+ All Categories
Home > Documents > Intestinal absorption of long-chain fatty acid

Intestinal absorption of long-chain fatty acid

Date post: 09-Mar-2016
Category:
Upload: zhang-haichao
View: 229 times
Download: 4 times
Share this document with a friend
Description:
Keywords: Small intestine Long-chain fatty acid Lipid-binding protein Chylomicron Health 0163-7827/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.plipres.2009.01.001 Physiologie de la Nutrition, UMR Inserm U866, Ecole Nationale Supérieure de Biologie Appliquée à la Nutrition et à l’Alimentation (ENSBANA), Université de Bourgogne, 1, Esplanade Erasme, F-21000 Dijon, France Contents lists available at ScienceDirect Contents a r t i c l e i n f o
Popular Tags:
15
Review Intestinal absorption of long-chain fatty acids: Evidence and uncertainties Isabelle Niot * , Hélène Poirier, Thi Thu Trang Tran, Philippe Besnard * Physiologie de la Nutrition, UMR Inserm U866, Ecole Nationale Supérieure de Biologie Appliquée à la Nutrition et à l’Alimentation (ENSBANA), Université de Bourgogne, 1, Esplanade Erasme, F-21000 Dijon, France article info Article history: Received 4 December 2008 Received in revised form 19 December 2008 Accepted 6 January 2009 Keywords: Small intestine Long-chain fatty acid Lipid-binding protein Chylomicron Health abstract Over the two last decades, cloning of proteins responsible for trafficking and metabolic fate of long-chain fatty acids (LCFA) in gut has provided new insights on cellular and molecular mechanisms involved in fat absorption. To this systematic cloning period, functional genomics has succeeded in providing a new set of surprises. Disruption of several genes, thought to play a crucial role in LCFA absorption, did not lead to clear phenotypes. This observation raises the question of the real physiological role of lipid-binding pro- teins and lipid-metabolizing enzymes expressed in enterocytes. The goal of this review is to analyze pres- ent knowledge concerning the main steps of intestinal fat absorption from LCFA uptake to lipoprotein release and to assess their impact on health. Ó 2009 Elsevier Ltd. All rights reserved. Contents 1. Introduction ......................................................................................................... 102 2. Cellular LCFA uptake: diffusion and/or protein-mediated transfer? ............................................................. 102 2.1. From the physicochemical standpoint ............................................................................... 103 2.2. From the biochemical standpoint................................................................................... 103 2.3. From the physiological standpoint .................................................................................. 103 2.4. What functions for the plasma membrane lipid-binding proteins? ........................................................ 104 2.4.1. FABPpm ............................................................................................................... 104 2.4.2. FATP4 ................................................................................................................. 104 2.4.3. CD36 ................................................................................................................. 104 3. Intracellular trafficking of LCFA.......................................................................................... 105 3.1. Why two different FABPs? ........................................................................................ 105 3.2. Acyl-CoA-binding protein: an housekeeping protein ................................................................... 107 4. Chylomicron assembly and trafficking .................................................................................... 107 4.1. Re-esterification step ............................................................................................ 107 4.2. TAG transfer in the endoplasmic reticulum........................................................................... 108 4.3. Pre-chylomicron synthesis ........................................................................................ 109 4.4. Transfer of prechylomicrons to the Golgi ............................................................................ 109 4.5. Generation of intestinal lipid droplets ............................................................................... 109 5. Lessons learned from genomics.......................................................................................... 110 6. Do dietary lipids affect the intestinal function? ............................................................................. 110 0163-7827/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.plipres.2009.01.001 Abbreviations: ACBP, acyl-CoA-binding protein; ACS, acyl-CoA synthetase; BBM, brush border membrane; CM, chylomicron; DGAT, acyl-CoA diacylglycerol acyltransferase; ER, endoplasmic reticulum; FABPpm, plasma membrane fatty acid-binding protein; FACoA, long-chain-acyl-CoA esters; FAT/CD36, fatty acid transporter; FATP, fatty acid transport protein; FFA, free (non-esterified) fatty acids; I-FABP, intestinal fatty acid-binding protein; LBP, lipid-binding proteins; LCFA, long-chain fatty acids; L-FABP, liver fatty acid-binding protein; LPL, lipoprotein lipase; mAspAT, mitochondrial aspartate amino-transferase; MTP, microsomal triacylglycerol (triglyceride) transfer protein; PCTV, pre-chylomicron transfer vesicles; TAG, triacylglycerols. * Corresponding authors. Tel.: +33 (0) 3 80 39 91 24; fax: +33 (0) 3 80 39 66 91. E-mail addresses: [email protected] (I. Niot), [email protected] (P. Besnard). Progress in Lipid Research 48 (2009) 101–115 Contents lists available at ScienceDirect Progress in Lipid Research journal homepage: www.elsevier.com/locate/plipres
Transcript
Page 1: Intestinal absorption of long-chain fatty acid

Progress in Lipid Research 48 (2009) 101–115

Contents lists available at ScienceDirect

Progress in Lipid Research

journal homepage: www.elsevier .com/locate /p l ipres

Review

Intestinal absorption of long-chain fatty acids: Evidence and uncertainties

Isabelle Niot *, Hélène Poirier, Thi Thu Trang Tran, Philippe Besnard *

Physiologie de la Nutrition, UMR Inserm U866, Ecole Nationale Supérieure de Biologie Appliquée à la Nutrition et à l’Alimentation (ENSBANA),Université de Bourgogne, 1, Esplanade Erasme, F-21000 Dijon, France

a r t i c l e i n f o

Article history:Received 4 December 2008Received in revised form 19 December 2008Accepted 6 January 2009

Keywords:Small intestineLong-chain fatty acidLipid-binding proteinChylomicronHealth

0163-7827/$ - see front matter � 2009 Elsevier Ltd. Adoi:10.1016/j.plipres.2009.01.001

Abbreviations: ACBP, acyl-CoA-binding protein;acyltransferase; ER, endoplasmic reticulum; FABPpmFATP, fatty acid transport protein; FFA, free (non-esterL-FABP, liver fatty acid-binding protein; LPL, lipoproteprotein; PCTV, pre-chylomicron transfer vesicles; TAG

* Corresponding authors. Tel.: +33 (0) 3 80 39 91 2E-mail addresses: [email protected] (I. Niot), pb

a b s t r a c t

Over the two last decades, cloning of proteins responsible for trafficking and metabolic fate of long-chainfatty acids (LCFA) in gut has provided new insights on cellular and molecular mechanisms involved in fatabsorption. To this systematic cloning period, functional genomics has succeeded in providing a new setof surprises. Disruption of several genes, thought to play a crucial role in LCFA absorption, did not lead toclear phenotypes. This observation raises the question of the real physiological role of lipid-binding pro-teins and lipid-metabolizing enzymes expressed in enterocytes. The goal of this review is to analyze pres-ent knowledge concerning the main steps of intestinal fat absorption from LCFA uptake to lipoproteinrelease and to assess their impact on health.

� 2009 Elsevier Ltd. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1022. Cellular LCFA uptake: diffusion and/or protein-mediated transfer? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

2.1. From the physicochemical standpoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1032.2. From the biochemical standpoint. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1032.3. From the physiological standpoint . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1032.4. What functions for the plasma membrane lipid-binding proteins?. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

2.4.1. FABPpm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1042.4.2. FATP4. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1042.4.3. CD36 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

3. Intracellular trafficking of LCFA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

3.1. Why two different FABPs? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1053.2. Acyl-CoA-binding protein: an housekeeping protein . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

4. Chylomicron assembly and trafficking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

4.1. Re-esterification step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1074.2. TAG transfer in the endoplasmic reticulum. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1084.3. Pre-chylomicron synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1094.4. Transfer of prechylomicrons to the Golgi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1094.5. Generation of intestinal lipid droplets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

5. Lessons learned from genomics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1106. Do dietary lipids affect the intestinal function?. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

ll rights reserved.

ACS, acyl-CoA synthetase; BBM, brush border membrane; CM, chylomicron; DGAT, acyl-CoA diacylglycerol, plasma membrane fatty acid-binding protein; FACoA, long-chain-acyl-CoA esters; FAT/CD36, fatty acid transporter;ified) fatty acids; I-FABP, intestinal fatty acid-binding protein; LBP, lipid-binding proteins; LCFA, long-chain fatty acids;in lipase; mAspAT, mitochondrial aspartate amino-transferase; MTP, microsomal triacylglycerol (triglyceride) transfer, triacylglycerols.4; fax: +33 (0) 3 80 39 66 [email protected] (P. Besnard).

Page 2: Intestinal absorption of long-chain fatty acid

FaceabL-tr

102 I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115

7. Intestinal contribution to dyslipidemia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1118. Conclusions and future directions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

1. Introduction

Lipids account for about 40% of the calories ingested in Westerncountries, whereas nutritional recommendations are 5–10% lower.This excessive lipid intake, associated with a qualitative imbalance(excess of saturated fatty acids and cholesterol, too high n�6/n�3ratio) strongly favours the development of obesity and associateddiseases (atherosclerosis, non insulin-dependent diabetes, hyper-tension, cancer). From organs involved in lipid homeostasis, smallintestine remains the most poorly known likely because it has beenonly considered for a long time as a selective barrier between out-side and inside body environment.

In human diet, around 95% of dietary lipids are triacylglycerols(TAG), mainly composed of long-chain fatty acids (LCFA, number ofcarbons � 16), the remaining being phospholipids (4.5%) and ster-ols. Because TAG cannot cross cellular membranes, they must behydrolyzed before their subsequent metabolic use. Acid-stable gas-tric lipase partially digests TAG to form diacylglycerols and freefatty acids (FFA) in the stomach. This low lipase activity is essentialfor an efficient lipid emulsification [1]. Digestion of TAG continuesin the small intestine mainly through the action of a colipase-dependent pancreatic lipase and leads to the release of 2-monoa-cylglycerols and LCFA (for a review, see [2]). In contrast to otherenergetic nutrients, LCFA are poorly soluble in aqueous solution.Moreover, they exhibit detergent properties potentially harmfulfor cellular integrity. To overcome these limitations, LCFA are suc-cessively dispersed into mixed micelles in intestinal lumen, boundto soluble lipid-binding proteins (LBP) in intestinal absorptive cellsand, after re-esterification, are secreted into lymph as TAG-rich

lymph

Dietary lipids

TG PL CE/CS

Biliary lipidsBA PL

CM

HYDROLYSIS

CM,VLDLnascent HDL

enterocyte

CM

remnants

LPL

AdipocyteTG

LCFA

myocyte

LCFA

ATP

blood

VLDL

blood hepatocyte

HLLCFA

ATP

LRP

lymph

Dietary lipids

TG PL CE/CS

Biliary lipidsBA PL

CM

HYDROLYSIS

CM,VLDLnascent HDL

enterocyte

HYDROLYSIS

CM,VLDLnascent HDL

enterocyte

CM

remnants

LPL

AdipocyteTG

LCFA

myocyte

LCFA

ATP

blood

VLDL

blood hepatocyte

HLLCFA

ATP

LCFA

ATP

LRP

Unw

Unw

A B

TAG

ig. 1. The metabolic fate of dietary lipid in the body. (A) Circulation and metabolic fate obsorption are depicted: (1) micellar dissociation due to LCFA protonation mediated byllular uptake, (3) intracellular trafficking involving soluble lipid-binding proteins and

cyl-CoA-binding protein; ACS, acyl-CoA synthetases; CS, cholesterol; ER, endoplasmic reinding protein; FAH, protonated long-chain fatty acids; FATP4, fatty acid transport proFABP, liver fatty acid-binding protein; HL, hepatic lipase; LPL, lipoprotein lipase; Liacylglycerol (triglycerides); PL, phospholipids; CE, cholesterol esters; VLDL, very low

lipoproteins (Fig. 1A). During the post-prandial period, the smallintestine produces and secretes chylomicrons (CM), whereas verylow density lipoproteins (VLDL) are mainly synthesized duringthe interprandial periods. Once in the blood, TAG-rich lipoproteinsare progressively hydrolyzed by the endothelial lipoprotein lipase(LPL) providing LCFA to peripheral tissues (muscles and adiposetissue). The TAG remaining in resulting small lipoproteins (rem-nants) are further hydrolyzed by the hepatic lipase; then remnantsare cleared from blood by liver mainly via the LDL receptor-relatedpeptide (Fig. 1A).

LCFA exert basic functions in the cell as membrane components,metabolic fuel, precursors of lipid mediators, regulators of ion-channels and modulators of gene expression [3]. They are also in-volved in various post-translational modifications of proteins (e.g.palmitoylation) affecting their cellular functions [4]. Therefore,digestion and absorption of dietary fat must be highly efficient toensure a correct LCFA supply to the body. Nevertheless, LCFA beinghydrophobic nutrients, their intestinal absorption remains com-plex. For didactic reason, it is classically depicted in three succes-sive steps: cellular uptake, intracellular trafficking andlipoprotein synthesis/release (Fig. 1B).

2. Cellular LCFA uptake: diffusion and/or protein-mediatedtransfer?

LCFA transfer through the plasma membrane is a highly contro-versial question at the origin of several reviews [5–11]. By reasonof their physicochemical properties, it was thought for a long timethat LCFA uptake by cells only took place by diffusion. However,

FAH

Lymph

ACBP

ACS

AG-CoA

TG, PL

ChylomicronsVLDL

Nascent HDL

Uptake

trafficking

Lipoproteinsynthesis

ABSORPTION

stirredater

layer

pH

> 6 Mixedmicelles

FA-

Micellardissociation

I-FABPL-FABP FA-

< 5H +

Na+

ER/golgi

passivediffusion

FABPpmFATP4

FAT/CD36

AG-CoA

MTP

FAH

Lymph

ACBP

ACS

AG-CoA

TG, PL

ChylomicronsVLDL

Nascent HDL

Uptake

trafficking

Lipoproteinsynthesis

ABSORPTION

stirredater

layer

pH

> 6 Mixedmicelles

FA-

Micellardissociation

I-FABPL-FABP FA-

< 5H +

Na+

ER/golgi

passivediffusion

FABPpmFATP4

FAT/CD36

AG-CoA

MTP

FACoA FACoA

TAG, PL, CE

Micellar dissociation

CM

f dietary lipid in the body. (B) The main steps and players involved in intestinal LCFAthe acidic microclimate lining the brush border membrane of enterocytes, (2) LCFA(4) triacylgycerol-rich lipoprotein synthesis and exocytosis into the lymph. ACBP,

ticulum; FA�, ionized long-chain fatty acids; FABPpm, plasma membrane fatty acid-tein 4; HDL, high density lipoprotein; I-FABP, intestinal fatty acid-binding protein;RP, LDL-related peptide; MTP, microsomal triacylglycerol transfer protein; TAG,

density lipoproteins.

张海超
Sticky Note
在人类饮食中,大约95%的饮食脂肪是三酰甘油(TAG),主要由长链脂肪酸(LCFA碳原子数大于16个)组成,剩余的是4.5%的磷脂和甾醇。因为TAG不能通过细胞膜,他们在被新陈代谢连续的利用前要水解。酸稳定的胃脂酶部分的消化TAG在胃中形成自由脂肪酸FFA和甘油二酯。这个较低的脂酶活性对有效的脂类乳化是必需的。在小肠中继续消化TAG主要是通过共脂肪酶依赖的胰腺脂肪酶的作用并导致2-monoa-cylglycerols与LCFA的释放。与其他营养物质相比较,LCFA在水中的溶解量较小。而且,它们有洗涤的特性可能会损伤细胞的完整性。为了克服这些限制,连续的将LCFA散布入小肠中的混合物中,结合到小肠吸收细胞的可溶的脂结合蛋白(LBP)上,经过再酯化后,作为富含TAG的脂蛋白分泌到淋巴中。在餐后的时期,小肠产生和分泌乳糜颗粒(CM),然而VLDL主要在饮食期间合成。一旦进入血液,富含TAG的脂蛋白逐渐的被内皮的脂蛋白脂酶逐渐的水解,为外周组织提供LCFA。剩余的TAG导致小脂蛋白被肝脏脂肪酶进一步的水解;然后血液中的小脂蛋白被肝脏通过LDL受体相关的肽清除了。 LCFA在细胞中的基本功能是组成膜,新陈代谢能量,脂类调节因子的前体,离子通道的调节因子和基因表达的调节器。LCFA也与多种蛋白质翻译后的修饰,改变它们的细胞功能有关。因此,消化和吸收脂肪必须高效的保持一个正确的身体LCFA供应。无论如何,LCFA是不溶于水的养分,它们的小肠吸收很复杂。为了教学,它被经典的描述为3个连续的步骤:细胞摄取,细胞内运输和脂蛋白合成/释放。 2。细胞LCFA摄取:被动扩散和/或蛋白介导的运输? LCFA穿过质膜的运输是一个有争议的问题。根据它们的物理化学特性,人们长期认为LCFA只通过被动扩散运输。
Page 3: Intestinal absorption of long-chain fatty acid

I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115 103

this concept has been challenged over the two last decades by bio-chemical approaches suggesting an involvement of a protein-med-iated transfer. It is likely that an efficient LCFA uptake by cellsrequires both spontaneous and facilitated transfer. However, therelative importance of these two mechanisms appears to be highlydependent from the microenvironment and phenotype of lipid-uti-lizing tissues. Thus, a direct extrapolation of a lipid-transfer modelfrom one to another tissue may constitute physiological nonsense.

2.1. From the physicochemical standpoint

Simple models, as phospholipid vesicles, have been largely usedto study the transbilayer movement of native or modified fattyacids in vitro, with or without albumin used as a donor molecule.This transfer can be viewed as three successive steps: adsorptionof LCFA upon the membrane surface, ‘‘flip–flop” movement fromexternal hemi-leaflet of bilayer to internal hemi-leaflet anddesorption from internal bilayer into the inner of vesicle. By reasonof their lipophilicity, LCFA adsorption is thought to be thermody-namically favourable and extremely fast [8,12]. Data on flip–flopstep and, in a lesser extent, desorption are more controversial.Using the pH-sensitive fluorophore pyranin trapped into lipo-somes, Kamp and co-workers report a fast movement of free fattyacids (FFA) across phospholipid bilayer. A flip–flop rate shorterthan 20 ms, independent of fatty acid chain length, was found.Interestingly, the transmembrane movement of anionic form ofFFA was several orders of magnitude slower than that of proton-ated ones [13,14]. Altogether these data strongly suggest thatFFA can spontaneously cross the cell surface especially when theyare protonated. In this model, flip–flop is faster than desorption[13]. Conversely, other studies indicate that the flip–flop, but notthe desorption step, is rate-limiting (for review see [11]). Indeed,half times found for FFA transbilayer transfer were within a70 ms to 10 s range in function of the fatty acid type, vesicle sizeand temperature. It was concluded that the time course of nativeFFA through the phospholipid bilayer might be insufficient to sup-port the metabolic activity in cells known to have high require-ments in LCFA, as cardiomyocytes. According to theseobservations, a model for LCFA transfer across a phospholipid bi-layer has been recently proposed [15]. Since the pKa of LCFA islower than 5, more than 99% of LCFA are ionized at physiologicalpH [6]. Therefore, LCFA insert their hydrophobic carbon backbonefirst inside the external hemi-leaflet of membrane, the ionized car-boxyl head remaining in contact with the lipid/water interface. Tocross the membrane, LCFA must undergo a rotation of 180� toreorient its carboxyl group toward the aqueous phase lining theinternal hemi-leaflet. In this model, LCFA rotation is a rate-limitingstep since it is dependent on both formation of a free volume insidethe phospholipid bilayer and LCFA folding. These steric limitationsmight explain why the efficiency of flip–flop appears to be tightlydependent of membrane curvature. Indeed, flip–flop half times areinversely correlated with the vesicle size, LCFA crossing small ves-icles more rapidly than larger ones [15–17]. This size dependencecan be accounted for by a larger free volume inside the phospho-lipid bilayer when the radius of curvature is small [15]. Efficiencyof the flip–flop is also dependent of the LCFA itself, unsaturatedfatty acids being transferred faster than saturated ones. It has beensuggested that the folded conformation of unsaturated fatty acidsfacilitates their rotation inside to the membrane [15]. Altogetherthese data indicate that protonated LCFA can easily cross phospho-lipid bilayers.

2.2. From the biochemical standpoint

Liposomes are useful models to dissect FFA fluxes throughphospholipid bilayers in a controlled environment. However, the

simplicity of this model and the lack of subsequent LCFA metabolicuse constitute major limitations of this technical approach. In iso-lated enterocytes, as well as in the enterocyte-like Caco-2 cell line,transport of LCFA in both apical and basolateral membranes ap-pears to be saturable, when low LCFA concentrations are used[18,19]. A competitive inhibition by structurally related LCFA hasalso been reported in Caco-2 cells [20,21]. These data are consis-tent with an involvement a protein-facilitated process in intestinalLCFA permeation (Fig. 1B). Identification of proteins exhibiting ahigh affinity for LCFA in the brush border membrane (BBM) ofenterocyte is in agreement with this assumption. Therefore, ithas been concluded that the intestinal LCFA uptake is likely a pro-tein-mediated event as reported for cardiac and skeletal muscles[22].

2.3. From the physiological standpoint

In contrast to other lipid-utilizing cells (i.e. adipocytes, myo-cytes and hepatocytes), intestinal absorptive cells (enterocytes)are daily subjected to dramatic changes in fat supply. Adequateluminal environment and morphological adaptations explain whythe intestinal fat absorption remains efficient despite of this chal-lenge. Firstly, the unstirred water layer lining BBM constitutes aunique microclimate which promotes the LCFA permeation(Fig. 1B). It is produced by the trapping of water molecules into acomplex glycoprotein network formed by mucus and glycocalyx.With a thickness from 50 to 500 lm [23], this weak renewal com-partment is characterized by a low pH gradient generated by anefficient H+/Na+ antiport exchange system located in the BBM[24,25]. By reason of their hydrophobicity, LCFA begin to cross thisaqueous diffusion barrier into mixed micelles which increase 100–1000 fold their aqueous concentration [26]. Since LCFA are ionizedat physiological pH in aqueous solution, the low pH microclimatefound near BBM induces their massive protonation as soon as localpH becomes lower than LCFA pKa. This event is essential for an effi-cient fat absorption. Indeed, by reducing the LCFA solubility intomixed micelles, protonation induces the release of LCFA near themicrovilli of absorptive cells [27]. Moreover, protonation facilitatestheir subsequent cellular uptake mainly by diffusion since proton-ated LCFA are known to have a greater membrane permeationcapacity than their corresponding ionized species [13,28]. The factthat pharmacological inhibition of the H+/Na+ antiporter by thediuretic amyloride decreases in dose-dependent manner LCFA up-take in rat jejunal sheets is consistent with this assumption [29]. Itis noteworthy that this acidic microclimate is not found in other li-pid-utilizing tissues. Secondly, the presence of microvilli in the api-cal side of enterocytes dramatically enhances their uptakecapacity. Moreover, the small membrane curvature found in thetop of microvilli is favourable to a fast LCFA flip–flop in thephospholipid bilayer [15–17]. Altogether these specificities explainwhy intestinal LCFA uptake is not limited during a lipid load (e.g.during the post-prandial period). Therefore, extrapolation of up-take data found in the muscles, adipose tissue or liver to the smallintestine is physiologically irrelevant. The same limitation also ex-ists for intestinal cell lines or isolated enterocytes which do notfully reproduce this complex extracellular microclimate and onlydisplay a weak density of small microvilli.

In brief, passive diffusion seems to play a significant role inLCFA uptake in the small intestine. This high capacity and lowaffinity system appears especially well adapted to intestinal chal-lenges since it remains efficient during the post-prandial period.Indeed, lipid fecal loss remains below 5% (w/w) even in situationof high fat supply in healthy humans, [30]. Therefore, lipid uptakedoes not appear to be a rate-limiting step for the intestinal fatabsorption, in contrast to what it is reported for other lipid-utiliz-ing tissues. This conclusion raises the question of the physiological

张海超
Sticky Note
然而,这个概念被最近20年的生物化学方法研究显示的有一个蛋白介导的运输所挑战。有效的LCFA细胞吸收需要自发的和协助的运输。然而,这两种机制的相关的重要性似乎高度依赖于微环境和利用脂类组织的表型。因此,一个直接的解释脂类从一个组织运输到另一个组织的可能构成生理学上的无差异。 2.1从物理化学观点看: 简单的模型,如磷脂小泡,以及被大量的利用来研究内源的和体内修改的脂肪酸的跨膜运输,利用或不利用白蛋白作为分子的捐赠者。这个运输可以被认为是3步连续的步骤:吸收LCFA到膜表面,“触发器”移动从双分子膜的外侧到双分子膜的内侧,双分子膜内侧的解吸附释放到囊泡中。由于他们的亲脂性,LCFA的吸收有助于热力学并且很快。触发器步骤的数据和解吸附在一个次要的范围内有争议。利用PH敏感的荧光pyranin捕获脂质体,kamp和同事的报的到一个FFA快速通过磷脂双分子层。触发器的比率小于20ms,与脂肪酸的长度无关。有趣的是,阴离子形式的FFA跨膜运输比阳离子慢几个数量级。总的考虑这些数据,显示FFA在质子化时可以自发的穿过细胞表面。在这种形势下,触发器快于解吸附作用。与此相反的是,有些研究显示触发器是受速率限制的。确实,一半的FFA跨膜运输在70ms到10s范围内是以脂肪酸的形式,囊泡的大小和温度。可以推断出内源的FFA跨越磷脂双分子层用的时间可能不足以支持需要LCFA的细胞如心肌细胞内新陈代谢的活性。根据这些观察,最近提出了一个LCFA穿过磷脂双分子层的模型。因为LCFA的pka值小于5,超过99%的LCFA在生理ph下都是离子。因此,LCFA将它们的疏水碳端首先进入双层膜的外层,质子化的羧基端仍然接触脂质/水界面。为了穿过膜,LCFA需要转过180°来使它的羧基进入内膜的水相。在这个模型中,LCFA的旋转是限速步骤,因为它依赖于磷脂双分子层的内部空间题解和LCFA的折叠。这些空间品类限制可以解释为什么触发器的效率似乎紧密的依赖于膜的曲度。确实,触发器在一半的情况下与囊泡体积有相反的关系,LCFA通过小囊泡更快。对体积的依赖性可以用当磷脂双分子层曲度的半径较小时有较大的内部体积来解释。触发器的效率还与LCFA有关,不饱和FA运输速度更快。折叠的非饱和FA的结构促进在膜内的旋转。这些数据显示质子化的LCFA可以较容易的穿过磷脂双分子层。 2.2从生物化学的观点 脂质体对仔细研究FFA在一个控制的环境中通过磷脂双分子层的流量很有用。然而,这个模型的简单和缺少随后的新陈代谢利用的LCFA组成了这个技术的主要限制因素。在分离的肠细胞中,当用低浓度的LCFA时,LCFA在绒毛顶端和刷装膜缘的运输似乎是饱和的。一个LCFA结构上的竞争性抑制剂在caco2细胞中发现。这些数据与小肠LCFA渗透的蛋白协助运输一致。蛋白质的鉴定显示出一个高度的LCFA相关,小肠刷状缘膜支持这一假设。因此,我们可以推论出小肠LCFA的吸收类似于一个蛋白介导的事件,如心脏和肌肉中报道的。
张海超
Sticky Note
2.3从生理学角度 与其它利用脂肪的细胞相比,小肠吸收细胞受脂肪供应的影响更大。调节内环境和形态学改变解释来了为什么小肠脂肪吸收在这种挑战下仍有较高的效率。首先,BBM表面的不动的水相组成了一个特殊的内环境,促进LCFA的渗透。这是由粘液和多糖蛋白复合物形成的复杂糖蛋白网将水分子围在里面产生的。厚度为50-500um,这个新产生的隔层的特点为:存在一个有BBM上的高效的H/Na双向运输交换系统产生的较低的pH变化率。由于它们的疏水性,LCFA开始穿过这个水相层到微团内,增加了他们100-1000的水浓度。因为LCFA在生理pH下的水溶液中质子化,在BBM附近发现的较低的pH微环境一旦pH低于LCFA的pKa就诱导它们质子化。这对脂肪吸收是很重要的。确实,通过减少LCFA溶入微团中的数量,质子化诱导LCFA在绒毛顶端吸收细胞附近的释放。而且,质子化使它们随后被细胞吸收主要通过扩散,因为质子化的LCFA有很强的穿透膜的能力。药理学抑制H/Na的反向转运载体减少了依赖于剂量的LCFA在空肠的摄取与这个假设一致。值得注意的一点是酸性微环境没有在其它脂肪利用组织中发现。第二,肠顶端出现微绒毛直接加强了它们的摄取能力。而且,微绒毛顶端小的膜曲度对快速在磷脂双分子层上LCFA的翻转有利。这些特点解释了为什么LCFA的吸收在脂质负载时不受限制。因此,由肌肉中得到的数据推断,小肠或肝脏的脂肪组织生理学上有一定的相关性。小肠细胞和分离的小肠中未完全再生这个复杂的细胞外微环境和较低的微绒毛密度时也有相同的限制存在。
张海超
Sticky Note
简单来说,自由扩散在LCFA的吸收中有重要作用。高能力和低亲和力的系统在小肠中调节的最好,因为它在采食后仍保持高效。确实,脂肪排泄物损失保持在5%以下,即使是在给健康人供给高脂肪是。因此,与其他报道的脂肪利用组织相比,脂质的吸收不是小肠脂肪吸收的限速步骤。这个结论提出了一个问题,即细胞膜脂肪结合蛋白在小肠中的作用。
Page 4: Intestinal absorption of long-chain fatty acid

104 I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115

role played by the plasma membrane lipid-binding proteins foundin enterocytes.

2.4. What functions for the plasma membrane lipid-binding proteins?

Three proteins displaying a high affinity for LCFA have been iso-lated in the BBM of enterocytes: the plasma membrane-associatedfatty acid-binding protein (FABPpm) [31], the fatty acid transportprotein 4 (FATP4) [32] and the fatty acid transporter (CD36) [33].

2.4.1. FABPpmFABPpm is a 43 kDa peripheral-associated membrane protein

expressed in organs with high lipid requirements [34]. It is thoughtto adhere to the membrane through a specific N-terminal peptide[35]. A body of evidence supports a role of FABPpm in cellular LCFAuptake. FABPpm can bind LCFA with an apparent dissociation con-stant (kd) of 80 nM [36]. In vitro use of polyclonal antibodies raiseagainst FABPpm leads to a drop in LCFA uptake in hepatocytes [37],adipocytes [38], cardiac myocytes [39] and heart giant vesicles[40]. Targeted-overexpression of FABPpm in rat soleus muscle byusing electroporation of the FABPpm expression vector leads toan increase in palmitate uptake [41].

In the small intestine, FABPpm expression is found in jejunumand, in a lesser extent, in ileum. Albeit plausible, implication ofFABPpm in intestinal lipid absorption remains questionable.Preincubation of jejunal explants with a mono-specific FABPpmantibody only results in a partial, but significant, inhibition of[3H]-oleate uptake [18]. Nevertheless, large amounts of antiserumwere required to achieve such an inhibition. Moreover, the factthat FABPpm is also found in cryptic cells not involved in the fatabsorption raises a doubt about a physiological role of this proteinin the intestinal LCFA uptake [42,43]. A specific involvement in cel-lular LCFA transfer is also challenged because the peptidic se-quence of FABPpm is identical to the mitochondrial aspartateamino-transferase (mAspAT). This enzyme, found in inner mito-chondrial membrane, catalyzes the reaction of transaminationlinking the ureogenesis pathway to the Krebs cycle [44]. Even if re-sults obtained during (i) fasting in red skeletal muscles [45], (ii)diabetes in adipocytes from Zucker rats [46] or (iii) ethanol loadin human hepatoma HepG2 cells [47] are consistent with aninvolvement of FABPpm/mAspAT in the FA metabolism, its precisephysiological role in intestinal fat absorption remains elusive. Gen-eration of FABPpm null mice might provide new insights on thephysiological role played by this plasma membrane LBP in intesti-nal lipid absorption.

2.4.2. FATP4Fatty acid transport proteins (FATP) are 63 kDa proteins firstly

identified in 3T3-L1 preadipocytes by expression cloning strategyon the basis of their facilitation to uptake LCFA [32]. Five and sixdifferent isoforms of FATP are found in rodents and humans,respectively [48]. Each FATP displays a specific pattern of expres-sion [48]. For example, FATP5 and FATP2 are highly expressed inthe liver [49], while FATP1 is essentially found in the adipose tissueand FATP4 in the gut [48]. Interestingly, the intestinal location ofFATP4 correlated quite well with fat absorption. Indeed, its expres-sion level is especially high in the jejunum and, to a lesser extent,in the duodenum, but is lacking in the colon [48]. Moreover, theexpression of FATP4 was initially found in BBM of mature entero-cytes located in the upper side of villi. In contrast, it was low orlacking in undifferentiated cryptic cells [48]. In vitro studiesstrongly also support an involvement of FATP4 in the LCFA uptake.Indeed, mediated-depletion of FATP4 protein by antisense strategyleads to a significant decrease in LCFA uptake in freshly isolatedenterocytes [48]. Similarly, a 40% decrease in the fatty acid uptakeis reported in enterocytes isolated from FATP4+/� mice which

display 48% reduction of FATP4 protein [50]. Finally, FATP4 antag-onists decrease LCFA uptake in cells stably transfected with aFATP4 expression vector [51]. The mechanism by which FATPsfacilitate the cellular fatty acid transfer is not yet fully established.Since FATP proteins do not contain a putative LCFA-binding site, arole as fatty acid pool former upon the surface of membranesseems unlikely. In contrast, a high amino-acid sequence identitybetween acyl-CoA synthetases (ACS) and FATPs, as well as a com-parable predictive structure including a highly conserved ATP-binding motif, strongly suggests that FATPs might be involved infatty acid acylation with a preferential substrate specificity forLCFA and very LCFA [52,53]. The fact that membrane extracts ofCos1 cells transfected with a murine FATP4 expression vector dis-play a high ACS activity is consistent with this assumption [54].Such a function might be crucial. Indeed, acylation of LCFA is abso-lutely required for their subsequent metabolic use. Moreover, plas-ma membranes are impermeable to fatty acyl-CoA (FA-CoA).Therefore, it has been proposed that FATP4 contributes to uptakeof LCFA by trapping them as CoA derivatives [10,53,55]. In brief,FATP4 might play an indirect role in intestinal lipid uptake througha vectorial acylation of LCFA. Nevertheless, this hypothesis hasbeen recently challenged by the demonstration that FATP4 israther expressed in the endoplasmic reticulum (ER) membranethan in BBM [55,56]. Moreover, the physiological impact of FATP4on fat absorption remains also elusive in vivo. FATP4 inhibitors donot display any effects on fat absorption in wild-type mice [51].Similarly, no change in lipid fecal elimination and weight gain isfound in FATP4+/� mice subjected to standard or high fat diets[50]. Unfortunately, deletion of both FATP4 alleles results inembryonic lethality by reason of dramatic perinatal skin defect[50]. To produce viable FATP4�/�mice and explore further its phys-iological role in intestinal fat absorption, the skin phenotype hasbeen recently rescued using an FATP4 transgene driven by a kerat-inocyte-specific promoter [57]. These FATP4-null mice and theirwild-type littermates display indistinguishable intestinal TAGabsorption or fecal fat output when they are fed with a standardchow. The lack of effect being not due to compensatory inductionof other LBP, it has been concluded that FATP4 is dispensable forfat absorption in the mouse. However, an increase in the entero-cyte TAG levels was found when FATP4�/� mice were subjectedto a high fat/cholesterol diet suggesting an involvement of FATP4in the TAG processing [57]. It is noteworthy that TAG re-esterifica-tion takes place in the ER membrane which is also the main expres-sion site for FATP4 gene. Therefore, it is likely that the ACS-likefunction of FATP4 is involved in the metabolic fate of TAG in themouse enterocyte. The fact that the human FATP4 Gly-209-Serpolymorphism is associated with a decrease in post-prandial tria-cylglycerolemia and chylomicron/remnant ratio may be consistentwith this assumption [58].

2.4.3. CD36CD36 is a 88 kDa transmembrane glycoprotein firstly identified

in rat adipocytes by labelling with sulfo-N-succinimidyl derivatesof LCFA under conditions where LCFA uptake was significantlyinhibited [33]. CD36 is a multifunctional protein homologous tothe class B scavenger receptor SR-B1. CD36 increases the uptakeof LCFA by cardiomyocytes and adipocytes [59,60] and that of oxi-dized-LDL by macrophages [61], modifies platelet aggregation bybinding to thrombospondin and collagen [62], facilitates thephagocytosis of apoptotic cells by macrophages [63] and increasesthe cyto-adhesion of erythrocytes infected with Plasmodium falci-parum [64,65]. In addition, CD36 has also recently been shown toplay a role in the taste-reception of dietary lipids on the tongue[66,67]. Analysis of its amino-acid sequence predicts a receptor-like structure. Two transmembrane domains located near theN- and C-terminal tails result in a hairpin configuration with a

张海超
Sticky Note
2.4细胞膜脂肪结合蛋白的功能? 小肠BBM中分离出了3种与LCFA有关的高表达的蛋白:FABPpm,FATP4,CD36。 2.4.1FABPpm FABPpm是一个43kDa的在有较高脂类要求的器官表达的外周的膜蛋白。它通过一个特殊的N端肽连接到膜上。有证据支持FABPpm在细胞LCFA吸收中有作用。FABPpm可以通过明显的连续分离的80nm来结合LCFA。在体外用多细胞系FABPpm抗体培养肝细胞、脂肪细胞、心肌细胞时,LCFA的吸收都减少了。FABPpm过度表达的小鼠肌肉中棕榈酸盐的吸收增加。 在小肠中,在空肠和回肠中发现了FABPpm。尽管有棕榈酸的发现,小肠脂类的吸收仍然有疑问。用FABPpm的特异性抗体培养的空肠只导致了一部分但是很显著的抑制3H-油酸盐的吸收。无论如何,为了达到这种抑制需要大量的抗血清。而且,在隐窝细胞中也发现了FABPpm没有参与脂类吸收提出了一个疑问,即这种蛋白质在小肠LCFA吸收中的作用。一个特殊的参与细胞LCFA运输也是有缺陷的,因为FABPpm肽的顺序对线粒体天冬氨酸转移酶是一致的。这种在线粒体内膜发现的酶,催化尿素生成通路连接到Kreb循环的转氨作用。即使这个结果在:红骨骼肌的禁食、小鼠脂肪细胞糖尿病、人肝脏细胞中乙醇负担中发现,与FABPpm/mAspAT在脂肪新陈代谢的参与一致,这是一个在小肠脂肪吸收中难以捉摸的精确生理学角色。产生FABPpm失效的小鼠可能提供一个研究小肠脂类吸收中细胞膜LBP的生理角色的机会。
张海超
Sticky Note
FATP是一个63kDa的蛋白质,首先在前脂肪细胞中在它们能帮助吸收LCFA的基础上通过克隆的方法确认的。FATP有五到六中不同的亚型在小鼠和人体中发现。每种FATP都显示一个特殊的表达方式。如,FATP5和FATP2在肝脏中高度表达,FATP1在脂肪组织中高度表达,FATP4在小肠中。有趣的是,FATP4在小肠中的分布与脂肪的吸收的相关性很好。确实,它在空肠中的表达量很高,十二指肠中的丰度次之,空肠缺少其表达。而且,FATP4的表达最初在成熟的肠BBM绒毛上层发现。在体外实验也显示FATP4参与LCFA的吸收。确实,用反义方法介导的FATP4耗尽显著减少了新分离的肠FATP4的吸收。相似的是,在FATP4表达减少48%的小鼠中发现脂肪酸吸收下降了40%。最后,FATP4的拮抗剂减少了细胞LCFA的吸收,通过引导FATP4的合成。FATP帮助LCFA吸收的机理还不清楚。因为FATP蛋白不含LCFA结合位点,似乎不是膜表面的脂肪酸共用模型的角色。相比较而言,一个高氨基酸连续确认在酰基CoA合成酶和FATPs(即一个类似包括一个高度保守的ATP结合的前体结构)显示FATPs能够通过LCFA和veryLCFA的优先的底物特性参与脂肪酸酰基。Cos1细胞细胞膜提取出的FATP4表达显示较高的ACS活性的结果支持这一假设。这样的功能很关键。确实,LCFA的酰基对其连续的新陈代谢利用是必需的。而且,细胞膜对FA-CoA是不可渗透的。因此,FATP4通过把它们作为CoA的派生物捕获来吸收LCFA。简而言之,FATP4可能在小肠脂类LCFA矢量酰化的吸收中有直接作用。然而,近来发现FATP4在内质网ER中表达比BBM中更多,这一发现挑战了这一假设。而且,FATP4对脂肪吸收的生理学影响在体内仍然不清楚。FATP4的抑制剂在野生型小鼠中没有抑制脂肪的吸收。相似的,FATP4+/-的小鼠在标准和高脂肪日粮中脂肪的排泄和体重增加都没有变化。遗憾的是,删除FATP4的等位基因导致夭折,因为显著的围产期皮肤缺陷。为了获得FATP4-/-的小鼠和探索其在小肠脂肪吸收中的生理角色,皮肤显性用角质细胞特异性启动因子引导的FATP4转基因解决。FATP4失效的小鼠和他们的野生型显示难以区分的小肠TAG吸收和粪便脂肪排出,当饲喂标准饲粮时。缺少影响不是因为补偿性的吸入其他LBP导致的,得出FATP4在小鼠吸收脂肪酸中不是必须的。然而,发现有增加的小肠TAG水平,当FATp-/-小鼠处于一个高脂肪/胆固醇日粮中,显示FATP4参与TAG进程。TAG的再酯化在ER产生,而ER也是FATP4表达的主要位置,这一点是值得注意的。因此,FATP4的ACS-相似功能可能参与了TAG在小鼠肠上皮细胞中的代谢。人FATP4的Gly-209-Ser的多态性与采食后三酰甘油和乳糜/残余比例有关也支持这一假设。
张海超
Sticky Note
2.4.3 CD36 CD36是一个88kDa的跨膜糖蛋白首先在小鼠脂肪细胞中发现,通过在LCFA吸收显著抑制的部位用磺基-N-琥珀派生物标记。CD36是一个有多种功能的蛋白质,与清道夫受体SR-B1同源。CD36增加了LCFA的吸收通过心肌细胞和脂肪细胞和巨噬细胞的氧化型LDL,通过结合糖蛋白G和胶原蛋白改变血小板的凝结性,有助于巨噬细胞吞噬凋亡的细胞,注入恶性疟原虫后增加红细胞的细胞粘附性。此外,CD36最近显示在舌头上的饮食脂类味觉感受器有作用。分析其氨基酸顺序显示一个类似受体的结构。2个跨膜区域位于接近N端和C端区域通过细胞外的疏水区域形成一个发夹结构。
Page 5: Intestinal absorption of long-chain fatty acid

I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115 105

large extracellular hydrophobic domain [68,33,69]. CD36 bindsionized LCFA with an affinity in the nanomolar range and a stoi-chiometry of 3 mol FA by 1 mol protein [70,71]. Compelling evi-dence shows that it plays a significant role in lipid transfer inadipose tissue and heart (reviewed [72]). Indeed, CD36 gene dis-ruption leads to a high blood FFA level and an inability to performintense efforts in the mouse [73,74]. Similar function might alsoexist in the small intestine. Indeed, CD36 expression level (mRNAand protein) is especially high in the duodeno–jejunum, the majorsite of fat absorption [75]. Immunocytochemical studies in ratsand humans demonstrate that CD36 is strictly localized in thebrush border membrane of well differentiated enterocytes[75,76]. Moreover, intestinal CD36 gene expression has beenshown to parallel lipids contents of diets. Indeed, jejunal mRNAlevels are significantly up-regulated, when rats are chronicallysubmitted to a high fat diet [75] and down-regulated, when theyare fed a low fat chow [77]. However, CD36 function in gut re-mains a matter of debate. In isolated enterocytes, CD36 geneinvalidation either impairs LCFA uptake [78] or has no effect[73]. In vivo, the lack of dramatic intestinal phenotype in CD36null mice, as well as in patients with type I CD36 deficiency,seems to preclude a critical role in the net LCFA intestinal absorp-tion [78–80]. Consistently with this assumption, no change inLCFA uptake is found in FAT/CD36-null mice using an in situ iso-lated jejunal loops, an in vivo technical known to respect bothintestinal microclimate (i.e. unstirred water layer, cell polariza-tion) and lymph/blood circulation (I.N., unpublished data). By con-trast, a disappearance of CD36 in BBM of enterocytes is observedwhen starved rats or mice are refed a standard laboratory chow(I.N. et al., unpublished data). This phenomenon is rapid and strictlylipid-dependent. The mechanism responsible for this membranechange is presently unknown. A cellular internalization of CD36via an endocytosic transport is possible. Indeed, it has been re-ported that CD36 is co-localized with Cav-1 in caveolae vesicles[81–83,56,84]. Such a transport might target LCFA toward the sitesof lipoprotein synthesis (ER and Golgi, Fig. 1B). However, the effi-ciency of this transport system is questionable, especially duringthe post-prandial period. An alternative hypothesis is an involve-ment of CD36 in cell signalisation. As numerous receptors [85], li-gand-induced internalization of CD36 is followed by an ubiquitin/proteasome-mediated degradation in the enterocyte (I.N. et al.,unpublished data). This finding seems more consistent with a roleas a receptor rather than an efficient lipid transporter. It is notewor-thy that cytoplasmic C-terminal tail of CD36 is able to physicallyinteract with Src kinases in various tissues [67,86–89]. This systemmight constitute a functional complex allowing a lipid sensing as ithas been recently shown in the mouse taste buds [67,89]. Indeed,LCFA binding to lingual CD36 induces neuromediator release bytaste receptor cells through a signalling cascade triggered by Scr ki-nases. Resulting lipid signal conveyed by gustatory pathway con-tributes to the spontaneous fat preference and cephalic phase ofthe digestion [67]. In the small intestine, a lipid-sensing might leadto metabolic adaptation at the origin of more efficient lipid han-dling by enterocytes. TAG retention in enterocytes and productionof smaller chylomicrons occurring in CD36-null mice subjected toa high fat diet correlate quite well with this assumption[73,80,90]. Therefore, intestinal CD36 seems rather play a role inthe lipoprotein synthesis and secretion than in LCFA uptake. Contri-bution of CD36 as an intestinal lipid sensor is also supported by re-cent data showing that CD36 can link dietary fat intake to satiety viathe synthesis of the endocannabinoid oleoyletanolamide [91]. Re-cent identification of a CD36-related receptor responsible for theolfactory detection of a fatty acid-derived pheromone in Drosophila[92] suggests that the lipid-sensing function of CD36, previously re-ported in the mouse [66], may be widespread throughout the ani-mal kingdom.

In conclusion, the direct contribution of plasma membrane LBP(FABPpm, FAPT4 and FAT/CD36) to the LCFA uptake seems unlikelyin the small intestine. However in vivo experiments using trans-genic mouse models suggest that they can affect the metabolic fateof LCFA in the enterocyte (Fig. 2).

3. Intracellular trafficking of LCFA

Once into the enterocyte, LCFA and FACoA are reversibly boundto fatty acid-binding proteins (FABPs) and acyl-CoA binding pro-tein (ACBP), respectively (Fig. 1B).

3.1. Why two different FABPs?

FABPs belong to a multigenic family of 14–15 kDa intracellularproteins exhibiting a high affinity for various hydrophobic mole-cules (LCFA, bile acids or retinoids) (for a review see [93]). Two dif-ferent FABPs are expressed in the small intestine: the intestinal-type (I-FABP/FABP2) and the liver-type (L-FABP/FABP1). There isan extensive inter-species conservation of FABP peptide sequencessince more than 80% homology exists between human and rat L-FABP. In contrast, there is less than 30% homology between I-FABPand L-FABP in a same species [94]. Despite specific peptide se-quences, members of FABP family display a similar tertiary struc-ture that consists of two a helices (aI, aII) and ten anti-parallel bstrands (bA-bJ) organized in two almost orthogonal b sheets form-ing an hydrophobic pocket [95]. A ‘‘portal” region constituted bythe a helices connected to the bC/bD and bE/bF strands allowsthe entry and exit of LCFA [95,96] (Fig. 3A). The small intestine con-stitutes a unique example of an organ in which two different FABPsare highly co-expressed. The physiological advantage of such spec-ificity remains elusive. It is likely that I-FABP and L-FABP exert bothcommon (redundant) and specific functions. I-FABP or L-FABPoverexpression increases LCFA influx in cell lines in which theseproteins are lacking, as fibroblastic L cells [97] and pluripotentmouse embryonic ES cells [98], or only one FABP-type is constitu-tively expressed, as hBRIE 380i (+I-FABP/�L-FABP) [99] and HepG2(�I-FABP/+L-FABP) [100]. This driving force, which might partici-pate in the high efficiency of intestinal LCFA uptake, is likely dueto common trapping properties of these FABPs.

I-FABP is strictly confined to the small intestine. It only bindsLCFA with a ratio of one mole for one mole of protein. Peptide YYspecifically induces I-FABP gene expression in enterocyte-likehBRIE380i cells [101]. This regulatory peptide is secreted by ilealendocrine cells when dietary fat reaches the distal part of thegut. The fact that peptide YY acts as a paracrine agent might ex-plain why I-FABP induction found in rodents subjected to a highfat diet is especially high in the ileum [102,103]. Using a resonanceenergy transfer assay, Hsu and Storch have demonstrated that thetransfer of fluorescent anthroyloxy-FA from I-FABP to acceptormembrane vesicles takes place through a direct collisional interac-tion [104]. This exchange requires ionic interactions between fewamino-acid residues of the helicoidal domain of I-FABP (Fig. 3Aand B) with anionic phospholipids of membranes. In the helix-lessmutant I-FABP obtained by site-directed mutagenesis, the colli-sional exchanges are fully suppressed. In these conditions, transferof anthroyloxy-FA occurs by aqueous diffusion as reported for theL-FABP [105]. I-FABP may also utilize the membrane-protein inter-action for LCFA acquisition [106]. Since LCFA delivered on the api-cal side of enterocytes are preferentially bound to I-FABP [107],this dynamic exchange system appears to be especially welladapted to facilitate a vectorial transport of dietary lipids fromthe BBM to cellular compartments devoted to lipoprotein synthe-sis. It is thought that the collision with a target membrane yieldsa conformational change in the flexible region of I-FABP backbone

张海超
Sticky Note
CD36结合离子化的LCFA通过一个同源性的1014(nanomolar)序列和3molFA和1mol蛋白质化学计量学。有强力证据显示其在脂肪组织和心脏中的脂类运输有重要作用。确实,CD36基因确实导致血液FFA水平升高和不能表现出紧张。相似的功能在小肠中可能也存在。CD36的表达水平在十二指肠和空肠(脂肪的主要吸收位点)中都很高。免疫细胞化学研究显示CD36完全分布在完全分化的肠上皮细胞的BBM上。而且,小肠CD36基因的表达与饮食中的脂类含量平行。当饲喂高脂肪日粮时,空肠mRA水平显著提高了,而当饲喂低脂肪日粮是表达降低。然而,胃肠道中CD36的功能还存在争论。在分离的肠上皮细胞中,CD36基因缺失不影响LCFA的吸收。在体内,在CD36失效的小鼠中和1型CD36不足病人中,缺少显著地小肠显型,似乎排出了其在LCFA小肠吸收中的关键作用。与这个假设一致的是,没有在FAT/CD36失效小鼠中发现LCFA吸收有差别,利用一个原位分离的空肠片段,这是一个体内技术,保留了小肠内微环境和淋巴血液循环。相比较,在禁饲的小鼠再饲喂标准实验室饲粮后发现肠上皮细胞BBMCD36消失。这个现象快速而完全是脂类依赖的。其机制目前还不清楚。一个细胞CD36内在化经过endocytosic运输是可能的。CD36被报道位于细胞膜穴样内陷泡中。这样的运输可能将LCFA运送到脂蛋白合成位点。然而,这个转运系统的效率还存在疑问,尤其是在采食后的时期。另一个假设是CD36参与细胞信号发送。与大多数受体相似,配体诱导的CD36内移伴随着小肠内遍在蛋白质/蛋白酶体介导的下降。这个发现更支持其作为一个受体,而不是高效的脂类载体。值得注意的是CD36的C端细胞质在生理上与Src激酶在不同组织中都有相互作用。这个系统可能构成一个功能复合体导致最近从小鼠味蕾中发现的感觉到脂类。LCFA与CD36结合诱导味觉受体细胞中神经介质的释放通过Scr激酶触发的信号。味觉通路应答的脂类信号运输导致自发的脂肪喜好和消化的头部阶段。在小肠中,一个脂类敏感的
Page 6: Intestinal absorption of long-chain fatty acid

Unstirred Waterlayer

FAH

FA-CoA

pH

> 6

+I-FABP/L-FABP

ACBP

CD36

CMVLDL

CD36

< 5H

+

Na

FATP4

LCFAProteasome-

Mediateddegradation

ACS

Signaling ?

Membrane Recycling ?

Pass

ive

trans

port

PassivetransportPassivediffusion

Fig. 2. Intestinal absorption of long-chain fatty acids: a working model. For a detailed explanation see Sections 2 and 3. ACBP, acyl-CoA-binding proteins; ACS, acyl-CoAsynthetases; CM, chylomicrons; FACoA, long-chain acyl-CoA; FAH, protonated long-chain fatty acids; FATP4, fatty acid transport protein 4; I-FABP, intestinal fatty acid-binding protein; LCFA, long-chain fatty acids; L-FABP, liver fatty acid-binding protein; VLDL, very low density lipoproteins.

ββA βBα I αII

MSFSGKYQLQSQENFEAFMKA IGLPEEL I QKGKDI KGVSEIVQNI-FABP 1

L-FABP 1

GKHFKFT I TA GSKV I QNEFTVGEECELETMTGEKVKTVVQLEβC β Eβ D β F

GNKFTVKESS A FRNIEV VFELGVTFNYNLADGTELRGTWSLE

L-FABP 45

I-FABP 44

β G β H β I β J

G NKLIGKFKRTDNGNELNTVREI I GDELVQTYV YEGVEAKRIFKKD 131

GDNKLVTTFK N I K SVTELNGD I I TNTMTL GDI VFKRISKRI 127

I-FABP 86

L-FABP 87

.

... ... .

AFDSTWKVDRSENYDKFMEKMGVN IVKRKLAAHDNL KLT I TQE

T

Ala54Thrpolymorphism

FA entry/exitportal

secondportal

Collisionalexchange

I-FABP/L-FABP

I-FABP

I-FABP

L-FABP

16:0 18:0 18:1, n-9

18:2,n-6

18:3,n-3

20:4,n-6

L-FABP

I-FABP

0

100

200

300

400

500

Kd(n

mol

/L)

A

B C

Fig. 3. Main features of fatty acid-binding proteins expressed in the small intestine. (A) Example of a typical tertiary structure of a FABP. (B) Amino-acid sequence alignmentof human I-FABP and L-FABP with the localization of a helices and b stands. Arrows indicate amino-acids responsible for the collisional exchange. The Thr54Alapolymorphism is shown. (C) Binding affinity of I-FABP and L-FABP for various LCFA.

106 I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115

constituted by the distal half of a-II helix and bC-bD turns (Fig. 3A).This event might trigger the ‘‘hinged opening” of the portal domainand facilitates LCFA release [108]. Therefore, a punctual modifica-

tion in this molecular region of I-FABP might have significant func-tional consequences. Human Ala54Thr polymorphism of I-FABPgene is consistent with this hypothesis. One base substitution in

Page 7: Intestinal absorption of long-chain fatty acid

I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115 107

codon 54 of FABP2 gene leads to the change of Ala54 by Thr54(Fig. 3A). Initially found in the Pima indians, this genetic polymor-phism is associated with a high TAG plasma levels [109], an insulinresistance [110] and an increase in the body mass index [111]. Thefact that Thr54 I-FABP displays a 2-fold greater affinity for LCFAthan the Ala54 wild-protein might explain these metabolic distur-bances [110]. Indeed, a greater avidity for LCFA of mutant I-FABPcould lead to an increase in both cellular fatty acid uptake andTAG-rich lipoproteins synthesis [110]. According to this assump-tion, a dramatic rise in LCFA transport and TAG secretion is foundin Thr54I-FABP-transfected Caco-2 cells as compared to cells trans-fected the wild-type isoform [112]. Similarly, organ-culture of jeju-nal explants expressing the Thr54 allele appears to be associatedwith a rise in TAG synthesis and chylomicron output [113]. Con-versely, the down-regulation of I-FABP gene is associated with adecrease in the [14C]-palmitate uptake, TAG synthesis and secre-tion in Caco-2 cells [114]. Collectively, these data suggest that achange in I-FABP binding properties and/or expression levels canaffect the intestinal fat absorption. Paradoxically, I-FABP geneinvalidation does not lead to a lipid malabsorption in the mouse[115]. Likewise, mouse I-FABP gene deficiency mimics humanThr54 mutation. Indeed, feeding a high fat diet increases weightgain, triacylglycerolemia and insulin resistance in male I-FABP-nullmice [115]. Origin of these discrepancies remains unclear. One pos-sibility might be that metabolic alterations induced by Thr54mutation are secondary to a drop in LCFA transfer activity ofI-FABP. Whatever the mechanism involved, these data clearly showthat I-FABP and, thereby small intestine plays a role in the meta-bolic syndrome.

In contrast to I-FABP, L-FABP is also expressed in liver and, in alesser extent, in kidneys. Because of a hydrophobic pocket muchlarger (440 Å3 versus 234 Å3 for the I-FABP) [94], L-FABP can bind2 LCFA as well as various bulky hydrophobic molecules as bileacids [116] or heme [117]. This difference in volume cavity alsoaffects LCFA binding affinities. Indeed, one of two binding sites ofL-FABP exhibits a higher avidity for unsaturated LCFA (foldedmolecules) than I-FABP [118] (Fig. 3C). L-FABP displays other spec-ificities not found for I-FABP. L-FABP-mediated transfer of LCFAtakes place through an aqueous diffusion [104]. Its intestinalexpression is up-regulated by LCFA through a PPAR-dependentmechanism [119]. L-FABP seems also to exert an active role as part-ner of the peroxisome proliferator-activated receptor (PPAR), nu-clear receptors considered as a cellular lipid sensor [120]. Thispositive regulatory loop associated to a binding ratio of LCFA toL-FABP/I-FABP of 3.3 [107] strongly suggest that L-FABP can actas an intracellular reservoir for LCFA. Through this action, L-FABPmight support lipid metabolism when lipid supply is low as wellas protect the cell against harmful effect of an excess of free fattyacid during post-prandial period. From a physiological point ofview, this last function might be essential to maintain the intesti-nal mucosa integrity required to an efficient lipid absorption. Final-ly, a role of L-FABP in lipoprotein processing has been recentlyreported in vitro. L-FABP seems to contribute to the budding ofER membrane facilitating the formation of pre-chylomicron trans-fer vesicles (PCTV) which are subsequently targeted towards theGolgi apparatus [121]. The physiological relevance of this findingis supported by in vivo studies. Indeed, L-FABP gene disruptionleads to a TAG retention in enterocytes associated with a decreasein intestinal TAG output in mice subjected to an acute lipid load[122]. The fact that L-FABP gene is tightly regulated by the lipidcontent of diet is also consistent with such a function. However,these effects are not reproduced when L-FABP-null mice were onlyfed a standard laboratory chow suggesting the existence of analternative mechanism [122]. It is noteworthy that L-FABP-nullmice grow normally what suggests a normal intestinal absorption

when the lipid content of diet is low as in the standard laboratorychow (4% lipids, w/w).

In summary, physiological functions of FABPs in the small intes-tine are likely more complex than those initially assigned (i.e. LCFAuptake and intracellular trafficking). Their respective structuralfeatures speak for a specialization at least in high dietary fat sup-ply: I-FABP being predominantly devoted to the TAG-rich lipopro-tein synthesis, while L-FABP might be involved in the formation ofPCTV, membrane protection and gene regulation. Because I-FABPand L-FABP gene disruption does not lead to a severe phenotypein mice fed a low fat diet, existence of a functional redundancyto ensure a correct fat absorption, especially during the post-pran-dial period is likely. Double I-FABP/L-FABP knock-out mice mightprovide new insights concerning their respective physiologicalfunctions.

3.2. Acyl-CoA-binding protein: an housekeeping protein

Thio-esterification of LCFA in FA-CoA is an obligatory step initi-ating the cellular lipid metabolism. It is catalyzed by a set of mem-brane-associated ACSs. The newly synthesized FA-CoA which arenot immediately metabolised are bound to a specific carrier pro-tein, the acyl-CoA binding protein (ACBP) (Fig. 1). This LBP is anubiquitous 10 kDa soluble FA-CoA transporter conserved fromthe yeast to the mammal (for a review see, [123]). The 86 amino-acid residues of ACBP are folded in four a-helix forming a boomer-ang structure [124,125]. The acyl chain of FA-CoA is buried in ahydrophobic groove of the binding pocket within it is totally pro-tected from the aqueous solvent by its acyl-CoA head [126,127].ACBP binds with a stoichiometry of 1:1 and an affinity in the nano-molar range both medium and long-chain FA-CoA [127,128]. In thesmall intestine from rats and mice, ACBP is co-expressed with FAB-Ps in absorptive cells [129]. In rat, a similar ACBP mRNA level isfound along the cephalo-caudal axis of the small intestine (I.N.et al., unpublished data) in contrast what is observed for L-FABPand I-FABP. The precise function of ACBP in this organ is not yetfully determined. ACBP might facilitate FA-CoA desorption fromFATP4 and other ACSs and, thus, contribute to the lipoprotein syn-thesis (Fig. 2). Indeed, ACBP has been reported to regulate the cel-lular FA-CoA disposal for microsomal TAG synthesis [126]. The factthat the ACBP gene disruption leads to a drastic reduction in TAGaccumulation in pre-adipocytes is consistent with this assumption[130]. The detection of ACBP and FA-CoA in the nucleus of differentcells [131] suggests a possible interference of ACBP expression and/or FA-CoA disposal with fundamental regulatory pathways [130].

A working model illustrating the main steps involved in LCFAuptake and trafficking in the absorptive intestinal cells is proposedin Fig. 2.

4. Chylomicron assembly and trafficking

FA-CoA are rapidly re-esterified in TAG in the ER before to formpre-CM then CM in the Golgi apparatus (Fig. 4). During the last pastyears, significant progress in the understanding of molecularevents responsible for CM assembly, trafficking and release havebeen realized (for recent reviews see, [132–134].

4.1. Re-esterification step

Re-esterification of FA-CoA takes place in the ER membrane viatwo multi-enzymatic systems [135]. The 2-monoglyceride path-way, specific of the small intestine, is located in the smooth ER(SER). It is responsible for up to 80% of newly synthesized TAG dur-ing the post-prandial period. Enzymes of the a-glycerophosphatepathway, also active in the liver, are found on rough ER (RER). Their

Page 8: Intestinal absorption of long-chain fatty acid

golgi

Mesenteric lymph

SERRER

DGAT2

DGAT1

Acyl-carnitinetransferaseMTP

Limiting steps

TAG

TAG

B48

B48

CMA1

A4

C2/3

B48

B48

TAG

Proteasome-Mediated

degradation

DGAT2

DGAT1

Acyl-carnitinetransferase MTP

B48

TG TAGA4

B48

FACoA

CMA1

A4

C2/3

B48

TAG

lipases

Acyl-carnitinetransferase

MTPDGAT1

TAG

L-FABP(budding)

COP-II (Sar1b)SNARE (VAMP7)

(fusion)

PCTV

fusion

Synt

hesi

sStorage

pool

Storagepool

FFA

Portal blood

1

9

6

2

3

4

5

7

8

10

TAG

TAG TAG

TAG

Fig. 4. Synthesis of intestinal lipoproteins: a working model. For a detailed explanation see Section 4. ApoB48, apolipoprotein B48; CM, chylomicrons; DGAT, acyl-CoA-diacylglycerol acyltransferases; FFA, free fatty acids; MTP, microsomal triacylglycerol transfer protein; RER, rough endoplasmic reticulum; SER, smooth endoplasmicreticulum; TAG, triacylglycerol (triglycerides).

108 I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115

activity is predominant during the interprandial and fasting peri-ods. In contrast to the 2-monoglyceride pathway, which only pro-duces TAG, a-glycerophosphate pathway leads also to thesynthesis of phospholipids [136]. Whatever the pathways, the finalreaction leading to TAG synthesis is catalyzed by the acyl-CoA-diacylglycerol acyltransferases (DGAT). Two unrelated DGAT,termed DGAT1 and DGAT2, are co-expressed in the small intestine[137,138]. Despite similar biochemical functions, these enzymesdisplay several differences. DGAT1 is mainly expressed in the prox-imal small intestine [137,139] in which it represents 85–90% of thetotal DGAT activity, whereas DGAT2 is more importantly found inadipose tissue and liver [140]. DGAT2 appears to be located in thecytosolic side of the ER membrane [141]. In contrast, the active siteof DGAT1 is located in the lumen of ER [138,142]. Therefore, intes-tinal TAG re-esterification takes place on the two sides of the ERmembrane suggesting the existence of two different pools of newlysynthesized TAG (Fig. 4). Consistently with this assumption, DGAT-1 over-expression in the liver mainly leads to an increase in lipo-protein secretion, while over-expression of DGAT2 essentially pro-duces a TAG accumulation in the cytoplasm as lipid droplets [143].

Paradoxically, DGAT1 gene disruption does not lead to a lipidmalabsorption when mice were fed a standard laboratory chow(3% lipids, w/w). DGAT2-mediated synthesis of TAG might explainthis data suggesting some functional redundancies between thesetwo enzymes. However, a reduction of post-absorptive chylomi-cronemia and an accumulation of cytosolic lipid droplets in entero-

cytes are found in DGAT1-null mice subjected to a high fat diet[139]. Therefore, DGAT2 appears to be insufficient to fully compen-sate the absence of DGAT1 in a high fat situation. Direct investiga-tion of DGAT2 gene role in the small intestine is presently lacking.Indeed, DGAT2 gene invalidation, leading to deep epidermal de-fects, is lethal soon after birth [140]. Generation of mice with tar-geted DGAT2 gene disruption in spatio-temporal manner will berequired to explore further its function. It is noteworthy that thediacylglycerol transacylase [144] and the lecithine–cholesterolacyltransferases [145] can also constitute alternative pathwaysfor the TAG synthesis in the small intestine. In brief, we can thinkthat during post-prandial period TAG generated by DGAT1 areimmediately available for the lipoprotein synthesis in contrast toTAG produced by DGAT2 which are mainly stored as a cytosolicpool.

4.2. TAG transfer in the endoplasmic reticulum

Existence of active DGAT on the two sides of ER membraneraises the question of mechanisms involved in the TAG deliveryinto ER citernae. Existence of two complementary transfer systemsis likely (Fig. 4). A first system is required for the delivering of FA-CoA to DGAT1 active site by reason to inability of these moleculesto cross the membranes (Fig. 4, reaction 1). The carnitine acyl-transferase-like system, initially identified in hepatic microsomes[146], might assume such a function in the small intestine. Indeed,

Page 9: Intestinal absorption of long-chain fatty acid

I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115 109

the intraduodenal infusion of the specific carnitine palmitoyltrans-ferase inhibitor etomoxir dramatically (�75%) decreases TAG out-put in lymph in rats subjected to a lipid load [142]. The secondsystem allowing the TAG transfer into the endoplasmic reticulumciternae involves the microsomal triacylglycerol transfer protein(MTP) (Fig. 4, reaction 2). Mainly expressed in the upper part ofthe villi of the proximal intestine [147], MTP is known to bindand shuttle neutral lipids within membranes [148]. It consists oftwo non convalently bound proteins. A 97 kDa subunit, which be-longs to the large lipid-transfer protein superfamily [149], contains3 functional domains allowing the association of MTP with ERmembranes, the adsorption of neutral lipids (TAG and cholesterolesters) and the binding of the apolipoprotein B48 (ApoB48). A55 kDa subunit, identified as the protein disulfide isomerase, pro-motes the proper folding of the large subunit promoting both itsretention in ER and its binding properties. MTP appears to be themost important factor regulating the intestinal ApoB lipoproteinassembly and secretion [132,150]. First MTP binds and shuttlesTAG molecules within ER membrane [151] because of its phospho-lipids and TAG transfer activity [152]. Second, the initial lipidationof ApoB48 occurs through a direct interaction with MTP. A physicalinteraction between MTP and ApoB48 has been demonstrated byusing co-precipitation experiments [153]. When MTP and ApoB48are co-expressed in Hela cells, that do not make lipoprotein parti-cles, small dense lipoproteins are produced. This phenomenon ap-pears to be strictly MTP-dependent since the lipoprotein is lackingin absence of the MTP expression vector [154]. It is thought thatMTP is needed for a proper folding of ApoB48 and the formationof a secretion competent particle. Third, MTP protects nascentApoB48 lipoprotein against a rapid proteolysis by the ubiquitin–proteasome pathway [155,156]. Consistently with these functions,mice with conditional intestine-specific MTP deficiency develop adramatic phenotype associating fat malabsorption, cytoplasmic li-pid droplet accumulation and virtual disappearance of ApoB48lipoproteins [157]. In human, a similar syndrome is found in pa-tients with abetalipoproteinemia, a rare autosomal recessive dis-ease due to structural defects in the MTP gene [158]. Conversely,the human polymorphism leading to an over-expression of MTPis associated with an increased blood level of Apo B48 bound toTAG-rich lipoprotein. [159]. Altogether these data demonstratethat MTP plays a significant rate-limiting role in lipoprotein syn-thesis (Fig. 4, reaction 2). MTP has also been identified in the Golgiapparatus [160] and in the BBM [161]. To date, the functional rel-evance of these observations remains elusive.

4.3. Pre-chylomicron synthesis

CM are large TAG-rich lipoprotein particles that transport die-tary lipids from the small intestine to other locations in the body.CM synthesis is considered to be a three step process which occursin the ER [132,152]. First, the primordial lipoprotein particles con-taining one molecule of Apo B48, phospholipids and small amountsof TAG are synthesized in the RER (Fig. 4, reaction 3). Synthesis ofprimordial lipoprotein particles requires the MTP-mediated lipida-tion of the newly synthesis ApoB48 which is an obligatory, nonexchangeable, CM partner in contrast to other constitutive apolipo-proteins. This fundamental role has been highlighted by using amouse model producing ApoB only in the liver. In these animals,the lack of ApoB48 in the small intestine totally prevents the for-mation of CM and results in a severe intestinal fat malabsorptionwith appearance of large cytoplasmic lipid droplets [162]. Simi-larly, a drop in lymph CM levels associated with an accumulationof TAG in cytosol is found in mice programmed to produce onlythe ApoB100 in the small intestine [163,164]. Second, re-esterifica-tion of TAG and cholesterol esters in the SER leads to the formationof lipid droplets (Fig. 4, reaction 4). Third, these two particles

merge leading to the core expansion of the primordial lipoproteinsand the synthesis of nascent lipoproteins (Fig. 4, reaction 5). Incor-poration of apolipoprotein AIV (ApoAIV) to the nascent CM isthought to stabilize the particle. Such a function might explainwhy (i) the ApoIV gene expression is tightly correlated to the lipidcontent of the diet [165] and (ii) its over-expression produces lar-ger CM [166,167].

4.4. Transfer of prechylomicrons to the Golgi

Transfer of pre-CM from SER to the Golgi constitutes the secondrate-limiting step in the intestinal fat absorption [168]. It is a vec-torial transport. In brief, the cargo buds from the SER membraneand forms pre-CM transfer vesicles (PCTV) which translocate toand fuse with the Golgi cis-membrane to deliver pre-CM in Golgiciternae (Fig. 4, reaction 6). This targeted transport, specific ofthe small intestine, appears to be tightly controlled by a set of pro-teins [169,170,170–173]. The molecular mechanism at the origin ofthis transport is progressively elucidated. It has been recentlyshown that L-FABP plays a role in the SER membrane deformationleading to PCTV budding [121]. This unexpected function might ex-plain why an impaired intestinal TAG secretion is found in L-FABP-null mice [122]. However, the mechanism involved remains elu-sive. It has been proposed that L-FABP organizes the PCTV bud-ding-machinery under the control of the phosphorylation of a9 KDa protein by PKCf [173]. The PCTV membrane also containsa vectorial protein: the vesicle-associated membrane protein-7(VAMP7) which belongs to the specific v-soluble-N-ethylmalei-mide-sensitive factor attachment protein receptor (v-SNARE) fam-ily. This protein, only found in the small intestine [172], seems tobe crucial for the targeting of PCTV to the Golgi membrane. Indeed,the use of an antibody raised against VAMP7 leads to 85% decreasein the PCTV transfer [172]. Finally, PCTV fusion with the cis-Golgimembrane requires the presence of Sar1b protein, a member ofcoatomer II proteins (COPII) system [174]. Indeed, missense muta-tions of SARA2 gene encoding for the Sar1b protein is found both inCM retention disease and Anderson’s disease, two rare humanpathologies characterized by a severe malabsorption, fat ladenenterocytes and the inability to produce chylomicrons [169]. Finalmaturation of CM takes place in the Golgi through post-transla-tional modification (i.e. glycolysation and addition of the minorapolipoproteins Apo A1, Apo CII and apo CIII (Fig. 4, reaction 7)).

4.5. Generation of intestinal lipid droplets

Both in the healthy mouse and human, a lipid load leads to theformation of lipid droplets in the cytoplasm of enterocytes [175–179]. Similar event can be reproduced in vitro in Caco-2 cells[180]. It is a transient intestinal steatosis due to the conjunctionof highly efficient LCFA uptake and TAG re-esterification levelswith the existence of two rate-limiting steps in the CM synthesis:MTP-mediated transfer of neutral lipids through ER membrane andvectorial targeting of PCTV into Golgi (Fig. 4). When the TAG syn-thetic rate is upper to their subsequent transfer into ER citernae,DGAT1-synthesized TAG pile up on the cytosolic face of the ERand/or are formed at the external side of the ER membrane byDGAT2 [177] (Fig. 4, reaction 8). It is likely that these intestinalcytosolic lipid droplets are surrounded with a phospholipid mono-layer as shown in liver cells [181]. Cytosolic TAG pool, accumulatedduring post-prandial period, is transient. It might provide sub-strates for the formation of TAG-rich lipoprotein during the interp-randial period [177,176]. This pathway requires an efficient TAGhydrolysis (Fig. 4, reaction 9). Interestingly, several cytosolic andmicrosomal lipases, including pancreatic TAG lipase [182], hor-mone-sensible lipase [183] and arylacetamide deacetylase [184],have been identified in the enterocyte. Some of FFA released from

Page 10: Intestinal absorption of long-chain fatty acid

Table 1Lessons from genomics.

Proteins Lipid fecal loss (lowand high fat diet)

Intestinal lipid metabolism Triacylglyceridemia after a lipid rich meal References

FATP4 Unchanged Subtle intestinal TAG retention on Westerndiet.

No change in blood TAG levels and kinetics [57]

Fatp-4�/�/Ivl-Fatp4tg:+)

miceUnchanged in a chow fed mice

Human Gly209polymorphism

Undetermined Decrease in triacylglyceridemia and ratiochylomicrons/remnants.

[58]

CD36 Unchanged Fat laden enterocytes. Decreased lymphaticTAG secretion on high fat diet

Increased blood TAG and FFA levels. Small CMleading to a decrease in LPL-mediated CMclearance

[79,80,90]CD36�/� mice

Human CD36 deficiency Undetermined Small CM. Increased blood TAG and FFA levels [80]

I-FABP Undetermined Undetermined Hypertriacylglyceridemia in male, only [112,115]I-FABP�/� miceHuman Ala54Thrpolymorphism

Undetermined Increased Apol B synthesis and CM secretionin vitro

Post-prandial hypertriglyceridemia [110,113,109]

L-FABP Unchanged Fat laden enterocytes. Decreased lymphaticTAG secretion on high fat diet. Unchanged ina chow fed mice

Decreased lymphatic TAG secretion [122]L-FABP�/� mice

DGAT1 Unchanged Fat laden enterocytes. Decreased lymphaticTAG secretion on high fat diet. Unchanged ina chow fed mice

Decreased blood CM and TAG levels [139]DGAT-1 (�/�) mice

MTP Steatorrhoea Massive fat laden enterocytes. Lack of CM-sized particles within the secretory pathway

Apobetalipoproteinemia [157]Conditional intestinalMTP�/� mice

Human MTP deficiency Steatorrhoea Idem Idem [158]

ApoB Steatorrhoea Massive fat laden enterocytes Absence of CM secretion [162]Mice producing Apo Bonly in the smallintestine

Lack of CM-sized particles within thesecretory pathway

No modification of plasma TAG levels after 4 hfasting

Sar1b Steatorrhoea Fat laden enterocytes.Human mutations inSARA 2 gene

inability to produce chylomicrons [169]

110 I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115

cytosolic TAG pool by lipases might also exit the enterocyte to betransported to the liver via the portal vein (Fig. 4, reaction 10).

5. Lessons learned from genomics

The generation of transgenic mouse models with a targetedknock-out or over-expression of genes involved in fat absorptionas well as the identification of related human polymorphisms hasevidenced several intestinal specific features (Table 1).

First, in contrast to what is demonstrated in muscles and adiposetissue, CD36 gene has only a neglected role in the intestinal LCFA up-take. This finding highlights that a direct generalization of a genefunction found in one organ to another one, without to take in ac-count of its own specificities (i.e. micro-environment and phenotype)is physiologically irrelevant. By contrast, intestinal CD36 plays a rolein the metabolic fate of re-esterified TAG. Indeed, CD36 gene invali-dation is associated with fat laden enterocytes and a decrease in lym-phatic CM secretion. The molecular mechanism at the origin of theseCD36-mediated effects must to be elucidated. Microclimate liningBBM explains why LCFA uptake by enterocytes remains efficient dur-ing a lipid load. In contrast, the rate-limiting steps in the fat absorp-tion takes place at the ER and Golgi levels and are dependent of MTPand proteins responsible for the vectorial targeting of pre-CM into theGolgi (e.g. L-FABP, Sar1b).

Second, invalidation of several genes, considered for a long timeas crucial for fat absorption (e.g. I-FABP, DGAT1), does not lead to aclear phenotype when animals are fed a standard chow. Metabolicredundancies (e.g. I-FABP versus L-FABP or DGAT1 versus DGAT2)are likely responsible for this observation. Existence of alternativepathways contributing to the efficiency of lipid absorption is of a

great physiological interest. Indeed, they prevent an excessive fecalfat output and, hence, energy loss.

Third, genomics has contributed to reveal unexpected genefunctions as, for instance, the role of L-FABP gene in the PCTV syn-thesis and transfer towards the Golgi. Targeted gene modificationsspecifically in the small intestine might provide new insights inintestinal fat absorption in a near future.

6. Do dietary lipids affect the intestinal function?

Chronic fat overconsumption increases the obesity risk in num-ber of species. Although the small intestine is responsible for fatbody disposal, its role in this phenomenon has been neglected.The fact that the small intestine has long been considered a simpleselective barrier able to efficiency absorb dietary fat explain thisparadox. Recent data strongly suggest that the high TAG bioavail-ability of intestine might be not attributable to inborn propertiesbut acquired properties. Indeed, intestinal fat absorption capacitycan be adapted to the fat content of the diet in the mouse. Thisfat-mediated adaptation takes place through two complementaryevents. First, there is a lipid-mediated induction of intestinal cellproliferation, which might lead to an increased absorptive area[103]. Consistent with this assumption, rat studies show that highfat diets increase the height of villi and induce the rate of entero-cyte migration along the crypt-to-villus axis [185,186]. It is note-worthy that the effect of lipids on intestinal trophism appears tobe more efficient than that of other nutrients. Second, mice sub-jected to a chronic high fat diet display a coordinate induction ofgenes (i.e. CD36, FATP4, I-FABP, L-FABP, MTP, Apo AIV). This li-pid-mediated modification is rapid and adaptative since the gene

Page 11: Intestinal absorption of long-chain fatty acid

I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115 111

expression returns to control values when mice are re-fed a low fatdiet [103,187]. Interestingly, these genes seem to play a significantrole in the final size of CM by facilitating the expansion core of TAGin the CM, and hence, in their subsequent blood clearance by LPL.Moreover, it has been recently reported that lipid content of thediet is also able to modulate gene expression of secreted signallingmolecules which could affect the metabolic homeostasis of periph-eral organs [188]. All together, these emergent data highlight thatthe lipid-mediated impact on the small intestine could contributeto the etiology of metabolic disorders linked to a high fat diet.

7. Intestinal contribution to dyslipidemia

Hypertriacylglycerodemiad is a metabolic disorder generallyassociated with insulin-resistance, obesity, metabolic syndrome.Adverse effects of a chronic post-prandial high TAG levels on car-dio-vascular function is well established. Although CM and theirremnants can be atherogenic [189], contribution of the small intes-tine in the genesis of this dyslipidemia is not much studied as com-pared to the liver implication. The fact that triacylglycerolemia isusually determined in fasted human can explain this paradox. How-ever, it is well known that the plasma TAG level remains elevatedfor the most of the day even in subjects with normal fasting TAG[190,191]. Post-prandial triacylglycerolemia results not only fromthe TAG-rich lipoprotein secretion but also from their subsequentblood clearance by endothelial LPL. CM size and number are knownto affect LPL activity and, hence, blood TAG clearance. A small num-ber of large CM are more rapidly cleared by LPL than a large numberof small CM [192–194]. CM size is highly dependent on expressionlevels of few genes expressed in the small intestine. For example,CD36-null mice secrete smaller CM than wild-type mice [90] andintestinal ApoAIV overexpression produces larger TAG-rich lipopro-teins [166,167]. LPL activity is also under control of apolipoproteinsCII (apoCII) and CIII (ApoCIII) acquired during the intestinal CM bio-genesis (Fig. 4, reaction 7) or by exchange with hepatic lipoproteins.Apo CII is a co-factor which activates LPL. For this reason, the lowlipolytic activity of LPL found in Apo CII-deficient human can be cor-rected by the addition of Apo CII [195]. In contrast, Apo CIII is an LPLinhibitor. Therefore, ApoCII/ApoCIII ratio found in CM determinesthe efficiency of their blood clearance by LPL.

Generally, fat feeding increases the size of the CM particles ratherthan their number [196]. However, this effect is highly dependentfrom the quality of dietary lipids. A diet rich in unsaturated fattyacids leads to larger CM than a diet containing mainly saturatedfatty acids [197–199]. The nature of absorbed fatty acids also affectstheir blood clearance. In the mouse, a high fat diet mainly composedof monounsaturated fatty acids specifically up-regulates the expres-sion of ApoCII gene but down-regulates the ApoCIII gene in the smallintestine. This change, not reproduced in the liver, might explainedwhy mice subjected to this diet displayed a lower post-prandial tria-cylglycerolemia than mice fed standard chow [103]. In humans,studies have shown that post-prandial blood lipid levels are alsohighly dependent of the fatty acid composition of the test meal(SFA > MUFA > PUFA) (for a review, see [200]).

The intestinal impact on the blood TAG levels seems to be stillmore important during insulin-resistance [201]. Enterocytes frominsulin-resistant hamsters secrete more TAG, more ApoB48 andthus a greater number of chylomicrons than controls [202]. Similaralterations has been also found in vivo in diabetic rabbit [203], andas initially suggested by the Tomkin’s team [204], in humans [205].

8. Conclusions and future directions

This review illustrates the remarkable capacity of the smallintestine to absorb dietary fat. Modulation of intestinal plasticity

(i.e. cell proliferation and coordinate gene regulation) by dietarylipids and the existence of compensatory redundant mechanismsallow the adaptation of the fat absorption to the lipid content ofthe diet. This lipid-mediated adaptability appears to be sufficientto prevent an excessive lipid elimination in faeces when fat supplyis enhanced. This physiological feature maximizes the absorptionof foods known to be rich in energy, in essential fatty acids andin lipid-soluble vitamins (A, D, E, and K) which display numerousfundamental biological functions. Likely inherited from evolution,these lipid-mediated intestinal adaptations might constitute anadvantage allowing the survival in an environment of food scarcity.Conversely, it might contribute to increase the prevalence of obes-ity and related diseases during periods of food abundance. Never-theless, molecular mechanisms responsible for this adaptation areyet not fully understood. For instance, the regulator(s) at the originthe coordinated modulation of genes responsible for intestinalmetabolic fate of dietary lipids remains to be determined. Thesmall intestine also secretes numerous substances that may alsostimulate intestinal adaptation under certain extreme conditionsas high fat feeding, and this idea is being explored. A better under-standing of the mechanisms leading to lipid-dependent intestinaladaptation may provide new therapeutic approaches and/or die-tary recommendations allowing the optimisation of blood CMclearance and, hence, decrease prevalence of fat-mediated diseasesin the population.

Acknowledgments

This work was supported by the National Institute of Agro-nomic Research (INRA) and the National Institute of Health andMedical Research (INSERM) through the Research Program in Hu-man Nutrition and Alimentation (PRNH to PB), and by grants fromBurgundy Council (to PB).

References

[1] Armand M, Borel P, Dubois C, Senft M, Peyrot J, Salducci J, et al.Characterization of emulsions and lipolysis of dietary lipids in the humanstomach. Am J Physiol 1994;266:G372–81.

[2] Mu H, Hoy CE. The digestion of dietary triacylglycerols. Prog Lipid Res2004;43:105–33.

[3] Eyster KM. The membrane and lipids as integral participants in signaltransduction: lipid signal transduction for the non-lipid biochemist. AdvPhysiol Educ 2007;31:5–16.

[4] Linder ME, Deschenes RJ. Palmitoylation: policing protein stability and traffic.Nat Rev Mol Cell Biol 2007;8:74–84.

[5] Berk PD, Stump DD. Mechanisms of cellular uptake of long chain free fattyacids. Mol Cell Biochem 1999;192:17–31.

[6] McArthur MJ, Atshaves BP, Frolov A, Foxworth WD, Kier AB, Schroeder F.Cellular uptake and intracellular trafficking of long chain fatty acids. J LipidRes 1999;40:1371–83.

[7] Hajri T, Abumrad NA. Fatty acid transport across membranes: relevance tonutrition and metabolic pathology. Annu Rev Nut 2002;22:383–415.

[8] Hamilton JA, Guo W, Kamp F. Mechanism of cellular uptake of long-chainfatty acids: do we need cellular proteins? Mol Cell Biochem 2002;239:17–23.

[9] Pohl J, Ring A, Stremmel W. Uptake of long-chain fatty acids in HepG2 cellsinvolves caveolae: analysis of a novel pathway. J Lipid Res 2002;43:1390–9.

[10] Mashek DG, Coleman RA. Cellular fatty acid uptake: the contribution ofmetabolism. Curr Opin Lipidol 2006;17:274–8.

[11] Kampf JP, Kleinfeld AM. Is membrane transport of FFA mediated by lipid,protein, or both? An unknown protein mediates free fatty acid transportacross the adipocyte plasma membrane. Physiol (Bethesda) 2007;22:7–14.

[12] Zakim D. Thermodynamics of fatty acid transfer. J Membr Biol2000;176:101–9.

[13] Kamp F, Zakim D, Zhang F, Noy N, Hamilton JA. Fatty acid flip–flop inphospholipid bilayers is extremely fast. Biochemistry 1995;34:11928–37.

[14] Brunaldi K, Miranda MA, Abdulkader F, Curi R, Procopio J. Fatty acid flip–flopand proton transport determined by short-circuit current in planar bilayers. JLipid Res 2005;46:245–51.

[15] Kampf JP, Cupp D, Kleinfeld AM. Different mechanisms of free fatty acid flip–flop and dissociation revealed by temperature and molecular speciesdependence of transport across lipid vesicles. J Biol Chem2006;281:21566–74.

[16] Kleinfeld AM, Storch J. Transfer of long-chain fluorescent fatty acids betweensmall and large unilamellar vesicles. Biochemistry 1993;32:2053–61.

Page 12: Intestinal absorption of long-chain fatty acid

112 I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115

[17] Kleinfeld AM, Chu P, Romero C. Transport of long-chain native fatty acidsacross lipid bilayer membranes indicates that transbilayer flip–flop is ratelimiting. Biochemistry 1997;36:14146–58.

[18] Stremmel W. Uptake of fatty acids by jejunal mucosal cells is mediated by afatty acid binding membrane protein. J Clin Invest 1988;82:2001–10.

[19] Trotter PJ, Ho SY, Storch J. Fatty acid uptake by Caco-2 human intestinal cells.J Lipid Res 1996;37:336–46.

[20] Storch J, Herr FM, Hsu KT, Kim HK, Liou HL, Smith ER. The role of membranesand intracellular binding proteins in cytoplasmic transport of hydrophobicmolecules: fatty acid-binding proteins. Comp Biochem Physiol1996;115B:333–9.

[21] Tranchant T, Besson P, Hoinard C, Pinault M, Alessandri JM, Delarue J, et al.Long-term supplementation of culture medium with essential fatty acidsalters alpha-linolenic acid uptake in Caco-2 clone TC7. Can J PhysiolPharmacol 1998;76:621–9.

[22] Chabowski A, Gorski J, Luiken JJ, Glatz JF, Bonen A. Evidence for concertedaction of FAT/CD36 and FABPpm to increase fatty acid transport across theplasma membrane. Prostaglandins, Leukot and Essent Fatty Acids2007;77:345–53.

[23] Wilson FA, Sallee VL, Dietschy JM. Unstirred water layers in intestine: ratedeterminant of fatty acid absorption from micellar solutions. Science1971;174:1031–3.

[24] Shiau YF, Fernandez P, Jackson MJ, McMonagle S. Mechanisms maintaining alow-pH microclimate in the intestine. Am J Physiol 1985;248:G608–17.

[25] Orsenigo MN, Tosco M, Zoppi S, Faelli A. Characterization of basolateralmembrane Na/H antiport in rat jejunum. Biochim Biophys Acta1990;1026:64–8.

[26] Tso P, Intestinal lipid absorption. Physiology of the gastrointestinal tract. 3rded. 1994. p. 1867–907.

[27] Shiau YF. Mechanisms of intestinal fat absorption. Am J Physiol1981;240:G1–9.

[28] Kamp F, Westterhoff HV, Hamilton JA. Movement of fatty acids, fatty acidanalogues and bile acids across phospholipid bilayers. Biochemistry1993;32:11074–86.

[29] Schoeller C, Keelan M, Mulvey G, Stremmel W, Thomson AB. Oleic acid uptakeinto rat and rabbit jejunal brush border membrane. Biochim Biophys Acta1995;1236:51–64.

[30] Ross AC. Overview of retinoid metabolism. J Nutr 1993;123:346–50.[31] Stremmel W, Lotz G, Strohmeyer G, Berk PD. Identification, isolation, and

partial characterization of a fatty acid binding protein from rat jejunalmicrovillous membranes. J Clin Invest 1985;75:1068–76.

[32] Schaffer JE, Lodish HF. Expression cloning and characterization of a noveladipocyte long chain fatty acid transport protein. Cell 1994;79:427–36.

[33] Abumrad NA, el-Maghrabi MR, Amri EZ, Lopez E, Grimaldi PA. Cloning of a ratadipocyte membrane protein implicated in binding or transport of long-chainfatty acids that is induced during preadipocyte differentiation. Homologywith human CD36. J Biol Chem 1993;268:17665–8.

[34] Potter BJ, Stump D, Schwieterman W, Sorrentino D, Jacobs LN, Kiang CL, et al.Isolation and partial characterization of plasma membrane fatty acid bindingproteins from myocardium and adipose tissue and their relationship toanalogous proteins in liver and gut. Biochem Biophys Res Commun1987;148:1370–6.

[35] Berk PD, Wada H, Horio Y, Potter BJ, Sorrentino D, Zhou SL, et al. Plasmamembrane fatty acid-binding protein and mitochondrial glutamic-oxaloacetic transaminase of rat liver are related. Proc Natl Acad Sci USA1990;87:3484–8.

[36] Stremmel W, Strohmeyer G, Borchard F, Kochwa S, Berk PD. Isolation andpartial characterization of a fatty acid-binding protein in rat liver plasmamembranes. Proc Natl Acad Sci USA 1985;82:4–8.

[37] Stremmel W, Theilmann L. Selective inhibition of long-chain fatty acid uptakein short-term cultured rat hepatocytes by an antibody to the rat liver plasmamembrane fatty acid-binding protein. Biochim Biophys Acta1986;877:191–7.

[38] Zhou SL, Stump D, Sorrentino D, Potter BJ, Berk PD. Adipocyte differentiationof 3T3–L1 cells involves augmented expression of a 43-kDa plasmamembrane fatty acid-binding protein. J Biol Chem 1992;267:14456–61.

[39] Sorrentino D, Stump D, Potter BJ, Robinson RB, White R, Kiang C-L, et al.Oleate uptake by cardiac myocytes is carrier mediated and involves a 40-kDplasma membrane fatty acid binding protein similar to that in liver, adiposetissue, and gut. J Clin Invest 1988;82:928–35.

[40] Luiken JJ, Turcotte LP, Bonen A. Protein-mediated palmitate uptake andexpression of fatty acid transport proteins in heart giant vesicles. J Lipid Res1999;40:1007–16.

[41] Clarke DC, Miskovic D, Han XX, Calles-Escandon J, Glatz JF, Luiken JJ, et al.Overexpression of membrane-associated fatty acid binding protein (FABPpm)in vivo increases fatty acid sarcolemmal transport and metabolism. PhysiolGenomics 2004;17:31–7.

[42] Isola LM, Zhou SL, Kiang CL, Stump DD, Bradbury MW, Berk PD. 3T3fibroblasts transfected with a cDNA for mitochondrial aspartateaminotransferase express plasma membrane fatty acid-binding protein andsaturable fatty acid uptake. Proc Natl Acad Sci USA 1995;92:9866–70.

[43] Zhou SL, Stump D, Isola L, Berk PD. Constitutive expression of a saturabletransport system for non- esterified fatty acids in Xenopus laevis oocytes.Biochem J 1994;297:315–9.

[44] Stump DD, Zhou SL, Berk PD. Comparison of plasma membrane FABP andmitochondrial isoform of aspartate aminotransferase from rat liver. Am JPhysiol 1993;265:G894–902.

[45] Turcotte LP, Srivastava AK, Chiasson JL. Fasting increases plasma membranefatty acid-binding protein (FABP(PM)) in red skeletal muscle. Mol CellBiochem 1997;166:153–8.

[46] Berk PD, Zhou SL, Kiang CL, Stump D, Bradbury M, Isola LM. Uptake of longchain free fatty acids is selectively up-regulated in adipocytes of Zucker ratswith genetic obesity and non-insulin- dependent diabetes mellitus. J BiolChem 1997;272:8830–5.

[47] Zhou S-L, Gordon RE, Bradbury M, Stump D, Kiang C-L, Berk PD. Ethanol up-regulates fatty acid uptake and plasma membrane expression and export ofmitochondrial aspartate aminotransferase in HepG2 cells. Hepatology1998;19998:1064–74.

[48] Stahl A, Hirsch DJ, Gimeno RE, Punreddy S, Ge P, Watson N, et al.Identification of the major intestinal fatty acid transport protein. Mol Cell1999;4:299–308.

[49] Stahl A, Gimeno RE, Tartaglia LA, Lodish HF. Fatty acid transport proteins: acurrent view of a growing family. Trends Endocrinol Metab 2001;12:266–73.

[50] Gimeno RE, Hirsch DJ, Punreddy S, Sun Y, Ortegon AM, Wu H, et al. Targeteddeletion of fatty acid transport protein-4 results in early embryonic lethality.J Biol Chem 2003;278:49512–6.

[51] Blackburn C, Guan B, Brown J, Cullis C, Condon SM, Jenkins TJ, et al.Identification and characterization of 4-aryl-3,4-dihydropyrimidin-2(1H)-ones as inhibitors of the fatty acid transporter FATP4. Bioorg Med Chem Lett2006;16:3504–9.

[52] Hall AM, Wiczer BM, Herrmann T, Stremmel W, Bernlohr DA. Enzymaticproperties of purified murine fatty acid transport protein 4 and analysis ofacyl-CoA synthetase activities in tissues from FATP4 null mice. J Biol Chem2005;280:11948–54.

[53] Jia Z, Pei Z, Maiguel D, Toomer CJ, Watkins PA. The fatty acid transport protein(FATP) family: very long chain acyl-CoA synthetases or solute carriers? J MolNeurosci 2007;33:25–31.

[54] Herrmann T, Buchkremer F, Gosch I, Hall AM, Bernlohr DA, Stremmel W.Mouse fatty acid transport protein 4 (FATP4): characterization of the geneand functional assessment as a very long chain acyl-CoA synthetase. Gene2001;270:31–40.

[55] Milger K, Herrmann T, Becker C, Gotthardt D, Zickwolf J, Ehehalt R, et al.Cellular uptake of fatty acids driven by the ER-localized acyl-CoA synthetaseFATP4. J Cell Sci 2006;119:4678–88.

[56] Ehehalt R, Sparla R, Kulaksiz H, Herrmann T, Fullekrug J, Stremmel W. Uptakeof long chain fatty acids is regulated by dynamic interaction of FAT/CD36with cholesterol/sphingolipid enriched microdomains (lipid rafts). BMC CellBiol 2008;9:45.

[57] Shim J, Moulson CL, Newberry EP, Lin MH, Xie Y, Kennedy SM, et al. Fatty acidtransport protein 4 is dispensable for intestinal lipid absorption in mice. JLipid Res 2008, doi:10.1194/jlr.M800400-JLR200.

[58] Gertow K, Bellanda M, Eriksson P, Boquist S, Hamsten A, Sunnerhagen M,et al. Genetic and structural evaluation of fatty acid transport protein-4 inrelation to markers of the insulin resistance syndrome. J Clin EndocrinolMetab 2004;89:392–9.

[59] Coburn CT, Knapp Jr FF, Febbraio M, Beets AL, Silverstein RL, Abumrad NA.Defective uptake and utilization of long chain fatty acids in muscle andadipose tissues of CD36 knockout mice. J Biol Chem 2000;275:32523–9.

[60] Hajri T, Ibrahimi A, Coburn CT, Knapp Jr FF, Kurtz T, Pravenec M, et al.Defective fatty acid uptake in the spontaneously hypertensive rat is a primarydeterminant of altered glucose metabolism, hyperinsulinemia, andmyocardial hypertrophy. J Biol Chem 2001;276:23661–6.

[61] Endemann G, Stanton LW, Madden KS, Bryant CM, White RT, Protter AA. CD36is a receptor for oxidized low density lipoprotein. J Biol Chem1993;268:11811–6.

[62] Chen CH, Cartwright Jr J, Li Z, Lou S, Nguyen HH, Gotto Jr AM, et al. Inhibitoryeffects of hypercholesterolemia and ox-LDL on angiogenesis-like endothelialgrowth in rabbit aortic explants. Essential role of basic fibroblast growthfactor. Arterioscl Thromb Vasc Biol 1997;17:1303–12.

[63] Ren Y, Silverstein RL, Allen J, Savill J. CD36 gene transfer confers capacity forphagocytosis of cells undergoing apoptosis. J Exp Med 1995;181:1857–62.

[64] Oquendo P, Hundt E, Lawler J, Seed B. CD36 directly mediates cytoadherenceof plasmodium falciparum parasitized erythrocytes. Cell 1989;58:95–101.

[65] Febbraio M, Hajjar DP, Silverstein RL. CD36: a class B scavenger receptorinvolved in angiogenesis, atherosclerosis, inflammation, and lipidmetabolism. J Clin Invest 2001;108:785–91.

[66] Laugerette F, Passilly-Degrace P, Patris B, Niot I, Febbraio M, Montmayeur JP,et al. CD36 involvement in orosensory detection of dietary lipids,spontaneous fat preference, and digestive secretions. J Clin Invest2005;115:3177–84.

[67] Gaillard D, Laugerette F, Darcel N, El-Yassimi A, Passilly-Degrace P, Hichami A,et al. The gustatory pathway is involved in CD36-mediated orosensoryperception of long-chain fatty acids in the mouse. Faseb J 2008;22:1458–68.

[68] Rac ME, Safranow K, Poncyljusz W. Molecular basis of human CD36 genemutations. Mol Med 2007;13:288–96.

[69] Greenwalt DE, Lipsky RH, Ockenhouse CF, Ikeda H, Tandon NN, Jamieson GA.Membrane glycoprotein CD36: a review of its roles in adherence, signaltransduction, and transfusion medicine. Blood 1992;80:1105–15.

[70] Baillie RA, Jump DB, Clarke SD. Specific effects of polyunsaturated fatty acidson gene expression. Curr Opin Lipidol 1996;7:53–5.

Page 13: Intestinal absorption of long-chain fatty acid

I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115 113

[71] Ibrahimi A, Sfeir Z, Magharaie H, Amri EZ, Grimaldi P, Abumrad NA.Expression of the CD36 homolog (FAT) in fibroblast cells: effects on fattyacid transport. Proc Natl Acad Sci USA 1996;93:2646–51.

[72] Ibrahimi A, Abumrad NA. Role of CD36 in membrane transport of long-chainfatty acids. Curr Opin Clin Nutr Metab Care 2002;5:139–45.

[73] Drover VA, Ajmal M, Nassir F, Davidson NO, Nauli AM, Sahoo D, et al. CD36deficiency impairs intestinal lipid secretion and clearance of chylomicronsfrom the blood. J Clin Invest 2005;115:1290–7.

[74] Hajri T, Hall AM, Jensen DR, Pietka TA, Drover VA, Tao H, et al. CD36-facilitated fatty acid uptake inhibits leptin production and signaling inadipose tissue. Diabetes 2007;56:1872–80.

[75] Poirier H, Degrace P, Niot I, Bernard A, Besnard P. Localization and regulationof the putative membrane fatty-acid transporter (FAT) in the small intestine.Comparison with fatty acid-binding proteins (FABP). Eur J Biochem1996;238:368–73.

[76] Lobo MV, Huerta L, Ruiz-Velasco N, Teixeiro E, de la Cueva P, Celdran A, et al.Localization of the lipid receptors CD36 and CLA-1/SR-BI in the humangastrointestinal tract: towards the identification of receptors mediating theintestinal absorption of dietary lipids. J Histochem Cytochem2001;49:1253–60.

[77] Sukhotnik I, Gork AS, Chen M, Drongowski RA, Coran AG, Harmon CM. Effectof low fat diet on lipid absorption and fatty-acid transport following bowelresection. Pediatr Surg Int 2001;17:259–64.

[78] Nassir F, Wilson B, Han X, Gross RW, Abumrad NA. CD36 is important for fattyacid and cholesterol uptake by the proximal but not distal intestine. J BiolChem 2007;282:19493–501.

[79] Drover VA, Nguyen DV, Bastie CC, Darlington YF, Abumrad NA, Pessin JE, et al.CD36 mediates both cellular uptake of very long chain fatty acids and theirintestinal absorption in mice. J Biol Chem 2008;283:13108–15.

[80] Masuda D, Hirano KI, Oku H, Sandoval JC, Kawase R, Yuasa-Kawase M, et al.Chylomicron remnants are increased in the postprandial state in CD36deficiency. J Lipid Res 2008, doi:10.1194/jlr.P700032-JLR200.

[81] Lisanti MP, Scherer PE, Vidugiriene J, Tang Z, Hermanowski-Vosatka A, Tu YH,et al. Characterization of caveolin-rich membrane domains isolated from anendothelial-rich source: implications for human disease. J Cell Biol1994;126:111–26.

[82] Tao N, Wagner SJ, Lublin DM. CD36 is palmitoylated on both N- and C-terminal cytoplasmic tails. J Biol Chem 1996;271:22315–20.

[83] Galbiati F, Volonte D, Meani D, Milligan G, Lublin DM, Lisanti MP, et al. Thedually acylated NH2-terminal domain of gi1alpha is sufficient to target agreen fluorescent protein reporter to caveolin-enriched plasma membranedomains. Palmitoylation of caveolin-1 is required for the recognition of duallyacylated g-protein alpha subunits in vivo. J Biol Chem 1999;274:5843–50.

[84] Greaves J, Salaun C, Fukata Y, Fukata M, Chamberlain LH. Palmitoylation andmembrane interactions of the neuroprotective chaperone cysteine-stringprotein. J Biol Chem 2008;283:25014–26.

[85] Miranda M, Sorkin A. Regulation of receptors and transporters byubiquitination: new insights into surprisingly similar mechanisms. MolInterv 2007;7:157–67.

[86] Huang M-M, Bolen JB, Barnwell JW, Shattil SJ, Brugge JS. Membraneglycoprotein IV (CD36) is physically associated with the fyn, lyn, and yesprotein-tyrosine kinases in human platelets. Proc Nat Acad Sci USA1991;88:7844–8.

[87] Jimenez B, Volpert OV, Crawford SE, Febbraio M, Silverstein RL, Bouck N.Signals leading to apoptosis-dependent inhibition of neovascularization bythrombospondin-1. Nat Med 2000;6:41–8.

[88] Rahaman SO, Lennon DJ, Febbraio M, Podrez EA, Hazen SL, Silverstein RL. ACD36-dependent signaling cascade is necessary for macrophage foam cellformation. Cell Metab 2006;4:211–21.

[89] El-Yassimi A, Hichami A, Besnard P, Khan NA. Linoleic acid induces calciumsignaling, SRC-kinase phosphorylation and neurotransmitters release inmouse CD36-positive gustatory cells. J Biol Chem 2008;283:12949–59.

[90] Nauli AM, Nassir F, Zheng S, Yang Q, Lo CM, Vonlehmden SB, et al. CD36 isimportant for chylomicron formation and secretion and may mediatecholesterol uptake in the proximal intestine. Gastroenterology2006;131:1197–207.

[91] Schwartz GJ, Fu J, Astarita G, Li X, Gaetani S, Campolongo P, et al. The lipidmessenger OEA links dietary fat intake to satiety. Cell Metab 2008;8:281–8.

[92] Benton R, Vannice KS, Vosshall LB. An essential role for a CD36-relatedreceptor in pheromone detection in Drosophila. Nature 2007;450:289–93.

[93] Storch J, Corsico B. The emerging functions and mechanisms of mammalianfatty acid-binding proteins. Annu Rev Nut 2008;28:73–95.

[94] Thompson J, Winter N, Terwey D, Bratt J, Banaszak L. The crystal structure ofthe liver fatty acid-binding protein. A complex with two bound oleates. J BiolChem 1997;272:7140–50.

[95] Sacchettini JC, Gordon JI. Rat intestinal fatty acid binding protein. A modelsystem for analyzing the forces that can bind fatty acids to proteins. J BiolChem 1993;268:18399–402.

[96] Thompson H, Zhu Z, Banni S, Darcy K, Loftus T, Ip C. Morphological andbiochemical status of the mammary gland as influenced by conjugatedlinoleic acid: implication for a reduction in mammary cancer risk. Cancer Res1997;57:5067–72.

[97] Prows DR, Schroeder F. Metallothionein-IIA promoter induction alters ratintestinal fatty acid binding protein expression, fatty acid uptake, and lipidmetabolism in transfected L-cells. Arch Biochem Biophys 1997;340:135–43.

[98] Atshaves BP, Foxworth WB, Frolov A, Roths JB, Kier AB, Oetama BK, et al.Cellular differentiation and I-FABP protein expression modulate fatty aciduptake and diffusion. Am J Physiol 1998;274:C633–44.

[99] Holehouse EL, Liu ML, Aponte GW. Oleic acid distribution in small intestinalepithelial cells expressing intestinal-fatty acid binding protein. BiochimBiophys Acta 1998;1390:52–64.

[100] Wolfrum C, Ellinghaus P, Fobker M, Seedorf U, Assmann G, Börchers T, et al.Phytanic acid is ligand and transcriptional activator of murine liver fatty acidbinding protein. J Lipid Res 1999;40:708–14.

[101] Hallden G, Aponte GW. Evidence for a role of the gut hormone PYY in theregulation of intestinal fatty acid-binding protein transcripts in differentiatedsubpopulations of intestinal epithelial cell hybrids. J Biol Chem1997;272:12591–600.

[102] Bass NM. The cellular fatty acid binding proteins: aspects of structure,regulation, and function. Int Rev Cytol 1988;111:143–84.

[103] Petit V, Arnould L, Martin P, Monnot MC, Pineau T, Besnard P, et al. Chronichigh-fat diet affects intestinal fat absorption and postprandial triglyceridelevels in the mouse. J Lipid Res 2007;48:278–87.

[104] Hsu M-H, Palmer CNA, Griffin KJ, Johnson EF. A single amino acid change inthe mouse peroxisome proliferator-activated receptor a alters transcriptionalresponses to peroxisome proliferators. Molr Pharmacol 1995;48:559–67.

[105] Corsico B, Cistola DP, Frieden C, Storch J. The helical domain of intestinal fattyacid binding protein is critical for collisional transfer of fatty acids tophospholipid membranes. Proc Natl Acad Sci USA 1998;95:12174–8.

[106] Storch J, Thumser AE. The fatty acid transport function of fatty acid-bindingproteins. Biochim Biophys Acta 2000;1486:28–44.

[107] Alpers DH, Bass NM, Engle MJ, DeSchryver-Kecskemeti K. Intestinal fatty acidbinding protein may favor differential apical fatty acid binding in theintestine. Biochim Biophys Acta 2000;1483:352–62.

[108] Hodsdon ME, Cistola DP. Ligand binding alters the backbone mobility ofintestinal fatty acid-binding protein as monitored by 15N NMR relaxation and1H exchange. Biochemistry 1997;36:2278–90.

[109] Agren JJ, Valve R, Vidgren H, Laakso M, Uusitupa M. Postprandial lipemicresponse is modified by the polymorphism at codon 54 of the fatty acid-binding protein 2 gene. Arterioscl Thromb Vas Biol 1998;18:1606–10.

[110] Baier LJ, Sacchettini JC, Knowler WC, Eads J, Paolisso G, Tataranni PA, et al. Anamino acid substitution in the human intestinal fatty acid binding protein isassociated with increased fatty acid binding, increased fat oxidation, andinsulin resistance. J Clin Invest 1995;95:1281–7.

[111] Hegele RA, Harris SB, Hanley AJ, Sadikian S, Connelly PW, Zinman B. Geneticvariation of intestinal fatty acid-binding protein associated with variation inbody mass in aboriginal Canadians. J Clin Endocrinol Metab 1996;81:4334–7.

[112] Baier LJ, Bogardus C, Sacchettini JC. A polymorphism in the human intestinalfatty acid binding protein alters fatty acid transport across Caco-2 cells. J BiolChem 1996;271:10892–6.

[113] Levy E, Menard D, Delvin E, Stan S, Mitchell G, Lambert M, et al. Thepolymorphism at codon 54 of the FABP2 gene increases fat absorption inhuman intestinal explants. J Biol Chem 2001;276:39679–84.

[114] Darimont C, Gradoux N, de Pover A. Epidermal growth factor regulates fattyacid uptake and metabolism in Caco-2 cells. Am J Physiol 1999;276:G606–12.

[115] Vassileva G, Huwyler L, Poirier K, Agellon LB, Toth MJ. The intestinal fatty acidbinding protein is not essential for dietary fat absorption in mice. Faseb J2000;14:2040–6.

[116] Thumser AE, Wilton DC. The binding of cholesterol and bile salts torecombinant rat liver fatty acid-binding protein. Biochem J1996;320:729–33.

[117] Thompson J, Ory J, Reese-Wagoner A, Banaszak L. The liver fatty acid bindingprotein – comparison of cavity properties of intracellular lipid-bindingproteins. Mol Cell Biochem 1999;192:9–16.

[118] Richieri GV, Ogata RT, Kleinfeld AM. Equilibrium constants for the binding offatty acids with fatty acid-binding proteins from adipocyte, intestine, heart,and liver measured with the fluorescent probe ADIFAB. J Biol Chem1994;269:23918–30.

[119] Poirier H, Braissant O, Niot I, Wahli W, Besnard P. 9-cis-Retinoic acidenhances fatty acid-induced expression of the liver fatty acid-binding proteingene. FEBS Lett 1997;412:480–4.

[120] Wolfrum C, Borrmann CM, Franke WW, Gorski J, Spener F. Fatty acids anddrugs interacting with FABP and PPAR in hepatocytes: a signaling ath to thenucleus. Chem Phys Lipids 1999;101:149.

[121] Neeli I, Siddiqi SA, Siddiqi S, Lagakos WS, Binas B, Gheyi T, et al. Liver fattyacid-binding protein initiates budding of pre-chylomicron transport vesiclesfrom intestinal endoplasmic reticulum. J Biol Chem 2007;282:17974–84.

[122] Newberry EP, Xie Y, Kennedy SM, Luo J, Davidson NO. Protection againstWestern diet-induced obesity and hepatic steatosis in liver fatty acid-bindingprotein knockout mice. Hepatology 2006;44:1191–205.

[123] Knudsen J, Neergaard TB, Gaigg B, Jensen MV, Hansen JK. Role of acyl-CoAbinding protein in acyl-CoA metabolism and acyl-CoA- mediated cellsignaling. J Nutr 2000;130:294S–8S.

[124] Knudsen J, Mandrup S, Rasmussen JT, Andreasen PH, Poulsen F, Kristiansen K.The function of acyl-CoA-binding protein (ACBP)/diazepam binding inhibitor(DBI). Mol Cell Biochem 1993;123:129–38.

[125] Kragelund BB, Andersen KV, Madsen JC, Knudsen J, Poulsen FM. Three-dimensional structure of the complex between acyl-coenzyme A bindingprotein and palmitoyl-coenzyme A. J Mol Biol 1993;230:1260–77.

Page 14: Intestinal absorption of long-chain fatty acid

114 I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115

[126] Knudsen J, Jensen MV, Hansen JK, Faergeman NJ, Neergaard TB, Gaigg B. Roleof acylCoA binding protein in acylCoA transport, metabolism and cellsignaling. Mol Cell Biochem 1999;192:95–103.

[127] Frolov A, Schroeder F. Acyl coenzyme A binding protein conformationalsensitivity to long chain fatty acyl-CoA. J Biol Chem 1998;273:11049–55.

[128] Rosendal J, Ertbjerg P, Knudsen J. Characterization of ligand binding to acyl-CoA-binding protein. Biochem J 1993;290:321–6.

[129] Yanase H, Shimizu H, Kanda T, Fujii H, Iwanaga T. Cellular localization of thediazepam binding inhibitor (DBI) in the gastrointestinal tract of mice and itscoexistence with the fatty acid binding protein (FABP). Arch Histol Cytol2001;64:449–60.

[130] Mandrup S, Sorensen RV, Helledie T, Nohr J, Baldursson T, Gram C, et al.Inhibition of 3T3-L1 adipocyte differentiation by expression of acyl-CoA-binding protein antisense RNA. J Biol Chem 1998;273:23897–903.

[131] Helledie T, Antonius M, Sorensen RV, Hertzel AV, Bernlohr DA, Kolvraa S, et al.Lipid-binding proteins modulate ligand-dependent trans-activation byperoxisome proliferator-activated receptors and localize to the nucleus aswell as the cytoplasm. J Lipid Res 2000;41:1740–51.

[132] Hussain MM, Fatma S, Pan X, Iqbal J. Intestinal lipoprotein assembly. CurrOpin Lipidol 2005;16:281–5.

[133] Mansbach 2nd CM, Gorelick F. Development and physiological regulation ofintestinal lipid absorption. II. Dietary lipid absorption, complex lipidsynthesis, and the intracellular packaging and secretion of chylomicrons.Am J Physiol Gastrointest Liver Physiol 2007;293:G645–50.

[134] Black DD. Development and physiological regulation of intestinal lipidabsorption. I. Development of intestinal lipid absorption: cellular events inchylomicron assembly and secretion. Am J Physiol Gastrointest Liver Physiol2007;293:G519–24.

[135] Bell RM, Ballas LM, Coleman RA. Lipid topogenesis. J Lipid Res1981;22:391–403.

[136] Tso P, Karlstad MD, Bistrian BR, DeMichele SJ. Intestinal digestion, absorption,and transport of structured triglycerides and cholesterol in rats. Am J Physiol1995;268:G568–577.

[137] Cases S, Smith SJ, Zheng YW, Myers HM, Lear SR, Sande E, et al. Identificationof a gene encoding an acyl CoA: diacylglycerol acyltransferase, a key enzymein triacylglycerol synthesis. Proc Natl Acad Sci USA 1998;95:13018–23.

[138] Cases S, Stone SJ, Zhou P, Yen E, Tow B, Lardizabal KD, et al. Cloning of DGAT2,a second mammalian diacylglycerol acyltransferase, and related familymembers. J Biol Chem 2001;276:38870–6.

[139] Buhman KK, Smith SJ, Stone SJ, Repa JJ, Wong JS, Knapp Jr FF, et al. DGAT1 isnot essential for intestinal triacylglycerol absorption or chylomicronsynthesis. J Biol Chem 2002;277:25474–9.

[140] Stone SJ, Myers HM, Watkins SM, Brown BE, Feingold KR, Elias PM, et al.Lipopenia and skin barrier abnormalities in DGAT2-deficient mice. J BiolChem 2004;279:11767–76.

[141] Stone SJ, Levin MC, Farese Jr RV. Membrane topology and identification of keyfunctional amino acid residues of murine acyl-CoA: diacylglycerolacyltransferase-2. J Biol Chem 2006;281:40273–82.

[142] Washington L, Cook GA, Mansbach 2nd CM. Inhibition of carnitinepalmitoyltransferase in the rat small intestine reduces export oftriacylglycerol into the lymph. J Lipid Res 2003;44:1395–403.

[143] Yamazaki T, Sasaki E, Kakinuma C, Yano T, Miura S, Ezaki O. Increased verylow density lipoprotein secretion and gonadal fat mass in miceoverexpressing liver DGAT1. J Biol Chem 2005;280:21506–14.

[144] Lehner R, Kuksis A. Triacylglycerol synthesis by an sn-1,2(2,3)-diacylglyceroltransacylase from rat intestinal microsomes. J Biol Chem 1993;268:8781–6.

[145] Oelkers P, Tinkelenberg A, Erdeniz N, Cromley D, Billheimer JT, Sturley SL. Alecithin cholesterol acyltransferase-like gene mediates diacylglycerolesterification in yeast. J Biol Chem 2000;275:15609–12.

[146] Abo-Hashema KA, Cake MH, Power GW, Clarke D. Evidence for triacylglycerolsynthesis in the lumen of microsomes via a lipolysis-esterification pathwayinvolving carnitine acyltransferases. J Biol Chem 1999;274:35577–82.

[147] Swift LL, Jovanovska A, Kakkad B, Ong DE. Microsomal triglyceride transferprotein expression in mouse intestine. Histochem Cell Biol 2005;123:475–82.

[148] Atzel A, Wetterau JR. Mechanism of microsomal triglyceride transfer proteincatalyzed lipid transport. Biochemistry 1993;32:10444–50.

[149] Smolenaars MM, Madsen O, Rodenburg KW, Van der Horst DJ. Moleculardiversity and evolution of the large lipid transfer protein superfamily. J LipidRes 2007;48:489–502.

[150] Gordon DA, Jamil H. Progress towards understanding the role of microsomaltriglyceride transfer protein in apolipoprotein-B lipoprotein assembly.Biochim Biophys Acta 2000;1486:72–83.

[151] Jamil H, Dickson Jr JK, Chu CH, Lago MW, Rinehart JK, Biller SA, et al.Microsomal triglyceride transfer protein. Specificity of lipid binding andtransport. J Biol Chem 1995;270:6549–54.

[152] Hussain MM, Rava P, Pan X, Dai K, Dougan SK, Iqbal J, et al. Microsomaltriglyceride transfer protein in plasma and cellular lipid metabolism. CurrOpin Lipidol 2008;19:277–84.

[153] Wu X, Zhou M, Huang LS, Wetterau J, Ginsberg HN. Demonstration of aphysical interaction between microsomal triglyceride transfer protein andapolipoprotein B during the assembly of ApoB-containing lipoproteins. J BiolChem 1996;271:10277–81.

[154] Gordon DA, Jamil H, Sharp D, Mullaney D, Yao Z, Gregg RE, et al. Secretion ofapolipoprotein B-containing lipoproteins from HeLa cells is dependent onexpression of the microsomal triglyceride transfer protein and is regulated bylipid availability. Proc Natl Acad Sci USA 1994;91:7628–32.

[155] Hussain MM, Shi J, Dreizen P. Microsomal triglyceride transfer protein and itsrole in apoB-lipoprotein assembly. J Lipid Res 2003;44:22–32.

[156] Hussain MM. A proposed model for the assembly of chylomicrons.Atherosclerosis 2000;148:1–15.

[157] Xie Y, Newberry EP, Young SG, Robine S, Hamilton RL, Wong JS, et al.Compensatory increase in hepatic lipogenesis in mice with conditionalintestine-specific Mttp deficiency. J Biol Chem 2006;281:4075–86.

[158] Berriot-Varoqueaux N, Aggerbeck LP, Samson-Bouma M. Microsomaltriglyceride transfer protein and abetalipoproteinemia. Ann Endocrinol(Paris) 2000;61:125–9.

[159] Karpe F, Olivecrona T, Hamsten A, Hultin M. Chylomicron/chylomicronremnant turnover in humans: evidence for margination of chylomicrons andpoor conversion of larger to smaller chylomicron remnants. J Lipid Res1997;38:949–61.

[160] Levy E, Stan S, Delvin E, Menard D, Shoulders C, Garofalo C, et al. Localizationof microsomal triglyceride transfer protein in the Golgi: possible role in theassembly of chylomicrons. J Biol Chem 2002;277:16470–7.

[161] Slight I, Bendayan M, Malo C, Delvin E, Lambert M, Levy E. Identification ofmicrosomal triglyceride transfer protein in intestinal brush-bordermembrane. Exp Cell Res 2004;300:11–22.

[162] Young SG, Cham CM, Pitas RE, Burri BJ, Connolly A, Flynn L, et al. A geneticmodel for absent chylomicron formation: mice producing apolipoprotein B inthe liver, but not in the intestine. J Clin Invest 1995;96:2932–46.

[163] Kendrick JS, Chan L, Higgins JA. Superior role of apolipoprotein B48 overapolipoprotein B100 in chylomicron assembly and fat absorption: aninvestigation of apobec-1 knock-out and wild-type mice. Biochem J2001;356:821–7.

[164] Lo CM, Nordskog BK, Nauli AM, Zheng S, Vonlehmden SB, Yang Q, et al. Whydoes the gut choose apolipoprotein B48 but not B100 for chylomicronformation? Am J Physiol Gastrointest Liver Physiol 2008;294:G344–52.

[165] Tso P, Liu M, Kalogeris TJ, Thomson AB. The role of apolipoprotein A-IV in theregulation of food intake. Annu Rev Nut 2001;21:231–54.

[166] Lu S, Yao Y, Meng S, Cheng X, Black DD. Overexpression of apolipoprotein A-IV enhances lipid transport in newborn swine intestinal epithelial cells. J BiolChem 2002;277:31929–37.

[167] Lu S, Yao Y, Cheng X, Mitchell S, Leng S, Meng S, et al. Overexpression ofapolipoprotein A-IV enhances lipid secretion in IPEC-1 cells by increasingchylomicron size. J Biol Chem 2006;281:3473–83.

[168] Kumar NS, Mansbach 2nd CM. Prechylomicron transport vesicle: isolationand partial characterization. Am J Physiol 1999;276:G378–86.

[169] Jones B, Jones EL, Bonney SA, Patel HN, Mensenkamp AR, Eichenbaum-VolineS, et al. Mutations in a Sar1 GTPase of COPII vesicles are associated with lipidabsorption disorders. Nat Genet 2003;34:29–31.

[170] Siddiqi SA, Gorelick FS, Mahan JT, Mansbach 2nd CM. COPII proteins arerequired for Golgi fusion but not for endoplasmic reticulum budding of thepre-chylomicron transport vesicle. J Cell Sci 2003;116:415–27.

[171] Siddiqi SA, Mahan J, Siddiqi S, Gorelick FS, Mansbach 2nd CM. Vesicle-associated membrane protein 7 is expressed in intestinal ER. J Cell Sci2006;119:943–50.

[172] Siddiqi SA, Siddiqi S, Mahan J, Peggs K, Gorelick FS, Mansbach 2nd CM. Theidentification of a novel endoplasmic reticulum to Golgi SNARE complex usedby the prechylomicron transport vesicle. J Biol Chem 2006;281:20974–82.

[173] Siddiqi SA, Mansbach 2nd CM. PKC{zeta}-mediated phosphorylation controlsbudding of the pre-chylomicron transport vesicle. J Cell Sci2008;121:2327–38.

[174] Shoulders CC, Stephens DJ, Jones B. The intracellular transport ofchylomicrons requires the small GTPase, Sar1b. Curr Opin Lipidol2004;15:191–7.

[175] Sabesin SM, Frase S, Ragland JB. Accumulation of nascent lipoproteins in rathepatic Golgi during induction of fatty liver by orotic acid. Lab Invest1977;37:127–35.

[176] Buschmann RJ, Manke DJ. Morphometric analysis of the membranes andorganelles of small intestinal enterocytes. II. lipid-fed hamster. J UltrastructRes 1981;76:15–26.

[177] Mansbach CM, Dowell R. Effect of increasing lipid loads on the ability of theendoplasmic reticulum to transport lipid to the Golgi. J Lipid Res2000;41:605–12.

[178] Cartwright IJ, Plonne D, Higgins JA. Intracellular events in the assembly ofchylomicrons in rabbit enterocytes. J Lipid Res 2000;41:1728–39.

[179] Robertson MD, Parkes M, Warren BF, Ferguson DJ, Jackson KG, Jewell DP, et al.Mobilisation of enterocyte fat stores by oral glucose in humans. Gut2003;52:834–9.

[180] Chateau D, Pauquai T, Delers F, Rousset M, Chambaz J, Demignot S. Lipidmicelles stimulate the secretion of triglyceride-enriched apolipoprotein B48-containing lipoproteins by Caco-2 cells. J Cell Physiol 2005;202:767–76.

[181] Tauchi-Sato K, Ozeki S, Houjou T, Taguchi R, Fujimoto T. The surface of lipiddroplets is a phospholipid monolayer with a unique fatty acid composition. JBiol Chem 2002;277:44507–12.

[182] Mahan JT, Heda GD, Rao RH, Mansbach 2nd CM. The intestine expressespancreatic triacylglycerol lipase: regulation by dietary lipid. Am J PhysiolGastrointest Liver Physiol 2001;280:G1187–96.

[183] Grober J, Lucas S, Sorhede-Winzell M, Zaghini I, Mairal A, Contreras JA, et al.Hormone-sensitive lipase is a cholesterol esterase of the intestinal mucosa. JBiol Chem 2003;278:6510–5.

[184] Trickett JI, Patel DD, Knight BL, Saggerson ED, Gibbons GF, Pease RJ.Characterization of the rodent genes for arylacetamide deacetylase, a

Page 15: Intestinal absorption of long-chain fatty acid

I. Niot et al. / Progress in Lipid Research 48 (2009) 101–115 115

putative microsomal lipase, and evidence for transcriptional regulation. J BiolChem 2001;276:39522–32.

[185] Thomson AB, Cheeseman CI, Keelan M, Fedorak R, Clandinin MT. Crypt cellproduction rate, enterocyte turnover time and appearance of transport alongthe jejunal villus of the rat. Biochim Biophys Acta 1994;1191:197–204.

[186] Thomson AB, Keelan M, Clandinin MT, Walker K. Dietary fat selectively alterstransport properties of rat jejunum. J Clin Invest 1986;77:279–88.

[187] Kondo H, Minegishi Y, Komine Y, Mori T, Matsumoto I, Abe K, et al.Differential regulation of intestinal lipid metabolism-related genes inobesity-resistant A/J vs. obesity-prone C57BL/6J mice. Am J PhysiolEndocrinol Metab 2006;291:E1092–9.

[188] de Wit NJ, Bosch-Vermeulen H, de Groot PJ, Hooiveld GJ, Bromhaar MM,Jansen J, et al. The role of the small intestine in the development of dietaryfat-induced obesity and insulin resistance in C57BL/6J mice. BMC MedGenomics 2008;1:14.

[189] Zilversmit DB. Atherogenesis: a postprandial phenomenon. Circulation1979;60:473–85.

[190] Ginsberg HN. Hypertriglyceridemia: new insights and new approaches topharmacologic therapy. Am J Cardiol 2001;87:1174–80. A1174.

[191] Lopez-Miranda J, Perez-Martinez P, Marin C, Moreno JA, Gomez P, Perez-Jimenez F. Postprandial lipoprotein metabolism, genes and risk ofcardiovascular disease. Curr Opin Lipidol 2006;17:132–8.

[192] Quarfordt SH, Goodman DS. Heterogeneity in the rate of plasma clearance ofchylomicrons of different size. Biochim Biophys Acta 1966;116:382–5.

[193] Martins IJ, Mortimer BC, Miller J, Redgrave TG. Effects of particle size andnumber on the plasma clearance of chylomicrons and remnants. J Lipid Res1996;37:2696–705.

[194] Xiang SQ, Cianflone K, Kalant D, Sniderman AD. Differential binding oftriglyceride-rich lipoproteins to lipoprotein lipase. J Lipid Res1999;40:1655–63.

[195] Olivecrona G, Beisiegel U. Lipid binding of apolipoprotein CII is required forstimulation of lipoprotein lipase activity against apolipoprotein CII-deficientchylomicrons. Arterioscl Thromb Vasc Biolo 1997;17:1545–9.

[196] Hayashi H, Fujimoto K, Cardelli JA, Nutting DF, Bergstedt S, Tso P. Fat feedingincreases size, but not number, of chylomicrons produced by small intestine.Am J Physiol 1990;259:G709–719.

[197] Levy E, Roy CC, Goldstein R, Bar-On H, Ziv E. Metabolic fate of chylomicronsobtained from rats maintained on diets varying in fatty acid composition. JAm Coll Nutr 1991;10:69–78.

[198] Sakr SW, Attia N, Haourigui M, Paul JL, Soni T, Vacher D, et al. Fatty acidcomposition of an oral load affects chylomicron size in human subjects. Br JNut 1997;77:19–31.

[199] Cartwright IJ, Higgins JA. Increased dietary triacylglycerol markedly enhancesthe ability of isolated rabbit enterocytes to secrete chylomicrons: an effectrelated to dietary fatty acid composition. J Lipid Res 1999;40:1858–66.

[200] Lairon D. Macronutrient intake and modulation on chylomicron productionand clearance. Atheroscler Suppl 2008;9:45–8.

[201] Adeli K, Lewis GF. Intestinal lipoprotein overproduction in insulin-resistantstates. Curr Opin Lipidol 2008;19:221–8.

[202] Haidari M, Leung N, Mahbub F, Uffelman KD, Kohen-Avramoglu R, Lewis GF,et al. Fasting and postprandial overproduction of intestinally derivedlipoproteins in an animal model of insulin resistance. Evidence that chronicfructose feeding in the hamster is accompanied by enhanced intestinal denovo lipogenesis and ApoB48-containing lipoprotein overproduction. J BiolChem 2002;277:31646–55.

[203] Phillips C, Bennett A, Anderton K, Owens D, Collins P, White D, et al. Intestinalrather than hepatic microsomal triglyceride transfer protein as a cause ofpostprandial dyslipidemia in diabetes. Metabolism 2002;51:847–52.

[204] Curtin A, Deegan P, Owens D, Collins P, Johnson A, Tomkin GH. Elevatedtriglyceride-rich lipoproteins in diabetes. A study of apolipoprotein B-48.Acta Diabetol 1996;33:205–10.

[205] Duez H, Lamarche B, Uffelman KD, Valero R, Cohn JS, Lewis GF.Hyperinsulinemia is associated with increased production rate of intestinalapolipoprotein B-48-containing lipoproteins in humans. Arterioscl ThrombVasc Biol 2006;26:1357–63.


Recommended