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 lecithin:cholesterol acyltransferase (LCAT)-induced transformation of two discrete species of model complexes that differ in number of apolipoprotein A-I (apoA-I) molecules per particle was investigated. One complex species (designated 3A-I(UC)-complexes) contained 3 apoA-I per particle, was discoidal (13.5 X 4.4 nm), and had a molar composition of 22:78:1 (unesterified cholesterol (UC):egg yolk phosphatidylcholine (egg yolk PC):apoA-I). The other complex species (designated 2A-I(UC)complexes) containing 2 apoA-I per particle was also discoidal (8.4 X 4.1 nm) and had a molar composition of 6:40:1. Transformation of 3A-I(UC)complexes by partially purified LCAT yielded a product (24 hr, 37 degrees C) with a cholesteryl ester (CE) core, 3 apoA-I, and a mean diameter of 9.2 nm. The 2A-I(UC)complexes were only partially transformed to a core-containing product (24 hr, 37 degrees C) which also had 3 apoA-I; this product, however, was smaller (diameter of 8.5 nm) than the product from 3A-I(UC)complexes. Transformation of 3A-I(UC)complexes appeared to result from build-up of core CE directly within the precursor complex. Transformation of 2A-I(UC)complexes, however, followed a stepwise pathway to the product with 3 apoA-I, apparently involving fusion of transforming precursors and release of one apoA-I from the fusion product. In the presence of low density lipoprotein (LDL), used as a source of additional cholesterol, conversion of 2A-I(UC)complexes to the product with 3 apoA-I was more extensive. The transformation product of 3A-I(UC)complexes in the presence of LDL also had 3 apoA-I but was considerably smaller in size (8.6 vs. 9.2 nm, diameter) and had a twofold lower molar content of PC compared with the product formed without LDL. LDL appeared to act both as a donor of UC and an acceptor of PC. Transformation products with 3 apoA-I obtained under the various experimental conditions in the present studies appear to be constrained in core CE content (between 13 to 22 CE per apoA-I; range of 9 CE molecules) but relatively flexible in content of surface PC molecules they can accommodate (between 24 to 49 PC per apoA-I; range of 25 PC molecules). The properties of the core-containing products with 3 apoA-I compare closely with those of the major subpopulation of human plasma HDL in the size range of 8.2-8.8 nm that contains the molecular weight equivalent of 3 apoA-I molecules.  相似文献   

2.
The protein heterogeneity of fractions isolated by immunoaffinity chromatography on anti-apolipoprotein A-I and anti-apolipoprotein A-II affinity columns was analyzed by high resolution two-dimensional gel electrophoresis. The two-dimensional gel electrophoresis profiles of the fractions were analyzed and automatically compared by the computer system MELANIE. Fractions containing apolipoproteins A-I + A-II and only A-I as the major protein components have been isolated from plasma and from high density lipoproteins prepared by ultracentrifugation. Similarities between the profiles of the fractions, as indicated by two-dimensional gel electrophoresis, suggested that those derived from plasma were equivalent to those from high density lipoproteins (HDL), which are particulate in nature. The established apolipoproteins (A-I, A-II, A-IV, C, D, and E) were visible and enriched in fractions from both plasma and HDL. However, plasma-derived fractions showed a much greater degree of protein heterogeneity due largely to enrichment in bands corresponding to six additional proteins. They were present in trace amounts in fractions isolated from HDL and certain of the proteins were visible in two-dimensional gel electrophoresis profiles of the plasma. These proteins are considered to be specifically associated with the immunoaffinity-isolated particles. They have been characterized in terms of Mr and pI. Computer-assisted measurements of protein spot-staining intensities suggest an asymmetric distribution of the proteins (as well as the established apolipoproteins), with four showing greater prominence in particles containing apolipoprotein A-I but no apolipoprotein A-II.  相似文献   

3.
Apolipoprotein (apo) A-I is the major protein in high density lipoproteins (HDL) and is found in two major subclasses of lipoproteins, those containing apolipoprotein A-II (termed LpA-I,A-II) and those without apoA-II (termed LpA-I). The in vivo kinetics of apoA-I on LpA-I and LpA-I,A-II were investigated in normolipidemic human subjects. In the first series of studies, radiolabeled apoA-I and apoA-II were reassociated with autologous plasma lipoproteins and injected into normal subjects. LpA-I and LpA-I,A-II were isolated from plasma at selected time points by immunoaffinity chromatography. By 24 h after injection, only 52.8 +/- 1.0% of the apoA-I in LpA-I remained, whereas 66.9 +/- 2.7% of apoA-I in LpA-I,A-II remained (P less than 0.01). In the second series of studies, purified apoA-I was labeled with either 131I or 125I and reassociated with autologous plasma. Isolated LpA-I and LpA-I,A-II particles differentially labeled with 131I-labeled apoA-I and 125I-labeled apoA-I, respectively, were simultaneously injected into study subjects. The plasma residence time of apoA-I injected on LpA-I (mean 4.39 days) was substantially shorter than that of apoA-I injected on LpA-I,A-II (mean 5.17 days), with a mean difference in residence times of 0.79 +/- 0.08 days (P less than 0.001). These data demonstrate that apoA-I injected on LpA-I is catabolized more rapidly than apoA-I injected on LpA-I,A-II. The results are consistent with the concept that LpA-I and LpA-I,A-II have divergent metabolic pathways.  相似文献   

4.
5.
Zhang X  Chen B 《Biological chemistry》2011,392(5):423-429
It has been shown that apolipoprotein A-V (apoA-V) over-expression significantly lowers plasma triglyceride levels and decreases atherosclerotic lesion development. To assess the feasibility of recombinant high density lipoprotein (rHDL) reconstituted with apoA-V and apolipoprotein A-I (apoA-I) as a therapeutic agent for hyperlipidemic disorder and atherosclerosis, a series of rHDL were synthesized in vitro with various mass ratios of recombinant apoA-I and apoA-V. It is interesting to find that apoA-V of rHDL had no effect on lipoprotein lipase (LPL) activation in vitro and very low density lipoprotein (VLDL) clearance in HepG2 cells and in vivo. By contrast, LPL activation and VLDL clearance were inhibited by the addition of apoA-V to rHDL. Furthermore, the apoA-V of rHDL could not redistribute from rHDL to VLDL after incubation at 37°C for 30 min. These findings suggest that an increase of apoA-V in rHDL could not play a role in VLDL clearance in vitro and in vivo, which could, at least in part, attribute to the lost redistribution of apoA-V from rHDL to VLDL and LPL binding ability of apoA-V in rHDL. The therapeutic application of rHDL reconstituted with apoA-V and apoA-I might need the construction of rHDL from which apoA-V could freely redistribute to VLDL.  相似文献   

6.
We have shown previously that low density lipoprotein (LDL) subjected to vortexing forms self-aggregates that are avidly phagocytosed by macrophages. That phagocytic uptake is mediated by the LDL receptor. We now show that LDL self-aggregation is strongly inhibited (80-95%) by the presence of high density lipoprotein (HDL) or apolipoprotein (apo) A-I. Another type of LDL aggregation, namely that induced by incubation of LDL with phospholipase C, was also markedly inhibited by HDL or apoA-I. The aggregation of LDL induced by vortexing was not inhibited by 2.5 M NaCl, and apoA-I was still able to block LDL aggregation at this high salt concentration, strongly suggesting hydrophobic interactions as the basis for the effect of apoA-I. The fact that apoA-I protected against LDL aggregation induced by two apparently quite different procedures suggests that the aggregation in these two cases has common features. We propose that these forms of LDL aggregation result from the exposure of hydrophobic domains normally masked in LDL and that the LDL-LDL association occurs when these domains interact. ApoA-I, because of its amphipathic character, is able to interact with the exposed hydrophobic domains of LDL and thus block the intermolecular interactions that cause aggregation.  相似文献   

7.
8.
We prepared and isolated defined, reconstituted high density lipoprotein (r-HDL) particles containing apolipoprotein A-I (apoA-I), palmitoyloleoylphosphatidylcholine, and cholesterol. The initial r-HDL were prepared by the sodium cholate method, then part of the preparation was depleted of phospholipid by exposure to LDL, and the resulting, stable r-HDL species were isolated by gel filtration. The isolated r-HDL were characterized in terms of their size, alpha-helix content, and the conformation of apoA-I as reported by the fluorescence properties of the tryptophan residues. Then the relative reactivity of the r-HDL with lecithin cholesterol acyltransferase was assessed. The isolated, discoidal r-HDL contained 2 and 3 apoA-I molecules/particle, and had 77 and 109 A diameters, respectively. Their spectral properties were essentially identical and were distinct from the larger particles in the class of r-HDL with 2 apoA-I molecules/particle (particles with diameters of 86 and 96 A). In addition, the reactivity of the 77 and 109 A particles with pure lecithin cholesterol acyltransferase was similar and about 10-fold lower than for the 86 and 96 A particles. We conclude that the stable, limiting r-HDL particles in each class (77 and 109 A) can arise from the larger particles of the same class by depletion of phospholipids. These limiting particles have very similar apoA-I conformations, with decreased alpha-helix contents and compact protein regions, that are very poor in activating lecithin cholesterol acyltransferase. Based on these results, we propose a model to explain the origin of the different classes and subclasses of the discoidal r-HDL particles.  相似文献   

9.
Human plasma HDLs are classified on the basis of apolipoprotein composition into those that contain apolipoprotein A-I (apoA-I) without apoA-II [(A-I)HDL] and those containing apoA-I and apoA-II [(A-I/A-II)HDL]. ApoA-I enters the plasma as a component of discoidal particles, which are remodeled into spherical (A-I)HDL by LCAT. ApoA-II is secreted into the plasma either in the lipid-free form or as a component of discoidal high density lipoproteins containing apoA-II without apoA-I [(A-II)HDL]. As discoidal (A-II)HDL are poor substrates for LCAT, they are not converted into spherical (A-II)HDL. This study investigates the fate of apoA-II when it enters the plasma. Lipid-free apoA-II and apoA-II-containing discoidal reconstituted HDL [(A-II)rHDL] were injected intravenously into New Zealand White rabbits, a species that is deficient in apoA-II. In both cases, the apoA-II was rapidly and quantitatively incorporated into spherical (A-I)HDL to form spherical (A-I/A-II)HDL. These particles were comparable in size and composition to the (A-I/A-II)HDL in human plasma. Injection of lipid-free apoA-II and discoidal (A-II)rHDL was also accompanied by triglyceride enrichment of the endogenous (A-I)HDL and VLDL as well as the newly formed (A-I/A-II)HDL. We conclude that, irrespective of the form in which apoA-II enters the plasma, it is rapidly incorporated into spherical HDLs that also contain apoA-I to form (A-I/A-II)HDL.  相似文献   

10.
It is well accepted that HDL has the ability to reduce risks for several chronic diseases. To gain insights into the functional properties of HDL, it is critical to understand the HDL structure in detail. To understand interactions between the two major apolipoproteins (apos), apoA-I and apoA-II in HDL, we generated highly defined benchmark discoidal HDL particles. These particles were reconstituted using a physiologically relevant phospholipid, 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) incorporating two molecules of apoA-I and one homodimer of apoA-II per particle. We utilized two independent mass spectrometry techniques to study these particles. The techniques are both sensitive to protein conformation and interactions and are namely: 1) hydrogen deuterium exchange combined with mass spectrometry and 2) partial acetylation of lysine residues combined with MS. Comparison of mixed particles with apoA-I only particles of similar diameter revealed that the changes in apoA-I conformation in the presence of apoA-II are confined to apoA-I helices 3-4 and 7-9. We discuss these findings with respect to the relative reactivity of these two particle types toward a major plasma enzyme, lecithin:cholesterol acyltransferase responsible for the HDL maturation process.  相似文献   

11.
Light-dependent Ca2+ efflux via the Ca2+/H+ antiport in the photosynthetic purple sulfur bacterium Chromatium vinosum was inhibited by three phenothiazines: chlorpromazine; trifluoperazine and phenothiazine. The inhibitors had no effect on Ca2+ uptake by C. vinosum in the dark nor any effect on the light-dependent efflux of either Na+ or Tl+ catalyzed, respectively, by the C. vinosum Na+/H+ or K+/H+ antiports. Ruthenium red and LaCl3, neither of which inhibited light-dependent Ca2+ efflux in C. vinosum, markedly inhibited Ca2+ uptake in the dark by C. vinosum cells. Ruthenium red had no effect on the uptake of either Na+or the K+ analog T1+ by C. vinosum cells in the dark. These results have been interpreted in terms of two separate Ca2+ transport systems in C. vinosum: (i) a phenothiazine-sensitive and ruthenium red, La3+-insensitive Ca2+/H+ antiport responsible for Ca2+ efflux in the light; and (ii) a ruthenium red and La3+-sensitive but phenothiazine-insensitive Ca2+ uptake system.  相似文献   

12.
Our aim was to characterize HDL subspecies and fat-soluble vitamin levels in a kindred with familial apolipoprotein A-I (apoA-I) deficiency. Sequencing of the APOA1 gene revealed a nonsense mutation at codon -2, Q[-2]X, with two documented homozygotes, eight heterozygotes, and two normal subjects in the kindred. Homozygotes presented markedly decreased HDL cholesterol levels, undetectable plasma apoA-1, tuboeruptive and planar xanthomas, mild corneal arcus and opacification, and severe premature coronary artery disease. In both homozygotes, analysis of HDL particles by two-dimensional gel electrophoresis revealed undetectable apoA-I, decreased amounts of small alpha-3 migrating apoA-II particles, and only modestly decreased normal amounts of slow alpha migrating apoA-IV- and apoE-containing HDL, while in the eight heterozygotes, there was loss of large alpha-1 HDL particles. There were no significant decreases in plasma fat-soluble vitamin levels noted in either homozygotes or heterozygotes compared with normal control subjects. Our data indicate that isolated apoA-I deficiency results in marked HDL deficiency with very low apoA-II alpha-3 HDL particles, modest reductions in the separate and distinct plasma apoA-IV and apoE HDL particles, tuboeruptive xanthomas, premature coronary atherosclerosis, and no evidence of fat malabsorption.  相似文献   

13.
To determine the apolipoprotein specificity of high density lipoprotein (HDL) receptor, apolipoprotein A-I (apo-AI) and apolipoprotein A-II (apo-AII) purified from high density lipoprotein3 (HDL3) were reconstituted into dimyristoyl phosphatidylcholine vesicles (DMPC) and their ability to bind to luteinized rat ovarian membranes was examined. Both 125I-apo-A-I.DMPC and 125I-apo-A-II.DMPC were shown to bind to ovarian membranes with Kd = 2.87 and 5.70 micrograms of protein/ml, respectively. The binding of both 125I-apo-A-I.DMPC and 125I-apo-A-II.DMPC was inhibited by unlabeled HDL3, apo-A-I.DMPC, apo-A-II.DMPC, apo-C-I.DMPC, apo-C-II.DMPC, apo-C-III1.DMPC, and apo-C-III2.DMPC, but not by DMPC vesicles, bovine serum albumin.DMPC or low density lipoprotein. Since the binding labeled apo-A-I.DMPC and apo-A-II.DMPC was inhibited by the DMPC complexes of apo-C groups, the direct binding of 125I-apo-C-III1.DMPC was also demonstrated with Kd = 9.6 micrograms of protein/ml. In addition, unlabeled apo-A-I.DMPC, and apo-A-II.DMPC, as well as apo-C.DMPC, inhibited 125I-HDL3 binding. 125I-apo-A-I, 125I-apo-A-II, and 125I-apo-C-III1 in the absence of DMPC also bind to the membranes. These results suggest that HDL receptor recognizes apolipoprotein AI, AII, and the C group and that the binding specificity of the reconstituted lipoproteins is conferred by their apolipoprotein moiety rather than the lipid environment. In vivo pretreatment of rats with human chorionic gonadotropin resulted in an increase of 125I-apo-A-I.DMPC, 125I-apo-A-II.DMPC, and 125I-apo-C-III1.DMPC binding activities. However, no induction of binding activity was observed when the apolipoprotein was not included in DMPC vesicles. An examination of the equilibrium dissociation constant and binding capacity for 125I-apo-A-I.DMPC and 125I-apo-A-II.DMPC after human chorionic gonadotropin treatment revealed that the increase in binding activity was due to an increase in the number of binding sites rather than a change in the binding affinity. These results further support our contention that apo-A-I, apo-A-II, and the apo-C group bind to HDL receptor. In conclusion, the HDL receptor of luteinized rat ovary recognizes apolipoproteins A-I, A-II, and the C group but not low density lipoprotein, and the binding is induced by human chorionic gonadotropin in vivo.  相似文献   

14.
Overexpression of human apolipoprotein A-II (apo A-II) in mice induced postprandial hypertriglyceridemia and marked reduction in plasma HDL concentration and particle size [Boisfer et al. (1999) J. Biol. Chem. 274, 11564-11572]. We presently compared lipoprotein metabolism in three transgenic lines displaying plasma concentrations of human apo A-II ranging from normal to 4 times higher, under ad libitum feeding and after an overnight fast. Fasting dramatically decreased VLDL and lowered circulating human apo A-II in transgenic mice; conversely, plasma HDL levels increased in all genotypes. The apo A-I content of HDL was inversely related to the expression of human apo A-II, probably reflecting displacement of apo A-I by an excess of apo A-II. Thus, the molar ratios of apo A-II/A-I in HDL were significantly higher in fed as compared with fasted animals of the same transgenic line, while endogenous LCAT activity concomitantly decreased. The number and size of HDL particles decreased in direct proportion to the level of human apo A-II expression. Apo A-II was abundantly present in all HDL particles, in contrast to apo A-I mainly present in large ones. Two novel findings were the presence of pre-beta migrating HDL transporting only human apo A-II in the higher-expressing mice and the increase of plasma HDL concentrations by fasting in control and transgenic mice. These findings highlight the reciprocal modifications of VLDL and HDL induced by the feeding-fasting transition and the key role of the molar ratio of apo A-II/A-I as a determinant of HDL particle metabolism and pre-beta HDL formation.  相似文献   

15.
High density lipoproteins (HDL) are heterogeneous particles consisting of about equal amounts of lipid and protein that are thought to mediate the transport of cholesterol from peripheral tissues to liver. We show that a previously identified polymorphism affecting HDL electrophoretic mobility in mice is due to a monogenic variation controlling HDL size and apolipoprotein composition. Thus, the HDL particles of various inbred strains of mice exhibit a striking difference in the ratio fo the two major apolipoproteins of HDL, apoA-I and apoA-II. HDL particles in all strains examined contain an average of about five apoA-I molecules; however, whereas the strains with small HDL contain two to three apoA-II molecules per particle, the strains with large HDL contain about five apoA-II molecules per particle. This increase in the protein content of the large HDL is also accompanied by increased lipid content. The HDL size polymorphism and apoA-II levels cosegregate with the apoA-II structural gene on mouse chromosome 1, indicating that a mutation of the apoA-II gene locus is responsible. The rates of synthesis of apoA-II are increased in the strains with large HDL and high apoA-II levels as compared to the strains with small HDL and low apoA-II levels. On the other hand, the fractional catabolic rates of both apoA-I and apoA-II among the strains are very similar, confirming that apoA-II concentrations are controlled at the level of synthesis. Despite the difference in rates of apoA-II synthesis between strains, the apoA-II mRNA levels in the strains are not discernibly different, suggesting that a mutation of the apoA-II structural gene controls apoA-II translational efficiency. This was confirmed by translating apoA-II mRNA in vitro using a rabbit reticulocyte lysate system. Sequencing of apoA-II cDNA from the strains revealed a number of nucleotide substitutions, which may affect translational efficiency. We conclude that the assembly of apoA-II into HDL does not have a set stoichiometry but, rather, is controlled by the production of apoA-II. As apoA-II levels increase, the HDL particles become larger and acquire more lipid, but apoA-I content per particle remains unchanged. These studies with mice provide a model for the metabolic relationships between apoA-I, apoA-II, and HDL lipid in humans.  相似文献   

16.
C Talussot  G Ponsin 《Biochimie》1991,73(9):1173-1178
Recent reports have shown that apolipoprotein A-I (apo A-I), the major protein of high density lipoprotein (HDL) may exist in different conformational states. We studied the effects of apolipoprotein A-II and/or cholesterol on the conformation of apo A-I in reassembled HDL. Analysis of tryptophan fluorescence quenching in the presence of iodine suggested that cholesterol increased the number of apo A-I tryptophan residues accessible to the aqueous phase, but decreased their mean degree of hydration. These observations cannot be totally explained on the basis of the effect of cholesterol on phospholipid viscosity as determined by fluorescence anisotropy of diphenyl hexatriene. We did not observe any effect of apo A-II on the conformation of apo A-I.  相似文献   

17.
Obese mice without leptin (ob/ob) or the leptin receptor (db/db) have increased plasma HDL levels and accumulate a unique lipoprotein referred to as LDL/HDL1. To determine the role of apolipoprotein A-I (apoA-I) in the formation and accumulation of LDL/HDL1, both ob/ob and db/db mice were crossed onto an apoA-I-deficient (apoA-I(-/-)) background. Even though the obese apoA-I(-/-) mice had an expected dramatic decrease in HDL levels, the LDL/HDL1 particle persisted. The cholesterol in this lipoprotein range was associated with both alpha- and beta-migrating particles, confirming the presence of small LDLs and large HDLs. Moreover, in the obese apoA-I(-/-) mice, LDL particles were smaller and HDLs were more negatively charged and enriched in apoE compared with controls. This LDL/HDL1 particle was rapidly remodeled to the size of normal HDL after injection into C57BL/6 mice, but it was not catabolized in obese apoA-I(-/-) mice even though plasma hepatic lipase (HL) activity was increased significantly. The finding of decreased hepatic scavenger receptor class B type I (SR-BI) protein levels may explain the persistence of LDL/HDL1 in obese apoA-I(-/-) mice. Our studies suggest that the maturation and removal of large HDLs depends on the integrity of a functional axis of apoA-I, HL, and SR-BI. Moreover, the presence of large HDLs without apoA-I provides evidence for an apoA-I-independent pathway of cholesterol efflux, possibly sustained by apoE.  相似文献   

18.
The purpose of this experiment was to characterize the high density lipoproteins (HDL) as a function of hydrated density. HDL was subfractionated on the basis of hydrated density by CsCl density gradient centrifugation of whole serum or the d 1.063-1.25 g/ml HDL fraction isolated from three men and three women. Apolipoprotein A-I and A-II quantitation by radial immunodiffusion showed that the A-I/A-II ratio varied with the lipoprotein hydrated density. The A-I/A-II molar ratio of HDL lipoproteins banding between d 1.106 and 1.150 g/ml was nearly constant at 2.2 +/- 0.2. In the density range 1.151-1.25 g/ml the A-I/A-II ratio increased as the density increased. On the other hand, in the density range between 1.077 and 1.105 the A-I/A-II ratio increased as the density decreased, ranging from 2.8 +/- 0.5 for the d 1.093-1.105 g/ml fraction to 5.6 +/- 1.3 for the d 1.077-1.082 g/ml fraction. The d 1.063-1.076 g/ml fraction and the d 1.077-1.082 g/ml fractions had comparable A-I/A-II ratios. Serum and the d 1.063-1.25 g/ml HDL fraction exhibited similar trends. The cholesterol/(A-I + A-II) ratio decreased as the density increased in all 12 samples (six serum and six HDL) examined. Gradient gel electrophoresis of the density gradient fractions showed that as the density increased from 1.063 to 1.200 g/ml the apparent molecular weight decreased from 3.9 x 10(5) to 1.1 x 10(5). HDL subfractions with the same hydrated densities had comparable molecular weights and A-I/A-II and cholesterol/(A-I + A-II) ratios when isolated from men or women. HDL contains subpopulations that differ in the A-I/A-II molar ratio.-Cheung, M. C., and J. J. Albers. Distribution of cholesterol and apolipoprotein A-I and A-II in human high density lipoprotein subfractions separated by CsCl equilibrium gradient centrifugation: evidence for HDL subpopulations with differing A-I/A-II molar ratios.  相似文献   

19.
Mice have a monodisperse high density lipoprotein (HDL) profile, whereas humans have two major subfractions designated HDL(2) and HDL(3). Human apoA-I transgenic mice exhibit a human-like HDL profile, indicating that the amino acid sequence of apoA-I is a determinant of the HDL profile. Comparison of the primary sequence of mouse and human apoA-I and the previously designated "hinge" domain of apoA-I led us to hypothesize that alpha-helices 7 and 8 (7/8) are determinants of HDL subclass distribution. The following proteins were expressed in Escherichia coli: human apoA-I, T7-hAI; mouse apoA-I, T7-mAI; chimeric human apoA-I containing murine helices 7/8 in place of human helices 7/8, T7-hAI(m7/8); and the reciprocal chimera, T7-mAI(h7/8). The recombinant proteins were examined for their association with human plasma HDL subclasses. The results demonstrated that T7-hAI bound HDL(2) and HDL(3) equally well, whereas T7-mAI bound to HDL(2) preferentially. T7-hAI(m7/8) behaved like T7-mAI, and T7-mAI(h7/8) behaved like T7-hAI. Thus, alpha-helices 7/8 are strong contributors to the pattern of HDL subclass association. Self-association, alpha-helicity, cholesterol efflux, and lecithin-cholesterol acyltransferase activity of the recombinant proteins were also assessed. Human apoA-I self-associates more and activates human lecithin-cholesterol acyltransferase better than mouse apoA-I. These differential characteristics of human and mouse apoA-I are not dependent on helices 7/8.  相似文献   

20.
Apolipoprotein E (apoE) enters the plasma as a component of discoidal HDL and is subsequently incorporated into spherical HDL, most of which contain apoE as the sole apolipoprotein. This study investigates the regulation, origins, and structure of spherical, apoE-containing HDLs and their remodeling by cholesteryl ester transfer protein (CETP). When the ability of discoidal reconstituted high density lipoprotein (rHDL) containing apoE2 [(E2)rHDL], apoE3 [(E3)rHDL], or apoE4 [(E4)rHDL] as the sole apolipoprotein to act as substrates for LCAT were compared with that of discoidal rHDL containing apoA-I [(A-I)rHDL], the rate of cholesterol esterification was (A-I)rHDL > (E2)rHDL approximately (E3)rHDL > (E4)rHDL. LCAT also had a higher affinity for discoidal (A-I)rHDL than for the apoE-containing rHDL. When the discoidal rHDLs were incubated with LCAT and LDL, the resulting spherical (E2)rHDL, (E3)rHDL, and (E4)rHDL were larger than, and structurally distinct from, spherical (A-I)rHDL. Incubation of the apoE-containing spherical rHDL with CETP and Intralipid(R) generated large fusion products without the dissociation of apoE, whereas the spherical (A-I)rHDLs were remodeled into small particles with the formation of lipid-poor apoA-I. In conclusion, i) apoE activates LCAT less efficiently than apoA-I; ii) apoE-containing spherical rHDLs are structurally distinct from spherical (A-I)rHDL; and iii) the CETP-mediated remodeling of apoE-containing spherical rHDL differs from that of spherical (A-I)rHDL.  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