed of saturated fatty acyl residue-carrying PC and cholesterol was decreased with increasing chain length of the acyl residues between 12 and 18 carbon atoms, although toxin-binding to the liposomes was not significantly affected by the length of fatty acyl residue; (2) that unsaturated fatty acyl residue in PC or sphingomyelin molecule conferred higher sensitivity to alpha-toxin on the phospholipid-cholesterol liposomes, compared with saturated fatty acyl residues; and (3) that hexamerization of alpha-toxin, estimated by SDS-polyacrylamide gel electrophoresis, occurred more efficiently on the liposomes composed of PC with shorter fatty acyl chain or unsaturated fatty acyl chain. Thus, hydrophobic domain of the phospholipids influences membrane-channel formation of alpha-toxin in the phospholipid-cholesterol membrane, perhaps by modulating packing of phospholipid, cholesterol and the toxin in membrane.  相似文献   

7.
The rates of exchange of [4-14C]cholesterol between lipid vesicles prepared with different phospholipids and with different sizes have been measured. The first-order rate constants were higher using vesicles prepared from phosphatidylcholines with highly branched or polyunsaturated fatty acyl chains than with saturated diacyl or di-O-alkyl chains. The rate measurements indicate that the affinity of cholesterol for phospholipid does not vary significantly on change of the type of linkage (ether or ester) in phosphatidylcholine (PC) or of the positions of the fatty acyl chains in 1,2-diacyl-PC bearing one saturated and one unsaturated chain; furthermore, egg phosphatidylglycerol and egg phosphatidylethanolamine appear to have comparable affinities for cholesterol. However, the molecular packing in the bilayer and nearest-neighbor interactions involving cholesterol appear tightened more by N-palmitoylsphingomyelin than by dipalmitoyl-PC; on incorporation of 44 mol % of these phospholipids (which have the same fatty acyl chain composition) into either small or large unilamellar vesicles prepared with egg phosphatidylglycerol, the exchange rates were strikingly slower when the donor species contained sphingomyelin compared with PC. The rate of cholesterol exchange was 100% faster with small unilamellar vesicles than with large unilamellar vesicles as donors, suggesting that the looser packing in the highly curved small vesicles facilitates cholesterol desorption. The cholesterol exchange rate did not vary with the size of the acceptor vesicles, which indicates that desorption is the rate-limiting step in the exchange process in the presence of excess acceptors.  相似文献   

8.
The reaction catalysed by phosphatidylcholine-sterol acyltransferase (EC 2.3.1.43) is believed to be the major source of cholesteryl ester in human plasma; the enzyme requires a protein activator. Several human apolipoproteins were found to exhibit an activator function, the major one being apolipoprotein A-I. Human apolipoprotein E exists in the population mainly in three different genetic isoforms; apolipoprotein E-2, E-3 and E-4. These isopeptides were isolated from subjects homozygous for one of the isoforms, incorporated into phospholipid/cholesterol/[14C]cholesterol complexes by the cholate dialysis procedure and used to measure capacity to activate phosphatidylcholine-sterol acyltransferase in comparison to apolipoprotein A-I lipid substrate particles prepared by the same procedure. Acyltransferase activity was measured by the formation of [14C]cholesteryl ester from [14C]cholesterol using purified enzyme. With egg yolk phosphatidylcholine as acyl donor, apo E was 15-19% as efficient as apolipoprotein A-I for activation of the acyltransferase. Apo-E-stimulated cholesteryl ester formation by the enzyme was enhanced when 1-oleoyl-2-palmitoyl-glycerophosphocholine was used as a substrate phospholipid (45% of apo A-I/phosphatidylcholine control) and most pronounced with dimyristoylglycerophosphocholine (75% of apo A-I/phosphatidylcholine control). No significant difference in activation was found between apo E isoforms. It is concluded that apolipoprotein E activates phosphatidylcholine-sterol acyltransferase in vitro and that apolipoprotein E isoforms are similarly effective.  相似文献   

9.
Human plasma apoproteins (apo) A-I and A-IV both activate the enzyme lecithin:cholesterol acyltransferase (EC 2.3.1.43). Lecithin:cholesterol acyltransferase activity was measured by the conversion of [4-14C] cholesterol to [4-14C]cholesteryl ester using artificial phospholipid/cholesterol/[4-14C]cholesterol/apoprotein substrates. The substrate was prepared by the addition of apoprotein to a sonicated aqueous dispersion of phospholipid/cholesterol/[4-14C]cholesterol. The activation of lecithin:cholesterol acyltransferase by apo-A-I and -A-IV differed, depending upon the nature of the hydrocarbon chains of the sn-L-alpha-phosphatidylcholine acyl donor. Apo-A-I was a more potent activator than apo-A-IV with egg yolk lecithin, L-alpha-dioleoylphosphatidylcholine, and L-alpha-phosphatidylcholine substituted with one saturated and one unsaturated fatty acid regardless of the substitution position. When L-alpha-phosphatidylcholine esterified with two saturated fatty acids was used as acyl donor, apo-A-IV was more active than apo-A-I in stimulating the lecithin:cholesterol acyltransferase reaction. Complexes of phosphatidylcholines substituted with two saturated fatty acids served as substrate for lecithin:cholesterol acyltransferase even in the absence of any activator protein. Essentially the same results were obtained when substrate complexes (phospholipid-cholesterol-[4-14C]cholesterol-apoprotein) were prepared by a detergent dialysis procedure. Apo-A-IV-L-alpha-dimyristoylphosphatidylcholine complexes thus prepared were shown to be homogeneous particles by column chromatography and density gradient ultracentrifugation. It is concluded that apo-A-IV is able to facilitate the lecithin:cholesterol acyltransferase reaction in vitro.  相似文献   

10.
Lipid composition of Mycoplasma orale was examined and compared with that of horse serum added to the growth medium. Ratios of cholesterol/cholesterol ester and sphingomyelin/phosphatidylcholine were much higher in M. orale than in the horse serum, indicating the organism incorporates selectively cholesterol and sphingomyelin. A distinct difference between the lipids from the two sources was that in phospholipids of M. orale almost all (greater than 95%) of the fatty acyl residues were saturated whereas nearly half of the residues were unsaturated in horse serum phospholipids. Approximately one third of M. orale phospholipids was phosphatidylglycerol, which was synthesized by the organism as was demonstrated by 32P-labeling experiment. Its acyl residues consisted mainly of C16:0 and were efficiently labeled with 14C-palmitate but not with 14C-acetate. These results clearly indicate the de novo synthesis of phosphatidylglycerol by M. orale is through acylation with exogenous saturated fatty acids. On the other hand, all the phosphatidylcholine and sphingomyelin of M. orale were derived from the medium. The 14C-labeling experiment demonstrates that no fatty acid synthesis takes place nor exogenous fatty acid can be incorporated so efficiently as phosphatidylglycerol, suggesting that extremely high proportion of saturated fatty acyl residues in these phospholipids is the consequence of saturation directed to the acyl chains of the incorporated phospholipids.  相似文献   

11.
The ESR spectra from different positional isomers of sphingomyelin and phosphatidylcholine spin-labeled in their acyl chain have been studied in sphingomyelin(cerebroside)-phosphatidylcholine mixed membranes that contain cholesterol. The aim was to investigate mechanisms by which cholesterol could stabilize possible domain formation in sphingolipid-glycerolipid membranes. The outer hyperfine splittings in the ESR spectra of sphingomyelin and phosphatidylcholine spin-labeled on the 5 C atom of the acyl chain were consistent with mixing of the components, but the perturbations on adding cholesterol were greater in the membranes containing sphingomyelin than in those containing phosphatidylcholine. Infrared spectra of the amide I band of egg sphingomyelin were shifted and broadened in the presence of cholesterol to a greater extent than the carbonyl band of phosphatidylcholine, which was affected very little by cholesterol. Two-component ESR spectra were observed from lipids spin-labeled on the 14 C atom of the acyl chain in cholesterol-containing membranes composed of sphingolipids, with or without glycerolipids (sphingomyelin/cerebroside and sphingomyelin/cerebroside/phosphatidylcholine mixtures). These results indicate the existence of gel-phase domains in otherwise liquid-ordered membranes that contain cholesterol. In the gel phase of egg sphingomyelin, the outer hyperfine splittings of sphingomyelin spin-labeled on the 14-C atom of the acyl chain are smaller than those for the corresponding spin-labeled phosphatidylcholine. In the presence of cholesterol, this situation is reversed; the outer splitting of 14-C spin-labeled sphingomyelin is then greater than that of 14-C spin-labeled phosphatidylcholine. This result provides some support for the suggestion that transbilayer interdigitation induced by cholesterol stabilizes the coexistence of gel-phase and "liquid-ordered" domains in membranes containing sphingolipids.  相似文献   

12.
The four peptide analogs of the amphipathic helix whose interactions with dimyristoyl phosphatidylcholine were described in the preceding paper were compared with apolipoproteins (apo) A-I and A-II in ability to displace native apolipoprotein from high density lipoprotein (HDL) and in ability to activate lecithin:cholesterol acyltransferase. The rank order of the ability of the four peptide analogs to displace apo-A-I from intact HDL was 18A-Pro-18A greater than 18A greater than des-Val10-18A greater than reverse-18A, the same order suggested in the preceding paper for relative lipid affinities. Modified HDL from which 40% of the apo-A-I had been displaced by 18A was indistinguishable from unmodified HDL in its ability to act as a lecithin:cholesterol acyltransferase substrate. This suggests that the easily displaced apo-A-I molecules in polydisperse HDL are relatively ineffectual as lecithin:cholesterol acyltransferase activators and/or 18A replaces the lecithin:cholesterol acyltransferase activity lost. The peptide analog 18A-Pro-18A was found to be a powerful activator of lecithin:cholesterol acyltransferase when incubated with unilamellar egg phosphatidylcholine (PC) vesicles, reaching 140% of the activity of apo-A-I at a 1:1.75 peptide-to-egg PC ratio. In another experiment, it was found that discoidal egg PC complexes of 18A-Pro-18A, 18A, and des-Val10-18A, formed by cholate dialysis, had 30-45% of the activity of apo-A-I/egg PC discoidal complexes, also formed by cholate dialysis, at the same peptide/lipid weight ratio. Examination of the structures formed when the 18A-Pro-18A peptide was incubated with unilamellar egg PC vesicles indicated that the ability of 18A-Pro-18A to exceed apo-A-I in lecithin:cholesterol acyltransferase activating ability is due to the spontaneous conversion by 18A-Pro-18A of egg PC vesicles to small protein annulus-bilayer disc structures. Apo-A-I, apo-A-II, nor any of the other three peptide analogs of the amphipathic helix studied were able to convert a significant fraction of egg PC unilamellar vesicles to discoidal structures.  相似文献   

13.
Selectivity factors (Vm/Km) for human and rat lecithin: cholesterol acyltransferases (LCAT) for the transfer of various acyl groups from the 2-position of phosphatidylcholine were determined. By multiplying these values by the proportions of acyl groups at the 2-position of phosphatidylcholine, one can predict the proportions of molecular species of cholesterol ester which will be synthesized by LCAT. In human subjects fasted overnight, the molecular composition of plasma cholesterol ester was found to reflect the LCAT selectivity relatively accurately. This result supports the concepts that hepatic acyl-CoA:cholesterol acyltransferase (ACAT) does not contribute significantly to the synthesis of plasma cholesterol ester and that removal of cholesterol ester from plasma is not selective with respect to molecular species under these conditions. In contrast to the results with humans, the molecular composition of plasma cholesterol ester formed in spontaneously hypertensive rats fed a high-cholesterol diet and then fasted overnight differs from that which is predicted from LCAT selectivity and the proportion of various fatty acids at the 2-position of phosphatidylcholine: these results suggest that cholesterol ester is formed mainly via the ACAT reaction.  相似文献   

14.
13C-NMR and permeability studies are described for sonicated vesicles of phosphatidylcholines bearing two 16-carbon saturated hydrocarbon chains with (a) one ether linkage at carbon 1 (3) or 2 of glycerol and one ester linkage at carbon 2 or 1 (3) of glycerol; (b) two ether linkages and (c) two ester linkages at carbons 1 (3) and 2 of glycerol. The results of 13C-NMR relaxation enhancement measurements using cholesterol enriched with 13C at the 4 position indicate that no significant relocation of the cholesterol molecules takes place in the bilayer when a methylene group is substituted for a carbonyl group in phosphatidylcholine. The 4-13C atom of cholesterol undergoes similar fast anisotropic motions in diester- and diether -phosphatidylcholine bilayers, as judged by spin-lattice relaxation time measurements in the liquid-crystalline phase; although the fast motions are unaltered, linewidth and spin-spin relaxation time measurements suggested some restriction of the slow motions of cholesterol molecules in bilayers from phosphatidylcholines containing an O-alkyl linkage at the sn-2 position instead of an acyl linkage. At temperatures above the gel to liquid-crystal phase transition, the kinetics of ionophore A23187-mediated 45Ca2+ efflux from vesicles prepared from each type of phosphatidylcholine molecule were the same; the kinetics of spontaneous carboxyfluorescein diffusion from diester- and diether -phosphatidylcholine vesicles were the same, whereas mixed ether/ester phosphatidylcholine molecules gave bilayers which are less permeable. The rate constants were reduced on cholesterol incorporation into the bilayers of each type of phosphatidylcholine molecule. The reductions were not statistically significant for 45Ca2+ release. The rate constants for carboxyfluorescein release were also reduced by cholesterol to the same extent in vesicles from diester-, diether -, and 1-ether, and 1-ether-2-ester-phosphatidylcholines; however, a smaller reduction was noted in bilayers from the 1-ester-2-ether analog. The results provide further evidence that there are no highly specific requirements for ester or ether linkages in phosphatidylcholine for cholesterol to reduce bilayer permeability. This is a reflection of the fact that in both diester- and diether -phosphatidylcholine bilayers, the 4-13C atom of cholesterol is located in the region of the acyl carboxyl group or the glyceryl ether oxygen atom.  相似文献   

15.
The phospholipase A1 activity of lipoprotein lipase (LpL) was determined with monomolecular phospholipid films. Rates of phospholipid hydrolysis were dependent on apolipoprotein C-II (the activator protein for LpL) phospholipid fatty acyl composition, and lipid-packing density. In sphingomyelin: cholesterol (2:1, molar) monolayers containing 5 mol % disaturated phosphatidylcholines (PC) and at a surface pressure of 22 mNm-1, rates of LpL hydrolysis of diC14:0PC, diC16:0PC, and diC18:0PC were 74, 207, and 65 nmol h-1 mg LpL-1, respectively. At 22 mNm-1, phospholipids containing unsaturated fatty acyl chains were hydrolyzed at rates 5-10 times greater than saturated lipids. At higher lipid packing densities, the difference in hydrolysis rates between saturated and unsaturated lipids was less apparent. Comparison of molecular areas indicate no simple dependency between the rate of LpL catalysis and phospholipid fatty acyl chain length and saturation/unsaturation.  相似文献   

16.
Lipid microemulsions with various core and surface lipid compositions were prepared by co-sonication of cholesteryl esters, triolein (TO), egg phosphatidylcholine (egg PC), and cholesterol. The heterogeneous emulsion particle mixture was purified by gel filtration and particles with the size and general organization of low density lipoproteins were obtained. These lipid microemulsion particles were used for studies of the cellular metabolism of lipoprotein-derived cholesterol and cholesteryl esters as catalyzed by the enzyme acid sterol ester hydrolase (EC 3.1.1.13). The hydrolysis of cholesteryl oleate (CO) was more than twice and that of cholesteryl linoleate (CL) more than three times faster than the hydrolysis of cholesteryl stearate (CS) over the temperature range 25-39.6 degrees C. Both the synthesis and hydrolysis of cholesteryl esters were insensitive to the physical state of the microemulsion cores. The synthesis of cholesteryl esters by this enzyme was also insensitive to the ratios of cholesterol and egg PC in the microemulsion surface layers. Incorporation of triolein into the microemulsion cholesteryl ester core slightly increased the rate of cholesteryl ester synthesis. A decreasing fatty acyl chain length (C18:0 to C14:0) and an increasing degree of unsaturation (C18:0 to C18:2) enhanced the synthesis rate. It is suggested that the hydrolysis and synthesis of cholesteryl esters in microemulsions (and lipoproteins) take place only in the particle surface layer and that the rate of catalysis is directly dependent on the amount of substrate in this surface layer.  相似文献   

17.
Phospholipids isolated from the plasma of monkeys fed a diet enriched in fish oil were poor substrates for cholesteryl ester (CE) synthesis by the lecithin:cholesterol acyltransferase (LCAT) reaction relative to those from animals fed a lard containing diet when the phospholipids were used for the preparation of recombinant particles by cholate dialysis (Parks, J. S., B. C. Bullock, and L. L. Rudel. 1989. J. Biol. Chem. 264: 2545-2551). The purpose of the present study was to directly test the influence of eicosapentaenoic acid (20:5 n-3) and docosahexaenoic acid (22:6 n-3) in the sn-2 position of phosphatidylcholine (PC) on the activity of LCAT. PC species containing 1-palmitoyl-2-oleoyl PC (POPC), 1-palmitoyl-2-linoleoyl PC (PLPC), 1-palmitoyl-2-arachidonoyl PC (PAPC), 1-palmitoyl-2-eicosapentaenoyl PC (PEPC), or 1-palmitoyl-2-docosahexaenoyl PC (PDPC) were purchased or synthesized and made into recombinant particles of uniform size and composition with [14C]cholesterol and apoA-I using the cholate dialysis procedure. The recombinant particles (PC:cholesterol:apoA-I molar ratio = 42:1.9:1) exhibited the following order of reactivity towards purified human LCAT in vitro: POPC greater than PLPC greater than PEPC = PAPC greater than PDPC. The apparent Vmax/Km for recombinant particles containing PEPC and PDPC was 17% and 7% that of particles containing POPC, respectively. There was a linear decrease in CE formation when the percentage of PEPC or PDPC was increased from 0 to 100% relative to POPC in recombinant particles with a constant PC:cholesterol:apoA-I molar ratio, suggesting that the PEPC and PDPC were competitive inhibitors of the LCAT reaction.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

18.
Here we have studied how the length of the pyrene-labeled acyl chain (n) of a phosphatidylcholine, sphingomyelin, or galactosylceramide affects the partitioning of these lipids between 1), gel and fluid domains coexisting in bovine brain sphingomyelin (BB-SM) or BB-SM/spin-labeled phosphatidylcholine (PC) bilayers or 2), between liquid-disordered and liquid-ordered domains in BB-SM/spin-labeled PC/cholesterol bilayers. The partitioning behavior was deduced either from modeling of pyrene excimer/monomer ratio versus temperature plots, or from quenching of the pyrene monomer fluorescence by spin-labeled PC. New methods were developed to model excimer formation and pyrene lipid quenching in segregated bilayers. The main result is that partition to either gel or liquid-ordered domains increased significantly with increasing length of the labeled acyl chain, probably because the pyrene moiety attached to a long chain perturbs these ordered domains less. Differences in partitioning were also observed between phosphatidylcholine, sphingomyelin, and galactosylceramide, thus indicating that the lipid backbone and headgroup-specific properties are not severely masked by the pyrene moiety. We conclude that pyrene-labeled lipids could be valuable tools when monitoring domain formation in model and biological membranes as well as when assessing the role of membrane domains in lipid trafficking and sorting.  相似文献   

19.
Nonhydrolyzable matrices of ether-linked phosphatidylcholines (PCs) and sphingomyelin have been used to study the mechanism of action of lipolytic enzymes. Since ether PCs, sphingomyelin, and ester PCs vary in the number of hydrogen bond donors and acceptors in the carbonyl region of the bilayer, we have examined several physical properties of ether PCs and sphingomyelin in model systems to validate their suitability as nonhydrolyzable lipid matrices. The intermolecular interactions of ether PCs with ester PCs, sphingomyelin, and cholesterol were investigated by differential scanning calorimetry. Phase diagrams constructed from the temperature dependence of the gel to liquid-crystalline phase transition of 1,2-O-dihexadecyl-sn-glycero-3-phosphocholine (DPPC-ether) and 1,2-O-ditetradecyl-sn-glycero-3-phosphocholine (DMPC-ether) with both 1,2-dimyristoyl-sn-glycero-3-phosphocholine (DMPC) and 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) demonstrated complete lipid miscibility in the gel and liquid-crystalline phases. Additionally, phase diagrams of egg yolk sphingomyelin (EYSM) with DMPC or DMPC-ether and 1,2-distearoyl-sn-glycero-3-phosphocholine (DSPC) or 1,2-O-dioctadecyl-sn-glycero-3-phosphocholine (DSPC-ether) demonstrated no major differences in miscibility of EYSM in ester and ether PCs. The effect of 10 mol % cholesterol on the thermal transitions of mixtures of ester and ether PCs also indicates little preference of cholesterol for either lipid. The fusion of small single bilayer vesicles of DMPC, DMPC-ether, DPPC, and DPPC-ether to larger aggregates as determined by gel filtration indicated that the ester PC vesicles were somewhat more stable.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
To examine the potential of high density lipoproteins (HDL) to ameliorate atherosclerotic plaques in vivo, we examined the ability of native HDL, lipid-free HDL apolipoproteins (apo HDL), cholesterol-free discoidal reconstituted HDL (R-HDL) comprised of apo HDL and phosphatidylcholine (PC) and PC liposomes to release cholesterol from cholesterol-rich insoluble components of plaques (ICP) isolated from atherosclerotic human aorta. Isolated ICP had a free cholesterol (FC) to phospholipid (PL) mass ratio (0.8-3.1) and a sphingomyelin (SPM) to PC mass ratio (1.2-4.2) that exceeded those of plasma membranes of cultured cells. Surprisingly, native HDL and its apolipoproteins were not able to release cholesterol from ICP. However, R-HDL and PC liposomes were effectively released cholesterol from ICP. The release of ICP cholesterol by R-HDL was dose-dependent and accompanied by the transfer of > 8 x more PC in the reverse direction (i.e., from R-HDL to ICP), resulting in a marked enrichment of ICP with PC. Compared to R-HDL, PC liposomes were significantly less effective in releasing cholesterol from ICP but were somewhat more effective in enriching ICP with PC. Native HDL was minimally effective in enriching ICP with PC, but became effective after prior in vitro enrichment of HDL with PC from multilamellar PC liposomes. The enrichment of ICP with PC resulted in the dissolution of cholesterol crystals on ICP and allowed the removal of ICP cholesterol by apo HDL and plasma. Our study revealed that the removal of cholesterol from ICP in vivo will be possible through a change in the level, composition, and physical state of ICP lipids mediated by PC-enriched HDL.  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