Editorial Open Access
Copyright ©2006 Baishideng Publishing Group Co., Limited. All rights reserved.
World J Gastroenterol. Oct 28, 2006; 12(40): 6429-6439
Published online Oct 28, 2006. doi: 10.3748/wjg.v12.i40.6429
Emerging roles of the intestine in control of cholesterol metabolism
Janine K Kruit, Folkert Kuipers, Department of Pediatrics, Center for Liver, Digestive, and Metabolic Diseases, University Medical Center Groningen, Groningen, The Netherlands
Albert K Groen, Department of Experimental Hepatology, Academic Medical Center, Amsterdam, The Netherlands
Theo J van Berkel, Division of Biopharmaceutics, Leiden/Amsterdam Center for Drug Research, Leiden University, Leiden, The Netherlands
Supported by grant 2001B043 from the Netherlands Heart Foundation
Correspondence to: Dr. Folkert Kuipers, Professor, Laboratory of Pediatrics, Center for Liver, Digestive and Metabolic Diseases, University Medical Center Groningen, Hanzeplein 1 9700 RB Groningen, The Netherlands. f.kuipers@med.umcg.nl
Telephone: +31-50-3632669 Fax: +31-50-3611746
Received: June 2, 2006
Revised: June 12, 2006
Accepted: July 7, 2006
Published online: October 28, 2006

Abstract

The liver is considered the major “control center” for maintenance of whole body cholesterol homeostasis. This organ is the main site for de novo cholesterol synthesis, clears cholesterol-containing chylomicron remnants and low density lipoprotein particles from plasma and is the major contributor to high density lipoprotein (HDL; good cholesterol) formation. The liver has a central position in the classical definition of the reverse cholesterol transport pathway by taking up periphery-derived cholesterol from lipoprotein particles followed by conversion into bile acids or its direct secretion into bile for eventual removal via the feces. During the past couple of years, however, an additional important role of the intestine in maintenance of cholesterol homeostasis and regulation of plasma cholesterol levels has become apparent. Firstly, molecular mechanisms of cholesterol absorption have been elucidated and novel pharmacological compounds have been identified that interfere with the process and positively impact plasma cholesterol levels. Secondly, it is now evident that the intestine itself contributes to fecal neutral sterol loss as a cholesterol-secreting organ. Finally, very recent work has unequivocally demonstrated that the intestine contributes significantly to plasma HDL cholesterol levels. Thus, the intestine is a potential target for novel anti-atherosclerotic treatment strategies that, in addition to interference with cholesterol absorption, modulate direct cholesterol excretion and plasma HDL cholesterol levels.

Key Words: Cholesterol metabolism, Intestine, High density lipoprotein, Cholesterol absorption, Reverse cholesterol transport



INTRODUCTION

Maintenance of cholesterol homeostasis in the body requires accurate metabolic cross-talk between processes that govern de novo cholesterol synthesis and turnover to adequately cope with (large) fluctuations in dietary cholesterol intake. Imbalance may lead to elevated plasma cholesterol levels and increased risk for cardiovascular diseases (CVD), the main cause of death in Western society. A multitude of epidemiological studies has shown the direct link between high plasma cholesterol, particularly of low density lipoprotein (LDL) cholesterol, and risk for CVD. Treatment of high plasma cholesterol has been focused for many years on interference with cholesterol synthesis by application of statins. Statins are competitive inhibitors of 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase, the rate-controlling enzyme in the cholesterol biosynthesis pathway[1]. Inhibition of cholesterol synthesis leads to reduced production of very low density lipoprotein (VLDL) particles by the liver and particularly, up-regulation of LDL receptor activity. Both processes contribute to lowering of plasma LDL-cholesterol levels[2]. Large clinical trials have established the beneficial effects of statin treatment[3]. However, a relative large number of hypercholesterolaemic patients do not adequately respond to statin therapy or remain at risk for CVD despite substantial reductions in LDL cholesterol[4,5]. Consequently, alternative strategies are currently actively pursued, particularly high density lipoprotein (HDL)-raising approaches. These approaches are considered particularly promising, as data from epidemiological studies indicate that every 1 mg/dL increase in HDL cholesterol reduces CVD risk by 2%-3%[6,7]. In addition, strategies aiming at interference with intestinal cholesterol metabolism are gaining interest. A major development has been the introduction of ezetimibe, a potent inhibitor of intestinal cholesterol absorption that reduces plasma LDL-cholesterol by approximately 20% in mildly hypercholesteroleamic patients[8]. Likewise, phytosterol/stanol (esters)-enriched functional foods have successfully been introduced for lowering of plasma cholesterol levels through interference with cholesterol absorption[9].

Recently obtained insights in intestinal cholesterol trafficking may open even more promising avenues for further developments. It appears that the intestine actively excretes cholesterol and thereby, significantly contributes to fecal sterol excretion. In addition, it appears that the intestine is an important source of HDL-cholesterol, also known as “good” cholesterol. Thus, the intestine is an attractive target for new therapeutic strategies aimed to alter plasma cholesterol profiles and to reduce the risk for CVD. This review summarizes the important new findings regarding the mechanism(s) of intestinal cholesterol absorption, with specific focus on newly identified transporter proteins, the novel concept of direct intestinal cholesterol secretion and the role of the intestine in HDL biogenesis.

SOME BASIC FEATURES OF CHOLESTEROL

Cholesterol is essential for mammalian life as a structural component of cellular membranes, influencing membrane organization and thereby membrane properties[10]. Cholesterol is the precursor molecule of steroid hormones and therefore, essential for metabolic control. In the liver, cholesterol can be converted into bile salts, which represents the major pathway for cholesterol metabolism in quantitative sense. Bile salts are amphipathic molecules that facilitate the absorption of dietary cholesterol, fats and fat-soluble vitamins in the small intestine. Recently, it has become clear that bile salts are able to regulate gene expression through activation of the nuclear receptor, the farnesoid X receptor (FXR)[11-13]. Cholesterol or more correctly, oxidized cholesterol acts as a ligand for the nuclear liver X receptor (LXR or NRH2 or NRH3) and directly contributes to regulation of expression of genes involved in cholesterol, lipid, and glucose metabolism. Accumulation of free cholesterol rather than cholesterylesters, has been shown to induce apoptosis in macrophages by activating the Fas pathway[14]. Thus, cholesterol is a key component in cellular and whole-body physiology and cholesterol homeostasis is tightly regulated at a variety of levels.

Body cholesterol derives from two sources, i.e., de novo biosynthesis and diet. Cholesterol is synthesized from two-carbon acetyl-CoA moieties. The rate-controlling enzyme in the synthetic pathway is HMG-CoA reductase, a highly regulated enzyme that catalyses the conversion of HMG-CoA into mevalonate. Cholesterol itself regulates feed-back inhibition of HMG-CoA reductase activity, as accumulation of sterols in the endoplasmic reticulum (ER) membrane triggers HMG-CoA reductase to bind to Insig proteins, which leads to ubiquitination and degradation of HMG-CoA reductase[15,16]. In addition, cholesterol regulates the gene expression of HMG-CoA reductase indirectly by blocking the activation of the transcription factor, sterol regulatory element-binding protein 2 (SREBP2). Under low-cholesterol conditions, SREBP2 in the ER binds to the SREBP cleavage activating protein (SCAP), which escorts SREBP2 to the Golgi. In the Golgi, SREBP2 is cleaved to generate its transcriptionally active form, which activates transcription of the HMG-CoA reductase encoding gene. Upon accumulation of sterols in the ER-membrane, binding of cholesterol to the sterol-sensing domain of SCAP causes a conformation change, which induces binding of SCAP to the ER anchor protein Insig, preventing exit of SCAP-SREBP2 complexes to the Golgi thereby preventing activation of SREBP2[17] .

The contribution of the two sources to the total pool of cholesterol differs between species and prevailing diet composition, but the total cholesterol pool is similar in rodents and humans when expressed on the basis of body weight[18]. Cholesterol synthesis in the liver is highly sensitive to the amount of (dietary) cholesterol that reaches the liver from the intestine via the chylomicron-remnant pathway[19]. The Western-type human diet provides approximately 400 mg of cholesterol per day. On top of this, the liver secretes approximately 1 gram of cholesterol into bile per day[20]. Intestinal cholesterol absorption efficiency in humans is highly variable, ranging from 15% to 85% in healthy subjects[21]. After uptake by enterocytes, cholesterol is packed with triglycerides into chylomicrons and secreted into the lymph. In the circulation, the triglycerides are rapidly hydrolyzed and free fatty acids are taken up by the peripheral tissues. Cholesterol-enriched chylomicron remnants are subsequently cleared by the liver. Since chylomicron remnants, which contain most of the cholesterol that is being absorbed from the intestine, are rapidly taken up by the liver, interference with the absorption process directly influences hepatic cholesterol synthesis.

The healthy liver is perfectly equipped for handling large amounts of cholesterol. When relatively large amounts of cholesterol reach the liver, de novo synthesis and LDL uptake are rapidly down-regulated. In addition, the liver can dispose excess cholesterol molecules in several ways. A rapid response involves esterification of cholesterol by Acyl CoA cholesterol acyltransferase (ACAT) 2 for storage as cholesterylesters in cytoplasmic lipid droplets. Cholesterylester can be hydrolyzed when necessary and this esterification/hydrolysis cycle provides cells with short-term buffering capacity for cholesterol. The liver, like the intestine, is able to produce and secrete VLDL particles, which consist of a neutral lipid core composed of cholesterylesters and triacylglycerols and a monolayer surface containing phospholipids, free cholesterol, and a variety of apolipoproteins. Finally, cholesterol can be converted into bile acids by the hepatocytes, followed by their secretion into the bile along with significant amounts of free cholesterol and phosphatidylcholine. In humans, cholesterol lost via the feces consists of approximately 50% acidic (= bile acids) and 50% neutral sterols, emphasizing the point that conversion into bile acids represents a major pathway for cholesterol elimination.

Peripheral cells, e.g., macrophages, muscle and fat cells, are not able to form lipoproteins or to metabolize cholesterol extensively. Therefore, these cell types depend massively on efflux pathways for removal of their excess cholesterol. It is generally assumed that HDL is the primary acceptor for cholesterol efflux from cells. HDL cholesterol can subsequently be taken up by the liver for further processing. This pathway is generally referred to as the reverse cholesterol transport (RCT) pathway. The RCT pathway is particularly important for removal of excess cholesterol from macrophages, as accumulation of esterified cholesterol in these cells is considered a primary step in the development of atherosclerosis. Several epidemiological studies have shown that plasma HDL is an independent, negative risk factor for the development of CVD. The common hypothesis is that high HDL cholesterol levels decrease the risk for CVD by removing the excess of cholesterol from the macrophages and enhancing RCT. Recent work, however, indicates that this is an oversimplification and that current concepts of RCT require re-definition[22]. In addition, the anti-inflammatory and anti-oxidant features of molecules rather than cholesterol associated with the HDL particles, like paraoxonase, platelet activating factor-acetylhydrolase or lysophospholipids, are becoming increasingly apparent[23-25].

TOWARDS UNDERSTANDING OF INTESTINAL CHOLESTEROL ABSORPTION

In the past years, insight in regulation of cholesterol absorption has greatly increased by identification of transporter proteins involved. In addition, unraveling of molecular regulation of their expression is progressing. Yet, it should be realized that besides transporter proteins, the presence of bile acids in the intestinal lumen is an essential prerequisite for absorption to occur[26]. Micellar solubilization of (dietary/biliary) cholesterol is necessary for its absorption as exemplified by the fact that fractional cholesterol absorption is virtually zero in bile-diverted rats and Cyp7a1-deficient mice with a strongly diminished bile acid pool size[26].

Identification of novel proteins involved in cholesterol absorption

Cholesterol absorption has long been considered a merely passive process, despite the fact that the process is clearly selective since dietary cholesterol is absorbed with a relative high efficiency whereas structurally similar phytosterols are not. Several candidate intestinal cholesterol transporters have been proposed during the past couple of years, e.g., SR-B1[27] and aminopeptidase N[28], but their role (if any) has remained elusive so far. The recent identification of the Niemann-Pick C1 like 1 (NPC1L1) protein as a crucial molecule involved in cholesterol uptake by enterocytes[29] and of Abcg5 and Abcg8 proteins as (intestinal) cholesterol efflux transporters[30-32], has provided definite proof that cholesterol absorption is a protein-mediated, selective and active process.

The identification of NPC1L1 is strongly facilitated by the discovery of a powerful cholesterol absorption inhibitor named ezetimibe[33]. Ezetimibe and analogs comprise a new class of sterol absorption inhibitors that reduce diet-induced hypercholesterolemia in mice, hamsters, rats, rabbits, dogs, monkeys and humans[8,33-37]. Using a bioinformatics approach, Altmann et al[29] have identified the NPC1L1 protein as a putative cholesterol transporter in intestinal cells. NPC1L1 is expressed in the intestine at the brush border membrane and Npc1l1- deficient mice show a 69% reduction in fractional cholesterol absorption. Importantly, treatment with ezetimibe could not further reduce fractional cholesterol absorption efficiency in these mice, indicating that NPC1L1 at least is involved in a pathway targeted by ezetimibe[29]. In support of this, recent studies have shown that ezetimibe glucuronide, the active molecule, indeed binds to cells expressing NPC1L1[38]. Using intestinal brush border membrane (BBM) fractions, the authors showed that ezetimibe binds specifically to a single site in the brush border membrane and that this binding is lost in BBM fractions of Npc1l1- deficient mice[38]. The exact cellular localization of NPC1L1 is, however, still under debate. Iyer et al[39] showed that NPC1L1 is glycosylated and enriched in the BBM of rat enterocytes. Davies et al[40] who were the first to identify NPC1L1 as a homolog of the Niemann Pick type C (NPC) protein[40], showed in HepG2 cells that NPC1L1 is localized to a subcellular vesicular compartment but not in the plasma membrane. Using immortalized fibroblasts from wild-type and Npc1l1 knock-out mice these authors also showed that lack of NPC1L1 activity causes dysregulation of caveolin transport and localization, suggesting that the observed sterol transport defect may be an indirect result of the inability of Npc1l1-deficient cells to properly target and/or regulate cholesterol transport in the cells.

Another possible mechanism of action of ezetimibe has been proposed by Smart and colleagues[41]. These authors described the presence of a stable complex of annexin (ANX) 2 and caveolin (CAV) 1 located in enterocytes of zebrafish and mouse. Disruption of this complex by morpholino antisense oligonucleotides in zebrafish prevented normal uptake of cholesterol. Ezetimibe treatment of zebrafish, C57Bl/6 mice fed a Western type diet and LDL receptor knock-out mice disrupted the ANX2-CAV1 complex, suggesting that ANX2 and CAV1 are components of an intestinal sterol transport complex and targets for ezetimibe. Interestingly, C57BL/6 mice fed a standard diet did not show disruption of the ANX2-CAV1 complex upon ezetimibe treatment, but did show decreased cholesterol absorption[41]. Moreover, recent research using CAV1-deficient mice revealed that inhibition of cholesterol absorption by ezetimibe did not require the presence of CAV1[42]. In addition, rabbits did not appear to form the ANX2-CAV1 complexes, yet, their cholesterol absorption efficiency was still inhibited by ezetimibe[43]. Collectively, these studies make a mode of action in which ezetimibe acts by deregulating the ANX2-CAV1 complex less likely.

Other proteins critical in control of sterol absorption are the ATP-binding cassette (ABC) transporter proteins, G5 and G8. ABCG5 and ABCG8 act as functional heterodimers[44] and are localized at the canalicular membrane of hepatocytes and at the brush border membrane of enterocytes. Mutations in the human genes encoding ABCG5 or ABCG8 have been shown to cause the inherited disease sitosterolemia[30-32], which is characterized by an accumulation of plant sterols (e.g., sitosterol, campesterol) in blood and tissues due to their enhanced intestinal absorption and decreased biliary removal. Thus, ABCG5/ABCG8 limit plant sterol absorption by effective efflux back into the intestinal lumen. Since ABCG5/ABCG8 also accommodate cholesterol, as evidenced from the fact that Abcg5/g8-deficient mice show a strongly reduced biliary cholesterol secretion[45], this system also provides a means to control cholesterol absorption efficiency. Yet, Abcg5 and/or Abcg8 deficiency in mice clearly enhances phytosterol absorption[45-47], but reported effects on cholesterol absorption efficiency are minimal[45,46]. On the other hand, overexpression of ABCG5 and ABCG8 in mice as well as pharmacological induction of their expression did lead to a strongly decreased fractional cholesterol absorption[46,48,49], indicating that ABCG5 and ABCG8 play a role in control of cholesterol absorption under certain conditions.

Other transporter proteins, like the scavenger receptor BI (SR-BI) and ABCA1 have been suggested to play a role in control of cholesterol absorption. In the small intestine, SR-BI is localized both at the apical membrane and at the basolateral membrane of enterocytes, with different expression levels along the length of the small intestine[50]. It was reported that mice deficient in SR-B1, however, show only a small increase in fractional cholesterol absorption efficiency and a small decrease in fecal neutral sterol output[51]. On the other hand, intestine-specific overexpression of SR-BI in mice did lead to increased cholesterol and triglyceride absorption in short-term absorption experiments[52], indicating that SR-BI might have a role in cholesterol absorption.

Although earlier reports[53] have suggested an apical localization, it is evident that ABCA1 is localized at the basolateral membranes of chicken enterocytes[54] and human CaCo-2 cells[55]. The conflicting results yielded in studies assessing intestinal cholesterol absorption in mice lacking Abca1[56,57], suggest that the overall effect of Abca1 on absorption is very minor. However, as will be described later, this protein does have an important function in intestinal cholesterol metabolism.

After uptake, cholesterol is esterified by the enzyme ACAT 2 in the endoplasmic reticulum (ER) of enterocytes. It was reported that Acat2-deficiency in mice on a low-cholesterol chow diet did not affect cholesterol absorption efficiency, however, Acat2-deficient mice did show a clear reduction in cholesterol absorption upon feeding a high-fat/high-cholesterol diet and as a consequence, are resistant to diet-induced hypercholesterolemia[58]. Other proteins crucial for cholesterol absorption are those involved in chylomicron formation, like apolipoprotein B (apoB) and microsomal triglyceride transfer protein (MTP), and proteins involved in intracellular chylomicron trafficking such as SARA2. Mutations in the MTP gene result in abetalipoproteinemia, an inherited human disease characterized by extremely low plasma cholesterol and triglyceride levels and absence of apoB-containing particles. Patients suffer from fat and cholesterol malabsorption and neurological diseases due to malabsorption of lipid-soluble vitamins. Mutations in SARA2 cause chylomicron retention disease or Anderson disease[59], both of which are characterized by the inability to secrete chylomicrons causing severe fat malabsorption and accumulation of chylomicron-like particles in enterocytes. SARA2 belongs to the Sar1-ADP-ribosylation factor family of small GTPases and is involved in intracellular trafficking of chylomicrons through the secretory pathway[59].

The major routes of cholesterol in enterocytes and the proteins involved are depicted schematically in Figure 1.

Figure 1
Figure 1 Schematic overview of the major routes of cholesterol in enterocytes. Dietary and biliary cholesterol are taken up via the action of NPC1L1. In the ER, cholesterol is esterified and incorporated into chylomicrons, which are subsequently secreted into lymph. Non-esterified sterols can be re-secreted into the intestinal lumen via the action of ABCG5/G8 or secreted towards ApoA1 via the action of ABCA1. ABCA1, ABCG5, ABCG8: ATP-binding cassette transporter A1, G5, G8; ACAT2: Acyl-coenzyme A: Cholesterol acyl transferase 2; ApoAI, ApoB48, apolipoprotein AI, B48; C: Cholesterol; CE: Cholesterylester; ER: Endoplasmatisch reticulum; MTP: Microsomal triglyceride transfer protein; NPC1L1: Niemann Pick C 1 like 1 protein; SARA2: Sar1-ADP-ribosylation GTPase 2; TG: Triglycerides.
Regulation of cholesterol absorption

As indicated above, cholesterol can be taken up from the intestinal lumen by NPC1L1 and effluxed back into the lumen via ABCG5 and ABCG8. When both processes are active and present in the same cells, a classical futile cycle arises, enabling very sensitive regulation. Interference with this system has a great potential for reducing plasma cholesterol.

An established application hereof is provided by ezetimibe that interferes with NPC1L1 activity[29,38]. Lowering of NPC1L1 expression provides another potential means to reduce cholesterol absorption. Mechanisms involved in transcriptional control of NPC1L1 are beginning to be unraveled. The nuclear receptor peroxisome proliferator-activated receptor (PPAR) δ/β (NR1C2) has been shown to decrease cholesterol absorption, presumably by decreasing NPC1L1 expression[60]. Activation of PPARδ/β by the synthetic agonist GW610742 resulted in a 43% reduction of cholesterol absorption in mice, which coincides with unchanged intestinal expression of Abcg5 and Abcg8 but a decreased intestinal expression of Npc1l1. Treatment of human colon-derived CaCo-2 cells with ligands for PPARδ/β, but not for PPARγ or PPARα, decreased NPC1L1 expression as well[60]. Whether PPARδ/β regulates NPC1L1 directly or indirectly via transcriptional repression, is still under investigation.

The major regulatory pathways in cholesterol metabolism are controlled by the nuclear receptor liver X receptor (LXR). Two LXR isotypes have been identified in mammals, i.e., LXRα (NR1H3) which is mainly expressed in the liver, kidney, intestine, spleen and adrenals, and LXRβ (NR1H2) which is expressed ubiquitously. Natural ligands for both LXRs are oxysterols. After activation, LXR heterodimerizes with retinoid X receptor (RXR)[61,62]. Activated RXR/LXR heterodimers bind to specific LXR response elements (LXREs) in the promoter regions of their target genes and activate gene transcription. LXR target genes include many genes involved in cellular cholesterol efflux like ABCA1, ABCG1, ABCG5, and ABCG8[53,63,64], genes involved in bile acid synthesis [cholesterol-7α-hydroxylase (Cyp7a1)] in rodent models and genes involved in lipogenesis like sterol regulatory element-binding protein (SREBP) 1C, fatty acid synthase (FAS) and acetyl-CoA carboxylase (ACC). Global LXR activation by synthetic agonists therefore has a plethora of effects including elevated HDL levels, hypertriglyceridemia, hepatic steatosis, increased biliary cholesterol excretion, reduced intestinal cholesterol absorption efficiency and increased neutral sterol loss via the feces[65,66]. The decreased intestinal cholesterol absorption is primarily due to increased cholesterol efflux of cholesterol towards the intestinal lumen due to increased Abcg5 and Abcg8 expression, as fractional cholesterol absorption was reduced upon LXR activation in wild-type mice but remained unaltered in Abcg5/g8-deficient mice[49] and Abcg5-deficient mice[46] under these conditions. Other mechanisms, such as reduced intestinal Npc1l1 expression after LXR activation contribute to reduced cholesterol absorption, as recently shown in ApoE2-knock-out mice[67].

Dietary phytosterols and phytostanols and their esters have been introduced in functional foods to suppress intestinal cholesterol absorption and hence to reduce the risk for CVD[9]. Phytosterols and stanols are thought to decrease cholesterol absorption by competing with cholesterol for incorporation into mixed micelles in the intestinal lumen[68]. However, several recent studies suggested additional mechanisms involving alterations of intestinal gene expression. Igel and colleagues[69] showed for the first time that phytosterols and stanols are actually taken up by the enterocytes and subsequently re-secreted into the gut lumen, most probably through the action of Abcg5/Abcg8 transporters. This finding indicated that phytosterols and stanols, in addition to modes of action within the intestinal lumen, exert metabolic actions from inside the enterocytes. Moreover, dietary phytostanol consumption (2.5 g) once a day reduces LDL cholesterol as effective as consumption of 2.5 g phytostanols ingested in three daily portions[70], suggesting that luminal concentrations may not be the key to the control of metabolic actions. The identification of a phytosterol-derived agonist for the nuclear receptor LXR[71] has led to the proposal that phytosterols and stanols decrease cholesterol absorption via activation of intestinal LXR. In vitro studies in CaCo-2 cells indicated that phytostanols indeed are able to induce the expression of ABCA1, an established LXR target gene[72]. Recent in vivo studies, however, showed that dietary phytosterols and phytostanols decrease cholesterol absorption without activating LXR in rodent models. Field et al[73] showed that addition of 2% phytostanols to a chow diet do not affect intestinal expression of ABC sterol transporters and Npc1l1 in male golden Syrian hamsters. Likewise, Calpe-Berdien et al[74] showed very recently that decreased cholesterol absorption upon addition of 2% phytosterol to a Western type diet is not associated with transcriptional changes in Abca1, Abcg5, Abcg8 or Npc1l1 in C57BL/6J, ApoE-/- and LDLr-/- mice. Plösch and colleagues[66] have shown similar results using 0.5% phytosterol or phytostanol in a semi-synthetic diet containing 0.2% cholesterol in C57BL mice. Additionally, these authors showed that the plant sterol/stanol-induced reduction of cholesterol absorption in mice is not influenced by Abcg5-deficiency (J. Nutr., in press), indicating that intra-luminal events are most relevant for the inhibitory effect of these dietary compounds.

The modes of action of the different cholesterol absorption decreasing compounds are schematically depicted in Figure 2.

Figure 2
Figure 2 Schematic overview of the regulation of cholesterol transport in enterocytes. Plant sterols, ezetimibe PPARδ/β agonists and LXR agonists all reduce cholesterol absorption through different mechanisms. Plant sterols interfere with micellisation of cholesterol. Ezetimibe binds to NPC1L1 and thereby interferes with the cholesterol uptake. Agonists for PPARδ/β reduce expression of NPC1L1 and thereby the amount of NPC1L1 protein. Agonists for LXR increase the expression of ABCG5 and ABCG8 and thereby enhance the efflux of cholesterol towards the intestinal lumen. LXR: Liver X Receptor; PPARδ/β: Peroxisome proliferators-activated receptor δ/β.
NOVEL ROLE OF THE INTESTINE IN REVERSE CHOLESTEROL TRANSPORT

It is clear that the intestine plays a major role in cholesterol homeostasis as a cholesterol absorbing organ. However, recent studies revealed that the intestine also acts as an excretory organ in the reverse cholesterol transport (RCT) pathway[66,75]. This pathway is classically defined as the HDL-mediated flux of cholesterol from peripheral cells to the liver, followed by its secretion into bile and disposal via the feces. RCT is extremely important in prevention of CVD as it removes excess cholesterol from macrophages present in the arterial vessel wall. The amount of cholesterol secreted into bile is substantial. As only part of it is absorbed by the intestine, it contributes significantly to cholesterol loss via the feces. However, a novel pathway that contributes to fecal cholesterol loss has recently been established.

Already in the nineteen-sixties, it was suggested that non-dietary cholesterol present in the intestinal lumen consists of a fraction secreted by the liver into the bile and a second fraction directly secreted by the intestine. Measuring dietary cholesterol, cholesterol absorption and cholesterol loss via the feces in patients with complete obstruction of common bile duct due to carcinoma of the head of the pancreas unequivocally established the presence of intestinally secreted cholesterol in the feces[76]. By intestinal perfusion studies in humans, Simmonds et al[77] have tried to quantify this route. In a triple lumen tube system, perfusion studies were carried out using micellar solutions with radio-labeled cholesterol. Decrease in specific activity was interpreted as secretion of endogenous cholesterol from the intestine and the contribution of endogenously secreted cholesterol from the intestine was estimated to be about 44% of total fecal output, but direct proof for the existence of this pathway could not be provided[77].

Since these early experiments, the focus of research has shifted more towards the liver. Biliary cholesterol and bile acid secretions are believed to represent the major pathways for removal of excess cholesterol. However, recent calculations of cholesterol fluxes in different mouse models again emphasize the relevance of intestinal cholesterol secretion. A striking example is provided by the Cyp7a1-deficient mouse[78]. Cyp7a1 is important for converting cholesterol into bile acids and catalyzing the formation of 7α-hydroxycholesterol[79]. As Cyp7a1 is rate-controlling in this pathway, it is regulated in a complex manner involving multiple nuclear receptors. Cyp7a1-deficiency in mice leads to significantly decreased fecal bile acid loss and bile acid pool size. Surprisingly, fecal neutral sterol output is increased two times in Cyp7a1-/- mice, although biliary cholesterol concentration remains unaffected[78]. As dietary intake and cholesterol absorption are known, it can be calculated from these data that direct intestinal cholesterol secretion contributes at least 30% to the increased fecal neutral sterol output.

Plösch and colleagues[66] showed that the pathway of intestinal cholesterol secretion can be induced in mice by treatment with the synthetic LXR agonist T0901317. In C57BL/6 mice, efflux of cholesterol from the intestinal epithelium into the lumen, calculated from the difference between dietary and biliary input minus fecal output, contributed up to 36% of the total fecal cholesterol loss. Pharmacological LXR activation in these mice tripled the intestinal cholesterol secretion, showing that this represents a valid, inducible pathway for removal of cholesterol in mice.

To further characterize this route, Kruit et al[75] studied the effects of LXR activation by the synthetic agonist GW3965 in wild-type and Mdr2-deficient mice. Mdr2-Pgp (or Abcb4 according to the new nomenclature) mediates the ATP-dependent translocation of phospholipids at the canalicular membrane of hepatocytes. Consequently, Mdr2-deficiency leads to the inability to secrete phospholipids into the bile. Due to the tight coupling of phospholipid and cholesterol secretion, these mice also show a severely impaired biliary cholesterol secretion[80,81]. Despite the impaired biliary cholesterol secretion, chow-fed Mdr2-/- mice showed a similar fecal neutral sterols loss as wild-type mice, suggesting that the intestine indeed contributes to the fecal neutral sterol loss. LXR activation increased fecal neutral sterol output to a similar extent in Mdr2-/- and wild-type mice, although biliary cholesterol secretion remained impaired in Mdr2-/- mice but increased in wild-type mice. These data show that the increased fecal cholesterol loss upon LXR activation is independent of biliary cholesterol secretion. Although fractional cholesterol absorption decreased to a greater extent in Mdr2-/- mice compared to wild-type mice upon LXR activation, it could be calculated that at least 57% of fecal cholesterol originates from intestinal secretion in Mdr2-/- mice[75].

The most intriguing question, namely the origin of intestine-derived cholesterol has remained unanswered so far. Part of the cholesterol could, in theory, originate from enhanced sloughing of intestinal cells or reflect a consequence of increased intestinal de novo cholesterol synthesis. Indeed, increased intestinal cholesterol synthesis has been found in Cyp7a1-/- mice[78]. Upon LXR activation, however, intestinal HMGCoA reductase gene expression remained unchanged[66,75], indicative for unchanged cholesterol synthesis, while fecal sterol loss increased 3 times. Staining for the proliferation marker Ki-67 has revealed no signs of increased intestinal cell proliferation upon LXR activation, making the possibility of enhanced cell shedding less likely[75]. Using intravenously injected radiolabeled cholesterol as a marker, Kruit and colleagues[75] additionally showed that fecal loss of plasma-derived cholesterol is 1.7-fold higher upon LXR activation in Mdr2-/- mice, suggesting that the intestine plays an important role independently of biliary cholesterol in cholesterol transport from plasma to the feces.

Further research should be done to identify the putative proteins involved in this pathway. The sterol efflux proteins, ABCG5/ABCG8, seem to be good candidates, as increased fecal neutral sterol output upon LXR activation requires the presence of Abcg5 and Abcg8[49] and transgenic mice overexpressing human ABCG5 and ABCG8 (hG5G8Tg) showed significantly-increased fecal neutral sterol loss[48]. However, deficiency of Abcg5 and/or Abcg8 leads to only mild[45,49] or no[46] decrease in fecal neutral sterol loss and the increased fecal neutral sterol excretion loss in the hG5G8Tg mice was inhibited in hG5G8Tg mice lacking Mdr2 (Mdr2-/-hG5G8Tg mice), suggesting that biliary cholesterol secretion is responsible for the increased fecal sterol loss in hG5G8Tg mice[82]. However, hG5G8Tg mice showed a high expression of human ABCG5 and ABCG8 in the liver but their expression in the intestine was far less pronounced[48]. Thus, the question whether intestinal ABCG5 and ABCG8 are important for intestinal cholesterol efflux under normal conditions still remains unanswered.

Virtually nothing is known about transporter systems involved in uptake of plasma cholesterol by enterocytes prior to its excretion into the intestinal lumen. LXR activation can upregulate a number of cholesterol transporters, of which only SR-BI is known to be involved in cholesterol uptake, at least in the liver. Chow-fed SR-B1-/- mice show only a small decrease in fecal neutral sterol loss, suggesting a relatively small contribution of intestinal SR-B1 to the control of fecal cholesterol excretion. However, basolaterally localized SR-B1 in enterocytes could theoretically play a role in cholesterol. When free cholesterol in enterocytes decreases due to activation of ABCG5 and ABCG8, uptake of the sterol from the plasma compartment may become energetically favorable.

INTESTINAL CONTRIBUTION TO HDL BIOGENESIS

The intestine along with the liver, has been known for many years to synthesize and secrete apolipoprotein A-I (ApoA-I), the principal apolipoprotein of HDL. Already in 1977, Glickman and Green[83] have described the synthesis of ApoA-I by the intestine of rats. One year later, Wu and Windmueller[84] estimated that intestinally synthesized ApoA-I contributes up to 56% of total plasma ApoA-I in rats. A potential role for the intestine in HDL particle assembly was initially suggested from experiments in hepatectomized dogs and studies describing the presence of HDL in mesenteric lymph[86-89]. More recently, in vitro studies using the human colon carcinoma cell line CaCo-2 showed that basolateral efflux of cholesterol occurs in high density ApoB-free, ApoA-I containing lipoproteins[90,91].

In addition to ApoA-I, ATP-binding cassette (ABC) transporter 1 (ABCA1) is of crucial importance for HDL formation. Three different groups have independently reported mutations of the ABCA1 gene as the cause of Tangier disease[92-94]. Tangier disease is characterized by almost complete absence of plasma HDL, abnormal accumulation of cholesteryl esters in reticuloendothelial cells of many tissues and early incidence of atherosclerosis. No abnormalities in the ApoA-I protein[95] or in protein synthesis have been found. These findings and the subsequent generation of Abca1-/- mice which also lack plasma HDL[57], underscore that ABCA1 is crucial for HDL formation.

ABCA1 performs the rate-controlling step in HDL formation by mediating the efflux of cholesterol and phospholipids to nascent ApoA-I. ABCA1 is widely expressed throughout the body[96], however not all tissues are important for the regulation of plasma HDL. Bone marrow transplantation studies in which bone marrow of wild-type and Abca1-/- mice was transplanted into Abca1-/- or wild-type mice, respectively, revealed that macrophage expression of Abca1 contributes only minimally to plasma HDL[97]. Macrophage ABCA1 is, however, important for the development of atherosclerosis because deficiency of Abca1 in bone marrow-derived cells increased the susceptibility to atherosclerosis in sensitive strains of mice[98,99]. Conversely, overexpression of ABCA1 in bone marrow-derived cells inhibited the progression of atherosclerotic lesions in such mice[100].

As both the liver and intestine synthesize ApoA-I and express significant levels of ABCA1, they are prone to contribute to plasma HDL levels. Indeed, mice overexpressing human ABCA1 in the liver and macrophages showed increased plasma HDL levels. Since macrophage ABCA1 can only minimally increase plasma HDL[97], this indicates that plasma HDL is controlled by hepatic ABCA1. A similar conclusion can be drawn from studies employing adenoviral Abca1 transfer to mouse liver in vivo[101,102]. Basso et al showed that treatment of C57BL/6 mice with adenovirus containing rABCA1-GFP resulted in a 2-fold increase in plasma HDL levels. Wellington et al[102] treated mice with increasing doses of ABCA1-containing adenoviruses, resulting in a dose-dependent increase in hepatic ABCA1 protein expression. HDL cholesterol was increased in mice injected with low doses of adABCA1, but surprisingly higher doses did not further raise plasma HDL levels[102]. Liver-specific Abca1 knockdown by 50% in mice using siRNA resulted in a 40% decrease of plasma HDL cholesterol levels, indicating that hepatic Abca1 expression correlates with plasma HDL levels in mice[103].

The creation of liver-specific Abca1 knock-out (Abca1-L/-L) mice definitively showed that the liver is the major contributor to plasma HDL as liver-specific deficiency of Abca1 results in a decrease of plasma HDL cholesterol levels by 80%. Further analysis revealed that in vivo catabolism of HDL ApoA-I isolated from wild-type mice was 2-fold higher in Abca1-L/-L mice due to a 2-fold higher rate of catabolism of ApoA-I in the kidneys[104]. These data unequivocally demonstrate that hepatic Abca1 is responsible for the maintenance of the circulating plasma HDL by direct lipidation of lipid-poor ApoA-1 containing particles. These data also show that, although the liver is the major organ responsible for HDL levels, additional extra-hepatic sites also contribute to HDL biogenesis.

To address the contribution of intestinal Abca1 to plasma HDL, intestine-specific Abca1 knockout (Abca1-i/-i) mice were created using the Cre/Lox system with the Cre transgene under the control of the villin promoter[105]. Brunham et al showed that intestinal Abca1 deficiency resulted in a 30% decrease in plasma HDL cholesterol levels, indicating that intestinal Abca1 is critically involved in HDL biogenesis. Combined deletion of both hepatic and intestinal Abca1 resulted in a 90% decrease of plasma HDL, which was similar to the level found in the whole-body Abca1-/- mice, proving that the liver and intestine are really the two major sites for HDL biogenesis. Absence of intestinal Abca1 resulted in decreased transport of dietary cholesterol into plasma HDL, but total intestinal cholesterol absorption was not affected. Surprisingly, lymphatic HDL content was hardly affected in Abca1-i/-i mice. In contrast, HDL was virtually absent in lymph of Abca1-L/-L mice, indicating that lymph HDL originates from the plasma compartment rather than directly from the intestine[105]. This finding has solved a long-lasting debate on the origin of lymphatic HDL[83,86-88,106,107]. It would be interesting to see whether lack of intestinal Abca1 influences the development of atherosclerosis.

Modulation of plasma HDL by intestine- specific LXR activation

As discussed above, LXR is a major regulator of cholesterol metabolism and LXR agonists are considered promising candidates for novel treatment strategies against atherosclerosis. Indeed, treatment of ApoE-/- and LDLr-/- mice, both are sensitive to atherosclerosis development, with synthetic LXR agonists inhibited the development of atherosclerosis[108,109]. However, general LXR activation also leads to increased lipogenesis, hypertriglyceridemia and hepatic steatosis in rodents[65] and is therefore not recommended for its use in humans. Specific LXR activation in the intestine may be beneficial in this respect, as it can theoretically lead to decreased cholesterol absorption, increased intestinal cholesterol excretion and plasma HDL levels. The preliminary data from our laboratory, using an intestine-specific LXR agonist in Wistar rats, showed that intestine-specific LXR activation indeed has the desired effect in this model without adverse effects on triglyceride metabolism.

CONCLUSION

During the past 5 years, a number of developments have greatly contributed to appreciation of the important role of the intestine in maintenance of cholesterol homeostasis (Figure 3). The most important developments include the identification of transporter proteins involved in uptake and secretion of cholesterol by enterocytes, the establishment of the direct cholesterol excretion pathway of the intestine, and the definition of the role of the intestine in HDL biogenesis.

Figure 3
Figure 3 Schematic overview of the involvement of the intestine in cholesterol homeostasis. The intestine is critically involved in the control of plasma cholesterol due to its role in intestinal cholesterol absorption (1), direct cholesterol excretion into the intestinal lumen (2), and HDL biogenesis (3). CM: Chylomicron; HDL: High density lipoprotein; LDL: Low density lipoprotein; VLDL: Very low density lipoprotein.

A wealth of data indicate that the intestine should be considered a promising target for development of anti-atherosclerotic drugs that, in addition to interference with cholesterol absorption, may directly modulate cholesterol excretion and plasma HDL cholesterol levels.

Footnotes

S- Editor Liu Y L- Editor Wang XL E- Editor Bai SH

References
1.  Endo A. The discovery and development of HMG-CoA reductase inhibitors. J Lipid Res. 1992;33:1569-1582.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Brown MS, Goldstein JL. A receptor-mediated pathway for cholesterol homeostasis. Science. 1986;232:34-47.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4078]  [Cited by in F6Publishing: 3884]  [Article Influence: 102.2]  [Reference Citation Analysis (0)]
3.  Ong HT. The statin studies: from targeting hypercholesterolaemia to targeting the high-risk patient. QJM. 2005;98:599-614.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 37]  [Cited by in F6Publishing: 38]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
4.  Illingworth DR, Crouse JR 3rd, Hunninghake DB, Davidson MH, Escobar ID, Stalenhoef AF, Paragh G, Ma PT, Liu M, Melino MR. A comparison of simvastatin and atorvastatin up to maximal recommended doses in a large multicenter randomized clinical trial. Curr Med Res Opin. 2001;17:43-50.  [PubMed]  [DOI]  [Cited in This Article: ]
5.  Jones PH, Davidson MH, Stein EA, Bays HE, McKenney JM, Miller E, Cain VA, Blasetto JW. Comparison of the efficacy and safety of rosuvastatin versus atorvastatin, simvastatin, and pravastatin across doses (STELLAR* Trial). Am J Cardiol. 2003;92:152-160.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 971]  [Cited by in F6Publishing: 942]  [Article Influence: 44.9]  [Reference Citation Analysis (0)]
6.  Castelli WP, Garrison RJ, Wilson PW, Abbott RD, Kalousdian S, Kannel WB. Incidence of coronary heart disease and lipoprotein cholesterol levels. The Framingham Study. JAMA. 1986;256:2835-2838.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1314]  [Cited by in F6Publishing: 1169]  [Article Influence: 30.8]  [Reference Citation Analysis (0)]
7.  Gordon T, Kannel WB, Castelli WP, Dawber TR. Lipoproteins, cardiovascular disease, and death. The Framingham study. Arch Intern Med. 1981;141:1128-1131.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 273]  [Cited by in F6Publishing: 276]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
8.  Sudhop T, Lutjohann D, Kodal A, Igel M, Tribble DL, Shah S, Perevozskaya I, von Bergmann K. Inhibition of intestinal cholesterol absorption by ezetimibe in humans. Circulation. 2002;106:1943-1948.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 495]  [Cited by in F6Publishing: 460]  [Article Influence: 20.9]  [Reference Citation Analysis (0)]
9.  Miettinen TA, Puska P, Gylling H, Vanhanen H, Vartiainen E. Reduction of serum cholesterol with sitostanol-ester margarine in a mildly hypercholesterolemic population. N Engl J Med. 1995;333:1308-1312.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 565]  [Cited by in F6Publishing: 572]  [Article Influence: 19.7]  [Reference Citation Analysis (0)]
10.  Maxfield FR, Tabas I. Role of cholesterol and lipid organization in disease. Nature. 2005;438:612-621.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 897]  [Cited by in F6Publishing: 909]  [Article Influence: 47.8]  [Reference Citation Analysis (0)]
11.  Makishima M, Okamoto AY, Repa JJ, Tu H, Learned RM, Luk A, Hull MV, Lustig KD, Mangelsdorf DJ, Shan B. Identification of a nuclear receptor for bile acids. Science. 1999;284:1362-1365.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2012]  [Cited by in F6Publishing: 1993]  [Article Influence: 79.7]  [Reference Citation Analysis (0)]
12.  Parks DJ, Blanchard SG, Bledsoe RK, Chandra G, Consler TG, Kliewer SA, Stimmel JB, Willson TM, Zavacki AM, Moore DD. Bile acids: natural ligands for an orphan nuclear receptor. Science. 1999;284:1365-1368.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1717]  [Cited by in F6Publishing: 1670]  [Article Influence: 66.8]  [Reference Citation Analysis (0)]
13.  Wang H, Chen J, Hollister K, Sowers LC, Forman BM. Endogenous bile acids are ligands for the nuclear receptor FXR/BAR. Mol Cell. 1999;3:543-553.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1168]  [Cited by in F6Publishing: 1166]  [Article Influence: 46.6]  [Reference Citation Analysis (0)]
14.  Yao PM, Tabas I. Free cholesterol loading of macrophages induces apoptosis involving the fas pathway. J Biol Chem. 2000;275:23807-23813.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 144]  [Cited by in F6Publishing: 148]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
15.  Sever N, Song BL, Yabe D, Goldstein JL, Brown MS, DeBose-Boyd RA. Insig-dependent ubiquitination and degradation of mammalian 3-hydroxy-3-methylglutaryl-CoA reductase stimulated by sterols and geranylgeraniol. J Biol Chem. 2003;278:52479-52490.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 222]  [Cited by in F6Publishing: 225]  [Article Influence: 10.7]  [Reference Citation Analysis (0)]
16.  Sever N, Yang T, Brown MS, Goldstein JL, DeBose-Boyd RA. Accelerated degradation of HMG CoA reductase mediated by binding of insig-1 to its sterol-sensing domain. Mol Cell. 2003;11:25-33.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 266]  [Cited by in F6Publishing: 276]  [Article Influence: 13.1]  [Reference Citation Analysis (0)]
17.  Goldstein JL, DeBose-Boyd RA, Brown MS. Protein sensors for membrane sterols. Cell. 2006;124:35-46.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1162]  [Cited by in F6Publishing: 1208]  [Article Influence: 67.1]  [Reference Citation Analysis (0)]
18.  Dietschy JM, Turley SD. Control of cholesterol turnover in the mouse. J Biol Chem. 2002;277:3801-3804.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 210]  [Cited by in F6Publishing: 216]  [Article Influence: 9.8]  [Reference Citation Analysis (0)]
19.  Dietschy JM, Turley SD, Spady DK. Role of liver in the maintenance of cholesterol and low density lipoprotein homeostasis in different animal species, including humans. J Lipid Res. 1993;34:1637-1659.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Grundy SM, Metzger AL. A physiological method for estimation of hepatic secretion of biliary lipids in man. Gastroenterology. 1972;62:1200-1217.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Grundy SM. Absorption and metabolism of dietary cholesterol. Annu Rev Nutr. 1983;3:71-96.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 189]  [Cited by in F6Publishing: 186]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
22.  Groen AK, Oude Elferink RP, Verkade HJ, Kuipers F. The ins and outs of reverse cholesterol transport. Ann Med. 2004;36:135-145.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 51]  [Cited by in F6Publishing: 46]  [Article Influence: 2.3]  [Reference Citation Analysis (0)]
23.  Watson AD, Berliner JA, Hama SY, La Du BN, Faull KF, Fogelman AM, Navab M. Protective effect of high density lipoprotein associated paraoxonase. Inhibition of the biological activity of minimally oxidized low density lipoprotein. J Clin Invest. 1995;96:2882-2891.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 859]  [Cited by in F6Publishing: 894]  [Article Influence: 30.8]  [Reference Citation Analysis (0)]
24.  Watson AD, Navab M, Hama SY, Sevanian A, Prescott SM, Stafforini DM, McIntyre TM, Du BN, Fogelman AM, Berliner JA. Effect of platelet activating factor-acetylhydrolase on the formation and action of minimally oxidized low density lipoprotein. J Clin Invest. 1995;95:774-782.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 309]  [Cited by in F6Publishing: 326]  [Article Influence: 11.2]  [Reference Citation Analysis (0)]
25.  Nofer JR, van der Giet M, Tölle M, Wolinska I, von Wnuck Lipinski K, Baba HA, Tietge UJ, Gödecke A, Ishii I, Kleuser B. HDL induces NO-dependent vasorelaxation via the lysophospholipid receptor S1P3. J Clin Invest. 2004;113:569-581.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Voshol PJ, Schwarz M, Rigotti A, Krieger M, Groen AK, Kuipers F. Down-regulation of intestinal scavenger receptor class B, type I (SR-BI) expression in rodents under conditions of deficient bile delivery to the intestine. Biochem J. 2001;356:317-325.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 19]  [Cited by in F6Publishing: 32]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
27.  Hauser H, Dyer JH, Nandy A, Vega MA, Werder M, Bieliauskaite E, Weber FE, Compassi S, Gemperli A, Boffelli D. Identification of a receptor mediating absorption of dietary cholesterol in the intestine. Biochemistry. 1998;37:17843-17850.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 190]  [Cited by in F6Publishing: 189]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
28.  Kramer W, Girbig F, Corsiero D, Pfenninger A, Frick W, Jähne G, Rhein M, Wendler W, Lottspeich F, Hochleitner EO. Aminopeptidase N (CD13) is a molecular target of the cholesterol absorption inhibitor ezetimibe in the enterocyte brush border membrane. J Biol Chem. 2005;280:1306-1320.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 96]  [Cited by in F6Publishing: 98]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
29.  Altmann SW, Davis HR Jr, Zhu LJ, Yao X, Hoos LM, Tetzloff G, Iyer SP, Maguire M, Golovko A, Zeng M. Niemann-Pick C1 Like 1 protein is critical for intestinal cholesterol absorption. Science. 2004;303:1201-1204.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1297]  [Cited by in F6Publishing: 1244]  [Article Influence: 62.2]  [Reference Citation Analysis (0)]
30.  Lu K, Lee MH, Hazard S, Brooks-Wilson A, Hidaka H, Kojima H, Ose L, Stalenhoef AF, Mietinnen T, Bjorkhem I. Two genes that map to the STSL locus cause sitosterolemia: genomic structure and spectrum of mutations involving sterolin-1 and sterolin- 2, encoded by ABCG5 and ABCG8, respectively. Am J Hum Genet. 2001;69:278-290.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 281]  [Cited by in F6Publishing: 255]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
31.  Lee MH, Lu K, Hazard S, Yu H, Shulenin S, Hidaka H, Kojima H, Allikmets R, Sakuma N, Pegoraro R. Identification of a gene, ABCG5, important in the regulation of dietary cholesterol absorption. Nat Genet. 2001;27:79-83.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 438]  [Cited by in F6Publishing: 480]  [Article Influence: 20.9]  [Reference Citation Analysis (0)]
32.  Berge KE, Tian H, Graf GA, Yu L, Grishin NV, Schultz J, Kwiterovich P, Shan B, Barnes R, Hobbs HH. Accumulation of dietary cholesterol in sitosterolemia caused by mutations in adjacent ABC transporters. Science. 2000;290:1771-1775.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1238]  [Cited by in F6Publishing: 1115]  [Article Influence: 46.5]  [Reference Citation Analysis (0)]
33.  Salisbury BG, Davis HR, Burrier RE, Burnett DA, Bowkow G, Caplen MA, Clemmons AL, Compton DS, Hoos LM, McGregor DG. Hypocholesterolemic activity of a novel inhibitor of cholesterol absorption, SCH 48461. Atherosclerosis. 1995;115:45-63.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 87]  [Cited by in F6Publishing: 90]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
34.  Davis HR Jr, Pula KK, Alton KB, Burrier RE, Watkins RW. The synergistic hypocholesterolemic activity of the potent cholesterol absorption inhibitor, ezetimibe, in combination with 3-hydroxy-3-methylglutaryl coenzyme a reductase inhibitors in dogs. Metabolism. 2001;50:1234-1241.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 72]  [Cited by in F6Publishing: 73]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
35.  Van Heek M, France CF, Compton DS, McLeod RL, Yumibe NP, Alton KB, Sybertz EJ, Davis HR Jr. In vivo metabolism-based discovery of a potent cholesterol absorption inhibitor, SCH58235, in the rat and rhesus monkey through the identification of the active metabolites of SCH48461. J Pharmacol Exp Ther. 1997;283:157-163.  [PubMed]  [DOI]  [Cited in This Article: ]
36.  van Heek M, Farley C, Compton DS, Hoos L, Alton KB, Sybertz EJ, Davis HR Jr. Comparison of the activity and disposition of the novel cholesterol absorption inhibitor, SCH58235, and its glucuronide, SCH60663. Br J Pharmacol. 2000;129:1748-1754.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 245]  [Cited by in F6Publishing: 250]  [Article Influence: 10.4]  [Reference Citation Analysis (0)]
37.  van Heek M, Compton DS, Davis HR. The cholesterol absorption inhibitor, ezetimibe, decreases diet-induced hypercholesterolemia in monkeys. Eur J Pharmacol. 2001;415:79-84.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 105]  [Cited by in F6Publishing: 108]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
38.  Garcia-Calvo M, Lisnock J, Bull HG, Hawes BE, Burnett DA, Braun MP, Crona JH, Davis HR Jr, Dean DC, Detmers PA. The target of ezetimibe is Niemann-Pick C1-Like 1 (NPC1L1). Proc Natl Acad Sci USA. 2005;102:8132-8137.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 569]  [Cited by in F6Publishing: 549]  [Article Influence: 28.9]  [Reference Citation Analysis (0)]
39.  Iyer SP, Yao X, Crona JH, Hoos LM, Tetzloff G, Davis HR Jr, Graziano MP, Altmann SW. Characterization of the putative native and recombinant rat sterol transporter Niemann-Pick C1 Like 1 (NPC1L1) protein. Biochim Biophys Acta. 2005;1722:282-292.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 66]  [Cited by in F6Publishing: 51]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
40.  Davies JP, Levy B, Ioannou YA. Evidence for a Niemann-pick C (NPC) gene family: identification and characterization of NPC1L1. Genomics. 2000;65:137-145.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 173]  [Cited by in F6Publishing: 164]  [Article Influence: 6.8]  [Reference Citation Analysis (0)]
41.  Smart EJ, De Rose RA, Farber SA. Annexin 2-caveolin 1 complex is a target of ezetimibe and regulates intestinal cholesterol transport. Proc Natl Acad Sci USA. 2004;101:3450-3455.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 108]  [Cited by in F6Publishing: 112]  [Article Influence: 5.6]  [Reference Citation Analysis (0)]
42.  Valasek MA, Weng J, Shaul PW, Anderson RG, Repa JJ. Caveolin-1 is not required for murine intestinal cholesterol transport. J Biol Chem. 2005;280:28103-28109.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 57]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
43.  Kramer W, Corsiero D, Girbig F, Jähne G. Rabbit small intestine does not contain an annexin II/caveolin 1 complex as a target for 2-azetidinone cholesterol absorption inhibitors. Biochim Biophys Acta. 2006;1758:45-54.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 9]  [Cited by in F6Publishing: 9]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
44.  Graf GA, Li WP, Gerard RD, Gelissen I, White A, Cohen JC, Hobbs HH. Coexpression of ATP-binding cassette proteins ABCG5 and ABCG8 permits their transport to the apical surface. J Clin Invest. 2002;110:659-669.  [PubMed]  [DOI]  [Cited in This Article: ]
45.  Yu L, Hammer RE, Li-Hawkins J, Von Bergmann K, Lutjohann D, Cohen JC, Hobbs HH. Disruption of Abcg5 and Abcg8 in mice reveals their crucial role in biliary cholesterol secretion. Proc Natl Acad Sci USA. 2002;99:16237-16242.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 558]  [Cited by in F6Publishing: 529]  [Article Influence: 24.0]  [Reference Citation Analysis (0)]
46.  Plösch T, Bloks VW, Terasawa Y, Berdy S, Siegler K, Van Der Sluijs F, Kema IP, Groen AK, Shan B, Kuipers F. Sitosterolemia in ABC-transporter G5-deficient mice is aggravated on activation of the liver-X receptor. Gastroenterology. 2004;126:290-300.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 114]  [Cited by in F6Publishing: 117]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
47.  Klett EL, Lu K, Kosters A, Vink E, Lee MH, Altenburg M, Shefer S, Batta AK, Yu H, Chen J. A mouse model of sitosterolemia: absence of Abcg8/sterolin-2 results in failure to secrete biliary cholesterol. BMC Med. 2004;2:5.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 115]  [Cited by in F6Publishing: 117]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
48.  Yu L, Li-Hawkins J, Hammer RE, Berge KE, Horton JD, Cohen JC, Hobbs HH. Overexpression of ABCG5 and ABCG8 promotes biliary cholesterol secretion and reduces fractional absorption of dietary cholesterol. J Clin Invest. 2002;110:671-680.  [PubMed]  [DOI]  [Cited in This Article: ]
49.  Yu L, York J, von Bergmann K, Lutjohann D, Cohen JC, Hobbs HH. Stimulation of cholesterol excretion by the liver X receptor agonist requires ATP-binding cassette transporters G5 and G8. J Biol Chem. 2003;278:15565-15570.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 225]  [Cited by in F6Publishing: 223]  [Article Influence: 10.6]  [Reference Citation Analysis (0)]
50.  Cai L, Eckhardt ER, Shi W, Zhao Z, Nasser M, de Villiers WJ, van der Westhuyzen DR. Scavenger receptor class B type I reduces cholesterol absorption in cultured enterocyte CaCo-2 cells. J Lipid Res. 2004;45:253-262.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 39]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
51.  Mardones P, Quiñones V, Amigo L, Moreno M, Miquel JF, Schwarz M, Miettinen HE, Trigatti B, Krieger M, VanPatten S. Hepatic cholesterol and bile acid metabolism and intestinal cholesterol absorption in scavenger receptor class B type I-deficient mice. J Lipid Res. 2001;42:170-180.  [PubMed]  [DOI]  [Cited in This Article: ]
52.  Bietrix F, Yan D, Nauze M, Rolland C, Bertrand-Michel J, Coméra C, Schaak S, Barbaras R, Groen AK, Perret B. Accelerated lipid absorption in mice overexpressing intestinal SR-BI. J Biol Chem. 2006;281:7214-7219.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 106]  [Cited by in F6Publishing: 101]  [Article Influence: 5.6]  [Reference Citation Analysis (0)]
53.  Repa JJ, Turley SD, Lobaccaro JA, Medina J, Li L, Lustig K, Shan B, Heyman RA, Dietschy JM, Mangelsdorf DJ. Regulation of absorption and ABC1-mediated efflux of cholesterol by RXR heterodimers. Science. 2000;289:1524-1529.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1013]  [Cited by in F6Publishing: 966]  [Article Influence: 40.3]  [Reference Citation Analysis (0)]
54.  Mulligan JD, Flowers MT, Tebon A, Bitgood JJ, Wellington C, Hayden MR, Attie AD. ABCA1 is essential for efficient basolateral cholesterol efflux during the absorption of dietary cholesterol in chickens. J Biol Chem. 2003;278:13356-13366.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 94]  [Cited by in F6Publishing: 94]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
55.  Ohama T, Hirano K, Zhang Z, Aoki R, Tsujii K, Nakagawa-Toyama Y, Tsukamoto K, Ikegami C, Matsuyama A, Ishigami M. Dominant expression of ATP-binding cassette transporter-1 on basolateral surface of Caco-2 cells stimulated by LXR/RXR ligands. Biochem Biophys Res Commun. 2002;296:625-630.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 54]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
56.  Drobnik W, Lindenthal B, Lieser B, Ritter M, Christiansen Weber T, Liebisch G, Giesa U, Igel M, Borsukova H, Buchler C. ATP-binding cassette transporter A1 (ABCA1) affects total body sterol metabolism. Gastroenterology. 2001;120:1203-1211.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 104]  [Cited by in F6Publishing: 104]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
57.  McNeish J, Aiello RJ, Guyot D, Turi T, Gabel C, Aldinger C, Hoppe KL, Roach ML, Royer LJ, de Wet J. High density lipoprotein deficiency and foam cell accumulation in mice with targeted disruption of ATP-binding cassette transporter-1. Proc Natl Acad Sci USA. 2000;97:4245-4250.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 439]  [Cited by in F6Publishing: 444]  [Article Influence: 18.5]  [Reference Citation Analysis (0)]
58.  Buhman KK, Accad M, Novak S, Choi RS, Wong JS, Hamilton RL, Turley S, Farese RV. Resistance to diet-induced hypercholesterolemia and gallstone formation in ACAT2-deficient mice. Nat Med. 2000;6:1341-1347.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 283]  [Cited by in F6Publishing: 264]  [Article Influence: 11.0]  [Reference Citation Analysis (0)]
59.  Jones B, Jones EL, Bonney SA, Patel HN, Mensenkamp AR, Eichenbaum-Voline S, Rudling M, Myrdal U, Annesi G, Naik S. Mutations in a Sar1 GTPase of COPII vesicles are associated with lipid absorption disorders. Nat Genet. 2003;34:29-31.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 297]  [Cited by in F6Publishing: 303]  [Article Influence: 14.4]  [Reference Citation Analysis (0)]
60.  van der Veen JN, Kruit JK, Havinga R, Baller JF, Chimini G, Lestavel S, Staels B, Groot PH, Groen AK, Kuipers F. Reduced cholesterol absorption upon PPARdelta activation coincides with decreased intestinal expression of NPC1L1. J Lipid Res. 2005;46:526-534.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 143]  [Cited by in F6Publishing: 151]  [Article Influence: 7.6]  [Reference Citation Analysis (0)]
61.  Janowski BA, Willy PJ, Devi TR, Falck JR, Mangelsdorf DJ. An oxysterol signalling pathway mediated by the nuclear receptor LXR alpha. Nature. 1996;383:728-731.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1374]  [Cited by in F6Publishing: 1346]  [Article Influence: 48.1]  [Reference Citation Analysis (0)]
62.  Janowski BA, Grogan MJ, Jones SA, Wisely GB, Kliewer SA, Corey EJ, Mangelsdorf DJ. Structural requirements of ligands for the oxysterol liver X receptors LXRalpha and LXRbeta. Proc Natl Acad Sci USA. 1999;96:266-271.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 726]  [Cited by in F6Publishing: 734]  [Article Influence: 29.4]  [Reference Citation Analysis (0)]
63.  Repa JJ, Berge KE, Pomajzl C, Richardson JA, Hobbs H, Mangelsdorf DJ. Regulation of ATP-binding cassette sterol transporters ABCG5 and ABCG8 by the liver X receptors alpha and beta. J Biol Chem. 2002;277:18793-18800.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 624]  [Cited by in F6Publishing: 620]  [Article Influence: 28.2]  [Reference Citation Analysis (0)]
64.  Venkateswaran A, Repa JJ, Lobaccaro JM, Bronson A, Mangelsdorf DJ, Edwards PA. Human white/murine ABC8 mRNA levels are highly induced in lipid-loaded macrophages. A transcriptional role for specific oxysterols. J Biol Chem. 2000;275:14700-14707.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 313]  [Cited by in F6Publishing: 323]  [Article Influence: 13.5]  [Reference Citation Analysis (0)]
65.  Grefhorst A, Elzinga BM, Voshol PJ, Plösch T, Kok T, Bloks VW, van der Sluijs FH, Havekes LM, Romijn JA, Verkade HJ. Stimulation of lipogenesis by pharmacological activation of the liver X receptor leads to production of large, triglyceride-rich very low density lipoprotein particles. J Biol Chem. 2002;277:34182-34190.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 377]  [Cited by in F6Publishing: 375]  [Article Influence: 17.0]  [Reference Citation Analysis (0)]
66.  Plōsch T, Kok T, Bloks VW, Smit MJ, Havinga R, Chimini G, Groen AK, Kuipers F. Increased hepatobiliary and fecal cholesterol excretion upon activation of the liver X receptor is independent of ABCA1. J Biol Chem. 2002;277:33870-33877.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 145]  [Cited by in F6Publishing: 156]  [Article Influence: 7.1]  [Reference Citation Analysis (0)]
67.  Duval C, Touche V, Tailleux A, Fruchart JC, Fievet C, Clavey V, Staels B, Lestavel S. Niemann-Pick C1 like 1 gene expression is down-regulated by LXR activators in the intestine. Biochem Biophys Res Commun. 2006;340:1259-1263.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 129]  [Cited by in F6Publishing: 138]  [Article Influence: 7.7]  [Reference Citation Analysis (0)]
68.  Mel’nikov SM, Seijen ten Hoorn JW, Eijkelenboom AP. Effect of phytosterols and phytostanols on the solubilization of cholesterol by dietary mixed micelles: an in vitro study. Chem Phys Lipids. 2004;127:121-141.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 86]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
69.  Igel M, Giesa U, Lutjohann D, von Bergmann K. Comparison of the intestinal uptake of cholesterol, plant sterols, and stanols in mice. J Lipid Res. 2003;44:533-538.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 81]  [Cited by in F6Publishing: 83]  [Article Influence: 4.0]  [Reference Citation Analysis (0)]
70.  Plat J, van Onselen EN, van Heugten MM, Mensink RP. Effects on serum lipids, lipoproteins and fat soluble antioxidant concentrations of consumption frequency of margarines and shortenings enriched with plant stanol esters. Eur J Clin Nutr. 2000;54:671-677.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 148]  [Cited by in F6Publishing: 152]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
71.  Kaneko E, Matsuda M, Yamada Y, Tachibana Y, Shimomura I, Makishima M. Induction of intestinal ATP-binding cassette transporters by a phytosterol-derived liver X receptor agonist. J Biol Chem. 2003;278:36091-36098.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 148]  [Cited by in F6Publishing: 151]  [Article Influence: 7.2]  [Reference Citation Analysis (0)]
72.  Plat J, Mensink RP. Increased intestinal ABCA1 expression contributes to the decrease in cholesterol absorption after plant stanol consumption. FASEB J. 2002;16:1248-1253.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 130]  [Cited by in F6Publishing: 132]  [Article Influence: 6.0]  [Reference Citation Analysis (0)]
73.  Field FJ, Born E, Mathur SN. Stanol esters decrease plasma cholesterol independently of intestinal ABC sterol transporters and Niemann-Pick C1-like 1 protein gene expression. J Lipid Res. 2004;45:2252-2259.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 74]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
74.  Calpe-Berdiel L, Escolà-Gil JC, Ribas V, Navarro-Sastre A, Garcés-Garcés J, Blanco-Vaca F. Changes in intestinal and liver global gene expression in response to a phytosterol-enriched diet. Atherosclerosis. 2005;181:75-85.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 72]  [Cited by in F6Publishing: 75]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
75.  Kruit JK, Plosch T, Havinga R, Boverhof R, Groot PH, Groen AK, Kuipers F. Increased fecal neutral sterol loss upon liver X receptor activation is independent of biliary sterol secretion in mice. Gastroenterology. 2005;128:147-156.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 118]  [Cited by in F6Publishing: 127]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
76.  Cheng SH, Stanley MM. Secretion of cholesterol by intestinal mucosa in patients with complete common bile duct obstruction. Proc Soc Exp Biol Med. 1959;101:223-225.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 37]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
77.  Simmonds WJ, Hofmann AF, Theodor E. Absorption of cholesterol from a micellar solution: intestinal perfusion studies in man. J Clin Invest. 1967;46:874-890.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 141]  [Cited by in F6Publishing: 146]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
78.  Schwarz M, Russell DW, Dietschy JM, Turley SD. Marked reduction in bile acid synthesis in cholesterol 7alpha-hydroxylase-deficient mice does not lead to diminished tissue cholesterol turnover or to hypercholesterolemia. J Lipid Res. 1998;39:1833-1843.  [PubMed]  [DOI]  [Cited in This Article: ]
79.  Russell DW, Setchell KD. Bile acid biosynthesis. Biochemistry. 1992;31:4737-4749.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 602]  [Cited by in F6Publishing: 590]  [Article Influence: 18.4]  [Reference Citation Analysis (0)]
80.  Oude Elferink RP, Ottenhoff R, van Wijland M, Frijters CM, van Nieuwkerk C, Groen AK. Uncoupling of biliary phospholipid and cholesterol secretion in mice with reduced expression of mdr2 P-glycoprotein. J Lipid Res. 1996;37:1065-1075.  [PubMed]  [DOI]  [Cited in This Article: ]
81.  Smit JJ, Schinkel AH, Oude Elferink RP, Groen AK, Wagenaar E, van Deemter L, Mol CA, Ottenhoff R, van der Lugt NM, van Roon MA. Homozygous disruption of the murine mdr2 P-glycoprotein gene leads to a complete absence of phospholipid from bile and to liver disease. Cell. 1993;75:451-462.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1080]  [Cited by in F6Publishing: 1026]  [Article Influence: 33.1]  [Reference Citation Analysis (0)]
82.  Langheim S, Yu L, von Bergmann K, Lütjohann D, Xu F, Hobbs HH, Cohen JC. ABCG5 and ABCG8 require MDR2 for secretion of cholesterol into bile. J Lipid Res. 2005;46:1732-1738.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 78]  [Cited by in F6Publishing: 75]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
83.  Glickman RM, Green PH. The intestine as a source of apolipoprotein A1. Proc Natl Acad Sci USA. 1977;74:2569-2573.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 141]  [Cited by in F6Publishing: 138]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
84.  Wu AL, Windmueller HG. Relative contributions by liver and intestine to individual plasma apolipoproteins in the rat. J Biol Chem. 1979;254:7316-7322.  [PubMed]  [DOI]  [Cited in This Article: ]
85.  Roheim PS, Gidez LI, Eder HA. Extrahepatic synthesis of lipoproteins of plasma and chyle: role of the intestine. J Clin Invest. 1966;45:297-300.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 69]  [Cited by in F6Publishing: 76]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
86.  Bearnot HR, Glickman RM, Weinberg L, Green PH, Tall AR. Effect of biliary diversion on rat mesenteric lymph apolipoprotein-I and high density lipoprotein. J Clin Invest. 1982;69:210-217.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 40]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
87.  Forester GP, Tall AR, Bisgaier CL, Glickman RM. Rat intestine secretes spherical high density lipoproteins. J Biol Chem. 1983;258:5938-5943.  [PubMed]  [DOI]  [Cited in This Article: ]
88.  Green PH, Tall AR, Glickman RM. Rat intestine secretes discoid high density lipoprotein. J Clin Invest. 1978;61:528-534.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 210]  [Cited by in F6Publishing: 224]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
89.  Holt PR, Wu AL, Clark SB. Apoprotein composition and turnover in rat intestinal lymph during steady-state triglyceride absorption. J Lipid Res. 1979;20:494-502.  [PubMed]  [DOI]  [Cited in This Article: ]
90.  Iqbal J, Anwar K, Hussain MM. Multiple, independently regulated pathways of cholesterol transport across the intestinal epithelial cells. J Biol Chem. 2003;278:31610-31620.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 88]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
91.  Iqbal J, Hussain MM. Evidence for multiple complementary pathways for efficient cholesterol absorption in mice. J Lipid Res. 2005;46:1491-1501.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 67]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
92.  Bodzioch M, Orsó E, Klucken J, Langmann T, Böttcher A, Diederich W, Drobnik W, Barlage S, Büchler C, Porsch-Ozcürümez M. The gene encoding ATP-binding cassette transporter 1 is mutated in Tangier disease. Nat Genet. 1999;22:347-351.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1230]  [Cited by in F6Publishing: 1167]  [Article Influence: 46.7]  [Reference Citation Analysis (0)]
93.  Brooks-Wilson A, Marcil M, Clee SM, Zhang LH, Roomp K, van Dam M, Yu L, Brewer C, Collins JA, Molhuizen HO. Mutations in ABC1 in Tangier disease and familial high-density lipoprotein deficiency. Nat Genet. 1999;22:336-345.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1359]  [Cited by in F6Publishing: 1298]  [Article Influence: 51.9]  [Reference Citation Analysis (0)]
94.  Rust S, Rosier M, Funke H, Real J, Amoura Z, Piette JC, Deleuze JF, Brewer HB, Duverger N, Denèfle P. Tangier disease is caused by mutations in the gene encoding ATP-binding cassette transporter 1. Nat Genet. 1999;22:352-355.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1146]  [Cited by in F6Publishing: 1079]  [Article Influence: 43.2]  [Reference Citation Analysis (0)]
95.  Assmann G, Smootz E, Adler K, Capurso A, Oette K. The lipoprotein abnormality in Tangier disease: quantitation of A apoproteins. J Clin Invest. 1977;59:565-575.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 87]  [Cited by in F6Publishing: 96]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
96.  Wellington CL, Walker EK, Suarez A, Kwok A, Bissada N, Singaraja R, Yang YZ, Zhang LH, James E, Wilson JE. ABCA1 mRNA and protein distribution patterns predict multiple different roles and levels of regulation. Lab Invest. 2002;82:273-283.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 229]  [Cited by in F6Publishing: 227]  [Article Influence: 10.3]  [Reference Citation Analysis (0)]
97.  Haghpassand M, Bourassa PA, Francone OL, Aiello RJ. Monocyte/macrophage expression of ABCA1 has minimal contribution to plasma HDL levels. J Clin Invest. 2001;108:1315-1320.  [PubMed]  [DOI]  [Cited in This Article: ]
98.  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;22:630-637.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 313]  [Cited by in F6Publishing: 301]  [Article Influence: 13.7]  [Reference Citation Analysis (0)]
99.  van Eck M, Bos IS, Kaminski WE, Orsó E, Rothe G, Twisk J, Böttcher A, Van Amersfoort ES, Christiansen-Weber TA, Fung-Leung WP. Leukocyte ABCA1 controls susceptibility to atherosclerosis and macrophage recruitment into tissues. Proc Natl Acad Sci USA. 2002;99:6298-6303.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 286]  [Cited by in F6Publishing: 296]  [Article Influence: 13.5]  [Reference Citation Analysis (0)]
100.  Van Eck M, Singaraja RR, Ye D, Hildebrand RB, James ER, Hayden MR, Van Berkel TJ. Macrophage ATP-binding cassette transporter A1 overexpression inhibits atherosclerotic lesion progression in low-density lipoprotein receptor knockout mice. Arterioscler Thromb Vasc Biol. 2006;26:929-934.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 132]  [Cited by in F6Publishing: 136]  [Article Influence: 7.6]  [Reference Citation Analysis (0)]
101.  Basso F, Freeman L, Knapper CL, Remaley A, Stonik J, Neufeld EB, Tansey T, Amar MJ, Fruchart-Najib J, Duverger N. Role of the hepatic ABCA1 transporter in modulating intrahepatic cholesterol and plasma HDL cholesterol concentrations. J Lipid Res. 2003;44:296-302.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 180]  [Cited by in F6Publishing: 172]  [Article Influence: 8.2]  [Reference Citation Analysis (0)]
102.  Wellington CL, Brunham LR, Zhou S, Singaraja RR, Visscher H, Gelfer A, Ross C, James E, Liu G, Huber MT. Alterations of plasma lipids in mice via adenoviral-mediated hepatic overexpression of human ABCA1. J Lipid Res. 2003;44:1470-1480.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 74]  [Cited by in F6Publishing: 76]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
103.  Ragozin S, Niemeier A, Laatsch A, Loeffler B, Merkel M, Beisiegel U, Heeren J. Knockdown of hepatic ABCA1 by RNA interference decreases plasma HDL cholesterol levels and influences postprandial lipemia in mice. Arterioscler Thromb Vasc Biol. 2005;25:1433-1438.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 46]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
104.  Timmins JM, Lee JY, Boudyguina E, Kluckman KD, Brunham LR, Mulya A, Gebre AK, Coutinho JM, Colvin PL, Smith TL. Targeted inactivation of hepatic Abca1 causes profound hypoalphalipoproteinemia and kidney hypercatabolism of apoA-I. J Clin Invest. 2005;115:1333-1342.  [PubMed]  [DOI]  [Cited in This Article: ]
105.  Brunham LR, Kruit JK, Iqbal J, Fievet C, Timmins JM, Pape TD, Coburn BA, Bissada N, Staels B, Groen AK. Intestinal ABCA1 directly contributes to HDL biogenesis in vivo. J Clin Invest. 2006;116:1052-1062.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 411]  [Cited by in F6Publishing: 378]  [Article Influence: 21.0]  [Reference Citation Analysis (0)]
106.  Oliveira HC, Nilausen K, Meinertz H, Quintão EC. Cholesteryl esters in lymph chylomicrons: contribution from high density lipoprotein transferred from plasma into intestinal lymph. J Lipid Res. 1993;34:1729-1736.  [PubMed]  [DOI]  [Cited in This Article: ]
107.  Quintão EC, Drewiacki A, Stechhaln K, de Faria EC, Sipahi AM. Origin of cholesterol transported in intestinal lymph: studies in patients with filarial chyluria. J Lipid Res. 1979;20:941-945.  [PubMed]  [DOI]  [Cited in This Article: ]
108.  Joseph SB, McKilligin E, Pei L, Watson MA, Collins AR, Laffitte BA, Chen M, Noh G, Goodman J, Hagger GN. Synthetic LXR ligand inhibits the development of atherosclerosis in mice. Proc Natl Acad Sci USA. 2002;99:7604-7609.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 762]  [Cited by in F6Publishing: 725]  [Article Influence: 33.0]  [Reference Citation Analysis (0)]
109.  Terasaka N, Hiroshima A, Koieyama T, Ubukata N, Morikawa Y, Nakai D, Inaba T. T-0901317, a synthetic liver X receptor ligand, inhibits development of atherosclerosis in LDL receptor-deficient mice. FEBS Lett. 2003;536:6-11.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 266]  [Cited by in F6Publishing: 251]  [Article Influence: 12.0]  [Reference Citation Analysis (0)]