当前位置: 首页 > 医学版 > 期刊论文 > 内科学 > 循环研究杂志 > 2005年 > 第6期 > 正文
编号:11255815
New Insights Into the Regulation of HDL Metabolism and Reverse Cholesterol Transport
     The Departments of Medicine and Physiology (G.F.L), University of Toronto, Canada

    The Institute for Translational Medicine and Therapeutics (D.J.R.), The Cardiovascular Institute

    the Institute for Diabetes, Obesity

    Metabolism, University of Pennsylvania School of Medicine, Philadelphia.

    Abstract

    The metabolism of high-density lipoproteins (HDL), which are inversely related to risk of atherosclerotic cardiovascular disease, involves a complex interplay of factors regulating HDL synthesis, intravascular remodeling, and catabolism. The individual lipid and apolipoprotein components of HDL are mostly assembled after secretion, are frequently exchanged with or transferred to other lipoproteins, are actively remodeled within the plasma compartment, and are often cleared separately from one another. HDL is believed to play a key role in the process of reverse cholesterol transport (RCT), in which it promotes the efflux of excess cholesterol from peripheral tissues and returns it to the liver for biliary excretion. This review will emphasize 3 major evolving themes regarding HDL metabolism and RCT. The first theme is that HDL is a universal plasma acceptor lipoprotein for cholesterol efflux from not only peripheral tissues but also hepatocytes, which are a major source of cholesterol efflux to HDL. Furthermore, although efflux of cholesterol from macrophages represents only a tiny fraction of overall cellular cholesterol efflux, it is the most important with regard to atherosclerosis, suggesting that it be specifically termed macrophage RCT. The second theme is the critical role that intravascular remodeling of HDL by lipid transfer factors, lipases, cell surface receptors, and non-HDL lipoproteins play in determining the ultimate metabolic fate of HDL and plasma HDL-c concentrations. The third theme is the growing appreciation that insulin resistance underlies the majority of cases of low HDL-c in humans and the mechanisms by which insulin resistance influences HDL metabolism. Progress in our understanding of HDL metabolism and macrophage reverse cholesterol transport will increase the likelihood of developing novel therapies to raise plasma HDL concentrations and promote macrophage RCT and in proving that these new therapeutic interventions prevent or cause regression of atherosclerosis in humans.

    Key Words: high density lipoprotein insulin resistance lipase lipoprotein reverse cholesterol transport

    Introduction

    Population studies have shown a highly consistent, inverse correlation between plasma concentrations of high-density lipoprotein cholesterol (HDL-c) and its major protein apolipoprotein A-I (apoA-I) and atherosclerotic cardiovascular disease risk in humans.1 HDL-c and apoA-I concentrations could simply be integrative markers of other atherosclerotic cardiovascular risk factors and not themselves causal in the disease process. However, studies in animals strongly suggest that HDL and apoA-I have direct antiatherogenic properties. HDL metabolism is somewhat more complex than that of the other major lipoprotein fractions, in that the individual lipid and apolipoprotein components of HDL are mostly assembled after secretion, are frequently exchanged with or transferred to other lipoproteins, are actively remodeled within the plasma compartment, and are cleared at least in part independent from one another. Although the concept of reverse cholesterol transport (RCT) from macrophages to liver and ultimately biliary excretion is the most popular mechanism to explain the ability of HDL to inhibit atherosclerosis, many other properties of HDL have been demonstrated in vitro that could contribute to its antiatherogenic effects.2 To systematically elucidate the relationship between HDL and atherosclerotic risk, we need to better understand the key regulatory factors that determine the net plasma concentration of HDL particles and the flux of lipids through the HDL and RCT pathways.

    In this review, we will attempt to integrate the various components of HDL metabolism, with an emphasis on 3 major evolving themes. The first theme is the concept that RCT does not necessarily imply a unidirectional flux of cholesterol from extrahepatic tissues to liver. HDL also promote substantial cholesterol egress from the liver, which may be a significant source of lipidation for newly secreted nascent HDL particles. HDL should perhaps be viewed as a universal plasma acceptor lipoprotein for cholesterol efflux from cells, playing an important role in unloading excess cholesterol from cells for the maintenance of cellular cholesterol homeostasis. Although efflux of cholesterol from macrophages represents only a tiny fraction of overall cellular cholesterol efflux, it is the most important with regard to atherosclerosis, suggesting that it be specifically termed macrophage RCT.

    The second major theme is the critical role that intravascular remodeling of HDL by lipid transfer factors, lipases, cell-surface receptors, and non-HDL lipoproteins play in determining the ultimate metabolic fate of HDL. Altered HDL composition as a result of enhanced intravascular remodeling results in more rapid clearance, and HDL clearance, rather than production, is the major determinant of plasma HDL-c concentrations.

    The third major theme is the growing appreciation that insulin resistance appears to be associated with the majority of cases of low HDL-c in humans,3,4 although the exact prevalence of the metabolic syndrome and insulin resistance among those with low HDL-c in the general population has not yet been clearly established. Understanding the mechanisms of HDL lowering in insulin resistant states, which remain incompletely understood, could provide insight into the relationship between low HDL and atherosclerosis. There are numerous atherogenic factors that comprise the metabolic syndrome, and it is difficult in this setting to know whether low HDL is simply an associated risk marker or an important cause per se of atherosclerosis. Overall, low HDL probably represents both a risk marker as well as a causal factor for atherosclerotic disease in the metabolic syndrome.

    Regulation of the Synthesis and Secretion of ApoA-I and Nascent HDL Particles

    ApoA-I is present on the majority of HDL particles and constitutes 70% of the apolipoprotein content of HDL particles; as a result, plasma apoA-I concentrations correlate closely with plasma HDL-c. ApoA-I is secreted predominantly by liver and intestine as lipid-poor apoA-I and nascent phopholipid-rich cholesterol-poor HDL particles (Figure 1). Nascent apoA-I/HDL acquire additional phospholipids and cholesterol via cellular efflux as well as by transfer of surface components of triglyceride (TG)-rich lipoproteins (TRL; chylomicrons and VLDL) during lipoprotein lipase (LPL)-mediated intravascular lipolysis of TRL.

    Human apoA-I transgenic mice have elevated levels of HDL containing human apoA-I and are protected against atherosclerosis,5 proving in principle that apoA-I overexpression can positively influence both plasma HDL concentrations as well as atherosclerosis progression. An important issue is whether the rate of apoA-I production is a major determinant of plasma HDL-c and apoA-I concentrations in humans. Regulation of the apoA-I gene occurs primarily at the transcriptional level, mediated by the cis-acting sequences in the proximal promoter of the apoA-I gene, and partly at the posttranscriptional level by increasing the stability of the partially spliced and unspliced nuclear apoA-I mRNA (reviewed in references 6 and 7). Dietary fat, alcohol, estrogen, androgens, thyroid hormone, retinoids, glucocorticoids, fibrates, niacin, and HMG-CoA reductase inhibitors are some of the many nutritional, hormonal, and pharmacological factors known to influence transcriptional induction of the apoA-I gene (reviewed in references 6 and 7). Promotion of apoA-I gene transcription and biosynthesis is an attractive therapeutic target for drug development. Based on animal studies, upregulation of apoA-I expression in humans would be expected to raise HDL-c concentrations and provide protection against atherosclerosis.

    In vivo studies of HDL metabolism in human populations over a wide range of body weights, plasma TG concentrations, and presumably insulin sensitivity, have shown that clearance of apoA-I, rather than its production rate, is the most important determinant of the variability of plasma HDL-c and apoA-I concentrations in populations.8eC11 Within phenotypically narrowly-defined populations (ie, humans within a narrow range of body weights, insulin sensitivity, and plasma TG concentrations), production rates of apoA-I also are an important determinant of the variability of plasma HDL concentrations. Nutritional interventions such as, a switch from high-carbohydrate to a high-fat diet appear to exert their major effect on production rates of apoA-I rather than on clearance.12

    In summary, apoA-I production is influenced by many factors and apoA-I transcriptional regulation has an impact on plasma HDL concentrations. Variation in plasma HDL and apoA-I concentrations in the general population, however, is primarily a function of variation in clearance rather than production rates.

    HDL Apolipoproteins and Protein Components Other Than ApoA-I

    ApoA-II is the second most abundant apolipoprotein of HDL, found on approximately two-thirds of HDL particles. Its physiologic role has not been fully defined. HDL also contains a variety of other proteins, including apoA-IV, apo C-I, apo C-II, apo C-III, apoD, apoE, apoJ, apo L-I, apoM, serum amyloid A proteins, ceruloplasmin, transferrin, and enzymes such as LCAT, PON1, and PAF-AH/Lp-PLA2. These proteins have qualitative effects on HDL function. HDL are comprised of a number of discrete, quantitatively minor species for which discrete metabolic roles are emerging. For a more detailed discussion of this topic please see the online data supplement available at http://circres.ahajournals.org.

    Acquisition of Unesterified Cellular Cholesterol by ApoA-I and Nascent HDL and Its Esterification in Plasma

    As mentioned previously, nascent HDL is secreted by liver and intestine as lipid-poor apoA-I or small particles containing apoA-I and phospholipids and with pre- mobility on electrophoresis. Pre- HDL generally contain 2 copies of apoA-I per particle and 10% by mass of lipid (free cholesterol and phospholipids).13 Similar particles are regenerated during the catabolism of TRL and mature HDL.14 Pre- HDL are present as minor components in plasma and are believed to play a role as initial acceptors of free cholesterol from cells.15 Distinct from pre- HDL is monomolecular, pre--migrating, lipid-poor or -free apoA-I, which promotes cholesterol and phospholipid efflux from cells via ABCA1 (see below). The presence of lipid-poor or -free apoA-I in vivo has not been definitively established, because it is rapidly either reincorporated into mature HDL, relipidated to form pre- HDL, irreversibly lost, or cleared by the kidneys. It is likely that pre- HDL particles are formed very rapidly when monomeric apoA-I is exposed to cells.

    Lipid efflux from cells to these acceptor lipoprotein particles can occur by a number of mechanisms, including regulated transporter-facilitated processes as well as aqueous diffusion (reviewed in reference 16). ABCA1 is an important cellular protein that facilitates efflux of cellular cholesterol to lipid-poor apoA-I as the preferred acceptor. The critical role that ABCA1 plays in determining plasma HDL-c concentrations has been well established. Humans with Tangier disease, who are homozygous for functional mutations of ABCA1, have virtually undetectable plasma concentrations of HDL-c and apoA-I, because of impaired lipidation and subsequent rapid catabolism of lipid-poor apoA-I. Even heterozygotes for functional ABCA1 mutations generally have HDL-c levels that are less than fifth percentile. Cholesterol efflux from skin fibroblasts is impaired in persons with functional ABCA1 mutations and correlates with plasma HDL-c concentrations.17eC19 An important question remains as to the prevalence of ABCA1 mutations in the general population with low HDL-c levels. Studies in a large cohort of Dutch patients with hypoalphalipoproteinemia revealed a low prevalence (4.5%) of defects in ABCA1-mediated cholesterol efflux from skin fibroblasts.20 Probands of French-Canadians with moderate to severe familial hypoalphalipoproteinemia were shown to have a higher prevalence of ABCA1 mutations (29%)21 but accounted for no cases of low HDL-c in those with the metabolic syndrome.22 Other studies have also failed to find functional ABCA1 mutations in populations with low plasma HDL-cholesterol concentrations.23,24 However, in a recent population-based study, 16% of individuals with HDL-c levels less than fifth percentile had nonsynonymous mutations in ABCA1 compared with only 2% of those with HDL-c levels greater than ninety-fifth percentile.25 Similar findings have been reported by others.26 Thus rare private ABCA1 mutations may contribute substantially to the low HDL-c state. Because isolated familial hypoalphalipoproteinemia is a less prevalent condition than the low HDL-c that occurs in conjunction with insulin resistance, it is likely that primary cellular lipid efflux defects account for the minority of cases of hypoalphalipoproteinemia in the general population.

    Animal studies have helped to clarify the physiologic role of ABCA1. ABCA1 knockout mice have an extremely low HDL-c phenotype similar to that of humans with Tangier disease (reviewed in references 27 and 28). Interestingly, macrophage-specific deficiency of ABCA1 had a minimal effect on reducing plasma HDL-c levels in mice, but it resulted in significantly increased atherosclerosis (presumably, although not proven, attributable to impaired macrophage RCT).29,30 Thus, macrophage ABCA1 appears to contribute little to bulk lipidation of HDL and therefore to plasma HDL-c levels31 but does seem to be important for protection from atherosclerosis. Conversely, liver-specific deficiency of ABCA1 in mice dramatically reduced HDL-c levels by 80%.32 Thus, hepatic ABCA1 appears to be critical for the initial lipidation of newly secreted lipid-poor apoA-I (Figure 1), protecting it from rapid degradation and allowing it to go on to form mature HDL. ABCA1 overexpression in liver increased HDL-c levels33,34 and in macrophages and liver was associated with protection against atherosclerosis.35eC37

    Macrophages have other pathways for effluxing excess cholesterol to HDL. ABCG1 and ABCG4 were reported to mediate net mass efflux of cellular cholesterol to mature HDL but not to lipid-poor apoA-I.38 ABCG1 knockout mice have marked macrophage cholesterol accumulation, and ABCG1 overexpression protects tissues from cholesterol accumulation.39 Thus ABCG1 may be another important pathway for macrophage cholesterol efflux. The contribution of ABCG1 to integrated macrophage reverse cholesterol transport or its effect on atherosclerosis has not been determined. Scavenger receptor class B, type I (SR-BI) may also play a role in mediating cellular cholesterol efflux to mature HDL.40 The influence of macrophage SR-BIeCmediated efflux on HDL metabolism and reverse cholesterol transport has not been definitively established. The SR-BI knockout mouse has increased HDL-c levels,41,42 probably because of the key role of hepatic SR-BI in HDL catabolism (see below). However, these mice have increased atherosclerosis,43 which could be due in part to impaired macrophage cholesterol efflux. Indeed, mice specifically lacking macrophage SR-BI have normal HDL-c levels but increased atherosclerosis,44 consistent with this concept.

    The discovery of ABCA1 as a major regulator of cellular cholesterol efflux and determinant of HDL-c levels has led to intense interest in the molecular regulation of ABCA1 expression45 and as a target for the development of new therapies. The liver X receptors (LXR) and are nuclear receptors that sense cellular cholesterol excess and whose physiological ligands are oxysterols.46,47 Synthetic LXR agonists have been shown to upregulate macrophage ABCA1 expression, increase cholesterol efflux in vitro, and reduce atherosclerosis in mice.48,49 Thus far, enthusiasm over potential upregulation of LXR as a therapeutic target has been tempered by the development of hypertriglyceridemia and fatty liver, in part because of LXR-mediated upregulation of hepatic SREBP-1c. Of note, synthetic agonists of peroxisome-proliferator-activated receptor (PPAR) and PPAR have been shown to upregulate LXR and ABCA1 and promote cholesterol efflux in vitro,50 but the in vivo relevance of this observation has not been established.

    Despite the enthusiasm for RCT as a mechanism for HDL protecting against atherosclerosis, it is difficult to measure integrated RCT from macrophage to liver and feces in vivo. Indeed, studies of whole body RCT have suggested that manipulation of apoA-I levels did not change the rate at which peripheral cholesterol was returned to the liver.51,52 However, a method in which macrophages were loaded with cholesterol and labeled with a cholesterol tracer ex vivo and injected intraperitoneally into mice showed a significantly greater rate of macrophage to feces RCT in apoA-I overexpressing mice, indicating that apoA-I does promote macrophage-specific RCT.53 Future studies are necessary to determine the roles of specific gene products and of pharmacological approaches on macrophage specific RCT.

    Intravascular Maturation and Remodeling of HDL Particles by Cholesterol Esterification, Lipid Exchange, and Lipolytic Modification

    The above discussion has focused on the biosynthesis and initial lipidation of nascent HDL particles. This is followed by dynamic intravascular maturation and remodeling of HDL. A number of lipid transfer factors and lipolytic enzymes play key roles in this process. Many of these factors have diverse functions and in many cases their effects on atherosclerosis may be unrelated to their effects on HDL. The following discussion of these factors will focus exclusively on their roles in HDL metabolism.

    Lecithin-cholesterol acyltransferase (LCAT) plays an important role in the maturation of nascent to mature HDL. ApoA-I and nascent HDL acquire cholesterol from cells in the unesterified (or free) form, but much of plasma HDL cholesterol is in the form of cholesteryl esters (CE). LCAT catalyzes the transfer of 2-acyl groups from lecithin to free cholesterol, generating CE and lysolecithin.54 Hydrophobic CE are retained in the HDL core forming larger mature HDL particles. The activity of LCAT is critical to normal HDL metabolism. In humans, genetic LCAT deficiency syndromes are associated with markedly reduced HDL-c and apoA-I levels55 and rapid catabolism of CE-poor apoA-I.56 The mouse LCAT knockout has a similar phenotype.27,57,58 However, the importance of LCAT to the process of RCT, although postulated, has not been firmly established. Absent functional LCAT activity does not necessarily result in a major defect in RCT, perhaps because much RCT may occur as unesterified cholesterol. Indeed, in humans unesterified cholesterol in HDL can be directly transferred to the liver and secreted in bile.59

    Cholesteryl ester transfer protein (CETP) is a hydrophobic glycoprotein made by liver and adipose that circulates in the plasma bound to lipoproteins.60 It promotes the redistribution and equilibration of hydrophobic lipids packaged within the lipoprotein core (CE and TG) between HDL and the apo BeCcontaining lipoproteins (LDL, IDL, VLDL, chylomicrons, and remnants; Figure 2). The net effect of CETP action on HDL is depletion of CE and enrichment with TG, with an overall net reduction in the size of the HDL particle. Under normal physiological conditions the amount of CETP mass in plasma is not the rate-limiting factor that determines CE distribution between the slowly catabolized LDL and HDL, whereas CETP mass is rate limiting in the transfer of lipids between HDL and the more rapidly turning over TRL. Heterotransfer of CEs and TGs between HDL and TRL is increased in hypertriglyceridemia61 and in the postprandial state.62 The preference of CETP for lipid exchange between HDL and TRL is attributable to suppression of LDL lipid exchange by the naturally-occurring lipid transfer inhibitor protein (LTIP).63 Overall, the magnitude of net flux of CE and TG between the lipoproteins is probably dependent to a greater extent on the relative sizes of the respective lipoprotein pools rather than on the amount of CETP mass per se.

    The importance of CETP in HDL metabolism was conclusively demonstrated by the discovery of CETP-deficient patients in Japan, who have extremely elevated levels of HDL-c64 and reduced turnover of apoA-I.65 Mice naturally lack CETP, and when it is transgenically expressed HDL-c levels are markedly reduced.66 Thus, the activity of CETP in exchanging CE out of HDL for TG has the net effect of reducing HDL-c levels. The impact of CETP deficiency on cardiovascular risk has not yet been resolved. Synthetic CETP inhibitors have been developed and have been shown to decrease atherosclerosis in rabbits67,68 and to effectively raise HDL-c levels in humans,69eC71 spurring substantial interest in this approach. However, there remains some uncertainty, because CETP-mediated transfer of CEs from HDL to apo B-containing lipoproteins may be one pathway of RCT. In humans, radiolabeled HDL CE that eventually was excreted into bile (as free cholesterol or bile acid) was transported to the liver almost entirely after transfer (presumably via CETP) to apoB-containing lipoproteins.59 In addition, there is increasing appreciation that the functionality of HDL particles may be an important determinant of their antiatherosclerotic effects, and it is not yet clear whether the large CE-rich HDL particles that accumulate as a result of CETP inhibition will be effective in inhibiting atherogenesis. Thus, it will be extraordinarily interesting and important to determine whether CETP inhibition will reduce atherosclerosis in humans.

    Phospholipid transfer protein (PLTP) transfers surface phospholipids from TRL to HDL during TG lipolysis72 and appears to account for most of the phospholipid transfer in human plasma. Targeted disruption of PLTP in mice results in an 60% reduction in HDL lipid and apoA-I levels73 because of enhanced clearance of phosphaditylcholine-depleted HDL particles. Human PLTP transgenic mice have increased levels of preeC-1 migrating HDL, apoA-I, and phospholipid.74 PLTP is able to remodel HDL particles into larger HDL particles by particle fusion, with concomitant release of lipid-poor apoA-I.72 In addition to PLTP activity, HDL apolipoprotein composition and core lipid composition play an important role in PLTP-mediated HDL particle conversion.72 Remodeling by PLTP is markedly enhanced in triglyceride-enriched HDL,75 perhaps because of destabilization of apoA-I in these particles.76 Genetic PLTP deficiency in humans has not been conclusively described, and the role of PLTP in human physiology and pathophysiology has yet to be clearly elucidated.

    Lipoprotein lipase (LPL) is secreted by many tissues of the body, particularly metabolically active adipose tissue and muscle (reviewed in reference 77). It is transferred to the luminal surface of endothelial cells, where it is bound as homodimers to heparan sulfate proteoglycans and can be released by administration of heparin. LPL has predominantly TG lipase activity with minor phospholipase activity. It is the principal enzyme responsible for the hydrolysis of triglycerides in TRL and the release of free fatty acids, transforming large TG-rich particles into smaller TG-depleted remnant lipoproteins. In the process of hydrolyzing TRL, redundant surface lipid (free cholesterol and phospholipid) and apolipoproteins are transferred from TRL to HDL particles, contributing significantly to plasma HDL-c and HDL-associated apoA-I concentrations. In addition, an accumulation of TRL secondary to LPL deficiency enhances CETP-mediated lipid exchange between TRL and HDL, thereby enriching HDL particles with triglycerides and depleting them of CE.78 Not only would this have the direct effect of lowering plasma HDL-c concentration but the HDL lipid compositional change itself predisposes HDL particles to PLTP-mediated phospholipid exchange and hepatic lipase-mediated clearance of HDL from the circulation, as will be discussed.

    Plasma HDL-c concentrations correlate positively with postheparin plasma LPL activity.79 Homozygous and heterozygous LPL deficiency syndromes in humans are associated with reduced plasma HDL concentration. Similarly, LPL knockout mice not only have marked hypertriglyceridemia but their HDL levels are also low.80 Transgenic overexpression of LPL in mice is associated with elevations of HDL. Interestingly, significant correlations between LPL activity and HDL levels in LPL transgenic mice were only evident in the presence of the CETP transgene,81 suggesting that lipid transfer between TRL and HDL is the dominant mechanism whereby LPL over- and under-expression influences HDL levels. Pharmacological upregulation of LPL elevates HDL,82 whereas in vivo monoclonal antibodyeCmediated inhibition of LPL in monkeys results in a lowering of HDL-c and apoA-I attributable to a marked increase in HDL apolipoprotein catabolic rate and degradation of apoA-I in the kidneys.83 As will be discussed for the other lipases, LPL has also been shown to mediate selective HDL CE uptake by hepatocytes in culture, independent of its lipolytic activity.84

    Hepatic lipase (HL) is a lipolytic enzyme synthesized by hepatocytes and has both TG lipase as well as phospholipase A1 activity (for more detailed reviews of HL structure/function relationships, synthesis, regulation, and role in atherosclerosis see references 85 through 87). It is bound to the surface of liver sinusoidal capillaries anchored by heparan sulfate proteoglycans. Several lines of evidence indicate that HL plays an important role in mediating HDL metabolism. In contrast to LPL, HL has greater activity against HDL than VLDL or chylomicrons and converts larger HDL particles to smaller HDL remnants, pre- HDL, and lipid-poor or -free apoA-I. The magnitude of the effect of HL on HDL is highly dependent on the composition of HDL. CETP-mediated TG enrichment and CE depletion of HDL, as occurs in hypertriglyceridemic conditions, greatly enhances HDL remodelling by HL (reviewed in reference 88).

    In contrast to the positive relationship between postheparin LPL activity and HDL, postheparin HL activity correlates inversely with low HDL-c levels in humans.89 Individuals with a functional mutation of the hepatic lipase gene have moderately raised HDL-c with enlarged TG-rich HDL particles.90 Mice with targeted disruption of the HL gene have a mild elevation of plasma HDL, which becomes more profound compared with wild-type mice when they are fed a high-fat and cholesterol-rich diet.91 Overexpression of HL in mice and rabbits results in marked reductions in HDL-c and reductions in HDL size (reviewed in reference 87). As is the case with LPL, HL possesses both a lipolytic function and a physiologically relevant nonlipolytic function, both of which play a role in mediating HDL metabolism (reviewed in reference 87).

    Endothelial lipase was discovered in 1999 as a new member of the lipase family that has considerable homology with both LPL and HL.92,93 EL is synthesized by endothelial cells, functions at the vascular endothelial surface, and has primarily phospholipase A1 activity, with relatively little triglyceride lipase activity compared with LPL and HL.94 EL hydrolyzes HDL more efficiently than the other lipoprotein fractions.94 There is convincing in vivo evidence that EL plays an important role in modulating HDL metabolism. Adenoviral vector-mediated overexpression of EL in mice resulted in significant reduction in HDL-c and apoA-I levels.92 This decrease in HDL and apoA-I was shown to be associated with the generation of smaller HDL particles, because of increased fractional catabolism, and to be clearly dose-dependent.95 Transgenic expression of human EL was also reported to be associated with reduced HDL-c levels.96 On the other hand, antibody inhibition of EL activity in mice significantly increased plasma HDL-c, phospholipids, and apoA-I levels and resulted in larger HDL particles.97 Gene knockout of EL in mice also results in increased HDL-c levels.96,98 As with HL and LPL, EL can bridge lipoproteins through binding to heparan sulfate proteoglycan99,100 and therefore may have physiologically relevant ligand-binding functions that influence HDL metabolism in vivo; however, lipolytic activity is required for the profound effects on HDL metabolism with overexpression.101 Of note, overexpression of EL also reduces levels of apoB-containing lipoproteins102 and the knockout mice have modest increases in levels of apoB-containing lipoproteins,103 so the action of EL may not be strictly limited to HDL in vivo. The physiological role of EL in humans and its effects on atherosclerosis remain to be determined. Variation in the EL gene may be associated with variation in HDL-c levels.104,105 In the apoE-deficient murine model of atherosclerosis, EL knockout mice have reduced atherosclerosis, indicating that EL may be a proatherogenic enzyme.103

    The relative or combined effects of HL and EL on HDL metabolism have yet to be established. Perhaps HL is most active in hydrolyzing the triglycerides in TG-rich HDL, whereas EL hydrolyzes predominantly HDL phospholipids. HL and EL may exert differential effects on HDL metabolism depending on the prevailing metabolic milieu and their effects may depend on the relative composition of HDL particles, as determined by other genetic and metabolic factors.

    The secretory phospholipase A2 (sPLA2) family consists of low-molecular-weight secreted phospholipases that exhibit sn-2 phospholipase activity,106 and some have the ability to hydrolyze HDL phospholipids and may be physiologically relevant in regulating HDL metabolism.107 The most-studied member of this family is sPLA2-group IIA, which is upregulated in acute and chronic inflammatory states.106 Transgenic overexpression of sPLA2-IIA in mice results in reduced levels of HDL-c, reduction in size of HDL, and more rapid catabolism of the HDL particles.108 Thus, sPLA2-IIA may play a role in reducing HDL-c levels in the setting of inflammation.

    Receptor-Mediated Selective Uptake of HDL-c and Catabolism of HDL Apolipoproteins by Tissues

    The kidney, liver, and steroidogenic tissues are major sites of HDL catabolism. Clearance of HDL may be either by the selective removal of cholesterol (mainly) and other lipids from the particle, without uptake of the whole particle, a process termed selective cholesterol uptake, or by endocytic uptake and degradation of the whole particle, including apoA-I, a process we will refer to as holoparticle HDL uptake.

    SR-BI, a member of the scavenger receptor superfamily of proteins, binds to a variety of ligands including high-affinity binding to HDL and has been shown to play a major role in HDL selective lipid uptake by tissues.43,109 Catabolism of all major HDL lipids can occur via SR-BI, with the highest uptake constants for CE and free cholesterol and lower uptake constants for phospholipids and triglycerides.110 The importance of SR-BI, which is highly expressed in liver, adrenal gland, and ovary, for HDL metabolism has been readily demonstrated in genetically altered animals (for a recent comprehensive review see Trigatti et al43). Lipid is transferred from the HDL core to tissues by a 2-step process (binding of HDL to the receptor followed by diffusion of lipid into the plasma membrane) without concomitant degradation of the particle. Small, dense HDL particles generated by interaction with SR-BI are rapidly remodeled in an ill-defined fashion in plasma to form HDL2 particles, thereby in large part protecting them from rapid clearance.111 There is, however, some shedding of apoA-I in the process, as well as possibly increased subsequent holoparticle uptake of the lipid-depleted HDL particles. Mice deficient in SR-BI have elevated levels of HDL-c but not of plasma apoA-I, consistent with a defect in selective uptake of HDL cholesterol. Overexpression of SR-BI, on the other hand, results in reduced plasma concentrations not only of HDL-c but also of apoA-I, because of accelerated renal and hepatic clearance of cholesterol-depleted HDL particles after their interaction with SR-BI.112eC114 In contrast to the accelerated catabolism of TG-enriched HDL promoted by HL as discussed above, elevated HDL TG content diminishes rather than enhances the capacity of HDL to deliver CEs via SR-BI.115 ApoA-I conformation and particle size also influence HDL interaction with SR-BI,116,117 as does the remodeling of HDL by CETP and hepatic lipase.118,119

    Holoparticle HDL and apoA-I endocytosis and lysosomal degradation are known to occur in both liver and kidney.120,121 In the kidney, lipid-poor apoA-I is probably filtered at the glomerulus and catabolized by the renal tubular cell via the cubilin/megalin system.122eC124 Receptors that endocytose whole HDL particles or apoA-I protein and lead to degradation in the liver have been somewhat elusive. For further discussion of receptors potentially leading to HDL holoparticle or apoA-I uptake and degradation, including cubilin, megalin, AI-BP, HB1, HB2, HBP, ectopic -chain of ATP synthase, please see the online data supplement.

    Mechanism of HDL Lowering in Hypertriglyceridemic and Insulin Resistant States

    Although low HDL is common, single gene disorders (such as those involving apoA-I, LCAT, and ABCA1 mutations) causing low HDL are rare. There is growing appreciation that insulin resistance appears to underlie the majority of cases of low HDL in humans,3,4 although the exact prevalence of the insulin resistance syndrome among those with low HDL-c in the general population has not yet been clearly established. Approximately one quarter of the North American population has evidence of insulin resistance and the prevalence is rising.125 Hypertriglyceridemia, low plasma concentrations of HDL-c, and qualitative changes in LDL comprise the typical dyslipidemia of insulin resistant states.126 In fact, a high TG:HDL-c ratio may be the single most characteristic feature of the insulin resistant syndrome, even more highly predictive of insulin resistance than the presence of abdominal obesity.127 The hypertriglyceridemia of insulin resistance is primarily attributable to increased hepatic production of VLDL particles, with an important additional component of postprandial hyperlipidemia.128,129 In humans, there are significant negative correlations between fasting and postprandial plasma TG levels and HDL-c and apoA-I concentrations,3 suggesting a close link between TG and HDL metabolism.

    In vivo lipoprotein turnover studies in humans have shown that hypertriglyceridemic individuals with low HDL-c have significantly increased catabolic rates of apoA-I but no major reduction in apoA-I production rates, in comparison with normolipidemic subjects.8,10,11,130,131 Similarly, in individuals with impaired glucose tolerance and type 2 diabetes, hypertriglyceridemia was shown to be associated with increased HDL catabolism.132,133 A number of potential mechanisms, many already discussed above, could explain the inverse relationship between the hypertriglyceridemia of insulin resistant states and increased HDL catabolism leading to low plasma HDL-c and apoA-I concentrations. One is a reduction in LPL activity, which would have the effect of impairing the maturation of HDL particles. The normal insulin-mediated stimulation of LPL activity, such as occurs in the postprandial state, has been shown to be blunted in insulin resistance.134 In type 2 diabetes, particularly when glycemic control is poor and in patients who are relatively insulin deficient, LPL activity is reduced.135

    Another theory accounting for increased HDL catabolism in hypertriglyceridemic and insulin resistant states relates to the enhanced CETP-mediated hetero-exchange of TG and CE between TRL and HDL.136,137 Fasting triglycerides need only be mildly elevated for there to be significant TG enrichment of HDL.138 The combination of increased CETP-mediated TG enrichment of HDL coupled with the elevated HL activity that occurs in insulin resistance139 enhances the remodelling of HDL in the circulation, with shedding of lipid-poor apoA-I, the generation of remnant HDL particles, and increased HDL catabolism.137 In vivo stimulation of intravascular lipolysis by heparin administration in hypertriglyceridemic but not normolipidemic humans results in the production of small, dense, atypical HDL particles,140 and this phenomenon can be attenuated by the effective treatment of these patients’ hypertriglyceridemia with TG-lowering fibrate therapy.141 The clearance of apoA-I associated with HDL that had been TG-enriched in vivo was significantly faster than non-TG enriched HDL isolated from the same individuals in the fasted state.142

    A number of lines of evidence suggest that lipid compositional change of HDL alone is not sufficient to destabilize the particle and that lipolytic modification of TG-rich HDL is necessary to promote rapid catabolism. In isolated perfused rabbit kidneys, the renal clearance of HDL apoA-I was not significantly enhanced unless the TG-enriched HDL was pretreated with partially purified lipases.11 Lipolytically-modified TG-enriched HDL are cleared faster from the circulation than TG-rich HDL that have not undergone lipolytic modification by hepatic lipase.143 In the rabbit, an animal model naturally somewhat deficient in hepatic lipase, TG-enrichment of HDL alone is not sufficient to enhance HDL particle clearance.144 However, both ex vivo145 as well as in vivo146 lipolysis of TG-enriched HDL by HL enhances the clearance of HDL-associated apoA-I from the circulation. Hepatic lipase acting on TG-enriched but not TG-poor HDL2 induces the shedding of lipid-poor apoA-I, which is then more rapidly cleared from the circulation, presumably primarily by the kidneys. HL action also results in the formation of a residual -migrating HDL particle named remnant HDL2, which has one less molecule of apoA-I, 60% less TG, and 15% less phospholipids and is distinctly different from HDL3.147 Remnant HDL have conformational changes of apoA-I, changes in the fluidity of the lipid environment, and a high free fatty acid concentration.143,147 which may account for their more avid binding to receptors and more rapid catabolism.

    HL activity is elevated in insulin-resistant states such as abdominal obesity and type 2 diabetes, conditions that are also commonly associated with high rates of CETP-mediated lipid exchange between HDL and TRL, and HL activity is correlated with the low HDL levels in these conditions.139,148eC151 HL activity, however, unlike LPL, is not upregulated in a clear-cut fashion by insulin: although studies in patients with type 2 diabetes have shown positive correlations between HL activity and hyperinsulinemia,148 others have shown that short-term insulin infusion causes a decline in HL activity.152 It is more likely that insulin resistance at the liver induces the increase in HL activity. Consistent with this idea, studies in normal and diabetic rats have shown that increases in liver HL activity are induced by chronic, but not acute, insulin administration.153 HL mRNA, protein, and plasma postheparin TG lipase activity is increased in fructose-fed Syrian golden hamsters, an animal model of insulin resistance, and reduced with rosiglitazone treatment, an insulin sensitizer with PPAR agonist activity.154

    The interaction between hepatic lipase and HDL that is TG-enriched appears to be one important mechanism (not necessarily the exclusive mechanism) whereby HDL catabolism is enhanced in insulin resistant states. Other factors associated with the insulin resistant state, such as a chronic inflammatory mileau,155 may also contribute to HDL-c lowering. For example, endothelial lipase is upregulated by cytokines, providing a potential mechanistic link between inflammation and reduced HDL-c levels.156 Future studies are required to elucidate more precisely the perturbations and contributions of the various HDL metabolic pathways in the lowering of HDL plasma concentrations in insulin resistance.

    Conclusions and Future Directions

    In summary, the metabolism of HDL involves a complex interplay of factors regulating the synthesis, intravascular remodelling, and catabolism of HDL. In contrast to apoB-lipoprotein metabolism, the different components of HDL are largely assembled extracellularly and are subject to continuous dynamic exchange, transfer, and lipolysis within the plasma compartment. Even the catabolism of HDL occurs as separate components and generally not as a single HDL particle. Although HDL promotes RCT from extrahepatic tissues to liver, the liver itself is a major source of lipidation of nascent HDL. Furthermore, the tiny pool of macrophage cholesterol that is effluxed to HDL and returned to the liver is probably most important for protection from atherosclerosis, hence the new term macrophage RCT. HDL metabolism and macrophage RCT are major targets for the development of new therapies for atherosclerotic cardiovascular disease. In addition to RCT, HDL may play a role in processes such as inhibition of endothelial inflammation, promotion of endothelial NO and prostacyclin production, and the sequestration and transport of amyloidogenic proteins, oxidized lipids, and lipids derived from exogenous pathogens. These emerging functions of HDL may affect processes such as atherogenesis. Progress in our understanding of HDL metabolism, RCT, and other HDL functions will increase the likelihood of finding these therapies and proving that they are effective in humans.

    Acknowledgments

    Dr Lewis is the recipient of a Career Investigator Award from the Heart and Stroke Foundation of Canada and a Canada Research Chair in Diabetes (http://www.chairs.gc.ca/). His research is supported by the Heart and Stroke Foundation of Ontario, the Canadian Institutes for Health Research, and the Canadian Diabetes Association. Dr Rader is a recipient of an American Heart Association Established Investigator Award, a Burroughs Wellcome Fund Clinical Scientist Award in Translational Research, and a Doris Duke Distinguished Clinical Investigator Award and is also funded by National Institutes of Health grants from the National Heart, Lung, and Blood Institute, National Institute of Diabetes and Digestive and Kidney Diseases, and National Center for Research Resources, including HL55323, HL70128, HL22633, and M01 RR00040.

    Dr Gary Lewis has acted as a consultant to the following pharmaceutical companies that are involved in the lipid-lowering and diabetes therapeutics fields: Pfizer, Merck Frosst, Astra Zeneca, Eli Lilly, Fournier Pharma, Bristol Myers Squibb.

    References

    Boden WE. High-density lipoprotein cholesterol as an independent risk factor in cardiovascular disease: assessing the data from Framingham to the Veterans Affairs HigheCDensity Lipoprotein Intervention Trial. Am J Cardiol. 2000; 86: 19LeC22L.

    Assmann G, Gotto AM Jr. HDL cholesterol and protective factors in atherosclerosis. Circulation. 2004; 109: III8eC14.

    Despres JP, Lemieux I, Dagenais GR, Cantin B, Lamarche B. HDL-cholesterol as a marker of coronary heart disease risk: the Quebec cardiovascular study. Atherosclerosis. 2000; 153: 263eC272.

    Rubins HB, Robins SJ, Collins D, Fye CL, Anderson JW, Elam MB, Faas FH, Linares E, Schaefer EJ, Schectman G, Wilt TJ, Wittes J. Gemfibrozil for the secondary prevention of coronary heart disease in men with low levels of high-density lipoprotein cholesterol. Veterans Affairs High-Density Lipoprotein Cholesterol Intervention Trial Study Group. N Engl J Med. 1999; 341: 410eC418.

    Rubin EM, Krauss RM, Spangler EA, Verstuyft JG, Clift SM. Inhibition of early atherogenesis in transgenic mice by human apolipoprotein AI. Nature. 1991; 353: 265eC267.

    Srivastava RA, Srivastava N. High density lipoprotein, apolipoprotein A-I, and coronary artery disease. Mol Cell Biochem. 2000; 209: 131eC144.

    Malik S. Transcriptional regulation of the apolipoprotein AI gene. Front Biosci. 2003; 8: d360eCd368.

    Brinton EA, Eisenberg S, Breslow JL. Increased apo A-I and apo A-II fractional catabolic rate in patients with low high density lipoprotein-cholesterol levels with or without hypertriglyceridemia. J Clin Invest. 1991; 87: 536eC544.

    Brinton EA, Eisenberg S, Breslow JL. Elevated high density lipoprotein cholesterol levels correlate with decreased apolipoprotein A-I and A-II fractional catabolic rate in women. J Clin Invest. 1989; 84: 262eC269.

    Brinton EA, Eisenberg S, Breslow JL. Human HDL cholesterol levels are determined by apoA-I fractional catabolic rate, which correlates inversely with estimates of HDL particle size. Effects of gender, hepatic and lipoprotein lipases, triglyceride and insulin levels, and body fat distribution. Arterioscler Thromb. 1994; 14: 707eC720.

    Horowitz BS, Goldberg IJ, Merab J, Vanni TM, Ramakrishnan R, Ginsberg HN. Increased plasma and renal clearance of an exchangeable pool of apolipoprotein A-I in subjects with low levels of high density lipoprotein cholesterol. J Clin Invest. 1993; 91: 1743eC1752.

    Velez-Carrasco W, Lichtenstein AH, Welty FK, Li Z, Lamon-Fava S, Dolnikowski GG, Schaefer EJ. Dietary restriction of saturated fat and cholesterol decreases HDL ApoA-I secretion. Arterioscler Thromb Vasc Biol. 1999; 19: 918eC924.

    Kunitake ST, La Sala KJ, Kane JP. Apolipoprotein A-I-containing lipoproteins with pre-beta electrophoretic mobility. J Lipid Res. 1985; 26: 549eC555.

    Hennessy LK, Kunitake ST, Kane JP. Apolipoprotein A-I-containing lipoproteins, with or without apolipoprotein A-II, as progenitors of pre-beta high-density lipoprotein particles. Biochemistry. 1993; 32: 5759eC5765.

    Fielding CJ, Fielding PE. Molecular physiology of reverse cholesterol transport. J Lipid Res. 1995; 36: 211eC228.

    Yancey PG, Bortnick AE, Kellner-Weibel G, Llera-Moya M, Phillips MC, Rothblat GH. Importance of different pathways of cellular cholesterol efflux. Arterioscler Thromb Vasc Biol. 2003; 23: 712eC719.

    Marcil M, Brooks-Wilson A, Clee SM, Roomp K, Zhang LH, Yu L, Collins JA, van Dam M, Molhuizen HO, Loubster O, Ouellette BF, Sensen CW, Fichter K, Mott S, Denis M, Boucher B, Pimstone S, Genest J, Kastelein JJ, Hayden MR. Mutations in the ABC1 gene in familial HDL deficiency with defective cholesterol efflux. Lancet. 1999; 354: 1341eC1346.

    Brooks-Wilson A, Marcil M, Clee SM, Zhang LH, Roomp K, van Dam M, Yu L, Brewer C, Collins JA, Molhuizen HO, Loubser O, Ouelette BF, Fichter K, Ashbourne-Excoffon KJ, Sensen CW, Scherer S, Mott S, Denis M, Martindale D, Frohlich J, Morgan K, Koop B, Pimstone S, Kastelein JJ, Hayden MR. Mutations in ABC1 in Tangier disease and familial high-density lipoprotein deficiency. Nat Genet. 1999; 22: 336eC345.

    Clee SM, Kastelein JJ, van Dam M, Marcil M, Roomp K, Zwarts KY, Collins JA, Roelants R, Tamasawa N, Stulc T, Suda T, Ceska R, Boucher B, Rondeau C, DeSouich C, Brooks-Wilson A, Molhuizen HO, Frohlich J, Genest J Jr, Hayden MR. Age and residual cholesterol efflux affect HDL cholesterol levels and coronary artery disease in ABCA1 heterozygotes. J Clin Invest. 2000; 106: 1263eC1270.

    Hovingh GK, Van Wijland MJ, Brownlie A, Bisoendial RJ, Hayden MR, Kastelein JJ, Groen AK. The role of the ABCA1 transporter and cholesterol efflux in familial hypoalphalipoproteinemia. J Lipid Res. 2003; 44: 1251eC1255.

    Mott S, Yu L, Marcil M, Boucher B, Rondeau C, Genest J Jr. Decreased cellular cholesterol efflux is a common cause of familial hypoalphalipoproteinemia: role of the ABCA1 gene mutations. Atherosclerosis. 2000; 152: 457eC468.

    Alenezi MY, Marcil M, Blank D, Sherman M, Genest J Jr. Is the decreased high-density lipoprotein cholesterol in the metabolic syndrome due to cellular lipid efflux defect J Clin Endocrinol Metab. 2004; 89: 761eC764.

    Woll PS, Hanson NQ, Tsai MY. Absence of ABCA1 mutations in individuals with low serum HDL-cholesterol. Clin Chem. 2003; 49: 521eC522.

    Kakko S, Kelloniemi J, von Rohr P, Hoeschele I, Tamminen M, Brousseau ME, Kesaniemi YA, Savolainen MJ. ATP-binding cassette transporter A1 locus is not a major determinant of HDL-C levels in a population at high risk for coronary heart disease. Atherosclerosis. 2003; 166: 285eC290.

    Cohen JC, Kiss RS, Pertsemlidis A, Marcel YL, McPherson R, Hobbs HH. Multiple rare alleles contribute to low plasma levels of HDL cholesterol. Science. 2004; 305: 869eC872.

    Frikke-Schmidt R, Nordestgaard BG, Jensen GB, Tybjaerg-Hansen A. Genetic variation in ABC transporter A1 contributes to HDL cholesterol in the general population. J Clin Invest. 2004; 114: 1343eC1353.

    Rader DJ. Regulation of reverse cholesterol transport and clinical implications. Am J Cardiol. 2003; 92: 42eC49.

    Aiello RJ, Brees D, Francone OL. ABCA1-deficient mice: insights into the role of monocyte lipid efflux in HDL formation and inflammation. Arterioscler Thromb Vasc Biol. 2003; 23: 972eC980.

    Aiello RJ, Brees D, Bourassa PA, Royer L, Lindsey S, Coskran T, Haghpassand M, Francone OL. Increased atherosclerosis in hyperlipidemic mice with inactivation of ABCA1 in macrophages. Arterioscler Thromb Vasc Biol. 2002; 22: 630eC637.

    Van Eck M, Bos IS, Kaminski WE, Orso E, Rothe G, Twisk J, Bottcher A, Van Amersfoort ES, Christiansen-Weber TA, Fung-Leung WP, van Berkel TJ, Schmitz G. Leukocyte ABCA1 controls susceptibility to atherosclerosis and macrophage recruitment into tissues. Proc Natl Acad Sci U S A. 2002; 99: 6298eC6303.

    Haghpassand M, Bourassa PA, Francone OL, Aiello RJ. Monocyte/macrophage expression of ABCA1 has minimal contribution to plasma HDL levels. J Clin Invest. 2001; 108: 1315eC1320.

    Lee JY, Parks JS. ATP-binding cassette transporter AI and its role in HDL formation. Curr Opin Lipidol. 2005; 16: 19eC25.

    Basso F, Freeman L, Knapper CL, Remaley A, Stonik J, Neufeld EB, Tansey T, Amar MJ, Fruchart-Najib J, Duverger N, Santamarina-Fojo S, Brewer HB Jr. Role of the hepatic ABCA1 transporter in modulating intrahepatic cholesterol and plasma HDL cholesterol concentrations. J Lipid Res. 2003; 44: 296eC302.

    Wellington CL, Brunham LR, Zhou S, Singaraja RR, Visscher H, Gelfer A, Ross C, James E, Liu G, Huber MT, Yang YZ, Parks RJ, Groen A, Fruchart-Najib J, Hayden MR. Alterations of plasma lipids in mice via adenoviral-mediated hepatic overexpression of human ABCA1. J Lipid Res. 2003; 44: 1470eC1480.

    Vaisman BL, Lambert G, Amar M, Joyce C, Ito T, Shamburek RD, Cain WJ, Fruchart-Najib J, Neufeld ED, Remaley AT, Brewer HB, Jr., Santamarina-Fojo S. ABCA1 overexpression leads to hyperalphalipoproteinemia and increased biliary cholesterol excretion in transgenic mice. J Clin Invest. 2001; 108: 303eC309.

    Brewer HB Jr, Santamarina-Fojo S. New insights into the role of the adenosine triphosphate-binding cassette transporters in high-density lipoprotein metabolism and reverse cholesterol transport. Am J Cardiol. 2003; 91: 3EeC11E.

    Singaraja RR, Fievet C, Castro G, James ER, Hennuyer N, Clee SM, Bissada N, Choy JC, Fruchart JC, McManus BM, Staels B, Hayden MR. Increased ABCA1 activity protects against atherosclerosis. J Clin Invest. 2002; 110: 35eC42.

    Wang N, Lan D, Chen W, Matsuura F, Tall AR. ATP-binding cassette transporters G1 and G4 mediate cellular cholesterol efflux to high-density lipoproteins. Proc Natl Acad Sci U S A. 2004; 101: 9774eC9779.

    Kennedy MA, Berrera GC, Nakamura K, Baldan A, Tarr P, Fishbein MC, Frank J, Francone OL, Edwards PA. ABCG1 has a critical role in mediating cholesterol efflux to HDL and preventing ceullular lipid acumulation. Cell Metabolism. 2005; 1: 121eC131.

    Rothblat GH, Llera-Moya M, Atger V, Kellner-Weibel G, Williams DL, Phillips MC. Cell cholesterol efflux: integration of old and new observations provides new insights. J Lipid Res. 1999; 40: 781eC796.

    Rigotti A, Trigatti BL, Penman M, Rayburn H, Herz J, Krieger M. A targeted mutation in the murine gene encoding the high density lipoprotein (HDL) receptor scavenger receptor class B type I reveals its key role in HDL metabolism. Proc Natl Acad Sci U S A. 1997; 94: 12610eC12615.

    Trigatti B, Rigotti A. Scavenger receptor class B type I (SR-BI) and high-density lipoprotein metabolism: recent lessons from genetically manipulated mice. Int J Tissue React. 2000; 22: 29eC37.

    Trigatti B, Covey S, Rizvi A. Scavenger receptor class B type I in high-density lipoprotein metabolism, atherosclerosis and heart disease: lessons from gene-targeted mice. Biochem Soc Trans. 2004; 32: 116eC120.

    Van Eck M, Bos IS, Hildebrand RB, Van Rij BT, van Berkel TJ. Dual role for scavenger receptor class B, type I on bone marrow-derived cells in atherosclerotic lesion development. Am J Pathol. 2004; 165: 785eC794.

    Wang N, Tall AR. Regulation and mechanisms of ATP-binding cassette transporter A1-mediated cellular cholesterol efflux. Arterioscler Thromb Vasc Biol. 2003; 23: 1178eC1184.

    Tontonoz P, Mangelsdorf DJ. Liver X receptor signaling pathways in cardiovascular disease. Mol Endocrinol. 2003; 17: 985eC993.

    Lund EG, Menke JG, Sparrow CP. Liver X receptor agonists as potential therapeutic agents for dyslipidemia and atherosclerosis. Arterioscler Thromb Vasc Biol. 2003; 23: 1169eC1177.

    Joseph SB, McKilligin E, Pei L, Watson MA, Collins AR, Laffitte BA, Chen M, Noh G, Goodman J, Hagger GN, Tran J, Tippin TK, Wang X, Lusis AJ, Hsueh WA, Law RE, Collins JL, Willson TM, Tontonoz P. Synthetic LXR ligand inhibits the development of atherosclerosis in mice. Proc Natl Acad Sci U S A. 2002; 99: 7604eC7609.

    Levin N, Bischoff ED, Daige CL, Thomas D, Vu CT, Heyman RA, Tangirala RK, Schulman IG. Macrophage Liver X Receptor Is Required for Antiatherogenic Activity of LXR Agonists. Arterioscler Thromb Vasc Biol. 2004.

    Chinetti G, Lestavel S, Bocher V, Remaley AT, Neve B, Torra IP, Teissier E, Minnich A, Jaye M, Duverger N, Brewer HB, Fruchart JC, Clavey V, Staels B. PPAR-alpha and PPAR-gamma activators induce cholesterol removal from human macrophage foam cells through stimulation of the ABCA1 pathway. Nat Med. 2001; 7: 53eC58.

    Osono Y, Woollett LA, Marotti KR, Melchior GW, Dietschy JM. Centripetal cholesterol flux from extrahepatic organs to the liver is independent of the concentration of high density lipoprotein-cholesterol in plasma. Proc Natl Acad Sci U S A. 1996; 93: 4114eC4119.

    Jolley CD, Woollett LA, Turley SD, Dietschy JM. Centripetal cholesterol flux to the liver is dictated by events in the peripheral organs and not by the plasma high density lipoprotein or apolipoprotein A-I concentration. J Lipid Res. 1998; 39: 2143eC2149.

    Zhang Y, Zanotti I, Reilly MP, Glick JM, Rothblat GH, Rader DJ. Overexpression of apolipoprotein A-I promotes reverse transport of cholesterol from macrophages to feces in vivo. Circulation. 2003; 108: 661eC663.

    Jonas A. Lecithin cholesterol acyltransferase. Biochim Biophys Acta. 2000; 1529: 245eC256.

    Kuivenhoven JA, Pritchard H, Hill J, Frohlich J, Assmann G, Kastelein J. The molecular pathology of lecithin:cholesterol acyltransferase (LCAT) deficiency syndromes. J Lipid Res. 1997; 38: 191eC205.

    Rader DJ, Ikewaki K, Duverger N, Schmidt H, Pritchard H, Frohlich J, Clerc M, Dumon MF, Fairwell T, Zech L. Markedly accelerated catabolism of apolipoprotein A-II (ApoA-II) and high density lipoproteins containing ApoA-II in classic lecithin: cholesterol acyltransferase deficiency and fish-eye disease. J Clin Invest. 1994; 93: 321eC330.

    Sakai N, Vaisman BL, Koch CA, Hoyt RF Jr, Meyn SM, Talley GD, Paiz JA, Brewer HB Jr, Santamarina-Fojo S. Targeted disruption of the mouse lecithin:cholesterol acyltransferase (LCAT) gene. Generation of a new animal model for human LCAT deficiency. J Biol Chem. 1997; 272: 7506eC7510.

    Ng DS. Insight into the Role of LCAT from Mouse Models. Rev Endocr Metab Disord. 2004; 5: 311eC318.

    Schwartz CC, VandenBroek JM, Cooper PS. Lipoprotein cholesteryl ester production, transfer, and output in vivo in humans. J Lipid Res. 2004; 45: 1594eC1607.

    Barter PJ, Brewer HB Jr, Chapman MJ, Hennekens CH, Rader DJ, Tall AR. Cholesteryl ester transfer protein: a novel target for raising HDL and inhibiting atherosclerosis. Arterioscler Thromb Vasc Biol. 2003; 23: 160eC167.

    Guerin M, Le Goff W, Lassel TS, van Tol A, Steiner G, Chapman MJ. Atherogenic role of elevated CE transfer from HDL to VLDL(1) and dense LDL in type 2 diabetes : impact of the degree of triglyceridemia. Arterioscler Thromb Vasc Biol. 2001; 21: 282eC288.

    Guerin M, Egger P, Soudant C, Le Goff W, van Tol A, Dupuis R, Chapman MJ. Cholesteryl ester flux from HDL to VLDL-1 is preferentially enhanced in type IIB hyperlipidemia in the postprandial state. J Lipid Res. 2002; 43: 1652eC1660.

    Paromov VM, Morton RE. Lipid transfer inhibitor protein defines the participation of high density lipoprotein subfractions in lipid transfer reactions mediated by cholesterol ester transfer protein (CETP). J Biol Chem. 2003; 278: 40859eC40866.

    Inazu A, Brown ML, Hesler CB, Agellon LB, Koizumi J, Takata K, Maruhama Y, Mabuchi H, Tall AR. Increased high-density lipoprotein levels caused by a common cholesteryl-ester transfer protein gene mutation. N Engl J Med. 1990; 323: 1234eC1238.

    Ikewaki K, Nishiwaki M, Sakamoto T, Ishikawa T, Fairwell T, Zech LA, Nagano M, Nakamura H, Brewer HB Jr, Rader DJ. Increased catabolic rate of low density lipoproteins in humans with cholesteryl ester transfer protein deficiency. J Clin Invest. 1995; 96: 1573eC1581.

    Agellon LB, Walsh A, Hayek T, Moulin P, Jiang XC, Shelanski SA, Breslow JL, Tall AR. Reduced high density lipoprotein cholesterol in human cholesteryl ester transfer protein transgenic mice. J Biol Chem. 1991; 266: 10796eC10801.

    Okamoto H, Yonemori F, Wakitani K, Minowa T, Maeda K, Shinkai H. A cholesteryl ester transfer protein inhibitor attenuates atherosclerosis in rabbits. Nature. 2000; 406: 203eC207.

    Morehouse LA, Sugarman ED, Bourassa P-A, and Milici AJ. HDL elevation by the CETP-inhibitor Torcetrapib prevents aortic atherosclerosis in rabbits. Circulation. 110 (III-243). 2004.

    Brousseau ME, Schaefer EJ, Wolfe ML, Bloedon LT, Digenio AG, Clark RW, Mancuso JP, Rader DJ. Effects of an inhibitor of cholesteryl ester transfer protein on HDL cholesterol. N Engl J Med. 2004; 350: 1505eC1515.

    Clark RW, Sutfin TA, Ruggeri RB, Willauer AT, Sugarman ED, Magnus-Aryitey G, Cosgrove PG, Sand TM, Wester RT, Williams JA, Perlman ME, Bamberger MJ. Raising high-density lipoprotein in humans through inhibition of cholesteryl ester transfer protein: an initial multidose study of torcetrapib. Arterioscler Thromb Vasc Biol. 2004; 24: 490eC497.

    de Grooth GJ, Kuivenhoven JA, Stalenhoef AF, de Graaf J, Zwinderman AH, Posma JL, van Tol A, Kastelein JJ. Efficacy and safety of a novel cholesteryl ester transfer protein inhibitor, JTT-705, in humans: a randomized phase II dose-response study. Circulation. 2002; 105: 2159eC2165.

    Huuskonen J, Olkkonen VM, Jauhiainen M, Ehnholm C. The impact of phospholipid transfer protein (PLTP) on HDL metabolism. Atherosclerosis. 2001; 155: 269eC281.

    Jiang XC, Bruce C, Mar J, Lin M, Ji Y, Francone OL, Tall AR. Targeted mutation of plasma phospholipid transfer protein gene markedly reduces high-density lipoprotein levels. J Clin Invest. 1999; 103: 907eC914.

    Jiang X, Francone OL, Bruce C, Milne R, Mar J, Walsh A, Breslow JL, Tall AR. Increased prebeta-high density lipoprotein, apolipoprotein AI, and phospholipid in mice expressing the human phospholipid transfer protein and human apolipoprotein AI transgenes. J Clin Invest. 1996; 98: 2373eC2380.

    Rye K-A, Jauhiainen M, Barter PJ, Ehnholm C. Triglyceride-enrichment of high density lipoproteins enhances their remodelling by phospholipid transfer protein. J Lipid Res. 1998; 39: 613eC622.

    Settasatian N, Duong M, Curtiss LK, Ehnholm C, Jauhiainen M, Huuskonen J, Rye KA. The mechanism of the remodeling of high density lipoproteins by phospholipid transfer protein. J Biol Chem. 2001; 276: 26898eC26905.

    Merkel M, Eckel RH, Goldberg IJ. Lipoprotein lipase: genetics, lipid uptake, and regulation. J Lipid Res. 2002; 43: 1997eC2006.

    Kaser S, Sandhofer A, Holzl B, Gander R, Ebenbichler CF, Paulweber B, Patsch JR. Phospholipid and cholesteryl ester transfer are increased in lipoprotein lipase deficiency. J Intern Med. 2003; 253: 208eC216.

    Tornvall P, Olivecrona G, Karpe F, Hamsten A, Olivecrona T. Lipoprotein lipase mass and activity in plasma and their increase after heparin are separate parameters with different relations to plasma lipoproteins. Arterioscler Thromb Vasc Biol. 1995; 15: 1086eC1093.

    Weinstock PH, Bisgaier CL, Aalto-Setala K, Radner H, Ramakrishnan R, Levak-Frank S, Essenburg AD, Zechner R, Breslow JL. Severe hypertriglyceridemia, reduced high density lipoprotein, and neonatal death in lipoprotein lipase knockout mice. Mild hypertriglyceridemia with impaired very low density lipoprotein clearance in heterozygotes. J Clin Invest. 1995; 96: 2555eC2568.

    Clee SM, Zhang H, Bissada N, Miao L, Ehrenborg E, Benlian P, Shen GX, Angel A, LeBoeuf RC, Hayden MR. Relationship between lipoprotein lipase and high density lipoprotein cholesterol in mice: modulation by cholesteryl ester transfer protein and dietary status. J Lipid Res. 1997; 38: 2079eC2089.

    Tsutsumi K, Inoue Y, Shima A, Iwasaki K, Kawamura M, Murase T. The novel compound NO-1886 increases lipoprotein lipase activity with resulting elevation of high density lipoprotein cholesterol, and long-term administration inhibits atherogenesis in the coronary arteries of rats with experimental atherosclerosis. J Clin Invest. 1993; 92: 411eC417.

    Goldberg IJ, Blaner WS, Vanni TM, Moukides M, Ramakrishnan R. Role of lipoprotein lipase in the regulation of high density lipoprotein apolipoprotein metabolism. Studies in normal and lipoprotein lipase-inhibited monkeys. J Clin Invest. 1990; 86: 463eC473.

    Rinninger F, Kaiser T, Mann WA, Meyer N, Greten H, Beisiegel U. Lipoprotein lipase mediates an increase in the selective uptake of high density lipoprotein-associated cholesteryl esters by hepatic cells in culture. J Lipid Res. 1998; 39: 1335eC1348.

    Perret B, Mabile L, Martinez L, Terce F, Barbaras R, Collet X. Hepatic lipase: structure/function relationship, synthesis, and regulation. J Lipid Res. 2002; 43: 1163eC1169.

    Jansen H, Verhoeven AJ, Sijbrands EJ. Hepatic lipase: a pro- or anti-atherogenic protein J Lipid Res. 2002; 43: 1352eC1362.

    Santamarina-Fojo S, Gonzalez-Navarro H, Freeman L, Wagner E, Nong Z. Hepatic lipase, lipoprotein metabolism, and atherogenesis. Arterioscler Thromb Vasc Biol. 2004; 24: 1750eC1754.

    Barter PJ. Hugh sinclair lecture: the regulation and remodelling of HDL by plasma factors. Atheroscler Suppl. 2002; 3: 39eC47.

    Blades B, Vega GL, Grundy SM. Activities of lipoprotein lipase and hepatic triglyceride lipase in postheparin plasma of patients with low concentrations of HDL cholesterol. Arterioscler Thromb. 1993; 13: 1227eC1235.

    Connelly PW, Hegele RA. Hepatic lipase deficiency. Crit Rev Clin Lab Sci. 1998; 35: 547eC572.

    Homanics GE, de Silva HV, Osada J, Zhang SH, Wong H, Borensztajn J, Maeda N. Mild dyslipidemia in mice following targeted inactivation of the hepatic lipase gene. J Biol Chem. 1995; 270: 2974eC2980.

    Jaye M, Lynch KJ, Krawiec J, Marchadier D, Maugeais C, Doan K, South V, Amin D, Perrone M, Rader DJ. A novel endothelial-derived lipase that modulates HDL metabolism. Nat Genet. 1999; 21: 424eC428.

    Hirata K, Dichek HL, Cioffi JA, Choi SY, Leeper NJ, Quintana L, Kronmal GS, Cooper AD, Quertermous T. Cloning of a unique lipase from endothelial cells extends the lipase gene family. J Biol Chem. 1999; 274: 14170eC14175.

    McCoy MG, Sun GS, Marchadier D, Maugeais C, Glick JM, Rader DJ. Characterization of the lipolytic activity of endothelial lipase. J Lipid Res. 2002; 43: 921eC929.

    Maugeais C, Tietge UJ, Broedl UC, Marchadier D, Cain W, McCoy MG, Lund-Katz S, Glick JM, Rader DJ. Dose-dependent acceleration of high-density lipoprotein catabolism by endothelial lipase. Circulation. 2003; 108: 2121eC2126.

    Ishida T, Choi S, Kundu RK, Hirata K, Rubin EM, Cooper AD, Quertermous T. Endothelial lipase is a major determinant of HDL level. J Clin Invest. 2003; 111: 347eC355.

    Jin W, Millar JS, Broedl U, Glick JM, Rader DJ. Inhibition of endothelial lipase causes increased HDL cholesterol levels in vivo. J Clin Invest. 2003; 111: 357eC362.

    Ma K, Cilingiroglu M, Otvos JD, Ballantyne CM, Marian AJ, Chan L. Endothelial lipase is a major genetic determinant for high-density lipoprotein concentration, structure, and metabolism. Proc Natl Acad Sci U S A. 2003; 100: 2748eC2753.

    Fuki IV, Blanchard N, Jin W, Marchadier DH, Millar JS, Glick JM, Rader DJ. Endogenously produced endothelial lipase enhances binding and cellular processing of plasma lipoproteins via heparan sulfate proteoglycan-mediated pathway. J Biol Chem. 2003; 278: 34331eC34338.

    Strauss JG, Zimmermann R, Hrzenjak A, Zhou Y, Kratky D, Levak-Frank S, Kostner GM, Zechner R, Frank S. Endothelial cell-derived lipase mediates uptake and binding of high-density lipoprotein (HDL) particles and the selective uptake of HDL-associated cholesterol esters independent of its enzymic activity. Biochem J. 2002; 368: 69eC79.

    Broedl UC, Maugeais C, Marchadier D, Glick JM, Rader DJ. Effects of nonlipolytic ligand function of endothelial lipase on HDL metabolism in vivo. J Biol Chem. 2003.

    Broedl UC, Maugeais C, Millar JS, Jin W, Moore RE, Fuki IV, Marchadier D, Glick JM, Rader DJ. Endothelial lipase promotes the catabolism of ApoB-containing lipoproteins. Circ Res. 2004; 94: 1554eC1561.

    Ishida T, Choi SY, Kundu RK, Spin J, Yamashita T, Hirata K, Kojima Y, Yokoyama M, Cooper AD, Quertermous T. Endothelial Lipase Modulates Susceptibility to Atherosclerosis in Apolipoprotein-E-deficient Mice. J Biol Chem. 2004; 279: 45085eC45092.

    deLemos AS, Wolfe ML, Long CJ, Sivapackianathan R, Rader DJ. Identification of genetic variants in endothelial lipase in persons with elevated high-density lipoprotein cholesterol. Circulation. 2002; 106: 1321eC1326.

    Mank-Seymour AR, Durham KL, Thompson JF, Seymour AB, Milos PM. Association between single-nucleotide polymorphisms in the endothelial lipase (LIPG) gene and high-density lipoprotein cholesterol levels. Biochim Biophys Acta. 2004; 1636: 40eC46.

    Murakami M, Kudo I. New phospholipase A(2) isozymes with a potential role in atherosclerosis. Curr Opin Lipidol. 2003; 14: 431eC436.

    Jin W, Marchadier D, Rader DJ. Lipases and HDL metabolism. Trends Endocrinol Metab. 2002; 13: 174eC178.

    Tietge UJ, Maugeais C, Lund-Katz S, Grass D, deBeer FC, Rader DJ. Human secretory phospholipase A2 mediates decreased plasma levels of HDL cholesterol and apoA-I in response to inflammation in human apoA-I transgenic mice. Arterioscler Thromb Vasc Biol. 2002; 22: 1213eC1218.

    Trigatti BL, Krieger M, Rigotti A. Influence of the HDL receptor SR-BI on lipoprotein metabolism and atherosclerosis. Arterioscler Thromb Vasc Biol. 2003; 23: 1732eC1738.

    Thuahnai ST, Lund-Katz S, Williams DL, Phillips MC. Scavenger receptor class B, type I-mediated uptake of various lipids into cells. Influence of the nature of the donor particle interaction with the receptor. J Biol Chem. 2001; 276: 43801eC43808.

    Webb NR, Cai L, Ziemba KS, Yu J, Kindy MS, van der Westhuyzen DR, de Beer FC. The fate of HDL particles in vivo after SR-BI-mediated selective lipid uptake. J Lipid Res. 2002; 43: 1890eC1898.

    Kozarsky KF, Donahee MH, Rigotti A, Iqbal SN, Edelman ER, Krieger M. Overexpression of the HDL receptor SR-BI alters plasma HDL and bile cholesterol levels. Nature. 1997; 387: 414eC417.

    Wang N, Arai T, Ji Y, Rinninger F, Tall AR. Liver-specific overexpression of scavenger receptor BI decreases levels of very low density lipoprotein ApoB, low density lipoprotein ApoB, and high density lipoprotein in transgenic mice. J Biol Chem. 1998; 273: 32920eC32926.

    Ueda Y, Royer L, Gong E, Zhang J, Cooper PN, Francone O, Rubin EM. Lower plasma levels and accelerated clearance of high density lipoprotein (HDL) and non-HDL cholesterol in scavenger receptor class B type I transgenic mice. J Biol Chem. 1999; 274: 7165eC7171.

    Greene DJ, Skeggs JW, Morton RE. Elevated triglyceride content diminishes the capacity of high density lipoprotein to deliver cholesteryl esters via the scavenger receptor class B type I (SR-BI). J Biol Chem. 2001; 276: 4804eC4811.

    de Beer MC, Durbin DM, Cai L, Jonas A, de Beer FC, van der Westhuyzen DR. Apolipoprotein A-I conformation markedly influences HDL interaction with scavenger receptor BI. J Lipid Res. 2001; 42: 309eC313.

    Thuahnai ST, Lund-Katz S, Dhanasekaran P, Llera-Moya M, Connelly MA, Williams DL, Rothblat GH, Phillips MC. Scavenger receptor class B type I-mediated cholesteryl ester-selective uptake and efflux of unesterified cholesterol. Influence of high density lipoprotein size and structure. J Biol Chem. 2004; 279: 12448eC12455.

    Collet X, Tall AR, Serajuddin H, Guendouzi K, Royer L, Oliveira H, Barbaras R, Jiang XC, Francone OL. Remodeling of HDL by CETP in vivo and by CETP and hepatic lipase in vitro results in enhanced uptake of HDL CE by cells expressing scavenger receptor B-I. J Lipid Res. 1999; 40: 1185eC1193.

    Lambert G, Chase MB, Dugi K, Bensadoun A, Brewer HB, Santamarina-Fojo S. Hepatic lipase promotes the selective uptake of high density lipoprotein-cholesteryl esters via the scavenger receptor B1. J Lipid Res. 1999; 40: 1294eC1303.

    Fidge NH. High density lipoprotein receptors, binding proteins, and ligands. J Lipid Res. 1999; 40: 187eC201.

    Steinberg D. A docking receptor for HDL cholesterol esters. Science. 1996; 271: 460eC461.

    Hammad SM, Barth JL, Knaak C, Argraves WS. Megalin acts in concert with cubilin to mediate endocytosis of high density lipoproteins. J Biol Chem. 2000; 275: 12003eC12008.

    Hammad SM, Stefansson S, Twal WO, Drake CJ, Fleming P, Remaley A, Brewer HB, Argraves WS. Cubilin, the endocytic receptor for intrinsic factor-vitamin B(12) complex, mediates high-density lipoprotein holoparticle endocytosis. Proc Natl Acad Sci U S A. 1999; 96: 10158eC10163.

    Moestrup SK, Kozyraki R. Cubilin, a high-density lipoprotein receptor. Curr Opin Lipidol. 2000; 11: 133eC140.

    Ford ES, Giles WH, Dietz WH. Prevalence of the metabolic syndrome among US adults: findings from the third National Health and Nutrition Examination Survey. J Am Med Assoc. 2002; 287: 356eC359.

    Lewis GF, Steiner G. Hypertriglyceridemia and its metabolic consequences as a risk factor for atherosclerotic cardiovascular disease in non-insulin- dependent diabetes mellitus. [Review] [204 refs]. Diabetes Metab Rev. 1996; 12: 37eC56.

    Reaven G. Metabolic syndrome: pathophysiology and implications for management of cardiovascular disease. Circulation. 2002; 106: 286eC288.

    Adeli K, Taghibiglou C, Van Iderstine SC, Lewis GF. Mechanisms of hepatic very low-density lipoprotein overproduction in insulin resistance. Trends Cardiovasc Med. 2001; 11: 170eC176.

    Lewis GF, Carpentier A, Adeli K, Giacca A. Disordered fat storage and mobilization in the pathogenesis of insulin resistance and type 2 diabetes. Endocr Rev. 2002; 23: 201eC229.

    Vajo Z, Terry JG, Brinton EA. Increased intra-abdominal fat may lower HDL levels by increasing the fractional catabolic rate of Lp A-I in postmenopausal women. Atherosclerosis. 2002; 160: 495eC501.

    Pont F, Duvillard L, Florentin E, Gambert P, Verges B. High-density lipoprotein apolipoprotein A-I kinetics in obese insulin resistant patients. An in vivo stable isotope study. Int J Obes Relat Metab Disord. 2002; 26: 1151eC1158.

    Frenais R, Ouguerram K, Maugeais C, Mahot P, Maugere P, Krempf M, Magot T. High density lipoprotein apolipoprotein AI kinetics in NIDDM: a stable isotope study. Diabetologia. 1997; 40: 578eC583.

    Pietzsch J, Julius U, Nitzsche S, Hanefeld M. In vivo evidence for increased apolipoprotein A-I catabolism in subjects with impaired glucose tolerance. Diabetes. 1998; 47: 1928eC1934.

    Eckel RH, Yost TJ, Jensen DR. Alterations in lipoprotein lipase in insulin resistance. Int J Obes Relat Metab Disord. 1995; 19 Suppl 1: S16eCS21.

    Yost TJ, Froyd KK, Jensen DR, Eckel RH. Change in skeletal muscle lipoprotein lipase activity in response to insulin/glucose in non-insulin-dependent diabetes mellitus. Metabolism. 1995; 44: 786eC790.

    Lamarche B, Rashid S, Lewis GF. HDL metabolism in hypertriglyceridemic states. An overview. Clin Chim Acta. 1999; 286: 145eC161.

    Rashid S, Watanabe T, Sakaue T, Lewis GF. Mechanisms of HDL lowering in insulin resistant, hypertriglyceridemic states: the combined effect of HDL triglyceride enrichment and elevated hepatic lipase activity. Clin Biochem. 2003; 36: 421eC429.

    O’Meara NM, Lewis GF, Cabana VG, Iverius PH, Getz GS, Polonsky KS. Role of basal triglyceride and high density lipoprotein in determination of postprandial lipid and lipoprotein responses. J Clin Endocrinol Metab. 1992; 75: 465eC471.

    Deeb SS, Zambon A, Carr MC, Ayyobi AF, Brunzell JD. Hepatic lipase and dyslipidemia: interactions among genetic variants, obesity, gender, and diet. J Lipid Res. 2003; 44: 1279eC1286.

    O’Meara NM, Cabana VG, Lukens JR, Loharikar B, Forte TM, Polonsky KS, Getz GS. Heparin-induced lipolysis in hypertriglyceridemic subjects results in the formation of atypical HDL particles. J Lipid Res. 1994; 35: 2178eC2190.

    Lewis GF, Uffelman KD, Lamarche B, Cabana VG, Getz GS. Production of small HDL particles after stimulation of in vivo lipolysis in hypertriglyceridemic individuals. Studies before and after triglyceride-lowering therapy. Metabolism. 1998; 47: 234eC242.

    Lamarche B, Uffelman KD, Carpentier A, Cohn JS, Steiner G, Barrett PHR, Lewis GF. Triglyceride-enrichment of HDL enhances the in vivo metabolic clearance of HDL-apo A-1 in healthy men. J Clin Invest. 1999; 103: 1191eC1199.

    Lewis GF, Lamarche B, Uffelman KD, Heatherington AC, Szeto LW, Honig MA, Barrett HR. Clearance of postprandial and lipolytically-modified human HDL in rabbits and rats. J Lipid Res. 1997; 38: 1771eC1781.

    Rashid S, Uffelman KD, Barrett PHR, Vicini P, Adeli K, Lewis GF. Triglyceride enrichment of HDL does not alter selective cholesteryl ester clearance in rabbits. J Lipid Res. 2001; 265eC271.

    Rashid S, Barrett PH, Uffelman KD, Watanabe T, Adeli K, Lewis GF. Lipolytically modified triglyceride-enriched HDLs are rapidly cleared from the circulation. Arterioscler Thromb Vasc Biol. 2002; 22: 483eC487.

    Rashid S, Trinh DK, Uffelman KD, Cohn JS, Rader DJ, Lewis GF. Expression of human hepatic lipase in the rabbit model preferentially enhances the clearance of triglyceride-enriched versus native high-density lipoprotein apolipoprotein A-I. Circulation. 2003; 107: 3066eC3072.

    Guendouzi K, Jaspard B, Barbaras R, Motta C, Vieu C, Marcel Y, Chap H, Perret B, Collet X. Biochemical and physical properties of remnant-HDL2 and of pre beta 1-HDL produced by hepatic lipase. Biochemistry. 1999; 38: 2762eC2768.

    Syvanne M, Ahola M, Lahdenpera S, Kahri J, Kuusi T, Virtanen KS, Taskinen MR. High density lipoprotein subfractions in non-insulin-dependent diabetes mellitus and coronary artery disease. J Lipid Res. 1995; 36: 573eC582.

    Despres JP, Ferland M, Moorjani S, Nadeau A, Tremblay A, Lupien PJ, Theriault G, Bouchard C. Role of hepatic-triglyceride lipase activity in the association between intra-abdominal fat and plasma HDL cholesterol in obese women. Arterioscler. 1989; 9: 485eC492.

    Sibley SD, Palmer JP, Hirsch IB, Brunzell JD. Visceral obesity, hepatic lipase activity, and dyslipidemia in type 1 diabetes. J Clin Endocrinol Metab. 2003; 88: 3379eC3384.

    Nie L, Wang J, Clark LT, Tang A, Vega GL, Grundy SM, Cohen JC. Body mass index and hepatic lipase gene (LIPC) polymorphism jointly influence postheparin plasma hepatic lipase activity. J Lipid Res. 1998; 39: 1127eC1130.

    Baynes C, Henderson AD, Richmond W, Johnston DG, Elkeles RS. The response of hepatic lipase and serum lipoproteins to acute hyperinsulinaemia in type 2 diabetes. Eur J Clin Invest. 1992; 22: 341eC346.

    Knauer TE, Woods JA, Lamb RG, Fallon HJ. Hepatic triacylglycerol lipase activities after induction of diabetes and administration of insulin or glucagon. J Lipid Res. 1982; 23: 631eC637.

    Lewis GF, Murdoch S, Uffelman K, Naples M, Szeto L, Albers A, Adeli K, Brunzell JD. Hepatic lipase mRNA, protein and plasma enzyme activity is increased in the insulin-resistant fructose-fed Syrian Golden hamster and is partially normalized by the insulin sensitizer, rosiglitazone. Diabetes. 2004; 53: 2893eC2900.

    Reilly MP, Rader DJ. The metabolic syndrome: more than the sum of its parts Circulation. 2003; 108: 1546eC1551.

    Broedl UC, Jin W, Rader DJ. Endothelial lipase: a modulator of lipoprotein metabolism upregulated by inflammation. Trends Cardiovasc Med. 2004; 14: 202eC206.(Gary F. Lewis, Daniel J. )