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 109 毫秒
1.
We sought to investigate effects of lipoprotein lipase (LpL) on cellular catabolism of lipoproteins rich in apolipoprotein B-100. LpL increased cellular degradation of lipoprotein(a) (Lp(a)) and low density lipoprotein (LDL) by 277% +/- 3.8% and 32.5% +/- 4.1%, respectively, and cell association by 509% +/- 8.7% and 83.9% +/- 4.0%. The enhanced degradation was entirely lysosomal. Enhanced degradation of Lp(a) had at least two components, one LDL receptor-dependent and unaffected by heparitinase digestion of the cells, and the other LDL receptor-independent and heparitinase-sensitive. The effect of LpL on LDL degradation was entirely LDL receptor-independent, heparitinase-sensitive, and essentially absent from mutant Chinese hamster ovary cells that lack cell surface heparan sulfate proteoglycans. Enhanced cell association of Lp(a) and LDL was largely LDL receptor-independent and heparitinase-sensitive. The ability of LpL to reduce net secretion of apolipoprotein B-100 by HepG2 cells by enhancing cellular reuptake of nascent lipoproteins was also LDL receptor-independent and heparitinase-sensitive. None of these effects on Lp(a), LDL, or nascent lipoproteins required LpL enzymatic activity. We conclude that LpL promotes binding of apolipoprotein B-100-rich lipoproteins to cell surface heparan sulfate proteoglycans. LpL also enhanced the otherwise weak binding of Lp(a) to LDL receptors. The heparan sulfate proteoglycan pathway represents a novel catabolic mechanism that may allow substantial cellular and interstitial accumulation of cholesteryl ester-rich lipoproteins, independent of feedback inhibition by cellular sterol content.  相似文献   

2.
Apolipoprotein A-I regulates lipid hydrolysis by hepatic lipase   总被引:2,自引:0,他引:2  
Association of hepatic lipase (HL) with pure heparan sulfate proteoglycans (HSPG) has little effect on hydrolysis of high density lipoprotein (HDL) particles, but significantly inhibits (>80%) the hydrolysis of low (LDL) and very low density lipoproteins (VLDL). Lipolytic inhibition is associated with a differential ability of the lipoproteins to remove HL from the HSPG. LDL and VLDL are unable to displace HL, whereas HDL readily displaces HL from the HSPG. These data show that HSPG-bound HL is inactive. Purified apolipoprotein (apo) A-I is more efficient than HDL at liberating HL from HSPG, and HL displacement is associated with the direct binding of apoA-I to HSPG. However, displacement of HL by apoA-I does not enhance hydrolysis of VLDL particles. This appears due to the direct inhibition of HL by apoA-I. Both apoA-I and HDL are able to inhibit VLDL lipid hydrolysis by up to 60%. Inhibition of VLDL hydrolysis is associated with the binding of apoA-I to the surface of the VLDL particle and a concomitant decreased affinity for HL. These data show that apoA-I can regulate lipid hydrolysis by HL by liberating/activating the enzyme from cell surface proteoglycans and by directly modulating lipoprotein binding and hydrolysis.  相似文献   

3.
Lipoprotein lipase (LpL) hydrolyzes chylomicron and very low density lipoprotein triglycerides to provide fatty acids to tissues. Aside from its lipolytic activity, LpL promotes lipoprotein uptake by increasing the association of these particles with cell surfaces allowing for the internalization by receptors and proteoglycans. Recent studies also indicate that LpL stimulates selective uptake of lipids from high density lipoprotein (HDL) and very low density lipoprotein. To study whether LpL can mediate selective uptake of lipids from low density lipoprotein (LDL), LpL was incubated with LDL receptor negative fibroblasts, and the uptake of LDL protein, labeled with (125)I, and cholesteryl esters traced with [(3)H]cholesteryl oleoyl ether, was compared. LpL mediated greater uptake of [(3)H]cholesteryl oleoyl ether than (125)I-LDL protein, a result that indicated selective lipid uptake. Lipid enrichment of cells was confirmed by measuring cellular cholesterol mass. LpL-mediated LDL selective uptake was not affected by the LpL inhibitor tetrahydrolipstatin but was nearly abolished by heparin, monoclonal anti-LpL antibodies, or chlorate treatment of cells and was not found using proteoglycan-deficient Chinese hamster ovary cells. Selective uptake from HDL, but not LDL, was 2-3-fold greater in scavenger receptor class B type I overexpressing cells (SR-BI cells) than compared control cells. LpL, however, induced similar increases in selective uptake from LDL and HDL in either control or SR-BI cells, indicative of the SR-BI-independent pathway. This was further supported by ability of LpL to promote selective uptake from LDL in human embryonal kidney 293 cells, cells that do not express SR-BI. In Chinese hamster ovary cell lines that overexpress LpL, we also found that selective uptake from LDL was induced by both endogenous and exogenous LpL. Transgenic mice that overexpress human LpL via a muscle creatine kinase promoter had more LDL selective uptake in muscle than did wild type mice. In summary LpL stimulates selective uptake of cholesteryl esters from LDL via pathways that are distinct from SR-BI. Moreover this process also occurs in vivo in tissues where abundant LpL is present.  相似文献   

4.
According to numerous studies low-density lipoproteins (LDL) are supposed to interact with the glycosaminoglycan chain(s) of proteoglycans, e.g. with decorin and biglycan, which themselves are subject to receptor-mediated endocytosis. We tested, therefore, whether complexes of LDL and small proteoglycans can be endocytosed by either the LDL- or the small proteoglycan uptake mechanism. However, neither was the endocytosis of LDL significantly influenced by proteoglycans nor that of proteoglycans by LDL. This negative result could be explained by the observation that in vitro complex formation takes place only in buffers of low ionic strength. Under physiological conditions additional molecules may be necessary for complex stabilization. Lipoprotein lipase (LpL) which binds LDL was also able to interact with high affinity with decorin and its glycosaminoglycan-free core protein, both interactions being heparin-sensitive. Regardless of the presence or absence of LDL, LpL stimulated the endocytosis of decorin 1.5-fold, whereas LpL mediated a 4-fold stimulation of LDL uptake in the absence of decorin. No significant additional effect was seen in the presence of small concentrations of proteoglycans whereas in the presence of 1 microM decorin the endocytosis of [125I]LDL was reduced in normal as well as in LDL receptor-deficient fibroblasts. These observations could best be explained by assuming that LpL/LDL complexes are internalized upon binding to membrane-associated heparan sulphate and that small proteoglycans interfere with this process. It could not be ruled out, however, that a small proportion of the complexes is also taken up by the small proteoglycan receptor.  相似文献   

5.
Characterization of the lipolytic activity of endothelial lipase   总被引:16,自引:0,他引:16  
Endothelial lipase (EL) is a new member of the triglyceride lipase gene family previously reported to have phospholipase activity. Using radiolabeled lipid substrates, we characterized the lipolytic activity of this enzyme in comparison to lipoprotein lipase (LPL) and hepatic lipase (HL) using conditioned medium from cells infected with recombinant adenoviruses encoding each of the enzymes. In the absence of serum, EL had clearly detectable triglyceride lipase activity. Both the triglyceride lipase and phospholipase activities of EL were inhibited in a dose-dependent fashion by the addition of serum. The ratio of triglyceride lipase to phospholipase activity of EL was 0.65, compared with ratios of 24.1 for HL and 139.9 for LPL, placing EL at the opposite end of the lipolytic spectrum from LPL. Neither lipase activity of EL was influenced by the addition of apolipoprotein C-II (apoC-II), indicating that EL, like HL, does not require apoC-II for activation. Like LPL but not HL, both lipase activities of EL were inhibited by 1 M NaCl. The relative ability of EL, versus HL and LPL, to hydrolyze lipids in isolated lipoprotein fractions was also examined using generation of FFAs as an end point. As expected, based on the relative triglyceride lipase activities of the three enzymes, the triglyceride-rich lipoproteins, chylomicrons, VLDL, and IDL, were efficiently hydrolyzed by LPL and HL. EL hydrolyzed HDL more efficiently than the other lipoprotein fractions, and LDL was a poor substrate for all of the enzymes.  相似文献   

6.
Proteoheparan sulfate can be adsorbed to a methylated silica surface in a monomolecular layer via its transmembrane hydrophobic protein core domain. Due to electrostatic repulsion, its anionic glycosaminoglycan side chains are stretched out into the blood substitute solution, representing a receptor site for specific lipoprotein binding through basic amino acid-rich residues within their apolipoproteins. The binding process was studied by ellipsometric techniques showing that HDL has a high binding affinity to the receptor and a protective effect on interfacial heparan sulfate proteoglycan layers, with respect to LDL and Ca2+ complexation. LDL was found to deposit strongly at the proteoheparan sulfate, particularly in the presence of Ca2+, thus creating the complex formation ‘proteoglycan–low density lipoprotein–calcium’. This ternary complex build-up may be interpreted as arteriosclerotic nanoplaque formation on the molecular level responsible for the arteriosclerotic primary lesion. On the other hand, HDL bound to heparan sulfate proteoglycan protected against LDL docking and completely suppressed calcification of the proteoglycan–lipoprotein complex. In addition, HDL and aqueous garlic extract were able to reduce the ternary complex deposition and to disintegrate HS-PG/LDL/Ca2+ aggregates. Although much remains unclear regarding the mechanism of lipoprotein depositions at proteoglycan-coated surfaces, it seems clear that the use of such systems offers possibilities for investigating lipoprotein deposition at a ‘nanoscopic’ level under close to physiological conditions. In particular, Ca2+-promoted LDL deposition and the protective effect of HDL, even at high Ca2+ and LDL concentrations, agree well with previous clinical observations regarding risk and beneficial factors for early stages of atherosclerosis. Therefore, we believe that the system can be of some use in investigations, e.g. of the interplay between different lipoproteins in arteriosclerotic plaque formation, as well as in high throughput screening of candidate drugs to atherosclerosis in a biosensor application.  相似文献   

7.
Lipoprotein lipase (LPL) and hepatic lipase (HL) are enzymatic activities involved in lipoprotein metabolism. The purpose of this study was to analyze the physicochemical modifications of plasma lipoproteins produced by LPL activation in two patients with apoC-II deficiency syndrome and by HL activation in two patients with LPL deficiency. LPL activation was achieved by the infusion of normal plasma containing apoC-II and HL was released by the injection of heparin. Lipoproteins were analyzed by ultracentrifugation in a zonal rotor under rate flotation conditions before and after lipase activation. The LPL activation resulted in: a reduction of plasma triglycerides; a reduction of fast-floating very low density lipoprotein (VLDL) concentration; an increase of intermediate density lipoprotein (IDL), which maintained unaltered flotation properties; an increase of low density lipoproteins (LDL) accompanied by modifications of their flotation rates and composition; no significant variations of high density lipoprotein (HDL) levels; and an increase of the HDL flotation rate. The HL activation resulted in: a slight reduction of plasma triglycerides; a reduction of the relative triglyceride content of slow-floating VLDL, IDL, LDL2, and HDL3 accompanied by an increase of phospholipid in VLDL and by an increase of cholesteryl ester in IDL; and a reduction of the HDL flotation rate. These experiments in chylomicronemic patients provide in vivo evidence that LPL and HL are responsible for plasma triglyceride hydrolysis of different lipoproteins, and that LPL is particularly involved in determining the levels and physicochemical properties of LDL. Moreover, in these patients, the LPL activation does not directly change the HDL levels, and LPL or HL does not produce a step-wise conversion of HDL3 to HDL2 (or vice versa) but rather modifies the flotation rates of all the HDL molecules present in plasma.  相似文献   

8.
The effect of lipoprotein electrostatic properties on the catalytic regulation of hepatic lipase (HL) was investigated. Enrichment of serum or very low density lipoprotein (VLDL) with oleic acid increased lipoprotein negative charge and stimulated lipid hydrolysis by HL. Similarly, enrichment of serum or isolated lipoproteins with the anionic phospholipids phosphatidylinositol (PI), phosphatidic acid, or phosphatidylserine also increased lipoprotein negative charge and stimulated hydrolysis by HL. Anionic lipids had a small effect on phospholipid hydrolysis, but significantly stimulated triacylglyceride (TG) hydrolysis. High density lipoprotein (HDL) charge appears to have a specific effect on lipolysis. Enrichment of HDL with PI significantly stimulated VLDL-TG hydrolysis by HL. To determine whether HDL charge affects the association of HL with HDL and VLDL, HL-lipoprotein interactions were probed immunochemically. Under normal circumstances, HL associates with HDL particles, and only small amounts bind to VLDL. PI enrichment of HDL blocked the binding of HL with HDL. These data indicate that increasing the negative charge of HDL stimulates VLDL-TG hydrolysis by reducing the association of HL with HDL. Therefore, HDL controls the hydrolysis of VLDL by affecting the interlipoprotein association of HL. Lipoprotein electrostatic properties regulate lipase association and are an important regulator of the binding and activity of lipolytic enzymes.  相似文献   

9.
Lipoprotein lipase (LPL) is produced by cells in the artery wall and can mediate binding of lipoproteins to cell surface heparan sulfate proteoglycans (HSPG), resulting in endocytosis (the bridging function). Active, dimeric LPL may dissociate to inactive monomers, the main form found in plasma. We have studied binding/internalization of human low density lipoprotein (LDL), mediated by bovine LPL, using THP-1 monocytes and macrophages. Uptake of (125)I-LDL was similar in monocytes and macrophages and was not affected by the LDL-receptor family antagonist receptor-associated protein (RAP) or by the phagocytosis inhibitor cytochalasin D. In contrast, uptake depended on HSPG and on membrane cholesterol. Incubation in the presence of dexamethasone increased the endogenous production of LPL by the cells and also increased LPL-mediated binding of LDL to the cell surfaces. Monomeric LPL was bound to the cells mostly in a heparin-resistant fashion. We conclude that the uptake of LDL mediated by LPL dimers is receptor-independent and involves cholesterol-enriched membrane areas (lipid rafts). Dimeric and monomeric LPL differ in their ability to mediate binding/uptake of LDL, probably due to different mechanisms for binding/internalization.  相似文献   

10.
Equilibrium-binding data of highly purified 125I-labeled avian lipoprotein lipase to cultured avian adipocytes demonstrate the presence of a class of high affinity binding sites. Analysis of the binding function yielded an association constant of 0.62 x 10(8)M-1 and a maximum binding capacity of 2.1 micrograms/60-mm dish. From a time course of dissociation of 125I-lipoprotein lipase from adipocytes at 4 degrees C, a dissociation rate constant of 6.1 x 10(-5)s-1 was obtained. Pretreatment of cells with heparinase and heparitinase resulted in a quantitative suppression of the high affinity binding component, establishing that lipoprotein lipase is bound to cell surface heparan sulfate proteoglycans. At 37 degrees C, cell surface-bound 125I-lipoprotein lipase is internalized and either degraded or recycled to the medium. The degradation rate constant for 125I-lipoprotein lipase was estimated to be 0.78 h-1. The degradation rate constant was reduced 6-fold when cells were exposed to 100 microM chloroquine, indicating that most of the degradation occurs within the lysosomal compartment. By using cells that had been pulsed with Trans35S-label for 1 h, it was demonstrated that acute treatment with endoglycosidases for up to 1 h resulted in a new lipoprotein lipase secretion rate which was 6-fold higher than that of control cells. Degradation of newly synthesized lipoprotein lipase was essentially blocked 30 min after the initiation of the chase. In other studies it was observed that there were no additive effects of chloroquine and either endoglycosidase or heparin treatment on total lipoprotein lipase levels (intracellular, cell surface, and medium) in adipocyte cultures. These experiments support the hypothesis that the release of lipoprotein lipase from its receptor prevents its internalization and degradation and enhances enzyme efflux from the adipocyte. A new model of lipoprotein lipase secretion in cultured adipocytes is proposed: Newly synthesized lipoprotein lipase is transported to the cell surface where it binds to specific heparan sulfate proteoglycan receptors. The enzyme is either released to the medium or internalized via the receptor, in which case the enzyme is degraded or recycled to the cell surface. Major determinants of enzyme efflux from the cell surface include the number and integrity of receptors, the association constant of the enzyme-receptor complex, and the presence in the medium of competing molecules with high affinity for lipoprotein lipase. In this model, modulation of lipoprotein lipase degradation rate may be a significant mechanism for acute regulation of enzyme efflux independent of changes in the rate of enzyme synthesis.  相似文献   

11.
We previously identified that four of five putative N-linked glycosylation sites of human endothelial lipase (EL) are utilized and suggested that the substitution of asparagine-116 (Asn-116) with alanine (Ala) (N116A) increased the hydrolytic activity of EL. The current study demonstrates that mutagenesis of either Asn-116 to threonine (Thr) or Thr-118 to Ala also disrupted the glycosylation of EL and enhanced catalytic activity toward synthetic substrates by 3-fold versus wild-type EL. Furthermore, we assessed the hydrolysis of native lipoprotein lipids by EL-N116A. EL-N116A exhibited a 5-fold increase in LDL hydrolysis and a 1.8-fold increase in HDL2 hydrolysis. Consistent with these observations, adenovirus-mediated expression of EL-N116A in mice significantly reduced the levels of both LDL and HDL cholesterol beyond the reductions observed by the expression of wild-type EL alone. Finally, we introduced Asn-116 of EL into the analogous positions within LPL and HL, resulting in N-linked glycosylation at this site. Glycosylation at this site suppressed the LPL hydrolysis of synthetic substrates, LDL, HDL2, and HDL3 but had little effect on HL activity. These data suggest that N-linked glycosylation at Asn-116 reduces the ability of EL to hydrolyze lipids in LDL and HDL2.  相似文献   

12.
The rabbit as an animal model of hepatic lipase deficiency   总被引:3,自引:0,他引:3  
A natural deficiency of hepatic lipase in rabbits has been exploited to gain insights into the physiological role of this enzyme in the metabolism of plasma lipoproteins. A comparison of human and rabbit lipoproteins revealed obvious species differences in both low-density lipoproteins (LDL) and high-density lipoproteins (HDL), with the rabbit lipoproteins being relatively enlarged, enriched in triacylglycerol and depleted of cholesteryl ester. To test whether these differences related to the low level of hepatic lipase in rabbits, whole plasma or the total lipoprotein fraction from rabbits was either kept at 4 degrees C or incubated at 37 degrees C for 7 h in (i) the absence of lipase, (ii) the presence of hepatic lipase and (iii) the presence of lipoprotein lipase. Following incubation, the lipoproteins were recovered and subjected to gel permeation chromatography to determine the distribution of lipoprotein components across the entire lipoprotein spectrum. An aliquot of the lipoproteins was subjected also to gradient gel electrophoresis to determine the particle size distribution of the LDL and HDL. Both hepatic lipase and lipoprotein lipase hydrolysed lipoprotein triacylglycerol and to a much lesser extent, also phospholipid. There were, however, obvious differences between the enzymes in terms of substrate specificity. In incubations containing hepatic lipase, there was a preferential hydrolysis of HDL triacylglycerol and a lesser hydrolysis of VLDL triacylglycerol. By contrast, lipoprotein lipase acted primarily on VLDL triacylglycerol. When more enzyme was added, both lipases also acted on LDL triacylglycerol, but in no experiment did lipoprotein lipase hydrolyse the triacylglycerol in HDL. Coincident with the hepatic lipase-induced hydrolysis of LDL and HDL triacylglycerol, there were marked reductions in the particle size of both lipoprotein fractions, which were now comparable to those of human LDL and HDL3, respectively.  相似文献   

13.
The biologic properties of two major proteoglycans of bovine aorta, heparan sulfate proteoglycan and chondroitin sulfate-dermatan sulfate proteoglycan were compared. The heparan sulfate proteoglycan was isolated either by elastase digestion or by 4.0 M guanidine hydrochloride extraction, of aorta tissue, fractionated by CsCl isopycnic centrifugation and purified by chondroitinase ABC treatment. The first method resulted in considerably greater yield (about 70% of the total heparan sulfate proteoglycan of the tissue) than the second procedure (12% of total). The chondroitin sulfate-dermatan sulfate proteoglycan was obtained by 4.0 M guanidine-HCl extraction of aorta tissue followed by CsCl isopycnic centrifugation. The chemical composition of both heparan sulfate proteoglycan preparations was similar. Unlike the chondroitin sulfate-dermatan sulfate proteoglycan, which eluted in the void volume of Sepharose CL-6B column, the heparan sulfate proteoglycan preparations were each resolved into a high molecular weight fraction (kav = 0.18 and 0.13) and a low molecular weight fraction (kav = 0.47 and 0.36). The heparan sulfate proteoglycan preparations exhibited significantly more potent anticoagulant and platelet aggregation inhibitory activities than the chondroitin sulfate-dermatan sulfate proteoglycan. The protein core of the proteoglycan molecules did not seem to be essential for their hemostatic properties. The complex forming ability of the heparan sulfate proteoglycan with serum low density lipoproteins (LDL) was much less than that of chondroitin sulfate-dermatan sulfate proteoglycan in the presence and absence of Ca2+. Interaction between heparan sulfate proteoglycan and LDL was also much more sensitive to changes in the ionic strength of the medium than that of chondroitin sulfate-dermatan sulfate proteoglycan and the lipoprotein. Since the total sulfate content of both proteoglycans is almost similar, the smaller molecular size and hence the lower overall charge density of the heparan sulfate proteoglycan appears to be partly responsible for its low affinity for LDL. The differences in biologic properties of the two proteoglycans might have implications in the pathophysiology of cardiovascular diseases.  相似文献   

14.
Endothelial lipase (EL) influences high density lipoprotein (HDL) metabolism in vivo and mediates bridging and uptake of HDL particles independent of its lipolytic activity in vitro. To determine whether EL has a nonlipolytic ligand function in HDL metabolism in vivo, 1 x 1011 particles of a recombinant adenovirus encoding human EL (AdEL), catalytically inactive human EL (AdELS149A), or control (Adnull) were injected into wild-type, apoA-I transgenic, and hepatic lipase knockout mice. ELS149A protein was expressed at higher levels than wild-type EL. EL and ELS149A protein were both substantially increased in the postheparin plasma compared with preheparin, indicating that both the wild-type and mutant EL were bound to cell-surface heparan sulfate proteoglycans. Overexpression of wild-type EL was associated with a significantly increased postheparin-plasma phospholipase activity and dramatically decreased levels of total cholesterol, HDL cholesterol, phospholipids, and apoA-I. Injection of AdELS149A did not result in increased phospholipase activity confirming that ELS149A was catalytically inactive. Expression of ELS149A did not decrease lipid or apoA-I levels in wild-type and apoA-I transgenic mice yet led to an intermediate reduction of total cholesterol, HDL cholesterol, and phospholipids in hepatic lipase-deficient mice compared with control and EL-expressing mice. Our study demonstrates for the first time that EL has both a lipolytic and nonlipolytic function in HDL metabolism in vivo. Lipolytic activity of EL, however, seems to be most important for its effects on systemic HDL metabolism.  相似文献   

15.
High-density lipoprotein 3 (HDL3) binds to capillary endothelial cells when their lumen surfaces are exposed to 125I-HDL3 by post-mortem perfusion of whole brain. Kinetic studies of binding of HDL3 to isolated membranes show that HDL3 binds only to endothelial membranes with high affinity (Kd = 7 micrograms/ml). Trypsin treatment of membranes abolishes HDL3 binding. High-affinity binding sites for HDL3 were recovered when endothelial cells from bovine brain capillaries were maintained in culture (Kd = 13 micrograms/ml HDL3 protein). The characteristics of the binding were preserved up to the 6th passage. Competition experiments using isolated luminal membranes or cultured endothelial cells indicate that only HDL3 and not LDL or methylated LDL, are able to compete binding of 125I-HDL3. Furthermore, the inhibition of 125I-HDL3 binding by lipoprotein A-I and lipoprotein A-I:A-II strongly suggests that apolipoprotein A-I is implicated in the formation of HDL3-receptor complexes. The binding is increased by loading cells with free cholesterol or LDL cholesterol. In addition, surface-bound 125I-HDL3 remains sensitive to mild trypsin treatment after subsequent incubation of BBCE at 37 degrees C. HDL3 bound to the cell surface is not endocytosed, but rather rapidly released into the medium after binding (t1/2 = 5 min).  相似文献   

16.
Primary cultures of rabbit hepatocytes which were preincubated for 20 h in a medium containing lipoprotein-deficient serum subsequently bound, internalized and degraded 125I-labeled high-density lipoproteins2 (HDL2). The rate of degradation of HDL2 was constant in incubations from 3 to 25 h. As the concentration of HDL2 in the incubation medium was increased, binding reached saturation. At 37 degrees C, half-maximal binding (Km) was achieved at a concentration of 7.3 micrograms of HDL2 protein/ml (4.06 X 10(-8)M) and the maximum amount bound was 476 ng of HDL2 protein/mg of cell protein. At 4 degrees C, HDL2 had a Km of 18.6 micrograms protein/ml (1.03 X 10(-7)M). Unlabeled low-density lipoproteins (LDL) inhibited only at low concentrations of 125I-labeled HDL2. Quantification of 125I-labeled HDL2 binding to a specific receptor (based on incubation of cells at 4 degrees C with and without a 50-fold excess of unlabeled HDL) yielded a dissociation constant of 1.45 X 10(-7)M. Excess HDL2 inhibited the binding of both 125I-labeled HDL2 and 125I-labeled HDL3, but excess HDL3 did not affect the binding of 125I-labeled HDL3. Preincubation of hepatocytes in the presence of HDL resulted in only a 40% reduction in specific HDL2 receptors, whereas preincubation with LDL largely suppressed LDL receptors. HDL2 and LDL from control and hypercholesterolemic rabbits inhibited the degradation of 125I-labeled HDL2, but HDL3 did not. Treatment of HDL2 and LDL with cyclohexanedione eliminated their capacity to inhibit 125I-labeled HDL2 degradation, suggesting that apolipoprotein E plays a critical role in triggering the degradative process. The effect of incubation with HDL on subsequent 125I-labeled LDL binding was time-dependent: a 20 h preincubation with HDL reduced the amount of 125I-labeled LDL binding by 40%; there was a similar effect on LDL bound in 6 h but not on LDL bound in 3 h. The binding of 125I-labeled LDL to isolated liver cellular membranes demonstrated saturation kinetics at 4 degrees C and was inhibited by EDTA or excess LDL. The binding of 125I-labeled HDL2 was much lower than that of 125I-labeled LDL and was less inhibited by unlabeled lipoproteins. The binding of 125I-labeled HDL3 was not inhibited by any unlabeled lipoproteins. EDTA did not affect the binding of either HDL2 or HDL3 to isolated liver membranes. Hepatocytes incubated with [2-14C]acetate in the absence of lipoproteins incorporated more label into cellular cholesterol, nonsaponifiable lipids and total cellular lipid than hepatocytes incubated with [2-14C]acetate in the presence of any lipoprotein fraction. However, the level of 14C-labeled lipids released into the medium was higher in the presence of medium lipoproteins, indicating that the effect of those lipoproteins was on the rate of release of cellular lipids rather than on the rate of synthesis.  相似文献   

17.
Hepatic lipase (HL) is a key player in lipoprotein metabolism by modulating, through its lipolytic activity, the triglyceride (TG) and phospholipid content of apolipoprotein B (apoB)-containing lipoproteins and of high density lipoproteins (HDL), thereby affecting their size and density. A new and separate role has been suggested for HL in cellular lipoprotein metabolism, in which it serves as a ligand promoting cellular uptake of apoB-containing remnant lipoproteins and HDL. We tested the hypothesis that HL has both a lipolytic and a nonlipolytic role in human lipoprotein metabolism, by measuring lipid plasma concentrations, lipoprotein density distribution by density gradient ultracentrifugation, and lipoprotein composition, in three subjects with HL deficiency: two of the patients (S-1 and S-3) were characterized as having neither plasma HL activity nor detectable HL protein; the third subject (S-2) had no plasma HL activity but a detectable amount (35.5 ng/ml) of HL protein. All HL-deficient subjects showed a severalfold increase in lipoprotein TG content across the lipoprotein density spectrum [very low density lipoprotein (VLDL), intermediate density lipoprotein (IDL), low density lipoprotein (LDL), and HDL] as compared with control subjects. They also had remarkably more buoyant LDL particles (LDL-R(f) = 0.342;-0.394) as compared with the control subjects (LDL-R(f) = 0.303). Subjects S-1 and S-3 (no HL activity or protein) presented with a distinct increase in cholesterol and apoB levels in the IDL and VLDL density range as compared with patient S-2, with detectable HL protein, and the control subjects.This study provides evidence in humans that HL indeed plays an important role in lipoprotein metabolism independent of its enzymatic activity: in particular, inactive HL protein appears to affect VLDL and IDL particle concentration, whereas HL enzymatic activity seems to influence VLDL-, IDL-, LDL-, and HDL-TG content and their physical properties.  相似文献   

18.
Apolipoprotein E (apoE) plays a major role in lipoprotein metabolism by mediating the binding of apoE-containing lipoproteins to receptors. The role of hepatic apoE in the catabolism of apoE-free lipoproteins such as low density lipoprotein (LDL) and high density lipoprotein-3 (HDL(3)) is however, unclear. We analyzed the importance of hepatic apoE by comparing human LDL and HDL(3) metabolism in primary cultures of hepatic cells from control C57BL/6J and apoE knockout (KO) mice. Binding analysis showed that the maximal binding capacity (Bmax) of LDL, but not of HDL(3), is increased by twofold in the absence of apoE synthesis/secretion. Compared to control hepatic cells, LDL and HDL(3) holoparticle uptake by apoE KO hepatic cells, as monitored by protein degradation, is reduced by 54 and 77%, respectively. Cleavage of heparan sulfate proteoglycans (HSPG) by treatment with heparinase I reduces LDL association by 21% in control hepatic cells. Thus, HSPG alone or a hepatic apoE-HSPG complex is partially involved in LDL association with mouse hepatic cells. In apoE KO, but not in normal hepatic cells, the same treatment increases LDL uptake/degradation by 2.4-fold suggesting that in normal hepatic cells, hepatic apoE increases LDL degradation by masking apoB-100 binding sites on proteoglycans. Cholesteryl ester (CE) association and CE selective uptake (CE/protein association ratio) from LDL and HDL(3) by mouse hepatic cells were not affected by the absence of apoE expression. We also show that 69 and 72% of LDL-CE hydrolysis in control and apoE KO hepatic cells, respectively, is sensitive to chloroquine revealing the importance of a pathway linked to lysosomes. In contrast, HDL(3)-CE hydrolysis is only mediated by a nonlysosomal pathway in both control and apoE KO hepatic cells. Overall, our results indicate that hepatic apoE increases the holoparticle uptake pathway of LDL and HDL(3) by mouse hepatic cells, that HSPG devoid of apoE favors LDL binding/association but impairs LDL uptake/degradation and that apoE plays no significant role in CE selective uptake from either human LDL or HDL(3) lipoproteins.  相似文献   

19.
To study the interaction between low-density lipoprotein (LDL) and granules from rat serosal mast cells in vitro, mast cells were stimulated with the degranulating agent 48/80 to induce exocytosis of the secretory granules. Subsequent incubation of the exocytosed granules with 125I-LDL resulted in binding of the labelled LDL to the granules. When increasing amounts of agent 48/80 were added to mast-cell suspensions, a dose-dependent release of granules was observed and a parallel increase in the amount of 125I-LDL bound to granules resulted. 125I-LDL bound to a single class of high-affinity binding sites on the granules. At saturation, 105 ng of LDL were bound per microgram of granule protein. The lipoprotein binding to mast-cell granules was apolipoprotein(apo)-B + E-specific. Thus 125I-LDL binding to the granules was effectively compared for by LDL (apo-B) or by dimyristoyl phosphatidylcholine vesicles containing apo-E, but not by high-density lipoprotein (HDL3) containing apo-AI as their major protein component. Neutralization by acetylation of the positively charged amino groups of apo-B of LDL or presence of a high ionic strength in the incubation medium prevented LDL from binding to the granules, indicating the presence of ionic interactions between the positively charged amino acids of LDL and negatively charged groups of the granules. It could be demonstrated that LDL bound to the negatively charged heparin proteoglycan of the granules. Thus treatment of granules with heparinase resulted in loss of their ability to bind LDL, and substances known to bind to heparin, such as Toluidine Blue, avidin, lipoprotein lipase, fibronectin and protamine, all effectively competed with LDL for binding to the granules. The results show that LDL is efficiently bound to the heparin proteoglycan component of mast-cell granules once the mast cells are stimulated to release their granules into the extracellular space.  相似文献   

20.
1. Male chicks were deprived of feed for 48 hr to study the effect of metabolic stress on hepatic membrane and lipoprotein fluidity and binding of radioiodinated lipoproteins to hepatic membranes. 2. Plasma levels of low density lipoprotein (LDL) and high density lipoprotein (HDL) were markedly and slightly elevated, respectively. 3. There was a reduction in lipoprotein and hepatic membrane fluidity. 4. Binding of [125I]LDL, but not [125I]HDL, to hepatic membranes was decreased. 5. It is suggested that a reduction in the fluidity of LDL and/or hepatic membranes impedes LDL catabolism in vivo.  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