The HDL Handbook Biological Functions and Clinical Implications
Tsugikazu Komoda
AMSTERDAM l BOSTON l HEIDELBERG l LONDON NEW YORK l OXFORD l PARIS l SAN DIEGO SAN FRANCISCO l SINGAPORE l SYDNEY l TOKYO Academic Press is an imprint of Elsevier
Academic Press is an imprint of Elsevier 32 Jamestown Road, London NW1 7BY, UK 30 Corporate Drive, Suite 400, Burlington, MA 01803, USA 525 B Street, Suite 1800, San Diego, CA 92101-4495, USA First edition 2010 Copyright Ó 2010 Elsevier Inc. All rights reserved No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means electronic, mechanical, photocopying, recording or otherwise without the prior written permission of the publisher. Permissions may be sought directly from Elsevier’s Science & Technology Rights Department in Oxford, UK: phone (+ 44) (0) 1865 843830; fax (+44) (0) 1865 853333; email:
[email protected]. Alternatively, visit the Science and Technology Books website at www.elsevierdirect.com/rights for further information Notice No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library Library of Congress Cataloging-in-Publication Data A catalog record for this book is available from the Library of Congress ISBN: 978-0-12-382171-3 For information on all Academic Press publications visit our website at elsevierdirect.com Typeset by TNQ Books and Journals Pvt Ltd. www.tnq.co.in Printed and bound in China 10 11 12 13 14 15 10 9 8 7 6 5 4 3 2 1
Preface
When I started to write this book “The HDL Handbook: Biological Functions and Clinical Applications”, Professor Takashi Miida of Juntendo University in Tokyo strongly supported my plan. Of course, he is also an excellent contributor to this book. In addition, Professor David Alpers, Professor Emeritus of Washington University, School of Medicine, well revised the present HDL book. Therefore, I want to thank him for his revision of this book. Unfortunately, since planning to publish this HDL book, two and half years have passed. Some contributors immediately accepted my planning, however, half of the contributors have not been able to complete the publication of “The HDL Handbook: Biological Functions and Clinical Applications”. However, the contents of this HDL book include up-to-date progress on HDL research. The contents of this book are a crystallization of the work from all the contributors, because, despite being busy, all the contributors carefully revised their chapters under the suitable comments from eight reviewers. Therefore, if you read this book, you will be fascinated by the renewal of development of HDL researches and the present book is a very useful tool not only for basic researchers in institutes or pharmaceutical companies but also practical physicians. In addition, this book will be evaluated as an HDL Bible for medical and co-medical graduate students by their counselors. Furthermore, since this book is small, it is portable. However, the contents of this HDL book contain the latest news of HDL molecules. Finally, I believe that this book should be read to give an excellent impression of the HDL fields. Tsugikazu Komoda, MD
Contributors
Aishah Al-Jarallah, Department of Biochemistry and Biomedical Sciences, McMaster University, Hamilton, Ontario, Canada G.M. Anantharamaiah, Departments of Medicine, and Biochemistry and Molecular Genetics, University of Alabama at Birmingham, Birmingham, AL, USA Rachelle Brunet, Department of Biochemistry and Biomedical Sciences, McMaster University, Hamilton, Ontario, Canada Giovanna Catalano, INSERM UMRS939, Hoˆpital de la Pitie´, Paris, France; UPMC Universite´ Pierre et Marie Curie, Hoˆpital de la Pitie´, Paris, France Eric Chabrie`re, Institut de Recherche Biome´dicale des Arme´es-Antenne CRSSA, De´partement de Toxicologie, Groupe Bioe´purateurs Catalytiques et Re´activateurs, La Tronche, France; Architecture et Fonction des Macromole´cules Biologiques, Groupe Biocristallographie, Biotechnologie et Enzymologie Structurale, Universite´ de la Me´diterrane´e, Marseille, France Geeta Datta, Department of Medicine, and Biochemistry, University of Alabama at Birmingham, Birmingham, AL, USA Maryse Guerin, INSERM UMRS939, Hoˆpital de la Pitie´, Paris, France; UPMC Universite´ Pierre et Marie Curie, Hoˆpital de la Pitie´, Paris, France Akira Hara, Laboratory of Biochemistry, Gifu Pharmaceutical University, Japan Hiroaki Hattori, Advanced Technology and Development Division, BML Inc., 1361-1 Matoba, Kawagoe, Saitama, Japan Neil J. Hime, Centre for Vascular Research, Sydney Medical School (Pathology) and Bosch Institute, The University of Sydney, Medical Foundation Building, Camperdown, NSW, Australia Satoshi Hirayama, Department of Laboratory Medicine, Juntendo University School of Medicine, Tokyo, Japan Akihiro Inazu, Department of Laboratory Sciences, School of Health Sciences, Institute of Medical, Pharmaceutical and Health Sciences, Kanazawa University, Ishikawa, Japan Zorana Jelic-Ivanovic, Institute of Medical Biochemistry, Faculty of Pharmacy, University of Belgrade, Belgrade, Serbia
x
Contributors
Tsugikazu Komoda, Nihon Medical Science Institute, II-4 Minami-Tohrimachi, Kawagoe, Saitama, Japan Jelena Kotur-Stevuljevic, Institute of Medical Biochemistry, Faculty of Pharmacy, University of Belgrade, Belgrade, Serbia Patrick Masson, Institut de Recherche Biome´dicale des Arme´es-Antenne CRSSA, De´partement de Toxicologie, Groupe Bioe´purateurs Catalytiques et Re´activateurs, La Tronche, France Akira Matsunaga, Department of Laboratory Medicine, Fukuoka University School of Medicine, Fukuoka, Japan Toshiyuki Matsunaga, Laboratory of Biochemistry, Gifu Pharmaceutical University, Japan Takashi Miida, Department of Laboratory Medicine, Juntendo University School of Medicine, Tokyo, Japan Takanari Nakano, Department of Biochemistry, Faculty of Medicine, Saitama Medical University, Saitama, Japan; Brentwood Biomedical Research Institute, Department of Medicine, School of Medicine, University of California Los Angeles, Los Angeles, CA, USA Daniel Rochu, Institut de Recherche Biome´dicale des Arme´es-Antenne CRSSA, De´partement de Toxicologie, Groupe Bioe´purateurs Catalytiques et Re´activateurs, La Tronche, France; Bundeswehr Institute of Pharmacology and Toxicology, Munich, Germany Keijiro Saku, Department of Cardiology, Fukuoka University School of Medicine, Fukuoka, Japan Makoto Seo, Department of Biochemistry, Faculty of Medicine, Saitama Medical University, Saitama, Japan Slavica Spasic, Institute of Medical Biochemistry, Faculty of Pharmacy, University of Belgrade, Belgrade, Serbia Vesna Spasojevic-Kalimanovska, Institute of Medical Biochemistry, Faculty of Pharmacy, University of Belgrade, Belgrade, Serbia Aleksandra Stefanovic, Institute of Medical Biochemistry, Faculty of Pharmacy, University of Belgrade, Belgrade, Serbia Naoki Terasaka, Biological Research Laboratories, Daiichi Sankyo Co., Ltd, Tokyo, Japan Bernardo Trigatti, Department of Biochemistry and Biomedical Sciences, McMaster University, Hamilton, Ontario, Canada Yoshinari Uehara, Department of Cardiology, Fukuoka University School of Medicine, Fukuoka, Japan Jelena Vekic, Institute of Medical Biochemistry, Faculty of Pharmacy, University of Belgrade, Belgrade, Serbia
Contributors
xi
C. Roger White, Department of Medicine, and Biochemistry, University of Alabama at Birmingham, Birmingham, AL, USA Aleksandra Zeljkovic, Institute of Medical Biochemistry, Faculty of Pharmacy, University of Belgrade, Belgrade, Serbia Bo Zhang, Department of Cardiology, Fukuoka University School of Medicine, Fukuoka, Japan
Chapter 1
Role of Phospholipid Transfer Protein in HDL Remodeling and Atherosclerosis Hiroaki Hattori Advanced Technology and Development Division, BML Inc., Kawagoe, Saitama, Japan
INTRODUCTION Phospholipid transfer protein (PLTP) plays an important role in the regulation of high density lipoprotein (HDL) metabolism. The regulatory role of PLTP is achieved via its two main functions, phospholipid transfer activity (Tall et al., 1983; Rao et al., 1997) and the ability to modulate HDL size and composition in a process called HDL remodeling (Rye and Barter, 1986; Tu et al., 1993; Jauhiainen et al., 1993). The regulation of HDL metabolism is achieved by the concerted action of a number of plasma and cellular factors. These include the cellular receptors, scavenger receptor class B type 1 (SR-B1) and ATP-binding cassette transporter A1 (ABC-A1), as well as plasma proteins such as cholesteryl ester transfer protein (CETP), lecithin-cholesterol acyltransferase (LCAT), and the endothelial-bound enzymes, lipoprotein lipase (LPL) and triglyceride (TG) hydrolase hepatic lipase (HL). As indicated by the inverse relationship between HDL cholesterol and incidence of coronary heart disease in many epidemiological studies (Gordon and Rifkind, 1989), the plasma HDL level has a major impact on the progression of atherosclerosis. Although the exact mechanism behind the athero-protective role of HDL is still not fully understood, the reverse cholesterol transport (RCT) hypothesis has been widely accepted (Curtiss et al., 2006). Reverse cholesterol transport is the process by which cholesterol is transported from peripheral cells to the liver for elimination (Eisenberg, 1984). Preb-HDL particles, a subpopulation of HDL, act as efficient acceptors in the efflux process of cholesterol at the plasma membrane of peripheral cells (Eisenberg, 1984). PLTP is able to generate preb-HDL particles through HDL remodeling, and has a major role also in maintaining The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
1
2
The HDL Handbook
FIGURE 1.1 The physiological role of PLTP in HDL metabolism. Participation of PLTP is illustrated by bold arrows. The functions of PLTP are: (i) transfer of surface remnants (phospholipids and cholesterol) upon lipolysis of triglyceride-rich lipoproteins; (ii) generation of preb-HDL during remodeling of HDL; and (iii) transfer of a-tocopherol from HDL particles to cell membranes. Abbreviations used: VLDL, very low density lipoprotein; LDL, low density lipoprotein; IDL, intermediate density lipoprotein; HDL, high density lipoprotein; Rem, remnant; PLTP, phospholipid transfer protein; CETP, cholesteryl ester transfer protein; LCAT, lecithin-cholesterol acyltransferase; HL, hepatic lipase; LPL, lipoprotein lipase; SR-BI, scavenger receptor class B type I; ABC-A1, ATP-binding cassette transporter A1; a-T, a-tocopherol.
plasma HDL levels owing to its ability to transport surface remnants produced by lipolysis of triglyceride-rich lipoproteins (Figure 1.1). Thus, PLTP can be envisioned to play an important role in the prevention of atherosclerosis (Castro and Fielding, 1988; von Eckardstein et al., 1996; van Haperen et al., 2000; Tall and Lalanne, 2003). In spite of these effects on HDL, studies in genetically modified mouse models have suggested that systemic PLTP deficiency is athero-protective in vivo, and that PLTP overexpression is pro-atherogenic. Recent studies have focused on this apparent inconsistency, and have examined the effects of local PLTP on atherogenesis, using transplanted macrophages.
MOLECULAR BIOLOGY AND STRUCTURE OF PHOSPHOLIPID TRANSFER PROTEIN The PLTP gene is located on chromosome 20 (20q12-q13.1), and has a length of 13.3 kilobases, including 15 introns. PLTP cDNA is 1750 bp in length, and encodes a 17 amino acid hydrophobic signal peptide and a 476 amino acid mature protein (Day et al., 1994). Most tissues show expression of PLTP
Chapter j 1
Role of Phospholipid Transfer Protein
3
mRNA, but liver and adipose tissue are probably the major contributors to plasma PLTP (Dusserre et al., 2000). Although the predicted molecular weight mass of the mature protein is 55 kDa, plasma PLTP appears as an 80-kDa protein by sodium dodecylsulfate-polyacrylamide gel electrophoresis (SDSPAGE) under reducing and non-reducing conditions (Oka et al., 2000a). The discrepancy between the calculated molecular weight mass and the mass estimated by SDS-PAGE may be explained by the fact that PLTP has six potential N-glycosylation sites and numerous O-glycosylation sites (Day et al., 1994). As the protein additionally contains four cysteine residues, it also has the potential to form two intra-chain disulfide bonds. In contrast to apolipoproteins that are primarily hydrophilic, PLTP has a high content of hydrophobic residues scattered throughout, with over 40% of the amino acids being hydrophobic. Three other hydrophobic proteins of the lipid transfer/lipopolysaccharide binding protein family, namely lipopolysaccharide-binding protein (LBP), neutrophil bactericidal permeability-increasing protein (BPI), and CETP share structural homology with PLTP. However, these proteins also exhibit significant structural differences (Albers et al., 1996). For example, the carboxyl terminal portion of CETP is the most hydrophobic of these four proteins, and its main function is to bind and transfer neutral lipids (Tall et al., 1983; Albers et al., 1984). Although the carboxyl terminal portion of PLTP is somewhat hydrophobic, it does not have the functional capacity to transfer neutral lipids (Tollefson et al., 1988). Unlike PLTP, BPI has a very basic amino-terminal domain, which is responsible for its cytotoxic activity, whereas the hydrophobic carboxyl-terminal domain is believed to anchor the protein in the granule membrane (Gray et al., 1989). LBP and BPI share 44% amino acid sequence identity to bind lipopolysaccharide (Schumann et al., 1990). Furthermore, both proteins have similar amino-terminal amino acids (Tobias et al., 1988). Secondary structure predictions suggest that PLTP has two potential transmembrane regions spanning from residues 169 through 181, and residues 288 through 304 (Albers et al., 1996). Two potential disulfide bonds exist between cysteine residues 5 and 129, and between cysteine residues 168 and 318. The cysteine residues 146 and 185 form a disulfide bridge that is essential for the correct folding and secretion of PLTP (Huuskonen et al., 1998, 1999; Qu et al., 1999). We have shown that two forms of PLTP exist in human plasma, one with high activity (HA-PLTP) and another with low activity (LA-PLTP) (Oka et al., 2000b; Ka¨rkka¨inen et al., 2002). The LA-PLTP is associated with apoA-I, and the HA-PLTP co-purifies with apoE (Murdoch et al., 2002). LA-PLTP is located between LDL and HDL on size-exclusion chromatography, having an apparent molecular mass of 520 kDa and a Stokes diameter of 12e17 nm (Oka et al., 2000a; Murdoch et al., 2002). In contrast, HA-PLTP is associated with an average molecular mass of 160 kDa and a Stokes diameter between 7.6 and 12.0 nm. HA-PLTP but not LA-PLTP is able to remodel HDL, resulting in the formation of two types of particles, preb-HDL and large fused HDL (Vikstedt
4
The HDL Handbook
et al., 2007a). The mechanism by which these two HDL subpopulations are generated is not known.
PLTP-MEDIATED LIPID TRANSFER PLTP is a rather non-specific lipid transfer protein. Several studies have shown that it is able to transfer all common phospholipid types (Massey et al., 1984; Huuskonen et al., 1996; Rao et al., 1997). Diacylglyceride, a significant lipid constituent of HDL particles, is also transferred efficiently (Rao et al., 1997). The PLTP-mediated transfer of phosphatidylcholine molecules varies with their fatty acyl chain length, decreasing with increasing length, a phenomenon similar to that observed for spontaneous transfer (Huuskonen et al., 1996). PLTP functions by forming a ternary complex between donor and acceptor lipoprotein particles, and may facilitate phospholipid transfer by lowering the energy barrier for lipid monomer dissociation and/or association (Huuskonen et al., 2001). Among lipoproteins, HDL particles are the most efficient as phospholipid donors and/or acceptors in PLTP-mediated phospholipid transfer (Rao et al., 1997). Thus, PLTP appears to display specificity towards HDL in the transfer reaction. As HDLs are by far the most numerous lipoprotein particles in plasma, collisions are more likely to occur between PLTP and them than between PLTP and other lipoproteins (Huuskonen et al., 2001). Two lipidbinding pockets in PLTP protein, at the entrance to and inside the pockets, participate in the phospholipid transfer process (Huuskonen et al., 1999). It has recently been demonstrated that the amphipathic a-helix in PLTP at amino acid residues 144e163, at the tip of the N-terminal barrel, is critical for removing the lipid domain formed by ABC-A1, interacting with phospholipids, and promoting cellular cholesterol and phospholipid transfer (Oram et al., 2008). This suggests that PLTP may shuttle lipids between cells and HDL particles. PLTP also binds and transfers several other amphipathic molecules, including a-tocopherol, diacylglycerides, cerebrosides and lipopolysaccharides (Lagrost et al., 1998; Desrumaux et al., 1999).
PLTP-MEDIATED HDL REMODELING In humans, both CETP and PLTP catalyze the conversion of HDL in vitro into larger and smaller particles, including preb-HDL (Albers et al., 1995; Pulcini et al., 1995; Rye et al., 1995, 1997; Curtiss et al., 2000). The physiological significance of this process is that it increases the production of preb-HDL and thereby enhances the capacity of HDL to accept cellular cholesterol (von Eckardstein et al., 1996). HDL remodeling is also catalyzed by pig and mouse PLTP proteins (Pussinen et al., 1995; Marques-Vidal et al., 1997). PLTP has also been shown to mediate the conversion of HDL2 particles with a concomitant release of lipid-poor apoA-I (Marques-Vidal et al., 1997), and to transform reconstituted discoidal HDL particles into vesicular structures (Nishida et al.,
Chapter j 1
Role of Phospholipid Transfer Protein
5
1997). By using HDL particles containing phospholipids in the surface lipid layer or cholesteryl esters in the core, it has been demonstrated that PLTP rapidly transfers the surface phospholipids and reaches equilibrium prior to mixing of the cholesteryl ester core (Lusa et al., 1996). It is thought that the initial reaction in PLTP-mediated HDL remodeling is phospholipid transfer, followed by the release of lipid-poor apoA-I, and subsequent fusion of the unstable particles (Lusa et al., 1996). CETP also catalyzes HDL remodeling (Pulcini et al., 1995; Rye et al., 1997). Although particle fusion is again involved (Pulcini et al., 1995), there are several differences between the PLTP-mediated and CETP-mediated processes. No release of lipid-poor apoA-I was observed in HDL remodeling by CETP (Albers et al., 1995). Furthermore, PLTP promoted preferentially the formation of large HDL particles, whereas CETP favored the production of small HDL (Pulcini et al., 1995; Lagrost et al., 1998). The process of HDL remodeling by PLTP is controlled by at least three different factors: the HDL apolipoprotein/protein composition; the core lipid composition of HDL; and the phospholipid transfer activity of PLTP (Pussinen et al., 1997, 2001; Rye et al., 1998; Huuskonen et al., 2000a). In pig HDL, which natively contains apoA-I but not apoA-II, both the formation of large particles and the release of apoA-I were inhibited by increasing the content of apoA-II (Pussinen et al., 1997). Addition of purified serum amyloid A (SAA), an acute-phase upregulated protein, into HDL3 particles facilitated their ability to undergo HDL remodeling, despite the lower content of apoA-I in SAA-HDL (Pussinen et al., 2001). HDL particles enriched in core triglycerides exhibited enhanced PLTP-mediated HDL remodeling (Rye et al., 1998). Phospholipid transfer is a prerequisite for efficient PLTP-mediated HDL remodeling (Huuskonen et al., 2000a). PLTP can also remodel triglyceride-rich lipoproteins (TGRLPs) (Rye et al., 1998). Particularly during the postprandial phase, lipolysis of TGRLPs is intricately linked to subsequent remodeling by PLTP, which results in the integration of lipolytic surface remnants into HDL (Huuskonen et al., 2001), profoundly altering HDL speciation (Jauhiainen et al., 1993). The ensuing imbalance between lipolysis and remodeling may contribute to the toxicity of lipolyzed TGRLPs, leading to damage to vascular endothelial cells and/or macrophage apoptosis (Wehinger et al., 2007). Moreover, PLTP remodeled spherical apoE-containing HDL into large and small particles without dissociation of apoE, suggesting that apoE enhances the capacity of PLTP to remodel HDL, but reduces the ability of HDL to participate in PLTP-mediated phospholipid transfers (Settasatian et al., 2008).
GENETICALLY MODIFIED ANIMALS Plasma from mice overexpressing human PLTP was more effective in preventing acetylated LDL-induced accretion of cholesteryl ester in macrophages
6
The HDL Handbook
(von Eckardstein et al., 1996). Nevertheless, systemic PLTP deficiency in mice is associated with a decrease in atherosclerosis susceptibility, despite a decrease in plasma HDL levels, and overexpression of PLTP is accompanied by an increase in susceptibility to the disease (von Eckardstein et al., 1996; Jiang et al., 2001, 2002; van Haperen et al., 2002; Yang et al., 2003). Overexpression of human PLTP in transgenic mice (hPLTPtg) also increases preb-HDL generation, but decreases plasma HDL cholesterol (Jiang et al., 1996; von Eckardstein et al., 1996). Although preb-HDL is a minor HDL subpopulation, it is believed to be a very efficient acceptor of cellular cholesterol (Oram and Yokoyama, 1996). Therefore an elevated production of preb-HDL might enhance RCT (Huuskonen et al., 2001). These results suggest that PLTP may act as a pro-atherogenic, rather than an anti-atherogenic factor in vivo. Overexpression of hPLTP in LDL receptor-deficient (LDLr/) mice resulted in a dose-dependent decrease in HDL cholesterol, and a stimulation of VLDL secretion (van Haperen et al., 2002). LDLr/ and hPLTPtg/LDLr/ mice fed an atherogenic diet showed similar levels of hypercholesterolemia, but the hPLTPtg/LDLr/ mice had lower plasma HDL levels and more atherosclerosis (van Haperen et al., 2002). In LDLrþ/ mice expressing both hCETP and hPLTP, PLTP expression was accompanied by a dose-dependent decrease in HDL cholesterol and an increase in atherosclerosis, which was positively related to the level of hPLTP expression (Lie et al., 2004). These findings were confirmed and extended in apoE-deficient (apoE/) mice transfected with mouse PLTP using an adenovirus vector (Yang et al., 2003). On a chow diet, transient expression of PLTP caused a 30% reduction of HDL cholesterol and phospholipids, as well as a 20% reduction of plasma a-tocopherol content, resulting in increased susceptibility of apoB-containing lipoproteins to peroxidation in vitro (Yang et al., 2003). However, while complete deficiency of PLTP decreased both plasma HDL, apoB-secretion and atherosclerosis (Jiang et al., 2001), no such metabolic effects were produced by an approximately 50% decrease in PLTP activity in PLTPþ/ mice (Tall and Lalanne, 2003). Pronounced overexpression of PLTP increased atherosclerosis (van Haperen et al., 2002; Yang et al., 2003). However, moderate expression by two-fold in transgenics (van Haperen et al., 2002), and overexpression using low-dose adeno-associated virus, did not do so (Yang et al., 2003). Transgenic mice overexpressing a mutant PLTP, lacking phospholipid transfer activity, showed no increase in atherosclerosis lesion size (Samyn et al., 2008). Mice expressing inducible PLTP showed a strongly atherogenic lipoprotein profile, an increase of pre-existing atherosclerotic lesion size, and a decrease in lesion stability (Moreland et al., 2008). The role of PLTP deficiency in atherosclerosis has been studied in several mouse models: human apoB transgenic, apoE/ and LDLr/ mice (Jiang et al., 2001). PLTP/ mice of different genetic backgrounds showed a reduction of atherosclerosis, which was related to the reduction of apoB lipoproteins. Mechanistically, depletion of apoB lipoproteins of vitamin E by
Chapter j 1
Role of Phospholipid Transfer Protein
7
PLTP was proposed to account for the reduction of atherosclerosis in PLTPdeficient mice (Jiang et al., 2002). While vitamin E reduces atherosclerosis in mice, vitamin E supplementation in humans has had no demonstrable effect on atherosclerosis (Navab et al., 2004), suggesting a more prominent role of vitamin E in murine than in human atherosclerosis. In PLTP-deficient mice, decreased LDL, HDL and apoE-rich lipoproteins, as well as increased biliary cholesterol secretion and hepatic ABCG5/ABCG8 gene expression and decreased intestinal cholesterol absorption, were observed (Shelly et al., 2008). Furthermore, PLTP-deficient mice showed reduced expression of the proinflammatory genes, intracellular adhesion molecule-1 (ICAM-1) and vascular cell adhesion molecule-1 (VCAM-1), and lower levels of interleukin-6 (IL-6), supporting the concept that PLTP may play a pro-atherogenic role. Taken together, the available data suggest that only very pronounced systemic variations of PLTP influence atherosclerosis in mice. Such variations have not been described in humans (Ja¨nis et al., 2004). Local PLTP effects have attracted attention in relation to the function of macrophages in atherosclerogenesis (Valenta et al., 2006, 2008; Liu et al., 2007; Vikstedt et al., 2007b). Bone marrow transplantation from PLTP/ mice into several background mice has been performed by three groups, yielding contradictory results (Table 1.1). In male LDLr/ mice, Valenta et al. (2006) demonstrated an increase in atherosclerosis after transplantation of PLTP/ bone marrow, suggesting an athero-protective potential of macrophage PLTP in the presence of normal apoA-I levels, but not in the presence of increased apoA-I levels. Moreover, double mutant LDLr/ PLTP/ mice transplanted with PLTP/ bone marrow showed decreased plasma total cholesterol, increased HDL cholesterol, and smaller atherosclerotic lesions (Valenta et al., 2008). In contrast, Vikstedt et al. (2007b) demonstrated a decrease in both plasma PLTP activity and atherosclerosis after transplantation of PLTP/ bone marrow in female LDLr/ mice, implying an atherogenic potential of macrophage PLTP. ApoE/ mice showed a pronounced increase in atherosclerosis after transplantation of PLTP/ bone marrow, associated with a lower secretion of apoE by macrophages, and consequently much higher plasma cholesterol levels (Liu et al., 2007). In male LDLr/ mice, macrophage-derived PLTP was athero-protective, suggesting that the influence of PLTP on atherogenesis is highly dependent upon its site of expression (Vikstedt et al., 2007b). Moreover, mechanisms affecting atherosclerosis, such as transport of vitamin E by lipoproteins, may profoundly differ between species. In C57Bl/6 mice transplanted with PLTP/ bone marrow, increased cholesterol content of the PLTP/ macrophages was observed, which was normalized by a-tocopherol supplementation, suggesting that macrophage-derived PLTP is anti-atherogenic owing to its ability to reduce cholesterol accumulation in macrophages through changes in the a-tocopherol content and antioxidative status of the cells (Valenta et al., 2008). More recently, van Haperen et al. (2008) demonstrated an atherogenic effect of PLTP
8
TABLE 1.1 Differences between mouse models in the effect of macrophage PLTP on atherogenesis after bone marrow transplantation BM donors
Diet/endpoint
Effect on atherosclerosis
Reference
msapoAI/ LDLr/ huapoAItg
PLTP/ or wild type
HFD [15.8% (w/w) fat/ 1.25% (w/w) cholesterol/ no cholate] 16 weeks
Anti-atherogenic 1.3-fold decrease in PLTPwt/wt LDLr/ mice (aorta and heart valve areas); no significance in huapoAItg msapoAI LDLr/ mice
Valenta et al. (2006)
LDLr/, female
PLTP/ or wild type, male
WTD [15% fat/0.25% cholesterol] 9 weeks
Atherogenic 29% reduction in aortic root lesion in PLTP/ LDLr/ mice
Vikstedt et al. (2007b)
ApoE/
PLTP/ or wild type
WTD [0.15% cholesterol/ 20% saturated fat] 7 months
Anti-atherogenic Increased atherosclerosis Aortic arch and root (403%), and aorta (298%) in PLTP/ apoE/ mice
Liu et al. (2007)
PLTP/ LDLr/
PLTP/ or wild type
HFD [15.8% (w/w) fat/ 12.5% (w/w) cholesterol/ no cholate] 8 weeks
Anti-atherogenic 62.8% reduction in en face aortas and 48% reduction in heart sinus valve in PLTPwt/wt LDLr / mice
Valenta et al. (2008)
LDLr/
huPLTPtg/tg or huPLTPtg/wt
HFHC [0.25% cholesterol/ 15% cocoa butter/40% sucrose] 9 weeks
Atherogenic Increased atherosclerosis 2.3-fold in huPLTPtg/wt >LDLr/ 4.5-fold in huPLTPtg/tg >LDLr/
van Haparen et al. (2008)
BM, bone marrow; LDLr, LDL receptor; HFD, high fat diet; HFHC, high fat high cholesterol diet; WTD, western diet; tg, transgenic; wt, wild type
The HDL Handbook
Strain
Chapter j 1
Role of Phospholipid Transfer Protein
9
in macrophages in LDLr/ mice transplanted with bone marrow from mice overexpressing human PLTP. An important issue is how increased PLTP expression stimulates atherosclerosis. Elevation of PLTP activity results in HDL hypercatabolism and enhanced hepatic VLDL secretion. In addition to PLTP-mediated phospholipid transfer, the effects on the transfer of cholesterol and a-tocopherol may be important in this context (Ogier et al., 2007). Moreover, O’Brien et al. (2003) suggested that PLTP may have anti-atherogenic potential by acting as a bridging protein between lipoproteins and biglycan, one of the major extracellular proteoglycans found in human atherosclerotic lesions.
PLTP AND ATHEROSCLEROSIS IN HUMANS PLTP activity is increased in subjects with several risk factors for atherosclerosis. It increases with age (Tahvanainen et al., 1999), body mass index (Dullaart et al., 1994a; Riemens et al., 1998; Huuskonen et al., 2000b; Ja¨nis et al., 2004; Kaser et al., 2004), cigarette smoking (Dullaart et al., 1994b; Cheung et al., 2006), and type 1 and type 2 diabetes mellitus (Colholm et al., 2001; Jonkers et al., 2003; Attia et al., 2007). Studying coronary artery disease (CAD) patients with stable angina, unstable angina and acute myocardial infarction in a crosssectional study, Schlitt et al. (2003) found that increased plasma PLTP activity is a risk factor for CAD. In a prospective study of CAD patients, however, PLTP activity was not predictive of future events (Rao et al., 1997). In a prospective Japanese study, plasma PLTP concentration was negatively associated with CAD (Yatsuya et al., 2004). Finally, a small cross-sectional study, comparing normolipidemic CAD patients with controls matched for plasma lipoproteins, found a non-significant trend for increased PLTP activity in the cases (Ruhling et al., 199). Schgoer et al. (2008) demonstrated that low PLTP activity was significantly decreased in peripheral artery disease (PAD) patients with no history of smoking or diabetes mellitus, and that the magnitude of the decrease increased with increasing PAD stage. PLTP activity was also a determinant of carotid intima media thickness in type 2 diabetes independent of plasma lipids, insulin resistance and C-reactive protein (de Vries et al., 2006). In human atherosclerotic lesions, PLTP was immunohistochemically colocalized in plaques with macrophages and smooth muscle foam cells (Desrumaux et al., 2003; Laffitte et al., 2003; O’Brien et al., 2003). In the atherosclerotic segments, PLTP had accumulated in the extracellular matrix, co-localizing with apoA-I, apoE and biglycan, suggesting that PLTP might promote HDL retention on extracellular matrix proteoglycans and reverse cholesterol transport (O’Brien et al., 2003). It has recently been demonstrated that PLTP-mediated remodeling of apoE-containing HDL generates large as well as small particles, which are potential acceptors of cellular phospholipids and cholesterol (Settasatian et al., 2008). This remodeling may be important for enhancing plasma cholesterol transport, especially under circumstances in
10
The HDL Handbook
which the activity of CETP is inhibited and the levels of apoE-containing HDL in plasma are increased. The PLTP-mediated remodeling of apoE-containing HDL may offset the deleterious effects of increased PLTP activity in type 2 diabetes (Dallinga-Thie et al., 2006; Tan et al., 2006). The relationship of elevated PLTP activity to human atherosclerosis needs to be clarified.
CONCLUSION Numerous studies in humans, animals and in vitro have addressed the importance of RCT and cholesterol efflux in atherogenesis. It is possible that augmentation of RCT and cholesterol efflux could be therapeutically useful. Potential major strategies include accelerating RCT and cholesterol efflux, which may be achieved by increasing HDL and apoA-I levels, or by stimulating PLTP or CETP. However, uncertainties remain about the impact of PLTP on RCT, cholesterol efflux and cardiovascular disease. Elevated activities in subjects with diabetes mellitus and CAD may result from partly elevated rates of VLDL turnover, inhibition of VLDL synthesis and increase of HDL catabolism. The regulation of PLTP and lipolytic enzyme levels, and their impacts on atherosclerosis, are not clear. Since the function of PLTP may influence both pro-atherogenic and anti-atherogenic factors, the net effect of these on plasma lipoproteins and atherosclerosis is not predictable in the current state of knowledge, and further investigations are warranted.
REFERENCES Albers, J. J., Tollefson, J. H., Chen, C. H., & Steinmetz, A. (1984). Isolation and characterization of human plasma lipid transfer proteins. Atherosclerosis, 4, 49e58. Albers, J. J., Tu, A. Y., Wolfbauer, G., Cheung, M. C., & Marcovina, S. M. (1996). Molecular biology of phospholipid transfer protein. Curr Opin Lipidol, 7, 88e93. Albers, J. J., Wolfbauer, G., Cheung, M. C., Day, J. R., Ching, A. F., & Tu, A. Y. (1995). Functional expression of human and mouse phospholipid transfer protein: effect of recombinant and plasma PLTP on HDL subspecies. Biochim Biophys Acta, 1258, 27e34. Attia, N., Nakbi, A., Smaoui, M., et al. (2007). Increased phospholipid transfer protein activity associated with the impaired cellular cholesterol efflux in type 2 diabetic subjects with coronary artery disease. Tohoku J Exp Med, 213, 129e137. Castro, G. R., & Fielding, C. J. (1988). Early incorporation of cell-derived cholesterol into prebeta-migrating high-density lipoprotein. Biochemistry, 27, 25e29. Cheung, M. C., Brown, B. G., Marino Larsen, E. K., Frutkin, A. D., O’Brien, K. D., & Albers, J. J. (2006). Phospholipid transfer protein activity is associated with inflammatory markers in patients with cardiovascular disease. Biochim Biophys Acta, 1762, 131e137. Colholm, H. M., Scheek, L. M., Rubens, M. B., et al. (2001). Lipid transfer protein activities in type I diabetic patients without renal failure and nondiabetic control subjects and their association with coronary artery calcification. Diabetes, 50, 652e659. Curtiss, L. K., Bonnet, D. J., & Rye, K. A. (2000). The conformation of apolipoprotein A-I in highdensity lipoproteins is influenced by core lipid composition and particle size: a surface plasmon resonance study. Biochemistry, 39, 5712e5721.
Chapter j 1
Role of Phospholipid Transfer Protein
11
Curtiss, L. K., Velenta, D. T., Hime, N. J., & Rye, K. A. (2006). What is so special about apolipoprotein AI in reverse cholesterol transport? Arteroscler Thromb Vasc Biol, 26, 12e19. Dallinga-Thie, G. M., van Tol, A., Hattori, H., Rensen, P. C., & Sijbrands, E. J. (2006). Plasma phospholipid transfer protein activity is decreased in type 2 diabetes during treatment with atrovastatin: a role for apolipoprotein E? Diabetes, 55, 1491e1496. Day, J. R., Albers, J. J., Lofton-Day, C. E., et al. (1994). Complete cDNA encoding human phospholipid transfer protein from human endotherial cells. J Biol Chem, 269, 9388e9391. de Vries, R., Dallinga-Thie, G. M., Smit, A. J., Wolffenbuttel, B. H., van Tol, A., & Dullaart, R. P. (2006). Elevated plasma phospholipid transfer protein activity is a determinant of carotid intima-media thickness in type 2 diabetes mellitus. Diabetologia, 49, 398e404. Desrumaux, C., Deckert, V., Athias, A., et al. (1999). Plasma phospholipid transfer protein prevents vascular endothelium dysfunction by delivering alpha-tocopherol to endothelial cells. FASEB J, 13, 883e892. Desrumaux, C. M., Mak, P. A., Boisvert, W. A., et al. (2003). Phospholipid transfer protein is present in human atherosclerotic lesions and is expressed by macrophages and foam cells. J Lipid Res, 44, 1453e1461. Dullaart, R. P., Hoogenberg, K., Dikkeschei, B. D., & van Tol, A. (1994b). Higher plasma lipid transfer protein activities and unfavorable lipoprotein changes in cigarette-smoking men. Arterioscler Thromb, 14, 1581e1585. Dullaart, R. P., Sluiter, W. J., Dikkeschei, L. D., Hoogenberg, K., & van Tol, A. (1994a). Effect of adiposity on plasma lipid transfer protein activities: a possible link between insulin resistance and high density lipoprotein metabolism. Eur J Clin Invest, 24, 188e194. Dusserre, E., Moulin, P., & Vidal, H. (2000). Differences in mRNA expression of the proteins secreted by adipocytes in human subcutaneous and visceral adipose tissues. Biochim Biophys Acta, 1500, 88e96. Eisenberg, S. (1984). High density lipoprotein metabolism. J Lipid Res, 25, 1017e1058. Gordon, D. J., & Rifkind, B. M. (1989). High-density lipoprotein e the clinical implications of recent studies. N Engl J Med, 321, 1311e1316. Gray, P. W., Flaggs, G., Leong, S. R., et al. (1989). Cloning of the cDNA of a human neutrophil bacterial protein. Structure and functional correlations. J Biol Chem, 264, 9505e9509. Huuskonen, J., Ekstrom, M., Tahvanainen, E., et al. (2000b). Quantification of human plasma phospholipid transfer protein (PLTP): relationship between PLTP mass and phospholipid transfer activity. Atherosclerosis, 151, 451e461. Huuskonen, J., Jauhiainen, M., Ehnholm, C., & Olkkonen, V. M. (1998). Biosynthesis and secretion of human plasma phospholipid transfer protein. J Lipid Res, 39, 2021e2030. Huuskonen, J., Olkkonen, V. M., Ehnholm, C., Metso, J., Julkunen, I., & Jauhiainen, M. (2000a). Phospholipid transfer is a prerequisite for PLTP-mediated HDL conversion. Biochemistry, 39, 16092e16098. Huuskonen, J., Olkkonen, V. M., Jauhiainen, M., & Ehnholm, C. (2001). The impact of phospholipid transfer protein (PLTP) on HDL metabolism. Atherosclerosis, 155, 269e281. Huuskonen, J., Olkkonen, V. M., Jauhiainen, M., Metso, J., Somerharju, P., & Ehnholm, C. (1996). Acyl chain and headgroup specificity of human plasma phospholipid transfer protein. Biochim Biophys Acta, 1303, 207e214. Huuskonen, J., Wohlfahrt, G., Jauhiainen, M., Ehnholm, C., Teleman, O., & Olkkonen, V. M. (1999). Structure and phospholipid transfer activity of human PLTP: analysis by molecular modeling and site-directed mutagenesis. J Lipid Res, 40, 1123e1130.
12
The HDL Handbook
Ja¨nis, M. T., Siggins, S., Tahvanainen, E., et al. (2004). Active and low-active forms of serum phospholipid transfer protein in a normal Finnish population sample. J Lipid Res, 45, 2303e2309. Jauhiainen, M., Metso, J., Pahlman, R., Blomqvist, S., van Tol, A., & Ehnholm, C. (1993). Human plasma phospholipid transfer protein causes high density lipoprotein conversion. J Biol Chem, 268, 4032e4036. Jiang, X. C., Francone, O. L., Bruce, R., et al. (1996). Increased prebeta-high density lipoprotein, apolipoprotein AI, and phospholipid in mice expressing the human phospholipid transfer protein and human apolipoprotein AI transgenes. J Clin Invest, 98, 2373e2380. Jiang, X. C., Qin, S., Qiao, C., et al. (2001). Apolipoprotein B secretion and atherosclerosis are decreased in mice with phospholipid-transfer protein deficiency. Nat Med, 7, 847e852. Jiang, X. C., Tall, A. R., Qin, S., et al. (2002). Phospholipid transfer protein deficiency protects circulating lipoproteins from oxidation due to the enhanced accumulation of vitamin E. J Biol Chem, 277, 31850e31856. Jonkers, I. J., Smelt, A. H., Hattori, H., et al. (2003). Decreased PLTP mass but elevated PLTP activity linked to insulin resistance in HTG: effects of bezafibrate therapy. J Lipid Res, 44, 1462e1469. Ka¨rkka¨inen, M., Oka, T., Olkkonen, V. M., et al. (2002). Isolation and partial characterization of the inactive and active forms of human plasma phospholipid transfer protein (PLTP). J Biol Chem, 277, 15413e15418. Kaser, S., Laimer, M., Sandhofer, A., Salzmann, K., Ebenbichler, C. F., & Patsch, J. R. (2004). Effects weight loss on PLTP activity and HDL particle size. Int J Obes Relat Metab Disord, 28, 1280e1282. Laffitte, B. A., Joseph, S. B., Chen, M., et al. (2003). The phospholipid transfer protein gene is a liver X receptor target expressed by macrophages in atherosclerotic lesions. Mol Cell Biol, 23, 2182e2191. Lagrost, L., Desrumaux, C., Masson, D., Deckert, V., & Gambert, P. (1998). Structure and function of the plasma phospholipid transfer protein. Curr Opin Lipidol, 9, 203e209. Lie, J., de Crom, R., van Gent, T., et al. (2004). Elevation of plasma phospholipid transfer protein increases the risk of atherosclerosis despite lower apolipoprotein B containing lipoproteins. J Lipid Res, 45, 805e811. Liu, R., Hojjati, M. R., Devlin, C. M., Hansen, I. H., & Jiang, X. C. (2007). Macrophage phospholipid transfer protein deficiency and ApoE secretion: impact on mouse plasma cholesterol levels and atherosclerosis. Arterioscler Thromb Vasc Biol, 27, 190e196. Lusa, S., Jauhiainen, M., Metso, J., Somerharju, P., & Ehnholm, C. (1996). The mechanism of human plasma phospholipid transfer protein-induced enlargement of high-density lipoprotein particles: evidence for particle fusion. Biochem J, 313, 275e282. Marques-Vidal, P., Jauhiainen, M., Metso, J., & Ehnholm, C. (1997). Transformation of high density lipoprotein 2 particles by hepatic lipase and phospholipid transfer protein. Atherosclerosis, 133, 87e95. Massey, J. B., Hickson, D., She, H. S., et al. (1984). Measurement and prediction of the rates of spontaneous transfer of phospholipids between plasma lipoproteins. Biochim Biophys Acta, 794, 274e280. Moreland, M., Samyn, H., van Gent, T., et al. (2008). Acute elevation of plasma PLTP activity strongly increases pre-existing atherosclerosis. Arterioscler Thromb Vasc Biol, 28, 1277e1282. Murdoch, S. J., Kahn, S. E., Albers, J. J., Brunzell, J. D., & Purnell, J. Q. (2003). PLTP activity decreases with weight loss: changes in PLTP are associated with changes in subcutaneous fat and FFA but not IAF or insulin sensitivity. J Lipid Res, 44, 1705e1712.
Chapter j 1
Role of Phospholipid Transfer Protein
13
Murdoch, S. J., Wolfbauer, G., Kennedy, H., Marcovina, S. M., Carr, M. C., & Albers, J. J. (2002). Differences in reactivity of antibodies to active versus inactive PLTP significantly impacts PLTP measurement. J Lipid Res, 43, 281e289. Navab, M., Ananthramaiah, G. M., Reddy, S. T., et al. (2004). The oxidation hypothesis of atherogenesis: the role of oxidized phospholipids and HDL. J Lipid Res, 45, 993e1007. Nishida, H. I., Klock, D. G., Guo, Z., Jakstys, B. P., & Nishida, T. (1997). Phospholipid transfer protein can transform reconstituted discoidal HDL into vesicular structures. Biochim Biophys Acta, 1349, 222e232. O’Brien, K. D., Vuletic, S., McDonald, T. O., et al. (2003). Cell-associated and extracellular phospholipid transfer protein in human coronary atherosclerosis. Circulation, 108, 270e274. Ogier, N., Klein, A., Deckert, V., et al. (2007). Cholesterol accumulation is increased in macrophages of phospholipid transfer protein-deficient mice: normalization by dietary alphatocopherol supplementation. Arterioscler Thromb Vasc Biol, 27, 2407e2412. Oka, T., Kujiraoka, T., Ito, M., et al. (2000a). Measurement of human plasma phospholipid transfer protein by sandwich ELISA. Clin Chem, 46, 1357e1364. Oka, T., Kujiraoka, T., Ito, M., et al. (2000b). Distribution of phospholipid transfer protein in human plasma: presence of two forms of phospholipid transfer protein, one catalytically active and the other inactive. J Lipid Res, 41, 1651e1657. Oram, J. F., Wolfbauer, G., Tang, C., Davidson, W. S., & Albers, J. J. (2008). An amphipathic helical region of the N-terminal barrel of phospholipid transfer protein is critical for ABCA1dependent cholesterol efflux. J Biol Chem, 283, 11541e11549. Oram, J. F., & Yokoyama, S. (1996). Apolipoprotein-mediated removal of cellular cholestrol and phospholipids. J Lipid Res, 37, 2473e2491. Pulcini, T., Terru, P., Sparrow, J. T., Pownall, H. J., & Ponsin, G. (1995). Plasma factors affecting the in vivo conversion of high-density lipoproteins labeled with a non-transferable marker. Biochim Biophys Acta, 1254, 13e21. Pussinen, P., Jauhiainen, M., & Ehnholm, C. (1997). ApoA-II/apoA-I molar ratio in the HDL particle influences phospholipid transfer protein-mediated HDL interconversion. J Lipid Res, 38, 12e21. Pussinen, P., Jauhiainen, M., Metso, J., Tyynela, J., & Ehnholm, C. (1995). Pig plasma phospholipids transfer protein facilitates HDL interconversion. J Lipid Res, 36, 975e985. Pussinen, P., Malle, E., Metso, J., Sattler, W., Raynes, J. G., & Jauhiainen, M. (2001). Acute-phase HDL in phospholipid transfer protein (PLTP)-mediated HDL conversion. Atherosclerosis, 155, 297e305. Qu, S. J., Fan, H. Z., Kilinc, C., & Pownall, H. J. (1999). Role of cysteine residues in human plasma phospholipid transfer protein. J Protein Chem, 18, 193e198. Rao, R., Albers, J. J., Wolfbauer, G., & Pownall, H. J. (1997). Molecular and macromolecular specificity of human plasma phospholipid transfer protein. Biochemistry, 36, 3645e3653. Riemens, S. C., van Tol, A., Sluiter, W. J., & Dullaart, R. P. (1998). Plasma phospholipid transfer protein activity is related to insulin resistance: impaired acute lowering by insulin in obese Type II diabetic patients. Diabetologia, 41, 929e934. Ruhling, K., Lang, A., Richard, F., et al. (1999). Net mass transfer of plasma cholesteryl esters and lipid transfer proteins in normolipidemic patients with peripheral vascular disease. Metabolism, 48, 1361e1366. Rye, K. A., & Barter, P. J. (1986). Changes in the size and density of human high-density lipoproteins promoted by a plasma-conversion factor. Biochim Biophys Acta, 875, 429e438. Rye, K. A., Hime, N. J., & Barter, P. J. (1995). The influence of cholesteryl ester transfer protein on the composition, size, and structure of spherical, reconstituted high density lipoproteins. J Biol Chem, 270, 189e196.
14
The HDL Handbook
Rye, K. A., Hime, N. J., & Barter, P. J. (1997). Evidence that cholesteryl ester transfer proteinmediated reductions in reconstituted high density lipoprotein size involve particle fusion. J Biol Chem, 272, 3953e3960. Rye, K. A., Jauhiainen, M., Barter, P. J., & Ehnholm, C. (1998). Triglyceride-enrichment of high density lipoproteins enhances their remodeling by phospholipid transfer protein. J Lipid Res, 39, 613e622. Samyn, H., Moreland, M., van Gent, T., et al. (2008). Plasma phospholipid transfer activity is essential for increased atherogenesis in PLTP transgenic mice: a mutation-inactivation study. J Lipid Res, 49, 2504e2512. Schgoer, W., Mueller, T., Jauhiainen, M., et al. (2008). Low phospholipids transfer protein (PLTP) is a risk factor for peripheral atherosclerosis. Atherosclerosis, 196, 219e226. Schlitt, A., Bickel, C., Thumma, P., et al. (2003). High plasma phospholipid transfer protein levels as a risk factor for coronary artery disease. Arterioscler Thromb Vasc Biol, 23, 1857e1862. Schumann, R. R., Leong, S. R., Flaggs, G. W., et al. (1990). Structure and function of lipopolysaccharide binding protein. Science, 249, 1429e1431. Settasatian, N., Barter, P. J., & Rye, K. A. (2008). Remodeling of apolipoprotein E-containing spherical reconstituted high density lipoproteins by phospholipid transfer protein. J Lipid Res, 49, 115e126. Shelly, L., Royer, L., Sand, T., Jensen, H., & Luo, Y. (2008). Phospholipid transfer protein deficiency ameliorates diet-induced hypercholesterolemia and inflammation in mice. J Lipid Res, 49, 773e781. Tahvanainen, E., Jauhiainen, M., Funke, H., Vartiainen, E., Sundvall, J., & Ehnholm, C. (1999). Serum phospholipid transfer protein activity and genetic variation of the PLTP gene. Atherosclerosis, 146, 107e115. Tall, A. R., Abreu, E., & Shuman, J. (1983). Separation of plasma phospholipid transfer protein from cholesterol ester/phospholipid exchange protein. J Biol Chem, 258, 2174e2180. Tall, A. R., & Lalanne, F. (2003). Phospholipid transfer protein and atherosclerosis. Arteroscler Thromb Vasc Biol, 23, 1484e1485. Tan, K. C., Shiu, S. W., Wong, Y., Wong, W. K., & Tam, S. (2006). Plasma apolipoprotein E concentration is an important determinant of phospholipid transfer protein activity in type 2 diabetes mellitus. Diabetes Metab Res Rev, 22, 307e312. Tobias, P. S., Mathison, J. C., & Ulevitch, R. J. (1988). A family of lipopolysaccharide binding proteins involved in responses to gram-negative sepsis. J Biol Chem, 263, 13479e13481. Tollefson, J. H., Ravnik, S., & Albers, J. J. (1988). Isolation and characterization of a phospholipid transfer protein (LTP-II) from human plasma. J Lipid Res, 29, 1593e1602. Tu, A. Y., Nishida, H. I., & Nishida, T. (1993). High density lipoprotein conversion mediated by human plasma phospholipid transfer protein. J Biol Chem, 268, 23098e23105. Valenta, D. T., Bulgrien, J. J., Bonnet, D. J., & Curtiss, L. K. (2008). Macrophage PLTP is atheroprotective in LDLr-deficient mice with systemic PLTP deficiency. J Lipid Res, 49, 24e32. Valenta, D. T., Ogier, N., Bradshaw, G., et al. (2006). Atheroprotective potential of macrophagederived phospholipid transfer protein in low-density lipoprotein receptor-deficient mice is overcome by apolipoprotein AI overexpression. Arterioscler Thromb Vasc Biol, 26, 1572e1578. van Haperen, R., Samyn, H., Moerland, M. et al. (2008). Elevated expression of phospholipid transfer protein in bone marrow derived cells causes Atherosclerosis. PLoS ONE, 3:e2255. van Haperen, R., van Tol, A., van Gent, T., et al. (2002). Increased risk of atherosclerosis by elevated plasma levels of phospholipid transfer protein. J Biol Chem, 277, 48938e48943.
Chapter j 1
Role of Phospholipid Transfer Protein
15
van Haperen, R., van Tol, A., Vermeulen, P., et al. (2000). Human plasma phospholipid transfer protein increases the anti-atherogenic potential of high density lipoproteins in transgenic mice. Arterioscler Thromb Vasc Biol, 20, 1082e1088. Vikstedt, R., Metso, J., Hakala, J., Olkkonen, V. M., Ehnholm, C., & Jauhiainen, M. (2007a). Cholesterol efflux from macrophage foam cells is enhanced phospholipid transfer protein through generation of two types of acceptor particles. Biochemistry, 46, 11979e11986. Vikstedt, R., Ye, D., Metso, J., et al. (2007b). Macrophage phospholipid transfer protein contributes significantly to total plasma phospholipid transfer activity and its deficiency leads to diminished atherosclerotic lesion development. Arterioscler Thromb Vasc Biol, 27, 578e586. von Eckardstein, A., Jauhiainen, M., Huang, Y., et al. (1996). Phospholipid transfer protein mediated conversion of high density lipoproteins generates pre beta 1-HDL. Biochim Biophys Acta, 1301, 255e262. Wehinger, A., Tancevski, I., Schgoer, W., et al. (2007). Phospholipid transfer protein augments apoptosis in THP-1-derived macrophages induced by lipolyzed hypertriglyceridemic plasma. Arterioscler Thromb Vasc Biol, 27, 908e915. Yang, X. P., Yan, D., Qiao, C., et al. (2003). Increased atherosclerotic lesions in apo E mice with plasma phospholipid transfer protein overexpression. Arterioscler Thromb Vasc Biol, 23, 1601e1607. Yatsuya, H., Tamakoshi, K., Hattori, H., et al. (2004). Serum phospholipid transfer protein mass as a possible protective factor for coronary heart diseases. Circ J, 68, 11e16.
Chapter 2
The Role of Cholesteryl Ester Transfer Protein (CETP) in HDL Metabolism Neil J. Hime Centre for Vascular Research, Sydney Medical School (Pathology) and Bosch Institute, The University of Sydney, Medical Foundation Building, Camperdown, NSW, Australia
INTRODUCTION Cholesteryl ester transfer protein (CETP) exchanges neutral lipids (triglycerides and cholesteryl esters) and phospholipids between lipoproteins. Through the in vitro observance of the capacity of lipoprotein-deficient plasma from rabbits (Zilversmit et al., 1975) and humans (Barter and Lally, 1978; Pattnaik et al., 1978) to transfer cholesteryl esters between lipoproteins, the existence of CETP was known long before crystallization enabled a detailed description of the protein (Qiu et al., 2007). In between these discoveries, many studies have used isolated preparations of CETP to shed light on the role of CETP in high density lipoprotein (HDL) metabolism. These studies will be addressed in this chapter. CETP-mediated exchange of neutral lipids results in the net mass transfer of triglycerides from triglyceride-rich very low density lipoproteins (VLDL) to HDL and the net transfer of cholesteryl esters from HDL to VLDL1. This was first observed with lipoprotein-deficient plasma in vitro (Hopkins and Barter, 1980) and subsequently shown to be facilitated by the one protein (Morton and Zilversmit, 1982). It is now known that these CETP-mediated net mass transfers occur in the lipoprotein milieu in blood. The CETP-mediated exchange of neutral lipids changes the molecular structure of lipoproteins. Because HDL is smallest in size of the various lipoprotein classes, and therefore has fewer neutral lipids per particle, it is not surprising that the movement of neutral lipids in and 1 Low density lipoprotein (LDL) can substitute for HDL as the cholesteryl ester donor and triglyceride recipient in net transfers (Albers et al., 1984).
The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
17
18
The HDL Handbook
out of these particles has particularly profound effects. These effects change HDL structure and consequently function and metabolism. Not all animals that transport lipids in the aqueous environment of plasma via lipoproteins have similar cholesteryl ester-triglyceride exchange activity. Those animals that are deficient in CETP activity include: rat, mouse, pig, cow, sheep, guinea pig and dog, while those that have high CETP activity include: rabbit, monkey and human (Barter and Lally, 1978; Barter et al., 1979; Rajaram et al., 1980; Ha and Barter, 1982; Tall, 1986). Of the latter three, rabbit lipoprotein-deficient plasma has the highest cholesteryl ester transfer activity when added to radiolabeled lipoprotein substrates. The comparative lipoprotein profiles and susceptibility to atherosclerosis of various species offer important clues as to the role of CETP in HDL metabolism and cardiovascular disease. These comparisons are discussed in this chapter. Much use has been made of genetic manipulation of CETP expression in animal models to elucidate the role of CETP in HDL metabolism. Also, the very few individuals that are known to be CETP-deficient have been intensively studied. These analyses are addressed in the following pages. If the effect of the action of CETP on HDL had to be described in as few words as possible, it is that it reduces the concentration of HDL cholesterol in the plasma. Given that a low concentration of HDL cholesterol is clearly recognized as an independent risk factor for cardiovascular disease, CETP inhibition has been targeted as a strategy to prevent/reduce disease. The results of attempts to inhibit CETP in humans and the current status of this novel therapy are discussed at the end of this chapter.
MOLECULAR BIOLOGY CETP is a 476 residue hydrophobic glycoprotein that has sequence similarity and identical exoneintron boundaries to other lipid-binding proteins: phospholipid transfer protein, lipopolysaccharide-binding protein and bactericidal/permeability-increasing protein (Bruce et al., 1998). This indicates a common ancestral protein and, although these proteins have different physiological functions, they can each bind lipopolysaccharides and phospholipids, and facilitate lipid transport in the bloodstream (Beamer et al., 1997; Yu et al., 1997). CETP can facilitate the bi-directional exchange of the same neutral lipid or phospholipid between lipoproteins, so called homoexchange, without any change in net lipid transfer. The physiological significance of such an exchange is unclear. What is clearly of physiological relevance is the heteroexchange, net mass transfer of cholesteryl ester and triglyceride between lipoproteins. The exchange of lipids appears to be a passive process with the rate of transfer influenced by the concentrations of respective lipids in lipoproteins (Morton and Steinbrunner, 1990). Therefore the net result of heteroexchange on triglyceride-rich lipoproteins (i.e., VLDL) is a reduction in triglycerides and an increase in cholesteryl esters. Conversely, the effect on cholesteryl ester-rich lipoproteins (i.e., HDL) is
Chapter j 2
The Role of Cholesteryl Ester Transfer Protein (CETP)
19
a reduction in cholesteryl esters and an increase in triglycerides. As will be discussed in the following pages, this has considerable effect on the structure of HDL particles and on the ability of HDL to protect from atherosclerosis. Binding of CETP to VLDL, LDL and HDL is reversible and saturable, and is necessary for lipid transfer (Morton, 1985). Binding and cholesteryl ester transfer rate is influenced by modifications to lipoproteins that change surface charge (Nishida et al., 1993). The mechanism by which CETP facilitates lipid transfer has been postulated and various models proposed. One model proposes that CETP shuttles lipids between lipoproteins by binding to a lipoprotein particle, “picking up” lipid(s), dissociating from the lipoprotein particle and then binding to another lipoprotein particle to off-load the lipid. This model fits well with experimental kinetic data (Barter and Jones, 1980; Epps et al., 1995). The “shuttle” model assumes that, in plasma, CETP exists in two forms: (1) freely circulating, and (2) bound to a single lipoprotein particle. This model also only fits well with the kinetic data if esterified cholesterol in HDL is more “available” for transfer than esterified cholesterol in LDL. This may well be the case since CETP associates more avidly with HDL than LDL (Pattnaik and Zilversmit, 1979). A second model of CETP-mediated lipid transfer involves the formation of a ternary complex consisting of CETP and two lipoprotein particles. This ternary complex is the transition state for lipid exchange. This model also fits with experimental kinetic data (Ihm et al., 1982). Experiments examining CETP-mediated structural changes in HDL also support the concept of the generation of a ternary complex (Rye et al., 1997). ˚ resolution The recent determination of the structure of CETP at 2.2-A has provided more definitive evidence of how CETP mediates lipid exchange (Qiu et al., 2007). The crystal structure of human CETP was solved in complex with four bound lipid molecules. The overall shape of CETP is akin to that of a boomerang with a concave and a convex side. Traversing the length of the protein is a tunnel with two distinct openings that are large enough to permit lipid access. In the crystallized protein, this tunnel was occupied by two cholesteryl esters and two phospholipid molecules. The ability of different combinations of lipids to occupy this tunnel may account for conflicting models of lipid exchange based on kinetic data. Synthesis of proteins with mutated tunnels suggests that lipids are required to traverse the tunnel in order to be transferred between lipoproteins (Qiu et al., 2007). The openings of the tunnel are on the concave side of the protein making this surface the most likely to bind lipoproteins. The concave surface of the crystallized protein complements a 10 nm diameter sphere. This fits well with biochemical studies showing CETP to have high binding affinity for 10 nm diameter discoidal HDL (Bruce et al., 1995). CETP structure does allow for the possibility of conformational change to accommodate larger VLDL (Qiu et al., 2007). The structure of the concave surface indicates that CETP can only bind one lipoprotein particle at a time. This only supports the “shuttling” mechanism of lipid transfer. It is suggested that each time CETP picks up neutral lipid from lipoproteins in the tunnel, the
20
The HDL Handbook
tunnel openings are filled with phospholipid molecules to permit the protein to return to the polar environment of the plasma (Qiu et al., 2007).
EFFECT OF CETP ACTION ON HDL STRUCTURE In human plasma, most CETP is found associated with HDL (Pattnaik and Zilversmit, 1979). From analysis of binding capacities of various lipoproteins for CETP and respective plasma concentrations of lipoproteins, it has been estimated that 74% of plasma CETP is bound to HDL (Nishida et al., 1993). This increased association between CETP and HDL compared to other lipoprotein classes may reflect complementing structures as discussed above and/or higher electrostatic attraction. In the simplistic environment of a test tube incubation (up to 24 hours), CETP-mediated transfer of cholesteryl esters from HDL to VLDL and LDL results in a reduction in HDL particle size and a dissociation of apolipoprotein A-I in a lipid-poor form (Liang et al., 1994). With the addition of other plasma factors, this CETP-generated lipid-poor apolipoprotein A-I results in the formation of new HDL particles (Clay et al., 1992). Interestingly, in an 8-hour incubation of HDL with CETP and VLDL, hepatic lipase is required to achieve a reduction in HDL particle size (Clay et al., 1992). This may relate to HDL enriched in triglyceride being the preferred substrate for hepatic lipase. In hypertriglyceridemic mice, the presence of a CETP transgene also decreases HDL particle size (Hayek et al., 1993). Plasma HDL are a heterogeneous group of particles that vary in chemical and physical modalities. In order to get a more detailed description of the changes in HDL structure mediated by CETP, use has been made of homogeneous, reconstituted HDL (rHDL) assembled in the laboratory. Although rHDL are not the same as any given plasma HDL particle, there is sufficient experimental data to suggest that the interactions of rHDL with plasma proteins is equivalent to that of plasma HDL. When rHDL are incubated with CETP and Intralipid (a source of triglycerides), the loss of cholesteryl esters from rHDL is greater than the gain of triglycerides (Rye et al., 1995). This results in a loss of core lipids from the rHDL, a decrease in particle size, and a redistribution of apolipoprotein A-I molecules from three per particle to two. As there is no loss in rHDL constituents in this process, it results in a 50% increase in the number of rHDL particles (Rye et al., 1997).
EFFECT OF CETP ACTIVITY ON PLASMA HDL LEVELS AND ATHEROSCLEROTIC DISEASE CETP activity reduces the plasma concentration of HDL cholesterol. In mice, an animal normally deficient in CETP and with a high concentration of plasma HDL relative to humans, the expression of a human CETP transgene results in a reduction in HDL cholesterol (Agellon et al., 1991). In humans, genetic
Chapter j 2
The Role of Cholesteryl Ester Transfer Protein (CETP)
21
deficiency in CETP results in marked elevation in HDL cholesterol among homozygotes and moderate elevation in heterozygotes (Inazu et al., 1990, 1994; Cefalu` et al., 2009). Mutations in the CETP gene leading to CETP deficiency are particularly prevalent in Japanese people and CETP deficiency is the most common cause of hyperalphalipoproteinemia (elevated HDL) in the Japanese (Nagano et al., 2004). The strong inverse association between HDL cholesterol concentration and atherosclerosis risk (Gordon et al., 1989) has fuelled the study of CETP as a modulator of cardiovascular disease. The expression of CETP in wild type, LDL receptor knockout and apolipoprotein E knockout, atherosclerosis prone, mice decreases plasma HDL cholesterol and results in an increase in atherosclerosis (Marotti et al., 1993; Plump et al., 1999). However, expression of CETP in mice deficient in the scavenger receptor BI (SR-BI) and mice expressing human lecithin:cholesterol acyltransferase protects against atherosclerosis (Fo¨ger et al., 1999; Harder et al., 2007). Polymorphisms in the CETP gene that result in reduced CETP expression are associated with higher plasma HDL cholesterol concentrations, a lower prevalence of cardiovascular disease and increased longevity in humans (Ordovas et al., 2000; Barzilai et al., 2003; Koropatnick et al., 2008). A prospective case-control study showed that elevated CETP levels occur with an increased risk of future coronary artery disease but only in people with high plasma triglyceride levels (Boekholdt et al., 2004). In another case-control study there was actually a trend towards high CETP levels reducing cardiovascular disease risk when triglyceride levels were low (Borggreve et al., 2007). The effect of triglyceride concentration on the association between CETP and atherosclerosis risk is important for triglyceride levels will influence the level of CETP activity for a given plasma CETP concentration. This highlights the need to make distinctions between CETP activity and CETP mass when assessing trial data. Intima-media thickness, a marker for subclinical atherosclerosis, is positively associated with plasma cholesteryl ester transfer rates in diabetic and non-diabetic subjects, however, no association is observed for CETP mass (de Vries et al., 2005). A comparison of species that have or lack CETP activity offers some clues as to the relationship between CETP, HDL and atherosclerosis. Relative to humans, monkeys, and rabbits, mice (strain C57BL6) have only a trace of cholesteryl ester transfer activity (Ha and Barter, 1982; Marotti et al., 1993). Mice have approximately two- and ten-fold greater concentration of plasma HDL cholesterol than humans and rabbits, respectively. Mice are highly resistant to diet-induced atherosclerosis compared to humans and rabbits. Thus, at least with these species, there is a consistent negative association between CETP activity and HDL cholesterol concentration, and a positive correlation between CETP activity and atherosclerosis susceptibility. So is it as simple as CETP activity lowers HDL which increases atherosclerosis risk? Unfortunately, this is too simplistic and there are examples that destroy this nice hypothesis. For example, CETP activity is low in rat, sheep and pig but HDL
22
The HDL Handbook
cholesterol is not particularly high in these species (Ha and Barter, 1982). However, as cautioned above, it needs to be considered that the plasma CETP activity in these species was measured ex vivo. Furthermore, hyperalphalipoproteinemia associated with CETP deficiency is, in certain circumstances, associated with atherosclerosis (Hirano et al., 1995). What is clear is that CETP activity can, although not always, lower plasma HDL concentrations. Also, CETP activity can change HDL in ways that affect the functionality of these particles. This effect of CETP is poorly understood but is likely to affect the anti-atherogenic properties of HDL.
EFFECT OF CETP ACTIVITY ON HDL FUNCTION AND CONSEQUENCES FOR ATHEROSCLEROTIC DISEASE The primary mechanisms by which HDL are understood to protect against atherosclerotic disease are defined as anti-inflammatory (Barter et al., 2004) and reverse cholesterol transport (Glomset, 1968; Rader et al., 2009). The antiinflammatory properties of HDL include the ability of HDL to inhibit endothelial cell adhesion molecules and chemokines. Also, HDL-associated proteins, such as paraoxonase-1, can act as antioxidants. These properties undoubtedly are influenced by the form in which HDL reside, but equally important may be the residence time of individual HDL particles in the plasma. Put simply, the higher the concentration of HDL the greater the anti-inflammatory potential. The anti-inflammatory properties of HDL are not well understood and, at present, it is difficult to speculate how CETP may affect these properties. Although CETP activity decreases HDL cholesterol, it also has the potential to increases the number of HDL particles (Clay et al., 1992; Rye et al., 1997). In increasing the number of HDL particles, rather than increasing the concentration of any one HDL constituent, this may increase the sum anti-inflammatory potential. Reverse cholesterol transport, on the other hand, may depend less on the absolute concentration of HDL at any given time and more on the forms of HDL present. Reverse cholesterol transport refers to the ability of HDL to transport excess cholesterol from peripheral tissues to the liver for excretion. It has been demonstrated that the rate of reverse cholesterol transport from cholesterol-loaded macrophages predicts atherosclerosis severity better than the concentration of HDL cholesterol (Rader et al., 2009). The rate-limiting factor in cholesterol efflux from cholesterol-laden cells is the availability of extracellular acceptors of cholesterol. The major acceptor of cellular cholesterol is lipid-poor apolipoprotein A-I (the major apolipoprotein of HDL). As mentioned above, the action of CETP on HDL can result in the dissociation of apolipoprotein A-I. It has been hypothesized that this dissociated apolipoprotein A-I may then be available for acceptance of cellular cholesterol (Curtiss et al., 2006). As an anti-atherogenic mechanism of CETP, this would be particularly relevant when the CETP is expressed by macrophages in
Chapter j 2
The Role of Cholesteryl Ester Transfer Protein (CETP)
23
FIGURE 2.1 Hypothesised mechanism by which the actions of CETP (and other proteins) on HDL in the interstitium promote removal of cholesterol from atherosclerotic lesions. (from Curtiss, L.K. et al Arterioscler Thromb Vasc Biol. 2006;26:12-19.) This illustration depicts cholesterol (yellow)-loaded macrophages (green) in an atherosclerotic lesion in the arterial intima. The circular inset shows the schema whereby the proteins: CETP, phospholipid transfer protein (PLTP) and hepatic lipase (HL) expressed by macrophages, act upon HDL to generate lipid-poor apolipoprotein A-I (preb-HDL). This lipid-poor apolipoprotein A-I accepts cholesterol from macrophages resulting in the generation of new HDL particles that move from the interstitium to the plasma. The net effect is a reduction in cholesterol in atherosclerotic lesions. From Curtiss, L.K., Valenta, D.T., Hime, N.J. and Rye, K-A. What is so special about apolipoprotein AI in reverse cholesterol transport? Arterioscler Thromb Vasc Biol 2006;26:12-19. Reprinted with permission. The original figure was drawn by D.T. Valenta.
24
The HDL Handbook
atherosclerotic lesions. However, transfer of bone marrow from CETP transgenic mice into atherosclerosis-prone LDL receptor-deficient mice actually increases atherosclerotic lesion size (Van Eck et al., 2007). This is despite bone marrow-derived CETP increasing lipid-poor HDL three-fold (as measured by immunoelectrophoresis). Bone marrow-derived CETP induces a pro-atherogenic lipoprotein profile with an increase in VLDL/LDL cholesterol and a decrease in HDL cholesterol. This change in lipoprotein profile may well offset any advantage to be gained from the generation of lipid-poor HDL. Cellular cholesterol efflux to either HDL or apolipoprotein A-I does not differ between CETP expressing and non-CETP expressing macrophages (Van Eck et al., 2007). However, these studies were conducted ex vivo using isolated human HDL and apolipoprotein A-I. It is feasible that, in certain environments, CETP and other plasma factors could generate sufficient acceptors of cellular cholesterol to decrease atherosclerosis. CETP-mediated neutral lipid transfer between VLDL and HDL generates triglyceride-rich HDL, the preferred substrate of hepatic lipase. Hydrolysis of triglyceride-enriched HDL lipids by hepatic lipase results in the dissociation of apolipoprotein A-I from HDL (Clay et al., 1991). The net result is that the combined actions of CETP and hepatic lipase on HDL promote the shedding of apolipoprotein A-I that becomes available for the formation of new HDL particles by associating with lipid (Clay et al., 1992). This lipid may come from cholesterol-loaded macrophages in the arterial intima and be transported to apolipoprotein A-I via the ATP-binding cassette transporter A1 (ABC-A1) e the first step in the reverse cholesterol transport pathway. All the necessary components are available in the intima for this to happen (Figure 2.1). HDL are able to enter the intimal space. CETP is present in atherosclerotic lesions (Zhang et al., 2001) and hepatic lipase is expressed by macrophages (Gonza´lezNavarro et al., 2002). Whether these processes decrease atherosclerosis is not known, however, it is interesting that bone marrow-derived hepatic lipase reduces atherosclerosis but only when CETP is present (Hime et al., 2008). HDL from CETP-deficient subjects have been shown to enhance cholesterol efflux from macrophages compared to control HDL (Matsuura et al., 2006). This cholesterol efflux is dependent on ABC-G1, a cell membrane transporter that transports cellular lipid to mature HDL. Thus, while CETP may be a promoter of cellular cholesterol efflux to lipid-poor apolipoprotein A-I, it may lessen the amount of cellular cholesterol efflux to mature HDL. If this were the case, the effect on reverse cholesterol transport could depend on the relative contributions of these two processes to the overall transport of cholesterol from peripheral tissues to the liver. Cholesterol movement from macrophages to HDL may not be the only stage of the reverse cholesterol transport under the influence of CETP. CETP could have a role in the transfer of cholesterol from HDL to the liver for excretion. Adenovirus expression of CETP in mouse hepatocytes enhances uptake of cholesteryl esters from HDL independent of lipoprotein receptors
Chapter j 2
The Role of Cholesteryl Ester Transfer Protein (CETP)
25
(Gauthier et al., 2005). The integral membrane protein SR-BI facilitates the uptake of cholesteryl esters from HDL in the liver. Cholesteryl esters from HDL isolated from mice expressing CETP are taken up more readily via SR-BI than cholesteryl esters from HDL from control mice (Collet et al., 1999). The remodeling of HDL involving CETP-mediated neutral lipid exchange and hepatic lipase lipid hydrolysis also enhances uptake of HDL cholesteryl esters by SR-BI (Collet et al., 1999). The overall effect of CETP on the rate of reverse cholesterol transport is complex. While CETP would appear to enhance some reverse cholesterol transport processes, the reduction in HDL cholesterol through CETP-mediated neutral lipid exchange with other lipoproteins may be detrimental to the efficient movement of cholesterol from the vasculature to the liver. Beyond the scope of this chapter but important nonetheless, CETP activity influences the clearance rate of cholesterol via apolipoprotein B-containing lipoproteins and the LDL receptor. The effect of CETP on models of impaired and intact apolipoprotein B lipoprotein clearance is well summarized in a review by Rader et al. (2009). As has been discussed, animal models have provided valuable information regarding the association between CETP, HDL and atherosclerosis. Human CETP expression in atherosclerosis prone apolipoprotein E and LDL receptor deficient mice increases atherosclerosis (Plump et al., 1999). These two mouse models of atherosclerosis have high concentrations of atherogenic LDL/VLDL particles and low levels of HDL. The presence of CETP exacerbates this atherogenic lipid profile, and any change in the functionality of HDL is likely to be of little influence with regard to atherosclerosis in a background of such an atherogenic lipid profile. The C57BL6 strain of mice fed an atherogenic diet also have a lipid profile that is more atherogenic relative to chow fed mice. CETP expression in atherogenic diet fed C57BL6 mice also increases atherosclerosis (Marotti et al., 1993). In hypertriglyceridemic human apolipoprotein CIII transgenic mice, expression of CETP reduces diet-induced atherosclerosis (Hayek et al., 1995). It is possible that, in situations where CETP activity promotes generation of lipid-poor apolipoprotein A-I, atherosclerosis is reduced, whereas when this is not the case and CETP activity reduces HDL levels, then atherosclerosis is increased (Table 2.1). As discussed above, the combined activity of CETP and hepatic lipase on HDL can generate lipid-poor apolipoprotein A-I. In a case-controlled study there is evidence that a reduction in CETP and hepatic lipase activity is associated with atherosclerosis, even when accompanied by high levels of HDL (Hirano et al., 1995). Pharmacological CETP-inhibition in atherogenic diet-fed rabbits significantly reduces atherosclerosis (Morehouse et al., 2007). This is a model in which CETP inhibition is particularly effective at raising HDL. Furthermore, since rabbits do not have hepatic lipase activity, CETP-mediated triglyceride enrichment of HDL might not promote generation of lipid-poor apolipoprotein A-I. Thus, this is a scenario where CETP inhibition may be particularly beneficial.
26
The HDL Handbook
TABLE 2.1 Effects of CETP on HDL metabolism and the potential for these effects to influence atherosclerosis risk
Effect of CETP on HDL metabolism
Potential change in atherosclerosis risk due to CETP-mediated effect*
Decrease in HDL cholesteryl esters (Agellon et al., 1991; Inazu et al., 1990, 1994; Ordovas et al., 2000; Barzilai et al., 2003; Nagano et al., 2004; Koropatnick et al., 2008; Cefalu` et al., 2009
Increase
Reduction in HDL particle size (Clay et al., 1992; Hayek et al., 1993; Liang et al., 1994; Rye et al., 1995)
Increase or decrease
Dissociation of apolipoprotein A-I (Clay et al., 1992; Liang et al., 1994)
Decrease
Formation of new HDL particles (Clay et al., 1992; Rye et al., 1997)
Decrease
* The specified change in atherosclerosis risk is that speculated by the author and does not represent the opinion of the authors of the references.
POLYMORPHISMS OF THE CETP GENE AND INTERACTIONS BETWEEN ENVIRONMENTAL INFLUENCES AND HDL A variety of polymorphisms in the human CETP gene have been identified that result in a decrease in CETP activity and an increase in HDL cholesterol (Thompson et al., 2008). Associations between these polymorphisms and a reduction in cardiovascular disease risk are generally weak. This may be because the increases in HDL cholesterol are modest (3e5%) or it may be that not all effects of CETP gene polymorphisms are beneficial with regards to cardiovascular disease. Another confounding factor is that most studies examining CETP gene polymorphisms have not involved the thousands of participants necessary to fully appreciate relatively modest changes. A polymorphism in intron 1 of the CETP gene (Drayna and Lawn, 1987), denoted TaqIB because of the name of the restriction enzyme used to detect the polymorphism, has been most intensively studied. This polymorphism was one of the first genetic variations associated with HDL levels (Kondo et al., 1989). One reason why this particular polymorphism has been studied more than others is that the less common variant of the polymorphism (B2), that results in lower CETP levels, occurs at a high frequency of approximately 40% in populations so far studied. A meta-analysis has clearly demonstrated that the B2 variant is associated with a small but highly statistically significant increase
Chapter j 2
The Role of Cholesteryl Ester Transfer Protein (CETP)
27
in HDL cholesterol (Boekholdt et al., 2005). After adjustment for study, sex, age, smoking, body mass index, diabetes, systolic blood pressure, LDL cholesterol and use of alcohol, the B2 variant was significantly associated with a reduction in coronary artery disease. The effect on disease may be predominantly associated with HDL levels as statistical significance is lost when adjustment is made for HDL cholesterol. This meta-analysis was confined to white subjects only. A variety of environmental factors, including alcohol consumption, dietary fat intake, smoking and exercise, are known to influence HDL levels. Several studies have examined the interaction between the CETP gene TaqIB polymorphism, environmental factors and HDL. The association between the TaqIB polymorphism and HDL (lower HDL with the B1 genotype and higher HDL with the B2 genotype) has been shown to be more evident in diabetic patients with a high intake of fat in their diet than those with a low fat intake (Li et al., 2007). Alcohol consumption is associated with an increase in HDL cholesterol and this association is stronger for individuals with the B2B2 genotype (Tsujita et al., 2007; Jensen et al., 2008). Smoking is associated with decreased HDL levels and increased CETP activity and multiple regression analysis suggests that the decreased HDL with smoking is explained by the increase in CETP activity (Dullaart et al., 1994). Despite evidence that the effect of the TaqIB polymorphism on HDL levels is limited to non-smokers and ex-smokers (Kondo et al., 1989; Freeman et al., 1994), this polymorphism does influence the increased risk of myocardial infarction associated with smoking (Goldenberg et al., 2007). Smoking is associated with a reduction in the age of onset of first myocardial infarction in B1B1 and B1B2 carriers but not in B2B2 carriers. Associations between CETP gene polymorphisms, HDL levels and exercise are not clear. Polymorphisms aside, long-term exercise training is known to increase HDL and one study showed that endurance exercise training decreased plasma CETP mass (Seip et al., 1993).
INHIBITORS OF CETP ACTIVITY AS THERAPY There is not a single clinically used drug that effectively raises HDL cholesterol. It has been thought that such a drug could be used either where cholesterol lowering drugs are ineffective or used in combination with LDL cholesterol lowering drugs to improve efficacy (Hausenloy and Yellon, 2008). There has been great hope that inhibitors of CETP could fill this void. While CETP inhibition (torcetrapib) in combination with statin therapy (atorvastatin) effectively raises HDL cholesterol, it does not efficaciously reduce atherosclerosis in either hypercholesterolemic patients or patients with established coronary disease (Kastelein et al., 2007; Nissen et al., 2007). A post hoc analysis of the latter of these two studies found that there was a relationship between torcetrapib-mediated change in HDL cholesterol and atherosclerosis. Patients with regression of atheroma volume had greater increases in HDL
28
The HDL Handbook
cholesterol (Nicholls et al., 2008). This supports HDL raising as a therapeutic target, but leaves undecided the effectiveness of CETP inhibitors. The phase III clinical trial ILLUMINATE (Investigation of Lipid Level Management to Understand its Impact on Atherosclerotic Events) was prematurely terminated because of an increase in mortality in the torceptrapib/atorvastatin treatment arm compared with atorvastatin treatment alone (Barter et al., 2007). The increased mortality may relate to associated increases in blood pressure with torceptrapib treatment. It has recently been shown that torceptrapib induces aldosterone and cortisol production independent of CETP inhibition (Hu et al., 2009). Another CETP inhibitor (anacetrapib) raises HDL cholesterol and shows no signs of blood pressure raising (Bloomfield et al., 2009). Thus, there is justification for further investigations of CETP inhibitors as therapy. It is, however, unclear what effect CETP inhibitors have on the many functions of HDL.
CONCLUSION A combination of data obtained from transgenic research animals, different animal species, CETP-deficient humans, epidemiology, inhibitors of CETP and in vitro studies provides ample evidence that CETP activity has fundamental effects on HDL metabolism. The most consistent effect is that CETP activity reduces plasma HDL cholesteryl ester concentration via neutral lipid exchange with other lipoproteins. This is primarily viewed as an atherosclerosispromoting effect due to the independent inverse association between HDL cholesterol and atherosclerosis risk. However, the link between CETP, HDL metabolism and atherosclerosis is not independent of other modulators of lipid metabolism and atherogenesis. It remains to be seen whether CETP inhibitors are effective anti-atherosclerotic agents. Regardless of the outcome of CETP inhibitors as therapy, these drugs will provide valuable information on the effect of CETP on individual HDL functions; this is an area that is poorly understood.
REFERENCES Agellon, L. B., Walsh, A., Hayek, T., et al. (1991). Reduced high density lipoprotein cholesterol in human cholesteryl ester transfer protein transgenic mice. J Biol Chem, 266, 10796e10801. Albers, J. J., Tollefson, J. H., Chen, C. H., & Steinmetz, A. (1984). Isolation and characterization of human lipid transfer proteins. Arteriosclerosis, 4(1), 49e58. Barter, P. J., Caufield, M., Eriksson, M., et al. for the ILLUSTRATE Investigators. (2007). Effects of torcetrapib in patients at high risk for coronary events. N Engl J Med, 357, 2109e2122. Barter, P. J., Gooden, J. M., & Rajaram, O. V. (1979). Species differences in the activity of a serum triglyceride transferring factor. Atherosclerosis, 33, 165e169.
Chapter j 2
The Role of Cholesteryl Ester Transfer Protein (CETP)
29
Barter, P. J., & Jones, M. E. (1980). Kinetic studies of the transfer of esterified cholesterol between human plasma low and high density lipoproteins. J Lipid Res, 21, 238e249. Barter, P. J., & Lally, J. I. (1978). The activity of an esterified cholesterol transferring factor in human and rat serum. Biochim Biophys Acta, 531, 233e236. Barter, P. J., Nicholls, S., Rye, K.-A., Anantharamaiah, G. M., Navab, M., & Fogelman, A. M. (2004). Antiinflammatory properties of HDL. Circ Res, 95, 764e772. Barzilai, N., Atzmon, G., Schechter, C., et al. (2003). Unique lipoprotein phenotype and genotype associated with exceptional longevity. J Am Med Assoc, 290, 2030e2040. Beamer, L. J., Carroll, S. F., & Eisenberg, D. (1997). Crystal structure of human BPI and two bound phospholipids at 2.4 angstrom resolution. Science, 276, 1861e1864. Bloomfield, D., Carlson, G. L., Sapre, A., et al. (2009). Efficacy and safety of the cholesteryl ester transfer protein inhibitor anacetrapib as monotherapy and coadministered with atorvastatin in dyslipidemic patients. Am Heart J, 157, 352e360. Boekholdt, S. M., Kuivenhoven, J.-A., Wareham, N. J., et al. (2004). Plasma levels of cholesteryl ester transfer protein and the risk of future coronary artery disease in apparently healthy men and women: the prospective EPIC (European Prospective Investigation into Cancer and Nutrition) e Norfolk population study. Circulation, 110, 1418e1423. Boekholdt, S. M., Sacks, F. M., Jukema, J. W., et al. (2005). Cholesteryl ester transfer protein TaqIB variant, high-density lipoprotein cholesterol levels, cardiovascular risk, and efficacy of pravastatin treatment: individual patient meta-analysis of 13 677 subjects. Circulation, 111, 278e287. Borggreve, S. E., Hillege, H. L., Dallinga-Thie, G. M., et al. on behalf of the PREVEND study group. (2007). High plasma cholesteryl ester transfer protein levels may favour reduced incidence of cardiovascular events in men with low triglycerides. Eur Heart J, 28, 1012e1018. Bruce, C., Beamer, L. J., & Tall, A. R. (1998). The implications of the structure of the bactericidal/ permeability-increasing protein on the lipid-transfer function of the cholesteryl ester transfer protein. Curr Opin Struct Biol, 8, 426e434. Bruce, C., Davidson, W. S., Kussie, P., et al. (1995). Molecular determinants of plasma cholesteryl ester transfer protein binding to high density lipoproteins. J Biol Chem, 270, 11532e11542. Cefalu`, A. B., Noto, D., Magnolo, L., et al. (2009). Novel mutations of CETP gene in Italian subjects with hyperalphalipoproteinemia. Atherosclerosis, 204, 202e207. Clay, M. A., Newnham, H. H., & Barter, P. J. (1991). Hepatic lipase promotes a loss of apolipoprotein A-I from triglyceride-enriched human high density lipoproteins during incubation in vitro. Arterioscler Thromb Vasc Biol, 11, 415e422. Clay, M. A., Newnham, H. H., Forte, T. M., & Barter, P. J. (1992). Cholesteryl ester transfer protein and hepatic lipase activity promote shedding of apo A-I from HDL and subsequent formation of discoidal HDL. Biochim Biophys Acta, 1124, 52e58. Collet, X., Tall, A. R., Serajuddin, H., et al. (1999). Remodeling of HDL by CETP in vivo and by CETP and hepatic lipase in vitro results in enhanced uptake of HDL CE by cells expressing scavenger receptor B-I. J Lipid Res, 40, 1185e1193. Curtiss, L. K., Valenta, D. T., Hime, N. J., & Rye, K.-A. (2006). What is so special about apolipoprotein AI in reverse cholesterol transport? Arterioscler Thromb Vasc Biol, 26, 12e19. de Vries, R., Perton, F. G., Dallinga-Thie, G. M., et al. (2005). Plasma cholesteryl ester transfer is a determinant of intima-media thickness in type 2 diabetic and nondiabetic subjects: role of CETP and triglycerides. Diabetes, 54, 3554e3559. Drayna, D., & Lawn, R. (1987). Mutiple FRLPs at the human cholesteryl ester transfer protein (CETP) locus. Nucleic Acids Res, 15, 4698.
30
The HDL Handbook
Dullaart, R. P., Hoogenberg, K., Dikkeschei, B. D., & van Tol, A. (1994). Higher plasma lipid transfer protein activities and unfavorable lipoprotein changes in cigarette-smoking men. Arterioscler Thromb Vasc Biol, 14, 1581e1585. Epps, D. E., Greenlee, K. A., Harris, J. S., et al. (1995). Kinetics and inhibition of lipid exchange catalyzed by plasma cholesteryl ester transfer protein (lipid transfer protein). Biochemistry, 34, 12560e12569. Fo¨ger, B., Chase, M., Amar, M. J., et al. (1999). Cholesteryl ester transfer protein corrects dysfunctional high density lipoproteins and reduces aortic atherosclerosis in lecithin cholesterol acyltransferase transgenic mice. J Biol Chem, 274, 36912e36920. Freeman, D. J., Griffin, B. A., Holmes, A. P., et al. (1994). Regulation of plasma HDL cholesterol and subfraction distribution by genetic and environmental factors. Associations between the TaqI B RFLP in the CETP gene and smoking and obesity. Arterioscler Thromb Vasc Biol, 14, 336e344. Gauthier, A., Lau, P., Zha, X., Milne, R., & McPherson, R. (2005). Cholesteryl ester transfer protein directly mediates selective uptake of high density lipoprotein cholesteryl esters by the liver. Arterioscler Thromb Vasc Biol, 25, 2177e2184. Glomset, J. A. (1968). The plasma lecithin:cholesterol acyltransferase reaction. J Lipid Res, 9, 155e167. Goldenberg, I., Moss, A. J., Block, R., et al. (2007). Polymorphism in the cholesteryl ester transfer protein gene and the risk of early onset myocardial infarction among cigarette smokers. Ann Noninvasive Electrocardiol, 12, 364e374. Gonza´lez-Navarro, H., Nong, Z., Freeman, L., Bensadoun, A., Peterson, K., & SantamarinaFojo, S. (2002). Identification of mouse and human macrophages as a site of synthesis of hepatic lipase. J Lipid Res, 43, 671e675. Gordon, D. J., Probstfield, J. L., Garrison, R. J., et al. (1989). High-density lipoprotein cholesterol and cardiovascular disease: four prospective American studies. Circulation, 79, 8e15. Ha, Y. C., & Barter, P. J. (1982). Differences in plasma cholesteryl ester transfer activity in sixteen vertebrate species. Comp Biochem Physiol, 71B, 265e269. Harder, C., Lau, P., Meng, A., Whitman, S. C., & McPherson, R. (2007). Cholesteryl ester transfer protein (CETP) expression protects against diet induced atherosclerosis in SR-BI deficient mice. Arterioscler Thromb Vasc Biol, 27, 858e864. Hausenloy, D. J., & Yellon, D. M. (2008). Targeting residual cardiovascular risk: raising highdensity lipoprotein cholesterol levels. Heart, 94, 706e714. Hayek, T., Azrolan, N., Verdery, R. B., et al. (1993). Hypertriglyceridemia and cholesteryl ester transfer protein interact to dramatically alter high density lipoprotein levels, particle sizes, and metabolism: studies in transgenic mice. J Clin Invest, 92, 1143e1152. Hayek, T., Masucci-Magoulas, L., Jiang, X., et al. (1995). Decreased early atherosclerotic lesions in hypertriglyceridemic mice expressing cholesteryl ester transfer protein transgene. J Clin Invest, 96, 2071e2074. Hime, N. J., Black, A. S., Bulgrien, J. J., & Curtiss, L. K. (2008). Leukocyte-derived hepatic lipase increases HDL and decreases en face aortic atherosclerosis in LDLr/ mice expressing CETP. J Lipid Res, 49, 2113e2123. Hirano, K.-i., Yamashita, S., Kuga, Y., et al. (1995). Atherosclerotic disease in marked hyperalphalipoproteinemia: combined reduction of cholesteryl ester transfer protein and hepatic triglyceride lipase. Arterioscler Thromb Vasc Biol, 15, 1849e1856. Hopkins, G. J., & Barter, P. J. (1980). Transfers of esterified cholesterol and triglyceride between high density and very low density lipoproteins: in vitro studies of rabbits and humans. Metabolism, 29, 546e550.
Chapter j 2
The Role of Cholesteryl Ester Transfer Protein (CETP)
31
Hu, X., Dietz, J. D., Xia, C., et al. (2009). Torcetrapib induces aldosterone and cortisol production by an intracellular calcium-mediated mechanism independently of cholesteryl ester transfer protein (CETP) inhibition. Endocrinology, 150, 2211e2219. Ihm, J., Quinn, D. M., Busch, S. J., Chataing, B., & Harmony, J. A. K. (1982). Kinetics of plasma protein-catalyzed exchange of phosphatidylcholine and cholesteryl ester between plasma lipoproteins. J Lipid Res, 23, 1328e1341. Inazu, A., Brown, M. L., Hesler, C. B., et al. (1990). Increased high-density lipoprotein levels caused by a common cholesteryl-ester transfer protein gene mutation. N Engl J Med, 323, 1234e1238. Inazu, A., Jiang, X.-C., Haraki, T., et al. (1994). Genetic cholesteryl ester transfer protein deficiency caused by two prevalent mutations as a major determinant of increased levels of high density lipoprotein cholesterol. J Clin Invest, 94, 1872e1882. Jensen, M. K., Mukamal, K. J., Overvad, K., & Rimm, E. B. (2008). Alcohol consumption, TaqIB polymorphism of cholesteryl ester transfer protein, high-density lipoprotein cholesterol, and risk of coronary heart disease in men and women. Eur Heart J, 29, 104e112. Kastelein, J. J. P., van Leuven, S. I., Burgess, L., et al. for the RADIANCE 1 Investigators. (2007). Effect of torcetrapib on carotid atherosclerosis in familial hypercholesterolemia. N Engl J Med, 356, 1e11. Kondo, I., Berg, K., Drayna, D., & Lawn, R. (1989). DNA polymorphism at the locus for human cholesteryl ester transfer protein (CETP) is associated with high density lipoprotein cholesterol and apolipoprotein levels. Clin Genet, 35, 49e56. Koropatnick, T. A., Kimbell, J., Chen, R., et al. (2008). A prospective study of high-density lipoprotein cholesterol, cholesteryl ester transfer protein gene variants, and healthy aging in very old Japanese-American men. J Gerontol, 63A, 1235e1240. Li, T. Y., Zhang, C., Asselbergs, F. W., et al. (2007). Interaction between dietary fat intake and the cholesterol ester transfer protein TaqIB polymorphism in relation to HDL-cholesterol concentrations among US diabetic men. Am J Clin Nutr, 86, 1524e1529. Liang, H.-Q., Rye, K.-A., & Barter, P. J. (1994). Dissociation of lipid-free apolipoprotein A-I from high density lipoproteins. J Lipid Res, 35, 1187e1199. Marotti, K. R., Castle, C. K., Boyle, T. P., Lin, A. H., Murray, R. W., & Melchlor, G. W. (1993). Severe atherosclerosis in transgenic mice expressing simian cholesteryl ester transfer protein. Nature, 364, 73e75. Matsuura, F., Wang, N., Chen, W., Jiang, X.-C., & Tall, A. R. (2006). HDL from CETP-deficient subjects shows enhanced ability to promote cholesterol efflux from macrophages in an apoEand ABCG1-dependent pathway. J Clin Invest, 116, 1435e1442. Morehouse, L. A., Sugarman, E. D., Bourassa, P.-A., et al. (2007). Inhibition of CETP activity by torcetrapib reduces susceptibility to diet-induced atherosclerosis in New Zealand white rabbits. J Lipid Res, 48, 1263e1272. Morton, R. E. (1985). Binding of plasma-derived lipid transfer protein to lipoprotein substrates: the role of binding in the lipid transfer process. J Biol Chem, 260, 12593e12599. Morton, R. E., & Steinbrunner, J. V. (1990). Concentration of neutral lipids in the phospholipid surface of substrate particles determines lipid transfer protein activity. J Lipid Res, 31, 1559e1567. Morton, R. E., & Zilversmit, D. B. (1982). Purification and characterization of lipid transfer protein(s) from human lipoprotein-deficient plasma. J Lipid Res, 23, 1058e1067. Nagano, M., Yamashita, S., Hirano, K.-i., et al. (2004). Molecular mechanisms of cholesteryl ester transfer protein deficiency in Japanese. J Athero Thromb, 11, 110e121. Nicholls, S. J., Tuzcu, E. M., Brennan, D. M., Tardif, J.-C., & Nissen, S. E. (2008). Cholesteryl ester transfer protein inhibition, high-density lipoprotein raising, and progression of coronary
32
The HDL Handbook
atherosclerosis: insights from ILLUSTRATE (Investigation of Lipid Level Management Using Coronary Ultrasound to Assess Reduction of Atherosclerosis by CETP Inhibition and HDL Elevation). Circulation, 118, 2506e2514. Nishida, H. I., Arai, H., & Nishida, T. (1993). Cholesterol ester transfer mediated by lipid transfer protein as influenced by changes in the charge characteristics of plasma lipoproteins. J Biol Chem, 268, 16352e16360. Nissen, S. E., Tardif, J.-C., Nicholls, S. J., et al. for the ILLUSTRATE Investigators. (2007). Effect of torcetrapib on the progression of coronary atherosclerosis. N Engl J Med, 356, 1304e1316. Ordovas, J. M., Cupples, L. A., Corella, D., et al. (2000). Association of cholesteryl ester transfer protein-TaqIB polymorphism with variations in lipoprotein subclasses and coronary heart disease risk: the Framingham study. Artieroscler Thromb Vasc Biol, 20, 1323e1329. Pattnaik, N. M., Montes, A., Hughes, L. B., & Zilversmit, D. B. (1978). Cholesteryl ester exchange protein in human plasma isolation and characterization. Biochim Biophys Acta, 530, 428e438. Pattnaik, N. M., & Zilversmit, D. B. (1979). Interaction of cholesteryl ester exchange protein with human plasma lipoproteins and phospholipid vesicles. J Biol Chem, 254, 2782e2786. Plump, A. S., Masucci-Magoulas, L., Bruce, C., Bisgaier, C. L., Breslow, J. L., & Tall, A. R. (1999). Increased atherosclerosis in apoE and LDL receptor gene knock-out mice as a result of human cholesteryl ester transfer protein transgene expression. Arterioscler Thromb Vasc Biol, 19, 1105e1110. Qiu, X., Mistry, A., Ammirati, M. J., et al. (2007). Crystal structure of cholesteryl ester transfer protein reveals a long tunnel and four bound lipid molecules. Nature Struct Mol Biol, 14, 106e113. Rader, D. J., Alexander, E. T., Weibel, G. L., Billheimer, J., & Rothblat, G. H. (2009). The role of reverse cholesterol transport in animals and humans and relationship to atherosclerosis. J Lipid Res, 50, S189eS194. Rajaram, O. V., White, G. H., & Barter, P. J. (1980). Partial purification and characterization of a triacylglycerol-transfer protein from rabbit serum. Biochim Biophys Acta, 617, 383e392. Rye, K.-A., Hime, N. J., & Barter, P. J. (1995). The influence of cholesteryl ester transfer protein on the composition, size, and structure of spherical, reconstituted high density lipoproteins. J Biol Chem, 270, 189e196. Rye, K.-A., Hime, N. J., & Barter, P. J. (1997). Evidence that cholesteryl ester transfer proteinmediated reductions in reconstituted high density lipoprotein size involve particle fusion. J Biol Chem, 272, 3953e3960. Seip, R. L., Moulin, P., Cocke, T., et al. (1993). Exercise training decreases plasma cholesteryl ester transfer protein. Arterioscler Thromb Vasc Biol, 13, 1359e1367. Tall, A. R. (1986). Plasma lipid transfer proteins. J Lipid Res, 27, 361e367. Thompson, A., Di Angelantonio, E., Sarwar, N., et al. (2008). Association of cholesteryl ester transfer protein genotypes with CETP mass and activity, lipid levels, and coronary risk. J Am Med Assoc, 299, 2777e2788. Tsujita, Y., Nakamura, Y., Zhang, Q., et al. (2007). The association between high-density lipoprotein cholesterol level and cholesteryl ester transfer protein TaqIB gene polyporphism is influenced by alcohol drinking in a population-based sample. Atherosclerosis, 191, 199e205. Van Eck, M., Ye, D., Hildebrand, R. B., et al. (2007). Important role for bone marrow-derived cholesteryl ester transfer protein in lipoprotein cholesterol redistribution and atherosclerotic lesion development in LDL receptor knockout mice. Circ Res, 100, 678e685. Yu, B., Hailman, E., & Wright, S. D. (1997). Lipopolysaccharide binding protein and soluble CD14 catalyze exchange of phospholipids. J Clin Invest, 99, 315e324.
Chapter j 2
The Role of Cholesteryl Ester Transfer Protein (CETP)
33
Zhang, Z., Yamashita, S., Hirano, K.-i., et al. (2001). Expression of cholesteryl ester transfer protein in human atherosclerotic lesions and its implication in reverse cholesterol transport. Atherosclerosis, 159, 67e75. Zilversmit, D. B., Hughes, L. B., & Balmer, J. (1975). Stimulation of cholesterol ester exchange by lipoprotein-free rabbit plasma. Biochim Biophys Acta, 409, 393e398.
Chapter 3
Plasma Cholesteryl Ester Transfer Protein (CETP) in Relation to Human Pathophysiology Akihiro Inazu Department of Laboratory Sciences, School of Health Sciences, Institute of Medical, Pharmaceutical and Health Sciences, Kanazawa University, Ishikawa, Japan
INTRODUCTION Plasma low density lipoprotein (LDL) transports cholesterol from liver to peripheral tissues including the adrenal glands and gonads. On the other hand, high density lipoprotein (HDL) transports cholesterol from peripheral tissues including atheroma to liver, subsequently to bile and feces via the so-called reverse cholesterol transport (RCT) pathway (Figure 3.1). The structure of cholesterol is resistant to enzymatic degradation in the human body, hydroxylation is the only pathway to modify cholesterol for excretion from the body. In humans, HDL consists of particles heterogeneous in size, density and apolipoprotein composition. HDL is a vehicle for cholesterol, triglyceride, and phospholipids. Also, HDL has several apolipoproteins and enzymes on its surface that either promote or inhibit triglyceride or phospholipid lipolysis, inhibit hydroperoxidation of lipids, and promote lipid transfer among lipoproteins. In addition, HDL may be a platform for complement regulation, coagulation and inflammation (Scanu and Edelstein, 2008). Plasma HDL content levels are usually measured as cholesterol level concentration, but its particle numbers are better assessed by apolipoprotein A-I levels. This distinction may be due in part to the fact that as interindividual differences of plasma HDL-cholesterol, HDL2-cholesterol levels appeared to be highly variable, but HDL3 remains constant. Smoking and male sex decrease HDL2 levels, but alcohol intake and exercise increase them. HDL2 levels were determined from the catabolic rate of apolipoprotein The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
35
36
The HDL Handbook
CM VLDL LPL ApoA-I
CE
CETP
TG
PLTP
HL HDL1
HDL2 ApoE
Cells ABCA1
HDL3
FC, PL
LCAT ApoA-II α 1 HDL α 2-3 HDL Preβ1-HDL
FIGURE 3.1 Schema for HDL metabolism. Plasma cholesteryl ester transfer protein (CETP) facilitates exchange of neutral lipids of CE and TG between chylomicron (CM)/VLDL and HDL2. HDL-TG is provided by CETP, and it is subsequently hydrolyzed by hepatic lipase (HL). The synthetic rate of preb1-HDL is positively correlated with lipoprotein lipase (LPL)-mediated lipolysis or PLTP-mediated PL/FC transfer and increased cholesterol efflux by the ABCA1 transporter. On the other hand, the catabolic rate of preb1-HDL is correlated with cholesterol esterification rate by lecithin:cholesterol acyltransferase (LCAT). Thus, preb1-HDL levels are determined by activities of LPL, PLTP, ABCA1, and LCAT.
A-I and A-II rather than altered synthesis rate. The catabolic rate of HDL apolipoproteins is determined by HDL particle size. The smaller HDL tends to be catabolized faster in the kidney or other tissues. One of the determinants of HDL neutral lipid composition is plasma cholesteryl ester transfer protein (CETP). In incubated human plasma, transfer and equilibration of lecithin:cholesterol acyltransferase (LCAT)-generated CE is found, but the transfer of CE among lipoproteins was not found in rat (Barter and Lally, 1978). Similarly, mice, dogs, and pigs are members of a group of low plasma CETP activity, but rabbits and monkeys belong to a group of high CETP activity. Humans, hamsters, guinea pigs and chickens belong to a group with intermediate CETP activity. Interestingly, more phospholipid transfer protein activity is found in plasmas of low CETP activity animals (Ha and Barter, 1982; Cheung et al., 1996). Plasma CETP binds neutral lipids (cholesteryl ester (CE) or triglyceride (TG)) and phospholipid (PL) on HDL3, but CETP selectively promotes an exchange of CE and TG among lipoproteins. On the one hand, HDL-TG can be hydrolyzed by hepatic lipase and, on the other hand, plasma CETP decreases HDL particle size via CE/TG exchange between chylomicron/VLDL and HDL. Thus, CETP thereby accelerates the catabolic rate of HDL apolipoproteins (Lamarche et al., 1999).
Chapter j 3
CETP and Atherosclerosis
37
STRUCTURE OF CETP Plasma CETP was initially isolated as a highly purified 74 kD protein (Pattnaik et al., 1978). The human CETP gene is located at chromosome 16q13, near the locus of the LCAT gene. The CETP gene consists of 16 exons, spanning 25 kb (Agellon et al., 1990). The CETP mRNA encodes 476 amino acids (Drayna et al., 1987). The mature CETP contains four N-linked sugars (88, 240, 341, and 396) with a variable glycosylation site of 341Asn (Stevenson et al., 1993). CETP mRNA is expressed in various tissues, but liver cells, adipocytes and macrophages are abundant sources. Exon 9 works a cassette exon to generate short mRNA missing the sequences in frame in addition to full-length mRNA, but the splice-out variant is not efficiently secreted (Inazu et al., 1992). The C-terminal 26 amino acids of CETP form an amphipathic helix. Hydrophic residues bind to surface lipoproteins, and hydrophobic residues such as Leu, Phe are essential for binding neutral lipids such as CE and TG (Wang et al., 1993). The crystal structure of CETP shows that CETP forms a long tunnel occupied by four lipid molecules, two of CE or TG located inside the tunnel and two of PL plugging both sides of the tunnel openings. CETP is one of the lipopolysaccharide binding protein (LBP) family members. CETP exhibits an elongated boomerang shape located on the lipoprotein surface. Based on molecular size, CETP might prefer CE transfer rather than TG because of steric hindrance to TGs at the tunnel neck around residues 433, 443, 457, and 459 (Qiu et al., 2007).
REGULATION OF CETP EXPRESSION Cholesterol-rich and saturated fat-rich diets increased CETP expression via a liver X receptor (LXR) element in the promoter, a direct repeat of a nuclear receptor binding sequence separated by four nucleotides (DR4) (Luo and Tall, 2000). Among drugs that lower lipid levels, probucol increased plasma CETP activity (þ20%), but pravastatin decreased it (20%) (Inazu et al., 1999). Statins decrease both cholesterol and oxysterol, the latter being a ligand for LXRa activity (Masson et al., 2004). Thus, statins could decrease CETP mRNA levels through diminished LXR activity. However, the molecular effect of probucol on CETP expression is unknown, but it may be associated with increased cholesterol content in the liver by the remnant pathway. Unlike bezafibrate, fenofibrate decreased plasma CETP activity (20e30%) (Guerin et al., 1996; Watts et al., 2006). Since a putative peroxisome proliferator response element (PPRE) is located just upstream of the LXRasite, peroxisome proliferator-activated receptor a (PPARa) could suppress CETP promoter activity by antagonizing LXR activity (Cheema et al., 2005).
38
The HDL Handbook
Nicotinic acid is a well-established lipid-lowering agent. Side effects such as flushing may restrict drug usage, however, recent identification of a G protein-coupled receptor GPR109A and of a prostaglandin 2 (PGD2) receptor antagonist (laropiprant) may provide strategies to control side effects. Nicotinic acid is a powerful inhibitor of fat-mobilizing lipolysis via hormone-sensitive lipase in adipose tissue, and therefore limits free fatty acid (FFA) flux into the liver. Nicotinic acid lowers TG as well as Lp(a) levels, and increases HDL-C levels by z20e40%. Nicotinic acid may induce PPARg expression but it is also reported to lower CETP activity. Nicotinic acid increased HDL cholesterol levels by reducing hepatic CETP mRNA only in mice expressing the human CETP transgene (Hernandez et al., 2007; Van der Hoorn et al., 2008).
FUNCTION OF CETP LCAT promotes free cholesterol (FC) esterification in HDL3 and CETP transfers newly-esterified CE from HDL3 to VLDL or chylomicrons. Thus, these tandem reactions appear to be physiological. However, it is unclear whether or not CETP is pro-atherogenic, but it is likely that its atherogenicity is dependent on the metabolic context of lipoprotein receptors expressed in the liver, which are major determinants of RCT pathways to the liver.
LCAT, cholesterol esterification, and CE transfer rate Table 3.1 shows simultaneously determined plasma exogenous LCAT activity and cholesterol esterification rate (CER) and exogenous CETP activity in
TABLE 3.1 Plasma endogenous and exogenous LCAT activity and exogenous CETP activity in young women (n ¼ 38) HDL enzyme/protein
Method
Mean (SD)
Range
LCAT, endogenous
Self substrate method (Nagasaki and Akanuma, 1977)
110 (20)
80e180
LCAT, exogenous
Common substrate method (Manabe et al., 1987)
590 (110)
330e900
CETP
NBD-cholesteryl ester transfer activity between proteoliposome to VLDL
210 (30)
150e290
All units are nmol/ml/h. Endogenous LCAT activity of cholesterol esterification rate is only z20% of exogenous LCAT activity, the latter is correlated with plasma LCAT mass. Endogenous LCAT activity is only z50% of plasma CETP activity, therefore HDL-FC/CE ratio could be altered in heterozygous CETP deficiency.
Chapter j 3
CETP and Atherosclerosis
39
young women. Plasma CER is an endogenous LCAT reaction, which shows only z20% of maximal enzymatic activity of LCAT. Net CE transfer rate from HDL to VLDL is only 20e50 nmol/ml/h despite exogenous CETP activity of z200 nmol/ml/h (Pruneta et al., 1999). Since the net CE transfer rate is smaller than that of CER, CE in HDL needs to be directly catabolized in the liver. Therefore, plasma VLDL levels appear to be a rate-limiting step of net CE transfer rate in the fasting state. However, in the post-prandial state, net CE transfer is accelerated because the increased VLDL/chylomicrons provide increased CE acceptor capacity, and the clearance of LDL-CE or remnant-CE is dependent on LDL-receptor activity or remnant receptor (LRP) in the liver. Since HDL-FC is more rapidly catabolized in the liver than HDL-CE in a monkey study, selective uptake of FC without endocytosis of HDL apolipoproteins appears to be a predominant pathway of HDL-C catabolism in the liver (Scobey et al., 1989). Thus, HDL-CE pathways play minor roles in human HDL-cholesterol catabolism (Schwartz et al., 2004). Also, FC from HDL is efficiently secreted in bile, but not from other lipoproteins (Robins and Fasulo, 1997). Thus, cholesterol esterification is not necessarily required for the selective uptake of HDL-cholesterol in the liver via hepatic lipase and scavenger receptor class BI (SR-BI) mediated RCT pathways. The lipoprotein phenotype of high HDL and the low CER appear to be anti-atherogenic as long as efficient RCT is maintained in the liver.
Modulators of lipid transfer CETP-mediated lipid transfer is not preferably directed toward a specific lipoprotein in a reconstituted system. Because CE is generated in HDL via the LCAT reaction, higher CE concentrations are found in HDL. Therefore, net CE transfer operates from HDL to other lipoproteins in vivo. Similarly, because chylomicrons and VLDL are rich in TG, net TG transfer is found from chylomicron/VLDL to other lipoproteins via heteroexchange of CE and TG. In addition, some specific apolipoproteins and TG lipolysis occurring during the post-prandial state would modify the direction of lipid transfer among lipoproteins. As a modulator of CE transfer, apoF was identified as lipid transfer inhibitor protein (LTIP). LTIP inhibits CE transfer between VLDL and LDL, whereas it increases CE transfer from HDL to VLDL (Wang et al., 1999). Phospholipid transfer protein (PLTP) promotes PL transfer from VLDL to HDL, in addition to the PL transfer activity of CETP. Also, PLTP possesses free cholesterol and vitamin E transfer activity. As CE acceptors of the CE transfer reaction, VLDL and chylomicrons are active when lipolysis has occurred. VLDL-bound LPL and FFA levels may have a positive effect on the binding between CETP and lipoproteins, thereby accelerating CE mass transfer.
40
The HDL Handbook
Effects of CETP on LDL subclass remodeling An association of large LDL and low CETP activity with TaqIB polymorphism was found in men, but not in women in the genetic epidemiological survey of the Framingham Study (Ordovas et al., 2000). In remodeling of apoBcontaining lipoproteins, addition of CE increases lipoprotein size and deletion of PL and FC decrease its size, resulting in two homogeneous LDL subclasses (Musliner et al., 1991). Complete CETP deficiency produced unique characteristics of broad LDL band with (at least five) distinct IDL-LDL subclasses on a native polyacrylamide gel (Sakai et al., 1991), but partial CETP deficiency increased LDL size.
Effects of CETP on macrophage-specific RCT in mice and hamsters In radioactive cholesterol-labeled macrophage methodology, Rader et al. (Tanigawa et al., 2007) have shown direct RCT from peripheral macrophages to liver, bile and feces. In LDLR-KO mice, CETP cDNA adeno-associated virus mediated transfection promotes cholesterol transport to the liver, but not to bile and feces. In contrast, in SRBI-KO mice, CETP cDNA transfection increased cholesterol loss in the feces, indicating induction of overall RCT via active LDL-R activity despite diminished selective uptake of HDL-CE (or FC) in the liver (Tanigawa et al., 2007). The former model is similar to the setting of familial hypercholesterolemia (FH) or downregulated LDL receptor activity by a saturated-fat diet, while the latter model of decreased SR-BI activity reflects conditions found in hormone replacement therapy (Zhang et al., 2007). Thus, macrophage-specific RCT is dependent on CETP activity and active LDL receptors in the liver, and the efficacy of fecal sterol excretion is compatible with effects of anti-oxidative agents such as vitamin E and probucol, which would increase plasma CETP activity, in cholesterol-fed rabbits (Jeon et al., 2005). In hamsters, torcetrapib, a CETP inhibitor, elevated both HDL-C levels and the amounts of cholesterol and bile acids secreted in feces, indicating an overall increase in RCT (Tchoua et al., 2008). Such a difference may be explained (at least in part) by the presence of natural CETP activity and the inducible CYP7A gene found in hamsters (Zhang et al., 2004).
Effects on cholesterol efflux and preb-HDL formation Subjects with complete CETP deficiency have more preb-HDL despite less remodeling from large HDL to small subclasses via CETP (Asztalos et al., 2004). Thus, increased preb-HDL levels are caused by impaired maturation to large HDL due to decreased endogenous LCAT activity (Oliveira et al., 1997) or increased lipolysis of TG-rich lipoproteins (Miyazaki et al., 2009). Since LCAT mass and exogenous LCAT activity remain at normal levels, impaired LCAT activity is
Chapter j 3
CETP and Atherosclerosis
41
explained either by: (1) end-product inhibition, namely excess CE in large HDL or (2) altered phospholipid composition, such as sphingomyelin (SM) levels, in HDL. Plasma cholesterol esterification rate was decreased in CETP deficiency, which is compatible with altered lipid composition found in HDL fractions of homozygous CETP deficiency; i.e., high CE/TG ratio and low PL/FC ratio (Koizumi et al., 1991). Since CER is inversely associated with SM/PC ratio in HDL, SM itself may be an unsuspected link between low cholesterol esterification rate and low CETP activity (Noguchi and Inazu, unpublished data). However, SM-rich lipoproteins are not always pro-atherogenic because SM avidly binds cholesterol, and HDL with increased SM levels may be good acceptors for cholesterol efflux from atherosclerotic plaques (Fournier et al., 1997). Increased HDL levels found in CETP deficiency had no beneficial effect on the ABCA1-mediated cholesterol efflux but did enhance SR-BI-mediated efflux (Miwa et al., 2009).
Lipoprotein metabolism in CETP deficiency from a kinetic study Initially, Ikewaki et al. reported delayed catabolism of apoA-I and apoA-II in human subjects with CETP deficiency (Ikewaki et al., 1993). Also, they reported an increased catabolic rate of LDL-apoB in addition to a decreased production rate of VLDL-apoB (Ikewaki et al., 1995). In a CETP-deficient dog, Ouguerram et al. reported that VLDL and LDL CE metabolism was coupled to apoB catabolism without enrichment of CE during VLDL-LDL conversion and that 60% of HDL CE turnover was mediated by a selective uptake pathway (Bailhache et al., 2004; Ouguerram et al., 2004). As compared to other CETP-deficient animals, dogs have higher selective uptake of HDLCE (60% vs. 25e30% in rats and mice). The cholesterol esterification rate of dog plasma is 160 nmol/ml/h, which is between the rates in humans (30e80 nmol/ml/h) and in rats (300 nmol/ml/h). Thus, dogs may have an efficient RCT due to high activities of SR-BI and LCAT in addition to CETP deficiency.
ROLE OF CETP IN APOE-RICH HDL FORMATION Structure of apoE-rich HDL Plasma HDL is classified as HDL1 (density 1.08e1.09 g/ml), HDL2 (1.09e1.15), and HDL3 (1.15e1.18). HDL1 is apoE-rich with a diameter of 13e19 nm and increased LCAT activity compared with HDL2 and HDL3 (Schmitz and Assmann, 1982). HDL1 is also identified in cholesterol-fed CETP-deficient animals such as dogs and pigs.
Function of apoE-rich HDL In cultured smooth muscle cells, cholesterol from HDLc, lipoproteins with apoE only (density 1.006e1.02), present in cholesterol-fed canine plasma,
42
The HDL Handbook
was efficiently delivered to the cells as well as LDL (Mahley et al., 1977). ApoE-rich HDL appears in various situations such as genetic dyslipidemia, but its characteristics may not be uniform. In cholesterol-fed canines, plasma cholesterol increases, HDL loses apoA-I but it gains apoE. HDLc appears (as well as occurrence of LDL and b-migrating VLDL) when cholesterol exceeds 700 mg/dl. Thus, HDL1 and HDLc appeared to suppress apoB-containing lipoprotein formation in the liver, such as LDL and b-VLDL. Thus, one would predict that these lipoproteins would inhibit atherogenesis in canine models. ApoE-rich HDL have dual roles in atherogenicity. ApoE can serve as an LRP ligand, and therefore canine HDLc inhibits clearance of chylomicrons (Hussain et al., 1995). However, post-prandial lipemia is diminished in homoand heterozygous CETP deficiency (Inazu et al., 2008). VLDL lipolysis and hepatic uptake of CM/VLDL remnant appear to be increased probably due to apoE transfer from HDL to CM/VLDL during post-prandial periods (Krimbou et al., 2003). None the less, apoE-rich HDL reduces LPL-mediated retention of LDL by subendothelial matrix, and therefore could play an anti-atherogenic role in artery walls. Also, apoE-rich lipoproteins protect cells from apoptosis via the LRP signaling pathway (Hayashi et al., 2007). In SB-BI knockout mice, LCAT activity was impaired and oxidative stress was increased in large HDL (Lee et al., 2007; Van Eck et al., 2007). Remnant-like particle cholesterol (RLP-C) levels reflect cholesterol levels (10e15 mg/dl) of large apoE-rich HDL (probably apoE only particles) in homozygous CETP deficiency (Inazu et al., 2008). The apoE-rich HDL contains apoA-IV as well as apoA-I (Bisgaier et al., 1991), but the RLP fraction of homozygous CETP deficiency had large amounts of apoE with a trace of apoA-I and apoA-IV.
MOLECULAR GENETICS AND THE ETHNIC DIFFERENCE IN THE FREQUENCY OF HUMAN CETP DEFICIENCY Plasma CETP deficiency was originally reported in Japanese siblings with hyperalphalipoproteinemia (HALP) (Koizumi et al., 1985). The first mutation was found in a splice donor site mutation in intron 14 (intron 14 G(þ1)-to-A), resulting in non-translation of exon 14 and production of a stop codon in the fourth codon encoded by exon 15. These changes resulted in decreased mRNA levels to one-third of controls and a truncated protein that appeared to be rapidly degraded (Brown et al., 1989; Gotoda et al., 1997). So far, 20 different mutations have been found both in Asian and Caucasian populations, but predominantly in Asians (Nagano et al., 2004; Thompson et al., 2009). Two mutations were found in both ethnic groups (R268X and intron 14 G(þ1)-to-A), suggesting multiple origins of these mutations (de novo mutations) (Ai et al., 2009). Both mutations indeed have CpG sequences as mutational hot spots for deamination of the cytosine. Although many mutations
Chapter j 3
CETP and Atherosclerosis
43
are nonsense or splicing mutations, four missense mutations are reported to be associated with decreased CETP activity (L151P, L261R, R282C, and D442G). Only one promoter mutation was reported at e69G>A. Large differences in the frequency of CETP deficiency in various populations appear to be related to the frequency of two variants. The intron 14 G (þ1)-to-A mutation is the Japanese-type mutation with the higher gene frequency (0.8% in the general population of Japan). Homozygotes of this mutation were reported in >50 cases reflecting relatively higher frequency of consanguinity in Japan in past generations.
CLINICAL CHEMISTRY OF LDL-CHOLESTEROL AND HDL-CHOLESTEROL IN CETP DEFICIENCY Homozygotic mutations result in complete CETP deficiency with a phenotype of very high HDL-C levels and relatively low LDL-C levels (mean levels of 164 mg/dL and 77 mg/dL, respectively) (Inazu et al., 1990). Heterozygotes have a moderate increase in HDL-C (mean 66 mg/dL) and a decrease in plasma CETP levels (mean 1.4 mg/L) as compared to unaffected controls (53 mg/dL and 2.3 mg/L). Asp 442 Gly (D442G) is another highly prevalent mutation in Japan (3.4% in the general population of Japan) as well as in other Asian populations (1.7e5.9%), although it is only partially defective in CETP activity (Inazu et al., 1994). The compound heterozygotes of intron 14 G(þ1)to-A and D442G produce a less severe phenotype of CETP deficiency (n ¼ 9, CETP 0.9 0.3 (SD) mg/L, HDL-C 130 24 mg/dL) as compared to mean levels of plasma CETP 1.8 0.6 mg/L (SD) in Japanese men and 2.0 0.5 in women (Kiyohara et al., 1998).
LDL-C measurement The Friedewald formula, LDL-C ¼ TC HDL-C (TG/5), is used for estimation of LDL-C, but accurate measurement for HDL-C is required. For the precipitation method for HDL-C, the Cholesterol Reference Method Laboratory Network (CRMLN) using a heparin, Mn2þ supernatant cholesterol of plasma d > 1.006 (Centers for Disease Control and Prevention (CDC)) is better than the Designed Comparison Method (DCM) using dextran-sulfate, Mg2þ supernatant cholesterol levels, since the latter precipitates apoE-rich HDL in addition to apoB-containing lipoproteins, but the former does not. Even if accurate measurement of HDL-C is accomplished, cholesterol levels in VLDL are relatively decreased in CETP deficiency (Koizumi et al., 1991). Therefore, the Friedewald formula would underestimate LDL-C. However, since the density between 1.019 and 1.063 includes apoE-rich large HDL such as HDL1, the LDL-C separated by ultracentrifugation would overestimate LDL-C. Suitability for homogeneous LDL-C assays has not been reported in homozygous CETP deficiency.
44
The HDL Handbook
ApoE-rich HDL-C determination As a more suitable precipitation method for HDL-C in CETP deficiency, Chiba et al. reported that 13% polyethylene glycol (PEG) allows recovery of total HDL in the supernatant (Chiba et al., 1997). In that study, patients with complete CETP deficiency had a mean HDL-C level of 121 mg/dL detected by a commercial polyanionic reagent (dextran sulfate, sodium phosphotungstate, Mg2þ), but 176 mg/dL of total HDL-C using supernatants produced by the PEG method. The difference (z55 mg/dL) may indicate cholesterol levels in apoErich HDL.
EPIDEMIOLOGY OF INCREASED HDL CHOLESTEROL LEVELS AND CETP DEFICIENCY HDL-cholesterol could be excreted from bile as consequence of reverse cholesterol transport (RCT) involving HDL maturation from preb-HDL to apoE-rich HDL. However, CETP would bypass the cholesterol flow from HDL to VLDL-LDL without involving the liver. Thus, CETP-mediated CE transfer would increase indirect cholesterol transport to the liver via VLDL-IDL-LDL through LDL receptor or remnant receptor pathways. In addition, HDLcholesterol is directly transported to the liver by selective uptake of HDL-CE or FC via hepatic lipase and/or SR-BI pathway. Thus, the role of the CETP pathway appears to be anti-atherogenic when LDL levels are low and TRL clearance is rapid on a low-fat diet. However, the Western-type saturated-fat diet would suppress LDL receptor expression, and the flow of HDL-derived cholesterol back to the liver would be diminished via the LDL pathway. Indeed, subjects with high CETP activity may manifest lower coronary risk in the presence of low plasma TG levels (Borggreve et al., 2007). The role of CETP in LDL-receptor deficiency (familial hypercholesterolemia) is controversial, since double heterozygotes with FH and CETP deficiency are not protected from CHD (Haraki et al., 1997). It remains to be discussed whether slightly increased HDL-C (60 mg/ dL vs. 46 mg/dL) is not sufficient to prevent CHD or whether lower CETP is disadvantageous in FH.
Epidemiology of HDL cholesterol In earlier studies by Gofman et al. (Gofman et al., 1966), levels of HDL2 and HDL3 were significantly decreased in patients with CHD, but HDL1 levels were not changed. In heparin-sepharose chromatography, HDL-apoE levels were significantly decreased in survivors with myocardial infarction (Wilson et al., 1993). In a proteomic study of HDL proteins, apoE levels in the HDL3 fraction were increased in patients with CHD (Vaiser et al., 2007), but unfortunately no data were available on HDL2 or VLDL.
Chapter j 3
45
CETP and Atherosclerosis
Inconsistency of anti-atherogenicity of HDL might be explained by how much large HDL or apoE-rich HDL are increased, as these particles are believed to have less anti-atherogenic effects compared to small HDL. When apoA-I and apoB are kept constant, HDL-C and HDL particle size may confer risk at very high values (van der Steeg et al., 2008). On the contrary, apoA-I is a negative risk factor even when corrected for HDL-C and apoB, suggesting that HDL number assessed by apoA-I concentration is statistically more important than HDL size for anti-atherogenicity effects. The debate over whether HDL size or its components are more important for atherogenesis should be answered by measuring specific HDL-related lipid component levels such as sphingomyelin, sphingosine-1-phosphate, and dolichol in HDL of various dyslipidemias (Kontush et al., 2007).
The role of confounding factors in increased HDL state In many reports, a high HDL-C state is a negative risk factor for CHD and stroke (Kurth et al., 2007). Since low HDL-C is inversely associated with increased TG levels, low HDL appeared to be a marker for disturbed TG metabolism (Schaefer et al., 1994). However, some reports suggest a U-shape relationship with HDL-C and vascular events (Chien et al., 2002). Such a relationship may be associated with some confounding factors associated with increased HDL-C levels: alcohol, estrogen and exercise (Williams, 1996). There are reports of an adverse interaction between alcohol and hypertension on stroke (Leppala et al., 1999). Others have suggested an interaction between increased levels of TG and HDL-C on CHD (Jeppesen et al., 1998). A large genetic epidemiological survey is warranted to find associations between CETP or hepatic lipase polymorphisms and CHD events, especially by interacting with environmental factors such as alcohol consumption and hormone replacement therapy. Since lower activities of hepatic lipase and SR-BI and higher CETP are characteristics of premenopausal women, consideration of gender difference is necessary in unraveling the interactions between HDL and CHD (Jansen et al., 2002) (Table 3.2).
TABLE 3.2
Gender difference in HDL-associated biological activities Transfer proteins, enzyme, receptor
Lipoprotein phenotype
CETP
PLTP
Hepatic lipase
SR-BI
HDL2b
preb-HDL
Men
Low
High
High
High
Low
High
Women*
High
Low
Low
Low
High
Low
LPL activity is not different between men and women. * Premenopausal state
46
The HDL Handbook
The role of CETP mutations and polymorphisms on CHD risk A meta-analysis of studies including CETP gene single nucleotide polymorphisms (SNPs) of TaqIB2, -629C>A and Ile 405 Val (I405V) showed that the genotypes with low CETP may have anti-atherogenic effects (Thompson et al., 2008). Our data suggested that e1337C>T is responsible for the antiatherogenicity of the well-investigated TaqIB2 allele in the Japanese population (Lu et al., 2003; Takata et al., 2006). Similarly, anti-atherogenicity of lower CETP levels was also suggested in heterozygous CETP deficiency (Curb et al., 2004). CHD prevalence appears to be low in homozygous CETP deficiency, which is compatible with findings of the Kochi Study cross-sectional survey of disease prevalence stratified by increased HDL-cholesterol levels >80 mg/dL and >100 mg/dL (Moriyama et al., 1998). Three hundred subjects with HDL-C > 100 mg/dL were found in that paper, but no case with CHD was found. Indeed, high HDL-cholesterol and intron 14 G(þ1)>A variant may increase the odds for healthy aging in the Honolulu Heart Program Study (Koropatnick et al., 2008). Consistently, recent case reports of Caucasian CETP deficiency have shown the rarity of atherosclerotic disease even though Western diets were consumed (Teh et al., 1998; Rhyne et al., 2006). However, some investigators believe pro-atherogenicity in some cases with homozygous CETP deficiency (Nagano et al., 2005). In contrast, there is no defined CHD, but two cases with cerebrovascular disease were found in our cohort of homozygous CETP deficiency (n ¼ 53).
Malignancy association In earlier studies, Keys suggested a possible association between increased HDL-C levels and malignancy incidence (Keys, 1983). The failure of torcetrapib is a reminder of such a possible association. This issue should be assessed in studies using other CETP inhibitors.
DEVELOPMENT OF A CETP INHIBITOR Three compounds are currently in clinical trials, torcetrapib (CP-529414), anacetrapib (MK-859) and dalcetrapib (JTT-705/ Roche R1658). Phase III of torcetrapib was terminated in December, 2006 due to an unexpected excess of mortality in the torcetrapib arm. The early termination was partially explained by hypertension due to aldosterone excess. However, the role of CETP inhibition on the increased mortality was not clearly shown, but it may be rather associated with infection or malignancy than CHD (Barter et al., 2007). The vascular endpoints of carotid atherosclerosis and coronary atheroma volume assessed by intravascular ultrasound showed no benefit from torcetrapib over a background of atorvastatin treatment, despite increased levels
Chapter j 3
CETP and Atherosclerosis
47
of HDL and further decreased levels of LDL and TG (Kastelein et al., 2007; Nissen et al., 2007). HDL-cholesterol might be excreted from bile as consequence of reverse cholesterol transport (RCT) involving HDL maturation from preb-HDL to apoE-rich HDL. Using a CETP inhibitor, CE uptake in liver was not decreased in rabbits, but fecal sterol excretion was not increased in patients taking torcetrapib, indicating that overall RCT was not significantly induced (Brousseau et al., 2005; Kee et al., 2006; Catalano et al., 2009). However, torcetrapib did increase overall RCT assessed by cholesterol and bile acids in feces of hamsters (Tchoua et al., 2008). Such a difference in the response to CETP inhibitors definitely needs to be clarified.
Effects on small HDL subclasses Hyperalphalipoproteinemia (HALP) caused by prednisone plus cyclosporine was ineffective in producing HDL acceptors for cholesterol efflux. The ABCA1-dependent efflux was maintained, but the non-ABCA1-dependent route appeared to be impaired (Sviridov et al., 2006). CETP inhibition may disturb apoA-I liberation from HDL in atherosclerotic lesions. Therefore, ABCA1-mediated cholesterol efflux activity to small HDL or liberated apoA-I could be compromised. However, recent studies suggest that the ABCG1 transporter may favorably induce cholesterol efflux from cells to large HDL (Yvan-Charvet et al., 2007). Torcetrapib would increase this large HDL level, which is an active cholesterol acceptor for ABCG1 or SR-BImediated efflux, although the role of SR-BI-mediated cholesterol efflux remains controversial (Yvan-Charvet et al., 2008). Although small HDL, such as HDL3 subclass, is known to protect LDL from oxidation (Davidson et al., 2009), HDL3 levels were not increased in genetic CETP deficiency, but they were moderately increased in patients with CETP inhibitors.
Effects on apoB-containing lipoproteins By inhibiting neutral lipid transfer among lipoproteins, CE transfer from HDL to VLDL in exchange with TG was diminished. Therefore, relatively CE-poor, TG-rich VLDLs were lipolyzed to LDL and VLDL-IDL-LDL were rapidly removed from the circulation probably due to LDL-receptor upregulation (Millar et al., 2006). In LDL subclasses, small-and-dense LDL levels were decreased but large LDL levels were increased in patients with torcetrapib (Brousseau et al., 2004), which is compatible with a phenotype in low CETP subjects with a TaqIB2 polymorphism (Ordovas et al., 2000). Plasma Lp(a) levels were decreased in CETP deficiency, and z50% reduction of plasma Lp(a) levels was achieved by anacetrapib (Bloomfield et al., 2009). Large HDL contains multiple apoE molecules, but such lipoproteins could be efficiently removed from the circulation via increased LDL receptor
48
The HDL Handbook
expression. Thus, increased levels of apoE-rich HDL produced by a CETP inhibitor could be offset by combination therapy with a statin, which induces LDL receptor expression and increases hepatic uptake of apoE-rich HDL. Increased RCT was found when both CETP and LDL receptor are upregulated in the liver. Thus, combination therapy with a CETP inhibitor and a statin would result in opposite responses in the RCT pathway, a finding that would be compatible with a proposed adverse pharmacogenetic interaction between a statin and a CETP inhibitor (Regieli et al., 2008).
ROLE OF CETP IN AGING AND LONGEVITY CETP enhances HDL remodeling from large HDL to small subclasses including pre-HDL. However, CETP deficiency would decrease cholesterol esterification rate, thereby inhibiting maturation of preb-HDL to a-migrating spherical HDL. Therefore, in CETP deficiency, large-to-small HDL remodeling is decreased and preb-HDL catabolism is also decreased. The levels of prebHDL were increased in homozygous CETP deficiency, but were decreased in the heterozygotes (Asztalos et al., 2004), indicating that maturation of the small HDL subclass is preserved in heterozygotes, but not in homozygotes. The difference is dependent on the magnitude of low CER and low ABCA1mediated efflux activity (Figure 3.2). Recent studies suggested PAF-AH (lipoprotein-associated phospholipase A2) inhibitors could inhibit sdLDL formation, thereby preventing atherosclerosis in animal models and humans. Plasma paraoxonase activity was decreased in HALP with hepatic lipase deficiency (Kontush et al., 2004). Thus, the antioxidant activity of HDL needs to be evaluated in patients treated with CETP inhibitors.
Effects on aging and Alzheimer’s disease A promising effect on longevity has been reported in Ashkenazi Jews, as increased homozygosity of I405V was found in offspring of individuals with exceptional longevity (mean age 98 years). These subjects had high HDL, low LDL and large LDL size, and low prevalence of hypertension and metabolic syndrome (Barzilai et al., 2003). A different CETP polymorphism (D442G) may have a protective effect against the development of Alzheimer’s disease (AD), especially in apoE4 carriers in the Chinese population (Chen et al., 2008). But, the opposite relationship between another CETP polymorphism (I405V) and AD was found in the Dutch population (Arias-Vasquez et al., 2007). Indeed, the CETP gene haplotype was associated with both markers of cholesterol synthesis and degradation in the cerebrospinal fluid and CETP may have neuronal repair effects through PL transfer activity. Data regarding possible AD associations between genes for CETP and apoE are conflicting, and need to be resolved.
Chapter j 3
49
CETP and Atherosclerosis
A
CM VLDL FC, PL
TG
CE
Cells
ApoA-I
HDL2
HDL2
HDL3 ApoA-II
α 1 HDL
α 2-3 HDL
Preβ1-HDL
B CM VLDL FC, PL ApoA-I
CE
HDL1
HDL2
ApoE
Cells
TG
HDL3 ApoA-II
α 1 HDL
α 2-3 HDL
Preβ1-HDL
FIGURE 3.2 Differential metabolic fate of HDL in heterozygous and homozygous CETP deficiency. (A) In heterozygotes, both CM/VLDL lipolysis and cellular ABCA1-mediated FC/PL efflux are maintained. Also, LCAT reaction is relatively preserved. Thus, decrease in preb1-HDL indicates that HDL maturation is not disturbed. (B) In homozygotes, CM/VLDL lipolysis is enhanced and the cellular efflux is diminished. Also, LCAT reaction is severely suppressed. Thus, preb1-HDL is accumulated in plasma.
Susceptibility to infectious disease As CETP belongs to the lipopolysaccharide binding protein (LBP) gene family, a role for CETP has been suggested in infection. Since LBP and lipoprotein may be associated with the detoxification of endotoxin, apoB-containing lipoprotein levels may reflect the efficacy of LBP function (Vreugdenhil et al., 2001). Along with an increase of LBP, plasma cholesterol, PL, LDL-C, HDL-C
50
The HDL Handbook
decrease, whereas plasma TG, VLDL and apoE-rich HDL tend to be increased after intravenous endotoxin (Hudgins et al., 2003; Li et al., 2008). In experimental endotoxemia, increased CRP levels are found with reciprocal decreases of LCAT and CETP activities (Levels et al., 2007). CETP expression was suppressed by cytokines of TNFa and IL-1. PLTP deficiency led to a significant increase in LPS-induced mortality in mice (Gautier et al., 2008). Thus, we need to consider possible disadvantages of CETP deficiency in terms of endotoxemia because CETP would enhance the LPS binding to HDL/ LDL. The liver uptake of LPS was greater in CETP-transgenic mice than controls, suggesting accelerated clearance of LPS from the circulation (Cazita et al., 2008). On the other hand, large HDL found in CETP deficiency might be protective against Schistosoma japonicum (Okamura-Noji et al., 2001).
ROLE OF CETP IN DYSLIPIDEMIA ASSOCIATED WITH DIABETES AND METABOLIC SYNDROME In hyperlipidemic patients, increased production of VLDL and/or decreased catabolism of LDL are major risk factors in addition to low HDL-C. Decreased production rate of VLDL appears to be associated with decreased CETP activity in patients with metabolic syndrome treated by fenofibrate (Watts et al., 2006). A CETP inhibitor may be especially useful for combined hyperlipidemia of high VLDL and low HDL levels. LDL catabolic rate is increased with CETP inhibitors, but effects on VLDL production rate have been less established. Plasma CETP levels are increased in metabolic syndrome (Sandhofer et al., 2006). Plasma PLTP levels are increased, but the increase in CETP is somewhat controversial in diabetes (Dallinga-Thie et al., 2007). Plasma CETP activity was positively related with CETP mass, and negatively to HbA1c (Dullaart et al., 2004). In diabetes, decreased sterol regulatory element binding protein (SREBP) expression may lead to lower CETP expression in the liver (MacLean et al., 2005). Phosphoinositide 3-kinase activity is decreased in diabetes, and liver SRBI expression is decreased (Shetty et al., 2006). Since increased CETP activity may be beneficial in diabetes as shown in a db/db mouse study (MacLean et al., 2003), such a complex relationship should be examined in human diabetes. In the Copenhagen City Heart Study, elevated HDL-C caused by an I405V polymorphism is a risk for CHD in women without hormone replacement therapy but not in men (Agerholm-Larsen et al., 2000). Atherogenicity of CETP may be related to SR-BI expression in terms of RCT to bile cholesterol excretion. Hepatic SR-BI is induced by a diet rich in polyunsaturated fat, but suppressed by cholesterol, vitamin E and estradiol. As shown in knock-out mice, SR-BI deficiency is associated with increased HDL levels but it is proatherogenic (Trigatti et al., 1999). Thus, specific attention may be required for diabetic and female patients when a CETP inhibitor is considered, because reduced SR-BI expression is
Chapter j 3
CETP and Atherosclerosis
51
assumed in those conditions. Also, usefulness of combination therapy with a statin (HMG-CoA reductase inhibitor) and a CETP inhibitor needs to be validated experimentally.
CLINICAL TRIALS OF CETP INHIBITORS More studies are needed for development of HDL intervention through inhibiting plasma CETP activity. Especially, it is important to assess how to suppress plasma CETP activity. Antisense CETP therapy is of greater interest than chemical compounds because the antisense therapy would decrease plasma CETP mass. In protection of neuronal disease, more study of cerebrospinal fluid lipoprotein is needed in terms of compositions of apoE and lipid transfer proteins and pharmaceutical changes in those lipoproteins. In addition to CE/TG transfer, CETP may transfer estrogen-ester and retinyl-ester but not vitamin E. Antioxidative local effects of CETP may be more important under oxidative stress or combined metabolic conditions of PTLP deficiency, which is defective in vitamin E transport. Thus, the role of CETP in terms of lipoprotein oxidation needs to be clarified in various settings of concurrent hyperlipoproteinemia or hormonal exposure such as estrogen. LPS is associated not only with endotoxemia, but also with vascular oxidative stress and inflammation. Therefore, the role of CETP in LPS metabolism needs to be clarified in CETP deficiency and patients treated with CETP inhibitors.
LOW CETP STATUS, GENETIC OR ENVIRONMENTAL? Recent publications suggested that low CETP mass is associated with lipidlowering drugs, history of myocardial infarction, diabetes, smoking, and inflammation with elevations of CRP and IL-6 (Ritsh et al., 2010). Indeed, CETP expression was decreased in leukocytes and macrophages in acute coronary syndrome (Ye et al., 2008), suggesting downregulation of CETP expression during acute inflammation. In addition, Vasan et al. have shown that low CETP activity was associated with greater cardiovascular risk in a prospective study of the Framingham Heart Study (Vasan et al., 2009). Thus, the cause of low CETP activity needs to be clarified to insight conflicting data between CETP activity and cardiovascular risk. Thus, both genetic and environmental factors need to be assessed in a cardiovascular health study including plasma CETP mass or activity.
REFERENCES Agellon, L. B., Quinet, E. M., Gillette, T. G., Drayna, D. T., Brown, M. L., & Tall, A. R. (1990). Organization of the human cholesteryl ester transfer protein gene. Biochemistry, 29, 1372e1376.
52
The HDL Handbook
Agerholm-Larsen, B., Nordestgaard, B. G., Steffensen, R., Jensen, G., & Tybjaerg-Hansen, A. (2000). Elevated HDL cholesterol is a risk fator for ischemic heart disease in white women when caused by a common mutation in the cholesteryl ester transfer protein gene. Circulation, 101, 1907e1912. Ai, M., Tanaka, A., Shimokado, K., et al. (2009). A deficiency of cholesteryl ester transfer protein whose serum remnant-like particle (RLP)-triglyceride significantly increased, but RLPcholesterol did not after an oral fat load. Ann Clin Biochem, 46, 457e463. Arias-Vasquez, A., Isaacs, A., Aulchenko, Y. S., et al. (2007). The cholesteryl ester transfer protein (CETP) gene and the risk of Alzheimer’s disease. Neurogenetics, 8, 189e193. Asztalos, B., Horvath, K. V., Kajinami, K., et al. (2004). Apolipoprotein composition of HDL in cholesteryl ester transfer protein deficiency. J Lipid Res, 45, 448e455. Bailhache, E., Briand, F., Nguyen, P., Krempf, M., Magot, T., & Ouguerram, K. (2004). Metabolism of cholesterol ester of apolipoprotein B100-containing lipoproteins in dogs: evidence for disregarding cholesterol ester transfer. Eur J Clin Invest, 34, 527e534. Barter, P. J., Caulfield, M., Eriksson, M., et al., for the ILLUMINATE Investigators. (2007). Effects of torcetrapib in patients at high risk for coronary events. N Engl J Med, 357, 2109e2122. Barter, P. J., & Lally, J. I. (1978). The activity of an esterified cholesterol transferring factor in human and rat serum. Biochim Biophys Acta, 531, 233e236. Barzilai, N., Atzmon, G., Schechter, C., et al. (2003). Unique lipoprotein phenotype and genotype associated with exceptional longevity. J Am Med Assoc, 290, 2030e2040. Bisgaier, C. L., Siebenkas, M. V., Brown, M. L., et al. (1991). Familial cholesteryl ester transfer protein deficiency is associated with triglyceride-rich low density lipoproteins containing cholesteryl esters of probable intracellular origin. J Lipid Res, 32, 21e33. Bloomfield, D., Carlson, G. L., Sapre, A., et al. (2009). Efficacy and safety of the cholesteryl ester transfer protein inhibitor anacetrapib as monotherapy and coadministered with atorvastatin in dyslipidemic patients. Am Heart J, 157, 352e360. Borggreve, S. E., Hillege, H. L., Dallinga-Thie, G. M., et al., on behalf of the PREVEND Study Group. (2007). High plasma cholesteryl ester transfer protein levels may favour reduced incidence of cardiovascular events in men with low triglycerides. Eur Heart J, 28(8), 1012e1018. Brousseau, M. E., Diffenderfer, M. R., Millar, J. S., et al. (2005). Effects of cholesteryl ester transfer protein inhibition on high-density lipoprotein subspecies, apolipoprotein A-I metabolism, and fecal sterol excretion. Arterioscler Thromb Vasc Biol, 25, 1e8. Brousseau, M. E., Schaefer, E. J., Wolfe, M. L., et al. (2004). Effects of an inhibitor of cholesteryl ester transfer protein on HDL cholesterol. N Engl J Med, 350, 1505e1515. Brown, M. L., Inazu, A., Hesler, C. B., et al. (1989). Molecular basis of lipid transfer protein deficiency in a family with increased high-density lipoproteins. Nature, 342(6248), 448e451. Catalano, G., Julia, Z., Frisdal, E., et al. (2009). Torcetrapib differentially modulates the biological activities of HDL2 and HDL3 particles in the reverse cholesterol transport pathway. Arterioscler Thromb Vasc Biol, 29, 268e275. Cazita, P. M., Barbeiro, D. F., Moretti, A. I. S., Quitao, E. C. R., & Soriano, F. G. (2008). Human cholesteryl ester transfer protein expression enhances the mouse survival rate in an experimental systemic inflammation model: a novel role for CETP. Shock, 30, 590e595. Cheema, S. K., Agarwal-Mawal, A., Murray, C. M., & Tucher, S. (2005). Lack of stimulation of cholesteryl ester transfer protein by cholesterol in the presence of a high-fat diet. J Lipid Res, 46, 2356e2366.
Chapter j 3
CETP and Atherosclerosis
53
Chen, D. W., Yang, J. F., Tang, Z., et al. (2008). Cholesteryl ester transfer protein polymorphism D442G associated with a potential decreased risk for Alzheimer’s disease as a modifier for APOE 34 in Chinese. Brain Res, 1187, 52e57. Cheung, M. C., Wolfbauer, G., & Albers, J. J. (1996). Plasma phospholipids mass transfer rate: relationship to plasma phospholipid and cholesteryl ester transfer activities and lipid parameters. Biochim Biophys Acta, 1303, 103e110. Chiba, H., Akita, H., Tsuchihashi, K., et al. (1997). Quantitative and compositional changes in high density lipoprotein subclasses in patients with various genotypes of cholesteryl ester transfer protein deficiency. J Lipid Res, 38, 1204e1216. Chien, K. L., Sung, F. C., Hsu, H. C., Su, T. C., Lin, R. S., & Lee, Y. T. (2002). Apolipoprotein A-I and B and stroke events in a community-based cohort in Taiwan. Report of the Chin-Shan Community Cardiovascular Study. Stroke, 33, 39e44. Curb, J. D., Abbott, R. D., Rodriguez, B. L., et al. (2004). A prospective study of HDL-C and cholesteryl ester transfer protein gene mutations and the risk of coronary heart disease in the elderly. J Lipid Res, 45, 948e953. Dallinga-Thie, G. M., Dullaart, R. P. F., & van Tol, A. (2007). Concerted actions of cholesteryl ester transfer protein and phospholipid transfer protein in type 2 diabetes: effects of apolipoproteins. Curr Opin Lipidol, 18, 251e257. Davidson, W. S., Silva, R. A. G. D., Chantepie, S., Lago, W. R., Chapman, M. J., & Kontush, A. (2009). Proteomic analysis of defined HDL subpopulations reveals particle-specific protein clusters. Arterioscler Thromb Vasc Biol, 29, 870e876. Drayna, D., Jarnagin, A. S., McLean, J., et al. (1987). Cloning and sequencing of human cholesteryl ester transfer protein cDNA. Nature, 327, 632e634. Dullaart, R. P. F., de Vries, R., Scheek, L., et al. (2004). Type 2 diabetes mellitus is associated with differential effects on plasma cholesteryl ester transfer protein and phospholipid transfer protein activities and concentrations. Scand J Clin Lab Invest, 64, 205e216. Fournier, N., Paul, J. L., Atger, V., et al. (1997). HDL phospholipid content and composition as a major factor determining cholesterol efflux capacity from Fu5AH cells to human serum. Arterioscler Thromb Vasc Biol, 17, 2685e2691. Gautier, T., Klein, A., Deckert, V., et al. (2008). Effect of plasma phospholipid transfer protein deficiency on lethal endotoxemia in mice. J Biol Chem, 283, 18702e18710. Gofman, J. W., Young, W., & Tandy, R. (1966). Ischemic heart disease, atherosclerosis, and longevity. Circulation, 34, 679e697. Gotoda, T., Kinoshita, M., Ishibashi, S., et al. (1997). Skipping of exon 14 and possible instability of both the mRNA and the resultant truncated protein underlie a common cholesteryl ester transfer protein deficicency in Japan. Arterioscler Thromb Vasc Biol., 17, 1376e1381. Guerin, M., Bruckert, E., Dolphin, P. J., Turpin, G., & Chapman, M. J. (1996). Fenofibrate reduces plasma cholesteryl ester transfer from HDL to VLDL and normalizes the atherogenic, dense LDL profile in combined hyperlipidemia. Arterioscler Thromb Vasc Biol., 16, 763e772. Ha, Y. C., & Barter, P. J. (1982). Differences in plasma cholesteryl ester transfer activity in sixteen vertebrate species. Comp Biochem Physiol, 71B, 265e269. Haraki, T., Inazu, A., Yagi, K., Kajinami, K., Koizumi, J., & Mabuchi, H. (1997). Clinical characteristics of double heterozygotes with familial hypercholesterolemia and cholesteryl ester transfer protein deficiency. Atherosclerosis, 132, 229e236. Hayashi, H., Campenot, R. B., Vance, D. E., & Vance, J. E. (2007). Apolipoprotein E-containing lipoproteins protect neurons from apoptosis via a signaling pathway involving low-density lipoprotein receptor-related protein-1. J Neurosci, 27, 1933e1941.
54
The HDL Handbook
Hernandez, M., Wright, S. D., & Cai, T. Q. (2007). Critical role of cholesterol ester transfer protein in nicotinic acid-mediated HDL elevation in mice. Biochem Biophys Res Commun, 355, 1075e1080. Hudgins, L. C., Parker, T. S., Levine, D. M., et al. (2003). A single intravenous dose of endotoxin rapidly alters serum lipoproteins and lipid transfer proteins in normal volunteers. J Lipid Res, 44, 1489e1498. Hussain, M. M., Innerarity, T. L., Brecht, W. J., & Mahley, R. W. (1995). Chylomicron metabolism in normal, cholesterol-fed, and Watanabe heritable hyperlipidemic rabbits. J Biol Chem, 270, 8578e8587. Ikewaki, K., Nishiwaki, M., Sakamoto, T., et al. (1995). Increased catabolic rate of low density lipoproteins in humans with cholesteryl ester transfer protein deficiency. J Clin Invest, 96, 1573e1581. Ikewaki, K., Rader, D. J., Sakamoto, T., et al. (1993). Delayed catabolism of high density lipoprotein apolipoprotein A-I and A-II in human cholesteryl ester transfer protein deficiency. J Clin Invest, 92, 1650e1658. Inazu, A., Brown, M. L., Hesler, C. B., et al. (1990). Increased high-density lipoprotein levels caused by a common cholesteryl-ester transfer protein gene mutation. N Engl J Med, 323(18), 1234e1238. Inazu, A., Jiang, X. C., Haraki, T., et al. (1994). Genetic cholesteryl ester transfer protein deficiency caused by two prevalent mutations as a major determinant of increased levels of high density lipoprotein cholesterol. J Clin Invest, 94(5), 1872e1882. Inazu, A., Koizumi, J., Kajinami, K., Kiyohara, T., Chichibu, K., & Mabuchi, H. (1999). Opposite effects on serum cholesteryl ester transfer protein levels between long-term treatments with pravastatin and probucol in patients with primary hypercholesterolemia and xanthoma. Atherosclerosis, 145(2), 405e413. Inazu, A., Nakajima, K., Nakano, T., et al. (2008). Decreased post-prandial triglyceride response and diminished remnant lipoprotein formation in cholesteryl ester transfer protein (CETP) deficiency. Atherosclerosis, 196, 953e957. Inazu, A., Quinet, E. M., Wang, S., et al. (1992). Alternative splicing of the mRNA encoding the human cholesteryl ester transfer protein. Biochemistry, 31(8), 2352e2358. Jansen, H., Verhoeven, A. J. M., & Sijbrands, E. J. G. (2002). Hepatic lipase: a pro- or antiatherogenic protein? J Lipid Res, 43, 1352e1362. Jeon, S.-M., Park, Y. B., Kwon, O. S., et al. (2005). Vitamin E supplementation alters HDLcholesterol and paranoxase activity in rabbits fed high cholesterol diet: comparison with probucol. J. Biochem Molecular Taxicology, 19, 336e346. Jeppesen, J., Hein, H. O., Suadicani, P., & Gyntelberg, F. (1998). Triglyceride concentration and ischemic heart disease. An eight-year follow-up in the Copenhagen Male Study. Circulation, 97, 1029e1036. Kastelein, J. J. P., van Leuven, S. I., Burgess, L., et al., for the RADIANCE 1 Investigators. (2007). Effects of torcetrapib on carotid atherosclerosis in familial hypercholesterolemia. N Engl J Med, 356, 1620e1630. Kee, P., Caiazza, D., Rye, K. A., Barrett, P. H. R., Morehouse, L. A., & Barter, P. J. (2006). Effects of inhibiting cholesteryl ester transfer protein on the kinetics of high-density lipoprotein cholesteryl ester transport in plasma. In vivo studies in rabbits. Arterioscler Thromb Vasc Biol, 26, 884e890. Keys, A. (1983). Lipoprotein profile e its value in prediction. Prev Med, 12, 25e31. Kiyohara, T., Kiriyama, R., Zamma, S., Inazu, A., Koizumi, J., Mabuchi, H., & Chichibu, K. (1998). Enzyme immunoassay for cholesteryl ester transfer protein in human serum. Clin Chim Acta, 271(2), 109e118.
Chapter j 3
CETP and Atherosclerosis
55
Koizumi, J., Inazu, A., Yagi, K., et al. (1991). Serum lipoprotein lipid concentration and composition in homozygous and heterozygous patients with cholesteryl ester transfer protein deficiency. Atherosclerosis, 90(2-3), 189e196. Koizumi, J., Mabuchi, H., Yoshimura, A., et al. (1985). Deficiency of serum cholesteryl-ester transfer activity in patients with familial hyperalphalipoproteinaemia. Atherosclerosis, 58, 175e186. Kontush, A., de Faria, E. C., Chantepie, S., & Chapman, M. J. (2004). Antioxidative activity of HDL particle subspecies is impaired in hyperalphalipoproeinemia: Relevance of enzymatic and physicochemical properties. Arterioscler Thromb Vasc Biol, 24, 526e533. Kontush, A., Therond, P., Zerrad, A., et al. (2007). Preferential sphingosine-1-phosphate enrichment and sphingomyelin depletion are key features of small dense HDL3 particles. Arterioscler Thromb Vasc Biol, 27, 1843e1849. Koropatnick, T. A., Kimbell, J., Chen, R., et al. (2008). A prospective study of high-density lipoprotein cholesterol, cholesteryl ester transfer protein gene variants, and healthy aging in very old Japanese-American men. J Gerontol, 63A, 1235e1240. Krimbou, L., Marcil, M., Chiba, H., & Genest, J., Jr. (2003). Structural and functional properties of human plasma high density-sized lipoprotein containing only apoE particles. J Lipid Res, 44, 884e892. Kurth, T., Everett, B. M., Buring, J. E., Kase, C. S., Ridker, P. M., & Gaziano, J. M. (2007). Lipid levels and the risk of ischemic stroke in women. Neurology, 68, 556e562. Lamarche, B., Uffelman, K. D., Carpentier, A., et al. (1999). Triglyceride enrichment of HDL enhances in vivo metabolic clearance of HDL apoA-I in healthy men. J Clin Invest, 103, 1191e1199. Lee, J. Y., Badeau, R. M., Mulya, A., et al. (2007). Functional LCAT deficiency in human apolipoprotein A-I transgenic, SR-BI knockout mice. J Lipid Res, 48, 1052e1061. Leppala, J. M., Paunio, M., Virtamo, J., et al. (1999). Alcohol consumption and stroke incidence in male smokers. Circulation, 100, 1209e1214. Levels, J. H. M., Pajkrt, D., Schultz, M., et al. (2007). Alterations in lipoprotein homeostasis during human experimental endotoxemia and clinical sepsis. Biochim Biophys Acta, 1771, 1429e1438. Li, L., Thompson, P. A., & Kitchens, R. L. (2008). Infection induces a positive acute phase apolipoprotein E response from a negative acute phase gene: role of hepatic LDL receptors. J Lipid Res, 49, 1782e1793. Lu, H., Inazu, A., Moriyama, Y., et al. (2003). Haplotype analyses of cholesteryl ester transfer protein gene promoter: a clue to an unsolved mystery of TaqIB polymorphism. J Mol Med, 81 (4), 246e255. Luo, Y., & Tall, A. R. (2000). Sterol upregulation of human CETP expression in vitro and in transgenic mice by an LXR element. J Clin Invest, 105, 513e520. MacLean, P. S., Bower, J. F., Vadlamudi, S., et al. (2003). Cholesteryl ester transfer protein expression prevents diet-induced atherosclerosis lesions in male db/db mice. Arterioscler Thromb Vasc Biol, 23, 1412e1415. MacLean, P. S., Vadlamudi, S., MacDonald, K. G., Pories, W. J., & Barakat, H. A. (2005). Suppression of hepatic cholesteryl ester transfer protein expression in obese humans with the development of type 2 diabetes mellitus. J Clin Endocrinol Metab, 90, 2250e2258. Mahley, R. W., Innerarity, T. L., Weisgraber, K. H., & Fry, D. L. (1977). Canine hyperlipoproteinemia and atherosclerosis. Am J Pathol, 87, 205e226. Manabe, M., Abe, T., Nozawa, M., Maki, A., Hirata, M., & Itakura, H. (1987). New substrate for determination of serum lecithin:cholesterol acyltransferase. J Lipid Res, 28, 1206e1215. Masson, D., Staels, B., Gautier, T., et al. (2004). Cholesteryl ester transfer protein modulates the effect of liver X receptor agonists on cholesterol transport and excretion in the mouse. J Lipid Res, 45, 543e550.
56
The HDL Handbook
Millar, J. S., Brousseau, M. E., Diffenderfer, M. R., et al. (2006). Effects of cholesteryl ester transfer protein inhibitor torcetrapib on apolipoprotein B100 metabolism in humans. Arterioscler Thromb Vasc Biol, 26, 1350e1356. Miwa, K., Inazu, A., Kawashiri, M., et al. (2009). Cholesterol efflux from J774 macrophages and Fu5AH hepatoma cells to serum is preserved in CETP-deficient patients. Clin Chim Acta, 402, 19e24. Miyazaki, O., Fukamachi, I., Mori, A., et al. (2009). Formation of prebeta1-HDL during lipolysis of triglyceride-rich lipoprotein. Biochem Biophys Res Commun, 379, 55e59. Moriyama, Y., Okamura, T., Inazu, A., et al. (1998). A low prevalence of coronary heart disease in subjects with increased high-density lipoprotein cholesterol levels including those with plasma cholesteryl ester transfer protein deficiency. Prev Med, 27(5Pt1), 659e667. Musliner, T. A., Long, M. D., Forte, T. M., & Krauss, R. M. (1991). Size transformations of intermediate and low density lipoproteins induced by unesterified fatty acids. J Lipid Res, 32, 903e915. Nagano, M., Nakamura, M., Kobayashi, N., Kamata, J., & Hiramori, K. (2005). Effort angina in a middle-aged woman with abnormally high levels of serum high-density lipoprotein cholesterol. Circ J, 69, 609e612. Nagano, M., Yamashita, S., Hirano, K. I., et al. (2004). Molecular mechanisms of cholesteryl ester transfer protein deficiency in Japanese. J Atheroscler Thromb, 11, 110e121. Nagasaki, T., & Akanuma, Y. (1977). A new colorimetric method for the determination of plasma lecithin-cholesterol acyltransferase activity. Clin Chim Acta, 75, 371e375. Nissen, S. E., Tardif, J. C., Nicholls, S. J., et al., for the ILLUSTRATE Investigators. (2007). Effects of torcetrapib on the progression of coronary atherosclerosis. N Engl J Med, 356, 1304e1316. Okamura-Noji, K., Sasai, K., Zhan, R., et al. (2001). Cholesteryl ester transfer protein deficiency causes slow egg embryonation of Schistosoma japonicum. Biochim Biophys Res Commun, 286, 305e310. Oliveira, H. C. F., Ma, L., Milne, R., et al. (1997). Cholesteryl ester transfer protein (CETP) activity enhances plasma cholesteryl ester formation: studies in CETP transgenic mice and human genetic CETP deficiency. Arterioscler Thromb Vasc Biol, 17(6), 1045e1052. Ordovas, J. M., Cupples, L. A., Corella, D., et al. (2000). Association of cholesteryl ester transfer protein-TaqIB polymorphism with variations in lipoprotein subclasses and coronary heart disease risk: the Framingham study. Arterioscler Thromb Vasc Biol., 20(5), 1323e1329. Ouguerram, K., Nguyen, P., Krempf, M., et al. (2004). Selective uptake of high density lipoproteins cholesteryl ester in the dog, a species lacking in cholesteryl ester transfer protein activity. An in vivo approach using stable isotopes. Comparative Biochem Physiol, Part B, 138, 339e345. Pattnaik, N. M., Montes, A., Hughes, L. B., & Zilversmit, D. B. (1978). Cholesteryl ester exchange protein in human plasma isolation and characterization. Biochim Biophys Acta, 530, 428e438. Pruneta, V., Pulcini, T., Lalanne, F., et al. (1999). VLDL-bound lipoprotein lipase facilitates the cholesteryl ester transfer protein-mediated transfer of cholesteryl esters from HDL to VLDL. J Lipid Res, 40, 2333e2339. Qiu, X., Mistry, A., Ammirati, M. J., et al. (2007). Crystal structure of cholesteryl ester transfer protein reveals a long tunnel and four bound lipid molecules. Nature Struct Molec Biol, 14(2), 106e113. Regieli, J. J., Jukema, J. W., Grobbee, D. E., et al. (2008). CETP genotype predicts increased mortality in statin-treated men with proven cardiovascular disease: an adverse pharmacogenetic interaction. Eur Heart J, 29, 2792e2799.
Chapter j 3
CETP and Atherosclerosis
57
Rhyne, J., Ryan, M. J., White, C., Chimonas, T., & Miller, M. (2006). The two novel CETP mutations Gln87X and Gln165X in a compound heterozygous state are associated with marked hyperalphalipoproteinemia and absence of significant coronary artery disease. J Mol Med, 84(8), 647e650. Ritsch, A., Scharnagl, H., Eller, P., et al. (2010). Cholesteryl ester transfer protein and mortality in patients undergoing coronary angiography. The Ludwigshafen Risk and Cardiovascular Health Study. Circulation, 121, 366e374. Robins, S. J., & Fasulo, J. M. (1997). High density lipoproteins, but not other lipoproteins, provide a vehicle for sterol transport to bile. J Clin Invest, 99, 380e384. Sakai, N., Matsuzawa, Y., Hirano, K., et al. (1991). Detection of two species of low density lipoprotein particles in cholesteryl ester transfer protein deficiency. Arterioscler Thromb, 11, 71e79. Sandhofer, A., Kaser, S., Ritsch, A., et al. (2006). Cholesteryl ester transfer protein in metabolic syndrome. Obesity, 14, 812e818. Scanu, A. M., & Edelstein, C. (2008). HDL: bridging past and present with a look at the future. FASEB J, 22, 4044e4054. Schaefer, E. J., Lamon-Fava, S., Ordovas, J. M., et al. (1994). Factors associated with low and elevated plasma high density lipoprotein cholesterol and apolipoprotein A-I levels in the Framingham Offspring Study. J Lipid Res, 35, 871e882. Schmitz, G., & Assmann, G. (1982). Isolation of human serum HDL1 by zonal ultracentrifugation. J Lipid Res, 23, 903e910. Schwartz, C. C., VandenBroek, J. M., & Cooper, P. S. (2004). Lipoprotein cholesteryl ester production, transfer, and output in vivo in humans. J Lipid Res, 45, 1594e1607. Scobey, M. W., Johnson, F. L., & Rudel, L. L. (1989). Delivery of high-density lipoprotein free and esterified cholesterol to bile by the perfused monkey liver. Am J Physiol, 257(4 Pt 1), G644e52. Shetty, S., Eckhardt, E. R. M., Post, S. R., & van der Westhuyzen, D. R. (2006). Phosphatidylinositol-3-kinase regulates scavenger receptor class B type I subcellular localization and selective lipid uptake in hepatocytes. Arterioscler Thromb Vasc Biol, 26, 2125e2131. Stevenson, S., Wang, S., Deng, L., & Tall, A. R. (1993). Human plasma cholesteryl ester transfer protein consists of a mixture of two forms reflecting variable glycosylation at Asparagine341. Biochemistry, 32, 5121e5126. Sviridov, D., Chin-Dusting, J., Nestel, P., et al. (2006). Elevated HDL cholesterol is functionally ineffective in cardiac transplant recipients: Evidence for impaired reverse cholesterol transport. Transplantation, 81, 361e366. Takata, M., Inazu, A., Katsuda, S., et al. (2006). CETP (cholesteryl ester transfer protein) promoter -1337 C>T polymorphism protects against coronary atherosclerosis in Japanese patients with heterozygous familial hypercholesterolaemia. Clin Sci (Lond), 111(5), 325e331. Tanigawa, H., Billheimer, J. T., Tohyama, J., Zhang, Y., Rothblat, G., & Rader, D. J. (2007). Expression of cholesteryl ester transfer protein in mice promotes macrophage reverse cholesterol transport. Circulation, 116(11), 1267e1273. Tchoua, U., D’Souza, W., Mukhamedova, N., et al. (2008). The effect of cholesteryl ester transfer protein overexpression and inhibition on reverse cholesterol transport. Cardiovasc Res, 77, 732e739. Teh, E. M., Dolphin, P. J., Breckenridge, W. C., & Tan, M. H. (1998). Human plasma CETP deficiency: identification of a novel mutation in exon 9 of the CETP gene in a Caucasian subject from North America. J Lipid Res, 39, 442e456.
58
The HDL Handbook
Thompson, A., Angelantonio, E. D., Sarwar, N., et al. (2008). Association of cholesteryl ester transfer protein genotypes with CETP mass and activity, lipid levels, and coronary risk. J Am Med Assoc, 299, 2777e2788. Thompson, J. F., Reynolds, J. M., Williams, S. P., Wood, L. S., Paciga, S. A., & Lloyd, D. B. (2009). Frequency and function of CETP variants among individuals of Asian ancestry. Atherosclerosis, 202, 241e247. Trigatti, B., Rayburn, H., Vinals, M., et al. (1999). Influence of the high density lipoprotein receptor SR-BI on reproductive and cardiovascular pathophysiology. Proc Natl Acad Sci USA, 96, 9322e9327. Vaisar, T., Pennathur, S., Green, P. S., et al. (2007). Shotgun proteomics implicates protease inhibition and complement activation in the anti-inflammatory properties of HDL. J Clin Invest, 117, 746e756. Van der Hoorn, J. W. A., de Haan, W., Berbee, J. F. P., et al. (2008). Niacin increases HDL by reducing hepatic expression and plasma levels of cholesteryl ester transfer protein in APOE*3Leiden.CETP mice. Arterioscler Thromb Vasc Biol, 28, 2016e2022. van der Steeg, W. A., Holme, I., Boekholdt, S. M., et al. (2008). High-density lipoprotein cholesterol, high-density lipoprotein particle size, and apolipoprotein A-I: Significance for cardiovascular risk. J Am Med Assoc, 51, 634e642. Van Eck, M., Hoekstra, M., Hildebrand, R. B., et al. (2007). Increased oxidative stress in scavenger receptor BI knockout mice with dysfunctional HDL. Arterioscler Thromb Vasc Biol, 27, 2413e2419. Vasan, R. S., Pencina., M. J., Robins, S. J., et al. (2009). Association of circulating cholesteryl ester transfer protein activity with incidence of cardiovascular disease in the community. Circulation, 120, 2414e2420. Vreugdenhil, A. C. E., Snoek, A. M. P., van’t Veer, C., Greve, J. W. M., & Buurman, W. A. (2001). LPS-binding protein circulates in association with apoB-containing lipoproteins and enhances endotoxin-LDL/VLDL interaction. J Clin Invest, 107, 225e234. Wang, S., Wang, X., Deng, L., Rassart, E., Milne, R. W., & Tall, A. R. (1993). Point mutagenesis of carboxyl-terminal amino acids of cholesteryl ester transfer protein. J Biol Chem, 268, 1955e1959. Wang, X., Driscoll, D. M., & Morton, R. E. (1999). Molecular cloning and expression of lipid transfer inhibitor protein reveals its identity with apolipoprotein F. J Biol Chem, 274, 1814e1820. Watts, G. F., Ji, J., Chan, D. C., et al. (2006). Relationships between changes in plasma lipid transfer proteins and apolipoprotein B-100 kinetics during fenofibrate treatment in the metabolic syndrome. Clin Sci, 111, 193e199. Williams, P. T. (1996). High-density lipoprotein cholesterol and other risk factors for coronary heart disease in female runners. N Engl J Med, 334, 1298e1303. Wilson, H. M., Patel, J. C., Russell, D., & Skinner, E. R. (1993). Alterations in the concentration of an apolipoprotein E-containing subfraction of plasma high density lipoprotein in coronary heart disease. Clin Chim Acta, 220, 175e187. Ye, D., Kraaijeveld, A. O., Grauss, R. W., et al. (2008). Reduced leucocyte cholesteryl ester transfer protein expression in acute coronary syndromes. J Intern Med, 246, 571e585. Yvan-Charvet, L., Matsuura, F., Wang, N., et al. (2007). Inhibition of cholesteryl ester transfer protein by torcetrapib modestly increases macrophage cholesterol efflux to HDL. Arterioscler Thromb Vasc Biol, 27, 1132e1138. Yvan-Charvet, L., Pagler, T. A., Wang, N., et al. (2008). SR-BI inhibits ABCG1-stimulated net cholesterol efflux from cells to plasma HDL. J Lipid Res, 49, 107e114.
Chapter j 3
CETP and Atherosclerosis
59
Zhang, A., Gao, S., Fan, J., Huang, W., Zhao, T., & Liu, G. (2004). Increased plasma HDL cholesterol levels and biliary cholesterol excretion in hamsters by LCAT overexpression. FEBS Letters, 570, 25e29. Zhang, X., Moor, A., Merkler, K., Liu, Q., & McLean, M. (2007). Regulation of alternative splicing of liver scavenger receptor class B gene by estrogen and the involved regulatory splicing factors. Endcrinology, 148, 5295e5304.
Chapter 4
HDL and Reverse Cholesterol Transport: Physiological Modulation Giovanna Catalano and Maryse Guerin INSERM UMRS939, Hoˆpital de la Pitie´, Paris, France; UPMC Universite´ Pierre et Marie Curie, Hoˆpital de la Pitie´, Paris, France
INFLUENCE OF AGING ON THE REVERSE CHOLESTEROL TRANSPORT PATHWAY The incidence of cardiovascular disease (CVD) due to atherosclerosis increases with aging (Grundy, 1995). Aging is characterized by the occurrence of several physical and biochemical modifications that may affect high density lipoprotein (HDL) structure and functions (Walter, 2009). Analysis of the effects of aging on the reverse cholesterol transport (RCT) process revealed that HDL3 from elderly subjects displayed a reduced capacity to mediate free cholesterol efflux via the ATP-binding cassette A1 (ABCA-1) pathway from THP-1 cells as compared to HDL3 from younger subjects, as a result of decreased HDL bilayer fluidity, increased apolipoprotein A (apoA-I) oxidation, decreased phosphatidylcholine to sphingomyeline ratio (Berrougui, 2007). In elderly subjects, endogenous vitamin E levels in native HDL particles were reduced as compared with those in HDL isolated from young and middle aged subjects (Khalil, 1998). This reduced vitamin E content in HDL could be responsible for an increased susceptibility of these particles to oxidation. HDL antioxidant activity decreases with aging as a consequence of decreased paraoxonase 1 protection effect towards low density lipopreotein (LDL) peroxidation (Seres et al., 2004; Jaouad et al., 2006).
INFLUENCE OF HORMONES ON THE REVERSE CHOLESTEROL TRANSPORT PATHWAY That atherosclerosis development and reverse cholesterol transport (RCT) are related to gender is evident by the distinct incidence of coronary death between men and women. The epidemiology of coronary heart disease demonstrates The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
61
62
The HDL Handbook
a striking sex difference in its age-specific incidence and severity even independently of environmental risk factors. If atherogenesis is one of the consequences of altered RCT, then sex difference in RCT could explain the different predisposition of men and women to develop cardiovascular disease. The net effect of hormones on protection from atherosclerosis is difficult to assess since these effects depend on whether associations with endogenous or exogenous hormones have been investigated. Much of the evidence for sex hormone effects on lipid metabolism comes from studies of changes associated with administration of synthetic gonadal steroids or with changes in gonadal function. Studies on differences in lipoprotein metabolism in normal men and women are extremely limited.
Estrogens-mediated effect Numerous epidemiological studies suggest that estrogens protect women against cardiovascular disease before the age of menopause. After menopause, CVD risk for women becomes progressively closer to that of men, suggesting an atheroprotective role for estrogens (Arnal et al., 2004). There are three main naturally occurring estrogens: 17b estradiol (E2), estrone and estriol. E2 is mainly produced by the ovary whereas estrone and estriol are produced in the liver from estradiol (Gruber et al., 2002). Estriol is also produced by the placenta during pregnancy. Women usually display higher HDL cholesterol (HDL-C) levels than men (Schaefer et al., 1982). Prepubertal boys and girls do not differ significantly in their serum lipid levels and lipoprotein profile. In contrast with girls, in whom HDL-C levels do not change with puberty despite the rise in estrogen concentrations, sexually maturing boys display a decrease in HDL-C and an increase in both LDL-C and triglyceride (TG) levels (Godsland et al., 1987) which occur concomitantly with an elevation in testosterone levels. Evidence of how sex hormones influence lipid metabolism is given by studies in pregnant women throughout the gestational period (Sattar et al., 1997). Estradiol elevation during pregnancy favors an increase in VLDL particles, especially the VLDL2 subpopulation, a decreased hepatic lipase (HL) activity and variations in plasma cholesteryl ester transfer protein (CETP) activity (Iglesias et al., 1994). During pregnancy LDL-C, HDL-C and TG rise progressively. LDL-C remains elevated until after delivery, but TGs fall to baseline at this time (LaRosa, 1992). CETP activity has been shown to increase significantly during the second trimester of pregnancy before declining just before delivery (Silliman et al., 1993). In vivo studies conducted in intact female rats demonstrated that physiological doses of E2 increased apoA-I and HDL-C levels (Parini et al., 2000). An increase in plasma HDL-C and apoA-I levels in women as compared to men has been also described (Patsch et al., 1980). Equally, in vitro studies demonstrated that secretion of apoA-I-containing particles from HepG2 cells was stimulated by estradiol as a result of increased transcription rate of apoA-I mRNA
Chapter j 4
HDL and Reverse Cholesterol Transport
63
(Jin et al., 1998). These observations are consistent with in vivo studies demonstrating that increased HDL levels in postmenopausal women under estrogen therapy result primarily from an increase of HDL production rate rather than from a decrease of HDL catabolism (Schaefer et al., 1982; Kauma et al., 1996). Concerning CETP activity, a clear sex difference was observed in the genotype effect on HDL-C level. HDL-C levels were strongly associated with TaqIB polymorphism (Kauma et al., 1996) in women and weakly in postmenopausal women without hormone replacement therapy. Attenuation of the gene dosage effect after menopause suggests that the gender difference in HDL-C may be due in part to sex hormones. Furthermore, a study conducted in hCETP transgenic rats suggests that sex steroids potentiate the regulation of CETP expression even at low plasma levels of estrogen (Vadlamudi et al., 1998). Estrogen therapy in atherosclerotic mice has consistently demonstrated a dose dependent inhibitory effect of E2 on lesion initiation and progression in ovariectomized females. When these mice were treated with E2 in a physiological range, atheroma plaques rarely progressed beyond uncomplicated fatty streaks. In the apoE / mouse, castration is followed by an increase of lesion area and E2 administration prevents fatty streak deposition but only at concentrations similar to those in the gestational period. However, owing to high doses of estrogens used in animal studies, it is unclear whether estrogen effects are physiologically relevant. The different predisposition between men and premenopausal women in developing atherosclerosis, suggests hormonal replacement therapy as a tool for prevention in postmenopausal women. Treatment of rats with pharmacological doses of E2 increased hepatic LDL-R protein and mRNA (Parini et al., 1997). In fasting women, LDL-R expression was positively correlated with plasma estrone concentration (Nanjee et al., 1990). In rat liver, SR-BI expression is suppressed by estrogen administration as well as in cultured human hepatic cells (Landschulz et al., 1996; Fluiter et al., 1998; Graf et al., 2001). Recently, it has been demonstrated that estrogen regulation of SR-BI expression is due to differential modulation of SR-BI RNA splicing and glycosylation, a process that is important in folding and export of SR-BI to the membrane (Zhang et al., 2007). By contrast with results obtained in hepatic cells, estrogen treatment increases SR-BI expression and selective lipid uptake in adrenal and ovarian luteal cells as well as in macrophages (Fluiter et al., 1998). SR-BII is an alternatively spliced product of the SR-BI gene that differs only in the C-terminal cytoplasmic domain. Concomitant with SR-BI downregulation, estrogen induces an increase in SR-BII expression. Both non-ovariectomized female rats and ovariectomized estrogen-treated female rats present an increased expression of SR-BII with a nearly irrelevant expression of SR-BI in the liver, while ovariectomized female and male rats showed an opposite profile of expression (Graf et al., 2001). Hepatic lipase (HL) is also subject to hormonal regulation. Its activity is decreased after the peak of estrogen in the reproductive cycle (Tikkanen et al., 1986) and women have lower HL than men (Bersot et al., 1999)
64
The HDL Handbook
which may account for the higher HDL-C levels in women. Estrogen-mediated inhibition of HL has been recently demonstrated to occur via repression of the HL gene promoter (Jones et al., 2002).
Androgens-mediated effect Testosteron (T) can exert its metabolic effects either directly or through its metabolites E2 and dihydro-testosterone. The effect of T and E2, in particular, can be either additive or counter regulatory. T is clearly associated with a lowering effect on HDL-C. Zitzmann et al. (2001) found significant correlation between HDL-C levels and the number of CAG repeat polymorphisms in the androgen receptor gene, which determines the sensitivity of target cells towards testosterone. T upregulates SR-BI and HL expression in macrophages, even if upregulation of HL in HepG2 cells was observed only with supraphysiological concentration of T. Macrophages with increased testosterone-induced SR-BI expression, showed enhanced cholesterol efflux capacity that could reflect facilitated RCT in vivo (Langer et al., 2002). Castration or T suppression in men with prostate cancer was found to increase HDL-C by about 20%. The effect of T on HDL metabolism is marked on the large HDL subclass but the mechanism by which T regulates HDL metabolism is not well understood.
Thyroid-mediated effect Thyroid hormones affect serum cholesterol by altering lipid metabolism. Hypothyroidism frequently causes hyperlipidemia with increased TC, TG, LDL-C and apoB levels. On the contrary, hyperthyroidism is responsible for low levels of TC and LDL-C and increased HDL-C levels. Changes in LDL-C are mainly due to altered clearance of LDL as a result of the presence of a thyroid hormone responsive element in the LDL-r gene promoter. The effects on HDL-C level are mainly due to modification of CETP and HL activities, that decrease with hypothyroidism and increase with hyperthyroidism (Kuusi et al., 1988). In hyperthyroid states, HDL decrease is mainly observed in the HDL2b subfraction (Muls et al., 1982). Changes in apoA-I in hyper- and hypothyroidism are mainly caused by changes in the concentration of LpAI rather than LpAI:AII particles since apoA-II levels remain unaltered. HL appears to be the main factor influencing HDL concentrations in these patients but it is not clear if modulation of HDL levels results from variations in HDL synthesis or catabolism. ApoA-I mRNA synthesis is reduced in hypothyroidism and apoA-I transcription is stimulated by thyroid hormone in the rat (Staels et al., 1990).
Insulin-mediated effect It is clear that a strong link exists between perturbations in insulin signal transduction and alterations in lipid and lipoprotein metabolism. Insulin and
Chapter j 4
HDL and Reverse Cholesterol Transport
65
glucose are known to directly regulate lipid synthesis and secretion (Kok et al., 1996). Insulin controls, at the transcriptional level, hepatic sterol regulatory element binding protein (SREBP) expression which is a key transcription factor responsible for regulating fatty acid and cholesterol biosynthesis. SREBP can activate a cascade of enzymes involved in cholesterol biosynthetic pathways, such as HMG-CoA reductase and fatty acid synthase (Bennett et al., 1995; Brown and Goldstein, 1997). SREBP binds to sterol response elements (SRE) in the promoter/enhancer regions of target genes. Expression of SREBP is enhanced by insulin in liver, adipose tissue and skeletal muscle. Under normal conditions, insulin stimulates cellular uptake of glucose, as well as free fatty acids (FFA) esterification and storage. Equally, insulin suppresses the hormone sensitive lipase, the major regulator of FFA release from adipose tissue. At the same time, insulin and glucose are believed to stimulate LPL activity from adipose tissue, thus regulating the first step of FFA delivery to this tissue for storage and TG plasma removal (Goldberg and Merkel, 2001). In the insulin resistant state, there is an increased release of FFA from adipose tissue concomitant with a decrease in TG-derived FFA uptake by muscle tissue. Since only visceral fat is linked to the liver through the portal vein, this may result in an increased influx of FFA into the liver (Arner, 2002). This may result in a vicious cycle which results in typical dyslipidemia of insulin resistance characterized by increased TG level, decreased HDL and compositional changes in LDL. The elevated FFA and glucose levels regulate the VLDL output from the liver and elevated TG in the liver inhibit apoB degradation and result in an increased assembly and secretion of VLDL particles. Furthermore, LPL levels and VLDL clearance are decreased (Howard, 1999; Avramoglu et al., 2006). Possible mechanisms of decreased HDL in insulin resistance include impaired VLDL lipolysis, which impairs HDL maturation by decreased transfer of TGRL remnants, increased activity of HL, alterations in hepatic function, which inhibit production of apoA-I and secretion of nascent HDL, enhanced CETP activity as a result of increased CE acceptors, i.e., VLDL (Howard, 1999).
INFLUENCE OF PHYSICAL ACTIVITY ON THE REVERSE CHOLESTEROL TRANSPORT PATHWAY Physical activity favorably affects a number of cardiovascular disease risk factors such as obesity, insulin resistance and blood pressure and it also has a beneficial effect on lipoprotein profile. The major effect of physical exercise is on HDL particle number and HDL cholesterol content. The exercise-induced changes in HDL-C level are the result of the interaction among exercise intensity, frequency, duration and length of the exercise period and could also be gender specific. Athletes had increased HDL-C and apoA-I concentrations, increased LCAT activity and preb-HDL levels as compared to sedentary controls (Olchawa et al., 2004). Furthermore, cholesterol efflux from
66
The HDL Handbook
macrophages to total plasma was increased in athletes as a result of increased HDL particle number. Fitness increases HDL-C but this increase is not correlated to very high level physical activity, thus indicating an upper threshold value of the HDL-C increase with physical activity. HDL-C was increased, mainly in the HDL2 subfraction, in a group of soccer players as compared to sedentary controls (Brites et al., 2004). However, no relevant differences were observed for all the other HDL-related parameters, thus suggesting that increased HDL-C was due to enrichment of the particles in cholesterol and not to increased particle number. Increase in HDL level could be due to either increased synthesis or catabolism. It is not clear whether increased HDL in fit subjects is due to one or the other reason. Interestingly, ABCA1 expression in humans is associated with physical activity. Mechanisms underlying the effect of exercise on HDL are multiple. Free cholesterol and PL are transferred to HDL during the catabolism of TGRL by LPL. Furthermore, muscles are TG depleted during exercise and it is possible that, in this situation, cholesterol is mobilized by muscle cells to its primary acceptor, HDL. As demonstrated by Sviridov et al. (2003), preb-HDL could be formed extrahepatically at the level of the muscle after exercise, thus contributing to the RCT process.
INFLUENCE OF DIET ON THE REVERSE CHOLESTEROL TRANSPORT PATHWAY The protective role of the consumption of monounsaturated fatty acids (MUFA) or polyunsaturated fatty acids (PUFA) against the development of atherosclerosis has been known for many years (Psota et al., 2006). Indeed, epidemiologic studies show that the Mediterranean diet, in which olive oil (enriched in oleic acid) predominates, is associated with a reduction in cardiovascular pathology. n-6 PUFA (linoleic and arachidonic acids) decrease plasma LDL-cholesterol concentrations while saturated fatty acids increase it. Moreover, n-3 PUFA (alpha-linolenic acid), reduce triglyceridemia and have anti-thrombotic and anti-atherogenic properties. Recently, a study conducted in normolipidemic subjects without cardiovascular disease showed that consumption of PUFA improves the anti-inflammatory properties of HDL particles while consumption of SFA reduces their anti-inflammatory activity and degrades endothelial function (Nicholls et al., 2006). These studies suggest that the functionality of HDL particles in patients presenting a high cardiovascular risk can be modulated by fatty acid consumption. Dietary fats induce physicochemical changes in HDL particles, which can modulate cellular cholesterol efflux. Contrastingly, when fat intake is limited to the NCEP (National cholesterol education program) (25e30% of total calories intake in the form of fat) and when the proportion of MUFA in the diet is fixed, differential percentages of PUFA, SFA or trans fatty acids (TFA) do not influence the capacity of HDL2 or HDL3 to promote cholesterol efflux capacity (Buonacorso et al., 2007). In monkeys,
Chapter j 4
HDL and Reverse Cholesterol Transport
67
dietary fatty acids modified HDL-PL content without altering cholesterol efflux capacity from hepatoma cells (Gillotte et al., 1998). Consumption of MUFA increases the capacity of whole plasma from normolipidemic subjects to promote cellular free cholesterol efflux via the SR-BI pathway as a result of an enrichment in phospholipids (Sakr et al., 1996), a greatest fluidity, a higher cholesteryl ester content and a smaller HDL3 particle size (Sola et al., 1993). Changes in the FA composition of the diet can have several effects on the FA composition of membrane lipids. The fact that both n-6 and n-3 PUFA classes cannot be synthesized de novo in humans, suggests that the fatty acid composition of membrane lipids is strongly influenced by the relative abundance of these PUFA in the diet (Hulbert et al., 2005). Variations in membrane composition can alter processes through which cholesterol desorbs or absorbs on membranes such as free cholesterol efflux or cholesteryl esters uptake, thus influencing RCT. At the same time, when membranes are in contact with FA, these latter can act as modulators of cholesterol transporters at different levels. ABCA1 and ABCG1 are sterol transporters critical for cholesterol efflux and HDL formation, both at the liver and intestine level. Wang and Oram (2007) observed that incubation of mouse macrophages in an oleate-rich medium resulted in more than 50% reduction in cholesterol efflux to apoA-I even if cells were stimulated with 8-Br cAMP or LXR-RXR ligand to induce ABCA1 expression. The unsaturated fatty acids palmitoleate, oleate, linoleate and arachidonate decreased apoA-I mediated cholesterol and PL efflux and apoA-I binding as a result of an increase in degradation of ABCA1. Preincubation of cells in the presence of unsaturated fatty acids results in a reduction of cholesterol efflux via both the ABCA1 and ABCG1 pathways by a downregulation of ABC transporter gene expression (Wang and Oram, 2002; Wang et al., 2004; Uehara et al., 2002, 2007). In addition, enrichment of cellular membrane with n-3 PUFA enhanced HDL-mediated cholesterol efflux from smooth muscle cell (Dusserre et al., 1995) or from human fibroblast (Pal and Davis, 1990).
INFLUENCE OF ALCOHOL CONSUMPTION ON THE REVERSE CHOLESTEROL TRANSPORT PATHWAY Many lines of evidence support a link between moderate drinking and lower risk of CVD but definitive proof remains elusive. Alcohol consumption shifts HDL particle distribution towards large, lipid-rich HDL2b in the heavy alcohol drinkers (Lagrost et al., 1996; Makela et al., 2008). Several studies also document an increase in paraoxonase-1 (PON-1) activity with alcohol intake. In moderate alcohol users, CETP, PLTP and LCAT mass and activities are unchanged, while in heavy alcohol abusers CETP activity and mass are decreased and PLTP activity increased, thus increasing HDL-PL content (Hannuksela paraoxonase-1, 2004). Moderate alcohol intake with dinner affects the postprandial phase by increasing HDL-C, PL and TG (Hendriks
68
The HDL Handbook
et al., 1998; van Tol et al., 1998). In addition, fatty acid composition of PL and CE in HDL is affected by moderate alcohol intake. Alcohol effects on cholesterol efflux were evaluated in both moderate and heavy alcohol consumers. Senault et al. (2000) observed that serum from moderate red wine consumers showed increased capacity of efflux on Fu5AH cells as compared to non-consumers, and that this was correlated to HDL-C, HDL3-C and HDL-PL variations. In alcoholic abusers with or without liver disease, the capacity of HDL particles to mediate cholesterol efflux from macrophages was decreased approximately by 80% and HDLeCE liver uptake was decreased by 60% as compared to non-drinkers. Alcohol is responsible for apoE desialylation that leads to decreased association of apoE with HDL, thus possibly affecting HDL functionality (Rao et al., 2000). By contrast, Makela et al. (2008) recently showed that large PL-rich HDL2 are better acceptors for FC than controls. Furthermore, postprandial serum shows an increased capacity to mediate cellular free cholesterol efflux, independently of the type of alcoholic beverage as compared to fasting serum (van der Gaag et al., 2001). In addition to its action on cholesterol efflux, alcohol could increase another step in RCT such as LCAT-induced free cholesterol esterification or CETP-mediated cholesteryl ester transfer between plasma lipoproteins. Indeed, LCAT activity is increased during the postprandial phase and apoA-I levels increase with wine consumption. Recently, it has been suggested that alcohol may inhibit the glycosylation of CETP, which could affect the binding of CETP to lipoproteins in alcohol drinkers (Liinamaa et al., 2006). Given that the association of alcohol with HDL-C and possibly cardiovascular risk may be partly mediated by CETP activity, genetic variation in CETP may modulate these relationships. A lightto-moderate alcohol intake was associated with a lower risk of CHD among B2 carriers of the Taq1B polymorphism (Jensen et al., 2008) which has been reported to be associated with increased HDL cholesterol levels and decreased CETP activity and levels (Hannuksela et al., 1994).
INFLUENCE OF SMOKING ON THE REVERSE CHOLESTEROL TRANSPORT PATHWAY Cigarette smoking is associated with changes in the levels of plasma lipoproteins with a dose response relationship between the number of cigarettes smoked per day and the extent of lipid abnormalities. However, very little has been understood concerning the molecular mechanisms through which cigarette smoke could accelerate atherosclerosis leading to CHD. Smoking could accelerate atherosclerosis by promoting modifications of lipoproteins and thus influencing HDL functionality. Cigarette smokers have reduced concentrations of plasma HDL and LCAT compared to non-smokers and exposure of human plasma to cigarette smoke resulted in cross-linking of HDL, inhibition of LCAT activity, shift in HDL particle size and dissociation of CETP from particles (Bielicki et al., 1995). HDL that were exposed to whole cigarette smoke
Chapter j 4
HDL and Reverse Cholesterol Transport
69
extracts showed a remarkable increase in the level of conjugated dienes as in oxidized HDL, that inversely correlated with cholesterol efflux capacity. Cigarette smoke extracts contain a variety of free radicals and active oxygens that could be involved in FA peroxidation of HDL (Ueyama et al., 1998). Acrolein is a major component of cigarette smoke and one of the aldehydes that could modify HDL structure and charge. It has been demonstrated that this toxic aldehyde is able to modify apoA-I by site specific modification of lys-226 on HDL and to decrease ABCA1 mediated cholesterol efflux (Shao et al., 2005).
INFLUENCE OF POSTPRANDIAL STATE ON THE REVERSE CHOLESTEROL TRANSPORT PATHWAY Postprandial lipoprotein metabolism is characterized by transient accumulation of intestinally-derived chylomicrons (CM) and hepatically-derived VLDL and their remnants. Hydrolysis of the triglyceride core of CM and VLDL by lipoprotein lipase facilitates the delivery of free fatty acids to muscle and adipose tissue, with formation of CM- and VLDL-remnants which are efficiently removed from the circulation by receptor-mediated pathways in the liver (Cooper, 1997). During lipolysis of postprandial triglyceride-rich lipoproteins, an excess of surface components containing apolipoproteins, unesterified cholesterol and phospholipids is generated and sequesters to HDL potentially via the action of hepatic lipase and PLTP, thereby increasing the total circulating HDL pool and enhancing the transformation of small HDL3 to large CE-rich HDL2 particles (Zilversmit, 1995). Equally, CETP mediates heterotransfer of cholesteryl esters and triglycerides between HDL on the one hand, and apoB-containing lipoproteins on the other; such transfer is accelerated under postprandial conditions with CE enrichment of triglyceride-rich lipoprotein particles, and transient transformation of CE-enriched HDL into TG-enriched particles which become a substrate for hepatic lipase (Castro and Fielding, 1985; Tall et al., 1986; Contacos et al., 1998). In this way, HDL particle chemical composition and size are modulated during the postprandial phase. Elevated postprandial TG is associated with adverse events such as the formation of atherogenic CM remnants, small dense LDL particles and reduction of HDL-C. Fasting plasma TG levels account in part for the amplitude of the postprandial phase and elevated fasting TG is indicative of abnormal postprandial lipoprotein metabolism, for which delayed elimination of postprandial TGRL is associated with atherosclerosis. In patients with coronary artery disease (CAD), postprandial TG levels are elevated and remain so over a prolonged period in comparison with patients without CAD. In type IIb patients, plasma TG accumulation during the postprandial phase was increased as compared to normolipidemic subjects taking the same meal and the incremental area under the curve (AUC) for TG was elevated four-fold (Guerin et al., 2002). Since most people consume fat-containing meals or snack at regular
70
The HDL Handbook
4e5 h intervals, it is clear that the usual state of TG metabolism for most humans is a postprandial one (Lairon, 2007). The postprandial TG response refers to a series of metabolic events that occur after ingestion of a fatcontaining meal and the magnitude of this response is determined by several factors: it increases with fasting plasma TG concentration, age, sedentary life style, and is greater in males than females. Furthermore, the capacity of individuals to regulate circulating TG and clear TGRL is a reflection of their metabolic efficiency, which can be modulated by various gene polymorphisms for key proteins, such as apoE, apoB, apoA-IV and apoC-III, LDL receptor, microsomal transfer protein, fatty acid-binding protein, CETP, LPL and HL (Lairon et al., 2007). Efficient postprandial lipid metabolism with rapid clearance of CMs, promotes HDL formation and could stimulate RCT. After ingestion of dietary fat, an increase in plasma TG concentration is a normal metabolic consequence that is associated with CM and VLDL production and lipolysis. The small intestine delivers CM to the circulation packaged with a truncated form of apoB called apoB48. CM are very quickly transformed to CM remnants by the action of LPL with reduced TG content and particle size. The accumulation of these remnants in the circulation is thought to be one of the most detrimental consequences of the postprandial period, since they can accumulate in the plaque (Rapp et al., 1994) thereby promoting atherosclerosis. During the lipolytic process, surface components (FC, PL and apoE and apoA-I lipoproteins) of CM and VLDL are sequestered to HDL. In this state, CE transfer rate is modulated by changes in plasma lipoproteins and may increase as a result of an increase in CE acceptors and/or CETP concentrations (Tall et al., 1987). This process favors CE enrichment of TGRL and allows transformation of CE-rich HDL in TG-rich HDL that becomes a substrate for HL. On a quantitative basis, VLDL but not CM represent the preferential acceptors among TGRL during the postprandial phase (Lassel et al., 1998) and on a qualitative basis VLDL2 are the preferential ones. In NLS, LDL particles are the major acceptors of CE from HDL in both the fasting and postprandial state despite increases in TGRL (Lassel et al., 1999). However, LDL-CE are secondarily transferred to CM through CETP action, thus CE-enriched CM can be cleared by the liver and irreversibly removed from the body; in this case RCT in the postprandial state is enhanced, thus representing an anti-atherogenic process in normolipidemic subjects. PLTP could play an important role during the postprandial phase by transferring surface PL to the plasma HDL fraction and it is documented that CETP and PLTP activities are modulated by a large intravenous fat load in healthy men (Riemens et al., 1999). Syeda et al. (2003) observed that, during the late postprandial period (6e8 h after meal intake), CETP activation was accompanied by a large increase in HDL-PL but no change in PLTP activity. Furthermore, serum cholesterol efflux capacity was increased during the late postprandial phase in normolipidemic subjects. Guerin et al. (2002) observed that four hours after meal intake, the capacity of postprandial plasma from type IIb subjects to mediate FC efflux was increased
Chapter j 4
HDL and Reverse Cholesterol Transport
71
as compared with the baseline value obtained before meal intake. However, efflux rates to IIb and control plasmas were similar at 2, 4, and 8 h postprandially.
CONCLUSION Interventional studies with lipid-lowering drugs have demonstrated a reduction of up to 50% in LDL-cholesterol levels with concomitant diminution in cardiovascular morbi-mortality of up to approximately 40% in dyslipidemic patients at high risk in both primary and secondary prevention. However, approximately 60% of CV risk, the residual risk, remains to be corrected by nutritional, pharmacological and behavioral approaches. A high proportion of patients exhibiting significant residual risk frequently present with a low HDLcholesterol level (<40 mg/dL). Several specific biological activities underlie the cardioprotective effects of HDL, in particular those in the reverse cholesterol transport pathway (Kontush, 2006). The exit of intracellular cholesterol from the target cell, the foam cell macrophage, and its integration into HDL particles and/or lipid-poor apoA-I, constitutes the first step of the reverse transport of cholesterol, the mechanism by which cholesterol derived from the macrophage and foam cells of the plaque is transported to the liver for its elimination. Additional studies for understanding HDL biology, HDL-mediated atheroprotective pathways in the reverse cholesterol transport pathway, as well as novel biomarkers are needed for the complex assessment of RCT and HDL function/dysfunction in vivo.
ACKNOWLEDGMENTS We are indebted to INSERM (Institut National de la Sante´ et de la Recherche Me´dicale) for generous support of these studies. Giovanna Catalano was the recipient of a Research Fellowship from the New French Atherosclerosis Society (NSFA-2007) and of a Research Fellowship from Fondation pour la Recherche Me´dicale (FRM-2008).
REFERENCES Arnal, J. F., Gourdy, P., et al. (2004). Estrogens and atherosclerosis. Eur J Endocrinol, 150(2), 113e117. Arner, P. (2002). Insulin resistance in type 2 diabetes: role of fatty acids. Diabetes Metab Res Rev, 18(Suppl. 2), S5e9. Avramoglu, R. K., Basciano, H., et al. (2006). Lipid and lipoprotein dysregulation in insulin resistant states. Clin Chim Acta, 368(1-2), 1e19. Bennett, M. K., Lopez, J. M., et al. (1995). Sterol regulation of fatty acid synthase promoter. Coordinate feedback regulation of two major lipid pathways. J Biol Chem, 270(43), 25578e25583.
72
The HDL Handbook
Berrougui, H., Isabelle, M., et al. (2007). Age-related impairment of HDL-mediated cholesterol efflux. J Lipid Res, 48(2), 328e336. Bersot, T. P., Vega, G. L., et al. (1999). Elevated hepatic lipase activity and low levels of high density lipoprotein in a normotriglyceridemic, nonobese Turkish population. J Lipid Res, 40(3), 432e438. Bielicki, J. K., McCall, M. R., et al. (1995). Copper and gas-phase cigarette smoke inhibit plasma lecithin:cholesterol acyltransferase activity by different mechanisms. J Lipid Res, 36(2), 322e331. Brewer, H. B., Jr. (2003). Benefit-risk assessment of Rosuvastatin 10 to 40 milligrams. Am J Cardiol, 92(4B), 23Ke29K. Brites, F., Verona, J., et al. (2004). Enhanced cholesterol efflux promotion in well-trained soccer players. Metabolism, 53(10), 1262e1267. Brown, M. S., & Goldstein, J. L. (1997). The SREBP pathway: regulation of cholesterol metabolism by proteolysis of a membrane-bound transcription factor. Cell, 89(3), 331e340. Buonacorso, V., Nakandakare, E. R., et al. (2007). Macrophage cholesterol efflux elicited by human total plasma and by HDL subfractions is not affected by different types of dietary fatty acids. Am J Clin Nutr, 86(5), 1270e1277. Castro, G. R., & Fielding, C. J. (1985). Effects of postprandial lipemia on plasma cholesterol metabolism. J Clin Invest, 75(3), 874e882. Chapman, M. J., Assmann, G., Fruchart, J. C., Shepherd, J., Sirtori, C., & European Consensus Panel on HDL-C. (2004). Raising high-density lipoprotein cholesterol with reduction of cardiovascular risk: the role of nicotinic acid e a position paper developed by the European Consensus Panel on HDL-C. Curr Med Res Opin, 20(8), 1253e1268. Contacos, C., Barter, P. J., Vrga, L., & Sullivan, D. R. (1998). Cholesteryl ester transfer in hypercholesterolaemia: fasting and postprandial studies with and without pravastatin. Atherosclerosis, 141(1), 87e98. Cooper, A. D. (1997). Hepatic uptake of chylomicron remnants. J Lipid Res, 38(11), 2173e2192. Dusserre, E., Pulcini, T., Bourdillon, M. C., Ciavatti, M., & Berthezene, F. (1995). Omega-3 fatty acids in smooth muscle cell phospholipids increase membrane cholesterol efflux. Lipids, 30(1), 35e41. Fluiter, K., van der Westhuijzen, D. R., et al. (1998). In vivo regulation of scavenger receptor BI and the selective uptake of high density lipoprotein cholesteryl esters in rat liver parenchymal and Kupffer cells. J Biol Chem, 273(14), 8434e8438. Gillotte, K. L., Lund-Katz, S., et al. (1998). Dietary modification of high density lipoprotein phospholipid and influence on cellular cholesterol efflux. J Lipid Res, 39(10), 2065e2075. Godsland, I. F., Wynn, V., et al. (1987). Sex, plasma lipoproteins, and atherosclerosis: prevailing assumptions and outstanding questions. Am Heart J, 114(6), 1467e1503. Goldberg, I. J., & Merkel, M. (2001). Lipoprotein lipase: physiology, biochemistry, and molecular biology. Front Biosci, 6, D388e405. Graf, G. A., Roswell, K. L., et al. (2001). 17beta-Estradiol promotes the up-regulation of SR-BII in HepG2 cells and in rat livers. J Lipid Res, 42(9), 1444e1449. Gruber, C. J., Tschugguel, W., et al. (2002). Production and actions of estrogens. N Engl J Med, 346(5), 340e352. Grundy, S. M. (1995). Role of low-density lipoproteins in atherogenesis and development of coronary heart disease. Clin Chem, 41(1), 139e146. Guerin, M., Egger, P., et al. (2002). Cholesteryl ester flux from HDL to VLDL-1 is preferentially enhanced in type IIB hyperlipidemia in the postprandial state. J Lipid Res, 43(10), 1652e1660.
Chapter j 4
HDL and Reverse Cholesterol Transport
73
Hannuksela, M. L., Liinamaa, M. J., et al. (1994). Relation of polymorphisms in the cholesteryl ester transfer protein gene to transfer protein activity and plasma lipoprotein levels in alcohol drinkers. Atherosclerosis, 110(1), 35e44. Hannuksela, M. L., Ramet, M. E., et al. (2004). Effects of ethanol on lipids and atherosclerosis. Pathophysiology, 10(2), 93e103. Hendriks, H. F., Veenstra, J., et al. (1998). Moderate doses of alcoholic beverages with dinner and postprandial high density lipoprotein composition. Alcohol Alcohol, 33(4), 403e410. Howard, B. V. (1999). Insulin resistance and lipid metabolism. Am J Cardiol, 84(1A), 28Je32J. Hulbert, A. J., Turner, N., et al. (2005). Dietary fats and membrane function: implications for metabolism and disease. Biol Rev Camb Philos Soc, 80(1), 155e169. Iglesias, A., Montelongo, A., et al. (1994). Changes in cholesteryl ester transfer protein activity during normal gestation and postpartum. Clin Biochem, 27(1), 63e68. Jaouad, L., de Guise, C., et al. (2006). Age-related decrease in high-density lipoproteins antioxidant activity is due to an alteration in the PON1’s free sulfhydryl groups. Atherosclerosis, 185(1), 191e200. Jensen, M. K., Mukamal, K. J., et al. (2008). Alcohol consumption, TaqIB polymorphism of cholesteryl ester transfer protein, high-density lipoprotein cholesterol, and risk of coronary heart disease in men and women. Eur Heart J, 29(1), 104e112. Jin, F. Y., Kamanna, V. S., et al. (1998). Estradiol stimulates apolipoprotein A-I- but not A-IIcontaining particle synthesis and secretion by stimulating mRNA transcription rate in Hep G2 cells. Arterioscler Thromb Vasc Biol, 18(6), 999e1006. Jones, D. R., Schmidt, R. J., et al. (2002). Estrogen receptor-mediated repression of human hepatic lipase gene transcription. J Lipid Res, 43(3), 383e391. Kauma, H., Savolainen, M. J., et al. (1996). Sex difference in the regulation of plasma high density lipoprotein cholesterol by genetic and environmental factors. Hum Genet, 97(2), 156e162. Khalil, A., Jay-Gerin, J. P., et al. (1998). Age-related increased susceptibility of high-density lipoproteins (HDL) to in vitro oxidation induced by gamma-radiolysis of water. FEBS Lett, 435(2-3), 153e158. Kok, N., Roberfroid, M., et al. (1996). Involvement of lipogenesis in the lower VLDL secretion induced by oligofructose in rats. Br J Nutr, 76(6), 881e890. Kontush, A., & Chapman, M. J. (2006). Functionally defective high-density lipoprotein: a new therapeutic target at the crossroads of dyslipidemia inflammation, and atherosclerosis. Pharmacol. Rev, 58(3), 342e374. Kuusi, T., Taskinen, M. R., et al. (1988). Lipoproteins, lipolytic enzymes, and hormonal status in hypothyroid women at different levels of substitution. J Clin Endocrinol Metab, 66(1), 51e56. Lagrost, L., Athias, A., et al. (1996). Opposite effects of cholesteryl ester transfer protein and phospholipid transfer protein on the size distribution of plasma high density lipoproteins. Physiological relevance in alcoholic patients. J Biol Chem, 271(32), 19058e19065. Lairon, D., Lopez-Miranda, J., et al. (2007). Methodology for studying postprandial lipid metabolism. Eur J Clin Nutr, 61(10), 1145e1161. Landschulz, K. T., Pathak, R. K., et al. (1996). Regulation of scavenger receptor, class B, type I, a high density lipoprotein receptor, in liver and steroidogenic tissues of the rat. J Clin Invest, 98(4), 984e995. Langer, C., Gansz, B., et al. (2002). Testosterone up-regulates scavenger receptor BI and stimulates cholesterol efflux from macrophages. Biochem Biophys Res Commun, 296(5), 1051e1057. LaRosa, J. C. (1992). Lipids and cardiovascular disease: do the findings and therapy apply equally to men and women? Womens Health Issues, 2(2), 102e111, discussion 111e113.
74
The HDL Handbook
Lassel, T. S., Guerin, M., et al. (1998). Preferential cholesteryl ester acceptors among triglyceriderich lipoproteins during alimentary lipemia in normolipidemic subjects. Arterioscler Thromb Vasc Biol, 18(1), 65e74. Lassel, T. S., Guerin, M., et al. (1999). Evidence for a cholesteryl ester donor activity of LDL particles during alimentary lipemia in normolipidemic subjects. Atherosclerosis, 147(1), 41e48. Liinamaa, M. J., Hannuksela, M. L., et al. (2006). Defective glycosylation of cholesteryl ester transfer protein in plasma from alcohol abusers. Alcohol Alcohol, 41(1), 18e23. Makela, S. M., Jauhiainen, M., et al. (2008). HDL2 of heavy alcohol drinkers enhances cholesterol efflux from raw macrophages via phospholipid-rich HDL 2b particles. Alcohol Clin Exp Res, 32(6), 991e1000. Muls, E., Blaton, V., et al. (1982). Serum lipids and apolipoproteins A-I, A-II, and B in hyperthyroidism before and after treatment. J Clin Endocrinol Metab, 55(3), 459e464. Nanjee, M. N., Koritnik, D. R., et al. (1990). Hormonal determinants of apolipoprotein B, E receptor expression in human liver. Positive association of receptor expression with plasma estrone concentration in middle-aged/elderly women. Biochim Biophys Acta, 1046(2), 151e158. Nicholls, S. J., Lundman, P., Harmer, J. A., et al. (2006). Consumption of saturated fat impairs the anti-inflammatory properties of high-density lipoproteins and endothelial function. J Am Coll Cardiol, 48(4), 715e720. Olchawa, B., Kingwell, B. A., et al. (2004). Physical fitness and reverse cholesterol transport. Arterioscler Thromb Vasc Biol, 24(6), 1087e1091. Pal, S., & Davis, P. J. (1990). N-3 polyunsaturated fatty acids enhance cholesterol efflux from human fibroblasts in culture. Biochem Biophys Res Commun, 173(2), 566e570. Parini, P., Angelin, B., et al. (1997). Importance of estrogen receptors in hepatic LDL receptor regulation. Arterioscler Thromb Vasc Biol, 17(9), 1800e1805. Parini, P., Angelin, B., et al. (2000). Biphasic effects of the natural estrogen 17beta-estradiol on hepatic cholesterol metabolism in intact female rats. Arterioscler Thromb Vasc Biol, 20(7), 817e823. Patsch, W., Kim, K., et al. (1980). Effects of sex hormones on rat lipoproteins. Endocrinology, 107 (4), 1085e1094. Psota, T. L., Gebauer, S. K., & Kris-Etherton, P. (2006). Dietary omega-3 fatty acid intake and cardiovascular risk. Am J Cardiol, 98(4A), 3ie18i. Rao, M. N., Liu, Q. H., et al. (2000). High-density lipoproteins from human alcoholics exhibit impaired reverse cholesterol transport function. Metabolism, 49(11), 1406e1410. Rapp, J. H., Lespine, A., et al. (1994). Triglyceride-rich lipoproteins isolated by selected-affinity anti-apolipoprotein B immunosorption from human atherosclerotic plaque. Arterioscler Thromb, 14(11), 1767e1774. Riemens, S. C., Van Tol, A., et al. (1999). Acute and chronic effects of a 24-hour intravenous triglyceride emulsion challenge on plasma lecithin: cholesterol acyltransferase, phospholipid transfer protein, and cholesteryl ester transfer protein activities. J Lipid Res, 40(8), 1459e1466. Sakr, S. W., Senault, C., et al. (1996). Oleic acid-rich fats increase the capacity of postprandial serum to promote cholesterol efflux from Fu5AH cells. Biochim Biophys Acta, 1300(1), 49e55. Sattar, N., Greer, I. A., et al. (1997). Lipoprotein subfraction changes in normal pregnancy: threshold effect of plasma triglyceride on appearance of small, dense low density lipoprotein. J Clin Endocrinol Metab, 82(8), 2483e2491.
Chapter j 4
HDL and Reverse Cholesterol Transport
75
Schaefer, E. J., Zech, L. A., et al. (1982). Human apolipoprotein A-I and A-II metabolism. J Lipid Res, 23(6), 850e862. Senault, C., Betoulle, D., et al. (2000). Beneficial effects of a moderate consumption of red wine on cellular cholesterol efflux in young men. Nutr Metab Cardiovasc Dis, 10(2), 63e69. Seres, I., Paragh, G., et al. (2004). Study of factors influencing the decreased HDL associated PON1 activity with aging. Exp Gerontol, 39(1), 59e66. Shao, B., O’Brien, D. K., et al. (2005). Acrolein modifies apolipoprotein A-I in the human artery wall. Ann N Y Acad Sci, 1043, 396e403. Silliman, K., Tall, A. R., et al. (1993). Unusual high-density lipoprotein subclass distribution during late pregnancy. Metabolism, 42(12), 1592e1599. Sola, R., Motta, C., et al. (1993). Dietary monounsaturated fatty acids enhance cholesterol efflux from human fibroblasts. Relation to fluidity, phospholipid fatty acid composition, overall composition, and size of HDL3. Arterioscler Thromb, 13(7), 958e966. Staels, B., Van Tol, A., et al. (1990). Alterations in thyroid status modulate apolipoprotein, hepatic triglyceride lipase, and low density lipoprotein receptor in rats. Endocrinology, 127(3), 1144e1152. Sviridov, D., & Nestel, P. (2002). Dynamics of reverse cholesterol transport: protection against atherosclerosis. Atherosclerosis, 161(2), 245e254. Sviridov, D., Kingwell, B., et al. (2003). Single session exercise stimulates formation of pre beta 1-HDL in leg muscle. J Lipid Res, 44(3), 522e526. Syeda, F., Senault, C., et al. (2003). Postprandial variations in the cholesteryl ester transfer protein activity, phospholipid transfer protein activity and plasma cholesterol efflux capacity in normolipidemic men. Nutr Metab Cardiovasc Dis, 13(1), 28e36. Tall, A., Granot, E., et al. (1987). Accelerated transfer of cholesteryl esters in dyslipidemic plasma. Role of cholesteryl ester transfer protein. J Clin Invest, 79(4), 1217e1225. Tall, A., Sammett, D., & Granot, E. (1986). Mechanisms of enhanced cholesteryl ester transfer from high density lipoproteins to apolipoprotein B-containing lipoproteins during alimentary lipemia. J Clin Invest, 77(4), 1163e1172. Tikkanen, M. J., Kuusi, T., et al. (1986). Variation of postheparin plasma hepatic lipase by menstrual cycle. Metabolism, 35(2), 99e104. Uehara, Y., Engel, T., Li, Z., et al. (2002). Polyunsaturated fatty acids and acetoacetate downregulate the expression of the ATP-binding cassette transporter A1. Diabetes, 51(10), 2922e2928. Uehara, Y., Miura, S., et al. (2007). Unsaturated fatty acids suppress the expression of the ATPbinding cassette transporter G1 (ABCG1) and ABCA1 genes via an LXR/RXR responsive element. Atherosclerosis, 191(1), 11e21. Ueyama, K., Yokode, M., et al. (1998). Cholesterol efflux effect of high density lipoprotein is impaired by whole cigarette smoke extracts through lipid peroxidation. Free Radic Biol Med, 24(1), 182e190. Vadlamudi, S., MacLean, P., et al. (1998). Role of female sex steroids in regulating cholesteryl ester transfer protein in transgenic mice. Metabolism, 47(9), 1048e1051. van der Gaag, M. S., van Tol, A., et al. (2001). Alcohol consumption stimulates early steps in reverse cholesterol transport. J Lipid Res, 42(12), 2077e2083. van Tol, A., van der Gaag, M. S., et al. (1998). Changes in postprandial lipoproteins of low and high density caused by moderate alcohol consumption with dinner. Atherosclerosis, 141(Suppl. 1), S101e103. von Eckardstein, A., Nofer, J. R., et al. (2001). High density lipoproteins and arteriosclerosis. Role of cholesterol efflux and reverse cholesterol transport. Arterioscler Thromb Vasc Biol, 21(1), 13e27.
76
The HDL Handbook
Walter, M. (2009). Interrelationships among HDL metabolism, aging and atherosclerosis. Arterioscler Thromb Vasc Biol, 29, 1244e1250. Wang, Y., & Oram, J. F. (2002). Unsaturated fatty acids inhibit cholesterol efflux from macrophages by increasing degradation of ATP-binding cassette transporter A1. J Biol Chem, 277(7), 5692e5697. Wang, Y., & Oram, J. F. (2007). Unsaturated fatty acids phosphorylate and destabilize ABCA1 through a protein kinase C delta pathway. J Lipid Res, 48(5), 1062e1068. Wang, Y., Kurdi-Haidar, B., & Oram, J. F. (2004). LXR-mediated activation of macrophage stearoyl-CoA desaturase generates unsaturated fatty acids that destabilize ABCA1. J Lipid Res, 45(5), 972e980. Zhang, X., Moor, A. N., et al. (2007). Regulation of alternative splicing of liver scavenger receptor class B gene by estrogen and the involved regulatory splicing factors. Endocrinology, 148(11), 5295e5304. Zilversmit, D. B. (1995). Atherogenic nature of triglycerides, postprandial lipidemia, and triglyceride-rich remnant lipoproteins. Clin Chem, 41(1), 153e158. Zitzmann, M., Brune, M., et al. (2001). The CAG repeat polymorphism in the AR gene affects high density lipoprotein cholesterol and arterial vasoreactivity. J Clin Endocrinol Metab, 86(10), 4867e4873.
Chapter 5
Serum Paraoxonase (PON1) and its Interactions with HDL: Relationship between PON1 and Oxidative Stress Jelena Vekic, Jelena Kotur-Stevuljevic, Aleksandra Zeljkovic, Aleksandra Stefanovic, Zorana Jelic-Ivanovic, Slavica Spasic and Vesna Spasojevic-Kalimanovska Institute of Medical Biochemistry, Faculty of Pharmacy, University of Belgrade, Belgrade, Serbia
INTRODUCTION Perturbations in lipid homeostasis in combination with conditions favoring oxidative stress form the main conjecture of atherosclerosis. Both assign a central role to low density lipoproteins (LDL) in initiating atherogenesis and to high density lipoproteins (HDL) in preventing this process. The current concept of atherosclerosis as a low-grade inflammation supports the role of oxidatively modified lipoproteins as the trigger of the cascade of events that finally lead to plaque formation (Ross, 1999). A critical step in atherogenesis is the entry and accumulation of LDLassociated lipids within the vessel wall. When trapped in the subendothelial space, LDL particles undergo oxidation through multiple pathways. Once formed, oxidized LDL possesses significant pro-inflammatory (Cominacini et al., 1997), pro-thrombotic (Ishii et al., 1996) and pro-apoptotic (Heermeier et al., 2001) potential which together further aggravate the imbalanced microenvironment of the vascular intima. For a long time the removal of lipids from the vascular wall was considered as the key anti-atherogenic mechanism of HDL particles. It is now recognized that atherosclerosis is more than a simple over-supply of cholesterol. Furthermore, extensive research on HDL has uncovered a wide spectrum of its atheroprotective functions. The present chapter will focus on the antioxidative properties of HDL particles and will discuss how the in vitro evidence translates The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
77
78
The HDL Handbook
into in vivo confirmation of HDL-mediated protection, which may be relevant in the prevention or development of atherosclerosis.
HDL e STRUCTURE, HETEROGENEITY AND ANTI-ATHEROGENIC PROPERTIES HDL particles, although the smallest lipoproteins in plasma, remain one of the most intriguing topics in science and they undoubtedly represent attractive lines of research in the future. Ever since low HDL cholesterol (HDL-C) concentration was recognized as a potent risk factor for cardiovascular disease (CVD) (Miller and Miller, 1975), preserving its optimal level became the imperative of preventive medicine. Although the measurement of HDL-C has proven clinical utility, its cholesterol content is only one of the features that characterizes HDL particles, a feature that does not reflect their diversity nor dynamic nature. HDL represents a highly heterogeneous mixture of lipoproteins which all have in common a high density (>1.063 kg/L) and a small size (5e17 nm). Regarding the basic composition, HDL particles have the highest proportion of protein relative to lipid compared to other plasma lipoproteins. The major proteins are apolipoprotein A-I (apoA-I) and apoA-II, with small amounts of apo C, E, A-IV, D and J. The principal lipid components are the phospholipids, with lesser amounts of cholesterol esters, unesterified cholesterol and triglycerides. HDL particles also carry several important enzymes: paraoxonase-1 (PON1), platelet-activating factor acetyl hydrolase (PAF-AH), glutathione selenoperoxidase (GSPx), lecithin:cholesterol acyltransferase (LCAT) and phospholipid transfer protein (PLTP).
HDL subclasses Variations in the qualitative and quantitative content of lipids and apolipoproteins result in the presence of various HDL subclasses, characterized by different shape, charge, density and size (Skinner, 1994). This remarkable heterogeneity arises from their complex synthesis and continuous intravascular remodeling processes. The understanding that structure and composition of each HDL subclass reflect its different metabolic function challenges the traditional approach of using only HDL-C level for the management of cardiovascular risk. To date, investigators have described several different approaches to separate HDL subclasses based on a variety of their physicochemical properties including particle density and size, charge and apolipoprotein composition. Here, we address the most commonly used techniques. Density gradient ultracentrifugation represents the original procedure from which subsequent methods have been calibrated and validated. When fractionated by ultracentrifugation, HDL particles are usually separated into two major subclasses; HDL2 (density range 1.063e1.125 kg/L) and HDL3 (density
Chapter j 5
PON1 and HDL
79
range 1.125e1.210 kg/L) (Chapman et al., 1981). However, during the ultracentrifugation process labile lipoproteins may be substantially altered by the high salt conditions and centrifugal forces which can lead to loss of loosely bound apolipoproteins and enzymes (Kunitake and Kane, 1982; Cheung and Wolf, 1988). In spite of numerous disadvantages, ultracentrifugation is still an irreplaceable preparative technique in numerous research laboratories. Electrophoretic methods separate HDL particles on the basis of their charge and size. In agarose gel electrophoresis, most HDL particles have a-mobility with the exception of a minor fraction that migrates in the pre-b region (Kunitake et al., 1985). Non-denaturing polyacrylamide gradient gel electrophoresis separates HDL into five distinct subpopulations ranging from smaller, denser HDL3c (7.2e7.8 nm), HDL3b (7.8e8.2 nm) and HDL3a (8.2e8.8 nm) to larger HDL2a (8.8e9.7 nm) and HDL2b (9.7e12 nm) subclasses (Blanceh et al., 1981). Finally, HDL subclasses are often classified based on their particular apolipoprotein composition into HDL particles that contain only apoA-I (LpAI) and those that contain both apoA-I and apoA-II (LpAI:AII) (Cheung and Albers, 1982). In our laboratory, we separate HDL subclasses using a modified gradient (3e31%) gel electrophoretic method originally developed by Rainwater et al. (1997). In addition to HDL, this method also provides the separation of LDL particles on the same gradient gel (Figure 5.1). From our own experience, this procedure is reasonably reproducible, relatively economical and therefore suitable for use in population and clinical studies (Vekic et al., 2007a). A key question remains to be answered. Do the various HDL subclasses equally contribute to the anti-atherogenic properties of HDL? The main controversy centers around the role of smaller, denser HDL particles. Numerous studies have investigated whether smaller, denser HDL particles are more or less protective than their larger counterparts. However, attempts to resolve the question have provided somewhat ambiguous results. It is worth mentioning that the use of different methodological approaches and, consequently, different nomenclature for the same class of particles may be the explanation of inconsistent and confusing results. Another reason could be the fact that HDL particles can undergo dramatic changes during inflammatory conditions which considerably alter or attenuate their anti-atherogenic effects (Ansell et al., 2005). Thus, smaller, denser HDL particles appear to have a dual role in atherogenesis; functional particles could be protective, whereas alterations in their structure could implicate pro-atherogenic properties.
Anti-atherogenic properties of HDL The protective role of HDL is well documented, but the precise mechanisms by which it exerts its anti-atherogenic effects are still being characterized. HDL particles prevent the development of atherosclerosis through multiple pathways including both reverse cholesterol transport (RCT) and cholesterol-independent
80
The HDL Handbook
FIGURE 5.1 Separation of LDL and HDL particles on a polyacrylamide gradient (3e31%) gel. Electrophoresis was performed at 8 C using Tris-boric acid-Na2EDTA buffer, pH 8.35 for 20 hours. Separated fractions were stained with Sudan Black. The relative distribution of each subfraction can be estimated by determining the areas under the peaks of the densitometric scan of the sample.
mechanisms. Although this chapter mostly concerns antioxidative properties of HDL particles, such a discussion would be incomplete without briefly considering their other significant roles.
Reverse cholesterol transport The process involves the unique ability of HDL particles to remove excess cholesterol from peripheral tissues to the liver. To start the process the liver and intestine synthesize lipid-poor apoA-I. Once in the bloodstream, circulating nascent HDL particles receive free cholesterol from tissues via their interactions with the ATP-binding membrane cassette transport proteins A1 (ABCA1) and G1/G4 (ABCAG1/G4) expressed on macrophages, scavenger receptor class B type 1 (SR-B1), caveolin and even via passive diffusion (Yancey et al., 2003). Subsequently, LCAT esterifies free cholesterol within nascent HDL and produces mature HDL3 (smaller, denser HDL) and HDL2 (larger, less dense HDL) particles (Jonas, 1991). Mature HDL has two metabolic pathways. In the direct pathway, cholesterol esters within HDL undergo selective uptake by hepatocytes via SR-B1 receptors or LDL receptors (if HDL contains apoE) and subsequent excretion into the bile. In the indirect pathway, cholesterol esters within HDL are exchanged for triglycerides in apolipoprotein B
Chapter j 5
PON1 and HDL
81
(apoB)-containing particles (LDL and VLDL) through the action of cholesterol ester transfer protein (CETP). Finally, apoB-rich lipoproteins deliver cholesterol esters to the liver via LDL receptors. After removal of cholesterol esters, triglyceride-rich HDL undergoes hydrolysis by hepatic lipase (HL) and endothelial lipase (EL). The remaining smaller, denser HDL releases apoA-I which then either participates in the next lipidation cycle or is catabolized in the kidneys (Lewis and Rader, 2005). Although the mechanism of the RCT process is still not fully elucidated, its efficiency is considered to be an important determinant of HDL’s anti-atherogenic potential.
Anti-inflammatory effects Some authors have suggested that HDL is capable of inhibiting cytokineinduced expression of adhesion molecules (vascular cell adhesion molecule (VCAM)-1, intracellular adhesion molecule (ICAM)-1 and E-selectin) thereby impeding monocyte adhesion on the endothelial surface (Cockerill et al., 2001). However, we failed to confirm the proposed association between low HDL-C concentrations and adhesion molecule overexpression in normolipidemic subjects (Bogavac-Stanojevic et al., 2005). Furthermore, there is evidence suggesting that the main protein in HDL (apoA-I) is able to bind to bacterial endotoxin thereby inhibiting endotoxin-induced cellular activation (Ma et al., 2004). In addition, the significant anti-inflammatory potential of HDL is a consequence of its antioxidative effects. Vasoprotective and anti-thrombotic effects HDL preserves the integrity of the endothelium via stimulation of endothelial cell migration and suppression of their apoptosis (Nofer et al., 2001). Incubation of cultured endothelial cells with HDL particles activates nitric oxide (NO) synthase (eNOS), with subsequent NO release and NO-dependent vasodilatation (Kuvin et al., 2002). Furthermore, by stimulating endothelial prostacyclin synthesis, HDL can further enhance vasorelaxation (Spector et al., 1985). The inhibitory action of HDL on platelet activation and aggregation can be accomplished directly (Lerch et al., 1998), or by modulating the production of the above-mentioned endothelial factors. In addition, HDL may prevent thrombogenesis via regulation of thrombomodulin synthesis and enhancement of protein C and protein S activities (Griffin et al., 1999; Nicholls et al., 2005). Antioxidative effects HDL is also believed to protect against atherosclerosis by inhibiting the oxidative modification of LDL thereby attenuating the biological activities of oxidized LDL (Ng et al., 2005). Such antioxidative properties have been attributed to various proteins associated with HDL, mainly a set of enzymes: LCAT, PAF-AH, PON1, proteinase (elastase-like activity) and phospholipase D but also due to albumin and apoA-I (Durrington et al., 2001). PON1 was found
82
The HDL Handbook
to be more effective than LCAT or apoA-I in protecting LDL against oxidation, although it was documented that the antioxidative effect of PON1 was slightly enhanced in the presence of LCAT and apoA-I (Arrol et al., 1996). The activity of PAF-AH is believed to be similar to that of PON1. However, studies using PON1-knockout mice have demonstrated that PAF-AH was unable to perform its function in the absence of PON1, indicating that PON1 is the primary enzyme responsible for LDL protection (Rodrigo et al., 2001). HDL is the most abundant lipoprotein in tissue fluid and the only lipoprotein in the central nervous system. Based on the evidence that PON1 is present not only in the circulation but also in other tissue fluids, some authors suspect that the antioxidative function of PON1 has evolved to protect against the development of atherosclerosis, a disease prevalent for less than a century (Durrington et al., 2001).
PARAOXONASE-1 (PON1) PON1 (EC 3.1.8.1, aryldialkylphosphatase) is specifically an HDL-associated serum enzyme (Durrington et al., 2001). Its primary function was first described as an organophosphate hydrolase, meaning that it could degrade commonly used insecticides and nerve gases. Recent reports have suggested that PON1 is largely responsible for the breakdown of lipid peroxides before they accumulate in LDL particles (Mackness et al., 1991a). In addition, PON1 also renders HDL resistant to oxidation (Durrington et al., 2001). According to Aviram et al. (2000) PON1 is able to hydrolyze lipid peroxides in human atherosclerotic lesions. Recent evidence has confirmed that PON1 is primarily a lactonase acting upon certain lactones/hydroxyl acids. Therefore, one of the physiological roles of PON1 is the metabolism of lipid species arising from the oxidation of polyunsaturated fatty acids (Draganov et al., 2005).
PON1 polymorphism Research regarding the PON1 gene sequence has demonstrated nearly 200 polymorphic sites (La Du, 2003). PON1 has two common coding polymorphisms: a methionine (M) to leucine (L) substitution at position 55 and a glutamine (Q) to arginine (R) substitution at position 192. In addition, PON1 has at least three structural variants in the same region. The polymorphic sites are also found in the non-coding regions of the PON1 gene: 7 are in the 5’-untranslated region, 171 are in intronic regions and 15 are in the 3’-untranslated region. However, the most prominent coding polymorphisms (Q192R and L55M) have particular pathophysiological significance. Numerous studies have indicated that both polymorphisms could be associated with CVD, stroke, familial hypercholesterolemia, type 2 diabetes mellitus and Parkinson’s disease (Ruiz et al., 1995; Schmidt et al., 1998; Akhmedova et al., 2001; Ng et al., 2005). Of particular note is the fact that PON1Q192R isoforms have
Chapter j 5
PON1 and HDL
83
substrate-dependent differences in their enzyme activity; the Q variant has a higher diazoxonase activity (DZOase) whereas paraoxonase activity (POase) predominates in the R variant (Serrato and Marian, 1995). By measuring the DZOase/POase activity ratio it is possible to discriminate between the PON1192QQ, PON1192QR and PON1192RR phenotypes.
PON1 status Possible involvement of the PON1Q192R polymorphism in vascular disease development has often been evaluated only by determination of PON1 genotype without considering PON1 concentration and/or activity (Richter et al., 2003). Richter and Furlong (1999) have characterized “PON1 status” as PON1 enzymatic activities towards two non-physiological organophosphate substrates e paraoxon and diazoxon in combination with PON1Q192R phenotype determination. By plotting the rates of PON1’s ability to hydrolyze diazoxon (DZOase PON1 activity) vs. paraoxon (POase PON1 activity), the study population can be divided into three distinct groups: individuals functionally homozygous for PON1192QQ, heterozygotes (PON1192QR) and individuals homozygous for PON1192RR. Therefore, PON1 status determination provides information not only on the PON1Q192R polymorphism but also on the two PON1 enzymatic activities. Richter et al. (2003) found that the PON1 status of individuals with different PON1Q192R genotypes can be quite similar, suggesting that PCR analysis was unable to discover inactive alleles. This finding indicates that PON1 status determination provides insight into the functional state of PON1 genomics in any given individual (Richter et al., 2003).
THE ASSOCIATION BETWEEN PON1 AND HDL The relationship between PON1 and HDL relies on a unique reciprocal arrangement. HDL facilitates the secretion of PON1 and offers a hydrophobic harbor for its anchoring. It also provides a hydrophobic environment that is essential for optimal PON1 activity (James and Deakin, 2004). In return, the enzyme prevents the oxidation of HDL and stimulates the cholesterol efflux from cells which enhances the capacity of RCT (Rosenblat et al., 2005). Therefore, the association between PON1 and HDL should be considered in both normal and pathological conditions.
PON1 and HDL in healthy subjects Variations in PON1 activity and phenotype distribution may be the cause of different predispositions to atherosclerosis within populations (Brophy et al., 2002). Such diversity in PON1 activities is largely genetically determined but it can also be modulated by various dietary (Shih et al., 1996; Kleemola et al.,
84
The HDL Handbook
2002) and lifestyle factors (Ferre´ et al., 2003). Recently, we studied the PON1Q192R polymorphism in Serbia (Kotur-Stevuljevic et al., 2006). We concluded that the frequencies of PON1 phenotypes in our healthy population were in Hardy-Weinberg equilibrium and comparable to those previously reported in other European populations (Table 5.1). A recent study by Blatter Garin et al. (2006) confirmed a significant positive association between PON1 activity and apoA-I and HDL-C concentrations and considerable influence of PON1 polymorphism on apoA-I and HDL-C levels. With the current appreciation of HDL variability, it has been postulated that PON1 is not equally distributed among HDL particles but, rather, it has a specific association with particular HDL subpopulations. Zech et al. (1993) suggested that the PON1-HDL association follows a normal remodeling process of HDL particles. Initially, PON1 is associated with smaller, denser HDL3 subfractions which will subsequently be converted into larger HDL2 particles. However, different studies have reported conflicting results. PON1 has been found to be associated with larger HDL particles isolated by gel filtration (Blatter et al., 1993) or with smaller HDL particles fractionated by ultracentrifugation (Kontush et al., 2003). On the other hand, it has been suggested that PON1 is preferentially associated with larger HDL particles in vivo and that it can be displaced to smaller HDL particles (Kontush et al., 2003; Bergmeier et al., 2004) or even accumulated in the lipoprotein-free plasma (Rosenblat et al., 2006) upon ultracentrifugation. Knowledge concerning the mechanisms of antioxidative protection by HDL is mainly derived from the results of basic science studies. Recently, we studied the association of oxidative stress and PON1 status and HDL particle size
TABLE 5.1
PON1Q192R allele frequencies in European populations PON1 alleles
Population
N
Q
R
Reference
Serbian
105
0.72
0.28
Kotur-Stevuljevic et al. (2006)
Italian
162
0.70
0.30
Ombres et al. (1998)
Austrian
144
0.75
0.25
Schmidt et al. (1998)
German
971
0.73
0.27
Cascorbi et al. (1999)
Swiss
273
0.73
0.27
James et al. (2000)
British
152
0.75
0.25
Mackness et al. (2000)
Spanish
310
0.70
0.30
Senti et al. (2000)
Russian
117
0.74
0.26
Akhmedova et al. (2001)
Chapter j 5
PON1 and HDL
85
heterogeneity in a clinical setting (Vekic et al., 2007b). All study participants were deemed clinically healthy and were middle-aged. We found that oxidative stress parameters positively correlated with the levels of smaller, denser HDL3b and HDL3c but negatively with larger HDL2a subclasses. In addition, PON1 activity exhibited a profound inverse relationship with HDL size. Our findings could represent in vivo evidence that oxidative stress could alter the maturation of HDL particles, perhaps due to impaired LCAT activity (Mertens et al., 2003) and also to attenuate their antioxidative potential by inactivation of PON1 (Mertens et al., 2003). Another interesting finding was significantly reduced HDL particle size among the carriers of the PON1192R allele (Vekic et al., 2007b). According to Mackness at al. (1998), the expression of the R allele is directly related to a lower capacity of PON1 to protect LDL from oxidation. Thus, we concluded that smaller, denser HDL particles could have a lower antioxidative potential. Furthermore, in a recent study, smaller HDL particles were found to be associated with elevated uric acid and markers of inflammation in apparently healthy subjects (Vekic et al., 2009). There is evidence to suggest that uric acid (UA) may function as a potent antioxidant by scavenging free radicals and also by stabilizing ascorbate in biological fluids. Therefore, elevated UA could also be considered to be a compensatory mechanism that counteracts oxidative stress related to the preponderance of smaller HDL particles.
PON1 and HDL in cardiovascular disease Reports that sought associations between the PON1192R allele (or PON1192RR genotype) and CVD are considered controversial. Some studies reported a positive association (Humbert et al., 1993; Zama et al., 1997; Sanghera et al., 1998), whereas others failed to show any relationship (Antikainen et al., 1996; Suehiro et al., 1996; Rice et al., 1997; Ombres et al., 1998). There is also evidence to suggest that low PON1 activity is associated with atherosclerosis irrespective of the PON1 phenotype (Mackness et al., 1991b, 2003; Jarvik et al., 2003). Jarvik et al. (2003) recently showed that PON1 activity was a better predictor of disease development than the PON1 genotype. In our own study, we found a significantly lower POase activity (24% reduced) in combination with reduced, but marginally significant, DZOase activity in a group of coronary heart disease (CHD) patients when compared with a healthy population (Kotur-Stevuljevic et al., 2006). Moreover, after the patients were separated according to the results of coronary angiography (CAD and CADþ), both groups had lower POase and DZOase activities compared with the controls (Figure 5.2). The distribution of the PON1Q192R phenotypes (estimation based on Figure 5.3) in the CHD patients was not significantly different from the control population. Moreover, the phenotype distribution within the CADþ and CAD groups did not differ between themselves nor from the control population.
86
The HDL Handbook
FIGURE 5.2 PON1 POase and DZOase activities in coronary heart disease (CHD) patients and controls. CHD patients were also separated according to coronary angiography results into CADþ and CAD groups (*P<0.05 respective group vs. control group, **P<0.01 respective group vs. control group). Salt-stimulated PON1 activities toward paraoxon (POase) and diazoxon (DZOase) were measured spectrophotometrically.
Despite the fact that the frequency distributions were re-evaluated according to gender, age, smoking habits and an HDL-C cut-off value of 1.03 mmol/L (Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults, 2001), we failed to detect differences between groups in the control and CHD populations. Logistic regression analysis confirmed that PON1 activity, but not the PON1192RR phenotype, was the significant determinant of the risk for CHD in our population. The observed reduction in PON1 activity in CHD patients was proposed to be through partial inactivation of the enzyme as a consequence of increased oxidative stress and inflammation (Kotur-Stevuljevic et al., 2007). Oxidative stress plays an important role in the pathogenesis of several diseases, such as atherosclerosis, CHD, stroke, diabetes mellitus, Parkinson’s disease, renal disease, AIDS and cancer. Increased oxidative stress could be defined as a condition where the cellular antioxidant system is incompetent to neutralize ROS completely due to an excessive ROS formation and/or weakening of the antioxidative defense system. The influence of ROS may be harmful to virtually all biomolecular species, it leads to structure/function changes and even to necrotic or apoptotic cell death. The most relevant and deleterious ROS originated in biosystems include the superoxide anion radical (O. 2 ), hydroxyl radical (HO$), nitric oxide (NO$) and lipid radicals (Table 5.2). Hydrogen peroxide (H2O2), peroxynitrite (ONOO) and hypochlorous acid (HOCl) are not free radicals per se, but they are also involved in free radical reactions and thus contribute to oxidative stress. Deleterious effects of free radical reactions are based on the fact that generation of one ROS may lead to the production of
Chapter j 5
PON1 and HDL
87
FIGURE 5.3 Distribution plots of PON1192QQ (A), PON1192QR (B) and PON1192RR (C) phenotypes in controls and CHD patients. The PON1Q192R phenotype was predicted after examination of the two-dimensional plot of diazoxon vs. paraoxon hydrolysis rates and also by calculating the DZOase/POase activity ratio. PON1Q192R phenotyping was also performed using the anti-mode of the histogram of the DZOase/POase activity ratio (Kotur-Stevuljevic et al., 2006).
88
The HDL Handbook
TABLE 5.2
Reactive oxygen species in the vascular wall
Reactive oxygen species Superoxide anion radical
(O$ 2 )
Sources in the vascular wall NADPH oxidase
Hydroxyl radical (HO$)
Xanthine oxidase
Hydrogen peroxide (H2O2)
Uncoupled nitric oxide synthase (eNOS)
Nitric oxide (NO)
Cyclooxygenase
Peroxynitrite (ONOO )
Lipoxygenase
Hypochlorous acid (HOCl)
Mitochondrial respiratory chain
several others via radical chain reactions. This was particularly evidenced in the case of peroxidative reactions on cell membrane’s polyunsaturated fatty acids. Their consequences are reflected as changes in cellular function, including cell membrane and receptor dysfunction. Moreover, membrane-lipid peroxidation releases chemically reactive and toxic products, such as malondialehyde (MDA), 4-hydroxy-2-nonenal (HNE), 2-propenal (acrolein) and isoprostanes (Tarpey et al., 2004). Above-mentioned lipid peroxidation products are indicated as “secondary cytotoxic messengers” since they are more stable compared to basic free radicals. In addition, they can move far from the site of origin and damage distant cellular targets (Clempus and Griendling, 2006). Several studies have consistently reported elevated oxidative stress (Bridges et al., 1992; Jung et al., 2004), as well as decreased antioxidative defense mechanisms (Wang et al., 1998; Landmesser et al., 2000) in patients with CAD. Dyslipidemia is a traditional risk factor for atherosclerosis development, closely associated with increased endothelial production of ROS (Rizzo et al., 2009). Clinical studies have documented strong positive associations between plasma levels of oxidative stress parameters and atherogenic lipoproteins in patients with CVD (Rizzo et al., 2009). In addition, we have also reported that plasma levels of O. 2 and MDA are positively associated with markers of inflammation (Kotur-Stevuljevic et al., 2007). Similar results were published by Rankinen et al. (2000). These results also suggest that the burden of oxidative stress cannot reflect the extent of CAD (Kotur-Stevuljevic et al., 2007); by contrast, a recent finding that C-reactive protein concentration was associated with both the presence and the extent of CAD (Memon et al., 2006) highlights the pivotal role of inflammation in all phases during CAD development and progression. In order to clarify the connection between PON1, oxidative stress and inflammation we performed additional studies. Oxidative stress status parameters and inflammatory markers were compared between PON1Q192R
Chapter j 5
PON1 and HDL
89
phenotype subgroups of CHD patients (Kotur-Stevuljevic et al., 2008). Among the CHD patients, the carriers of the PON1192RR phenotype exhibited significantly higher oxidative stress and reduced antioxidative defense parameters in comparison with the PON1192QQ phenotype group. The concentrations of inflammatory markers did not vary between the PON1 phenotypes. Thus, we concluded that our CHD patients were in a state of exacerbated oxidative stress which is compellingly connected with inflammation (Kotur-Stevuljevic et al., 2007) and mostly evident in the PON1192RR phenotype group. This is in agreement with the findings from previous studies that the PON1192QQ phenotype had better antioxidative capability (Mackness et al., 1999; Aviram et al., 2000). Any consideration of the relationship between PON1 and HDL in conditions such as CVD needs to include a variety of possible interactions in order to elucidate every aspect of it. CVD patients usually have decreased HDL-C concentration which implicates a less efficient anti-atherogenic defense system. In our recent study, the HDL-C concentration was one of the parameters in the model with the greatest discriminatory ability for CVD (Bogavac-Stanojevic et al., 2007). The inter-relationships among HDL particle size, its lipid content and its function are not still fully defined, but it is likely that the anti-atherogenic action of HDL reflects functional and biological properties of its subclasses rather then the absolute HDL-C plasma level (Stein and Stein, 1999). In addition, there is a growing body of evidence suggesting that smaller, denser HDL particles might be positively correlated with CVD (Rosenson et al., 2002; Rosenson, 2004; Asztalos et al., 2004, 2005). Our own results from a case-control study also confirmed that CVD patients had significantly reduced HDL size and HDL subclass distribution markedly shifted towards smaller, denser HDL3 particles when compared to healthy controls (Zeljkovic et al., 2008). Furthermore, our previous study in apparently healthy middle-aged subjects showed a tendency towards smaller HDL size and greater prevalence of HDL3 particles among the individuals with higher CHD risk (Vekic et al., 2007a). This provokes the question of why these particular HDL subclasses fail to provide adequate anti-atherogenic defense? The most likely explanation is that smaller HDL particles contain less cholesterol, implicating less effective RCT (Rosenson, 2004). However, knowing that oxidation is central to the initiation and propagation of atherosclerosis, one of the possible reasons for this anti-atherogenic defectiveness could also be the diminished antioxidative capacity of HDL3 particles. As it was previously mentioned, it appears that PON1 is not evenly distributed within the total HDL population implicating that both the shape and size of HDL particles are critical for PON1 binding. Based on the findings in asymptomatic individuals, we raised the possibility that reduced HDL size might be associated with lack of PON1 activity and that the PON1192RR phenotype could be related to the formation of smaller, denser HDL particles (Vekic et al., 2007b).
90
The HDL Handbook
When considering PON1 activity in specific clinical conditions such as CVD one should not forget that HDL particles themselves are susceptible to structural and functional modifications. Progression of CVD is a complex process involving oxidative stress, inflammation and dyslipidemia. Therefore, each contributory process can lead to alterations in HDL structure and function (Vekic et al., 2008). Thus, it is possible that it is not the size, but a conformational change within HDL that could be the cause of less preferable binding of PON1. The most common lipid disorders associated with CVD are hypercholesterolemia and combined hyperlipidemia characterized by a concomitant increase in cholesterol and triglyceride concentrations, conditions usually associated with low LCAT and high CETP activities (Jia et al., 2007). As a result, HDL particles become enriched with triglycerides. Such a change of HDL composition, and subsequently its conformation, could be a possible reason for altered PON1 binding. In addition, one must not neglect the fact that HDL particles themselves are susceptible to oxidative modifications. A recent study showed that oxidized HDL can paradoxically enhance LDL oxidation as a result of the changes in its enzyme and protein components (Ansell et al., 2005). Thus, it is possible that a decrease in PON1 activity might be a result of HDL oxidative modification. Another point raised in previous investigations was the eventual replacement of PON1 with serum amyloid A (SAA) in conditions associated with progression of CVD (Van Lenten et al., 2001). Since SAA is preferably associated with smaller, denser HDL particles this can also be a potential explanation for the detected reduction of PON1 activity within individuals having predominantly HDL3 subclasses.
PON1 and HDL in diabetes Numerous studies have confirmed that patients with type 2 diabetes mellitus are in an exacerbated state of oxidative stress (Nourooz-Zadeh et al., 1995; Buyukkocak et al., 2000; Stefanovic et al., 2007). In addition, Rabini et al. (1994) found that LDL particles in diabetic patients were more susceptible to oxidation. It is well known that PON1 is an important determinant of the capacity of HDL to prevent oxidative modifications of LDL particles. Patients with type 2 diabetes mellitus have markedly reduced serum PON1 activity (Stefanovic et al., 2009a), despite the fact that their HDL-C concentration is within the recommended range (Mackness et al., 1991b). However, the cause of a lower PON1 activity in diabetes is not completely understood. In vitro experiments involving the incubation of human HDL in the presence of 25 mmol/L glucose has been shown to result in a significant decrease in PON1 activity (Ferretti et al., 2004). This suggests the possibility that glycation of PON1 could compromise its enzyme activity in diabetes. However, under the experimental conditions, no significant correlations were found between the Hb A1C level and PON1 activity (Abbott et al., 1995). This finding suggests that long-term glycemic control is not directly involved in this process. On the other
Chapter j 5
PON1 and HDL
91
hand, lower PON1 activity could be the consequence of altered HDL composition which is usually observed in diabetics. Abbot et al. (1995) suggested that the conformation of PON1 within the hydrophobic environment of HDL may be crucial for its enzymatic activity. In type 2 diabetics, HDL particles are enriched in triglycerides and deficient in both cholesterol and apoA-I (Ferretti et al., 2001). An altered HDL composition could therefore affect the binding of PON1 to HDL and thereby change the availability of the substrates to the hydrophobic region of HDL (Biesbroeck et al., 1982). In addition, Ferretti et al. (2001) confirmed that glycation of HDL is accompanied by increased lipid peroxidation. This suggests that glycation clearly impairs the antioxidative potential of HDL. Abbot et al. (1995) examined the possible influence of PON1 genotypic variations on decreased PON1 activity in diabetics. However, no significant differences in PON1 phenotypes between healthy subjects and type 2 diabetics were found. We examined the association between PON1 activity and HDL subfractions in type 2 diabetics (Stefanovic et al., 2009b). We found that the patients had significantly lower PON1 activity than controls, even when no difference in HDL-C concentrations was evident. We also evaluated the PON1Q192R polymorphism in both groups; however, we failed to find differences in PON1Q192R phenotype distributions between controls and patients. This finding, together with results from the field of CVD pathology, suggests that the connection between PON1 and disease development requires more than just phenotyping. Gowri et al. (1999) studied the ability of compositionally abnormal HDL particles from type 2 diabetics to protect LDL particles from oxidation. By comparing the antioxidative effects of HDL (by measuring the activities of PAF-AH and PON1 in HDL particles) from control and diabetics, they found that HDL 2 subclasses from diabetics exhibited a reduced potential to limit LDL oxidation. In contrast, no differences between HDL3 fractions from control and diabetics were observed. Furthermore, HDL2 particles from control subjects had a significantly greater potential to limit LDL oxidation than HDL3 particles. This was not the case in diabetics. Such findings suggest that HDL2 particles are more potent inhibitors of LDL oxidation but they probably lose their efficiency in diabetics. In contrast, Valabhji et al. (2001) found that smaller HDL particles had higher PON1 activity and more potent antioxidative effects in diabetics. However, the limitation of the study was that HDL was fractionated by ultracentrifugation which may alter PON1 distribution within HDL. Our recent results appear to provide in vivo support for such a conclusion (Stefanovic et al., 2009b). We compared PON1 activity between type 2 diabetics and controls, both with predominantly larger HDL2 particles. As we expected PON1 activity was significantly higher in controls. Such a difference was not observed between the groups with predominantly smaller HDL 3 subclasses. Furthermore, both patients and controls with more HDL3 subclasses had lower PON1 activity. Therefore, we concluded that reduced PON1 activity
92
The HDL Handbook
and altered HDL composition might also contribute to the acceleration of atherosclerosis in type 2 diabetics.
CONCLUSION It is clear that the benefit of a high HDL-C concentration originates from the different anti-atherogenic properties of HDL particles. Certainly, one of the most important roles of HDL is PON1-mediated antioxidative protection. There is a growing body of evidence suggesting that the structure and composition of HDL particles are extremely important for PON1 activity. The results of numerous studies (discussed above) indicate that the antioxidative properties of HDL can be compromised under conditions associated with accelerated atherosclerosis. However, researchers have not precisely determined how the alterations in HDL structure and composition affect its antioxidative capacity. Future research specifically directed to understand these basic processes will hopefully yield routes for new therapeutic strategies for the prevention and treatment of conditions including atherosclerosis.
ACKNOWLEDGMENT This work was supported by a grant from the Ministry of Science and Technological Development, Republic of Serbia (Project No. 145036B) and COST B35 Action.
REFERENCES Abbott, C. A., Mackness, M. I., Kumar, S., Boulton, A. J., & Durrington, P. N. (1995). Serum paraoxonase activity, concentration, and phenotype distribution in diabetes mellitus and its relationship to serum lipids and lipoproteins. Arterioscler Thromb Vasc Biol, 15, 1812e1818. Akhmedova, S. N., Yakimovsky, A. K., & Schwartz, E. I. (2001). Paraoxonase 1 MeteLeu 54 polymorphism is associated with Parkinson’s disease. J Neurol Sci, 184, 179e182. Ansell, B. J., Watson, K. E., Fogelman, A. M., Navab, M., & Fonarow, G. C. (2005). High-density lipoprotein function: recent advances. J Am Coll Cardiol, 46, 1792e1798. Antikainen, M., Murtoma¨ki, S., Syva¨nne, M., et al. (1996). The GlneArg 191 polymorphism of the human paraoxonase gene (HUMPONA) is not associated with the risk of coronary artery disease in Finns. J Clin Invest, 98, 883e885. Arrol, S., Mackness, M. I., & Durrington, P. N. (1996). High-density lipoprotein associated enzymes and the prevention of low-density lipoprotein oxidation. Eur J Lab Med, 4, 33e38. Asztalos, B. F., Cupples, L. A., Demissie, S., et al. (2004). High-density lipoprotein subpopulation profile and coronary heart disease prevalence in male participants of the Framingham Offspring Study. Arterioscler Thromb Vasc Biol, 24, 2181e2187. Asztalos, B. F., Collins, D., Cupples, L. A., et al. (2005). Value of high-density lipoprotein (HDL) subpopulations in predicting recurrent cardiovascular events in the Veterans Affairs HDL Intervention Trial. Arterioscler Thromb Vasc Biol, 25, 2185e2191. Aviram, M., Hardak, E., Vaya, J., et al. (2000). Human serum paraoxonases (PON1) Q and R selectively decrease lipid peroxides in human coronary and carotid atherosclerotic lesions: PON1 esterase and peroxidase-like activities. Circulation, 101, 2510e2517.
Chapter j 5
PON1 and HDL
93
Bergmeier, C., Siekmeier, R., & Gross, W. (2004). Distribution spectrum of paraoxonase activity in HDL fractions. Clin Chem, 50, 2309e2315. Biesbroeck, R. C., Albers, J. J., Wahl, P. W., Weinberg, C. R., Basset, M. L., & Bierman, E. L. (1982). Abnormal composition of high density lipoproteins in non-insulin-dependent diabetics. Diabetes, 31, 126e131. Blanche, P. J., Gong, E. L., Forte, T. M., & Nichols, A. V. (1981). Characterization of human high density lipoproteins by gradient gel electrophoresis. Biochim Biophys Acta, 665, 408e419. Blatter, M. C., James, R. W., Messmer, S., Barja, F., & Pometta, D. (1993). Identification of distinct human high-density lipoprotein subspecies defined by a lipoprotein-associated protein, K-45. Identity of K-45 with paraoxonase. Eur J Biochem, 211, 871e879. Blatter Garin, M. C., Moren, X., & James, R. W. (2006). Paraoxonase-1 and serum concentrations of HDL-cholesterol and apoA-I. J Lipid Res, 47, 515e520. Bogavac-Stanojevic, N., Jelic-Ivanovic, Z., Djurovic, S., et al (2005). Lack of association between low HDL-cholesterol and elevated circulating cellular adhesion molecules in normolipidemic CAD patients and healthy subjects. Int Heart J, 46, 593e600. Bogavac-Stanojevic, N., Jelic-Ivanovic, Z., Spasojevic-Kalimanovska, V., Spasic, S., & Kalimanovska-Ostric, D. (2007). Lipid and inflammatory markers for the prediction of coronary artery disease: a multi-marker approach. Clin Bioch, 40, 1000e1006. Bridges, A. B., Scott, N. A., Pringle, T. H., McNeill, G. P., & Belch, J. (1992). Relationship between the extent of coronary artery disease and indicators of free radical activity. Clin Cardiol, 15, 169e174. Brophy, V. H., Jarvik, G. P., & Furlong, C. E. (2002). PON1 polymorphisms. In L. G. Costa, & C. E. Furlong (Eds.), Paraoxonase in health and disease: basic and clinical aspects (pp. 53e57). Norwell, MA: Kluwer Acad. Buyukkocak, S., Ozturk, H. S., Tamer, M. N., Kacmaz, M., Cimen, M. Y., & Durak, I. (2000). Erythrocyte oxidant/antioxidant status of diabetic patients. J Endocrinol Invest, 23, 228e230. Cascorbi, I., Laule, M., Mrozikiewicz, P. M., et al. (1999). Mutations in the human paraoxonase 1 gene: frequencies, allelic linkages, and association with coronary artery disease. Pharmacogenetics, 9, 755e761. Chapman, M. J., Goldstein, S., Lagrange, D., & Laplaud, P. M. (1981). A density gradient ultracentrifugal procedure for the isolation of the major lipoprotein classes from human serum. J Lipid Res, 22, 339e358. Cheung, M. C., & Albers, J. J. (1982). Distribution of high density lipoprotein particles with different apoprotein composition: particles with A-I and A-II and particles with A-I but no A-II. J Lipid Res, 23, 747e753. Cheung, M. C., & Wolf, A. C. (1988). Differential effect of ultracentrifugation on apolipoprotein A-I-containing lipoprotein subpopulations. J Lipid Res, 29, 15e25. Clempus, R. E., & Griendling, K. K. (2006). Reactive oxygen species signalling in vascular smooth muscle cells. Cardiovasc Res, 71, 216e225. Cockerill, G. W., Huehns, T. Y., Weerasinghe, A., et al. (2001). Elevation of plasma high-density lipoprotein concentration reduces interleukin-1-induced expression of E-selectin in an in vivo model of acute inflammation. Circulation, 103, 108e112. Cominacini, L., Garbin, U., Pasini, A. F., et al. (1997). Antioxidants inhibit the expression of intercellular cell adhesion molecule-1 and vascular cell adhesion molecule-1 induced by oxidized LDL on human umbilical vein endothelial cells. Free Radic Biol Med, 22, 117e127. Draganov, D. I., Teiber, J. F., Speelman, A., Osawa, Y., Sunahara, R., & La Du, B. N. (2005). Human paraoxonases (PON1, PON2, and PON3) are lactonases with overlapping and distinct substrate specificities. J Lipid Res, 46, 1239e1247.
94
The HDL Handbook
Durrington, P. N., Mackness, B., & Mackness, M. I. (2001). Paraoxonase and atherosclerosis. Arterioscler Thromb Vasc Biol, 21, 473e480. Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults. (2001). Executive summary of the third report of the National Cholesterol Education Program (NCEP) expert panel on detection, evaluation, and treatment of high blood cholesterol in adults (Adult Treatment Panel III). J Am Med Assoc, 285, 2486e2497. Ferre´, N., Camps, J., Ferna´ndez-Ballart, J., Arija, V., et al. (2003). Regulation of serum paraoxonase activity by genetic, nutritional, and lifestyle factors in the general population. Clin Chem, 49, 1491e1497. Ferretti, G., Bacchetti, T., Busni, D., Rabini, R. A., & Curatola, G. (2004). Protective effect of paraoxonase activity in high-density lipoproteins against erythrocyte membranes peroxidation: a comparison between healthy subjects and type 1 diabetic patients. J Clin Endocrinol Metab, 89, 2957e2962. Ferretti, G., Bacchetti, T., Marchionni, C., Caldarelli, L., & Curatola, G. (2001). Effect of glycation of high density lipoproteins on their physicochemical properties and on paraoxonase activity. Acta Diabetol, 38, 163e169. Gowri, M. S., Van der Westhuyzen, D. R., Bridges, S. R., & Anderson, J. W. (1999). Decreased protection by HDL from poorly controlled type 2 diabetic subjects against LDL oxidation may be due to the abnormal composition of HDL. Arterioscler Thromb Vasc Biol, 19, 2226e2233. Griffin, J. H., Kojima, K., Banka, C. L., Curtiss, L. K., & Fernandez, J. A. (1999). High-density lipoprotein enhancement of anticoagulant activities of plasma protein S and activated protein C. J Clin Invest, 103, 219e227. Hedrick, C. C., Thorpe, S. R., Fu, M. X., et al. (2000). Glycation impairs high-density lipoprotein function. Diabetologia, 43, 312e320. Heermeier, K., Leicht, W., Palmetshofer, A., Ullrich, M., Wanner, C., & Galle, J. (2001). Oxidized LDL suppresses NF-kappaB and overcomes protection from apoptosis in activated endothelial cells. J Am Soc Nephrol, 12, 456e463. Humbert, R., Adler, D. A., Disteche, C. M., Hasset, C., Omiecinski, C. J., & Furlong, C. E. (1993). The molecular basis of the human serum paraoxonase activity polymorphism. Nat Genet, 3, 73e76. Ishii, H., Kizaki, K., Horie, S., & Kazama, M. (1996). Oxidized low density lipoprotein reduces thrombomodulin transcription in cultured human endothelial cells through degradation of the lipoprotein in lysosomes. J Biol Chem, 271, 8458e8465. James, R. W., & Deakin, S. P. (2004). The importance of high-density lipoproteins for paraoxonase-1 secretion, stability, and activity. Free Radic Biol Med, 37, 1986e1994. James, R. W., Leviev, I., Ruiz, J., Passa, P., Froguel, P., & Garin, M. C. (2000). Promoter polymorphism T(-107)C of the paraoxonase PON1 gene is a risk factor for coronary heart disease in type 2 diabetic patients. Diabetes, 49, 1390e1393. Jarvik, G. P., Hatsukami, T. S., Carlson, C., et al. (2003). Paraoxonase activity, but not haplotype utilizing the linkage disequilibrium structure, predicts vascular disease. Arterioscler Thromb Vasc Biol, 23, 1465e1471. Jia, L., Fu, M., Tian, Y., et al. (2007). Alterations of high-density lipoprotein subclasses in hypercholesterolemia and combined hyperlipidemia. Int J Cardiol, 120, 331e337. Jonas, A. (1991). Lecithin-cholesterol acyltransferase in the metabolism of high-density lipoproteins. Biochim Biophys Acta, 1084, 205e220. Jung, H. H., Choi, D. H., & Lee, S. H. (2004). Serum malondialdehyde and coronary artery disease in hemodialisis patients. Am J Nephrol, 24, 537e542. Kleemola, P., Freese, R., Jauhiainen, M., Pahlman, R., Alfthan, G., & Mutanen, M. (2002). Dietary determinants of serum paraoxonase activity in healthy humans. Atherosclerosis, 160, 425e432.
Chapter j 5
PON1 and HDL
95
Kontush, A., Chantepie, S., & Chapman, M. J. (2003). Small, dense HDL particles exert potent protection of atherogenic LDL against oxidative stress. Arterioscler Thromb Vasc Biol, 23, 1881e1888. Kotur-Stevuljevic, J., Memon, L., Stefanovic, A., et al. (2007). Correlation of oxidative stress parameters and inflammatory markers in coronary artery disease patients. Clin Biochem, 40, 181e187. Kotur-Stevuljevic, J., Spasic, S., Jelic-Ivanovic, Z., et al. (2008). PON1 status is influenced by oxidative stress and inflammation in coronary heart disease patients. Clin Biochem, 41, 1067e1073. Kotur-Stevuljevic, J., Spasic, S., Stefanovic, A., et al. (2006). Paraoxonase-1 (PON) activity, but not PON1Q192R phenotype, is a predictor of coronary artery disease in a middle-aged Serbian population. Clin Chem Lab Med, 44, 1206e1213. Kunitake, S. T., & Kane, J. P. (1982). Factors affecting the integrity of high density lipoproteins in the ultracentrifuge. J Lipid Res, 23, 936e940. Kunitake, S. T., La Sala, K. J., & Kane, J. P. (1985). Apolipoprotein A-I-containing lipoproteins with pre-beta electrophoretic mobility. J Lipid Res, 26, 549e555. Kuvin, J. T., Ramet, M. E., Patel, A. R., Pandian, N. G., Mendelsohn, M. E., & Karas, R. H. (2002). A novel mechanism for the beneficial vascular effects of high-density lipoprotein cholesterol: enhanced vasorelaxation and increased endothelial nitric oxide synthase expression. Am Heart J, 144, 165e172. La Du, B. N. (2003). Future studies of low-activity PON1 phenotype subjects may reveal how PON1 protects against cardiovascular disease. Arterioscler Thromb Vasc Biol, 23, 1317e1318. Landmesser, U., Merten, R., Spiekermann, S., Buttner, K., Drexler, H., & Hornig, B. (2000). Vascular extracellular superoxide dismutase activity in patients with coronary artery disease: relation to endothelium-dependent vasodilation. Circulation, 101, 2264e2270. Lerch, P. G., Spycher, M. O., & Doran, J. E. (1998). Reconstituted high density lipoprotein (rHDL) modulates platelet activity in vitro and ex vivo. Thromb Haemost, 80, 316e320. Lewis, G. F., & Rader, D. J. (2005). New insights into the regulation of HDL metabolism and reverse cholesterol transport. Circ Res, 96, 1221e1232. Ma, J., Liao, X. L., Lou, B., & Wu, M. P. (2004). Role of apolipoprotein A-I in protecting against endotoxin toxicity. Acta Biochim Biophys Sin, 36, 419e424. Mackness, M. I., Arrol, S., & Durrington, P. N. (1991a). Paraoxonase prevents accumulation of lipoperoxides in low-density lipoprotein. FEBS Lett, 286, 152e154. Mackness, M. I., Harty, D., Bhatnagar, D., et al. (1991b). Serum paraoxonase activity in familial hypercholesterolaemia and insulin-dependent diabetes mellitus. Atherosclerosis, 86, 193e199. Mackness, B., Mackness, M. I., Arrol, S., Turkie, W., & Durrington, P. N. (1998). Effect of the human serum paraoxonase 55 and 192 genetic polymorphisms on the protection by high density lipoprotein against low density lipoprotein oxidative modification. FEBS Lett, 423, 57e60. Mackness, B., Durrington, P. N., & Mackness, M. I. (1999). Polymorphisms of paraoxonase genes and low-density lipoprotein lipid peroxidation. Lancet, 353, 468e469. Mackness, B., Durrington, P. N., & Mackness, M. I. (2000). Low paraoxonase in Persian Gulf War Veterans self-reporting Gulf War Syndrome. Biochem Biophys Res Commun, 276, 729e733. Mackness, B., Durrington, P., McElduff, P., et al. (2003). Low paraoxonase activity predicts coronary events in the Caerphilly Prospective Study. Circulation, 107, 2775e2779. Memon, L., Spasojevic-Kalimanovska, V., Bogavac-Stanojevic, N., Kalimanovska-Ostric, D., Jelic-Ivanovic, Z., Spasic, S., & Topic, A. (2006). Association of C-reactive protein and fibrinogen with the presence and extent of angiographically verified coronary artery disease. Tohoku J Exp Med, 209, 197e206.
96
The HDL Handbook
Mertens, A., Verhamme, P., Bielicki, J. K., et al. (2003). Increased low-density lipoprotein oxidation and impaired high-density lipoprotein antioxidant defense are associated with increased macrophage homing and atherosclerosis in dyslipidemic obese mice: LCAT gene transfer decreases atherosclerosis. Circulation, 107, 1640e1646. Miller, G. J., & Miller, N. E. (1975). Plasma high-density lipoprotein concentration and development of ischaemic heart-disease. Lancet, 1, 16e19. Ng, C. J., Shih, D. M., Hama, S. Y., Villa, N., Navab, M., & Reddy, S. T. (2005). The paraoxonase gene family and atherosclerosis. Free Radic Biol Med, 38, 153e163. Nicholls, S. J., Cutri, B., Worthley, S. G., et al. (2005). Impact of short-term administration of high-density lipoproteins and atorvastatin on atherosclerosis in rabbits. Arterioscler Thromb Vasc Biol, 25, 2416e2421. Nofer, J. R., Levkau, B., Wolinska, I., et al. (2001). Suppression of endothelial cell apoptosis by high density lipoproteins (HDL) and HDL-associated lysosphingolipids. J Biol Chem, 276, 34480e34485. Nourooz-Zadeh, J., Tajaddini-Sarmadi, J., McCarthy, S., Betteridge, D. J., & Wolff, S. P. (1995). Elevated levels of authentic plasma hydroperoxides in NIDDM. Diabetes, 44, 1054e1058. Ombres, D., Pannitteri, G., Montali, A., et al. (1998). The GlneArg 192 polymorphism of human paraoxonase gene is not associated with coronary artery disease in Italian patients. Arterioscler Thromb Vasc Biol, 18, 1611e1616. Rabini, R. A., Fumelli, P., Galassi, R., et al. (1994). Increased susceptibility to lipid oxidation of low-density lipoproteins and erythrocyte membranes from diabetic patients. Metabolism, 43, 1470e1474. Rainwater, D. L., Moore, P. H., Shelledy, W. R., Dyer, T. D., & Slifer, S. H. (1997). Characterization of a composite gradient gel for the electrophoretic separation of lipoproteins. J Lipid Res, 38, 1261e1266. Rankinen, T., Hietanen, E., Va¨isa¨nen, S., et al. (2000). Relationship between lipid peroxidation and plasma fibrinogen in middle-aged men. Thromb Res, 99, 453e459. Rice, G., Ossei-Gerning, N., Stickland, M. H., & Grant, P. J. (1997). The paraoxonase GlneArg 192 polymorphism in subjects with ischemic heart disease. Coron Artery Dis, 8, 677e682. Richter, R. J., & Furlong, C. E. (1999). Determination of paraoxonase (PON1) status requires more than genotyping. Pharmacogenetics, 9, 745e753. Richter, R. J., Jampsa, R. L., Jarvik, G. P., Costa, L. G., & Furlong, C. E. (2003). Determination of paraoxonase 1 status and genotypes at specific polymorphic sites. In M. D. Maines, L. G. Costa, E. Hodgson, D. J. Reed, & I. G. Sipes (Eds.), Current protocols in toxicology (pp. 12e19). John Wiley & Sons. Rizzo, M., Kotur-Stevuljevic, J., Berneis, K., et al. (2009). Atherogenic dyslipidemia and oxidative stress: a new look. Transl Res, 153, 217e223. Rodrigo, L., Mackness, B., Durrington, P. N., Hernandez, A., & Mackness, M. I. (2001). Hydrolisis of platelet-activating factor by human serum paraoxonase. Biochem J, 354, 1e7. Rosenblat, M., Karry, R., & Aviram, M. (2006). Paraoxonase 1 (PON1) is a more potent antioxidant and stimulant of macrophage cholesterol efflux, when present in HDL than in lipoprotein-deficient serum: relevance to diabetes. Atherosclerosis, 187, 74e81. Rosenblat, M., Vaya, J., Shih, D., & Aviram, M. (2005). Paraoxonase 1 (PON1) enchances HDLmediated macrophage cholesterol efflux via the ABCA1 transporter in association with increased HDL binding to the cells: a possible role for lysophosphatidylcholine. Atherosclerosis, 179, 69e77. Rosenson, R. S., Otvos, J. D., & Freedman, D. S. (2002). Relations of lipoprotein subclass levels and low-density lipoprotein size to progression of coronary artery disease in the Pravastatin Limitation of Atherosclerosis in the Coronary Arteries (PLAC-I) trial. Am J Cardiol, 90, 89e94.
Chapter j 5
PON1 and HDL
97
Rosenson, R. S. (2004). Clinical role of LDL and HDL subclasses and apolipoprotein measurement. ACC Curr J Rev, 13, 33e37. Ross, R. (1999). Mechanisms of disease e atherosclerosis: an inflammatory disease. N Engl J Med, 340, 115e126. Ruiz, J., Blanche, H., James, R. W., et al. (1995). GlneArg 192 polymorphism of paraoxonase and coronary heart disease in type 2 diabetes. Lancet, 346, 869e872. Sanghera, D. K., Aston, C. E., Saha, N., & Kamboh, M. I. (1998). DNA polymorphisms in two paraoxonase genes (PON1 and PON2) are associated with the risk of coronary heart disease. Am J Hum Genet, 62, 36e44. Schmidt, H., Schmidt, R., Niederkorn, K., et al. (1998). Paraoxonase PON1 polymorphism LeueMet54 is associated with carotid atherosclerosis: results of the Austrian Stroke Prevention Study. Stroke, 29, 2043e2048. Senti, M., Aubo, C., & Tomas, M. (2000). Differential effects of smoking on myocardial infarction risk according to the Gln/Arg 192 variants of the human paraoxonase gene. Metabolism, 49, 557e559. Serrato, M., & Marian, A. J. (1995). A variant of human paraoxonase / arylesterase (HUMPONA) gene is a risk factor for coronary artery disease. J Clin Invest, 96, 3005e3008. Shih, D. M., Gu, L., Hama, S., et al. (1996). Genetic-dietary regulation of serum paraoxonase expression and its role in atherogenesis in a mouse model. J Clin Invest, 97, 1630e1639. Skinner, E. R. (1994). High-density lipoprotein subclasses. Curr Opin Lipidol, 5, 241e247. Spector, A. A., Scanu, A. M., Kaduce, T. L., Figard, P. H., Fless, G. M., & Czervionke, R. L. (1985). Effect of human plasma lipoproteins on prostacyclin production by cultured endothelial cells. J Lipid Res, 26, 288e297. Stefanovic, A., Kotur-Stevuljevic, J., Spasic, S., Bogavac-Stanojevic, N., & Bujisic, N. (2008). The influence of obesity on the oxidative stress status and the concentration of leptin in diabetes mellitus type 2 patients. Diab Res Clin Pract, 79, 156e163. Stefanovic, A., Kotur-Stevuljevic, J., Spasic, S., Vekic, J., & Bujisic, N. (2009a). Association of oxidative stress and paraoxonase status with PROCAM risk score. Clin Biochem, 42, 617e623. Stefanovic, A., Kotur-Stevuljevic, J., Vekic, J., Zeljkovic, A., & Spasic, S. (2009b). Oxidative stress and antioxidative capacity of HDL particles in diabetes mellitus type 2. Clin Chem Lab Med, 47, S197. Stein, O., & Stein, Y. (1999). Atheroprotective mechanisms of HDL. Atherosclerosis, 144, 285e301. Suehiro, T., Nakauchi, Y., Yamamoto, M., et al. (1996). Paraoxonase gene polymorphism in Japanese subject with coronary heart disease. Int J Cardiol, 57, 69e73. Tarpey, M. M., Wink, D. A., & Grisham, M. B. (2004). Methods for detection of reactive metabolites of oxygen and nitrogen: in vitro and in vivo considerations. Am J Physiol Regul Integr Comp Physiol, 286, 431e444. Valabhji, J., McColl, A. J., Schachter, M., Dhanjil, S., Richmond, W., & Elkeles, R. S. (2001). High-density lipoprotein composition and paraoxonase activity in Type 1 diabetes. Clin Sci, 101, 659e670. Van Lenten, B. J., Wagner, A. C., Nayak, D. P., Hama, S., Navab, M., & Fogelman, A. M. (2001). High-density lipoprotein loses its anti-inflammatory properties during acute influenza infection. Circulation, 103, 2283e2288. Vekic, J., Topic, A., Zeljkovic, A., Jelic-Ivanovic, Z., & Spasojevic-Kalimanovska, V. (2007a). LDL and HDL subclasses and their relationship with Framingham risk score in middle-aged Serbian population. Clin Biochem, 40, 310e316.
98
The HDL Handbook
Vekic, J., Jelic-Ivanovic, Z., Spasojevic-Kalimanovska, V., et al. (2009). High serum uric acid and low-grade inflammation are associated with smaller LDL and HDL particles. Atherosclerosis, 203, 236e242. Vekic, J., Kotur-Stevuljevic, J., Jelic-Ivanovic, Z., Spasic, S., et al. (2007b). Association of oxidative stress and PON1 with LDL and HDL particle size in middle-aged subjects. Eur J Clin Invest, 37, 715e723. Vekic, J., Kotur-Stevuljevic, J., Zeljkovic, A., Stefanovic, A., Bogavac-Stanojevic, N., & Vujovic, A. (2008). Oxidative stress, HDL and atherosclerosis. Anti-Inflammatory Anti-Allergy Agents Med Chem, 7, 158e165. Wang, X. L., Adachi, T., Sim, A. S., & Wilcken, D. E. I. (1998). Plasma extracellular superoxide dismutase levels in an Australian population with coronary artery disease. Arterioscler Thromb Vasc Biol, 18, 1915e1921. Yancey, P. G., Bortnick, A. E., Kellner-Weibel, G., De la Llera-Moya, M., Phillips, M. C., & Rothblat, G. H. (2003). Importance of different pathways of cellular cholesterol efflux. Arterioscler Vasc Biol, 23, 712e719. Zama, T., Murata, M., Matsubara, Y., et al. (1997). A 192Arg variant of the human paraoxonase (HUMPONA) gene polymorphism is associated with an increased risk for coronary artery disease in the Japanese. Arterioscler Thromb Vasc Biol, 17, 3565e3569. Zech, R., Rockseisen, M., Kluge, K., Dewald, K., Armstrong, V. W., & Chemnitius, J. M. (1993). Lipoproteins and hydrolysis of organophosphorus compounds. Chem Biol Interact, 87, 85e94. Zeljkovic, A., Spasojevic-Kalimanovska, V., Vekic, J., et al. (2008). Does simultaneous determination of LDL and HDL particle size improve prediction of coronary artery disease risk? Clin Exp Med, 8, 109e116.
Chapter 6
Paraoxonase-1 and its Interactions with HDL: Molecular Structures of PON1 and HDL Daniel Rochu 1, 2*, Eric Chabrie`re 1, 3 and Patrick Masson 1 1 Institut de Recherche Biome´dicale des Arme´es-Antenne CRSSA, De´partement de Toxicologie, Groupe Bioe´purateurs Catalytiques et Re´activateurs, La Tronche, France, 2 Bundeswehr Institute of Pharmacology and Toxicology, Munich, Germany, 3 Architecture et Fonction des Macromole´cules Biologiques, Groupe Biocristallographie, Biotechnologie et Enzymologie Structurale, Universite´ de la Me´diterrane´e, Marseille, France
INTRODUCTION PON1 as a link between cardiovascular medicine and toxicology Paraoxonase-1 (PON1) is a plasma protein associated with high density lipoprotein (HDL) particles. As the protein was shown to be involved in the protection against atherosclerosis (Lusis, 2000) and has become a player in vascular physiology, a new chapter in the enzyme’s history was opened. Moreover, an HDL-associated phosphate transporter, termed human phosphate binding protein (HPBP), was recently discovered and found to be a partner of human PON1 (Morales et al., 2006). In the absence of its natural environment (or mimicry by detergents), PON1 is highly unstable. A recent review highlighted the importance of the HDL-macromolecular environment for stability and activity of PON1 (James and Deakin, 2004). As a result, biochemical and physiological characterization of this environment allowing the retention of the functional state(s) of enzyme and partners in HDL are mandatory. Researches on PON1 and on HDL are intimately linked. The enzyme was named for its ability to hydrolyze paraoxon, the active metabolite of the organophosphorus (OP) insecticide parathion. Actually, PON1 displays a hydrolase activity toward a broad range of aromatic carboxylic acid esters and lactones, and at The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
99
100
The HDL Handbook
much lower rates toward other OP including nerve agents and pesticides. Now, PON1 is considered as an enzyme with promiscuous functions; its physiological activity is likely to be a lactonase (Aharoni et al., 2004; Khersonsky and Tawfik, 2005). Furthermore, PON1 has an anti-atherogenic activity closely linked to its localization on HDL particles. OPs are among the most toxic substances that have been identified. Although a number of organic phosphorus compounds were synthesized in the 1800s, nerve agents were synthesized in the late 1930s and early 1940s as insecticides. The German chemist Gerhard Schrader (1950) specializing in the discovery of new insecticides, hoping to make progress in the fight against hunger in the world, is credited for the discovery of the general chemical structure of anticholinesterase OPs and for the synthesis of the first commercialized OP insecticide, “Bladan”, and for the most well known, “Schraden” (parathion), in 1944. Following the synthesis of the “trilon group” (tabun in 1936, sarin in 1939 and soman in 1944), the value of OP compounds as potent chemical warfare agents was rapidly documented. Fortunately, during World War II, nerve agents were not used on battlefields, but some of them were stockpiled in large amounts by different countries, and new ones were synthesized during the Cold War. More than one hundred different OPs, commercialized worldwide in a variety of formulations, are currently used as insecticides and drugs (e.g., trichlorfon for schistosomiasis, echothiophate for glaucoma), or as nerve agents in chemical warfare (e.g., sarin or soman). The widespread use and easy accessibility of OPs in agricultural communities result in a huge number of accidental, suicidal and homicidal intoxications leading to several hundred thousands of fatalities annually (Jeyaratnam, 1990; Kwong, 2002; Eddleston et al., 2008). The acute toxicity of OPs results from inhibition of acetylcholinesterase, leading to generalized cholinergic over-stimulation due to accumulation of acetylcholine in synaptic clefts. Medical countermeasures against OP poisoning suffer from certain deficiencies. Hence, the concept of a biological scavenger has emerged as an alternative approach to overcome these limitations. Bioscavengers are proteins capable of sequestering and neutralizing highly toxic OPs in the blood before they reach biological targets, reducing their concentration to a toxicologically irrelevant level. High doses of stoichiometric scavengers need to be administered for rapid and efficient protection against OP toxicity, while similar protection is expected following administration of much smaller doses of catalytic scavengers that display a turnover with OPs (Lenz et al., 2007). Although having numerous advantages, enzymes appeared only recently as possible catalytic scavengers against highly toxic compounds. As a naturally occurring enzyme present in the plasma of mammals, and playing a role in natural defense against OPs, human PON1 is the most promising catalytic scavenger candidate for pre-treatment and therapy of OP poisoning (Rochu et al., 2007a). Strong arguments in favor of human PON1 are: (1) no immunological response following injections, propitious for a long
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
101
residence time in the bloodstream; (2) a catalytic turnover against sarin and soman similar to that of bacterial phosphotriesterase (PTE), likely improvable through protein engineering; (3) a role of PON1 as protector against OP toxicity in rodents has already been demonstrated (Shih et al., 1998). Directed evolution of PTE showed that only three amino acid changes enhanced the catalytic efficiency for an analog of soman by approximately three orders of magnitude (Hill et al., 2003). Certain rPON1-variants exhibit z12-fold enhanced OPhydrolase activity (Amitai et al., 2006). Such an approach opened new prospects for improving this activity of PON1 (Masson and Rochu, 2009), as well as its other activities related to protection against atherosclerosis.
HDL, as the familiarly dubbed good cholesterol Atherosclerosis, one of the major human diseases, is associated with cholesterol and lipid metabolism. The 1930s were not only the unwelcome period for the synthesis of OPs; they were also fruitful for the study of the metabolism of cholesterol in vivo (Schoenheimer, 1949). In the years following World War II, characterization and biochemical assays for lipoprotein fractions, as well as population studies were initiated. The well-known Ni-Hon-San Study tracking ethnic Japanese as they migrated from Japan to the Hawaiian Islands and to continental USA (Robertson et al., 1977) indicated that coronary heart disease (CHD) death rates tracked with the environmentally induced rise in total serum cholesterol levels. The data from cohort studies, such as that begun in Framingham in 1950, reinforced the relationship between cholesterol and CHD incidence (Kannel et al., 1961). The causal link between high cholesterol and atherosclerosis is now established, a large number of epidemiological studies having confirmed that HDL levels are inversely correlated to risk of atherosclerotic cardiovascular disease (CVD). Functionally, HDL are dynamic molecules involved in reverse cholesterol transport (they are believed to play a key role in the process by promoting the efflux of excess cholesterol from peripheral tissues and returning it to the liver for excretion in the bile), improvement of endothelial function, anti-thrombic activity, inhibition of low density lipoprotein (LDL) oxidation, promotion of atheroma regression, as well as possessing various anti-inflammatory and antioxidative properties. HDL, which are the smallest (7e12 nm diameter) and densest (1.063 < d < 1.25 g/mL) plasma lipoproteins are a heterogeneous collection of particles consisting of proteins and lipids. Approximately one third to one half of the HDL mass consists of highly specialized proteins, termed apolipoproteins (apo-), predominantly apoA-I and apoA-II. Other proteins (apolipoproteins, enzymes, lipid transfer factors) may also be present. The remaining half of the mass of HDL is composed of phospholipids, free and esterified cholesterol, triglycerides, and minor components like liposoluble vitamins and antioxidants. The structural arrangement consists of a hydrophobic core (mainly cholesteryl esters plus a small amount of triglycerides and
102
The HDL Handbook
unesterified cholesterol) surrounded by a surface monolayer of phospholipids, unesterified cholesterol in which are embedded various proteins. HDL offer a transient or durable transport vehicle for plasma proteins that are involved in plasma lipid metabolism, and a large number of less abundant proteins including PON1. Structurally, HDL are heterogeneous pseudo-micellar proteinelipid complexes varying in density, size and chemical composition that can be categorized by several methods (ultracentrifugation, nuclear magnetic resonance and electrophoresis). Density measurements have been the most widely used method for determining HDL subclasses. In humans, they exist predominantly as two major density species, HDL2 (d ¼ 1.063e1.125 g/mL) and HDL3 (d ¼ 1.125e1.210 g/mL) with diameters ranging from 7 to 12 nm. Minor, but important, subspecies include lipid-poor apoA-I and nascent discoidal particles. Clearly, the relationship between HDL and cardioprotective effects is complex due to the heterogeneity of the HDL lipoprotein class. General features (e.g., concentration, composition, shape and size) of plasma HDL are determined by proteins that influence their biogenesis, remodeling and catabolism. Thus, it is unlikely that the HDL antioxidant function has evolved to protect humans against atheroma, a disease that appears to have been prevalent for less than a century (Herrick, 1912). Finally, the discoveries of the HDL receptor (scavenger receptor class B type I, SR-BI) (Krieger, 1999, 2001) and the ABCA1 (ATP-binding cassette transporter A1) lipid transporter (Brooks-Wilson et al., 1999) provided missing links helpful to a better understanding of the biogenesis and some of the functions of HDL.
PON1 AS A PROMISCUOUS HUMAN PLASMA ENZYME Overall features of human PON1 PON1 is the main means of protection of the nervous system against the neurotoxicity of OPs entering the circulation. Although it was discovered because of its ability to hydrolyze OPs, there are numerous exogenous and endogenous esters, such as arylesters, homocysteine thiolactone (HTL), other lactones, and cyclic carbonates that are hydrolyzed by PON1 (La Du, 1996; Billecke et al., 2000) (Figure 6.1). PON2 and PON3 (the two other members of the PON family) have no significant OP-hydrolase activity. However, all PONs share the ability to hydrolyze aromatic and long-chain aliphatic lactones (including naturally-occurring lactone metabolites), thus, the term “lactonase” may be more appropriate than “paraoxonase”. Finally, PON1 is as an enzyme with promiscuous functions (Aharoni et al., 2005; Khersonsky and Tawfik, 2005), whose physiological substrates are currently under investigation. PON1 is a member of a family of mammalian enzymes that includes PON2 and PON3, which share z60% sequence identity with PON1. The PON1 gene is located on chromosome 7q21.22 with other members of its supergene family. Next to the PON1 gene, there is a gene that codes for a pyruvate dehydrogenase kinase;
Chapter j 6
103
Paraoxonase-1 and its Interactions with HDL
kcat
A
Km
B
HCyT Lactone
Paraoxon
Sarin
γ-Butyrolactone
Soman
HG Acid Lactone
Chlorpyrifos-oxon
Phenylacetate
Dihydrocoumarin
Coumaranone
kcat/Km
C
FIGURE 6.1 Catalytic parameters of human PON1 (Q192) for some substrates. kcat (A) Km (B) and kcat/Km (C) illustrate respectively the turnover, binding affinity and efficiency of the enzyme for lactones (blue), arylester (green) and OP compounds (red).
104
The HDL Handbook
this may explain the linkage of PON genotypes with diabetic glycemic control (Hegele et al., 1997). The catalytic activity of PON1 spans five orders of magnitude between phenylacetate (the best substrate known) and HTL, a metabolite known as a risk factor for atherosclerotic vascular disease. Human PON1 is a z45 kDa calcium-dependent enzyme. Its concentration in plasma (z60 mg/L) varies between individuals by as much as 13-fold. PON1 level is determined by a combination of complex genetic interactions and environmental/dietary factors, giving rise to a 40-fold interindividual variation in activity. The human PON1 gene shows several polymorphisms in promoter and coding regions. Four polymorphisms, PON1C-107T, PON1G-162A, PON1G-824A and PON1G-907C concern the functional promoter region (Leviev and James, 2000; Brophy et al., 2001; Deakin et al., 2003a); the first one is the main genetic determinant of PON1 levels. The coding region contains two polymorphic sites, PON1L55M and PON1Q192R (Adkins et al., 1993; Humbert et al., 1993). The L55M polymorphism influences the expression of plasma PON1 via interaction with the C-107T promoter polymorphism (Brophy et al., 2001). Paraoxon hydrolytic activity is greatest with HDL and with purified PON1 from PON1 192 RR and 55 LL individuals, and least with PON1 192 QQ and 55 MM individuals (Adkins et al., 1993; Davies et al., 1996; Mackness et al., 1997a). Conversely, the capacity of PON1 alloenzymes to protect LDL from oxidation is opposite to paraoxon hydrolytic activity. Thus, PON1 55 MM/192 QQ individuals have HDL and PON1 associated with the greatest protective capacity (Mackness et al., 1997b; Aviram et al., 1998a). These alloenzymes are also most active in hydrolyzing diazoxon, sarin and soman (Davies et al., 1996). There is yet another group of substrates, such as phenyl acetate, chlorpyrifos-oxon, and 2-naphtyl acetate, against which all PON1 alloenzymes have similar hydrolytic activity (Davies et al., 1996). Besides, the lactonase activity of the R192 PON1 variant was shown to be up to two-fold higher than that of Q192 (Khersonsky and Tawfik, 2006a). Age is a major determinent of PON1 activity. PON1 activity is very low at birth, increases until the age of six months and presents a plateau thereafter (Cole et al., 2003). Children two years or more in age have the same arylesterase activity as normal adults (Augustinsson and Barr, 1963). The PON1 192 variant is involved in the definition of the longevity phenotype. A meta-analysis on 11 studies in different populations, including a total number of 5962 subjects, showed that PON1 gene variants at codon 192 impact on the probability of attaining longevity, and subjects carrying RR or QR phenotypes (Rþ carriers) are favored to reaching extreme ages (Lescai et al., 2009).
Structural features of PON1 The crystal structure of human PON1 has not been solved. Yet, knowledge of the 3D structure is essential for understanding catalytic mechanism(s) of human
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
105
PON1 and designing mutants with optimized performances. However, the structure of a mammalian hybrid rPON1 (a synthetic construct issued from shuffling of rabbit, mouse, rat and human PON1 genes expressed in E. coli) has been determined (Harel et al., 2004) (Figure 6.2). The structure shows that the protein is a six-bladed b-propeller. Two calcium atoms are present, one is catalytic calcium, and the other is likely involved in stabilization of the structure. An unexpected phosphate ion is bound to the catalytic calcium. Three a helices located at the top of the propeller are likely involved in anchoring to HDL particles. Accordingly, the deep and hydrophobic enzyme active site must be oriented towards the phospholipid surface of the HDL particle and may accommodate a broad range of substrates. It is noteworthy that the polymorphic residue 192 is located in the wall of the active site. The rPON1 structure corresponds to an enzyme differing by z16% of amino acids from the human enzyme, some of the mutations being in the vicinity of the region involved in substrate binding, and in binding to HDL and partner lipoproteins. Although representative of the overall structure of natural PON1, this structure provides no clue about the OP-hydrolase catalytic mechanism of PON1. Indeed, while the H115-H134 dyad was shown to mediate the lactonase activity of mammalian rPON1 (Rosenblat et al., 2006), H115 is not meant to participate in the OP-hydrolase activity (Khersonsky and Tawfik, 2006b; Hu et al., 2009).
H115
Ca-2
Ca-1 PO4
H134 K192
FIGURE 6.2 Overall structure of rPON1. Cartoon representation of the mammalian hybrid recombinant PON1 variant G2E6 (Harel et al. 2004), showing the six-bladed b-propeller 3D structure, the three a-helices (red) involved in PON1 anchoring to HDL (and/or other lipoproteins), the structural calcium cation (Ca-2, red ball) and the catalytic calcium cation (Ca-1, magenta ball) to which a phosphate anion (sticks) is bound. The catalytic histidine dyad H115-H134 and the position of the R/Q192 polymorphism are indicated (sticks).
106
The HDL Handbook
Thus, all the residues involved in this activity have not yet been identified. Finally, parallel and complete investigations of molecular and catalytic properties of both enzymes have to be carried out. As regards natural PON1, its structural and functional relationships with other HDL-associated proteins are not sufficiently depicted. These relationships must be explored to evaluate possible iatrogenic effects following PON1 injection. Indeed, because PON1 is involved in protection from atherogenesis and CHD, possible dysfunction in physiological activity(ies) of PON1 caused by high concentration of the enzyme in blood must be thoroughly investigated.
HDL MOLECULAR ENVIRONMENT OF PON1 Catalog of the PON1 neighbors in the HDL milieu Recent data indicate that both activity and stability of PON1 are highly dependent on the molecular environment of HDL components (James and Deakin, 2004). A small subset of HDL particles with a complex dynamic structure is the major PON1 acceptor in plasma. Representing roughly 70% of the HDL protein mass, apolipoproteinA-I (apoA-I), a 243 amino acid, 28 kDa single polypeptide lacking glycosylation or disulfide linkages, is the main resident apolipoprotein component of HDL. Structurally, apoA-I contains a 1e43 residue N-terminal globular domain followed by a lipid-binding C-terminal domain (residues 44e243), containing amphipathic a-helices. The helices likely float among the phospholipid molecules with their hydrophobic faces penetrating past the phosphate group to interact with the acyl chains. The highly dynamic molecule is capable of adopting an array of conformations along the reverse cholesterol transport pathway. The relationship between this structural flexibility and the potential for functional flexibility makes understanding the apoA-I structure a critical issue. Several other proteins described below can be associated to PON1-bearing HDL. Platelet-activating factor acetylhydrolase (PAF-AH) hydrolyzes PAF, an active phospholipid involved in inflammation, anaphylaxis, and reproduction. By regulating blood PAF levels, PAF-AH has both anti-inflammatory and anti-allergenic properties. Lecithin: cholesterol acyl transferase (LCAT) converts cholesterol and phosphatidylcholines of HDL into cholesteryl esters and lysophosphatidylcholines; a role in hydrolyzing oxidized polar phospholipids has also been described. Apolipoprotein J (apoJ), also known as clusterin, is a protein present in an HDL subpopulation together with PON1 and apoA-I (Kelso et al., 1994). Its multiple functions include roles in complement regulation, lipid transport, apoptosis, and protection of membranes at fluidetissue interfaces. A recent review described clusterin as a forgotten player in Alzheimer’s disease (Nuutinen et al., 2009). An inverse relationship exists between PON1 activity and apoJ concentration, and not all HDL particles containing PON1 activity contained apoJ. ApoE, a crucial ligand for receptor-mediated uptake of chylomicron
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
107
FIGURE 6.3 Schematic representations for human PON1 in vitro and in vivo. (A) Purified PON1 sample in vitro: a buffer (Tris), a detergent (Triton) and calcium are mandatory. (B) A PON1containing HDL particle in vivo: some of the durably or transiently associated proteins having a propensity to contaminate purified PON1 fractions are shown. Unsolved structures (italicized) are modeled.
remnants and intermediate density lipoproteins, may play a significant role in the biological life of PON1 in blood. Ghrelin, a circulating brainegut peptide with growth-hormone-releasing and appetite-inducing effects, was described to bind to a species of HDL associated with PON1 (Beaumont et al., 2003). Adiponectin, an adipose tissue secreted protein whose plasma concentration is positively correlated with HDL levels, was described as enhancing the hepatic production of PON1 (Bajnok et al., 2008). Glutathione selenoperoxidase (GPX3) is an HDL-associated protein that could be important in protecting the endothelium from the toxic effects of oxidized lipids (Chen et al., 2000). Ceruloplasmin is an HDL-associated acute phase protein (Gutteridge and Stocks, 1981). All these transiently or durably HDL associated proteins e the above-mentioned list is not exhaustive e have been susceptible to contaminate PON1 preparations and thereby to promote erroneous interpretation of the enzyme functions (Figure 6.3). Finally, the issue of contaminating activities requires multiple experimental approaches and replication among laboratories before it can be considered to be solved.
HPBP, the unexpected but essential PON1 partner Research on human PON1 took an unexpected turn when a novel protein was isolated from human plasma by co-purification with PON1. This serendipitous discovery led to the crystal structure of a protein that showed a fold and a binding
108
The HDL Handbook
site similar to the prokaryotic periplasmic phosphate binding protein associated with ABC transmembrane transporters (Luecke and Quiocho, 1990). The ˚ resolution (Figure 6.4), has a site that tightly described structure, solved at 1.9 A binds inorganic phosphate (Pi) with a submicromolar dissociation constant, as the prokaryotic protein, via the “Venus fly trap” motion model. In this model, two domains engulf Pi as they close the binding cleft via a hinge-bend (and twist) (Mao et al., 1982). HPBP is the first identified transporter capable of binding phosphate ions in human plasma. It belongs to the DING proteins, identified in a wide range of eukaryotic organisms, on the basis of conserved DINGGG N-terminal sequences. Although eukaryotic genomes are extensively sequenced and an increasing number of them are complete or almost complete, DING proteins are systematically absent from eukaryotic databases. As a consequence, the complete and exact amino acids sequence of HPBP had to be determined by tandem use of X-ray crystallography and mass spectrometry (Diemer et al., 2008). Evidence that HPBP interacts tightly with PON1 gave rise to various assumptions. The interaction between both proteins could be fortuitous, due to the purification conditions, impinging directly on the function(s) of each protein, or forms part of a complex control mechanism modulating the affinity of the proteins or directing them to a particular location. Their separation is achievable by chromatography on hydroxyapatite (i.e., a sintered form of calcium phosphate) (Renault et al., 2006). The association of PON1 with HPBP was found to be essential for preserving the enzyme active conformation(s) (Rochu et al., 2007b). Even though PON1 can bind on HDL devoid of apoA-I, preferential association of PON1 with apoA-I-containing HDL particles was described in vivo and in vitro (James and Deakin, 2004). Meanwhile, this tendency cannot be considered as essential for all PON1 activities. A reconstituted in vitro system, based on purified lipid components and recombinant proteins expressed in E. coli, showed that OP-hydrolase and arylesterase activities of rPON1 were stimulated two- to five-fold by anchoring onto HDL, regardless of the apolipoprotein content. In contrast, apoA-I stimulated the lactonase activity by 20-fold (Gaidukov and Tawfik, 2005). Thus, due to the observed differences in the effects of partners on lactonase and esterase activities, it is essential to verify whether rPON1 displays properties identical to those of natural PON1. Complementary PON1 behaviors can be caused by functionally associated protein partners such as HPBP. In this context, surrogates of the natural environment of human PON1 could be cautiously considered.
IMPEDIMENTS FOR STUDYING THE PASSENGER PON1 AND THE VEHICLE HDL PON1 as a promiscuous enzyme requiring helpers Structural studies of a protein require purity, but the understanding of its functional behavior needs to take into account the physiological environment.
Chapter j 6
109
Paraoxonase-1 and its Interactions with HDL
A N C
S-S
S-S
B
Asp 61
Gly145
Ser144 Ser32
Thr146
Arg140
Thr8
FIGURE 6.4 Structure of HPBP. (A) Cartoon mode representation showing the two structural domains, the phosphate ion (spheres), and the two disulfide bridges (S-S). N and C represent the protein termini; (B) The hydrogen bond network (dashed lines), between the protein and the phosphate ion.
110
The HDL Handbook
Abundant biochemical, biological and toxicological information has been collected for a decade, leading to partial characterization of the enzyme function (Costa and Furlong, 2002; Mackness et al., 2008), but recurrent attempts at solving the PON1 structure failed. Determination of the structure of a mammalian rPON1 and discovery of HPBP gave momentum for research to aim to understand and overcome the difficulty in obtaining pure PON1. In vivo, HDL offers a hydrophobic harbor and a stabilizing environment for PON1 function(s). The particular HDL environment of multiple interacting lipids and proteins may explain why study of PON1 in solution is difficult. Akin to PON1 that displays multiple activities, HDL particles display long-lasting or transient binding capacity for numerous proteins involved in multiple biological mechanisms (Vaisar et al., 2007). HDL proteomics was described as a pot of gold or Pandora’s box (Reilly and Tall, 2007), and the weight of the HDLassociation for the biological function(s) of PON1 was underscored (James and Deakin, 2004; Gaidukov and Tawfik, 2005). The huge heterogeneity of this vehicle makes it difficult to easily study the physiological role of PON1. While the anti-atherosclerotic activity of PON1 is closely linked to its localization on HDL, the mode of binding of PON1 and other associated proteins to HDL is still unknown. Purification procedures of PON1 derive from those described in 1991 (Furlong et al., 1991; Gan et al., 1991), that assumed to provide PON1 at z95% purity. To solve the 3D structure of human PON1, such preparations were used and produced crystals providing the structure of HPBP (Fokine et al., 2003; Contreras-Martel et al., 2006). The presence of HPBP in allegedly pure PON1 preparations was not disputed, since numerous works substantiated the presence of various contaminants in usual PON1 preparations, and afterwards was found to be responsible for certain catalytic activities previously attributed to PON1 (Teiber et al., 2004; Connelly et al., 2005; Draganov et al., 2005). These results strengthened current literature data indicating that PON1 activities and stability are highly dependent on the molecular environment, and stipulated the importance of developing media that mimic its in vivo milieu. Since HDL are in continuous rearrangement with time, knowledge of partner lipoprotein(s) or/and hydrophobic cofactor(s) able to surrogate the natural stabilizing environment of active PON1 is essential. Optimizing the purification process and the biochemical characterization of PON1 became main concerns. As well, the harsh conditions needed to separate HPBP and PON1 indicated that HPBP is strongly associated with PON1. The presence of a phosphate ion bound to the catalytic calcium in the crystal structure of the rPON1 was another unexpected fact. Both observations argue for a possible functional link between PON1 and HPBP, implying phosphate transfer. Because PON1 is a promising OP-scavenger and, moreover, displays anti-atherogenic properties, it is crucial to establish whether the PON1eHPBP association is physiologically and pharmacologically relevant, and whether HPBP could be involved in phosphatemia-related disorders, including atherosclerosis. Thus, biochemical and physiological characterization of this PON1
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
111
partner is mandatory, as well as the environment stabilizing the enzyme in solution. A modified purification protocol providing highly purified PON1 and pure HPBP (Renault et al., 2006) allowed studies that emphasized the consequence of the molecular environment for the catalytic activity(ies) and stability of PON1. Storage and thermal stability studies provided evidence that the association of PON1 with HPBP is crucial for preserving its active conformation(s) (Rochu et al., 2007b). This supplied data for optimizing the stability of future human rPON1-based mutants with operational catalytic activity against OP and without detrimental effects on anti-atherogenic activity. Finally, scrutiny of PON1 as a nerve agents bioscavenger is not dissociable from the other enzyme activities. This is endorsed by the fact that, without losing sight of its toxicological links, researches shifted progressively to its involvement in vascular disease.
HDL as a multiform and dynamic vehicle HDL particles are heterogeneous in terms of their lipid and apolipoprotein composition, size, and density. This heterogeneity of HDL is critically modulated by abnormalities in HDL metabolism. In association with such metabolic alterations, HDL particles progressively lose normal biological activities and acquire novel properties. This led to the attribution of “chameleon-like” properties to such dysfunctional HDL (Navab et al., 2004). HDL metabolism involves a complex interplay of factors that control their synthesis, intravascular remodeling by lipases, and catabolism. Most HDL apolipoproteins are readily exchanged between plasma lipoproteins, including HDL subspecies. These proteins are in equilibrium with free apolipoproteins present in the plasma, may penetrate cells and organs, and may be cleared separately from the HDL particle. Some diseases, lifestyle and medications can induce low HDL levels, in addition to many genetic causes (mutations in associated proteins, sitosterolemia and Tangier disease) (Assmann et al., 1993). HDL metabolism is complicated and, in fact, HDL metabolism, HDL kinetics, concentration of HDL subclasses, and genetic factors affecting HDL functionality may all contribute to the anti-atherogenic properties of HDL. In this circumstance, HDL proteomics, although being in its infancy, may provide novel insights into HDL physiology and for bioassays of HDL functions. For example, while using shotgun proteomics with stringent criteria, 48 proteins were identified in HDL isolated from apparently healthy control and/or CAD subjects (Table 6.1), a collection of data indicated more than 90 proteins associated to HDL (Reilly and Tall, 2007). Possibly, the most prominent aspect of such early studies is not just the multiplicity of identified HDL proteins and peptides but also the overrepresentation of proteins involved in several nonlipid transport functions, including the acute-phase response, complement regulation, proteolysis, and coagulation, suggesting novel HDL functions. Although persuasive substantiation of functional roles for HDL in these
112
The HDL Handbook
TABLE 6.1 Proteins detected in HDL from plasma of apparently healthy control subjects Protein family
Proteins
Lipid metabolism
ApoA-I, A-II, A-IV, E, C-I, C-II, C-III, C-IV, L-I, M, F, D, H, J (clusterin), Serum amyloid A1, A2, A4, Paraoxonase-1, Paraoxonase-3, Lecithin:cholesterol acyltransferase, Phospho Lipid transfer protein, Lipoprotein lipase, Cholesteryl ester transfer protein
Acute-phase response
a2-HS-Glycoprotein, Haptoglobin-related protein, a1Antitrypsin, Bikunin, Kininogen-1, a1Acid glycoprotein 2, Transthyretin, Inter-a-trypsin inhibitor H4, Inter-a-trypsin inhibitor H2, Retinol binding protein, Transferrin, Fibrinogen, Hemopexin, Serum amyloid A1, A2, ApoA-I, Paraoxonase-1, ApoH, Apo A-IV, Complement component C3, C4A, C4B, C9, Vitronectin
Protease inhibitors
Angiotensin, Serpin peptidase inhibitor clade F-1, a2Antiplasmin, a2-HS-Glycoprotein, Haptoglobin-related protein, a1-Antitrypsin, Bikunin, Kininogen-1
Complement pathway Others
C3, C4A, C4B, C9, Vitronectin, Clusterin Albumin, Vitamin-D binding protein, Phenyl cystein oxidase, ApoB-100, SAA, a2 Antiplasmin, PAF-AH, Ghrelin
From Reilly and Tall, (2007). Proteins participating in different activities are italicized.
pathways is partial, past and emerging studies have shown that HDL and/or apoA-I can attenuate response to experimental endotoxemia, inhibit complement activation, and inhibit platelet activation, serpins, and thrombosis (Hamilton et al., 1993; Wu et al., 2004; Mineo et al., 2006). Discrepancies in the number and identity of HDL-associated proteins is likely related to methodology differences in HDL isolation. Finally, the complex and dynamic behavior of HDL adds a difficulty level to the tricky study of human plasma PON1 functions.
PON1 ACTIVITY IN NORMAL HDL CONDITIONS PON1 activity and HDL status First described in HDL immunoprecipitates after electrophoresis of human serum (Uriel, 1961), and remaining with HDL upon serum centrifugation (Mackness et al., 1985), PON1 activity was regarded as borne by an intrinsic HDL component. Several groups have confirmed and extended the findings that virtually all PON1 activity is associated with HDL-cholesterol, the common assumption being that PON1 is the main beneficiary of the
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
113
association, with HDL governing serum level of the enzyme. A mechanistic justification for this dependency was given by showing the importance of HDL for PON1 release from cells and maintenance of enzyme activity (Deakin et al., 2002). HDL provides a vector that facilitates the enzyme secretion by offering a hydrophobic harbor for the retained signal peptide of PON1 and coincidentally stabilizing the enzyme. The lipoprotein also furnishes the hydrophobic environment important for PON1 function. The circulating form of PON1 retains its hydrophobic signal peptide which is thought to play a key role in targeting PON1 to HDL particles (Sorenson et al., 1999). PON1 will target to phospholipid vesicles in the absence of apolipoprotein, though apoA-I is thought to stabilize this association. The colocalization of PON1 and apoA-I reflects most likely the fact that both proteins prefer a phospholipid-rich environment. PON1 can associate with HDL either because of its surface phospholipids and/or apolipoprotein components, or because it prefers phospholipid surfaces of a particular narrow range of radii of curvature. ApoA-I generates HDL particles of different sizes, and because there are z40 times as many apoA-I molecules as PON1 molecules in the plasma (Kujiraoka et al., 2000), PON1 can only be associated with a small subset of particles (Cabana et al., 2003). Among HDL subfractions, HDL3 carries the highest PON1 activity. To sum up, PON1 selectively associates with HDL or HDL-sized particles, regardless of their specific apolipoprotein composition. In most healthy populations, plasma PON1 activity and concentration are correlated with the HDL and apoA-I concentration. In severe HDL deficiency (e.g., Tangier and Fish eye diseases), PON1 activity is significantly lowered (Mackness et al., 1987, 1989), but not in other deficiencies (James et al., 1998). Undoubtedly, there are major influences on the activity of PON1 in plasma that are concentration-independent of those governing HDL as a whole. The 40-fold interindividual difference in PON1 activity is, for a major part, determined by common PON1 gene polymorphisms. Despite the large impact of PON1 genetic variation on PON1 activity, this variation is inconsistently associated to vascular disease status (Jarvik et al., 2003; Mackness and Mackness, 2004). Meta-analyses have not detected a significant predictive effect for a set of PON1 genotypes that were not attributable to publication bias (Lohmueller et al., 2003; Wheeler et al., 2004). Cholesterol efflux is attenuated with aging due to the change in composition and structure of HDL, including the activity of PON1. The propensity to oxidation of LDL and HDL increases with age. The factors involved in the increased susceptibility to lipid peroxidation have not yet been fully elucidated, although it is thought that PONl is an important player. The antioxidant activity of PONl was shown to decrease significantly with age (Jaouad et al., 2006). However, this decrease was not due to a decrease in concentration or synthesis of PON1 or a reduction in the concentration of circulating HDL (Milochevitch and Khalil, 2001; Seres et al., 2004).
114
The HDL Handbook
PON1 and anti-atherogenic activity of HDL Most of the HDL antioxidant properties are related to their associated enzymes, PON1, PAH-AH, LCAT and GPX3. Mackness et al. (1991a) first showed that PON1 had the capacity to block the accumulation of lipid peroxides in LDL under oxidizing conditions in vitro. While PON1 has been described as capable of hydrolyzing LDL-derived oxidized phospholipids (Rodrigo et al., 2001), recent studies indicated that PAF-AH, rather than PON1, is the oxidized phospholipid hydrolase in HDL (Marathe et al., 2003). The role of PON1 in atherosclerosis was substantiated in studies using knockout mice for PON1, and in mice over-expressing PON1 (Shih et al., 1998, 2000; Oda et al., 2002; Tward et al., 2002; Rozenberg et al., 2003). PON1 deficiency results in increased oxidative stress in serum and macrophages, and HDL isolated from PON1deficient mice did not protect LDL from oxidation, whereas HDL isolated from human PON1 transgenic mice (having 2e4-fold increased PON1 plasma levels) was more protective against LDL oxidation in a dose-dependent manner. All these studies are consistent with a role for PON1 in vascular disease. One of the main difficulties could result from the general use of paraoxon or phenylacetate as the substrate for measuring PON1 activity, in the continued absence of a defined physiological substrate. Thus, some studies reported positive correlations between these PON1 activities and the antioxidant activity of PON1 (Aviram et al., 1998b; Kontush et al., 2003; van Himbergen et al., 2005), while other studies showed no correlation (Cao et al., 1999; Sarandol et al., 2004). The use of rPON1 showed that arylesterase activity best reflected the antioxidant function, even though it was not directly responsible for it. In contrast, OP-hydrolase activity was a relatively poor substitute for the antioxidant activity. The lower PON1 activity was documented in patients with vascular disease or in high-risk populations, such as CHD (Mackness et al., 2001), diabetes (Abbott et al., 1995; Boemi et al., 2001, 2004), familial hypercholesterolemia (Mackness et al., 1991b; Tomas et al., 2000), the metabolic syndrome (Senti et al., 2003; Garin et al., 2005), systemic vasculitis (Que´me´neur et al., 2007), smokers (James et al., 2000; Senti et al., 2000), patients with erectile dysfunction (Ciftci et al., 2007), iron deficiency anemia (Aslan et al., 2007), and infection by Helicobacter pylori (Aslan et al., 2008). Due to the above-mentioned uncertainty about the value of enzyme activities of PON1 as surrogates for antioxidant function, an alternative, such as estimation of the plasma PON1 peptide mass, would be appropriate, since it was shown to be reduced in high-risk patients. The discovery of HPBP led to questions as to what could be the function of this protein in human plasma. Its structure is closely related to the prokaryotes periplasmic phosphate solute binding proteins (SBP) associated with bacterial ABC transporters. SBP in eukaryotes have never been predicted or characterized, despite the fact that eukaryotic ABC transporters obviously exist. The structure of a bacterial ABC transporter and that of an archaeal ABC
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
115
transporter in complex with its cognate binding protein ModA (a tungstate/ molybdate-binding protein whose structure is similar to that of other SBP) were recently solved (Ferretti et al., 2001; Dawson and Locher, 2006). In humans, the discovery of ABCA1 in 1999 provoked intense interest in unraveling the molecular regulation of the early steps of cholesterol transport from cells to HDL. Since then, an impressive body of work has provided us with a clear picture of the ABCA1 role in lipoprotein production, cholesterol adsorption and atherosclerosis. In this context, it was suggested that PON1 may contribute to the attenuation of atherosclerosis development by its ability to act on macrophage phospholipids, to form lysophosphatidylcholine, which in turn, stimulates HDL binding and HDL-mediated macrophage cholesterol efflux via the ABCA1 transporter (Dantoine et al., 2002). For HPBP that only came to light in an unexpected way, we can only speculate on the physiological function of this phosphate binding protein in human plasma. Recent investigations allowed asserting that HPBP is a major actor for maintaining the in vitro functional state, storage and thermal stability of PON1 (Rochu et al., 2007b; Cle´ry-Barraud et al., 2009), and showed that the association of PON1 and HPBP is believed to be physiologically relevant as their oligomerization is modulated by calcium and phosphate concentration (Renault et al., 2010). Even if HPBP physiological involvements are not yet clearly established, its structural relationship with proteins interacting with ABC transporters suggests HPBP could be regarded as a new predictor or as a possible therapeutic agent for phosphate-related disorders, including atherosclerosis (Morales et al., 2006; Diemer et al., 2008). As a result, a synergistic role of the HDL-associated PON1-HPBP duo may be suggested, expanding the promiscuous character of PON1 beyond its enzymatic properties and comforting the resourceful character HDL.
PON1 ACTIVITY IN ALTERED HDL CONDITIONS Due to the key role of HDL as the main vehicle for PON1 transport in human plasma, disorders affecting these particles and their subsequent therapeutics were supposed to interfere with the enzyme activity. In addition to the wide range of atherosclerosis-related disorders for which significantly lower serum PON1 activity is documented (see above), there are other cases of amended PON1 activity due to altered HDL. In diabetes, PON1 activity was mostly found reduced in type 1 and 2 patients. The mechanism by which PON1 is reduced may be associated with an increase in blood glucose concentration. Glycation is suggested to inactivate PON1 and favor lipid peroxidation in HDL (Hedrick et al., 2000; Ferretti et al., 2001). PON1 activity was also found reduced in patients with the metabolic syndrome, symptoms of which include abnormal fasting glucose levels and increased insulin resistance. Oxidative stress is deemed as a risk factor for the development of dementia. PON1 activity is reduced in patients with vascular dementia and Alzheimer’s disease,
116
The HDL Handbook
but it is not established whether this is the cause or the consequence of increased oxidation (Dantoine et al., 2002; Paragh et al., 2002). Chronic renal failure patients frequently have lipoprotein abnormalities. Putative causes for reduced PON1 activity in these patients are reduced HDL concentrations, altered HDL subfraction distribution, lowered PON1 concentration and different PON1 phenotype distribution (Dirican et al., 2004). It was shown that hemodialysis (HD) results in a significant, consistent increase in PON1 activity while there is no significant change in either apoA-I or HDL subclass distribution (Gugliucci et al., 2007), suggesting that HD can remove low molecular weight fractions involved in PON1 inhibition. As regards inflammation, Van Lenten (1995) showed in a rabbit model that as acute-phase HDL-containing serum amyloid A and ceruloplasmin increase, apoA-I, PON1 and PAF-AH decrease severely. These changes could result in the conversion of HDL from an anti-iflammatory/antioxidant complex into a proinflammatory/pro-oxidant one. In HDL level disorders, such as extreme deficiency (the Fish eye and Tangier diseases), depressed PON1 activity could simply be due to the absence of carrier HDL. In infectious diseases, there is some evidence that HDL can exert anti-infectious effects. The ability of plasma to inhibit the pyrogenic activity of bacterial lipopolysaccharide lost after delipidation, was shown to be restored after adding HDL (Ulevitch et al., 1979). Equally of note is the capacity of HDL to induce lysis of the parasite Trypanosoma brucei (Hajduk et al., 1989), and of PON1 to provide protection against trypanosome infection (Bhasin et al., 2006). Another PON1 trait results from its ability to hydrolyze lactones. The pathogenic bacterium Pseudomonas aeruginosa uses acyl-HSL quorum-sensing signals to regulate genes controlling virulence and biofilm formation. PON1 was found to hydrolyze the P. aeruginosa acyl-HSL 3OC12HSL and thus inhibit biofilm formation (Ozer et al., 2005). If PON1 is a naturally occurring inhibitor of quorum-sensing, then modulation of its activity or overexpression may represent a potential therapeutic target (Stoltz et al., 2007). Patients affected by the human immunodeficiency virus (HIV) infection often develop long-term pro-atherogenic metabolic alterations, the phenomenon being explained possibly by the infection itself or by the secondary effects of antiretroviral therapies (Maggi et al., 2000; Depairon et al., 2001). Activity and concentration of PON1 are influenced by HIV-infection, and these are related to alteration in HDL composition and the immunological status of the patients (Parra et al., 2007). In degenerative diseases, alterations of physicochemical properties of HDL could potentially contribute to accelerated atherosclerosis. Thus, antioxidant therapy and agents scavenging reactive oxygen species (ROS) or lipid peroxidation derivatives could be of interest in preventing these diseases. Omura (1980) found significantly lower activity in 236 Japanese with a variety of chronic diseases, and systemic amyloidosis was associated with lower PON1 activity in another study (Maury et al., 1984). The serum PON1 activity was significantly lower in the patients with age-related macular degeneration (AMD) than that in the controls. Increased serum homocysteine
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
117
and malondialdehyde levels may be responsible for the decreased PON1 activity in these patients. The exact mechanism of the decline of the serum PON1 activity in AMD patients is unclear; it may be attributed to an increased inactivation of PON1 due to the increased generation of ROS in AMD patients (Ates et al., 2009). The proximal promoter and 50 sequence of the PON1 gene may harbor unknown functional variant(s) associated with the risk of developing Alzheimer’s disease (AD) (Chapuis et al., 2009). Q192R polymorphism in the PON1 gene was shown to be associated with AD in a Chinese Han ethnic population. The frequency of the PON1 R allele was lower in AD patients than in controls (He et al., 2006). Aging is an irreversible biological process characterized by a general progressive decline in physiological functions and an increased morbidity and mortality. In accordance with the free radical theory proposed by Harman (1956) several decades ago, aging is a progressive, unavoidable process in part related to the growth of oxidative damage to biological molecules. In agreement with this theory is the idea that oxidative damage to nucleic acids, lipids, proteins, and carbohydrates is attributable to a tipping of the pro-oxidant/antioxidant equilibrium toward the pro-oxidant processes (Harman, 1956; Mariani et al., 2005). Substantiation from in vitro and in vivo studies has suggested that the LDL oxidation may contribute to early atherosclerotic lesion formation (Reaven and Witztum, 1996; Napoli et al., 1997). Oxidative stress increases dramatically with age and may play a major role in the observed prevalence of CVD, particularly in the elderly. Atherosclerosis begins early in life, and lipid streaks have been observed even in fetal coronary arteries. In spite of decreased apoA-I, no altered expression of PON1 in serum or association with HDL during aging were found (Jarvik et al., 2002; Seres et al., 2004). Since PON1 expression does not change with aging, the dramatic increase in CVD with aging can only be explained by other changes altering the anti-atherogenic properties of HDL. In fact, there is a dramatic decrease in PON1 activity with aging (Milochevitch and Khalil, 2001; Seres et al., 2004), both arylesterase and paraoxonase activities being reduced (Figure 6.5) (Marchegiani et al., 2006). This PON1 inactivation might result from an interaction of oxLDL associated lipids or with other oxidants (e.g., hydrogen peroxide) with the PON1 free sulfhydryl group at C284. Recently, it was shown that C284 is not part of the PON1 active site, but participates in PON1 stability. Environmental factors can also have an impact on PON1 activity. Patients exposed to OP poisoning, workers exposed to radiation, and Gulf War veterans, who may have been exposed to subtoxic doses of various OPs, all had lower PON1 activity than control groups (Sozmen et al., 2002; Hotopf et al., 2003; Serhatlioglu et al., 2003). Together with yet incompletely understood predisposing genotype/phenotype elements, exposure to agricultural insecticides notably increase the risk of Parkinson’s disease (PD) (Kondo and Yamamoto, 1998). It was thus suggested that inherited interactive weakness of PON1 expression increases the insecticide-induced occurrence of PD (Benmoyal-Segal et al., 2005).
118
A
The HDL Handbook
B150
500
R-
R+
C R-
R+
R-
R+
1,75
125
400
2
1,5 100
1,25
75
200
mmol/L
U/ml
µg/ml
300
50
1 0,75 0,5
100
0
25
OPH
Aryl
OPH
Aryl
0
PON1 concentration
0,25 0
HDL Cholesterol level
FIGURE 6.5 Effect of aging on PON1 activity and PON1 level in three age groups according to PON1 192 genotypes (from Marchegiani et al. 2006). (A) OP-hydrolase (OPH), Arylesterase (Aryl) activities of PON1; (B) PON1 concentration; (C) HDL level in young (22e65 y, blue), elderly (66e89 y, green) and nonagenarians/centenarians (>90 y, red). Genotypes were grouped as Rþ carriers (QRþRR) and R carriers (QQ) for the codon 192 polymorphism.
Finally, PON1 as an HDL-bound promiscuous enzyme is involved in a range of diseases. Thus, PON1 can be considered as a potential therapeutic target for treatment of these diverse illnesses.
PON1 ACTIVITY IN THERAPEUTIC HDL Manipulating HDL and/or PON1 level Positive lifestyle effects: HDL levels may be increased by 10% by implementing therapeutic lifestyle changes including weight reduction, exercise, smoking cessation and moderate alcohol consumption. These changes influence the PON1 in vivo status. Vitamin C and E intake (Jarvik et al., 2002) and polyphenol consumption (Aviram et al., 2000) were shown to increase PON1 activity. Cigarette smoke inhibits PON1 activity in vitro (Nishio and Watanabe, 1997) and, in agreement, PON1 activity is lower in smokers than in nonsmokers (James et al., 2000). Moderate consumption of beer, wine or spirits is associated with increased PON1 activity (Van der Gaag et al.,1999; Sierksma et al., 2002). Nutritional polyphenols, e.g., the pomegranate hydrolyzable tannin punicalagin, the flavonoid quercetin and phytoalexin resveratrol of red wine, the licorice root isoflavan glabridin, were shown to increase expression of the PON1 gene (Aviram et al., 2000, 2004; Goue´dard et al., 2004a,b), as well as usual consumption of virgin olive and argan oil that was described as improving
Chapter j 6
119
Paraoxonase-1 and its Interactions with HDL
PON1 activities and antioxidant status (Cherki et al., 2005). These protective roles in CVD were often referred to as the “French paradox” and the “Mediterranean diet”. Regulation by inflammatory cytokines: lipopolysaccharides, tumor necrosis factor and interleukin-1 were found to cause a fall in PON1 activity in hamsters (Feingold et al., 1998), whereas the reduction in PON1 mRNA by oxidized phospholipids was not seen when IL-6 knockout mice were used, suggesting an effect mediated through IL-6 (Van Lenten et al., 2001). Pharmaceutics for improving HDL protective effects: promising current and emerging HDL-based therapies include the use of statins to lower LDL levels relative to HDL level, as well as raising HDL-cholesterol levels by inhibition of cholesterol ester transfer protein, or by treatment with fibrates or peroxisome proliferative activating receptor agonists, and infusion of HDL mimetics (Table 6.2). Stimulation of the biosynthesis of natural PON1: most studies on PON1 modulation by drugs, mainly focused on statins and fibrates, have yielded conflicting results. A cell culture study showed that fibrates induced PON1 expression and that statins had the opposite effect (Goue´dard et al., 2003). Another in vitro study in HepG2 cells showed statin-induced upregulation of the PON1 gene (Deakin et al., 2003b), a result corroborated in vivo by using simvastatin (Deakin and James, 2004). On the other hand, attempts at combined therapy with simvastatin and ezetimide (a cholesterol-adsorption inhibitor) failed to reduce intima-media thickness of the carotid artery wall in patients with familial hypercholesterolemia (Kastelein et al., 2008). Dietary polyphenols were also shown to increase the expression of PON1 in HuH7 human hepatoma cell line, by stimulating the promoter (Goue´dard et al., 2004a). These data are of clinical relevance in high-risk individuals. Regulation of PON1 promoter by drugs may have potential applications in toxicology, i.e., administration of promoter inducers to increase PON1-based natural defense against
TABLE 6.2
Drugs as modulators of HDL level
Therapy
Effect on HDL
Statins
➔ up to 16%
Fibrates
➔ up to 10%
Nicotinic acid
➔ up to 35%
Thiazolidinediones (PPARg agonists)
➔ up to 20%
ApoA-1Milano, HDL mimetic peptides
transient
CETP inhibitors (JTT-705 and torcetrapib)
➔ from 34 to 106%
120
The HDL Handbook
OPs. Although the most potent inducers are toxic and non-specific, this complementary way has to be explored. The recent report that resveratrol induces PON1 expression and protects against nerve agent toxicity in human cell lines is a striking encouragement (Curtin et al., 2008). Gene therapy: certain Gulf War veterans with neurological symptoms have a low level of Q192 PON1 activity (Haley et al., 1999), and an association between self-reported Gulf War related symptomatology and low levels of PON1 was reported (Mackness et al., 2000). This suggested that genetic polymorphism and PON1 deficiency could have significant clinical consequences for individuals exposed to OPs, and prompted to ask whether gene therapy could be used to challenge OPs. It was also shown that producing rPON1 from a gene delivery vector can increase PON1 level in mice serum, and that the rPON1 can reduce, or even prevent, the entry of an OP into the brain (Cowan et al., 2001). To be used in humans, this strategy needs the approval of a delivery vector allowing transient non-toxic gene expression without subject-to-subject variability. Another approach, using administration of naked PON1 plasmid DNA to mice showed dose-dependent hydrolytic actions on injected soman in the blood, and reduced neurological symptoms (Fu et al., 2005). In addition, facilitated replacement of Kupffer cells expressing a PON1 transgene was shown capable of reducing atherosclerosis signs in mice (Bradshaw et al., 2005). A recent observation suggests that cellular and humoral increases in the PON1 gene are beneficial for inhibiting the development of atherosclerosis. Local delivery of PON1 gene using Sendai virus vector inhibited neonatal hyperplasia after arterial ballooninjury in rabbits fed a high-fat diet (Miyoshi et al., 2007). Besides, the therapeutic potential of human PON1 gene transfer in old apolipoprotein E-deficient (apoE/) mice with advanced atherosclerosis was confirmed (Guns et al., 2008). Thus, PON1 gene therapy could be beneficial for the different functions of the enzyme. This approach could even be attractive using a mutated PON1 gene coding for an enzyme with high OP-hydrolase activity against nerve agents and retaining anti-atherosclerotic capacity. Finally, the HDL complexity makes it apparent that strategies for reducing atherosclerosis by manipulating HDL metabolic pathways must be undertaken cautiously.
PON1 and HDL partners, from atherosclerosis marker to pharmaceutical Homocysteinylation could render HDL less protective against oxidative damage and against toxicity of homocysteine thiolactone (HTL). Besides, it was suggested that the HTLase activity of PON1 could be responsible for its antiatherosclerotic action. Generally, PON1 activity is determined by using paraoxon (a toxic organophosphate), and hydrolysis of this compound is strongly influenced by mutations in the PON1 gene. Thus, an assay that uses HTL, an endogenous substrate, would be more suitable for evaluating the status of atherosclerosis than the assays that use paraoxon, an exogenous substrate.
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
121
Such a PON1 assay with a non-toxic, more physiological substrate could make the assay amenable to automation in routine laboratories. An HTLase activity assay, suitable for use with a routine analyzer, was proposed to represent the status of atherosclerosis in patients with type 2 diabetes mellitus (Kosada et al., 2005). Recently, it was reported that PON1 mass inversely predicts mortality in patients on hemodialysis (Ikeda et al., 2007), and an immunoblot assay to quantify PON1 in serum and plasma was developed (Connelly et al., 2008), making PON1 a biomarker to predict risk of vascular disease. Measurement of PON1 activity was also proposed as a biomarker for liver fibrosis in hepatitis C patients, chronic liver impairment, acute pancreatitis, oxidative damage of kidney during ischemia reperfusion injury, and active sarcoidosis (Camps et al., 2007; Gangadharan et al., 2007; Kulah et al., 2007; Franco-Pons et al., 2008; Uzun et al., 2008). Finally, added to the natural protective role of PON1 against OP exposure and atherosclerosis, the multiple beneficial effects of improving its expression and activities make the enzyme a promising prophylactic therapeutic mean.
CONCLUSION PON1 was first investigated through its toxicological role in protection against environmental poisoning by organophosphates, and secondly focused on a protective role in vascular disease. This latter has been related to its preferential association with a category of HDL particles. Finally, the establishment that the primary activity of the enzyme is a lactonase one provided the basis for a role in drug metabolism and degradation of endogenous lactones and thiolactones, reinforcing early hypotheses for a general and widespread protective function. The general protective activity of PON1 has implications for the management of patients who have, or are at risk for, atherosclerosis, as PON1 could form the basis of treatment or preventive strategies. PON1 would seem worthy of further study as an etiologic factor in the development of atherosclerosis and perhaps other diseases. Additional comprehensive structuree function analyses are prerequisites to determine the relationship between arylesterase, OP-hydrolase, lactonase, peroxidase, phospholipase and antioxidative actions of PON1, and to their interrelations with other HDL proteins. Thus, the basis for the special HDL association of PON1 and partners, such as the DING protein HPBP, merits further attention. Some DING proteins were isolated in association with pathological circumstances (e.g., rheumatoid arthritis, kidney stones, cancer, AIDS (Darbinian-Sarkissian et al., 2006)), for which PON1 activity was often shown to be altered. Finally, depiction of the known role and/or the possible synergy of PON1 and HPBP is a stimulating new objective. To sum up, study of the interactions of PON1 with HDL is an important task but a complex issue due to the promiscuity of the enzyme and the multiform and dynamic character of the HDL particles. Overcoming this challenge is critical for the design and development of safe and efficient therapies to affect HDL metabolism or to protect against poisoning by OPs.
122
The HDL Handbook
REFERENCES Abbott, C. A., Mackness, M. I., Kumar, S. Z., Boulton, A. J., & Durrington, P. N. (1995). Serum paraoxonase activity, concentration, and phenotype distribution in diabetes mellitus and its relationship to serum lipids and lipoproteins. Arterioscler Thromb Vasc Biol, 15, 1812e1818. Adkins, S., Gan, K. N., Mody, M., & La Du, B. N. (1993). Molecular basis for the polymorphic forms of human serum paraoxonase/arylesterase: glutamine or arginine at position 191, for the respective A or B allozymes. Am J Hum Genet, 52, 598e608. Aharoni, A., Gaidukov, L., Khersonsky, O., Gould, S. M., Roodveldt, C., & Tawfik, D. S. (2005). The ‘evolvability’ of promiscuous protein functions. Nat Genet, 37, 73e76. Aharoni, A., Gaidukov, L., Yagur, S., Toker, L., Silman, I., & Tawfik, D. S. (2004). Directed evolution of mammalian paraoxonases PON1 and PON3 for bacterial expression and catalytic specialization. Proc Natl Acad Sci USA, 101, 482e487. Amitai, G., Gaidukov, L., Adani, R., et al. (2006). Enhanced stereoselective hydrolysis of toxic organophosphates by directly evolved variants of mammalian serum paraoxonase. FEBS J, 273, 1906e1919. Aslan, M., Kosecik, M., Horoz, M., Selek, S., Celik, H., & Erel, O. (2007). Assessment of paraoxonase and arylesterase activities in patients with iron deficiency anemia. Atherosclerosis, 191, 397e402. Aslan, M., Nazligul, Y., Horoz, M., et al. (2008). Serum paraoxonase-1 activity in Helicobacter pylori infected subjects. Atherosclerosis, 196, 270e274. Assmann, G., von Eckardstein, A., & Funke, H. (1993). High density lipoproteins, reverse transport of cholesterol, and coronary artery disease. Insights from mutations. Circulation, 87, 28e34. Ates, O., Azizi, S., Alp, A. A., et al. (2009). Decreased serum paraoxonase 1 activity and increased serum homocysteine and malondialdehyde levels in age-related macular degeneration. Tohoku J Exp Med, 217, 17e22. Augustinsson, A. B., & Barr, M. (1963). Age variation in plasma arylesterase activity in children. Clin Chim Acta, 8, 568e573. Aviram, M., Billecke, S., Sorenson, R., et al. (1998a). Paraoxonase active site required for protection against LDL oxidation involves its free sulfhydryl group and is different from that required for its arylesterase/paraoxonase activities: selective action of human paraoxonase allozymes Q and R. Arterioscler Thromb Vasc Biol, 10, 1617e1624. Aviram, M., Dornfeld, L., Rosenblat, M., et al. (2000). Pomegranate juice consumption reduces oxidative stress, atherogenic modifications to LDL, and platelet aggregation: studies in humans and in atherosclerotic apolipoprotein E-deficient mice. Am J Clin Nutr, 71, 1062e1076. Aviram, M., Rosenblat, M., Bisgaier, C. L., Newton, R. S., Primo-Parmo, S. L., & La Du, B. N. (1998b). Paraoxonase inhibits high-density lipoprotein oxidation and preserves its functions. A possible peroxidative role for paraoxonase. J Clin Invest, 101, 1581e1590. Aviram, M., Rosenblat, M., Gaitini, D., et al. (2004). Pomegranate juice consumption for 3 years by patients with carotid artery stenosis reduces common carotid intima-media thickness, blood pressure and LDL oxidation. Clin Nutr, 23, 423e433. Bajnok, L., Csongradi, E., Seres, I., et al. (2008). Relationship of adiponectin to serum paraoxonase-1. Atherosclerosis, 197, 363e367. Beaumont, N. J., Skinner, V. O., Tan, T. M. M., et al. (2003). Ghrelin can bind to a species of high density lipoprotein associated with paraoxonase. J Biol Chem, 278, 8877e8880.
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
123
Benmoyal-Segal, L., Vander, T., Shifman, S., et al. (2005). Acetylcholinesterase/paraoxonase interactions increase the risk of insecticide-induced Parkinson’s disease. FASEB J, 19, 452e454. Bhasin, K. K., Yu, J. M., Tward, A., Shih, D., Campbell, D. A., & Lusis, A. J. (2006). Trypanosoma congolense: paraoxonase 1 prolongs survival of infected mice. Exp Parasitol, 114, 240e245. Billecke, S., Draganov, D., Counsell, R., et al. (2000). Human serum paraoxonase (PON1) isoenzymes Q and R hydrolyze lactones and cyclic carbonate esters. Drug Metab Dispos, 28, 1335e1342. Boemi, M., Leviev, I., Sirolla, C., Pieri, C., Marra, M., & James, R. W. (2001). Serum paraoxonase is reduced in type 1 diabetic patients compared to non-diabetic, first degree relatives; influence on the ability of HDL to protect LDL from oxidation. Atherosclerosis, 155, 229e235. Boemi, M., Sirolla, C., Testa, R., Cenerelli, S., Fumelli, P., & James, R. W. (2004). Smoking is associated with reduced serum levels of the antioxidant enzyme, paraoxonase, in Type 2 diabetic patients. Diabet Med, 21, 423e427. Bradshaw, G., Gutierrez, A., Miyake, J. H., et al. (2005). Facilitated replacement of Kupffer cells expressing a paraoxonase-1 transgene is essential for ameliorating atherosclerosis in mice. Proc Natl Acad Sci USA, 102, 11029e11034. Brooks-Wilson, A., Marcil, M., Clee, S. M., et al. (1999). Mutations in ABC1 in Tangier disease and familial high-density lipoprotein deficiency. Nat Genet, 22, 336e345. Brophy, V. H., Jampsa, R. L., Clendenning, J. B., McKinstry, L. A., Jarvik, G. P., & Furlong, C. E. (2001). Effects of 5’ regulatory region polymorphism on paraoxonase gene (PON1) expression. Am J Hum Genet, 68, 1428e1436. Cabana, V. G., Reardon, C. A., Feng, N., Neath, S., Lukens, J., & Getz, G. S. (2003). Serum paraoxonase: effect of the apolipoprotein composition of HDL and the acute phase response. J Lipid Res, 44, 780e792. Camps, J., Marsillach, J., & Joven, J. (2007). Measurement of serum paraoxonase-1 activity as a potential biomarker for chronic liver impairment. Clin Chim Acta, 386, 114e115. Cao, H., Girard-Globa, A., Berthezene, F., & Moulin, P. (1999). Paraoxonase protection of LDL against peroxidation is independent of its esterase activity towards paraoxon and is unaffected by the Qe>R genetic polymorphism. J Lipid Res, 40, 133e139. Chapuis, J., Bosher, M., Bensemain, F., Cottel, D., Amouyel, P., & Lambert, J. C. (2009). Association of the paraoxonase 1 gene with the risk of developing Alzheimer’s desease. Neurobiol Aging, 30, 152e156. Chen, N., Liu, Y., Greiner, C. D., & Holtzman, J. L. (2000). Physiologic concentrations of homocysteine inhibit the human plasma GSH peroxidase that reduces organic hydroperoxides. J Lab Clin Med, 136, 58e65. Cherki, M., Derouiche, A., Drissi, A., et al. (2005). Consumption of argan oil may have an antiatherogenic effect by improving paraoxonase activities and antioxidant status: Intervention study in healthy men. Nutr Metab Cardiovasc Dis, 15, 352e360. Ciftci, H., Yeni, E., Savas, M., Verit, A., & Celik, H. (2007). Paraoxonase activity in patients with erectile dysfunction. Int J Impot Res, 19, 517e520. Cle´ry-Barraud, C., Renault, F., Leva, J., El Bakdouri, N., Masson, P., & Rochu, D. (2009). Exploring the structural and functional stabilities of different paraoxonase-1 formulations through electrophoretic mobilities and enzyme activity parameters under hydrostatic pressure. Biochim Biophys Acta, 1794, 680e688. Cole, T. B., Jampsa, R. L., Walter, B. J., et al. (2003). Expression of human paraoxonase (PON1) during development. Pharmacogenetics, 13, 1e8.
124
The HDL Handbook
Connelly, P. W., Draganov, D., & Maguire, G. F. (2005). Paraoxonase-1 does not reduce or modify oxidation of phospholipids by peroxinitrite. Free Rad Biol Med, 38, 164e174. Connelly, P., Maguire, G. F., Picardo, C. M., Teiber, J. F., & Draganov, D. (2008). Development of an immunoblot assay with infrared fluorescence to quantify paraoxonase 1 in serum and plasma. J Lipid Res, 49, 245e250. Contreras-Martel, C., Carpentier, P., Morales, R., et al. (2006). Crystallization of the human phosphate binding protein. Acta Cryst, F62, 67e69. Costa, L. G., & Furlong, C. E. (2002). Paraoxonase (PON1) in health and disease: basic and clinical aspects. Dordrecht, The Netherlands: Kluwer Academic Publishers. Cowan, J., Sinton, C. M., Varley, A. W., Wians, F. H., Haley, R. W., & Munford, R. S. (2001). Gene therapy to prevent organophosphate intoxication. Toxicol Appl Pharmacol, 173, 1e6. Curtin, B. F., Seetharam, K. I., Dhoieam, P., Gordon, R. K., Doctor, B. P., & Nambiar, M. P. (2008). Resveratrol induces catalytic bioscavenger paraoxonase 1 expression and protects against chemical warfare nerve agent toxicity in human cell lines. J Cell Biochem, 103, 1524e1535. Dantoine, T. F., Debord, J., Merle, L., Lacroix-Ramiandrisoa, H., Bourzeix, L., & Charmes, J. P. (2002). Paraoxonase 1 activity: a new vascular marker of dementia? Ann NY Acad Sci, 977, 96e101. Darbinian-Sarkissian, N., Darbinian, A., Otte, J., et al. (2006). p27SJ, a novel protein in St John’s Wort, that suppresses expression of HIV-1 genome. Gene Ther, 13, 288e295. Davies, H. G., Richter, R. J., Keifer, M., Broomfield, C. A., Sowalla, J., & Furlong, C. E. (1996). The effect of the human serum paraoxonase polymorphism is reversed with diazoxon, soman and sarin. Nat Genet, 14, 334e336. Dawson, R. J., & Locher, K. P. (2006). Structure of a bacterial multidrug ABC transporter. Nature, 443, 180e185. Deakin, S., & James, R. W. (2004). Genetic and environmental factors modulating serum concentrations and activities of the antioxidant enzyme paraoxonase-1. Clin Sci, 107, 435e447. Deakin, S., Leviev, I., Bruhlart-Meynet, M. C., & James, R. W. (2003a). Paraoxonase-1 promoter haplotypes and serum paraoxonase: a predominant role in vivo for polymorphic position -107 implicating the transcription factor Sp1. Biochem J, 372, 643e649. Deakin, S., Leviev, I., Gomaraschi, M., Calabresi, L., Franceschini, G., & James, R. W. (2002). Enzymatically active paraoxonase-1 is located at the external membrane of producing cells and released by a high affinity, saturable, desorption mechanism. J Biol Chem, 277, 4301e4308. Deakin, S., Leviev, I., Guernier, S., & James, R. W. (2003b). Simvastatin modulates expression of the PON1 gene and increases serum paraoxonase: a role for sterol regulatory element-binding protein-2. Arterioscler Thromb Vasc Biol, 23, 2083e2089. Depairon, M., Chessex, S., Sudre, P., et al. (2001). Premature atherosclerosis in HIV-infected individuals e focus on protease inhibitor therapy. AIDS, 15, 329e334. Diemer, H., Elias, M., Renault, F., et al. (2008). Tandem use of X-ray crystallography and mass spectrometry to obtain ab initio the complete and exact amino acids sequence of HPBP, a human 38-kDa apolipoprotein. Proteins: Struct Funct Bioinform, 71, 1708e1720. Dirican, M., Akca, R., Sarandol, E., & Dilek, K. (2004). Serum paraoxonase activity in uremic predialysis and hemodialysis patients. J Nephrol, 17, 813e818. Draganov, D. I., Teiber, J. F., Speelman, A., Osawa, Y., Sunahara, R., & La Du, B. N. (2005). Human paraoxonases (PON1, PON2, and PON3) are lactonases with overlapping and distinct substrate specificities. J Lipid Res, 46, 1239e1247.
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
125
Eddleston, M., Buckley, N. A., Eyer, P., & Dawson, A. H. (2008). Management of acute organophosphorus poisoning. Lancet, 371, 597e607. Feingold, K. R., Memon, R. A., Moser, A. H., & Grunfeld, C. (1998). Paraoxonase activity in the serum and hepatic mRNA levels decrease during the acute phase response. Atherosclerosis, 139, 307e315. Ferretti, G., Bacchetti, T., Marchionni, C., Caldarelli, L., & Curatola, G. (2001). Effect of glycation of high density lipoproteins on their physicochemical properties and on paraoxonase activity. Acta Diabetol, 38, 163e169. Fokine, A., Morales, R., Contreras-Martel, C., et al. (2003). Direct phasing at low resolution of a protein co-purified with human paraoxonase (PON1). Acta Cryst D, 59, 2083e2087. Franco-Pons, N., Marsillach, J., Joven, J., Camps, J., & Closa, D. (2008). Serum paraoxonase undergoes inhibition and proteolysis during experimental acute pancreatitis. J Gastrointest Surg, 12, 891e899. Fu, A. L., Wang, Y. X., & Sun, M. J. (2005). Naked DNA prevents soman intoxication. Biochem Biophys Res Commun, 328, 901e905. Furlong, C. E., Richter, R. J., Chapline, C., & Crabb, J. W. (1991). Purification of rabbit and human serum paraoxonase. Biochemistry, 30, 10133e10140. Gaidukov, L., & Tawfik, D. S. (2005). High affinity, stability, and lactonase activity of serum paraoxonase PON1 anchored on HDL with apoA-I. Biochemistry, 44, 11843e11854. Gan, K. N., Smolen, A., Eckerson, H. W., & La Du, B. N. (1991). Purification of human serum paraoxonase/arylesterase. Evidence for one esterase catalyzing both activities. Drug Metabol Dispos, 19, 100e106. Gangadharan, B., Antrobus, R., Dwek, R. A., & Zitzmann, N. (2007). Novel serum biomarker candidates for liver fibrosis in hepatitis C patients. Clin Chem, 53, 1792e1799. Garin, M. C., Kalix, B., Morabia, A., & James, R. W. (2005). Small, dense lipoprotein particles and reduced paraoxonase-1 in patients with the metabolic syndrome. J Clin Endocrinol Metab, 90, 2264e2269. Goue´dard, C., Koum-Besson, N., Barouki, R., & Morel, Y. (2003). Opposite regulation of the human paraoxonase-1 gene PON-1 by fenofibrate and statins. Mol Pharmacol, 63, 945e956. Goue´dard, C., Barouki, R., & Morel, Y. (2004a). Dietary polyphenols increase paraoxonase 1 gene expression by an aryl hydrocarbon receptor-dependent mechanism. Mol Cell Biol, 24, 5209e5222. Goue´dard, C., Barouki, R., & Morel, Y. (2004b). Induction of the paraoxonase-1 gene expression by resveratrol. Arterioscler Thromb Vasc Biol, 24, 2378e2383. Gugliucci, A., Mehlhaff, K., Kinugasa, E., et al. (2007). Paraoxonase-1 concentrations in end-stage renal disease patients increase after hemodialysis: correlation with low molecular AGE adduct clearance. Clin Chim Acta, 377, 213e220. Guns, P.-J., Van Assche, T., Verreth, W., et al. (2008). Paraoxonase 1 gene transfer lowers vascular oxidative stress and improves vasomotor function in apolipoprotein E-deficient mice with preexisting atherosclerosis. Brit J Pharmacol, 153, 508e516. Gutteridge, J. M., & Stocks, J. (1981). Caeruloplasmin: physiological and pathological perspectives. Crit Rev Clin Lab Sci, 14, 257e329. Hajduk, S. L., Moore, D. R., Vasudevacharya, J., et al. (1989). Lysis of Trypanosoma brucei by a toxic subspecies of human high density lipoprotein. J Biol Chem, 264, 5210e5217. Haley, R. W., Billecke, S., & La Du, B. N. (1999). Association of low PON1 type Q (type A) arylesterase activity with neurologic symptom complexes in Gulf War veterans. Toxicol Appl Pharmacol, 157, 227e233.
126
The HDL Handbook
Hamilton, K. K., Zhao, J., & Sims, P. J. (1993). Interaction between apolipoproteins A-I and A-II and the membrane attack complex of complement. Affinity of the apoproteins for polymeric C9. J Biol Chem, 268, 3632e3638. Harel, M., Aharoni, A., Gaidukov, L., et al. (2004). Structure and evolution of the serum paraoxonase family of detoxifying and anti-atherosclerotic enzymes. Nat Struct Mol Biol, 11, 412e419. Harman, D. (1956). Aging: A theory based on free radical and radiation chemistry. J Gerontol, 11, 298e300. He, X.-M., Zhang, Z. X., Zhang, J. W., et al. (2006). Gln192Arg polymorphism in paraoxonase 1 gene is associated with Alzheimer disease in a Chinese Han ethnic population. Chin Med J, 119, 1204e1209. Hedrick, C. C., Thorpe, S. R., Fu, M. X., et al. (2000). Glycation impairs high-density lipoprotein function. Diabetologia, 43, 312e320. Hegele, R. A., Connelly, P. W., Scherer, S. W., et al. (1997). Paraoxonase-2 gene (PON2) G148 variant associated with elevated fasting plasma glucose in noninsulin-dependent diabetes mellitus. J Clin Endocrinol Metab, 82, 3373e3377. Herrick, J. B. (1912). Clinical features of sudden obstruction of the coronary arteries. J Am Med Assoc, 59, 2015e2020. Hill, C. M., Li, W. S., Thoden, J. B., Holden, H. M., & Raushel, F. M. (2003). Enhanced degradation of chemical warfare agents through molecular engineering of the phosphotriesterase active site. J Am Chem Soc, 125, 8990e8991. Hotopf, M., Mackness, M. I., Nikolaou, V., et al. (2003). Paraoxonase in Persian Gulf War veterans. J Occup Environ Med, 45, 668e675. Hu, X., Jiang, X., Lenz, D. E., Cerasoli, D. M., & Wallqvist, A. (2009). In silico analyses of substrate interactions with human serum paraoxonase 1. Proteins, 75, 486e498. Humbert, R., Adler, D. A., Disteche, C. M., Hassett, C., Omiecinski, C. J., & Furlong, C. E. (1993). The molecular basis of the human serum paraoxonase activity polymorphism. Nat Genet, 3, 73e76. Ikeda, Y., Suehiro, T., Itahara, T., et al. (2007). Human serum paraoxonase concentration predicts cardiovascular mortality in hemodialysis patients. Clin Nephrol, 67, 358e365. James, R. W., & Deakin, S. P. (2004). The importance of high-density lipoproteins for paraoxonase-1 secretion, stability, and activity. Free Rad Biol Med, 37, 1986e1994. James, R. W., Leviev, I., & Righetti, A. (2000). Smoking is associated with reduced serum paraoxonase activity and concentration in patients with coronary artery disease. Circulation, 101, 2252e2257. James, R. W., Blatter-Garin, M. C., Calabresi, L., et al. (1998). Modulated serum activities and concentrations of paraoxonase in high density lipoprotein deficiency states. Atherosclerosis, 139, 77e82. Jaouad, L., de Guise, C., Berrougui, H., et al. (2006). Age-related decrease in high-dernity lipoprotcins antioxidant activity is due to an alteration in the PONl’s free sulfhydryl groups. Atherosclerosis, 185, 191e200. Jarvik, G. P., Tsai, N. T., McKinstry, L. A., et al. (2002). Vitamin C and E intake is associated with increased paraoxonase activity. Arterioscler Thromb Vasc Biol, 22, 1329e1333. Jarvik, G. P., Hatsukami, T. S., Carlson, C., et al. (2003). Paraoxonase activity, but not haplotype utilizing the linkage disequilibrium structure, predicts vascular disease. Arterioscler Thromb Vasc Biol, 23, 1465e1471. Jeyaratnam, J. (1990). Acute pesticide poisoning: a major global health problem. World Health Stat Q, 43, 139e144.
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
127
Kannel, W. B., Dawber, T. R., Kagan, A., Revotskie, N., & Stokes, J. (1961). Factors of risk in the development of coronary heart disease e six year follow-up experience. The Framingham Study. Ann Intern Med, 55, 33e50. Kastelein, J. J., Akdim, F., Stroes, E. S., et al. (2008). Simvastatin with or without ezetimibe in familial hypercholesterolemia. N Engl J Med, 358, 1431e1443. Kelso, G. J., Stuart, W. D., Richter, R. J., Furlong, C. E., Jordan-Starck, T. C., & Harmony, J. A. (1994). Apolipoprotein J is associated with paraoxonase in human plasma. Biochemistry, 33, 832e839. Khersonsky, O., & Tawfik, D. S. (2005). Structure-reactivity studies of serum paraoxonase PON1 suggest that its native activity is lactonase. Biochemistry, 44, 6371e6382. Khersonsky, O., & Tawfik, D. S. (2006). The histidine 115-histidine 134 dyad mediates the lactonase activity of mammalian serum paraoxonases. J Biol Chem, 281, 7649e7656. Khersonsky, O., & Tawfik, D. S. (2006a). Chromogenic and fluorogenic assays for the lactonase activity of serum paraoxonases. Chem BioChem, 7, 49e53. Kondo, I., & Yamamoto, M. (1998). Genetic polymorphism of paraoxonase 1 (PON1) and susceptibility to Parkinson’s disease. Brain Res, 806, 271e273. Kontush, A., Chantepie, S., & Chapman, M. J. (2003). Small, dense HDL particles exert potent protection of atherogenic LDL against oxidative stress. Arterioscler Thromb Vasc Biol, 23, 1881e1888. Kosada, T., Yamagushi, M., Motomura, T., & Mizuno, K. (2005). Investigation of the relationship between atherosclerosis and paraoxonase or homocysteine thiolactonase activity in patients with type 2 diabetes mellitus using a commercially available assay. Clin Chim Acta, 359, 156e162. Krieger, M. (1999). Charting the fate of the “good cholesterol”: identification and characterization of the high-density lipoprotein receptor SR-BI. Annu Rev Biochem, 68, 523e558. Krieger, M. (2001). Scavenger receptor class B type I is a multiligand HDL receptor that influences diverse physiologic systems. J Clin Invest, 108, 793e797. Kujiraoka, T., Oka, T., Ishihara, M., et al. (2000). A sandwich enzyme-linked immunosorbent assay for human serum paraoxonase concentration. J Lipid Res, 41, 1358e1363. Kulah, E., Tascilar, O., Acikgoz, S., et al. (2007). Oxidized LDL accumulation in experimental renal ischemia reperfusion injury model. Ren Fail, 29, 409e415. Kwong, T. C. (2002). Organophosphate pesticides: biochemistry and clinical toxicology. Ther Drug Monit, 24, 144e149. La Du, B. N. (1996). Structural and functional diversity of paraoxonases. Nat Med, 2, 1186e1187. Lenz, D. E., Broomfield, C. A., Yeung, D. T., Masson, P., Maxwell, D. M., & Cerasoli, D. M. (2007). Nerve agent bioscavengers: progress in development of a new mode of protection against organophosphorus exposure. In J. A. Romano, B. Luckey, & H. Salem (Eds.), Chemical warfare agents: chemistry, pharmacology and therapeutics (pp. 175e202). Boca Raton: CRC Press. Lescai, F., Marchegiani, F., & Franceschi, C. (2009). PON1 as a longevity gene: results of a metaanalysis. Ageing Res Rev, 8, 277e284. Leviev, I., & James, R. W. (2000). Promoter polymorphisms of the human paraoxonase PON1 gene and serum paraoxonase activities and concentrations. Arterioscler Thromb Vasc Biol, 20, 516e521. Lohmueller, K. E., Pearce, C. L., Pike, M., Lander, E. S., & Hirschorn, J. N. (2003). Meta-analysis of genetic association studies supports a contribution of common variants to susceptibility to common disease. Nat Genet, 33, 177e182.
128
The HDL Handbook
Luecke, H., & Quiocho, F. A. (1990). High specificity of a phosphate transport protein determined by hydrogen bonds. Nature, 347, 402e406. Lusis, A. J. (2000). Atherosclerosis. Nature, 407, 233e241. Mackness, M., & Mackness, B. (2004). Paraoxonase 1 and atherosclerosis: is the gene or the protein more important? Free Radic Biol Med, 37, 1317e1323. Mackness, M. I., Arrol, S., & Durrington, P. N. (1991). Paraoxonase prevents accumulation of lipoperoxides in low-density lipoprotein. FEBS Lett, 286, 152e154. Mackness, M. I., Walker, C. H., & Carlson, L. A. (1987). Low A-esterase activity in serum of patients with fish-eye disease. Clin Chem, 33, 587e588. Mackness, M. I., Halam, S. D., Peard, T., Warner, S., & Walker, C. H. (1985). The separation of sheep and human serum “A”eesterase activity into the lipoprotein fraction by ultracentrifugation. Comp Biochem Physiol, 82, 675e677. Mackness, M. I., Harty, D., Bhatnagar, D., et al. (1991b). Serum paraoxonase activity in familial hypercholesterolaemia and insulin-dependent diabetes mellitus. Atherosclerosis, 86, 193e199. Mackness, M. I., Peuchant, E., Dumon, M. F., Walker, C. H., & Clerc, M. (1989). Absence of “A”esterase activity in the serum of a patient with Tangier disease. Clin Biochem, 22, 475e478. Mackness, B., Mackness, M. I., Arrol, S., Turkie, W., & Durrington, P. N. (1997a). Effect of the molecular polymorphisms of human paraoxonase (PON1) on the rate of hydrolysis of paraoxon. Br J Pharmacol, 122, 265e268. Mackness, M. I., Arrol, S., Mackness, B., & Durrington, P. N. (1997b). Alloenzymes of paraoxonase and effectiveness of high-density lipoproteins in protecting low-density lipoprotein against lipid peroxidation. Lancet, 349, 851e852. Mackness, B., Durrington, P. N., & Mackness, M. I. (2000). Low paraoxonase in Persian Gulf War Veterans self-reporting Gulf War syndrome. Biochim Biophys Res Commun, 276, 729e733. Mackness, B., Davies, G. K., Turkie, W., Lee, E., Roberts, D. H., Hill, E., et al. (2001). Paraoxonase status in coronary heart disease: are activity and concentration more important than genotype? Arterioscler Thromb Vasc Biol, 21, 1451e1457. Mackness, B., Mackness, M., Aviram, M., & Paragh, G. (2008). The paraoxonase: their role in disease development and xenobiotic metabolism. Dordrecht, The Netherlands: Springer. Maggi, P., Serio, G., Epifani, G., et al. (2000). Premature lesions of the carotid vessels in HIV-1infected patients treated with protease inhibitors. AIDS, 14, 123e128. Mao, B., Pear, M. R., McCammon, J. A., & Quiocho, F. A. (1982). Hinge-bending in L-arabinosebinding protein. The “Venus’s-flytrap” model. J Biol Chem, 257, 1131e1133. Marathe, G. K., Zimmerman, G. A., & McIntyre, T. M. (2003). Platelet-activating factor acetylhydrolase, and not paraoxonase-1, is the oxidized phospholipid hydrolase of high density lipoprotein particles. J Biol Chem, 278, 3937e3947. Marchegiani, F., Marra, M., Spazzafumo, L., et al. (2006). Paraoxonase activity and genotype predispose to successful aging. J Gerontol, 61A, 541e546. Mariani, E., Polidori, M. C., Cherubini, A., & Mecocci, P. (2005). Oxidativc stress in brain aging, neurodegenerative and vascular diseases: An overview. J Chromatogr B, 827, 65e75. Masson, P., & Rochu, D. (2009). Catalytic bio scavengers: The next generation of bioscavengerbased medical countermeasures. In R. Gupta (Ed.), Handbook of the toxicology of chemical warfare agents (pp. 1053e1065). San Diego: Academic Press. Maury, C. P., Junge, W., & Teppo, A. M. (1984). Serum esterase activity in reactive systemic amyloidosis and its relation to amyloid A degrading activity. J Lab Clin Med, 104, 761e766. Milochevitch, C., & Khalil, A. (2001). Study of the paraoxonase and platelet-activating factor acetylhydrolase activities with aging. Prostaglandins Leukot Essent Fatty Acids, 65, 241e246.
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
129
Mineo, C., Deguchi, H., Griffin, J. H., & Shaul, P. W. (2006). Endothelial and antithrombotic actions of HDL. Circ Res, 98, 1352e1364. Miyoshi, M., Nakano, Y., Sakagushi, T., et al. (2007). Gene delivery of paraoxonase-1 inhibits neointimal hyperplasia after arterial balloon-injury in rabbits fed a high-fat diet. Hypertens Res, 30, 85e91. Morales, R., Berna, A., Carpentier, P., et al. (2006). Serendipitous discovery and X-ray structure of a human phosphate binding apolipoprotein. Structure, 14, 601e609. Napoli, C., Abete, P., Corso, G., et al. (1997). Increased low-density lipoprotein peroxidation in elderly men. Coron Artery Dis, 8, 129e136. Navab, M., Ananthramaiah, G. M., Reddy, S. T., et al. (2004). The oxidation hypothesis of atherogenesis: the role of oxidized phospholipids and HDL. J Lipid Res, 45, 993e1007. Nishio, E., & Watanabe, Y. (1997). Cigarette smoke extract inhibits plasma paraoxonase activity by modification of the enzyme’s free thiols. Biochem Biophys Res Commun, 236, 289e293. Nuutinen, T., Suuronen, T., Kauppinen, A., & Salminen, A. (2009). Clusterin: a forgotten player in Alzheimer’s disease. Brain Res Rev in press. doi:10.1016/ j.brainresrev.2009.05.007. Oda, M. N., Bielicki, J. K., Ho, T. T., Berger, T., Rubin, E. M., & Forte, T. M. (2002). Paraoxonase 1 overexpression in mice and its effect on high-density lipoproteins. Biochem Biophys Res Commun, 290, 921e927. Omura, O. (1980). Serum-paraoxonase-polymorphisms in Japan. MD Thesis. Germany: University of Erlanger. Ozer, E. A., Pezzulo, A., Shih, D. M., et al. (2005). Human and murine paraoxonase 1 are host modulators of Pseudomonas aeruginosa quorum-sensing. FEMS Microbiol Lett, 253, 29e37. Paragh, G., Balla, P., Katona, E., Seres, I., Egerhazi, A., & Degrell, I. (2002). Serum paraoxonase activity changes in patients with Alzheimer’s disease and vascular dementia. Eur Arch Psychiatry Clin Neurisci, 252, 63e67. Parra, S., Alonso-Villaverde, C., Coll, B., et al. (2007). Serum paraoxonase-1 activity and concentration are influenced by human immunodeficiency virus infection. Atherosclerosis, 194, 175e181. Que´me´neur, T., Martin-Nizard, F., Kandoussi, A., et al. (2007). PON1, a new biomarker of cardiovascular disease, is low in patients with systemic vasculitis. Semin Arthritis Rheum, 37, 149e155. Reaven, P. D., & Witztum, J. L. (1996). Oxidized low-density lipoproteins in atherogenesis: Role of dietary modification. Annu Rev Nutr, 16, 51e71. Reilly, M. P., & Tall, A. R. (2007). HDL proteomics: pot of gold or Pandora’s box? J Clin Invest, 117, 595e598. Renault, F., Chabrie`re, E., Andrieu, J. P., Dublet, B., Masson, P., & Rochu, D. (2006). Tandem purification of two HDL-associated partner proteins in human plasma, paraoxonase (PON1) and phosphate binding protein (HPBP) using hydroxyapatite chromatography. J Chromatogr B, 836, 15e21. Renault, F., Carus, T., Cle´ry-Barrand, C., et al. (2010). Integrative analytical approach by capillary electrophoresis and kinetics under high pressure optimized for deciphering intrinsic and extrinsic cofactors that modulate activity and stability of human paraoxonase (PON1). J Chromatogr. B. doi:10.1016/ j.jchromb.2009.11.027. Robertson, T. L., Kato, H., Rhoads, G. G., et al. (1977). Epidemiologic studies of coronary heart disease and stroke in Japanese men living in Japan, Hawaii and California. Incidence of myocardial infarction and death from coronary heart disease. Am J Cardiol, 39, 239e243. Rochu, D., Chabrie`re, E., & Masson, P. (2007a). Human paraoxonase: a promising approach for pre-treatment and therapy of organophosphorus poisoning. Toxicology, 233, 47e59.
130
The HDL Handbook
Rochu, D., Renault, F., Cle´ry-Barraud, C., Chabrie`re, E., & Masson, P. (2007b). Stability of highly purified human paraoxonase (PON1): Association with human phosphate binding protein (HPBP) is essential for preserving its active conformation(s). Biochim Biophys Acta, 1774, 874e883. Rodrigo, L., Mackness, B., Durrington, P. N., Hernandez, A., & Mackness, M. I. (2001). Hydrolysis of platelet-activating factor by human serum paraoxonase. Biochem J, 354, 1e7. Rosenblat, M., Gaidukov, L., Khersonsky, O., et al. (2006). The catalytic histidine dyad of high density lipoprotein-associated serum paraoxonase-1 (PON1) is essential for PON1-mediated inhibition of low density lipoprotein oxidation and stimulation of macrophage cholesterol efflux. J Biol Chem, 281, 7657e7665. Rozenberg, O., Rosenblat, M., Coleman, R., Shih, D. M., & Aviram, M. (2003). Paraoxonase (PON1) deficiency is associated with increased macrophage oxidative stress: studies in PONknockout mice. Free Radic Biol Med, 34, 774e784. Sarandol, E., Safak, O., Dirican, M., & Uncu, G. (2004). Oxidizability of apolipoprotein Bcontaining lipoproteins and serum paraoxonase/arylesterase activities in preeclampsia. Clin Biochem, 37, 990e996. Schoenheimer, R. (1949). The dynamic state of body constituents. Cambridge MA: Harvard University Press. Schrader, G. (1950). Organische phosphor verbindungen als neuartige Insektizide. Angew Chem, 62, 471e473. Senti, M., Aubo, C., & Tomas, M. (2000). Differential effects of smoking on myocardial infarction risk according to the Gln/Arg 192 variants of the human paraoxonase gene. Metabolism, 49, 557e559. Senti, M., Tomas, M., Fito, M., et al. (2003). Antioxidant paraoxonase 1 activity in the metabolic syndrome. J Clin Endocrinol Metab, 88, 5422e5426. Seres, I., Paragh, C., Deschene, E., Fulop, T., Jr., & Khalil, A. (2004). Study of factors influencing the decreased HDL associated PONl activity with aging. Exp Gerontol, 39, 59e66. Serhatlioglu, S., Gursu, M. F., Gulcu, F., Canatan, H., & Godekmerdan, A. (2003). Levels of paraoxonase and arylesterase activities and malondialdehyde in workers exposed to ionizing radiation. Cell Biochem Funct, 21, 371e375. Shih, D. M., Gu, L., Xia, Y.-R., et al. (1998). Mice lacking serum paraoxonase are susceptible to organophosphate toxicity and atherosclerosis. Nature, 394, 284e287. Shih, D. M., Xia, Y. R., Miller, E., et al. (2000). Combined serum paraoxonase knockout/apolipoprotein E knockout mice exhibit increased lipoprotein oxidation and atherosclerosis. J Biol Chem, 275, 17527e17535. Sierksma, A., Van der Gaag, M. S., van Tol, A., & James, R. W. (2002). Kinetics of HDL cholesterol and paraoxonase activity in moderate alcohol consumers. Alcohol Clin Exp Res, 26, 1430e1435. Sorenson, R. C., Bisgaier, C. L., Aviram, M., Hsu, C., Billecke, S., & La Du, B. N. (1999). Human serum paraoxonase/arylesterase’s retained hydrophobic N-terminal leader sequence associates with HDLs by binding phospholipids: apolipoprotein A-I stabilizes activity. Arterioscler Thromb Vasc Biol, 19, 2214e2225. Sozmen, E. Y., Mackness, B., Sozmen, B., et al. (2002). Effect of organophosphate intoxication on human serum paraoxonase. Hum Exp Toxicol, 21, 247e252. Stoltz, D. A., Ozer, E. A., Ng, C. J., et al. (2007). Paraoxonase-2 deficiency enhances Pseudomonas aeruginosa quorum sensing in murine tracheal epithelia. Am J Physiol Lung Cell Mol Physiol, 292, L852e860.
Chapter j 6
Paraoxonase-1 and its Interactions with HDL
131
Teiber, J. F., Draganov, D. I., & La Du, B. N. (2004). Purified human serum PON1 does not protect LDL against oxidation in the in vitro assays initiated with copper or AAPH. J Lipid Res, 45, 2260e2268. Tomas, M., Senti, M., Garcia-Faria, F., et al. (2000). Effect of simvastatin therapy on paraoxonase activity and related lipoproteins in familial hypercholesterolemic patients. Arterioscler Thromb Vasc Biol, 20, 2113e2119. Tward, A., Xia, Y. R., Wang, X. P., et al. (2002). Decreased atherosclerotic lesion formation in human serum paraoxonase transgenic mice. Circulation, 106, 484e490. Ulevitch, R. J., Johnston, A. R., & Weinstein, D. B. (1979). New function for high density lipoproteins. Their participation in intravascular reactions of bacterial lipopolysaccharides. J Clin Invest, 64, 1516e1524. Uriel, J. (1961). Characterization of cholinesterase and other carboxylic esterases after electrophoresis and immunoelectrophoresis on agar. I. Application to the study of esterases of normal human serum. Ann Inst Pasteur, 101, 104e119. Uzun, H., Yanardag, H., Gelisgen, R., et al. (2008). Levels of paraoxonase, an index of antioxidant defense, in patients with active sarcoidosis. Curr Med Res Opin, 24, 1651e1657. Vaisar, T., Pennathur, S., Green, P. S., et al. (2007). Shotgun proteomics implicates protease inhibition and complement activation in the antiinflammatory properties of HDL. J Clin Invest, 117, 746e756. Van der Gaag, M. S., van Tol, A., Scheek, L. M., et al. (1999). Daily moderate alcohol consumption increases serum paraoxonase activity; a diet-controlled, randomised intervention study in middle-aged men. Atherosclerosis, 147, 405e410. van Himbergen, T. M., van Tits, L. J. H., Hectors, M. P. C., de Graaf, J., Roest, M., & Stalenhoef, A. F. H. (2005). Paraoxonase-1 and linoleic acid oxidation in familial hypercholesterolemia. Biochem Biophys Res Commun, 333, 787e793. Van Lenten, B. J., Hama, S. Y., de Beer, F. C., et al. (1995). Anti-inflammatory HDL becomes proinflammatory during the acute phase response. Loss of protective effect of HDL against LDL oxidation in aortic wall cell cocultures. J Clin Invest, 96, 2758e2767. Van Lenten, B. J., Wagner, A. C., Navab, M., & Fogelman, A. M. (2001). Oxidized phospholipids induce changes in hepatic paraoxonase and ApoJ but not monocyte chemoattractant protein-1 via interleukin-6. J Biol Chem, 276, 1923e1929. Wheeler, J. G., Keavney, B. D., Watkins, H., Collins, R., & Danesh, J. (2004). Four paraoxonase gene polymorphisms in 11212 cases of coronary heart disease and 12786 controls: metaanalysis of 43 studies. Lancet, 363, 689e695. Wu, A., Hinds, C. J., & Thiemermann, C. (2004). High-density lipoproteins in sepsis and septic shock: metabolism, actions, and therapeutic applications. Shock, 21, 210e221.
Chapter 7
Apolipoprotein A-I Mutations Akira Matsunaga, Yoshinari Uehara, Bo Zhang and Keijiro Saku Department of Laboratory Medicine, Department of Cardiology, Fukuoka University School of Medicine, Fukuoka, Japan
INTRODUCTION High density lipoprotein (HDL) is synthesized via a complex pathway. Although the underlying genetic defects in many cases of primary low HDL cholesterolemia are not clearly understood, mutations in three pivotal genes, namely apolipoprotein A-I (apoA-I), ATP binding cassette transporter A1 (ABCA1) and lecithin:cholesterol acyltransferase (LCAT), are associated with low plasma HDL cholesterol (HDL-C) levels (Miller et al., 2003). Some mutations of these genes are also associated with an increased risk of premature coronary artery disease (CAD). The most severe form of HDL deficiency is Tangier disease, which is a recessive disorder characterized by cholesteryl ester (CE) accumulation in macrophages, orange-yellow tonsils, hepatosplenomegaly, peripheral nerve neuropathy, corneal opacifications and premature CAD in 50% of cases (Hobbs and Rader, 1999). Tangier disease is caused by mutations of the gene encoding ABCA1, a 2261-amino acid peptide (Hobbs and Rader, 1999; Uehara et al., 2007). LCAT catalyzes the transfer of a preferentially unesterified fatty acid from phosphatidylcholine to free cholesterol, thereby producing a CE. The LCAT gene contains six exons and encodes a 440-amino acid protein including a 24-amino acid signal peptide (Kuivenhoven et al., 1997). Depending on the mutation, homozygous or compound heterozygous patients with LCAT mutations present with one of two clinical phenotypes, namely classical LCAT deficiency or fish-eye disease. Both disorders are characterized by the occurrence of corneal opacifications as hallmarks (Moriyama et al., 1995; Miida et al., 2004). However, premature CAD is uncommon, even in the presence of cardiovascular risk factors. ApoA-I is the major protein component of HDL, and plays essential roles in the biogenesis and functions of HDL. Approximately 70% of the HDL protein The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
133
134
The HDL Handbook
mass is comprised of apoA-I with another 15e20% comprised of apoA-II (Mahley et al., 1984; Davidson and Thompson, 2007). HDL assembly initially involves cell surface ABCA1 transporter-mediated transfer of cellular phospholipids (PL) and unesterified cholesterol to extracellular lipid-poor apoA-I particles (preb-HDL particles). Overall, 61 mutations of the ApoA-I gene have been reported to date (45 missense, five nonsense and six frameshift mutations, and five in-frame deletions), and these are mostly located in the coding region. The nonsense and frameshift mutations, which affect the complete translation of apoA-I mRNA, are always associated with reduced levels of apoA-I and HDL-C. In contrast, only half of the missense mutations are associated with low levels of apoA-I and HDL-C. However, only a minority of these apoA-I mutations are associated with premature CAD. Most of the mutations of the ApoA-I gene have been found in heterozygous subjects. The wide allelic heterogeneity and condition of heterozygosity make it difficult to clearly define the phenotypic expression of ApoA-I gene mutations in heterozygotes with respect to the development of premature CAD. For these reasons, the identification of individuals homozygous for mutations of the ApoA-I gene, leading to severe apoA-I deficiency, is very useful for better understanding of the roles of apoA-I and HDL in protection against CAD. Some homozygous subjects with virtually no HDL or apoA-I have been reported (Funke et al., 1991; Matsunaga et al., 1991).
GENOMIC AND PROTEIN STRUCTURE OF APOA-I Mature apoA-I in plasma is a 243-amino acid peptide primarily synthesized in the liver and intestine. ApoA-I is a 28-kDa single polypeptide that lacks glycosylation or disulfide linkages. The ApoA-I gene resides on chromosome 11q23-q24, contains four exons and encodes a 267-amino acid peptide chain including an 18-amino acid prepeptide and a 6-amino acid propeptide (Fielding and Fielding, 1995). Exon 3 of the ApoA-I gene encodes the N-terminal region of mature apoA-I (residues 1-43) and the remaining regions are encoded by exon 4. Large parts of the mature protein form homologous repeats of amphipathic alpha-helices, each encompassing 11 or 22 amino acids (Nolte and Atkinson, 1992; Brouillette and Anantharamaiah, 1995; Borhani et al., 1997).
PHYSIOLOGICAL FUNCTIONS OF APOA-I As the quantitatively predominant component of HDL, apoA-I is crucial for HDL formation. It is also required to activate LCAT and mediate the interactions of HDL with cell surface receptors, such as scavenger receptor B1 or plasma membrane transporters including ABCA1 or G1 (Gelissen et al., 2006; Zanis et al., 2006). In mice and rabbits, knockout of the ApoA-I gene causes HDL deficiency and, conversely, transgenic overexpression of apoA-I increases HDL-C in a gene-dose-dependent manner. In susceptible animals with an
Chapter j 7
Apolipoprotein A-I Mutations
135
atherogenic lipoprotein profile, atherosclerosis is enhanced by apoA-I deficiency and decreased by transgenic overexpression of apoA-I. Human subjects with apoA-I deficiency and apoA-I-deficient mice fail to form normal HDL particles (Williamson et al., 1992).
APOA-I DEFICIENCIES WITH LARGE DELETION, NONSENSE OR FRAMESHIFT MUTATIONS In total, 13 types of ApoA-I deficiency caused by large deletion, nonsense or frameshift mutations have been reported to date (Table 7.1) (Norum et al., 1982; Schaefer et al., 1985; Funke et al., 1991; Matsunaga et al., 1991, 1999; Lackner et al., 1993; Nakata et al., 1993; Ng et al., 1994; Romling et al., 1994; Takata et al., 1995; Miccoli et al., 1996; Moriyama et al., 1996; Yokota et al., 2002; Ikewaki et al., 2004; Dastani et al., 2006). The probands of ApoA-I/C-III deficiency were two sisters aged 29 and 31 years, who had skin and tendon xanthomas, corneal opacities and severe CAD (Norum et al., 1982). The rearrangement in these probands consisted of a DNA inversion involving portions of the 30 ends of the ApoA-I and ApoC-III genes, including the DNA region between these genes (Karathanasis et al., 1987). Their plasma levels of HDL-C were 4 and 7 mg/dL (0.10 and 0.18 mmol/L, respectively), and only traces of apoA-I were detected in whole plasma. The proband of ApoA-I/C-III/A-IV deficiency was a 45-year-old white female (Schaefer et al., 1985). The defect in this proband was complete deletion of the ApoA-I, ApoC-III, ApoA-IV gene complex on chromosome 11 (Ordovas et al., 1989). She had severe CAD and corneal opacities, but no hepatosplenomegaly or xanthomas were detected. Her plasma level of HDL-C was 1 mg/dL. She died after coronary bypass surgery. The proband of Q-2X was a 34-year-old Caucasian female who had xanthelasma, Achilles xanthomas, bilateral cataracts, retinal detachments, mild midline cerebellar ataxia and asymmetric bilateral neurosensory hearing loss (Ng et al., 1994). Her plasma HDL-C was 2.3 mg/dL. This proband and four of her 11 siblings had a homozygous mutation of Q-2X, severe deficiency of HDL (<3.9 mg/dL) and undetectable plasma apoA-I. At 38 years of age, a homozygous sister had premature CAD. An 11-year-old girl of Turkish descent was the proband of Q5fsX36 (Lackner et al., 1993). She was presented to her pediatrician due to unusual yellow-orange xanthomas of several joints at 7 years of age. She had no corneal opacities, no enlarged tonsils and no splenomegaly. Her plasma HDL-C was 2 mg/dL. Another proband of heterozygous Q5fsX36 was a 10-year-old Japanese girl (Nakata et al., 1993). Her plasma HDL-C and apoA-I concentrations were 27 and 76 mg/dL, respectively. The proband of a compound heterozygote of Q5fsX11 and L141RPisa was a 47-year-old Italian man (Miccoli et al., 1996). He was referred to his physician due to corneal opacities and complete HDL deficiency (0 mg/dL). He underwent three-vessel coronary bypass surgery. His plasma apoA-I level was
136
TABLE 7.1
Apolipoprotein A-I deficiencies with large deletion, nonsense or frameshift mutations
Mutation
Genomic change Allele
Age
Sex
Ethnicity
CAD Characteristics
Reference
1. A-I/C-III
gene inversion
Homozygote Homozygote
29 31
F F
American American
þ þ
Two sisters had skin and tendon xanthomas and corneal opacities
Norum et al. (1982)
2. A-I/C-III/A-IV
gene deletion
Homozygote
45
F
American
þ
She died of CAD at age 45
Schaefer et al. (1985)
3. Q-2X
CAG->TAG at codon-2
Homozygote
34
F
Canadian
þ
She had retinopathy Ng et al. (1994) cataracts, spinocerebellar ataxia and tendon xanthomas
4.Q5fsX36)
Insertion of C at codon 5
Homozygote Heterozygote
11 10
F F
Turkish Japanese
She had planar xanthomas
5. Q5fsX11)
Deletion of C at codon 5
Compound))
47
M
Italian
þ
Another was apoA-I L141RPisa. He had corneal opacities
6. W8XWakayama
TGG->TGA at codon 8
Homozygote
39
M
Japanese
He had corneal opacities
Takata et al. (1995)
7. Q32X
CAG->TAG at codon 32
Homozygote
31
F
Italian
She had perioirbital xanthelasmas
Romling et al. (1994)
Lackner et al. (1993) Nakata et al. (1993)
The HDL Handbook
She had planar xanthomas Another was a missense mutation in the TATA box
Matsunaga et al. (1991) Matsunaga et al. (1999)
French Canadian
þ
E136X mutations were found in 17 carriers
Dastani et al. (2006)
F
Japanese
þ
She had corneal opacities and planar xanthomas
Ikewaki et al. (2004)
50
F
Japanese
She had corneal opacities and skin and tendon xanthomas
Moriyama et al. (1996)
Homozygote
69
F
Japanese
She had corneal opacities and corneal rings
Yokota et al. (2002)
Homozygote
42
M
German
He had corneal opacities
Funke et al. (1991)
Homozygote Compound))
9. E136X
Deletion of C at codon 136
Heterozygote
10. A154fsX178 Shinbashi)
Deletion of GC at codon 154
Homozygote
51
11. H162fsX208 Sasebo
Insertion of 23 pb from codon 162
Homozygote
12. K184FSX200)
Deletion of C at codon 184
13. T202fsX200)
Deletion of G at codon 202
)
Frameshift mutation;
60 13
F M
Apolipoprotein A-I Mutations
þ
CAG->TAG at codon 84
Chapter j 7
Japanese Japanese
8. Q84X
))
Compound: compound heterozygote; CAD: coronary artery disease
137
138
The HDL Handbook
3 mg/dL. The patient of W8XWakayama was a 39-year-old Japanese male (Takata et al., 1995). He had corneal opacities, but no CAD or xanthomas. His plasma HDL-C and apoA-I levels were 6 and <3 mg/dL, respectively. The proband of Q32X was a 31-year-old female of Italian origin (Romling et al., 1994). She had bilateral periorbital xanthelasmas, but did not suffer from clinical signs of CAD. The proband of Q84X was a 60-year-old Japanese woman (Matsunaga et al., 1991). She had planar xanthomas at 18 years of age and her CAD symptoms started at 52 years of age. She had no measurable plasma HDL-C or apoA-I. Another proband of Q84X was a 12-year-old Japanese boy, who was a compound heterozygote with a point mutation at nucleotide position -27 that changed ATAAATA of the putative ApoA-I gene TATA box signal sequence to ATACATA (Matsunaga et al., 1999). His plasma HDL-C and apoA-I concentrations were 5.4 and 13.3 mg/dL, respectively. The heterozygous E136X carriers (n ¼ 17) had markedly low plasma HDL-C levels (males, 23.2 7.7 mg/dL; females, 23.2 3.7 mg/dL), and among nine carriers aged 35 years, five men developed premature CAD (Dastani et al., 2006). The proband of A154fsX178Shinbashi was a 51-year-old woman, who was hospitalized due to severe heart failure with CAD (Ikewaki et al., 2004). She exhibited corneal opacities and planar xanthomas on the eyelids and elbows. Her plasma HDL-C and apoA-I concentrations were 3.1 mg/dL (0.08 mmol/L) and 1 mg/dL, respectively. A 50-year-old Japanese woman was the proband of H162fsX208Sasebo (Moriyama et al., 1996). She had corneal opacities, and bilateral upper palpebral, elbow, neck, knee and Achilles tendon xanthomas. Her plasma HDL-C and apoA-I concentrations were 5.4 mg/dL (0.14mmol/L) and 0.8 mg/dL, respectively. The patient of K184fsX200 was a 69-year-old Japanese woman (Yokota et al., 2002). Her HDL-C level was 5.0 mg/dL (0.13 mmol/L). She exhibited corneal opacities and corneal rings, but had no xanthomas, thickening of Achilles tendons or CAD. The proband of T202fsX229 was a 42-year-old man (Funke et al., 1991). He had massive corneal opacities, complete absence of HDL-C (0 mg/dL) and half-normal LCAT activity, but no CAD. His plasma apoA-I concentration was 3 mg/dL. The ratio of plasma CE to total cholesterol is used to measure the capacity of LCAT to generate CE in plasma. Studies on plasma apoA-I deficiency have shown that the ratio of plasma CE to total cholesterol decreases by z30% compared to the ratios in controls (Ng et al., 1994; Romling et al., 1994; Parks et al., 1995). It has been argued that non-synthetic forms of apoA-I may be atherogenic. A-I/C-III gene inversion and gene deletion of the entire ApoA-I/ C-III/A-IV locus could prevent the synthesis of apoA-I, and premature CAD was observed. CAD was not always present in other patients with single gene defects of ApoA-I (see Table 7.1). Almost all cases with homozygous defects of ApoA-I were characterized by the absence of plasma HDL-C and apoA-I, but premature CAD was only present in four probands (Q-2X, Q5fsX11, Q84X and A154fsX178Shinbashi). However, four probands (Q5fsX36, W8XWakayama, Q32X
Chapter j 7
139
Apolipoprotein A-I Mutations
and T202fsX229) were <50 years of age, and these patients may have been too young for clinical manifestations of atherosclerosis to occur.
APOLIPOPROTEIN A-I VARIANTS Apolipoprotein A-I variants with or without low HDL-C The naturally occurring ApoA-I variants that produce pathological phenotypes are shown in Figure 7.1. It has been estimated that structural mutations in ApoA-I occur in 0.3% of the Japanese population and may affect the plasma HDL levels (Yamakawa-Kobayashi et al., 1999). Among a total of 50 reported natural missense or in-frame deletion mutations of ApoA-I, 16 variants are associated with low HDL levels (Weisgraber et al., 1980; Mahley et al., 1984; Rall et al., 1984; Utermann et al., 1984; von Eckardstein et al., 1989; Deeb et al., 1991; Tilly-Kiesi et al., 1995; Leren et al., 1997; Miccoli et al., 1997; Miettinen et al., 1997; Huang et al., 1998; Lindholm et al., 1998; Miller et al., 1998; Daum et al., 1999a,b; Han et al., 1999; Recalde et al., 2001; MartinCampos et al., 2002; Hovingh et al., 2004). Moreover, nine mutations between Amyloidosis
a
b c,d e 66
44
NH2
f g
h 88
j kl m
i 99
121
Helix 1* Helix 2 Helix 3 Helix 4
Reduced HDL-C levels 1
No apparent effect
1 23
4 5
6
7
Reduced HDL-C levels 1. K107del [Rall et al., 1984, Tilly-Kiesi et al., 1995] 2. L141R Pisa [Miccoli et al., 1997] 3. P143R Giessen [Utermann et al., 1984] 4. L144R Zaragoza [Recalde et al., 2001] 5. E146-R160del Seattle [Deeb et al., 1991, Lindholm et al., 1998] 6. R151C Paris [Daum et al., 1999a] 7. V156E Oita** [Huang et al., 1998] 8. A158E [Mahley et al., 1984] 9. L159R Fin [Miettinen et al., 1997] 10. L159P Zavalla [Miller et al., 1998] 11. R160L Oslo [Leren et al., 1997] 12. P165R [von Eckardstein et al., 1989, Daum et al., 1999b] 13. P165-A175del Mallorca [Martin-Campos et al., 2002] 14. R173C Milano [Weisgraber et al., 1980] 15. L178P [Hovingh et al., 2004] 16. E235del Nichinan [Han et al., 1999]
8
9 1011 1314 12
143
Helix 5
165
Helix 6
188
Helix 7
209
220
11 12 13 14 15 2 4 6 8 3 5 7 910
1516
17
18
Amyloidosis a. G26R Iowa [Nichols et al., 1988] b. W50R [Booth et al., 1995] c. L60R [Soutar et al., 1992] d. L60-F71del;INS, V,T [Booth et al., 1996] e. L64P [Murphy et al., 2004] f. E70-W72del [Persey et al., 1998] g. L75P [Coriu et al., 2003, Obici et al., 2004] h. L90P [Hamidi et al., 1999a] i. K107del [Amarzguioui et al., 1998, Mucchiano et al., 2001] j. R173P [Hamidi et al., 1999b] k. L174S [Obici et al.,1999] l. A175P [Lachmann et al., 2002] m. L178H [de Sousa et al., 2000]
19 20
243
Helix 8 Helix 9 Helix 10
COOH
16
21
22
No apparent effect
1. P3R Munster 3C [von Echardstein et al., 1989] 2. P3H [von Echardstein et al., 1989] 3. P4R Munster 3B [von Echardstein et al., 1989] 4. R10L Baltimore [Ladias et al., 1990] 5. D13Y Yame [Takada et al., 1991] 6. A37T [Matsunaga et al., 1991, Araki et al., 1994] 7. D51V Kaho [Moriyama et al., 1996a] 8. D89E [von Echardstein et al., 1990] 9. A95D Hita [Araki et al., 1994] 10. Y100H Karatsu [Moriyama et al., 1996b] 11. D103N Munster 3A [von Echardstein et al., 1990] 12. K107M [Breslow et al., 1988] 13. W108R Tsushima [Araki et al., 1994] 14. E110K Fukuoka [Takada et al., 1990] 15. E136K Norway [Mahley et al., 1984] 16. E139G [von Echardstein et al., 1990] 17. E147V [von Echardstein et al., 1990] 18. H162Q Kurume [Moriyama et al., 1996b] 19. E169Q [von Echardstein et al., 1990] 20. R177H [Assmann et al., 1993] 21. E198K Munster 4 [Strobl et al., 1988] 22. D213G Munster 3D [Mahely et al., 1984]
FIGURE 7.1 Naturally occurring apoA-I missense mutations or in-frame deletions that produce pathological phenotypes. * COOH-terminal residues 44e243 of apoA-I contain 22- and 11-amino acid repeats, based on X-ray crystallography and computer modeling, that are organized in 10 amphipathic a-helices (Nolte and Atkinson, 1992). ** ApoA-I V156EOita was identified as a homozygous mutation associated with apoA-I deficiency.
140
The HDL Handbook
residues 26 and 107 and four mutations between residues 173 and 178 are associated with low HDL levels and amyloidosis. The apoA-I V156EOita variant was associated with apoA-I deficiency. Missense or in-frame deletion mutations were found as heterozygotes excluding V156EOita. The carriers of the homozygous V156EOita variant were Japanese brothers aged 67 and 71 years (Huang et al., 1998). They had corneal opacities, but no xanthomas. The younger brother underwent aortocoronary bypass surgery, but the elder brother had no signs of CAD. Their plasma HDL-C and apoA-I concentrations were 7.3e9.3 mg/dL (0.19e0.24 mmol/L) and 11 mg/dL, respectively. The LCAT activity and cholesterol esterification rate were about 40% of the corresponding normal control values. The mean plasma HDL-C concentration (26.7 mg/dL) of 10 heterozygous carriers of apoA-I K107del was 36% lower than that of controls (Tilly-Kiesi et al., 1995). After screening for this apoA-I mutation in autopsy cases with aortic amyloidosis, one case with extensive aortic intimal amyloid deposits showed apoA-I K107del (Amarzguioui et al., 1998). The proband and carriers of L159RFin had diminished concentrations of plasma HDL-C (7.3e15.5 mg/ dL) and apoA-I (21.9e47.8 mg/dL) (Miettinen et al., 1997). There was no clinical evidence of CAD. Heterozygous carriers (n ¼ 19) of apoA-I L159PZavalla had reduced mean HDL-C (10.4 mg/dL) and apoA-I (38.7 mg/dL) levels (Miller et al., 1998). LCAT activity was slightly lower in some carriers, but the cholesterol esterification rate did not differ from that of controls. Two sisters aged 68 and 72 years with apoA-IZavalla had clinically manifested CAD. Four subjects with apoA-I E235delNichinan were heterozygous carriers, and their mean plasma concentrations of apoA-I and HDL-C were 30% and 32% lower than those in control subjects, respectively (Han et al., 1999). The apoA-I E235delNichinan variant was associated with reduced lipid-binding properties (Huang et al., 2000). Some apoA-I mutations supposedly underlie CAD (e.g., L141RPisa, L159PZavalla) (Miccoli et al., 1996; Miller et al., 1998), whereas others show no effect (e.g., L144RZaragoza, P165-A175delMallorca, E235delNichinan) (Han et al., 1999; Recalde et al., 2001; Martin-Campos et al., 2002) or even protect against CAD (e.g., R173CMilano). A catalog of the naturally occurring human apoA-I mutations associated with low HDL concentrations suggests that disruption of the native structure of apoA-I dramatically alters the intravascular metabolism of HDL (see Figure 7.1) (Sorci-Thomas and Thomas, 2002). While most of the apoA-I helices participate in lipid binding, various helical regions also have other functions. For example, the helices spanning residues 143e187 are critical for LCAT activation (Minnich et al., 1992; Sorci-Thomas et al., 1993; Holvoet et al., 1995), while the helices from residue 192 to the COOH-terminus are involved in interactions with cells and PL bilayers (Ji and Jonas, 1995; Sviridov et al., 1996). Among the documented apoA-I mutants associated with low HDL-C levels that span residues 107 to 235 of apoA-I but are predominantly clustered on or in the vicinity of helix 6, 81% of the mutations are amino acid
Chapter j 7
Apolipoprotein A-I Mutations
141
substitutions/deletions within helices 6 and 7 (see Figure 7.1) (Sorci-Thomas and Thomas, 2002). The associations between mutant forms of helices 6 and 7 and aberrant HDL metabolism have been studied in mice by several different approaches. For example, transgenic mice expressing a mutant form of apoA-I that lacked the entire proline-punctuated 22-amino acid helix 6 (residues 143e165) were created (Baralle and Baralle, 2000; Sorci-Thomas et al., 2000). In other studies, an adenoviral construct expressing a dominant-negative helix 6 point mutation, L159RFin apoA-I, was expressed in apoA-I-knockout mice, and a targeted replacement for the dominant-negative mutant apoA-I R173CMilano, a helix 7 point mutation, was examined (McManus et al., 2001; Parolini et al., 2005). Both of these mutants decreased the plasma HDL-C and apoA-I levels. Recently, Koukos et al. (2007a) used adenovirus-mediated gene transfer of mutants in apoA-I-knockout mice and revealed that apoA-I L141RPisa and L159RFin inhibit an early step in the biogenesis of HDL due to inefficient esterification of the cholesterol in preb1-HDL particles by endogenous LCAT, and that both defects can be corrected by treatment with LCAT. They also showed that the apoA-I levels in apoA-I-knockout mice expressing the apoA-I R160LOslo and R151CParis mutants were reduced by 68% and 55%, respectively, and that apoA-I R151CParis generated subpopulations of different sizes that migrated between preb- and a-HDL particles and formed mostly spherical and a few discoidal particles (Koukos et al., 2007b). An artificial point mutation (pR160V/H162A) and deletion (L144-P165del) in helix 6 were also associated with abolished LCAT activity and reduced plasma spherical HDL (Chroni et al., 2005). In each of these studies, the data suggest that mutations within apoA-I helices 6 and/or 7 cause decreased production of HDL and contribute to the dominant-negative phenotype observed in individuals heterozygous for these mutations. Besides the structural apoA-I mutations, plasma apoA-I is subjected to post-translational modifications in plasma that alter its functions. ApoA-I bound to HDL is subject to oxidation by products of lipid peroxidation at its methionine residues 112 and 148 (von Eckardstein et al., 1991; Panzenbock and Stocker, 2005). The frequencies of apoA-I mutations without any apparent effect on HDL-C may not be that low in the general population. A total of 22 apoA-I mutations have been reported to show no association with low HDL-C concentrations or amyloidosis (see Figure 7.1) (Breslow, 1988; Strobl et al., 1988; von Eckardstein et al., 1989, 1990; Ladias et al., 1990; Takada et al., 1990; 1991; Matsunaga et al., 1991; Assmann et al., 1993; Araki et al., 1994; Moriyama et al., 1996a,b). For example, the frequency of the heterozygous A37T variant was found to be 8% in normal Japanese high school students (Araki et al., 1994).
ApoA-I variants related to amyloidosis ApoA-I amyloidosis is an autosomal dominant amyloidosis caused by mutations in the ApoA-I gene. The first apoA-I variant associated with amyloidosis
142
The HDL Handbook
was apoA-I G26RIowa found in an Iowa kindred with British ancestry (Van Allen et al., 1969). Since the identification of apoA-I G26RIowa, 12 additional apoA-I variants have been found to be associated with amyloidosis, of which nine are single amino acid substitutions, two are deletions and one is a deletion/ insertion (see Figure 7.1) (Nichols et al., 1988; Soutar et al., 1992; Booth et al., 1995; Amarzguioui et al., 1998; Persey et al., 1998; Hamidi Asl et al., 1999a,b; Obici et al., 1999, 2004; de Sousa et al., 2000; Mucchiano et al., 2001; Lachmann et al., 2002; Coriu et al., 2003; Murphy et al., 2004). The mechanisms by which variants of apoA-I form amyloid deposits are unknown. Normal apoA-I is itself weakly amyloidogenic, and acts as the precursor of small amyloid deposits that occur quite frequently in aortic atherosclerotic plaques (Amarzguioui et al., 1998). In a mutated form, apoA-I represents the amyloidogenic precursor in some cases of familial amyloidosis. N-terminal fragments of apoA-I, corresponding to the first 83e93 residues, have been identified as the main components of apoA-I amyloid fibrils (Gregorini et al., 2005). Many amyloidogenic apoA-I mutations that are associated with deposition of amyloid fibrils predominantly in the liver, kidney and heart have been described, while other tissues and organs that are less frequently involved include the skin, testes, larynx and peripheral nerves. Although an extensive visceral amyloid load is frequently observed as a postmortem finding or by means of serum amyloid P component scintigraphy in patients with progressive kidney or heart disease (Soutar et al., 1992; Persey et al., 1998; Obici et al., 1999), deterioration of liver function is seldom observed. The majority of patients eventually develop renal failure, but liver function usually remains well preserved, despite extensive hepatic amyloid deposition. Penetrance in hereditary apoA-I amyloidosis is high, although very substantial visceral amyloid deposits may be present for decades before any symptoms develop. Depending on the mutation, patients can present with massive abdominal visceral amyloid involvement, predominant cardiomyopathy or a familial amyloidotic polyneuropathy-like syndrome. The features of amyloidosis with apoA-I G26RIowa included peripheral neuropathy, peptic ulcer disease, nephropathy and death from renal failure (Van Allen et al., 1969). The amyloid fibrils from an affected member of the original Iowa kindred showed a protein that was identical to the NH2terminal portion of wild-type apoA-I, apart from a substitution at position 26 (G26R), with no normal apoA-I (Nichols et al., 1988). The next two variants discovered were apoA-I W50R and L60R (Soutar et al., 1992; Booth et al., 1996). These variants had the following similarities with apoA-I G26RIowa: the affected subjects were heterozygotes; in each subject an arginine replaced a neutral residue in the NH2-terminal, shifting the net charge by one positive unit; the amyloid fibrils consisted of NH2-terminal fragments of the variant apoA-I; each mutation was associated with hereditary non-neuropathic systemic amyloidosis. The apoA-I L60-F71del;INS, V,T
Chapter j 7
Apolipoprotein A-I Mutations
143
and E70-W72del mutants also resulted in a net charge gain of þ1. Although the charge substitution does not preclude a conformational change, as has been predicted for many mutations of transthyretin as another amyloidogenic protein, it is obvious that a change in charge is not required for fibrillogenesis. Furthermore, the mutant residue is not structurally required for fibril formation, based on the absence of mutant fragments in the amyloid fibrils of skin deposits, but their presence in the cardiac amyloid (Hamidi Asl et al., 1999a). K107del has been associated with increased susceptibility to amyloid deposition in the aortic intima and ischemic heart disease. Several variants, including proline substitutions at residues 90, 173 and 175 and histidine substitution at residue 178, are associated with hoarseness due to laryngeal amyloid deposits which may be misdiagnosed as localized primary amyloidosis (AL type) (Hamidi Asl et al., 1999a,b; de Sousa et al., 2000; Lachmann et al., 2002).
Cysteine mutants of apoA-I Apolipoprotein A-IMilano was the first described mutant of apoA-I. In the 1980s, a small Italian community was found to have a mutant version of apoA-I that was associated with a decreased risk of arteriosclerosis, heart attack and stroke. Individuals with this mutation are characterized by very low HDL-C, but no increased risk of atherosclerosis (Weisgraber et al., 1980). ApoA-I is a lipid-binding protein that forms complexes with other proteins (like apoA-II) and lipids to form HDL particles. ApoA-I can self-associate, and normally forms modest amounts of dimers, trimers and quadramers as well as being in the monomeric form, and produces HDL particles in a range of sizes (Calabresi et al., 1997). In apoA-IMilano, the basic amino acid arginine at position 173 is mutated to the sulfur-containing amino acid cysteine (R173C). This results in a greater ability to form stable dimers than normal apoA-I, since two cysteines can form a chemical bond together via their sulfur groups (Calabresi et al., 1997). ApoA-IMilano can lead to the formation of homodimers and heterodimers with apoA-II. Since the affected individuals are heterozygous for this mutation, the full impact of apoA-IMilano on HDL-C metabolism is unknown. ApoA-IMilano shows faster catabolism than normal apoA-I, thereby explaining the accompanying low HDL-C (Chiesa and Sirtori, 2003). Reduced plasma HDL-C and a decreased risk of vascular disease were observed in another cysteine variant, apoA-IParis, in which Arg151 is substituted by cysteine (Bruckert et al., 1997; Daum et al., 1999a). Both apoA-I R173CMilano and apoA-I R151CParis are rare cysteine variants that produce low HDL-C in the absence of cardiovascular disease in humans. Although there are no differences between lipid-free apoA-IMilano, apoA-IParis and normal apoA-I in mediating the efflux of cholesterol from macrophages, apoA-IMilano is twice as effective as apoA-IParis in preventing lipoxygenasemediated oxidation of PL (Bielicki and Oda, 2002).
144
The HDL Handbook
Antiatherogenic apoA-I and apoA-IMilano Liver-specific overexpression of a wild-type ApoA-I transgene was found to increase HDL-C and apoA-I levels, thereby reducing atherosclerosis in C57BL/6 mice (Rubin et al., 1991) and hyperlipidemic mice (Paszty et al., 1994; Plump et al., 1994). A significant reduction in atherosclerosis was found after somatic gene transfer of human apoA-I into low-density lipoprotein (LDL) receptor-knockout mice fed a high-fat diet, and mice lacking the ApoA-I gene had significantly increased atherosclerosis. Over the last few decades, several sets of apoA-I variants with different functions and structural correlations to HDL have been generated and characterized, and their clinical potentials have been assessed in animal and human models. This essential field is generally referred to as HDL-therapy (Newton and Krause, 2002; Brewer, 2004). Recombinant apoA-IMilano displays remarkable atheroprotective activities and suggests the possibility of directly reducing the burden of atherosclerosis in experimental models. ApoA-IMilano has been studied extensively with regard to its effects on atherosclerosis, including infusion and genetic expression in animals (Shah et al., 1998; Chiesa and Sirtori, 2003). It was previously shown that intravenous recombinant apoA-IMilano inhibits progression and induces rapid regression and remodeling of atherosclerosis in hypercholesterolemic rabbits and mice, while attenuating endothelial dysfunction (Ameli et al., 1994; Shah et al., 1998, 2001; Chiesa et al., 2002; Kaul et al., 2004). Based on these promising preclinical results, a small phase II human trial of patients with acute coronary syndromes was conducted, in which once-weekly intravenous injections of recombinant apoA-IMilano/PL complexes in patients with CAD were shown to induce rapid coronary atheroma regression within 5 weeks (Nissen et al., 2003). Gene transfer with hepatic expression of wild-type apoA-I or apoA-IMilano was found to result in plasma levels of apoA-I that were sufficient to significantly slow the development of atherosclerosis in mice, and both forms were similarly effective in their ability to retard atherosclerosis development in LDL receptor-knockout mice (Lebherz et al., 2007). A recent study revealed a superior atheroprotective effect of apoAIMilano compared to wild-type apoA-I using transduced macrophages after bone marrow transplantation into ApoA-I/ApoE double-knockout mice (Wang et al., 2006). It has been suggested that apoA-IMilano may be a gain-of-function mutation and the prototype of apoA-I mimetics.
REFERENCES Amarzguioui, M., Mucchiano, G., Haggqvist, B., et al. (1998). Extensive intimal apolipoprotein A1-derived amyloid deposits in a patient with an apolipoprotein A1 mutation. Biochem Biophys Res Commun, 242, 534e539. Ameli, S., Hultgardh-Nilsson, A., Cercek, B., et al. (1994). Recombinant apolipoprotein A-I Milano reduces intimal thickening after balloon injury in hypercholesterolemic rabbits. Circulation, 90, 1935e1941.
Chapter j 7
Apolipoprotein A-I Mutations
145
Araki, K., Sasaki, J., Matsunaga, A., et al. (1994). Characterization of two new human apolipoprotein A-I variants: apolipoprotein A-I Tsushima (Trp-108e>Arg) and A-I Hita (Ala-95e>Asp). Biochim Biophys Acta, 1214, 272e278. Assmann, G., von Eckardstein, A., & Funke, H. (1993). High density lipoproteins, reverse transport of cholesterol, and coronary artery disease. Insights from mutations. Circulation, 87, III28e34. Baralle, M., & Baralle, F. E. (2000). Genetics and molecular biology. Curr Opin Lipidol, 11, 653e656. Bielicki, J. K., & Oda, M. N. (2002). Apolipoprotein A-I(Milano) and apolipoprotein A-I(Paris) exhibit an antioxidant activity distinct from that of wild-type apolipoprotein A-I. Biochemistry, 41, 2089e2096. Booth, D. R., Tan, S. Y., Booth, S. E., et al. (1995). A new apolipoprotein Al variant, Trp50Arg, causes hereditary amyloidosis. QJM, 88:695e702. Booth, D. R., Tan, S. Y., Booth, S. E., Tennent, G. A., Hutchinson, W. L., Hsuan, J. J., et al. (1996). Hereditary hepatic and systemic amyloidosis caused by a new deletion/insertion mutation in the apolipoprotein AI gene. J Clin Invest, 97, 2714e2721. Borhani, D. W., Rogers, D. P., Engler, J. A., & Brouillette, C. G. (1997). Crystal structure of truncated human apolipoprotein A-I suggests a lipid-bound conformation. Proc Natl Acad Sci USA, 94, 12291e12296. Breslow, J. L. (1988). Apolipoprotein genetic variation and human disease. Physiol Rev, 68, 85e132. Brewer, H. B., Jr. (2004). Focus on high-density lipoproteins in reducing cardiovascular risk. Am Heart J, 148, S14e18. Brouillette, C. G., & Anantharamaiah, G. M. (1995). Structural models of human apolipoprotein A-I. Biochim Biophys Acta, 1256, 103e129. Bruckert, E., von Eckardstein, A., Funke, H., Beucler, I., Wiebusch, H., Turpin, G., et al. (1997). The replacement of arginine by cysteine at residue 151 in apolipoprotein A-I produces a phenotype similar to that of apolipoprotein A-Imilano. Atherosclerosis, 128, 121e128. Calabresi, L., Vecchio, G., Frigerio, F., Vavassori, L., Sirtori, C. R., & Franceschini, G. (1997). Reconstituted high-density lipoproteins with a disulfide-linked apolipoprotein A-I dimer: evidence for restricted particle size heterogeneity. Biochemistry, 36, 12428e12433. Chiesa, G., & Sirtori, C. R. (2003). Apolipoprotein A-I(Milano): current perspectives. Curr Opin Lipidol, 14, 159e163. Chiesa, G., Monteggia, E., Marchesi, M., et al. (2002). Recombinant apolipoprotein A-I(Milano) infusion into rabbit carotid artery rapidly removes lipid from fatty streaks. Circ Res, 90, 974e980. Chroni, A., Duka, A., Kan, H. Y., Liu, T., & Zannis, V. I. (2005). Point mutations in apolipoprotein A-I mimic the phenotype observed in patients with classical lecithin:cholesterol acyltransferase deficiency. Biochemistry, 44, 14353e14366. Coriu, D., Dispenzieri, A., Stevens, F. J., et al. (2003). Hepatic amyloidosis resulting from deposition of the apolipoprotein A-I variant Leu75Pro. Amyloid, 10, 215e223. Dastani, Z., Dangoisse, C., Boucher, B., et al. (2006). A novel nonsense apolipoprotein A-I mutation (apoA-I(E136X)) causes low HDL cholesterol in French Canadians. Atherosclerosis, 185, 127e136. Daum, U., Langer, C., Duverger, N., et al. (1999a). Apolipoprotein A-I (R151C)Paris is defective in activation of lecithin: cholesterol acyltransferase but not in initial lipid binding, formation of reconstituted lipoproteins, or promotion of cholesterol efflux. J Mol Med, 77, 614e622.
146
The HDL Handbook
Daum, U., Leren, T. P., Langer, C., et al. (1999b). Multiple dysfunctions of two apolipoprotein A-I variants, apoA-I(R160L)Oslo and apoA-I(P165R), that are associated with hypoalphalipoproteinemia in heterozygous carriers. J Lipid Res, 40, 486e494. Davidson, W. S., & Thompson, T. B. (2007). The structure of apolipoprotein A-I in high density lipoproteins. J Biol Chem, 282, 22249e22253. de Sousa, M. M., Vital, C., Ostler, D., et al. (2000). Apolipoprotein AI and transthyretin as components of amyloid fibrils in a kindred with apoAI Leu178His amyloidosis. Am J Pathol, 156, 1911e1917. Deeb, S. S., Cheung, M. C., Peng, R. L., et al. (1991). A mutation in the human apolipoprotein A-I gene. Dominant effect on the level and characteristics of plasma high density lipoproteins. J Biol Chem, 266, 13654e13660. Fielding, C. J., & Fielding, P. E. (1995). Molecular physiology of reverse cholesterol transport. J Lipid Res, 36, 211e228. Funke, H., von Eckardstein, A., Pritchard, P. H., Karas, M., Albers, J. J., & Assmann, G. (1991). A frameshift mutation in the human apolipoprotein A-I gene causes high density lipoprotein deficiency, partial lecithin: cholesterol-acyltransferase deficiency, and corneal opacities. J Clin Invest, 87, 371e376. Gelissen, I. C., Harris, M., Rye, K. A., et al. (2006). ABCA1 and ABCG1 synergize to mediate cholesterol export to apoA-I. Arterioscler Thromb Vasc Biol, 26, 534e540. Gregorini, G., Izzi, C., Obici, L., et al. (2005). Renal apolipoprotein A-I amyloidosis: a rare and usually ignored cause of hereditary tubulointerstitial nephritis. J Am Soc Nephrol, 16, 3680e3686. Hamidi Asl, L., Liepnieks, J. J., Hamidi Asl, K., et al. (1999a). Hereditary amyloid cardiomyopathy caused by a variant apolipoprotein A1. Am J Pathol, 154, 221e227. Hamidi Asl, K., Liepnieks, J. J., Nakamura, M., Parker, F., & Benson, M. D. (1999b). A novel apolipoprotein A-1 variant, Arg173Pro, associated with cardiac and cutaneous amyloidosis. Biochem Biophys Res Commun, 257, 584e588. Han, H., Sasaki, J., Matsunaga, A., et al. (1999). A novel mutant, ApoA-I nichinan (Glu235e>0), is associated with low HDL cholesterol levels and decreased cholesterol efflux from cells. Arterioscler Thromb Vasc Biol, 19, 1447e1455. Hobbs, H. H., & Rader, D. J. (1999). ABC1: connecting yellow tonsils, neuropathy, and very low HDL. J Clin Invest, 104, 1015e1017. Holvoet, P., Zhao, Z., Vanloo, B., et al. (1995). Phospholipid binding and lecithin-cholesterol acyltransferase activation properties of apolipoprotein A-I mutants. Biochemistry, 34, 13334e13342. Hovingh, G. K., Brownlie, A., Bisoendial, R. J., et al. (2004). A novel apoA-I mutation (L178P) leads to endothelial dysfunction, increased arterial wall thickness, and premature coronary artery disease. J Am Coll Cardiol, 44, 1429e1435. Huang, W., Sasaki, J., Matsunaga, A., et al. (1998). A novel homozygous missense mutation in the apo A-I gene with apo A-I deficiency. Arterioscler Thromb Vasc Biol, 18, 389e396. Huang, W., Sasaki, J., Matsunaga, A., Han, H., Li, W., Koga, T., et al. (2000). A single amino acid deletion in the carboxy terminal of apolipoprotein A-I impairs lipid binding and cellular interaction. Arterioscler Thromb Vasc Biol, 20, 210e216. Ikewaki, K., Matsunaga, A., Han, H., et al. (2004). A novel two nucleotide deletion in the apolipoprotein A-I gene, apoA-I Shinbashi, associated with high density lipoprotein deficiency, corneal opacities, planar xanthomas, and premature coronary artery disease. Atherosclerosis, 172, 39e45.
Chapter j 7
Apolipoprotein A-I Mutations
147
Ji, Y., & Jonas, A. (1995). Properties of an N-terminal proteolytic fragment of apolipoprotein AI in solution and in reconstituted high density lipoproteins. J Biol Chem, 270, 11290e11297. Karathanasis, S. K., Ferris, E., & Haddad, I. A. (1987). DNA inversion within the apolipoproteins AI/CIII/AIV-encoding gene cluster of certain patients with premature atherosclerosis. Proc Natl Acad Sci USA, 84, 7198e7202. Kaul, S., Coin, B., Hedayiti, A., et al. (2004). Rapid reversal of endothelial dysfunction in hypercholesterolemic apolipoprotein E-null mice by recombinant apolipoprotein A-I(Milano)phospholipid complex. J Am Coll Cardiol, 44, 1311e1319. Koukos, G., Chroni, A., Duka, A., Kardassis, D., & Zannis, V. I. (2007a). LCAT can rescue the abnormal phenotype produced by the natural ApoA-I mutations (Leu141Arg)(Pisa) and (Leu159Arg)(FIN). Biochemistry, 406, 167e174. Koukos, G., Chroni, A., Duka, A., Kardassis, D., & Zannis, V. I. (2007b). Naturally occurring and bioengineered apoA-I mutations that inhibit the conversion of discoidal to spherical HDL: the abnormal HDL phenotypes can be corrected by treatment with LCAT. Biochem J, 406, 167e174. Kuivenhoven, J. A., Pritchard, H., Hill, J., Frohlich, J., Assmann, G., & Kastelein, J. (1997). The molecular pathology of lecithin:cholesterol acyltransferase (LCAT) deficiency syndromes. J Lipid Res, 38, 191e205. Lachmann, H. J., Booth, D. R., Booth, S. E., et al. (2002). Misdiagnosis of hereditary amyloidosis as AL (primary) amyloidosis. N Engl J Med, 346, 1786e1791. Lackner, K. J., Dieplinger, H., Nowicka, G., & Schmitz, G. (1993). High density lipoprotein deficiency with xanthomas. A defect in reverse cholesterol transport caused by a point mutation in the apolipoprotein A-I gene. J Clin Invest, 92, 2262e2273. Ladias, J. A., Kwiterovich, P. O., Jr., Smith, H. H., Karathanasis, S. K., & Antonarakis, S. E. (1990). Apolipoprotein A1 Baltimore (Arg10e>Leu), a new apoA1 variant. Hum Genet, 84, 439e445. Lebherz, C., Sanmiguel, J., Wilson, J. M., & Rader, D. J. (2007). Gene transfer of wild-type apoAI and apoA-I Milano reduce atherosclerosis to a similar extent. Cardiovasc Diabetol, 6, 15. Leren, T. P., Bakken, K. S., Daum, U., et al. (1997). Heterozygosity for apolipoprotein A-I(R160L) Oslo is associated with low levels of high density lipoprotein cholesterol and HDL-subclass LpA-I/A-II but normal levels of HDL-subclass LpA-I. J Lipid Res, 38, 121e131. Lindholm, E. M., Bielicki, J. K., Curtiss, L. K., Rubin, E. M., & Forte, T. M. (1998). Deletion of amino acids Glu146e>Arg160 in human apolipoprotein A-I (ApoA-ISeattle) alters lecithin: cholesterol acyltransferase activity and recruitment of cell phospholipid. Biochemistry, 37, 4863e4868. Mahley, R. W., Innerarity, T. L., Rall, S. C., Jr., & Weisgraber, K. H. (1984). Plasma lipoproteins: apolipoprotein structure and function. J Lipid Res, 25, 1277e1294. Martin-Campos, J. M., Julve, J., Escola, J. C., et al. (2002). ApoA-I(MALLORCA) impairs LCAT activation and induces dominant familial hypoalphalipoproteinemia. J Lipid Res, 43, 115e123. Matsunaga, A., Sasaki, J., Han, H., et al. (1999). Compound heterozygosity for an apolipoprotein A1 gene promoter mutation and a structural nonsense mutation with apolipoprotein A1 deficiency. Arterioscler Thromb Vasc Biol, 19, 348e355. Matsunaga, T., Hiasa, Y., Yanagi, H., et al. (1991). Apolipoprotein A-I deficiency due to a codon 84 nonsense mutation of the apolipoprotein A-I gene. Proc Natl Acad Sci USA, 88, 2793e2797. McManus, D. C., Scott, B. R., Franklin, V., Sparks, D. L., & Marcel, Y. L. (2001). Proteolytic degradation and impaired secretion of an apolipoprotein A-I mutant associated with dominantly inherited hypoalphalipoproteinemia. J Biol Chem, 276, 21292e21302.
148
The HDL Handbook
Miccoli, R., Bertolotto, A., Navalesi, R., et al. (1996). Compound heterozygosity for a structural apolipoprotein A-I variant, apo A-I(L141R)Pisa, and an apolipoprotein A-I null allele in patients with absence of HDL cholesterol, corneal opacifications, and coronary heart disease. Circulation, 94, 1622e1628. Miccoli, R., Zhu, Y., Daum, U., et al. (1997). A natural apolipoprotein A-I variant, apoA-I (L141R) Pisa, interferes with the formation of alpha-high density lipoproteins (HDL) but not with the formation of pre beta 1-HDL and influences efflux of cholesterol into plasma. J Lipid Res, 38, 1242e1253. Miettinen, H. E., Gylling, H., Miettinen, T. A., Viikari, J., Paulin, L., & Kontula, K. (1997). Apolipoprotein A-IFin. Dominantly inherited hypoalphalipoproteinemia due to a single base substitution in the apolipoprotein A-I gene. Arterioscler Thromb Vasc Biol, 17, 83e90. Miida, T., Zhang, B., Obayashi, K., et al. (2004). T13M mutation of lecithin-cholesterol acyltransferase gene causes fish-eye disease. Clin Chim Acta, 343, 201e208. Miller, M., Aiello, D., Pritchard, H., Friel, G., & Zeller, K. (1998). Apolipoprotein A-I(Zavalla) (Leu159e>Pro): HDL cholesterol deficiency in a kindred associated with premature coronary artery disease. Arterioscler Thromb Vasc Biol, 18, 1242e1247. Miller, M., Rhyne, J., Hamlette, S., Birnbaum, J., & Rodriguez, A. (2003). Genetics of HDL regulation in humans. Curr Opin Lipidol, 14, 273e279. Minnich, A., Collet, X., Roghani, A., et al. (1992). Site-directed mutagenesis and structurefunction analysis of the human apolipoprotein A-I. Relation between lecithin-cholesterol acyltransferase activation and lipid binding. J Biol Chem, 267, 16553e16560. Moriyama, K., Sasaki, J., Arakawa, F., et al. (1995). Two novel point mutations in the lecithin: cholesterol acyltransferase (LCAT) gene resulting in LCAT deficiency: LCAT (G873 deletion) and LCAT (Gly344e>Ser). J Lipid Res, 36, 2329e2343. Moriyama, K., Sasaki, J., Matsunaga, A., & Arakawa, K. (1996a). Identification of two apolipoprotein variants, A-I Kaho (Asp 51e>Val) and A-I Lys 107 deletion. J Atheroscler Thromb, 3, 12e16. Moriyama, K., Sasaki, J., Matsunaga, A., Takada, Y., Kagimoto, M., & Arakawa, K. (1996b). Identification of two apolipoprotein variants, A-I Karatsu (Tyr 100e>His) and A-I Kurume (His 162e>Gln). Clin Genet, 49, 79e84. Moriyama, K., Sasaki, J., Takada, Y., et al. (1996). A cysteine-containing truncated apo A-I variant associated with HDL deficiency. Arterioscler Thromb Vasc Biol, 16, 1416e1423. Mucchiano, G. I., Haggqvist, B., Sletten, K., & Westermark, P. (2001). Apolipoprotein A-1derived amyloid in atherosclerotic plaques of the human aorta. J Pathol, 193, 270e275. Murphy, C. L., Wang, S., Weaver, K., Gertz, M. A., Weiss, D. T., & Solomon, A. (2004). Renal apolipoprotein A-I amyloidosis associated with a novel mutant Leu64Pro. Am J Kidney Dis, 44, 1103e1109. Nakata, K., Kobayashi, K., Yanagi, H., et al. (1993). Autosomal dominant hypoalphalipoproteinemia due to a completely defective apolipoprotein A-I gene. Biochem Biophys Res Commun, 196, 950e955. Newton, R. S., & Krause, B. R. (2002). HDL therapy for the acute treatment of atherosclerosis. Atheroscler Suppl, 3, 31e38. Ng, D. S., Leiter, L. A., Vezina, C., Connelly, P. W., & Hegele, R. A. (1994). Apolipoprotein A-I Q (-2)X causing isolated apolipoprotein A-I deficiency in a family with analphalipoproteinemia. J Clin Invest, 93, 223e229. Nichols, W. C., Dwulet, F. E., Liepnieks, J., & Benson, M. D. (1988). Variant apolipoprotein AI as a major constituent of a human hereditary amyloid. Biochem Biophys Res Commun, 156, 762e768.
Chapter j 7
Apolipoprotein A-I Mutations
149
Nissen, S. E., Tsunoda, T., Tuzcu, E. M., et al. (2003). Effect of recombinant ApoA-I Milano on coronary atherosclerosis in patients with acute coronary syndromes: a randomized controlled trial. J Am Med Assoc, 290, 2292e2300. Nolte, R. T., & Atkinson, D. (1992). Conformational analysis of apolipoprotein A-I and E-3 based on primary sequence and circular dichroism. Biophys J, 63, 1221e1239. Norum, R. A., Lakier, J. B., Goldstein, S., et al. (1982). Familial deficiency of apolipoproteins A-I and C-III and precocious coronary-artery disease. N Engl J Med, 306, 1513e1519. Obici, L., Bellotti, V., Mangione, P., et al. (1999). The new apolipoprotein A-I variant leu(174) e> Ser causes hereditary cardiac amyloidosis, and the amyloid fibrils are constituted by the 93-residue N-terminal polypeptide. Am J Pathol, 155, 695e702. Obici, L., Palladini, G., Giorgetti, S., et al. (2004). Liver biopsy discloses a new apolipoprotein A-I hereditary amyloidosis in several unrelated Italian families. Gastroenterology, 126, 1416e1422. Ordovas, J. M., Cassidy, D. K., Civeira, F., Bisgaier, C. L., & Schaefer, E. J. (1989). Familial apolipoprotein A-I, C-III, and A-IV deficiency and premature atherosclerosis due to deletion of a gene complex on chromosome 11. J Biol Chem, 264, 16339e16342. Panzenbock, U., & Stocker, R. (2005). Formation of methionine sulfoxide-containing specific forms of oxidized high-density lipoproteins. Biochim Biophys Acta, 1703, 171e181. Parks, J. S., Li, H., Gebre, A. K., Smith, T. L., & Maeda, N. (1995). Effect of apolipoprotein A-I deficiency on lecithin:cholesterol acyltransferase activation in mouse plasma. J Lipid Res, 36, 349e355. Parolini, C., Chiesa, G., Gong, E., et al. (2005). Apolipoprotein A-I and the molecular variant apoA-I(Milano): evaluation of the antiatherogenic effects in knock-in mouse model. Atherosclerosis, 183, 222e229. Paszty, C., Maeda, N., Verstuyft, J., & Rubin, E. M. (1994). Apolipoprotein AI transgene corrects apolipoprotein E deficiency-induced atherosclerosis in mice. J Clin Invest, 94, 899e903. Persey, M. R., Booth, D. R., Booth, S. E., et al. (1998). Hereditary nephropathic systemic amyloidosis caused by a novel variant apolipoprotein A-I. Kidney Int, 53, 276e281. Plump, A. S., Scott, C. J., & Breslow, J. L. (1994). Human apolipoprotein A-I gene expression increases high density lipoprotein and suppresses atherosclerosis in the apolipoprotein E-deficient mouse. Proc Natl Acad Sci USA, 91, 9607e9611. Rall, S. C., Jr., Weisgraber, K. H., Mahley, R. W., et al. (1984). Abnormal lecithin:cholesterol acyltransferase activation by a human apolipoprotein A-I variant in which a single lysine residue is deleted. J Biol Chem, 259, 10063e10070. Recalde, D., Velez-Carrasco, W., Civeira, F., et al. (2001). Enhanced fractional catabolic rate of apo A-I and apo A-II in heterozygous subjects for apo A-I(Zaragoza) (L144R). Atherosclerosis, 154, 613e623. Romling, R., von Eckardstein, A., Funke, H., et al. (1994). A nonsense mutation in the apolipoprotein A-I gene is associated with high-density lipoprotein deficiency and periorbital xanthelasmas. Arterioscler Thromb, 14, 1915e1922. Rubin, E. M., Krauss, R. M., Spangler, E. A., Verstuyft, J. G., & Clift, S. M. (1991). Inhibition of early atherogenesis in transgenic mice by human apolipoprotein AI. Nature, 353, 265e267. Schaefer, E. J., Ordovas, J. M., Law, S. W., et al. (1985). Familial apolipoprotein A-I and C-III deficiency, variant II. J Lipid Res, 26, 1089e1101. Shah, P. K., Nilsson, J., Kaul, S., et al. (1998). Effects of recombinant apolipoprotein A-I(Milano) on aortic atherosclerosis in apolipoprotein E-deficient mice. Circulation, 97, 780e785. Shah, P. K., Yano, J., Reyes, O., Chyu, K. Y., Kaul, S., Bisgaier, C. L., et al. (2001). High-dose recombinant apolipoprotein A-I(milano) mobilizes tissue cholesterol and rapidly reduces plaque lipid and macrophage content in apolipoprotein e-deficient mice. Potential implications for acute plaque stabilization. Circulation, 103, 3047e3050.
150
The HDL Handbook
Sorci-Thomas, M. G., & Thomas, M. J. (2002). The effects of altered apolipoprotein A-I structure on plasma HDL concentration. Trends Cardiovasc Med, 12, 121e128. Sorci-Thomas, M., Kearns, M. W., & Lee, J. P. (1993). Apolipoprotein A-I domains involved in lecithin-cholesterol acyltransferase activation. Structure:function relationships. J Biol Chem, 268, 21403e21409. Sorci-Thomas, M. G., Thomas, M., Curtiss, L., & Landrum, M. (2000). Single repeat deletion in apoA-I blocks cholesterol esterification and results in rapid catabolism of delta6 and wild-type apoA-I in transgenic mice. J Biol Chem, 275, 12156e12163. Soutar, A. K., Hawkins, P. N., Vigushin, D. M., et al. (1992). Apolipoprotein AI mutation Arg-60 causes autosomal dominant amyloidosis. Proc Natl Acad Sci USA, 89, 7389e7393. Strobl, W., Jabs, H. U., Hayde, M., Holzinger, T., Assmann, G., & Widhalm, K. (1988). Apolipoprotein A-I (Glu 198e>Lys): a mutant of the major apolipoprotein of high-density lipoproteins occurring in a family with dyslipoproteinemia. Pediatr Res, 24, 222e228. Sviridov, D., Pyle, L. E., & Fidge, N. (1996). Efflux of cellular cholesterol and phospholipid to apolipoprotein A-I mutants. J Biol Chem, 271, 33277e33283. Takada, Y., Sasaki, J., Ogata, S., Nakanishi, T., Ikehara, Y., & Arakawa, K. (1990). Isolation and characterization of human apolipoprotein A-I Fukuoka (110 Glue>Lys). A novel apolipoprotein variant. Biochim Biophys Acta, 1043, 169e176. Takada, Y., Sasaki, J., Seki, M., Ogata, S., Teranishi, Y., & Arakawa, K. (1991). Characterization of a new human apolipoprotein A-I Yame by direct sequencing of polymerase chain reactionamplified DNA. J Lipid Res, 32, 1275e1280. Takata, K., Saku, K., Ohta, T., et al. (1995). A new case of apoA-I deficiency showing codon 8 nonsense mutation of the apoA-I gene without evidence of coronary heart disease. Arterioscler Thromb Vasc Biol, 15, 1866e1874. Tilly-Kiesi, M., Zhang, Q., Ehnholm, S., et al. (1995). ApoA-IHelsinki (Lys107e>0) associated with reduced HDL cholesterol and LpA-I:A-II deficiency. Arterioscler Thromb Vasc Biol, 15, 1294e1306. Uehara, Y., Tsuboi, Y., Zhang, B., et al. (2007). POPC/apoA-I discs as a potent lipoprotein modulator in Tangier disease. Atherosclerosis, 197, 283e289. Utermann, G., Haas, J., Steinmetz, A., et al. (1984). Apolipoprotein A-IGiessen (Pro143e>Arg). A mutant that is defective in activating lecithin:cholesterol acyltransferase. Eur J Biochem, 144, 325e331. Van Allen, M. W., Frohlich, J. A., & Davis, J. R. (1969). Inherited predisposition to generalized amyloidosis. Clinical and pathological study of a family with neuropathy, nephropathy, and peptic ulcer. Neurology, 19, 10e25. von Eckardstein, A., Funke, H., Henke, A., Altland, K., Benninghoven, A., & Assmann, G. (1989). Apolipoprotein A-I variants. Naturally occurring substitutions of proline residues affect plasma concentration of apolipoprotein A-I. J Clin Invest, 84, 1722e1730. von Eckardstein, A., Funke, H., Walter, M., Altland, K., Benninghoven, A., & Assmann, G. (1990). Structural analysis of human apolipoprotein A-I variants. Amino acid substitutions are nonrandomly distributed throughout the apolipoprotein A-I primary structure. J Biol Chem, 265, 8610e8617. von Eckardstein, A., Walter, M., Holz, H., Benninghoven, A., & Assmann, G. (1991). Site-specific methionine sulfoxide formation is the structural basis of chromatographic heterogeneity of apolipoproteins A-I, C-II, and C-III. J Lipid Res, 32, 1465e1476. Wang, L., Sharifi, B. G., Pan, T., Song, L., Yukht, A., & Shah, P. K. (2006). Bone marrow transplantation shows superior atheroprotective effects of gene therapy with apolipoprotein A-I Milano compared with wild-type apolipoprotein A-I in hyperlipidemic mice. J Am Coll Cardiol, 48, 1459e1468.
Chapter j 7
Apolipoprotein A-I Mutations
151
Weisgraber, K. H., Bersot, T. P., Mahley, R. W., Franceschini, G., & Sirtori, C. R. (1980). A-Imilano apoprotein. Isolation and characterization of a cysteine-containing variant of the A-I apoprotein from human high density lipoproteins. J Clin Invest, 66, 901e907. Williamson, R., Lee, D., Hagaman, J., & Maeda, N. (1992). Marked reduction of high density lipoprotein cholesterol in mice genetically modified to lack apolipoprotein A-I. Proc Natl Acad Sci USA, 89, 7134e7138. Yamakawa-Kobayashi, K., Yanagi, H., Fukayama, H., et al. (1999). Frequent occurrence of hypoalphalipoproteinemia due to mutant apolipoprotein A-I gene in the population: a population-based survey. Hum Mol Genet, 8, 331e336. Yokota, H., Hashimoto, Y., Okubo, S., et al. (2002). Apolipoprotein A-I deficiency with accumulated risk for CHD but no symptoms of CHD. Atherosclerosis, 162, 399e407. Zannis, V. I., Chroni, A., & Krieger, M. (2006). Role of apoA-I, ABCA1, LCAT, and SR-BI in the biogenesis of HDL. J Mol Med, 84, 276e294.
Chapter 8
The Scavenger Receptor Class B Type I: An HDL Receptor Involved in Lipid Transport and HDL Dependent Signaling Aishah Al-Jarallah, Rachelle Brunet and Bernardo Trigatti Department of Biochemistry and Biomedical Sciences, McMaster University, Hamilton, Ontario, Canada
OVERVIEW The scavenger receptor, class B, type I (SR-BI) is a palmitylated cell surface glycoprotein that binds high and low density lipoproteins (HDL and LDL) with high affinity. It was the first molecularly well defined HDL receptor that was shown to mediate the selective uptake of HDL lipids and HDL-dependent cholesterol efflux from cells. This chapter will begin with a discussion of the impact of studies from genetically engineered mice in revealing the significance of SR-BI in HDL metabolism, atherosclerosis and coronary heart disease. We will then discuss the cellular biology of SR-BI, its activities in lipid transfer between HDL and cells and more recent insights into its role in mediating the activation of cellular signaling pathways by HDL in various cell types, and their possible contribution to atheroprotection.
ROLE OF SR-BI IN HDL METABOLISM, ATHEROSCLEROSIS AND CORONARY HEART DISEASE e STUDIES FROM GENETICALLY ALTERED MICE The scavenger receptor, class B type I was first reported by Acton et al. (1996) to be a high affinity receptor for HDL that could mediate selective HDL lipid uptake into cells. Since then, studies of gene-targeted SR-BI knockout mice have shed light on its physiological role in HDL cholesterol clearance by the liver, driving reverse cholesterol transport, and its role in both liver and bone The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
153
154
The HDL Handbook
marrow derived cells in protection against atherosclerosis and coronary heart disease. Mice that do not express SR-BI have increased plasma concentrations of cholesterol associated with enlarged HDL particles with increased unesterified cholesterol content and altered apolipoprotein composition (Rigotti et al., 1997; Trigatti et al., 1999; Miettinen et al., 2001; Braun et al., 2003; Van Eck et al., 2004). Hepatic concentrations of unesterified and esterified cholesterol are unchanged but biliary cholesterol is reduced (Trigatti et al., 1999; Mardones et al., 2001), consistent with studies demonstrating reduced hepatic clearance of HDL cholesterol from blood in mice with reduced hepatic SR-BI expression (Varban et al., 1998; Out et al., 2004; Brundert et al., 2005). In contrast, overexpression of SR-BI in livers of mice results in increased clearance of HDL cholesterol from blood and increased biliary secretion (Kozarsky et al., 1997; Wang et al., 1998; Ueda et al., 1999). Recent studies modeling reverse cholesterol transport from macrophages have demonstrated that hepatic SR-BI plays an important role in driving this process (Zhang et al., 2005b). Elimination of SR-BI expression also results in impaired cholesterol storage in steroidogenic cells in adrenals and ovaries (Rigotti et al., 1997; Trigatti et al., 1999) most likely due to the important role of SR-BI in HDL lipid uptake by steroidogenic cells (Temel et al., 1997). These studies and others demonstrated the important physiological role played by SR-BI in cellular uptake of HDL lipids including cholesterol ester, free cholesterol and bioactive lipids, such as a-tocopherol (vitamin E) (Mardones et al., 2002). Consistent with impaired reverse cholesterol transport (reduced hepatic HDL lipid uptake and secretion into bile), mice that are deficient in SR-BI are more susceptible to atherosclerosis. SR-BI KO mice exhibit increased development of early atherosclerotic plaques (fatty streaks) in the aortic sinus compared to wild type controls when fed atherogenic, high fat/high cholesterol diets (Van Eck et al., 2004; Zhang et al., 2005b). Similarly, mice deficient in both SR-BI and the LDL receptor (LDL-R) (SRBI/LDL-R double KO mice) develop increased levels of high fat diet induced atherosclerosis in the descending aorta compared to LDL-R single KO mice (Covey et al., 2003). SR-BI KO mice that are also deficient in apolipoprotein (apo) E (SR-BI/apoE dKO mice) develop greatly accelerated spontaneous aortic sinus atherosclerosis compared to apoE KO controls (Trigatti et al., 1999). Similar findings were reported in high fat diet fed SR-BI KO mice that contain a mutation (apoER61hypomorphic) that results in the reduced expression of a mutant form of apoE with impaired clearance (Raffai et al., 2001). The level of expression of the mutant form of apoE in these mice is only z5% of normal levels of apoE expression (Raffai et al., 2001; Raffai and Weisgraber, 2002). SR-BI/apoE dKO mice and high fat diet fed SR-BI KO/apoE-R61hypo mice also develop extensive, occlusive coronary artery disease, resulting in myocardial ischemia and infarction and which is associated with increased heart size, reduced heart function and death of mice within weeks of either birth (SR-BI/apoE dKO mice) or initiation of high fat diet feeding (SR-BI KO/apoE-R61hypo mice)
Chapter j 8
The HDL receptor, SR-BI
155
(Braun et al., 2002, 2003; Zhang et al., 2005a; Karackattu et al., 2006). These findings suggest that SR-BI normally plays an important role in protection against the development of atherosclerosis in coronary arteries and that SR-BI KO/apoE mutant mice may represent novel models of human coronary heart disease. The important role of hepatic SR-BI in protection against atherosclerosis is illustrated by the ability of hepatic SR-BI overexpression to reduce atherosclerosis in homozygous and heterozygous LDL-R KO mice (Arai et al., 1999; Kozarsky et al., 2000). Moderate levels of hepatic SR-BI overexpression also protect against diet induced atherosclerosis in apoB transgenic mice (Arai et al., 1999; Kozarsky et al., 2000; Ueda et al., 2000). To test the role of SR-BI in bone marrow derived cells, including macrophages in atherosclerosis, we and others have used bone marrow transplantation to generate chimeric mice with specific disruption of the gene encoding SR-BI in bone marrow derived cells (Covey et al., 2003; Zhang et al., 2003; Van Eck et al., 2004). Replacement of bone marrow stem cells of LDL-R KO mice with those from SR-BI KO mice resulted in increased development of diet induced atherosclerosis compared to mice receiving wild type bone marrow (Covey et al., 2003; Van Eck et al., 2004). Similar results were obtained when wild type mice were used as recipients, although the atherosclerotic plaques were much smaller in size (Covey et al., 2003). Similarly, transplantation of bone marrow stem cells from SR-BI/apoE dKO mice into apoE single KO recipients resulted in increased levels of spontaneous aortic atherosclerosis compared to control mice transplanted with apoE KO mouse bone marrow stem cells (Zhang et al., 2003). These studies indicate that in addition to its role in the liver in driving reverse cholesterol transport, SR-BI plays an important atheroprotective role in bone marrow derived cells, possibly macrophages, in atherosclerotic plaques. Together, these studies suggest that SR-BI plays a key role in multiple tissues/cell types in normal HDL metabolism, and protection against atherosclerosis.
SR-BI MEDIATED HDL LIPID UPTAKE AND EFFLUX SR-BI mediates the selective uptake of cholesterol ester and free cholesterol from HDL in various tissues without the net internalization and degradation of the lipoprotein particle (Acton et al., 1996; Ji et al., 1999; Silver et al., 2001a; Brundert et al., 2005; Shetty et al., 2006; Sun et al., 2006). Selective uptake involves the movement of cholesterol from plasma HDL to various tissues, including the liver, without degradation of the lipoprotein particles (Glass et al., 1983; Stein et al., 1983). SR-BI mediated selective uptake requires lipoproteinreceptor binding and is followed by lipid transfer into the cell (Gu et al., 1998; Williams et al., 2000; Nieland et al., 2002; Temel et al., 2002). Although the mechanisms of lipid transfer remain largely unknown, recent studies suggest that it may occur on the cell surface and does not require internalization of the lipoprotein particles (Liu and Krieger, 2002; Nieland et al., 2005a; Eckhardt
156
The HDL Handbook
et al., 2006; Harder et al., 2007). The internalization of SR-BI may, on the other hand, represent a mechanism by which its activity can be regulated (Eckhardt et al., 2006; Zhang et al., 2007a,b). Overexpression of SR-BI in cultured cells also increases the efflux of cholesterol tracer to HDL or phospholipid vesicles (Ji et al., 1997; Jian et al., 1998). This has led to the idea that SR-BI can mediate the bidirectional flux of free cholesterol between cells and lipoprotein particles; the direction of net flux being determined by the concentration gradient of cholesterol between the cells and HDL particles (Yancey et al., 2003). In contrast to selective lipid uptake, cholesterol efflux appears to involve the prior internalization and re-secretion of the HDL particle (Stangl et al., 1998). The importance of SR-BI-mediated free cholesterol efflux from macrophages to HDL for macrophage reverse cholesterol transport and protection against atherosclerosis, however, is not clear. Macrophages from mice lacking SR-BI do not appear to have significantly reduced capacity to efflux cholesterol to HDL acceptors either when measured in vitro or in a reconstituted efflux assay in mice (Zhang et al., 2003; Van Eck et al., 2004; Wang et al., 2007). On the other hand, macrophages deficient in both SR-BI and apoE do exhibit altered cholesterol homeostasis: they have increased cellular levels of both free cholesterol and cholesteryl ester and accumulate free cholesterol further upon incubation with HDL (Yancey et al., 2007). Thus, SR-BI may be required for cholesterol efflux and proper cholesterol homeostasis in macrophages under certain physiological contexts, such as the absence of apoE.
CELLULAR BIOLOGY OF SR-BI: LOCALIZATION SR-BI is a 509 amino acid and 82 kDa protein made up of short amino- and carboxyterminal cytoplasmic domains, each situated next to a transmembrane domain, with a large, heavily glycosylated extracellular region in between (Acton et al., 1994; Babitt et al., 1997). It is expressed in a variety of cell types involved in metabolism and in the development of atherosclerosis. These include liver hepatocytes and intestinal enterocytes, adipocytes, macrophages, vascular smooth muscle cells, endothelial cells, and steroidogenic cells in the adrenal cortex, testes and ovaries (Acton et al., 1996; Landschulz et al., 1996; Hatzopoulos et al., 1998; Hirano et al., 1999; Yuhanna et al., 2001; Yeh et al., 2002; Nakagawa-Toyama et al., 2005). Immunohistochemistry and immunofluorescence analyses of macrophages from cultured human monocytes or in sections of human atherosclerotic plaques and Kupffer cells in sections of human liver demonstrated that SR-BI is distributed in a heterogeneous punctate pattern (Hirano et al., 1999; Nakagawa-Toyama et al., 2005). Immunofluorescent detection of SR-BI in ldlA7 cells (Chinese hamster ovary (CHO) cells deficient in the LDL-R) stably expressing murine SR-BI (mSR-BI), and in Y1-BS1 murine adrenocortical cells also shows a heterogeneous, punctate as well as a cell peripheral distribution of the receptor (Babitt et al., 1997). Fluorescence microscopy images of CHO cells transfected with enhanced
Chapter j 8
The HDL receptor, SR-BI
157
green fluorescent protein (EGFP) tagged mSR-BI reveal a similar localization pattern (Eckhardt et al., 2004; Nieland et al., 2005b; Zhang et al., 2007a). However, immunofluorescence and immunohistochemical staining of sections of rat and human adrenal glands revealed a cell peripheral distribution of SR-BI (Landschulz et al., 1996; Nakagawa-Toyama et al., 2005). In 3T3-L1 adipocytes, SR-BI immunoreactivity was distributed in a punctate pattern throughout the cytoplasm and scarce cell surface labeling was reported (Tondu et al., 2005). Similarly, immunoelectronmicroscopy analyses of SR-BI in smooth muscle cells of rat aortic sections revealed that, although some cell-surface localization was observed, SR-BI was largely found in the cytoplasm (Yeh et al., 2002). Research conducted on human intestinal Caco 2 cells showed that SR-BI immunoreactivity was localized to the apical and basolateral surfaces of cells (Cai et al., 2001). In addition, immunofluorescence analyses of sections of intestine from mice revealed scarce SR-BI localization in the apical and basolateral membranes of the ileum, distribution of the receptor to both the apical and basolateral membranes in the jejunum, and strong SR-BI immunoreactivity in the apical membrane of the brush border of the duodenum (Cai et al., 2001). SR-BI in transfected CHO-derived ldlA7cells and in human microvascular endothelial cells co-fractionates and/or co-immunoprecipitates with caveolin-1, indicating that the receptor is localized to plasma membrane caveolae (Babitt et al., 1997; Uittenbogaard et al., 2000). Furthermore, in endothelial cells of rat aortic sections, immuno-electron microscopy analyses showed that SR-BI was distributed on the luminal and adluminal surfaces of the cell, but was not localized to the cytoplasm (Yeh et al., 2002). The localization of SR-BI has also been studied in rat, murine and human hepatocytes. Isolation of apical and basolateral membranes from rat livers, followed by immunoblotting for SR-BI revealed that it was localized to both the apical and basolateral surfaces of the cell (Harder et al., 2007). Immunofluorescence and immunohistochemical analyses of mouse liver sections have suggested apical or basolateral distributions in different studies (Sehayek et al., 2003; Yesilaltay et al., 2006). Analyses of primary murine hepatocyte couplets by immunofluorescence microscopy for SR-BI suggested it was localized to the apical membrane and co-localized with transferrin in the endosomal recycling compartment (ERC) (Silver et al., 2001b). In WIF-B cells, a model of polarized hepatocytes, yellow fluorescent protein (YFP)-tagged recombinant human SRBI localized to the apical and the basolateral membranes (Harder et al., 2007). Some of these apparent inconsistencies in the subcellular localization of SR-BI in different tissues and/or different studies may be explained by the dynamics of its localization. Research in both polarized Madin-Darby canine kidney (MDCK) cells and polarized hepatocytes revealed that SR-BI on the basolateral cell surface is continually internalized and recycled (Silver et al., 2001b; Burgos et al., 2004; Harder et al., 2007). Furthermore, recycling back
158
The HDL Handbook
to the basolateral membrane versus trafficking to the apical cell surface appears to be dependent on the cellular cholesterol content and regulated by protein kinase A (Burgos et al., 2004; Harder et al., 2007). Depletion of cellular cholesterol with cyclodextrin (CD) in polarized MDCK cells has been reported to divert the trafficking of internalized basolateral SR-BI from recycling to transcytosis to the apical cell surface via a pathway involving protein kinase A (PKA) (Burgos et al., 2004). On the other hand, cholesterol depletion in WIF-B cells resulted in the redistribution of SR-BI away from the apical (canalicular) membrane to an entirely basolateral localization (Harder et al., 2007). A similar process may take place in non-polarized cells such as adipocytes and transfected CHO-derived cells (Tondu et al., 2005; Zhang et al., 2007a), where SR-BI appears to be continually internalized and recycled back to the cell surface. In these cells, and in hepatocytes, SR-BI’s recycling back to the cell surface and possibly also its internalization appear to be regulated by insulin, angiotensin II (Ang II) and/or HDL via the phosphatidyl inositol 3-kinase (PI3K)/protein kinase B (Akt) signaling pathway (Tondu et al., 2005; Shetty et al., 2006; Zhang et al., 2007a). PI3K plays a major role in endosomal membrane trafficking which involves the recruitment of regulatory proteins to the plasma membrane and the endocytic uptake and recycling of receptors (Backer, 2000; Covera, 2001). The involvement of PI3K in regulating the recruitment of SR-BI to the plasma membrane in adipocytes and hepatocytes was demonstrated using wortmannin, a relatively specific PI3K inhibitor. This abolished Ang II and insulin induced SR-BI recruitment to the cell surface in adipocytes and hepatocytes as well as the HDL-dependent cell surface localization of SR-BI in hepatocytes and CHOderived cells (Tondu et al., 2005; Shetty et al., 2006; Zhang et al., 2007a). These alterations in the levels of cell surface SR-BI in wortmannin-treated cells were associated with reduced SR-BI activity, suggesting a physiological role for this pathway in regulating selective HDL lipid uptake and reverse cholesterol transport (Tondu et al., 2005; Shetty et al., 2006; Yvan-Charvet et al., 2007; Zhang et al., 2007a). Therefore, SR-BI internalization and resecretion appear to present opportunities for regulation of SR-BI activity. SR-BII, a splice variant of SR-BI with identical extracellular regions, N and C-terminal transmembrane segments and N-terminal cytoplasmic region but a different C-terminal cytoplasmic region, exhibits lower lipid uptake activity, related to its lower steady state level on the cell surface (Webb et al., 1998). SR-BII contains a di-Leu endocytic motif in its C-terminal cytoplasmic region and is efficiently endocytosed (Webb et al., 1998; Eckhardt et al., 2006). Transfer of the di-Leu motif from SR-BII’s cytoplasmic tail to SR-BI, which contains no identifiable endocytic motif (Zhang et al., 2007b) converts SR-BI into an endocytic receptor, resulting in a decrease in its selective uptake activity (Eckhardt et al., 2004, 2006). Indeed, unlike mouse SR-BI, disruption of endocytosis in cells expressing human
Chapter j 8
The HDL receptor, SR-BI
159
SR-BI appears to alter its selective uptake activity, possibly by altering the steady state level of SR-BI on the cell surface (Zhang at al., 2007b). Therefore, regulation of the recycling of SR-BI may be an important means of regulating cellular HDL lipid uptake and/or efflux.
SR-BI-MEDIATED ENDOCYTOSIS OF HDL In addition to selective HDL lipid uptake, SR-BI also mediates the endocytic uptake of HDL (Silver et al., 2001b; Burgos et al., 2004; Rhode et al., 2004; Wu¨stner et al., 2004; Nieland et al., 2005a; Pagler et al., 2006; Sun et al., 2006). The endocytosis of HDL is not blocked by inhibitors of SR-BI-mediated selective lipid uptake, such as blockers of lipid transport-1 and 4 (BLT-1 and BLT-4) or glyburide (Nieland et al., 2005a; Pagler et al., 2006; Sun et al., 2006). However, SR-BI mediated internalization of HDL is reduced by general inhibitors of endocytosis, such as hypertonic sucrose, potassium depletion or disruption of the actin cytoskeleton (Silver et al., 2001b; Nieland et al., 2005a; Harder et al., 2006; Zhang et al., 2007a). HDL endocytosis in CHO cells expressing SR-BI is not inhibited by the expression of either a dominant negative dynamin-1 mutant, which inhibits formation of clathrin-coated vesicles, or a dominant negative caveolin-1 mutant that disrupts caveolae-dependent endocytosis (Silver et al., 2001b). These results suggest that SR-BImediated endocytosis of HDL initially bound at the plasma membrane may occur by a pathway that is independent of either clathrin coated pits or caveolae. HDL particles and SR-BI are internalized and accumulate in perinuclear vesicles, that also accumulate transferrin, suggesting they comprise the ERC (Silver et al., 2001b; Rhode et al., 2004; Nieland et al., 2005a; Sun et al., 2006). In addition, fluorescently labeled SR-BI and HDL particles colocalize to the cell surface, and endocytosis of SR-BI occurs along a similar route and with similar velocity as that of HDL (Pagler et al., 2006). HDL particles internalized by endocytosis in an SR-BI dependent manner are resecreted from both polarized and non-polarized cells (Silver et al., 2001b; Rhode et al., 2004; Pagler et al., 2006; Sun et al., 2006). In polarized MDCK cells, cholesteryl ester and free cholesterol components of HDL, internalized via SR-BI from the basolateral membrane, are selectively sorted away from the HDL particle and secreted at the apical membrane, while the remaining HDL particle is re-secreted at the basolateral membrane (Silver et al., 2001). In contrast, in CHO-derived cells and in THP-1 macrophages, there is evidence that HDL is internalized, loaded with cholesterol during recycling, and then re-secreted, and that HDL internalization and re-secretion are required for SR-BI-mediated cholesterol efflux to HDL from cells (Pagler et al., 2006). These data suggest that SR-BI does mediate the internalization of HDL and possibly its shuttling, in both non-polarized and polarized cells.
160
The HDL Handbook
MULTIMERIZATION OF SR-BI SR-BI appears to exist as a monomer in cells and also forms dimers and possibly tetramers (Reaven et al., 2004; Sahoo et al., 2006, 2007). This was shown by co-expressing recombinant SR-BI fused to two different epitope tags in the same cells followed by immunoprecipitation of both tagged versions of SR-BI with an antibody specific to only one or the other epitope tag. To ensure that SR-BI molecules multimerize within the cells themselves, Sahoo and coworkers (2007) used fluorescence resonance energy transfer (FRET) to detect SR-BI dimerization in live cells in vivo. For these experiments, recombinant SR-BI tagged with either yellow or cyan fluorescent proteins (YFP or CFP) on their amino or carboxy termini were co-expressed in Cos-7 cells (Sahoo et al., 2007). Live cell FRET signals were detected only when both constructs encoded SR-BI tagged on the C-termini, indicating that the C-termini of SR-BI monomers were within 10 nm of one another (Sahoo et al., 2007). Since it has previously been shown that mutant SR-BI lacking its C-terminal tail can form oligomers, it has been proposed that oligomerization of the protein is mediated by the carboxy-terminal transmembrane domain or a portion of the extracellular loop in close proximity to this domain (Sahoo et al., 2006, 2007). The importance of SR-BI’s homo-oligomerization for its function is not currently known.
PDZK1, AN ADAPTOR PROTEIN THAT BINDS TO SR-BI’S CARBOXY TERMINAL CYTOPLASMIC TAIL PDZK1 is a 519 amino acid, 63 kDa adapter protein that contains four PDZ (postsynaptic density protein (PSD-95)/Drosophila discs-large (dlg)/tightjunction protein (ZO1)) proteineprotein interaction domains (Kocher et al., 1998; Silver, 2002; Yesilaltay et al., 2005). SR-BI interacts with PDZK1 through the last three amino acids, Ala-Lys-Leu, of its C-terminal tail (Ikemoto et al., 2000; Silver, 2002). Deletion of these three amino acids, or even the C-terminal amino acid, Leu, prevents the interaction between SR-BI and PDZK1 (Silver, 2002). Transgenic mice overexpressing a mutant SR-BI unable to interact with PDZK1, exhibited striking increases in SR-BI mRNA levels but only modest increases in SR-BI protein in the liver and no changes in plasma total cholesterol concentrations (Silver, 2002). PDZK1 KO mice, like SR-BI KO mice, exhibit increased plasma total cholesterol levels associated with HDL particles of increased size (Kocher et al., 2003; Yesilaltay et al., 2006). This is because these mice exhibit a 95% reduction in the level of SR-BI protein in the liver (Yesilaltay et al., 2006). They also exhibit a 50% reduction in SR-BI protein in small intestine but normal levels of SR-BI protein in steroidogenic tissues (Yesilaltay et al., 2006). Overexpression of wild-type SR-BI in livers of PDZK1 KO mice leads to the production of functional cell surface-localized SR-BI, and restores normal steady state HDL cholesterol levels and HDL
Chapter j 8
The HDL receptor, SR-BI
161
composition, despite the PDZK1 deficiency (Yesilaltay et al., 2006). These studies indicate that PDZK1 stabilizes SR-BI protein, protecting it from degradation, but does not affect its function or cell surface localization in liver hepatocytes (Yesilaltay et al., 2006). The degradation of SR-BI in mouse liver hepatocytes is stimulated by fibrates, dependent on PPARa and requires neither lysosomal proteases, calpain nor the proteasome (Mardones et al., 2003; Lan and Silver, 2005). Fibrates also stimulate the turnover of PDZK1 but this does not appear to be the trigger for SRBI turnover (Mardones et al., 2003; Lan and Silver, 2005). Together, these studies suggest that PDZK1 protects SR-BI from degradation in the liver. Regulation of PDZK1 may therefore represent another step at which SR-BI and reverse cholesterol transport can be regulated (Namgaladze and Brune, 2006). PDZK1 is also expressed in endothelial cells. Unlike liver hepatocytes, reduction of PDZK1 levels in endothelial cells does not increase the degradation of SR-BI (Kimura et al., 2006). Instead, SR-BI-mediated HDL induction of intracellular signaling appears to be impaired in the absence of PDZK1 in endothelial cells (Kimura et al., 2006). This will be discussed further below.
REGULATION OF SR-BI SR-BI expression is regulated by a variety of hormonal, metabolic, dietary, and pharmacological stimuli. The hormones estrogen (Stangl et al., 2002; Srivastava, 2003), adrenocorticotrophic hormone (Sun et al., 1999), Ang II (Yvan-Charvet et al., 2007; Yu et al., 2007) and insulin (Sporstol et al., 2007; Yvan-Charvet et al., 2007) have been shown to alter expression of SR-BI transcripts. Dietary cholesterol (Fluiter et al., 1998; Mardones et al., 2001), polyunsaturated fatty acids (Spady et al., 1999), saturated fatty acids (Loison et al., 2002), and vitamins A (Jeyakumar et al., 2006) and E (Witt et al., 2000) were also shown to either up- or downregulate SR-BI transcript and/or protein levels. The regulation of SR-BI gene expression in response to proinflammatory mediators such as lipopolyscaccharides (LPS), tumor necrosis factor (TNF)-a and inferon-g were also reported in various cell types (Buechler et al., 1999; Khovidhunkit et al., 2001). SR-BI gene regulation has been recently reviewed elsewhere (Rigotti et al., 2003; Rhainds and Brissette, 2004).
MOLECULAR MECHANISMS OF HDL-MEDIATED SIGNALING In addition to its role in cholesterol metabolism, HDL initiates signaling events in various cell types relevant to atherogenesis such as endothelial and vascular smooth muscle cells (VSMC), monocytes and platelets (Nofer et al., 2002; Grewal et al., 2003). These signaling events include the activation of phospholipases C and D (Walter et al., 1995), protein kinase C (PKC) (Mendez et al., 1991; Nofer et al., 1998; Grewal et al., 2003), mitogen
162
The HDL Handbook
activated protein kinases (MAPKs) (Deeg et al., 1997; Han et al., 2001) and PI3K/Akt (Li et al., 2002a; Mineo et al., 2003; Cao et al., 2004), the release of intracellular calcium (Porn et al., 1991), the phosphorylation of cellular proteins, the production of cyclic AMP (cAMP), ceramide and nitric oxide (NO) (Li et al., 2002a). HDL signaling has been most extensively studied in endothelial cells (Kimura et al., 2003; Mineo and Shaul, 2003; Mineo et al., 2003; Li et al., 2005). Based on their extensive research in this field, Shaul and his colleagues proposed a model for HDL-induced signaling in endothelial cells (Mineo et al., 2006). They have suggested that cholesterol flux is required for HDL-induced signal initiation and that SR-BI acts as a cholesterol sensor in the plasma membrane. This was based on a number of observations: Cholesterol depletion with CD induced endothelial NO synthase (eNOS) activation in bovine aortic endothelial cells (BAECs) while cholesterol loaded CD did not. An SR-BI neutralizing antibody that prevents SR-BI-mediated cholesterol efflux inhibited signaling by native HDL. Enrichment of HDL with phosphatidylcholine, which increases the efflux rate, resulted in enhanced signaling (Assanasen et al., 2005). It also has been shown that the C-terminal transmembrane and the Cterminal PDZK1 interacting domains of SR-BI are both required for HDL signaling and it has been suggested that perturbations in cholesterol binding to SR-BI results in conformational change(s) that modify its association with PDZK1 or other proteins critical for downstream signaling (Assanasen et al., 2005). The involvement of PDZK1 was demonstrated both by the inability of a mutant form of SR-BI, incapable of interacting with PDZK1, to mediate eNOS activation by HDL, and by knocking down PDZK1 using RNAi, which disrupted HDL-dependent eNOS activation (Assanasen et al., 2005; Kimura et al., 2006), reviewed in Mineo et al. (2006). Although HDL-dependent signaling in some instances has been linked to HDL-associated bioactive lipids, including ceramide (Li et al., 2002b; Nofer et al., 2004), estrogen (Gong et al., 2003), sphingosine-1-phosphate (Kimura et al., 2003, 2006; Nofer et al., 2004; Argraves and Argraves, 2007), and lysophospholipids (Nofer et al., 2004), it is also clear that they are not necessary since signaling can be triggered by reconstituted HDL consisting only of apoA1, phospholipids and cholesterol (Assanasen et al., 2005), reviewed in Mineo et al. (2006). This is consistent with the idea that HDL-dependent signaling may be the result of alterations in membrane cholesterol (Assanasen et al., 2005), reviewed in Mineo et al. (2006).
PROTEIN KINASE C (PKC) PATHWAY The PKC enzyme family regulates a variety of cellular responses, including cell proliferation, differentiation, apoptosis and cell metabolism, reviewed in Dempsey et al. (2000). There are three categories of PKC isoforms that are grouped based on their co-factor requirements. Classical PKCs (a, b and g)
Chapter j 8
The HDL receptor, SR-BI
163
require both diacylgylerol (DAG) and Ca2þ ions for their complete activation; novel PKCs (d, 3, h and q) require only DAG; atypical PKCs (z and i/l) are activated independently of either DAG or Ca2þ (Farhadi et al., 2006). HDL induces the activation of PKC which stimulates cholesterol mobilization in human skin fibroblasts (Walter et al., 1995; Francis et al., 1996). HDL induces the hydrolysis of phosphatidylcholine and phosphatidylinositol bisphosphate by phospholipases C and D generating DAG, phosphatidic acid and inositol phosphates (Walter et al., 1995). Mendenz et al. (1991) showed that the binding of HDL3 to cholesterol loaded fibroblasts stimulated the translocation of PKC to the cell membrane and that downregulation of PKC with sphingosine, a PKC inhibitor, or the long-term treatment with phorbol ester, inhibited HDL-induced cholesterol translocation and efflux. Furthermore, PKC activation, in rat VSMCs, has been shown to be a prerequisite for mobilization of intracellular cholesterol to a pool readily available for apolipoprotein-mediated efflux (Li and Yokoyama, 1995). HDL has been shown to increase intracellular calcium concentrations in CHO cells (Grewal et al., 2005), human skin fibroblasts (Porn et al., 1991; Nofer et al., 2000), smooth muscle cells (Bochkov et al., 1992) and endothelial cells (Honda et al., 1999; Nofer et al., 2004). Elevation of intracellular calcium upon HDL stimulation should result in inducible targeting of PKCa to the plasma membrane (Rentero et al., 2006). This may partially explain the PKCdependent activation of Raf-1/MEK/MAPK upon HDL-SR-BI interaction in annexin A6 overexpressing CHO cells (Rentero et al., 2006). Pilon et al. (2003) reported that SR-BI expression is increased in human NCI-H295R adrenocortical cell line in response to PKC activators, resulting in higher lipoprotein binding and specific cholesteryl ester uptake utilized for steroidogenesis. HDLdependent activation of PKC, mediated by HDL binding to SR-BI, in transfected CHO-derived cells, may also directly modulate the activity of SR-BI. This is based on the finding that a phorbol ester activator of PKC increased, while a PKC inhibitor, Ro31-8220, reduced SR-BI-dependent selective HDL lipid uptake, without affecting either the subcellular distribution or amount of SR-BI protein (Zhang et al., 2007a). Thus, SR-BI may be both a mediator and target of HDL-dependent PKC activation.
MITOGEN ACTIVATED PROTEIN KINASE (MAPK) PATHWAY HDL stimulates the proliferation of several different cell types, including VSMCs (Nofer et al., 2001), adrenocortical cells (Murao et al., 2006), endothelial cells (Nofer et al., 2002) and human lymphocytes (Cuthbert and Lipsky, 1989). HDL’s proliferative effects may be mediated by supply of nutrients such as cholesterol and fatty acids (reviewed in Von Eckardstein and Assmann, 2000) or by induction of specific signaling cascades. The mitogenic effects of HDL in the absence of growth factors such as insulin, fibroblast growth factor or fetal bovine serum were clearly demonstrated in VSMCs (Nofer et al., 2001).
164
The HDL Handbook
In these cells, HDL increased the proportion of cells in S and M phases of the cell cycle and decreased the number of cells in G0/G1 phase; it also induced the phosphorylation of Rb and the sequential expression of cyclins D1, A and E (Nofer et al., 2001). HDL itself is a strong mitogen and is capable of inducing cell proliferation via the activation of Raf-1, MEK1, and ERK1/2 components of the MAPK cascade (Nofer et al., 2001). Additionally, HDL increased the expression of c-fos, the downstream target of ERK1/2 in mitogen induced-cell proliferation. Although PKC can phosphorylate and activate Raf-1 (Kolch et al., 1993), it seems that HDL-dependent activation of MAPK and subsequent cell proliferation of VSMC and fibroblasts occurs in a PKC-independent manner and involves a pertussis toxin (PTX)-sensitive G-protein (Deeg et al., 1997; Nofer et al., 2001; Grewal et al., 2003). The contribution of p21ras (Ras), one of the best characterized activators of the Raf-1/MAPK pathway, to HDLinduced activation of the MAPK pathway was reported in CHO cells (Grewal et al., 2003). It was shown that the binding of HDL to SR-BI activates Ras in a PKC-independent manner with consequent activation of the MAPK cascade (Grewal et al., 2003). Han and colleagues (2002) have shown that HDL stimulates the phosphorylation of p42 and p44 components of the MAPK cascade in RAW264.7 cells and the incubation of J774 macrophages with CD resulted in the phosphorylation of these isoforms suggesting that the removal of cholesterol, not the addition of lipids or lipoproteins, is the key element in HDLmediated signaling. The rapid activation of ERK1/2 and p38 components of the MAPK pathway was also demonstrated in cells overexpressing human SR-BI after incubation with HDL (Baranova et al., 2005; Zhang et al., 2007a). The activation of p38 MAPK has been reported to be involved in the reorganization of the actin cytoskeleton and filamentous actin polymerization (Schafer et al., 1998; Wang and Doerschuk, 2001). This in turn could suggest the involvement of SR-BI in intracellular actin trafficking upon HDL binding (Baranova et al., 2005). Norata et al. (2004) have shown that HDL3 induces the expression of cyclooxygenase-2 and the subsequent production of prostacyclin in endothelial cells in a p38 MAPK/CREB-dependent manner. Finally, increased expression of c-fos, c-myc and erg-1, the ultimate nuclear effectors of the cascade, was detected in smooth muscle cells upon HDL stimulation in the absence of serum (Nofer et al., 2001). Reports identifying the mitogenic components of HDL are conflicting. Nofer et al. (2001) showed that neither apoA-I nor apoA-II alone are capable of inducing DNA synthesis or cell proliferation in VSMCs. Still, others showed that apoA-I and apoC-I components of HDL induce the growth of endothelial, lymphoblastic and adenocarcinoma cells (Favre et al., 1993). In contrast, others reported apoA-I and apoE inhibition of growth factor-induced proliferation and cell cycle progression in VSMCs and lymphocytes (Ishigami et al., 1998). On the other hand, sphingosylphosphorylcholine and lysosulfatide, bioactive lysosphingolipids present in HDL, were shown to exert strong mitogenic activities, activate phosphoinositide-specific phospholipase C and liberate
Chapter j 8
The HDL receptor, SR-BI
165
intracellular calcium (Nofer et al., 2000). These suggest that the lipid components of HDL may account for its mitogenic properties (Guha and Mackman, 2002).
PI3K/AKT PATHWAY PI3K is a ubiquitous lipid kinase that regulates the activity of various signaling cascades (Guha and Mackman, 2002). PI3K is a heterodimeric protein that consists of a regulatory, p85, and a catalytic, p110 subunit, which comprise Class IA isoforms of PI3K (Jones, 2000). The p110a and p110b are widely distributed in mammalian tissues; however, the p110d is usually restricted to leukocytes (Jones, 2000). A single Class IB variant containing the p110g catalytic subunit complexed with a 101-kDa regulatory subunit is thought to interact with G-proteins in leukocytes (Jones, 2000). In contrast, however, Class IA PI3Ks are activated by phosphorylation of tyrosine residues on the Src homology 2 domain of the regulatory subunit (Guha and Mackman, 2002). Phosphorylation stimulates docking of PI3K to the plasma membrane and results in allosteric modifications that increase its catalytic activity (Scheid and Woodgett, 2002). Activated PI3K phosphorylates membrane phosopholipids producing phosphatidylinositol-3,4,5-trisphosphate (PIP3) and phosphatidylinositol-3,4-bisphosphate (PIP2) in the plasma membrane. These membrane lipids stimulate the recruitment of protein kinase B/Akt, phosphatidylinositoldependent kinase 1 (PDK1) and other kinases which contain pleckstrin homology (PH) domains that mediate their binding to PIP2 and PIP3 (Klippel et al., 1997). PDK1 and other kinases phosphorylate Akt’s critical serine residues (Ser473 in Akt1) in their regulatory domains and threonine residues (Thr308 in Akt1) in their catalytic domains (Osaki et al., 2004), activating the Akt kinase (Franke et al., 1997). Like the MAPK pathway, HDL stimulation of PI3K/Akt has been extensively studied in endothelial cells. HDLeSR-BI interaction has been shown to activate the PI3K/Akt pathway and regulates endothelial cell apoptosis, proliferation and migration (Seetharam et al., 2006) (reviewed in Mineo et al., 2006). HDL has been shown to stimulate NO production via the activation of PI3K (Mineo and Shaul, 2003). Both receptor and non-receptor tyrosine kinases (Src) are involved in the activation of PI3K (reviewed in Mineo et al., 2006). Active PI3K induces the independent activation of both Akt and MAPK pathways (Mineo and Shaul, 2003). Akt phosphorylates eNOS at Ser1179, however, the molecular basis of MAPK mediated activation of eNOS is currently unclear (Mineo and Shaul, 2003). Unlike the phosphorylation at Ser1179, the phosphorylation at Thr497 attenuates enzyme activity and it has been reported that the effects of MAPK on eNOS can either be positive or negative (reviewed in Mineo et al., 2006). Evidence of HDL activation of these signaling pathways and the subsequent production of NO came from the use of dominant negative constructs, blocking antibodies and pharmacological antagonists that inhibit the activity of key signaling
166
The HDL Handbook
molecules. Dominant-negative PI3K and Akt inhibited both HDL-mediated phosphorylation and activation of the enzyme (Mineo et al., 2003). In addition, MEK inhibition by PD98059 blocked eNOS activation by HDL without affecting Akt or Ser-1179 phosphorylation (Mineo et al., 2003). These results suggest that HDL stimulates eNOS through a common upstream mediator, Src or PI3K, which leads to a parallel activation of Akt and MAPK and an independent modulation of the enzyme (Mineo et al., 2003). Deletion of the PDZK1 interaction domain on SR-BI or knocking down PDZK1 in endothelial cells prevented HDL-dependent eNOS activation suggesting that PDZK1 is essential for SR-BI-mediated signaling, possibly by acting as a scaffold linking the receptor to the downstream kinases required for the activation of eNOS by HDL (Assanasen et al., 2005; Kimura et al., 2006). Studies that investigated HDL components involved in the activation of eNOS revealed that apoA-I is necessary but not sufficient for eNOS stimulation (Mineo and Shaul, 2003). Evidence came from experiments in which anti-apoA-I antibodies inhibited HDL-induced eNOS activation, however, lipid free apoA-I did not activate eNOS (Yu et al., 2007). Recently, Yu et al. (2007) investigated the effects of Ang II on SR-BI expression in endothelial cells. Ang II decreased SR-BI protein and mRNA expression in human umbilical cord endothelial cells (HUVECs) (Yu et al., 2007). In addition, Ang II inhibited the activity of the human SR-BI promoter in a dose-dependent manner that was sensitive to the PI3K inhibitors, wortmannin and LY294002 (Yu et al., 2007). Therefore, PI3K appears to mediate Ang II-induced inhibition of SR-BI expression. The involvement of PI3K/Akt pathway was further confirmed by examining the phosphorylation state of Akt and by examining the response to Ang II treatment in the context of expression of either a dominant negative (DN) or a constitutively active form of Akt. Ag-II induced a time dependent phosporylation of Akt that was evident within 5 min and peaked at 15 min (Yu et al., 2007). The constitutively active Akt suppressed human SR-BI promoter activity in HUVECs, while the DN-Akt abolished the actions of Ang II (Yu et al., 2007). The involvement of FoxO1, a downstream target of Akt, was also examined. FoxO is a subfamily of forkhead transcription factors that when phosphorylated by Akt are excluded from the nucleus resulting in the inhibition of the transcription of their target genes (Anderson et al., 1998; Brunet et al., 1999). Ang II induced a time dependent phosphorylation of FoxO1. Furthermore, the mutation of the FoxO1 binding site blocked the effect of Ang II on the human SR-BI promoter activity, confirming the involvement of the PI3K/Akt/FoxO1 pathway in Ang II-mediated suppression of SR-BI gene expression (Yu et al., 2007). As discussed earlier, Ang II, insulin and HDL all stimulate the recruitment of SR-BI to the cell surface in a PI3K/ Akt dependent manner in hepatocytes, adipocytes and CHO cells (Tondu et al., 2005; Shetty et al., 2006; Zhang et al., 2007a). If this also occurs in endothelial cells, then SR-BI may be subject to differential short- and long-term regulation. On the other hand, Ang II results in upregulation of SR-BI gene expression in
Chapter j 8
The HDL receptor, SR-BI
167
adipocytes, contrasting with the downregulation in endothelial cells (Yu et al., 2007; Yvan-Charvet et al., 2007). Whether the differential cell type specific regulation of SR-BI gene expression in response to the same hormonal stimuli is also reflected in differential regulation of its subcellular distribution remains to be tested. It has also been shown that HDL signaling through SR-BI promotes endothelial cell migration by a pathway that is independent of eNOS activation (Seetharam et al., 2006). Seetharam et al. (2006) showed that HDL, via SR-BI, causes the rapid activation of Rac GTPase in concert with lamellipodia formation. HDL binding to SR-BI activates Src kinases, PI3K/Akt and p44/p42 MAPK cascades which independently activate Rac GTPase (Seetharam et al., 2006). Moreover, inhibition of PI3K or MAPK prevented HDL-induced Rac activation, however; these modulations did not alter lamellipodia formation suggesting the existence of alternative, Src-independent, pathway(s) that mediate initial lamellipodia formation (Seetharam et al., 2006). This process can also be mediated by lipoprotein particles reconstituted with different ratios of lipid free apoA-I, phospholipids and cholesterol (Seetharam et al., 2006). Cell apoptosis is a key regulator in the development of atherosclerotic lesions. Macrophage apoptosis occurs at all stages of atherosclerosis. In atherosclerotic lesions, macrophage foam cells undergo apoptosis induced by several factors (reviewed in Tobas, 2005). Examples include accumulation of large amounts of intracellular cholesterol and high concentrations of oxidized LDL, oxysterols and TNF-a (reviewed in Tabas, 2005). In early lesions, apoptotic foam cells are efficiently cleared by phagocytic macrophages and this is associated with diminished lesion cellularity and decreased lesion progression (Arai et al., 2005). In late lesions, macrophages also undergo apoptosis but tend to accumulate, suggesting a defect in the phagocytic clearance (Schrijvers et al., 2005). Defective phagocytosis would lead to secondary macrophage necrosis and enhanced inflammatory responses (Hegyi et al., 1996). Thus, the consequence of macrophage apoptosis on lesion development is dependent on the efficiency of apoptotic cell clearance by subendothelial phagocytes which are mostly macrophages. The activation of PI3K/Akt pathway has been shown to protect against cell death in response to a variety of apoptotic stimuli in endothelial cells (Piro et al., 2008), VSMCs (Wang et al., 2007) and in peritoneal macrophages, and macrophages from atherosclerotic lesions from high fat diet fed apoE KO and LDL-R KO mice (Boullier et al., 2006). On the other hand, inhibition of the PI3K/Akt pathway resulted in apoptosis in murine peritoneal macrophages (Koh et al., 1998) and primary human monocyte derived macrophages (Liu et al., 2001). Mechanisms by which Akt inhibition results in apoptosis involve the loss of mitochondrial transmembrane potential, the activation of caspases-9 and -3 and decreased expression of Mcl-1, suggesting the essential role of PI3K/Akt in macrophage survival (Liu et al., 2001). Moreover, activation of Scr/PI3K/Akt and Ras/Raf/MEK/ERK MAPK signaling cascades has been
168
The HDL Handbook
reported in granulocyte-macrophage colony stimulating factor (GM-CSF)induced microglial proliferation (Suh et al., 2005). More recently, Namgaladze and Brune (2006) have shown that phospholipase-A2 modified LDL promotes human and THP-1 monocytic cell survival via the activation of PI3K/Akt pathway. This suggests that HDL, via SR-BI, may activate Akt in macrophages as it does in endothelial cells, possibly involving a similar mechanism, triggering an anti-apoptotic response, potentially reducing the formation of a necrotic core. This has yet to be tested.
CONCLUSIONS The discovery of SR-BI by Acton et al. (1996) and further characterization of cellular biology of the receptor provided new insights into HDL metabolism and cellular responses to HDL. Studies from genetically engineered mice shed light onto the role of SR-BI in reverse cholesterol transport, in the protection against atherosclerosis and coronary heart disease. In addition, the more recently discovered interaction of SR-BI with the adaptor protein, PDZK1, (Ikemoto et al., 2000) clarified a complex picture of the cellular mechanisms regulating HDL metabolism and SR-BI induced signaling in various cell types. Studies on the molecular mechanisms of HDL/SR-BI induced signaling revealed the involvement of numerous pathways that mediate the atheroprotective actions of HDL in cells relevant to atherogenesis such as macrophages, endothelial cells and vascular smooth muscle cells.
ACKNOWLEDGMENTS Research in our laboratory is supported by grants NA6310 from the Heart and Stroke Foundation of Ontario and MOP74753 from the Canadian Institutions of Health Research, Institute of Genetics and a Program Grant (PRG 6502) from the Heart and Stroke Foundation of Ontario. BT has been a New Investigator of the Heart and Stroke Foundation of Canada. AA is supported by a graduate scholarship from Kuwait University. RB was supported by a graduate scholarship from the Heart and Stroke Foundation of Ontario.
REFERENCES Acton, S., Rigotti, A., Landschulz, K. T., Xu, S., Hobbs, H. H., & Krieger, M. (1996). Identification of scavenger receptor SR-BI as a high density lipoprotein receptor. Science, 271, 518e520. Acton, S. L., Scherer, P. E., Lodish, H. F., & Krieger, M. (1994). Expression cloning of SR-BI, a CD36-related class B scavenger receptor. J Biol Chem, 269, 21003e21009. Anderson, M., Viars, C. S., Czekay, S., Cavenee, W. K., & Arden, K. C. (1998). Cloning and characterization of three human forkhead genes that comprise an FKHR-like gene subfamily. Genomics, 47, 187e199. Arai, T., Wang, N., Bezouevski, M., Welch, C., & Tall, A. R. (1999). Decreased atherosclerosis in heterozygous low density lipoprotein receptor-deficient mice expressing the scavenger receptor BI transgene. J Biol Chem, 274, 2366e2371.
Chapter j 8
The HDL receptor, SR-BI
169
Arai, S., Shelton, J. M., Chen, M., et al. (2005). A role for the apoptosis inhibitory factor AIM/ Spalpha/Api6 in atherosclerosis development. Cell Metab, 1, 201e213. Argraves, K. M., & Argraves, W. S. (2007). HDL serves as an S1P signaling platform mediating a multitude of cardiovascular effects. J Lipid Res, 48, 2325e2333. Assanasen, C., Mineo, C., Seetharam, D., et al. (2005). Cholesterol binding, efflux, and a PDZinteracting domain of scavenger receptor-BI mediate HDL-initiated signaling. J Clin Invest, 115, 969e977. Babitt, J., Trigatti, B., Rigotti, A., et al. (1997). Murine SR-BI, a high density lipoprotein receptor that mediates selective lipid uptake, is N-glycosylated and fatty acylated and colocalizes with plasma membrane caveolae. J Biol Chem, 272, 13242e13249. Backer, J. (2000). Phosphoinositide 3-kinases and the regulation of vesicular trafficking. Mol Cell Biol Res Commun, 3, 193e204. Baranova, I., Vishnyakova, T. G., Bocharov, A. V., et al. (2005). Serum amyloid A binding to CLA-1 (CD36 and LIMPII analogous-1) mediates serum amyloid A protein-induced activation of ERK1/2 and p38 mitogen-activated protein kinases. J Biol Chem, 280, 8031e8040. Bochkov, V., Tkachuk, V., Buhler, F., & Resink, T. (1992). Phosphoinositide and calcium signalling responses in smooth muscle cells: comparison between lipoproteins, Ang II, and PDGF. Biochem Biophys Res Commun, 188, 1295e1304. Boullier, A., Li, Y., Quehenberger, O., et al. (2006). Minimally oxidized LDL offsets the apoptotic effects of extensively oxidized LDL and free cholesterol in macrophages. Arterioscler Thromb Vasc Biol, 26, 1169e1176. Braun, A., Trigatti, B. L., Post, M. J., et al. (2002). Loss of SR-BI expression leads to the early onset of occlusive atherosclerotic coronary artery disease, spontaneous myocardial infarctions, severe cardiac dysfunction, and premature death in apolipoprotein E-deficient mice. Circ Res, 90, 270e276. Braun, A., Zhang, S., Miettinen, H. E., et al. (2003). Probucol prevents early coronary heart disease and death in the high-density lipoprotein receptor SR-BI/apolipoprotein E double knockout mouse. Proc Natl Acad Sci USA, 100, 7283e7288. Brundert, M., Ewert, A., Heeren, J., et al. (2005). Scavenger receptor class B type I mediates the selective uptake of high-density lipoprotein-associated cholesteryl ester by the liver in mice. Arterioscler Thromb Vasc Biol., 25, 143e148. Brunet, A., Bonni, A., Zigmond, M. J., et al. (1999). Akt promotes cell survival by phosphorylating and inhibiting a Forkhead transcription factor. Cell, 96, 857e868. Buechler, C., Ritter, M., Quoc, C. D., Agildere, A., & Schmitz, G. (1999). Lipopolysaccharide inhibits the expression of the scavenger receptor Cla-1 in human monocytes and macrophages. Biochem Biophys Res Commun, 262, 251e254. Burgos, P., Klattenhoff, C., de la Fuente, E., Rigotti, A., & Gonza´lez, A. (2004). Cholesterol depletion induces PKA-mediated basolateral-to-apical transcytosis of the scavenger receptor class B type I in MDCK cells. Proc Natl Acad Sci USA, 101, 3845e3850. Cai, S., Kirby, R. J., Howles, P. N., & Hui, D. Y. (2001). Differentiation-dependent expression and localization of the class B type I scavenger receptor in intestine. J Lipid Res, 42, 902e909. Cao, W., Murao, K., Imachi, H., et al. (2004). A mutant high-density lipoprotein receptor inhibits proliferation of human breast cancer cells. Cancer Res, 64, 1515e1521. Covera, S. (2001). Phosphatidylinositol 3-kinase and the control of endosome dynamics: new players defined by structural motifs. Traffic, 2, 859e866. Covey, S., Krieger, M., Wang, W., Penman, M., & Trigatti, B. L. (2003). Scavenger receptor class B type I-mediated protection against atherosclerosis in LDL receptor-negative mice involves its expression in bone marrow-derived cells. Arterioscler Thromb Vasc Biol., 23, 1589e1594.
170
The HDL Handbook
Cuthbert, J., & Lipsky, P. E. (1989). Lipoproteins may provide fatty acids necessary for human lymphocyte proliferation by both low density lipoprotein receptor-dependent and -independent mechanisms. J Biol Chem, 264, 13468e13474. Deeg, M., Bowen, R. F., Oram, J. F., & Bierman, E. L. (1997). High density lipoproteins stimulate mitogen-activated protein kinases in human skin fibroblasts. Arterioscler Thromb Vasc Biol, 17, 1667e1674. Dempsey, E., Newton, A. C., Mochly-Rosen, D., et al. (2000). Protein kinase C isozymes and the regulation of diverse cell responses. Am J Physiol Lung Cell Mol Physiol, 279, L429e438. Eckhardt, E., Cai, L., Shetty, S., Zhao, Z., Szanto, A., Webb, N. R., & Van der Westhuyzen, D. R. (2006). High density lipoprotein endocytosis by scavenger receptor SR-BII is clathrindependent and requires a carboxyl-terminal dileucine motif. J Biol Chem, 281, 4348e4353. Eckhardt, E., Cai, L., Sun, B., Webb, N. R., & van der Westhuyzen, D. R. (2004). High density lipoprotein uptake by scavenger receptor SR-BII. J Biol Chem, 279, 14372e14381. Farhadi, A., Keshavarzian, A., Ranjbaran, Z., Fields, J. Z., & Banan, A. (2006). The role of protein kinase C isoforms in modulating injury and repair of the intestinal barrier. J Pharmacol Exp Ther, 316, 1e7. Favre, G., Tazi, K. A., Le Gaillard, F., Bennis, F., Hachem, H., & Soula, G. (1993). High density lipoprotein3 binding sites are related to DNA biosynthesis in the adenocarcinoma cell line A549. J Lipid Res, 34, 1093e1106. Fluiter, K., van der Westhuijzen, D. R., & van Berkel, T. J. (1998). In vivo regulation of scavenger receptor BI and the selective uptake of high density lipoprotein cholesteryl esters in rat liver parenchymal and Kupffer cells. J Biol Chem, 273, 8434e8438. Francis, G., Oram, J. F., Heinecke, J. W., & Bierman, E. L. (1996). Oxidative tyrosylation of HDL enhances the depletion of cellular cholesteryl esters by a mechanism independent of passive sterol desorption. Biochemistry, 35, 15188e15197. Franke, T., Kaplan, D. R., Cantley, L. C., & Toker, A. (1997). Direct regulation of the Akt protooncogene product by phosphatidylinositol-3,4-bisphosphate. Science, 275, 665e668. Glass, C., Pittman, R. C., Weinstein, D. B., & Steinberg, D. (1983). Dissociation of tissue uptake of cholesterol ester from that of apoprotein A-I of rat plasma high density lipoprotein: selective delivery of cholesterol ester to liver, adrenal, and gonad. Proc Natl Acad Sci USA, 80, 5435e5439. Gong, M., Wilson, M., Kelly, T., et al. (2003). HDL-associated estradiol stimulates endothelial NO synthase and vasodilation in an SR-BI-dependent manner. J Clin Invest, 111, 1579e1587. Grewal, T., de Diego, I., Kirchhoff, M. F., et al. (2003). High density lipoprotein-induced signaling of the MAPK pathway involves scavenger receptor type BI-mediated activation of Ras. J Biol Chem, 278, 16478e16481. Grewal, T., Evans, R., Rentero, C., et al. (2005). Annexin A6 stimulates the membrane recruitment of p120GAP to modulate Ras and Raf-1 activity. Oncogene, 24, 5809e5820. Gu, X., Trigatti, B., Xu, S., Acton, S., Babitt, J., & Krieger, M. (1998). The efficient cellular uptake of high density lipoprotein lipids via scavenger receptor class B type I requires not only receptor-mediated surface binding but also receptor-specific lipid transfer mediated by its extracellular domain. J Biol Chem, 273, 26338e26348. Guha, M., & Mackman, N. (2002). The phosphatidylinositol 3-kinase-Akt pathway limits lipopolysaccharide activation of signaling pathways and expression of inflammatory mediators in human monocytic cells. J Biol Chem, 277, 32124e32132. Han, J., Hajjar, D. P., Zhou, X., Gotto, A. M., Jr., & Nicholson, A. C. (2002). Regulation of peroxisome proliferator-activated receptor-gamma-mediated gene expression. A new mechanism of action for high density lipoprotein. J Biol Chem, 277, 23582e23586.
Chapter j 8
The HDL receptor, SR-BI
171
Han, J., Nicholson, A. C., Zhou, X., Feng, J., Gotto., A. M., Jr., & Hajjar, D. P. (2001). Oxidized low density lipoprotein decreases macrophage expression of scavenger receptor B-I. J Biol Chem, 276, 16567e16572. Harder, C., Vassiliou, G., McBride, H. M., & McPherson, R. (2006). Hepatic SR-BI-mediated cholesteryl ester selective uptake occurs with unaltered efficiency in the absence of cellular energy. J Lipid Res, 47, 492e503. Harder, C. J., Meng, A., Rippstein, P., McBride, H. M., & McPherson, R. (2007). SR-BI undergoes cholesterol-stimulated transcytosis to the bile canaliculus in polarized WIF-B cells. J Biol Chem, 282, 1445e1455. Hatzopoulos, A. K., Rigotti, A., Rosenberg, R. D., & Krieger, M. (1998). Temporal and spatial pattern of expression of the HDL receptor SR-BI during murine embryogenesis. J Lipid Res, 39, 495e508. Hegyi, L., Skepper, J. N., Cary, N. R., & Mitchinson, M. J. (1996). Foam cell apoptosis and the development of the lipid core of human atherosclerosis. J Pathol, 180, 423e429. Hirano, K., Yamashita, S., Nakagawa, Y., et al. (1999). Expression of human scavenger receptor class B type I in cultured human monocyte-derived macrophages and atherosclerotic lesions. Circ Res, 85, 108e116. Honda, H., Wakamatsu, B., Goldhaber, J., Berliner, J., Navab, M., & Weiss, J. (1999). Highdensity lipoprotein increases intracellular calcium levels by releasing calcium from internal stores in human endothelial cells. Atherosclerosis, 143, 299e306. Ikemoto, M., Arai, H., Feng, D., et al. (2000). Identification of a PDZ-domain-containing protein that interacts with the scavenger receptor class B type I. Proc Natl Acad Sci USA, 97, 6538e6543. Ishigami, M., Swertfeger, D. K., Granholm, N. A., & Hui, D. Y. (1998). Apolipoprotein E inhibits platelet-derived growth factor-induced vascular smooth muscle cell migration and proliferation by suppressing signal transduction and preventing cell entry to G1 phase. J Biol Chem, 273, 20156e20161. Jeyakumar, S., Vajreswari, A., & Giridharan, N. V. (2006). Impact of vitamin A on high-density lipoprotein-cholesterol and scavenger receptor class BI in the obese rat. Obesity, 15, 322e329. Ji, Y., Jian, B., Wang, N., et al. (1997). Scavenger receptor BI promotes high density lipoproteinmediated cellular cholesterol efflux. J Biol Chem, 272, 20982e20985. Ji, Y., Wang, N., Ramakrishnan, R., et al. (1999). Hepatic scavenger receptor BI promotes rapid clearance of high density lipoprotein free cholesterol and its transport into bile. J Biol Chem, 274, 33398e33402. Jian, B., de la Llera-Moya, M., Ji, Y., et al. (1998). Scavenger receptor class B type I as a mediator of cellular cholesterol efflux to lipoproteins and phospholipid acceptors. J Biol Chem, 273, 5599e5606. Jones, G. (2000). Cellular signaling in macrophage migration and chemotaxis. J Leukoc Biol, 68, 593e602. Karackattu, S. L., Trigatti, B., & Krieger, M. (2006). Hepatic lipase deficiency delays atherosclerosis, myocardial infarction, and cardiac dysfunction and extends lifespan in SR-BI/ apolipoprotein E double knockout mice. Arterioscler Thromb Vasc Biol, 26, 548e554. Khovidhunkit, W., Moser, A. H., Shigenaga, J. K., Grunfeld, C., & Feingold, K. R. (2001). Regulation of scavenger receptor class B type I in hamster liver and Hep3B cells by endotoxin and cytokines. J Lipid Res, 42, 1636e1644. Kimura, T., Sato, K., Malchinkhuu, E., et al. (2003). High-density lipoprotein stimulates endothelial cell migration and survival through sphingosine 1-phosphate and its receptors. Arterioscler Thromb Vasc Biol, 23, 1283e1288.
172
The HDL Handbook
Kimura, T., Tomura, H., Mogi, C., et al. (2006). Role of scavenger receptor class B type I and sphingosine 1-phosphate receptors in high density lipoprotein-induced inhibition of adhesion molecule expression in endothelial cells. J. Biol. Chem, 281, 37457e37467. Klippel, A., Kavanaugh, W. M., Pot, D., & Williams, L. T. (1997). A specific product of phosphatidylinositol 3-kinase directly activates the protein kinase Akt through its pleckstrin homology domain. Mol Cell Biol, 17, 338e344. Kocher, O., Comella, N., Tognazzi, K., & Brown, L. F. (1998). Identification and partial characterization of PDZK1: a novel protein containing PDZ interaction domains. Lab Invest, 78, 117e125. Kocher, O., Pal, R., Roberts, M., Cirovic, C., & Gilchrist, A. (2003). Targeted disruption of the PDZK1 gene by homologous recombination. Mol Cell Biol, 23, 1175e1180. Koh, J. S., Lieberthal, W., Heydrick, S., & Levine, J. S. (1998). Lysophosphatidic acid is a major serum noncytokine survival factor for murine macrophages which acts via the phosphatidyl inositol 3-kinase signaling pathway. J Clin Invest, 102, 716e727. Kolch, W., Heidecker, G., Kochs, G., et al. (1993). Protein kinase C alpha activates RAF-1 by direct phosphorylation. Nature, 364, 249e252. Kozarsky, K. F., Donahee, M. H., Glick, J. M., Krieger, M., & Rader, D. J. (2000). Gene transfer and hepatic overexpression of the HDL receptor SR-BI reduces atherosclerosis in the cholesterol-fed LDL receptor-deficient mouse. Arterioscler Thromb Vasc Biol, 20, 721e727. Kozarsky, K. F., Donahee, M. H., Rigotti, A., Iqbal, S. N., Edelman, E. R., & Krieger, M. (1997). Overexpression of the HDL receptor SR-BI alters plasma HDL and bile cholesterol levels. Nature, 387, 414e417. Lan, D., & Silver, D. L. (2005). Fenofibrate induces a novel degradation pathway for scavenger receptor B-I independent of PDZK1. J Biol Chem, 280, 23390e23396. Landschulz, K. T., Pathak, R. K., Rigotti, A., Krieger, M., & Hobbs, H. H. (1996). Regulation of scavenger receptor, class B, type I, a high density lipoprotein receptor, in liver and steroidogenic tissues of the rat. J Clin Invest, 98, 984e995. Li, Q., & Yokoyama, S. (1995). Independent regulation of cholesterol incorporation into free apolipoprotein-mediated cellular lipid efflux in rat vascular smooth muscle cells. J Biol Chem, 270, 26216e26223. Li, X., Titlow, W. B., Jackson, B. A., et al. (2002a). High density lipoprotein binding to scavenger receptor, Class B, type I activates endothelial nitric-oxide synthase in a ceramide-dependent manner. J Biol Chem, 277, 11058e11063. Li, X. A., Titlow, W. B., Jackson, B. A., et al. (2002b). High density lipoprotein binding to scavenger receptor, Class B, type I activates endothelial nitric-oxide synthase in a ceramidedependent manner. J Biol Chem, 277, 11058e11063. Li, X., Guo, L., Dressman, J. L., Asmis, R., & Smart, E. J. (2005). A novel ligand-independent apoptotic pathway induced by scavenger receptor class B, type I and suppressed by endothelial nitric-oxide synthase and high density lipoprotein. J Biol Chem, 280, 19087e19096. Liu, B., & Krieger, M. (2002). Highly purified scavenger receptor class B, type I reconstituted into phosphatidylcholine/cholesterol liposomes mediates high affinity high density lipoprotein binding and selective lipid uptake. J Biol Chem, 277, 34125e34135. Liu, H., Perlman, H., Pagliari, L. J., & Pope, R. M. (2001). Constitutively activated Akt-1 is vital for the survival of human monocyte-differentiated macrophages. Role of Mcl-1, independent of nuclear factor (NF)-kappaB, Bad, or caspase activation. J Exp Med, 194, 113e126. Loison, C., Mendy, F., Serougne, C., & Lutton, C. (2002). Dietary myristic acid modifies the HDLcholesterol concentration and liver scavenger receptor BI expression in the hamster. Br J Nutr, 87, 199e210.
Chapter j 8
The HDL receptor, SR-BI
173
Mardones, P., Pilon, A., Bouly, M., et al. (2003). Fibrates down-regulate hepatic scavenger receptor class B type I protein expression in mice. J. Biol. Chem, 278, 7884e7890. Mardones, P., Quinones, V., Amigo, L., et al. (2001). Hepatic cholesterol and bile acid metabolism and intestinal cholesterol absorption in scavenger receptor class B type I-deficient mice. J Lipid Res, 42, 170e180. Mardones, P., Strobel, P., Miranda, S., et al. (2002). Alpha-tocopherol metabolism is abnormal in scavenger receptor class B type I (SR-BI)-deficient mice. J Nutr, 132, 443e449. Mendez, A., Oram, J. F., & Bierman, E. L. (1991). Protein kinase C as a mediator of high density lipoprotein receptor-dependent efflux of intracellular cholesterol. J Biol Chem, 266, 10104e10111. Miettinen, H. E., Rayburn, H., & Krieger, M. (2001). Abnormal lipoprotein metabolism and reversible female infertility in HDL receptor (SR-BI)-deficient mice. J Clin Invest, 108, 1717e1722. Mineo, C., Deguchi, H., Griffin, J. H., & Shaul, P. W. (2006). Endothelial and antithrombotic actions of HDL. Circ Res, 98, 1352e1364. Mineo, C., & Shaul, P. W. (2003). HDL stimulation of endothelial nitric oxide synthase: a novel mechanism of HDL action. Trends Cardiovasc Med, 13, 226e231. Mineo, C., Yuhanna, I. S., Quon, M. J., & Shaul, P. W. (2003). High density lipoprotein-induced endothelial nitric-oxide synthase activation is mediated by Akt and MAP kinases. J Biol Chem, 278, 9142e9149. Murao, K., Imachi, H., Cao, W., et al. (2006). High-density lipoprotein is a potential growth factor for adrenocortical cells. Biochem Biophys Res Commun, 344, 226e232. Nakagawa-Toyama, Y., Hirano, K., et al. (2005). Human scavenger receptor class B type I is expressed with cell-specific fashion in both initial and terminal site of reverse cholesterol transport. Atherosclerosis, 183, 75e83. Namgaladze, D., & Brune, B. (2006). Phospholipase A2-modified low-density lipoprotein activates the phosphatidylinositol 3-kinase-Akt pathway and increases cell survival in monocytic cells. Arterioscler Thromb Vasc Biol, 26, 2510e2516. Nieland, T., Ehrlich, M., Krieger, M., & Kirchhausen, T. (2005a). Endocytosis is not required for the selective lipid uptake mediated by murine SR-BI. Biochim Biophys Acta, 1734, 44e51. Nieland, T. J., Ehrlich, M., Krieger, M., & Kirchhausen, T. (2005b). Endocytosis is not required for the selective lipid uptake mediated by murine SR-BI. Biochim Biophys Acta, 1734, 44e51. Nieland, T. J., Penman, M., Dori, L., Krieger, M., & Kirchhausen, T. (2002). Discovery of chemical inhibitors of the selective transfer of lipids mediated by the HDL receptor SR-BI. Proc Natl Acad Sci USA, 99, 15422e15427. Nofer, J., Fobker, M., Hobbel, G., et al. (2000). Activation of phosphatidylinositol-specific phospholipase C by HDL-associated lysosphingolipid. Involvement in mitogenesis but not in cholesterol efflux. Biochemistry, 39, 15199e151207. Nofer, J., Kehrel, B., Fobker, M., Levkau, B., Assmann, G., & von Eckardstein, A. (2002). HDL and arteriosclerosis: beyond reverse cholesterol transport. Atherosclerosis, 161, 1e16. Nofer, J., Walter, M., Kehrel, B., et al. (1998). HDL3-mediated inhibition of thrombin-induced platelet aggregation and fibrinogen binding occurs via decreased production of phosphoinositide-derived second messengers 1,2-diacylglycerol and inositol 1,4,5-tris-phosphate. Arterioscler Thromb Vasc Biol, 18, 816e819. Nofer, J. R., Junker, R., Pulawski, E., et al. (2001). High density lipoproteins induce cell cycle entry in vascular smooth muscle cells via mitogen activated protein kinase-dependent pathway. Thromb Haemost, 85, 730e735.
174
The HDL Handbook
Nofer, J. R., van der Giet, M., Tolle, M., et al. (2004). HDL induces NO-dependent vasorelaxation via the lysophospholipid receptor S1P3. J Clin Invest, 113, 569e581. Norata, G., Banfi, C., Pirillo, A., et al. (2004). Oxidised-HDL3 induces the expression of PAI-1 in human endothelial cells. Role of p38MAPK activation and mRNA stabilization. Br J Haematol, 127, 97e104. Osaki, M., Oshimura, M., & Ito, H. (2004). PI3K-Akt pathway: its functions and alterations in human cancer. Apoptosis, 9, 667e676. Out, R., Hoekstra, M., Spijkers, J. A., et al. (2004). Scavenger receptor class B type I is solely responsible for the selective uptake of cholesteryl esters from HDL by the liver and the adrenals in mice. J Lipid Res, 45, 2088e2095. Pagler, T., Rhode, S., & Neuhofer, A. (2006). SR-BI-mediated high density lipoprotein (HDL) endocytosis leads to HDL resecretion facilitating cholesterol efflux. J Biol Chem, 281, 1193e1204. Pilon, A., Martin, G., Bultel-Brienne, S., et al. (2003). Regulation of the scavenger receptor BI and the LDL receptor by activators of aldosterone production, angiotensin II and PMA, in the human NCI-H295R adrenocortical cell line. Biochim Biophys Acta, 8, 3. Piro, S., Spampinato, D., Spadaro, L., Oliveri, C. E., Purrello, F., & Rabuazzo, A. M. (2008). Direct apoptotic effects of free fatty acids on human endothelial cells. Nutr Metab Cardiovasc Dis, 18, 96e104. Porn, M., Akerman, K. E., & Slotte, J. P. (1991). High-density lipoproteins induce a rapid and transient release of Ca2þ in cultured fibroblasts. Biochem J, 279, 29e33. Raffai, R. L., & Weisgraber, K. H. (2002). Hypomorphic apolipoprotein E mice: a new model of conditional gene repair to examine apolipoprotein E-mediated metabolism. J Biol Chem, 277, 11064e11068. Raffai, R. L., Dong, L. M., Farese, R. V., Jr., & Weisgraber, K. H. (2001). Introduction of human apolipoprotein E4 “domain interaction” into mouse apolipoprotein E. Proc Natl Acad Sci USA, 98, 11587e11591. Reaven, E., Cortez, Y., Leers-Sucheta, S., Nomoto, A., & Azhar, S. (2004). Dimerization of the scavenger receptor class B type I: formation, function, and localization in diverse cells and tissues. J Lipid Res, 45, 513e528. Rentero, C., Evans, R., Wood, P., et al. (2006). Inhibition of H-Ras and MAPK is compensated by PKC-dependent pathways in annexin A6 expressing cells. Cell Signal, 18, 1006e1016. Rhainds, D., & Brissette, L. (2004). The role of scavenger receptor class B type I (SR-BI) in lipid trafficking, defining the rules for lipid traders. Int J Biochem Cell Biol, 36, 39e77. Rhode, S., Breuer, A., Hesse, J., et al. (2004). Visualization of the uptake of individual HDL particles in living cells via the scavenger receptor class B type I. Cell Biochem Biophys, 41, 343e356. Rigotti, A., Miettinen, H. E., & Krieger, M. (2003). The role of the high-density lipoprotein receptor SR-BI in the lipid metabolism of endocrine and other tissues. Endocr Rev, 24, 357e387. Rigotti, A., Trigatti, B. L., Penman, M., Rayburn, H., Herz, J., & Krieger, M. (1997). A targeted mutation in the murine gene encoding the high density lipoprotein (HDL) receptor scavenger receptor class B type I reveals its key role in HDL metabolism. Proc Natl Acad Sci USA, 94, 12610e12615. Sahoo, D., Darlington, Y. F., Pop, D., Williams, D. L., & Connelly, M. A. (2006). Scavenger receptor class B type I (SR-BI) assembles into detergent-sensitive dimers and tetramers. Biochim Biophys Acta, 1771, 807e817.
Chapter j 8
The HDL receptor, SR-BI
175
Sahoo, D., Peng, Y., Smith, J. R., Darlington, Y. F., & Connelly, M. A. (2007). Scavenger receptor class B, type I (SR-BI) homo-dimerizes via its C-terminal region: fluorescence resonance energy transfer analysis. Biochim Biophys Acta, 1771, 818e829. Schafer, C., Ross, S. E., Bragado, M. J., Groblewski, G. E., Ernst, S. A., & Williams, J. A. (1998). A role for the p38 mitogen-activated protein kinase/Hsp 27 pathway in cholecystokinininduced changes in the actin cytoskeleton in rat pancreatic acini. J Biol Chem, 273, 24173e24180. Scheid, M., & Woodgett, J. R. (2002). Protein kinases: six degrees of separation? Curr Biol, 10, R191e194. Schrijvers, D., De Meyer, G. R., Kockx, M. M., Herman, A. G., & Martinet, W. (2005). Phagocytosis of apoptotic cells by macrophages is impaired in atherosclerosis. Arterioscler Thromb Vasc Biol, 25, 1256e1261. Seetharam, D., Mineo, C., Gormley, A. K., et al. (2006). High-density lipoprotein promotes endothelial cell migration and reendothelialization via scavenger receptor-B type I. Circ Res, 98, 63e72. Sehayek, E., Wang, R., Ono, J. G., et al. (2003). Localization of the PE methylation pathway and SR-BI to the canalicular membrane: evidence for apical PC biosynthesis that may promote biliary excretion of phospholipid and cholesterol. J Lipid Res, 44, 1605e1613. Shetty, S., Eckhardt, E. R., Post, S. R., & van der Westhuyzen, D. R. (2006). Phosphatidylinositol3-kinase regulates scavenger receptor class B type I subcellular localization and selective lipid uptake in hepatocytes. Arterioscler Thromb Vasc Biol., 26, 2125e2131. Silver, D. L., Wang, N., Xiao, X., & Tall, A. R. (2001a). High density lipoprotein (HDL) particle uptake mediated by scavenger receptor class B type 1 results in selective sorting of HDL cholesterol from protein and polarized cholesterol secretion. J Biol Chem, 276, 25287e25293. Silver, D., Wang, N., Xiao, X., & Tall, A. R. (2001b). High density lipoprotein (HDL) particle uptake mediated by scavenger receptor class B type 1 results in selective sorting of HDL cholesterol from protein and polarized cholesterol secretion. J Biol Chem, 276, 25287e25293. Silver, D. L. (2002). A carboxyl-terminal PDZ-interacting domain of scavenger receptor B, type I is essential for cell surface expression in liver. J Biol Chem, 277, 34042e34047. Spady, D., Kearney, D. M., & Hobbs, H. H. (1999). Polyunsaturated fatty acids up-regulate hepatic scavenger receptor B1 (SR-BI) expression and HDL cholesteryl ester uptake in the hamster. J Lipid Res, 40, 1384e1394. Sporstol, M., Mousavi, S. A., Eskild, W., Roos, N., & Berg, T. (2007). ABCA1, ABCG1 and SRBI: hormonal regulation in primary rat hepatocytes and human cell lines. BMC Mol Biol, 8, 5. Srivastava, R. A. (2003). Scavenger receptor class B type I expression in murine brain and regulation by estrogen and dietary cholesterol. J Neurol Sci, 210, 11e18. Stangl, H., Cao, G., Wyne, K. L., & Hobbs, H. H. (1998). Scavenger receptor, class B, type I-dependent stimulation of cholesterol esterification by high density lipoproteins, low density lipoproteins, and nonlipoprotein cholesterol. J Biol Chem, 273, 31002e31008. Stangl, H., Graf, G. A., Yu, L., Cao, G., & Wyne, K. (2002). Effect of estrogen on scavenger receptor BI expression in the rat. J Endocrinol, 175, 663e672. Stein, Y., Dabach, Y., Hollander, G., Halperin, G., & Stein, O. (1983). Metabolism of HDLcholesteryl ester in the rat, studied with a nonhydrolyzable analog, cholesteryl linoleyl ether. Biochim Biophys Acta, 752, 98e105. Suh, H., Kim, M. O., & Lee, S. C. (2005). Inhibition of granulocyte-macrophage colony-stimulating factor signaling and microglial proliferation by anti-CD45RO: role of Hck tyrosine kinase and phosphatidylinositol 3-kinase/Akt. J Immunol, 174, 2712e2719.
176
The HDL Handbook
Sun, Y., Wang, N., & Tall, A. R. (1999). Regulation of adrenal scavenger receptor-BI expression by ACTH and cellular cholesterol pools. J Lipid Res, 40, 1799e1805. Sun, B., Eckhardt, E. R., Shetty, S., van der Westhuyzen, D. R., & Webb, N. R. (2006). Quantitative analysis of SR-BI-dependent HDL retroendocytosis in hepatocytes and fibroblasts. J Lipid Res., 47, 1700e1713. Tabas, I. (2005). Consequences and therapeutic implications of macrophage apoptosis in atherosclerosis: the importance of lesion stage and phagocytic efficiency. Arterioscler Thromb Vasc Biol, 25, 2255e2264. Temel, R. E., Trigatti, B., DeMattos, R. B., Azhar, S., Krieger, M., & Williams, D. L. (1997). Scavenger receptor class B, type I (SR-BI) is the major route for the delivery of high density lipoprotein cholesterol to the steroidogenic pathway in cultured mouse adrenocortical cells. Proc Natl Acad Sci USA, 94, 13600e13605. Temel, R., Walzem, R. L., Banka, C. L., & Williams, D. L. (2002). Apolipoprotein A-I is necessary for the in vivo formation of high density lipoprotein competent for scavenger receptor BImediated cholesteryl ester-selective uptake. J Biol Chem, 277, 26565e26572. Tondu, A., Robichon, C., Yvan-Charvet, L., et al. (2005). Insulin and angiotensin II induce the translocation of scavenger receptor class B, type I from intracellular sites to the plasma membrane of adipocytes. J Biol Chem, 280, 33536e33540. Trigatti, B., Rayburn, H., Vinals, M., et al. (1999). Influence of the high density lipoprotein receptor SR-BI on reproductive and cardiovascular pathophysiology. Proc Natl Acad Sci USA, 96, 9322e9327. Ueda, Y., Gong, E., Royer, L., Cooper, P. N., Francone, O. L., & Rubin, E. M. (2000). Relationship between expression levels and atherogenesis in scavenger receptor class B, type I transgenics. J Biol Chem, 275, 20368e20373. Ueda, Y., Royer, L., Gong, E., et al. (1999). Lower plasma levels and accelerated clearance of high density lipoprotein (HDL) and non-HDL cholesterol in scavenger receptor class B type I transgenic mice. J Biol Chem, 274, 7165e7171. Uittenbogaard, A., Shaul, P. W., Yuhanna, I. S., Blair, A., & Smart, E. J. (2000). High density lipoprotein prevents oxidized low density lipoprotein-induced inhibition of endothelial nitricoxide synthase localization and activation in caveolae. J Biol Chem, 275, 11278e11283. Van Eck, M., Bos, I. S., Hildebrand, R. B., Van Rij, B. T., & Van Berkel, T. J. (2004). Dual role for scavenger receptor class B, type I on bone marrow-derived cells in atherosclerotic lesion development. Am J Pathol, 165, 785e794. Varban, M. L., Rinninger, F., Wang, N., et al. (1998). Targeted mutation reveals a central role for SR-BI in hepatic selective uptake of high density lipoprotein cholesterol. Proc Natl Acad Sci USA, 95, 4619e4624. Von Eckardstein, A., & Assmann, G. (2000). Prevention of coronary heart disease by raising highdensity lipoprotein cholesterol? Curr Opin Lipidol, 11, 627e637. Walter, M., Reinecke, H., Nofer, J. R., Seedorf, U., & Assmann, G. (1995). HDL3 stimulates multiple signaling pathways in human skin fibroblasts. Arterioscler Thromb Vasc Biol, 15, 1975e1986. Wang, Q., & Doerschuk, C. M. (2001). The p38 mitogen-activated protein kinase mediates cytoskeletal remodeling in pulmonary microvascular endothelial cells upon intracellular adhesion molecule-1 ligation. J Immunol, 166, 6877e6884. Wang, N., Arai, T., Ji, Y., Rinninger, F., & Tall, A. R. (1998). Liver-specific overexpression of scavenger receptor BI decreases levels of very low density lipoprotein ApoB, low density lipoprotein ApoB, and high density lipoprotein in transgenic mice. J Biol Chem, 273, 32920e32926.
Chapter j 8
The HDL receptor, SR-BI
177
Wang, X., Collins, H. L., Ranalletta, M., et al. (2007). Macrophage ABCA1 and ABCG1, but not SR-BI, promote macrophage reverse cholesterol transport in vivo. J Clin Invest, 117, 2216e2224. Webb, N., Connell, P. M., Graf, G. A., et al. (1998). SR-BII, an isoform of the scavenger receptor BI containing an alternate cytoplasmic tail, mediates lipid transfer between high density lipoprotein and cells. J Biol Chem, 273, 15241e15248. Williams, D. L., de La Llera-Moya, M., Thuahnai, S. T., et al. (2000). Binding and cross-linking studies show that scavenger receptor BI interacts with multiple sites in apolipoprotein A-I and identify the class A amphipathic alpha-helix as a recognition motif. J Biol Chem, 275, 18897e18904. Witt, W., Kolleck, I., Fechner, H., Sinha, P., & Rustow, B. (2000). Regulation by vitamin E of the scavenger receptor BI in rat liver and HepG2 cells. J Lipid Res, 41, 2009e2016. Wu¨stner, D., Mondal, M., Huang, A., & Maxfield, F. R. (2004). Different transport routes for high density lipoprotein and its associated free sterol in polarized hepatic cells. J Lipid Res, 45, 427e437. Yancey, P. G., Bortnick, A. E., Kellner-Weibel, G., de la Llera-Moya, M., Phillips, M. C., & Rothblat, G. H. (2003). Importance of different pathways of cellular cholesterol efflux. Arterioscler Thromb Vasc Biol., 23, 712e719. Yancey, P., Jerome, W. G., Yu, H., et al. (2007). Severely altered cholesterol homeostasis in macrophages lacking apoE and SR-BI. J Lipid Res, 48, 1140e1149. Yeh, Y. C., Hwang, G. Y., Liu, I. P., & Yang, V. C. (2002). Identification and expression of scavenger receptor SR-BI in endothelial cells and smooth muscle cells of rat aorta in vitro and in vivo. Atherosclerosis, 161, 95e103. Yesilaltay, A., Kocher, O., Pal, R., et al. (2006). PDZK1 is required for maintaining hepatic scavenger receptor, class B, type I (SR-BI) steady state levels but not its surface localization or function. J Biol Chem, 281, 28975e28980. Yesilaltay, A., Kocher, O., Rigotti, A., & Krieger, M. (2005). Regulation of SR-BI-mediated highdensity lipoprotein metabolism by the tissue-specific adaptor protein PDZK1. Curr Opin Lipidol, 16, 147e152. Yesilaltay, A., Morales, M. G., Amigo, L. M., et al. (2006). Effects of hepatic expression of the high-density lipoprotein receptor SR-BI on lipoprotein metabolism and female fertility. Endocrinology, 147, 1577e1588. Yu, X., Murao, K., Imachi, H., et al. (2007). Regulation of scavenger receptor class BI gene expression by angiotensin II in vascular endothelial cells. Hypertension, 49, 1378e1384. Yuhanna, I., Zhu, Y., Cox, B. E., et al. (2001). High-density lipoprotein binding to scavenger receptor-BI activates endothelial nitric oxide synthase. Nat. Med, 7, 853e857. Yvan-Charvet, L., Bobard, A., Bossard, P., et al. (2007). In vivo evidence for a role of adipose tissue SR-BI in the nutritional and hormonal regulation of adiposity and cholesterol homeostasis. Arterioscler Thromb Vasc Biol, 27, 1340e1345. Zhang, W., Yancey, P. G., Su, Y. R., et al. (2003). Inactivation of macrophage scavenger receptor class B type I promotes atherosclerotic lesion development in apolipoprotein E-deficient mice. Circulation, 108, 2258e2263. Zhang, S., Picard, M. H., Vasile, E., et al. (2005a). Diet-induced occlusive coronary atherosclerosis, myocardial infarction, cardiac dysfunction, and premature death in scavenger receptor class B type I-deficient, hypomorphic apolipoprotein ER61 mice. Circulation, 111, 457e464. Zhang, Y., Da Silva, J. R., Reilly, M., Billheimer, J. T., Rothblat, G. H., & Rader, D. J. (2005b). Hepatic expression of scavenger receptor class B type I (SR-BI) is a positive regulator of macrophage reverse cholesterol transport in vivo. J Clin Invest, 115, 2870e2874.
178
The HDL Handbook
Zhang, Y., Ahmed, A. M., McFarlane, N., et al. (2007a). Regulation of SR-BI-mediated selective lipid uptake in Chinese hamster ovary-derived cells by protein kinase signaling pathways. J Lip Res, 48, 405e416. Zhang, Y., Ahmed, A. M., Tran, T., et al. (2007b). The inhibition of endocytosis affects HDL-lipid uptake mediated by the human scavenger receptor class B type I. Mol Membr Biol, 24, 442e454.
Chapter 9
HDL Mimetic Peptides: Novel Therapeutic Strategies for the Treatment of Inflammatory Vascular Disease C. Roger White 1, G.M. Anantharamaiah 1, 2 and Geeta Datta 1 1 Departments of Medicine, and 2 Biochemistry, and Molecular Genetics, University of Alabama at Birmingham, Birmingham, AL, USA
INTRODUCTION Despite numerous therapeutic advances, cardiovascular disease (CVD) remains the leading cause of death in the Western world. The Framingham Heart Study demonstrated that CVD risk directly correlates with circulating levels of low density lipoprotein (LDL) (Gordon et al., 1981). Statin (HMGCoA reductase inhibitors) therapy has been successful in lowering LDL and the risk for coronary events (Cannon et al., 2004; Brousseau, 2005). However, even high doses of statins are unable to completely eliminate CVD risk and events. Several studies have established that a reduction in high density lipoprotein (HDL) is associated with increased severity of coronary artery disease (CAD) and is an independent risk factor for CVD (Castelli, 1988). Epidemiologic studies have shown that each 1mg/dL decrease in plasma HDL cholesterol (HDL-C) enhances CVD risk by 2e3%, while an equivalent increase in HDL-C reduces risk by 6%, independent of LDL-C levels (Gordon and Rifkind, 1989). An increase in HDL also reduces vascular complications associated with sepsis, diabetes and atherosclerosis, conditions that occur with greater frequency in aging populations (Rosenson, 2005; Kontush and Chapman, 2008). Despite these findings, analysis of published data from the Framingham Heart Study indicates that a significant number of events occur in subjects with normal LDL and HDL levels (Ansell et al., 2003). More recently, studies in highrisk patients suggest that inflammatory/anti-inflammatory properties of HDL The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
179
180
The HDL Handbook
may be a better indicator of risk than HDL-C levels per se (Ansell et al., 2003). Therefore, therapeutic strategies aimed at increasing HDL and improving its functional properties are currently under development (LinselNitschke and Tall, 2005). Current approaches for raising HDL include both non-pharmacological (life-style) and pharmacological strategies (Singh et al., 2007). Exercise, weight loss, alcohol consumption and cessation of tobacco use are life-style modifications that are associated with a 5e10% increase in HDL-C (Roberts et al., 2006). HDL therapy represents an emerging field directed toward the development of new treatments that increase HDL-C levels and/or improve its anti-atherogenic properties. Pharmacological agents that increase HDL levels comprise a number of different drug classes. These drugs, however, may induce adverse reactions and/or side effects that may preclude their use. In this chapter, we will focus on agents that improve the function of HDL, specifically, apoA-I mimetic peptides.
HETEROGENEITY OF HDL HDL represents a class of heterogeneous particles that are characterized by high density (d > 1.06 g/mL) and low buoyancy. Subclasses of HDL also vary in their content of lipids, apolipoproteins, antioxidant enzymes, and lipid transfer proteins. Resulting differences in HDL size, density and composition ultimately influence particle function. The apolipoprotein composition of HDL determines, to a large extent, its function. Apolipoprotein A-I (apoA-I) accounts for 80e95% of the protein component of a-HDL, a fraction with a-mobility on agarose gels. In contrast, apoA-I comprises only 5e15% of the protein associated with lipid-poor preb-HDL. ApoA-I and apoE (a protein component of very low density lipoprotein (VLDL) and HDL) possess anti-inflammatory and atheroprotective properties. ApoA-II is another HDL-associated apolipoprotein that exerts antiatherogenic effects in numerous animal models (Meyers and Kashyap, 2004). The HDL-associated enzymes paraoxonase (PON) and platelet activating factor-acetyl hydrolase (PAF-AH) hydrolyze oxidized phospholipids and reduce lipid hydroperoxide levels in LDL and VLDL particles and thus render them non-atherogenic (Mackness et al., 2002; Oda et al., 2002). The relative concentration of these proteins per HDL particle will, in large part, reflect HDL function. In a recent study, the HDL proteome has been characterized and has revealed the presence of up to 48 proteins in human HDL (Vaisar et al., 2007). In addition to aforementioned proteins that regulate lipid transfer and metabolism, additional novel proteins that play a role in complement activation and proteolysis were identified (Vaisar et al., 2007). These observations underscore the complexity of HDL and suggest an additional important role for the particle in the regulation of immune function.
Chapter j 9
HDL Mimetic Peptides
181
PHYSIOLOGICAL MECHANISMS OF PROTECTIVE ACTION OF HDL AND APOA-I Anti-atherogenic effects of HDL and apoA-I may be explained by their functional properties. While a principal function of HDL is to mediate cholesterol efflux from peripheral cells, it also possesses anti-inflammatory and antioxidant properties which inhibit the initiation and progression of vascular disease.
Reverse Cholesterol Transport (RCT) RCT is the process by which excess cholesterol from non-hepatic tissues (especially cholesterol-laden, resident macrophages) is transferred to the liver for metabolism and excretion into the bile. Lipid-poor preb-HDL particles, produced in the liver or the intestine, initiate the efflux of cholesterol and phospholipids from cell membranes via interaction with the adenosine triphosphate-binding cassette transporter A1 (ABCA1). Subsequent action of lecithin-cholesterol acyl transferase (LCAT) esterifies cholesterol in preb-HDL particles and converts them to mature a-HDL particles. These particles can take up more cholesterol via the adenosine triphosphate-binding cassette transporter G1 (ABCG1). Mature HDL can deliver cholesterol to the liver either directly via the scavenger receptor type B1 (SR-B1) or indirectly by exchange of cholesteryl esters to apoB-containing particles for triglycerides (TG). In the latter pathway, cholesteryl esters can be exchanged for triglycerides in apoBrich particles (LDL and VLDL) by cholesteryl ester transfer protein (CETP). The uptake of apoB-rich particles via hepatic LDL receptors enables the delivery of cholesterol to the liver (approximately 50% of RCT). The lypolysis of TG in TG-rich HDL by hepatic lipase and endothelial lipase leads to a smaller HDL which re-enters the RCT cycle. Through this cycle, HDL mediates the delivery of cholesterol to the liver where it is metabolized and excreted into bile (Singh et al., 2007). Impairment of RCT due to dysfunctional or reduced HDL has been observed, among others, in the elderly and subjects with CAD, diabetes and Alzheimer’s disease (Clee et al., 2000; Singh-Manoux et al., 2008).
Endothelial function A decrease in nitric oxide (NO) bioavailability accompanied by an increase in leukocyte adhesion to the blood vessel wall are components of endothelial dysfunction. These represent critical early events in atherogenesis. In vitro studies show that HDL protects the endothelium by inhibiting the expression of adhesion molecules (VCAM-1, ICAM-1) and by inducing endothelial nitric oxide synthase (eNOS) activity and NO bioavailability (Assmann and Gotto, 2004). Beneficial effects of HDL have also been observed in vivo (Spiekeret al., 2002; Assmann and Nofer, 2003).
182
The HDL Handbook
Endothelial function in hypercholesterolemic patients is characterized by impairment of NO-dependent vasodilation compared to normolipemic subjects and can be restored by an infusion of cholesterol-free, reconstituted HDL (Nofer et al., 2002). However, this protective effect was not observed with free apoA-I, suggesting that either the lipid components or antioxidant enzymes of HDL contribute to the improvement in endothelial function. Several lines of evidence implicate bioactive sphingolipids (Nofer et al., 2002) in the initiation of signaling events leading to NO release and vasorelaxation.
Antioxidant properties of HDL Oxidized lipids and lipid hydroperoxides are associated with aging, CAD, diabetes and inflammation. Protective effects of HDL can be ascribed, in part, to its ability to inhibit LDL oxidation. Both metal ion- and 12-lipooxygenase-induced oxidation are inhibited by HDL (Nofer et al., 2002). Products of 12-lipooxygenase action, hydroperoxyeicosatetraenoic acid (HPETE) and hydroperoxyoctadecadienoic acid (HPODE), are “seeding molecules” that initiate the non-enzymatic oxidation of lipoprotein phospholipids (Assmann and Nofer, 2003). ApoA-I itself reduces HPETE and HPODE levels (Navab et al., 2000a,b, 2001). Oxidized phospholipids present in LDL are known to stimulate the production of cytokines and chemokines that induce monocyte adhesion to endothelial cells. PON1 and PAF-AH, present on HDL, effectively catalyze the degradation of oxidized phospholipids. This is further supported by transgenic animal studies. PON1 transgenic mice produce HDL that is resistant to oxidation, while PON1 deficient mice are more susceptible to atherosclerosis (Shih et al., 1998, 2000; Oda et al., 2002). Thus, HDL can scavenge oxidized lipids and lipid hydroperoxides and reduce LDL-associated inflammatory responses and atherogenicity (Navab et al., 2002).
Anti-inflammatory properties Inflammation is known to initiate atherosclerosis, diabetes, sepsis and is known to play a major role in aging (Chapman, 2007). Endothelial cells synthesize adhesion molecules (VCAM-1, ICAM-1) and release chemokines (eg., MCP-1) when induced by an inflammatory stimulus. HDL, apoA-I and reconstituted HDL (rHDL) have been shown to inhibit the upregulation of adhesion molecules and the transmigration of monocytes in vitro (Barter et al., 2004). The ability of HDL to modulate adhesion molecule expression in vivo was demonstrated by Dimayuga et al. (1999). Infusion of rHDL to apoE/ mice caused a 40% reduction in the expression of VCAM-1 within 1 week and significantly reduced monocyte infiltration and neointimal hyperplasia within 3 weeks.
Chapter j 9
HDL Mimetic Peptides
183
PRO-INFLAMMATORY HDL Despite the aforementioned protective properties of HDL, the concept that HDL can also be pro-inflammatory and dysfunctional has emerged (Vaisar et al., 2007). Alterations in the composition of HDL subclasses would impair the functional capacity of HDL to mediate cholesterol efflux from peripheral tissues and result in an increase in lipoprotein oxidation (Van Lenten et al., 1995; Navab et al., 2005a) and stimulation of inflammatory responses. Changes in HDL-associated proteins during inflammation induce the formation of pro-inflammatory HDL. Increased incorporation of ceruloplasmin, serum amyloid A (SAA), secretory phospholipase A2 (sPLA2), and apoJ in HDL reduces the capacity of the particle to inhibit LDL-induced chemokine production, monocyte adhesion and lipid hydroperoxide formation. A concomitant decrease in apoA-I, PON1 and PAF-AH further enhances the proinflammatory nature of HDL (Navab et al., 2006; White et al., 2008). Under inflammatory conditions, the neutrophilic enzyme myeloperoxidase (MPO) catalyzes the formation of the potent oxidant hypochlorous acid (HOCl) (Marcinkiewicz, 1997; Eiserich et al., 1998). Recent studies show that MPOderived HOCl oxidizes apoA-I and attenuates the ability of HDL to mediate RCT. Pro-inflammatory HDL has been observed in human subjects with dyslipidemia, hyperglycemia and hypertriglyceridemia (Kontush and Chapman, 2006, 2008). Results from The Framingham Heart Study indicated that 40% of CAD events were observed in subjects with “normal” or “high” HDL suggesting that HDL function, rather than HLD-C levels, is a better indicator of CVD risk. HDL function may be quantitatively assessed using the HDL inflammatory index (HII) (White et al., 2008). HII is the ratio of the LDL-induced monocyte chemotaxis in the presence or absence of test HDL. An increase in LDLinduced chemotaxis by the test HDL (HII > 1.0) indicates pro-inflammatory HDL and a decrease in chemotaxis by the test HDL (HII < 1.0) indicates antiinflammatory HDL (Navab et al., 2006; White et al., 2008). A study of 26 patients with CHD showed that HDL from patients prior to statin therapy had a high HII value which was reduced after treatment (Ansell et al., 2003; Navab et al., 2006). The pro-inflammatory nature of HDL was also noted in another study with 20 subjects who had CHD despite elevated HDL levels (Ansell et al., 2003). HII was found to be significantly elevated in these subjects. While measurement of the HII is useful in assessing functional properties of HDL, the identification of a readily measurable surrogate marker for pro-inflammatory HDL would be ideal. Identification of such a marker, to date, has been elusive.
HDL THERAPY Raising HDL-C levels and/or improving its anti-atherogenic function are principal goals of HDL therapy. The HDL-raising effects of several commonly
184
The HDL Handbook
prescribed medications have been tested and have shown mixed effects. Specifically, the efficacy of niacin, fibrates, and CETP inhibitors in raising serum HDL has been extensively reviewed (Assmann and Nofer, 2003; Singh et al., 2007; White et al., 2008) and will not be addressed by this chapter. The first demonstration that HDL administration served a protective function in experimental animals was reported by Badimon et al. (1990) who showed that administration of HDL to cholesterol-fed rabbits prevented the development of atherogenic lesions. HDL per se is not suitable as a drug since it is a heterogeneous mixture of proteins and lipids that is not easily duplicated by pharmaceutical preparations. In contrast, apoA-Ielipid complexes of defined composition (recombinant HDL or reconstituted HDL (rHDL)) have been tested in various dyslipidemic animal models (Chiesa et al., 2008) and shown prominent anti-atherogenic effects. Clinical studies with rHDL have been performed in small patient cohorts. The first study demonstrated that infusion of rHDL in four hypoalphalipoproteinemic (low HDL) patients induced an increase in plasma HDL-C and elimination of RCT-derived cholesterol degradation products (Eriksson et al., 1999; Chiesa et al., 2008). Patients with familial hypoalphalipo-proteinemia, such as seen in loss of functional mutations in ABCA1, have low HDL with normal levels of LDL and VLDL. Four hours after a single infusion of rHDL in such patients, forearm blood flow was restored, as measured by venous plethysmography. Hypercholesterolemia also causes endothelial dysfunction. In hypercholesterolemic patients, a single infusion of rHDL was able to increase flow-mediated dilatation in these patients and normalize endothelial dependent vasodilation by increasing NO bioavailability. In another small study, patients with established coronary disease received apoA-IMilano, a mutant of apoA-I that possesses enhanced antioxidant activity compared to wild type apoA-I (Bielicki and Oda, 2002). Five weekly treatments of apoAIMilano-phospholipid complexes (45 mg/kg each) reduced coronary atheroma burden as measured by intravascular ultrasound (IVUS). These studies underscore the efficacy of rHDL administration. Even though this treatment appears to have regressed the lesion, the large amount of the protein:lipid complexes required for this effect makes the approach impractical. In the recent ERASE study (Tardif et al., 2007), patients with acute coronary syndrome were given 40 mg or 80 mg/kg of rHDL. The changes in atheroma volume were measured by IVUS. In the 40 mg/kg group, an absolute reduction in atheroma volume of 5.34 mm3 was observed compared to baseline. At this dose, mild to moderate side effects were observed, while, at 80 mg/kg, hepatic dysfunction was significantly increased (Tardif et al., 2007). Reconstituted HDL infusion has also been shown to restore endothelial function in type 2 diabetes patients (Nieuwdorp et al., 2008). Although purified apoA-I appears to have many protective effects, the high concentrations of protein and lipid that are required to observe beneficial effects are financially and physically impractical. A new line of drugs, apolipoprotein mimetic peptides, whose
Chapter j 9
HDL Mimetic Peptides
185
design is based on the structure of apoA-I, represent a promising alternative to apoA-I.
STRUCTURAL CHARACTERISTICS OF APOA-I Exchangeable apolipoproteins possess lipid associating amphipathic helical domains that bind to lipids. This interaction is facilitated by their unique structure which consists of amphipathic helices arranged in tandem (Segrest et al., 1994). An amphipathic helix is defined as an alpha helix with opposing polar and non-polar faces oriented along its long axis. The sidedness of this structural motif imparts a structure complementary to that of phospholipids, thus promoting the interaction between proteins and lipids. Amphipathic helixes are grouped into seven different classes: A, H, L, G, K, C and M (Segrest et al., 1994), each with a distinct charge distribution and density. The most commonly found amphipathic helix in apolipoproteins is the Class A amphipathic a-helix, a zwitterionic helix characterized by a positively charged amino acid cluster at the polarenon-polar interface and the negatively charged residues at the center of the polar face. X-ray and molecular modeling studies of structures of D43-apo A-I (in which 1e43 residues of apo A-I are deleted) and full length apoA-I suggest a lipid bound structure in which two molecules of apoA-I form a belt-like structure in anti-parallel orientation with lipids in the center. The non-polar face of the protein is orientated toward the lipid acyl chains (Borhani et al., 1997; Li et al., 2004). Studies using other techniques such as mass spectroscopy have yielded similar structures, with some variations, depending on the protein/lipid ratios (Davidson and Thompson, 2007; Wu et al., 2007).
APOA-I MIMETIC PEPTIDES In 1985, an 18-residue Class A amphipathic peptide (DWLKAFYDKVAEKLKEAF), whose structure was based on helical repeating domains present in apoA-I, was designed in our laboratory (Anantharamaiah et al., 1985). This peptide was designated 18A and was found to mimic many of the properties of apoA-I. Similar to apoA-I, 18A interacted with lipids to form discoidal HDL-like structures; clarified suspensions of Dimyristoyl Phosphatidylcholine (DMPC) (Anantharamaiah et al., 1985); mediated cholesterol efflux from mouse fibroblasts and macrophages when complexed with DMPC (Mendez et al., 1994); and activated plasma enzyme lecithin cholesterol acyltransferase (LCAT) (Epand et al., 1987). 18A was subsequently modified by blocking the amino terminus with an acetyl group and the carboxy terminus with an amide group. These changes increased the helicity of the resulting peptide (Ac-18A-NH205 or 2F) and significantly enhanced its cholesterol efflux and LCAT activating properties compared to native 18A (Venkatachalapathi et al., 1993). Recently, high resolution NMR experiments on the structure of
186
The HDL Handbook
lipid-bound 2F demonstrated that the Class A peptide orients in a head to tail fashion around the edge of the lipids with the non-polar face toward the lipid (Mishra et al., 2006), in a manner similar to apoA-I (the belt model). Detailed structureefunction studies of Class A peptides were carried out by designing a family of peptides that differed in the content and position of aromatic amino acid residues (Datta et al., 2001). These peptides were generated by systematically increasing the hydrophobicity and lipid affinity of the peptide Ac-18A-NH2. This was achieved by replacing non-polar amino acids with phenylalanine (F) residues. Since Ac-18A-NH2 contains two F residues, it has been alternately designated 2F. The peptides 3F, 4F, 5F, 6F and 7F are additional analogs, all of which possess high lipid binding affinity. However, the increased hydrophobicity of 5F, 6F and 7F also leads to a decrease in solubility. This series of peptides mimicked many properties of apoA-I including the ability to mediate cholesterol efflux, LCAT activating ability and inhibition of monocyte chemotaxis. Although the physical properties of these peptides correlated well with their hydrophobicity, their biological properties were not directly related to hydrophobicity; 4F, 5F and 6F were equipotent in their ability to inhibit LDL-induced monocyte chemotaxis, while 5F had the highest LCAT activating property (Datta et al., 2001). Since all these peptides formed HDL-like structures when complexed to lipids and mimicked many properties of apoA-I, they have been classified as HDLmimetic or apoA-I mimetics. Studies with analogs of bioactive mimetics suggested that the position of the aromatic residues on the non-polar face plays an important role in conferring “biological activity” to these peptides (Data et al., 2004).
In vivo effects of apoA-I mimetic peptides Atheroprotective effects of apoA-I mimetic peptides have been demonstrated in dyslipidemic animal models. Several studies also show that they improve endothelial function, insulin sensitivity and attenuate weight gain in mouse models (Peterson et al., 2008). The peptide 18A protects rats from LPS-induced inflammatory responses and increases survival (Levine et al., 1993). The apoA-I mimetic peptide, 4F also inhibits inflammatory responses and improves survival in septic rats (Zhang et al., 2009). This was associated with an increase in HDL cholesterol and improvement in cardiac output. Recent studies also show an attenuation of vascular remodeling by apoA-I mimetics in transplantation models. These peptides protect against a wide range of inflammatory and vascular disorders.
ApoA-I mimetic peptides inhibit the formation of aortic lesions Since this class of peptides possessed many properties of apoA-I, we investigated their ability to inhibit aortic lesions in dyslipidemic mouse
Chapter j 9
HDL Mimetic Peptides
187
models. Although the peptide 2F possessed many of the properties of apoA-I, a significant inhibition in aortic lesion formation was not observed in C57BL/6J mice on a high fat diet (Datta et al., 2001). The more hydrophobic peptide, 5F, was the first Class A peptide that was shown to have atheroprotective activity. In vivo administration of 5F to C57BL/6J mice fed an atherogenic diet was shown to reduce aortic lesion area compared to mice treated with saline vehicle (Garber et al., 2001). No differences, however, were noted in total plasma cholesterol levels between groups, suggesting that reduction in fatty lesions observed in 5F-treated mice occurred by a mechanism that was independent of cholesterol levels. Examination of functional properties of HDL isolated from the plasma of 5F-treated mice revealed that HDL inhibited LDL-induced monocyte chemotaxis and LDL-associated lipid hydroperoxide levels, while the HDL from the vehicle treated mice did not (Garber et al., 2001). These results suggested that the protective effects of the peptide 5F were not due to changes in plasma HDL levels per se. Rather, it was suggested that antioxidative and anti-inflammatory properties of the peptide were related to an improvement in the functional properties of HDL. Although 5F had greater lipid affinity than 4F, they were both equipotent in inhibiting LDL-induced monocyte chemotaxis. Since 4F is more water soluble, subsequent studies have focused on the effects of 4F on atherosclerosis in apoE/ and LDLR/ mice. Oral administration of the peptide was made possible by the synthesis of 4F using D-amino acids resulting in a peptide (D-4F) that is resistant to the action of proteases. The presence of D-4F in the plasma of LDLR/ mice could be demonstrated 4h after administration by gavage (Navab et al., 2002). Chronic treatment with oral D-4F in LDLR/ mice fed a Western diet showed a 79% reduction in lesions. Similar to 5F treatment, there was no change in the plasma cholesterol or lipoprotein profiles compared to saline controls. However, HDL isolated from these mice showed greater inhibition of LDL-induced monocyte chemotaxis (Navab et al., 2002). Oral treatment with D-4F in apoE/ mice also significantly reduced aortic lesion formation (75% reduction compared to saline-treated controls) (Navab et al., 2004). It was shown that oral treatment with D-4F resulted in the rapid appearance (20 min) of small cholesterol containing particles in the plasma (Navab et al., 2004). These particles contained apoA-I, paraoxonase and had preb-HDL mobility on agarose gels. A decrease in plasma lipid hydroperoxides and an increase in cholesterol efflux were also noted in these experiments, suggesting that D-4F acts by improving HDL function (Navab et al., 2004). More interestingly, in older apoE/ mice, administration of low doses of D-4F in combination with a sub-therapeutic dose of pravastatin induced the formation of anti-inflammatory HDL and regression of aortic lesions in apoE/ mice and cynomologus monkeys (Navab et al., 2005b).
188
The HDL Handbook
ApoA-I mimetic peptides improve endothelial cell function Endothelial dysfunction and impaired vasodilation are early events in atherogenesis. Under hypercholesterolemic conditions, LDL impairs vasodilation by uncoupling eNOS, thus generating superoxide anion (O.2 ) instead of nitric oxide (NO). Using bovine aortic endothelial cells (BAECs), Ou et al. (2003a) showed that 4F, like HDL, improves endothelial vascular function by maintaining coupled eNOS activity to generate NO. Such protection by 4F has also been observed in vivo in LDLR/ (Ou et al., 2003b). Treatment of LDLR/ mice with 4F improved vasodilation in arterioles of these mice compared to vehicle controls. As in the cell culture studies, this improvement was associated with the inhibition of O.2 formation by 4F. The mechanism of O.2 generation in sickle cell disease (SCD) mice is different from the LDL-dependent mechanism observed in LDLR/ mice. In SCD mice, xanthine oxidase (XO) generates O.2 by its action on substrate xanthine and reduces NO activity (White et al., 1996). 4F treatment improved the vasodilation in SCD mice by preventing the binding of XO to endothelial cell surfaces and thus increasing NO bioavailability (Ou et al., 2003b). Recent studies with LDLR/ and LDLR//apoA-I/ double knockout mice demonstrated that 4F restores vascular endothelial cell eNOS function in both types of mice but reduces vessel wall thickness only in the LDLR/ mice suggesting that HDL containing apoA-I is required for 4F to reduce vessel wall thickness (Ou et al., 2005). The effects of 4F on vascular dilation were further corroborated by subsequent studies on tight-skin mice (Weihrauch et al., 2007). The tight-skin mouse (Tsk/þ) is a commonly used model for systemic sclerosis (SSc), an autoimmune disorder characterized by increased oxidant stress and impaired vascular function. Treatment with 4F improved vascular function and decreased myocardial inflammation and pro-inflammatory HDL.
ApoA-I mimetic peptides reduce complications associated with diabetes Oxidative stress plays an important role in diabetes and CVD. In the streptozotocin (STZ)-treated diabetic rat model, an increase in vascular O. 2 formation is associated with endothelial dysfunction and impaired relaxation (Peterson et al., 2007). The increase in superoxide anion and impaired vasodilation in diabetic rats were both rectified by the administration of 4F. These changes were associated with a 4F-induced increase in the antioxidant enzymes heme oxygenase-1 (HO-1) and extracellular superoxide dismutase (EC-SOD) and endothelial progenitor cells (EPC) (Peterson et al., 2007). In diabetic ob/ob mice, administration of 4F improved insulin sensitivity and glucose tolerance (Peterson et al., 2008). This was accompanied by an increase in plasma adiponectin levels and a decrease in inflammatory cytokines, IL-6 and IL-1b. Most
Chapter j 9
HDL Mimetic Peptides
189
interestingly, it also reduced visceral and subcutaneous fat content and restricted weight gain (Peterson et al., 2008).
ApoA-I mimetic peptides improve cognitive function Recent studies (Buga et al., 2006 and Handattu et al., 2009) have shown that administration of D-4F to LDLR null mice anf D-4F with pravastatin to APPSwe-PS1DE9 mice (a murine model of Alzheimer’s disease) significantly improved cognitive function. The treatment also reduced the amyloid burden in the APPSwe-PS1DE9 mice. These studies demonstrate that 4F (similar to HDL) improves symptoms of Alzheimer’s disease.
ApoA-I mimetic peptides inhibit rejection of organ transplantation Cardiac allograft vasculopathy (CAV) is characterized by lymphocyte infiltration and intimal lesion formation and is one of the main causes for the rejection of transplanted hearts. Inflammation and oxidative stress are implicated in the development of CAV. Since HDL and 4F have anti-inflammatory and antioxidant properties and have atheroprotective properties, the effect of 4F on CAV was studied (Weis and von Scheidt, 1997). Donor hearts transplanted in C57BL/ 6 mice treated with 4F or saline were examined. Treatment with 4F significantly reduced the severity of intimal lesions and lymphocyte infiltration after 24 days. These protective effects were, in part, due to the upregulation of HO-1. On the other hand, in vitro studies showed that 4F inhibited T-lymphocyte proliferation and cytokine production by an HO-1 independent pathway. These studies suggest that the protective effects of 4F on CAV are partly dependent on HO-1 but other HO-1 independent pathways are also invoked. Earlier studies showed that 4F administration inhibited lesion formation in vein grafts transplanted into apoE/ (Li et al., 2004). Daily administration of 4F for 4 weeks to mice in which a segment of the inferior vena cava was grafted into the right carotid artery significantly reduced lesions in the graft with a decrease in lipid content and inflammation (Li et al., 2004). However, it had no effect on established lesions in the aortic sinus. These results suggest that the time of 4F administration plays an important role. Most of the protective effects of 4F appear to be due to its anti-inflammatory and antioxidant properties (Figure 9.1). Since oxidized lipids initiate a number of inflammatory responses, the binding properties of 4F to oxidized lipids may be associated with these functions. A recent study examined the binding of 4F and apoA-I to oxidized 1-palmitoyl-2-arachidonoyl-sn-glycero-3-phosphatidylcholine (PAPC) by surface plasma resonance (Van Lenten et al., 2008). The binding constant for 4F was four to six orders of magnitude higher than that of apoA-I. These data suggest that perhaps 4F, with its greater affinity for oxidized lipids, can bind oxidized lipids and inhibit their inflammatory activity.
190
The HDL Handbook
FIGURE 9.1 Under conditions of stress, acute phase proteins can displace apoA-I and PON (antiinflammatory proteins) from HDL converting it to dysfunctional HDL. However, apo-mimetic peptides can reverse this and restore HDL functions.
APOE MIMETIC PEPTIDES Another group of apo-mimetic peptides are the apoE mimetics. ApoE, a protein component of VLDL and HDL, plays a key role in lipoprotein metabolism and atherosclerosis. It facilitated the hepatic uptake of atherogenic lipoproteins and analogous to apoA-I, mediated cholesterol efflux, and has anti-inflammatory and antioxidant properties. Although the anti-atherosclerotic properties of apoE can be attributed, in part, to its role in regulating cholesterol homeostasis, its vascular protective role has been shown to be independent of its cholesterol transport function (Hui, 2004). Ac-hE18A-NH2, designed to structurally mimic apoE, is a unique, dual-domain peptide with the receptor binding domain of apoE (residues 141e150) covalently bound to a lipid associating domain (the Class A amphipathic peptide, 18A) and has the sequence AcLRKLRKRLLRDWLKAFYDKVAEKLKEAF-NH2 (Datta et al., 2000, 2001). We have shown that it binds LDL and VLDL and enhances their uptake by HepG2 cells via the heparan sulfate proteoglycan (HSPG) pathway. This peptide dramatically lowered plasma cholesterol in apoE/ and Watanabe rabbits and improved endothelial function (Garber et al., 2003; Gupta et al., 2005). Ac-hE18A-NH2 also reduced superoxide formation and LPS-induced VCAM-1 expression and has been shown to recycle (Datta et al., 2010). Another apoE mimetic peptide that has shown suppression of systemic and brain inflammation is the peptide corresponding to the 133e149 receptor binding region of apoE (Lynch et al., 2003). Co-administration of this peptide with LPS to wild type C57BL6/J mice significantly reduced serum IL-6 and TNF-a. To enhance transmembrane permeability, this peptide was fused to
Chapter j 9
HDL Mimetic Peptides
191
a protein transduction domain to create COG112 (Li et al., 2006). This peptide inhibited bacterially induced expression of iNOS and the activation of NF-kB in colonic epithelial cells. These results suggest that apoE mimetic peptides with cholesterol reducing and potent anti-inflammatory properties, have significant potential in treating vascular and other inflammation-induced diseases.
OTHER APO-MIMETIC PEPTIDES ApoJ is an apolipoprotein present on HDL which also plays a protective role against inflammation (Navab et al., 1997). Similar to apoA-I, apoJ also inhibits LDL-induced lipid hydroperoxide formation and monocyte chemotaxis (Navab et al., 1997). Screening of short amphipathic peptide sequences from apoJ yielded a class G* amphipathic helix (corresponding to amino acids at positions 113e122 of apoJ) which exerts anti-atherogenic and antiinflammatory effects in mice and monkeys by reducing lipoprotein-associated lipid hydroperoxides and by enhancing the activity of PON (Navab et al., 2005c). Small tetrapeptides (KRES and FREL) that are too small to form helices but have amphipathic properties have also been shown to interact with HDL (Navab et al., 2005d). The four amino acid peptides KRES and FREL exert similar effects on HDL quality but also increase HDL levels in apoE null mice (Navab et al., 2005d). Despite these effects on HDL, KRES does not induce preb-HDL formation or directly stimulate cholesterol efflux from macrophages (Navab et al., 2005d). This is in contrast to the known effects of apoA-I mimetic peptides such as 4F. Ongoing studies are defining mechanisms of small peptide action.
CONCLUSION Peptide mimetics of apolipoproteins A-I, E and J exert anti-inflammatory and atheroprotective effects in dyslipidemic animal models (summarized in Table 9.1). These effects are ascribed, in large part, to their abilities to improve the functional properties of HDL and to remove seeding peroxides from atherogenic lipoproteins. Peptide mimetics increase circulating levels of lipid-poor preb-HDL particles that are effective mediators of cholesterol efflux from macrophages. They also increase the activity of the HDL-associated enzyme PON-1 which reduces LDL atherogenicity by hydrolyzing oxidized lipids. Vasoprotective effects of apolipoprotein mimetics also include an increase in the expression of the antioxidants HO-1 and EC-SOD (Peterson et al., 2007). These enzymes decrease oxidant stress and the formation of pro-inflammatory lipid peroxides, while improving NO bioavailability and endothelial cell function. Whether the induction of HO-1 and EC-SOD is a direct effect of 4F or is related to improvement in HDL function is currently unknown. The apoA-I mimetic 4F is currently undergoing clinical evaluation and has been shown to
192
The HDL Handbook
TABLE 9.1
Properties of apomimetic peptides Peptide
Effect
2F
ApoA-I mimetic
Inhibits LPS-associated toxicity in mice (Levine et al., 1993)
4F
ApoA-I mimetic
Inhibits aortic lesion formation in dyslipidemic animal models (Navab et al., 2002, 2004, 2005b) Sepsis-induced inflammatory responses (Zhang et al., 2009) Organ transplant (Weis and von Scheidt, 1997) Reduces insulin sensitivity (Peterson et al., 2007) Improves cognitive functions (Bulga et al., 2006 and Hattandu et al., 2009) Improves endothelial function (Ou et al., 2003a)
5F
ApoA-I mimetic
Inhibits aortic lesion formation in dyslipidemic mouse models (Garber et al., 2001)
Ac-hE18A-NH2
ApoE mimetic
Inhibits aortic lesion formation in dyslipidemic animal models (Garber et al., 2003; Gupta et al., 2005) Inhibits LPS-induced inflammatory responses (Datta et al., 2010) Recycles from macrophages and hepatocytes (Datta et al., 2010)
ApoJ peptide
ApoJ mimetic
Anti-atherogenic in mice and monkeys (Navah et al., 2005c)
KRES and FREL
Short peptides
Anti-atherogenic (Navab et al., 2005d)
be safe and well tolerated in phase I studies (Bloedon et al., 2008). Most importantly, it reduces the HII in high-risk subjects. These exciting results underscore the therapeutic potential of apolipoprotein mimetic peptides in the treatment of cardiovascular diseases.
REFERENCES Anantharamaiah, G. M., Jones, J. L., Brouillette, C. G., et al. (1985). Studies of synthetic peptide analogs of the amphipathic helix. Structure of complexes with dimyristoyl phosphatidylcholine. J Biol Chem, 260(18), 10248e10255. Ansell, B. J., Navab, M., Hama, S., et al. (2003). Inflammatory/antiinflammatory properties of high-density lipoprotein distinguish patients from control subjects better than high-density lipoprotein cholesterol levels and are favorably affected by simvastatin treatment. Circulation, 108(22), 2751e2756.
Chapter j 9
HDL Mimetic Peptides
193
Assmann, G., & Gotto, A. M., Jr. (2004). HDL cholesterol and protective factors in atherosclerosis. Circulation, 109(23 Suppl. 1), III8eII14. Assmann, G., & Nofer, J. R. (2003). Atheroprotective effects of high-density lipoproteins. Annu Rev Med, 54, 321e341. Badimon, J. J., Badimon, L., & Fuster, V. (1990). Regression of atherosclerotic lesions by high density lipoprotein plasma fraction in the cholesterol-fed rabbit. J Clin Invest, 85(4), 1234e1241. Barter, P. J., Nicholls, S., Rye, K. A., Anantharamaiah, G. M., Navab, M., & Fogelman, A. M. (2004). Antiinflammatory properties of HDL. Circ Res, 95(8), 764e772. Bielicki, J. K., & Oda, M. N. (2002). Apolipoprotein A-I(Milano) and apolipoprotein A-I(Paris) exhibit an antioxidant activity distinct from that of wild-type apolipoprotein A-I. Biochemistry, 41(6), 2089e2096. Bloedon, L. T., Dunbar, R., Duffy, D., et al. (2008). Safety, pharmacokinetics, and pharmacodynamics of oral apoA-I mimetic peptide D-4F in high-risk cardiovascular patients. J Lipid Res, 49(6), 1344e1352. Borhani, D. W., Rogers, D. P., Engler, J. A., & Brouillette, C. G. (1997). Crystal structure of truncated human apolipoprotein A-I suggests a lipid-bound conformation. Proc Natl Acad Sci USA, 94(23), 12291e12296. Brousseau, M. E. (2005). Emerging role of high-density lipoprotein in the prevention of cardiovascular disease. Drug Discov Today, 10(16), 1095e1101. Buga, G. M., Frank, J. S., Mottino, G. A., et al. (2006). D-4F decreases brain arteriole inflammation and improves cognitive performance in LDL receptor-null mice on a Western diet. J. Lipid Res, 47, 2148e2160. Cannon, C. P., Braunwald, E., McCabe, C. H., et al. (2004). Intensive versus moderate lipid lowering with statins after acute coronary syndromes. N Engl J Med, 350(15), 1495e1504. Castelli, W. P. (1988). Cholesterol and lipids in the risk of coronary artery disease e the Framingham Heart Study. Can J Cardiol, 4(Suppl A), 5Ae10A. Chapman, M. J. (2007). Metabolic syndrome and type 2 diabetes: lipid and physiological consequences. Diab Vasc Dis Res, 4(Suppl 3), S5eS8. Chiesa, G., Parolini, C., & Sirtori, C. R. (2008). Acute effects of high-density lipoproteins: biochemical basis and clinical findings. Curr Opin Cardiol, 23(4), 379e385. Clee, S. M., Kastelein, J. J., van Dam, M., et al. (2000). Age and residual cholesterol efflux affect HDL cholesterol levels and coronary artery disease in ABCA1 heterozygotes. J Clin Invest, 106(10), 1263e1270. Datta, G., Chaddha, M., Garber, D. W., et al. (2000). The receptor binding domain of apolipoprotein E, linked to a model class A amphipathic helix, enhances internalization and degradation of LDL by fibroblasts. Biochemistry, 39(1), 213e220. Datta, G., Chaddha, M., Hama, S., et al. (2001). Effects of increasing hydrophobicity on the physical-chemical and biological properties of a class A amphipathic helical peptide. J Lipid Res, 42(7), 1096e1104. Datta, G., Garber, D. W., Chung, B. H., et al. (2001). Cationic domain 141-150 of apoE covalently linked to a class A amphipathic helix enhances atherogenic lipoprotein metabolism in vitro and in vivo. J Lipid Res, 42(6), 959e966. Datta, G., Epand, R. F., Epand, R. M., et al. (2004). Aromatic residue position on the nonpolar face of class A amphipathic helical peptides determines biological activity. J. Biol Chem, 279(24), 26509e26517. Datta, G., White, C. R., Dashti, N., et al. (2010). Anti-inflammatory and recycling properties of an apolipoprotein mimetic peptide, Ac-hE18A-NH(2). Atherosclerosis, 208, 134e141.
194
The HDL Handbook
Davidson, W. S., & Thompson, T. B. (2007). The structure of apolipoprotein A-I in high density lipoproteins. J Biol Chem, 282(31), 22249e22253. Dimayuga, P., Zhu, J., Oguchi, S., et al. (1999). Reconstituted HDL containing human apolipoprotein A-1 reduces VCAM-1 expression and neointima formation following periadventitial cuff-induced carotid injury in apoE null mice. Biochem Biophys Res Commun, 264(2), 465e468. Eiserich, J. P., Hristova, M., Cross, C. E., et al. (1998). Formation of nitric oxide-derived inflammatory oxidants by myeloperoxidase in neutrophils. Nature, 391(6665), 393e397. Epand, R. M., Gawish, A., Iqbal, M., et al. (1987). Studies of synthetic peptide analogs of the amphipathic helix. Effect of charge distribution, hydrophobicity, and secondary structure on lipid association and lecithin:cholesterol acyltransferase activation. J Biol Chem, 262(19), 9389e9396. Eriksson, M., Carlson, L. A., Miettinen, T. A., & Angelin, B. (1999). Stimulation of fecal steroid excretion after infusion of recombinant proapolipoprotein A-I. Potential reverse cholesterol transport in humans. Circulation, 100(6), 594e598. Garber, D. W., Datta, G., Chaddha, M., et al. (2001). A new synthetic class A amphipathic peptide analogue protects mice from diet-induced atherosclerosis. J Lipid Res, 42(4), 545e552. Garber, D. W., Handattu, S., Aslan, I., Datta, G., Chaddha, M., & Anantharamaiah, G. M. (2003). Effect of an arginine-rich amphipathic helical peptide on plasma cholesterol in dyslipidemic mice. Atherosclerosis, 168(2), 229e237. Gordon, D. J., & Rifkind, B. M. (1989). High-density lipoprotein e the clinical implications of recent studies. N Engl J Med, 321(19), 1311e1316. Gordon, T., Kannel, W. B., Castelli, W. P., & Dawber, T. R. (1981). Lipoproteins, cardiovascular disease, and death. Arch Intern Med, 141(9), 1128e1131, The Framingham study. Gupta, H., White, C. R., Handattu, S., et al. (2005). Apolipoprotein E mimetic peptide dramatically lowers plasma cholesterol and restores endothelial function in watanabe heritable hyperlipidemic rabbits. Circulation, 111(23), 3112e3118. Handattu, S. P., Garber, D. W., Monroe, C. E., et al. (2009). Oral apolipoprotein A-I mimetic peptide improves cognitive function and reduces amyloid burden in a mouse model of Alzheimer’s disease. Neurobiol Dis, 34(3), 525e534. Hui, D. Y. (2004). Apolipoprotein E-induced cell signaling in the vessel wall. Rev Endocr Metab Disord, 5(4), 335e341. Kontush, A., & Chapman, M. J. (2006). Functionally defective high-density lipoprotein: a new therapeutic target at the crossroads of dyslipidemia, inflammation, and atherosclerosis. Pharmacol Rev, 58(3), 342e374. Kontush, A., & Chapman, M. J. (2008). Why is HDL functionally deficient in type 2 diabetes? Curr Diab Rep, 8(1), 51e59. Levine, D. M., Parker, T. S., Donnelly, T. M., Walsh, A., & Rubin, A. L. (1993). In vivo protection against endotoxin by plasma high density lipoprotein. Proc Natl Acad Sci USA, 90(24), 12040e12044. Li, L., Chen, J., Mishra, V. K., et al. (2004). Double belt structure of discoidal high density lipoproteins: molecular basis for size heterogeneity. J Mol Biol, 343(5), 1293e1311. Li, X., Chyu, K. Y., Faria Neto, J. R., et al. (2004). Differential effects of apolipoprotein AI-mimetic peptide on evolving and established atherosclerosis in apolipoprotein E-null mice. Circulation, 110(12), 1701e1705. Li, F. Q., Sempowski, G. D., McKenna, S. E., Laskowitz, D. T., Colton, C. A., & Vitek, M. P. (2006). Apolipoprotein E-derived peptides ameliorate clinical disability and inflammatory infiltrates into the spinal cord in a murine model of multiple sclerosis. J Pharmacol Exp Ther, 318(3), 956e965.
Chapter j 9
HDL Mimetic Peptides
195
Linsel-Nitschke, P., & Tall, A. R. (2005). HDL as a target in the treatment of atherosclerotic cardiovascular disease. Nat Rev Drug Discov, 4(3), 193e205. Lynch, J. R., Tang, W., Wang, H., et al. (2003). APOE genotype and an ApoE-mimetic peptide modify the systemic and central nervous system inflammatory response. J Biol Chem, 278(49), 48529e48533. Mackness, B., Durrington, P. N., Boulton, A. J., Hine, D., & Mackness, M. I. (2002). Serum paraoxonase activity in patients with type 1 diabetes compared to healthy controls. Eur J Clin Invest, 32(4), 259e264. Marcinkiewicz, J. (1997). Nitric oxide and antimicrobial activity of reactive oxygen intermediates. Immunopharmacology, 37(1), 35e41. Mendez, A. J., Anantharamaiah, G. M., Segrest, J. P., & Oram, J. F. (1994). Synthetic amphipathic helical peptides that mimic apolipoprotein A-I in clearing cellular cholesterol. J Clin Invest, 94(4), 1698e1705. Meyers, C. D., & Kashyap, M. L. (2004). Pharmacologic elevation of high-density lipoproteins: recent insights on mechanism of action and atherosclerosis protection. Curr Opin Cardiol, 19 (4), 366e373. Mishra, V. K., Anantharamaiah, G. M., Segrest, J. P., et al. (2006). Association of a model class A (apolipoprotein) amphipathic alpha helical peptide with lipid: high resolution NMR studies of peptide.lipid discoidal complexes. J Biol Chem, 281(10), 6511e6519. Navab, M., Anantharamaiah, G. M., & Fogelman, A. M. (2005a). The role of high-density lipoprotein in inflammation. Trends Cardiovasc Med, 15(4), 158e161. Navab, M., Anantharamaiah, G. M., Hama, S., et al. (2002). Oral administration of an Apo A-I mimetic peptide synthesized from D-amino acids dramatically reduces atherosclerosis in mice independent of plasma cholesterol. Circulation, 105(3), 290e292. Navab, M., Anantharamaiah, G. M., Reddy, S. T., et al. (2004). Oral D-4F causes formation of prebeta high-density lipoprotein and improves high-density lipoprotein-mediated cholesterol efflux and reverse cholesterol transport from macrophages in apolipoprotein E-null mice. Circulation, 109(25), 3215e3220. Navab, M., Anantharamaiah, G. M., Hama, S., et al. (2005b). D-4F and statins synergize to render HDL antiinflammatory in mice and monkeys and cause lesion regression in old apolipoprotein E-null mice. Arterioscler Thromb Vasc Biol, 25(7), 1426e1432. Navab, M., Anantharamaiah, G. M., Reddy, S. T., et al. (2005c). An oral apoJ peptide renders HDL antiinflammatory in mice and monkeys and dramatically reduces atherosclerosis in apolipoprotein E-null mice. Arterioscler Thromb Vasc Biol, 25(9), 1932e1937. Navab, M., Anantharamaiah, G. M., Reddy, S., et al. (2005d). Oral small peptides render HDL antiinflammatory in mice and monkeys and reduce atherosclerosis in ApoE null mice. Circ Res, 97(6), 524e532. Navab, M., Berliner, J. A., Subbanagounder, G., et al. (2001). HDL and the inflammatory response induced by LDL-derived oxidized phospholipids. Arterioscler Thromb Vasc Biol, 21(4), 481e488. Navab, M., Hama, S. Y., Anantharamaiah, G. M., et al. (2000a). Normal high density lipoprotein inhibits three steps in the formation of mildly oxidized low density lipoprotein: steps 2 and 3. J Lipid Res, 41(9), 1495e1508. Navab, M., Hama, S. Y., Cooke, C. J., et al. (2000b). Normal high density lipoprotein inhibits three steps in the formation of mildly oxidized low density lipoprotein: step 1. J Lipid Res, 41(9), 1481e1494. Navab, M., Hama-Levy, S., Van Lenten, B. J., et al. (1997). Mildly oxidized LDL induces an increased apolipoprotein J/paraoxonase ratio. J Clin Invest, 99(8), 2005e2019.
196
The HDL Handbook
Navab, M., Anantharamaiah, G. M., Reddy, S. T., Van Lenten, B. J., Ansell, B. J., & Fogelman, A. M. (2006). Mechanisms of disease: proatherogenic HDL e an evolving field. Nat Clin Pract Endocrinol Metab, 2(9), 504e511. Nieuwdorp, M., Vergeer, M., Bisoendial, R. J., et al. (2008). Reconstituted HDL infusion restores endothelial function in patients with type 2 diabetes mellitus. Diabetologia, 51(6), 1081e1084. Nofer, J. R., Kehrel, B., Fobker, M., Levkau, B., Assmann, G., & von Eckardstein, A. (2002). HDL and arteriosclerosis: beyond reverse cholesterol transport. Atherosclerosis, 161(1), 1e16. Oda, M. N., Bielicki, J. K., Ho, T. T., Berger, T., Rubin, E. M., & Forte, T. M. (2002). Paraoxonase 1 overexpression in mice and its effect on high-density lipoproteins. Biochem Biophys Res Commun, 290(3), 921e927. Ou, J., Ou, Z., Jones, D. W., et al. (2003b). L-4F, an apolipoprotein A-1 mimetic, dramatically improves vasodilation in hypercholesterolemia and sickle cell disease. Circulation, 107(18), 2337e2341. Ou, J., Wang, J., Xu, H., et al. (2005). Effects of D-4F on vasodilation and vessel wall thickness in hypercholesterolemic LDL receptor-null and LDL receptor/apolipoprotein A-I doubleknockout mice on Western diet. Circ Res, 97(11), 1190e1197. Ou, Z., Ou, J., Ackerman, A. W., Oldham, K. T., & Pritchard, K. A., Jr. (2003a). L-4F, an apolipoprotein A-1 mimetic, restores nitric oxide and superoxide anion balance in low-density lipoprotein-treated endothelial cells. Circulation, 107(11), 1520e1524. Peterson, S. J., Drummond, G., Kim, D. H., et al. (2008). L-4F treatment reduces adiposity, increases adiponectin levels, and improves insulin sensitivity in obese mice. J Lipid Res, 49(8), 1658e1669. Peterson, S. J., Husney, D., Kruger, A. L., et al. (2007). Long-term treatment with the apolipoprotein A1 mimetic peptide increases antioxidants and vascular repair in type I diabetic rats. J Pharmacol Exp Ther, 322(2), 514e520. Roberts, C. K., Ng, C., Hama, S., Eliseo, A. J., & Barnard, R. J. (2006). Effect of a short-term diet and exercise intervention on inflammatory/anti-inflammatory properties of HDL in overweight/obese men with cardiovascular risk factors. J Appl Physiol, 101(6), 1727e1732. Rosenson, R. S. (2005). HDL-C and the diabetic patient: target for therapeutic intervention? Diabetes Res Clin Pract, 68(Suppl 2), S36eS42. Segrest, J. P., Garber, D. W., Brouillette, C. G., Harvey, S. C., & Anantharamaiah, G. M. (1994). The amphipathic alpha helix: a multifunctional structural motif in plasma apolipoproteins. Adv Protein Chem, 45, 303e369. Shih, D. M., Gu, L., Xia, Y. R., et al. (1998). Mice lacking serum paraoxonase are susceptible to organophosphate toxicity and atherosclerosis. Nature, 394(6690), 284e287. Shih, D. M., Xia, Y. R., Wang, X. P., et al. (2000). Combined serum paraoxonase knockout/ apolipoprotein E knockout mice exhibit increased lipoprotein oxidation and atherosclerosis. J Biol Chem, 275(23), 17527e17535. Singh, I. M., Shishehbor, M. H., & Ansell, B. J. (2007). High-density lipoprotein as a therapeutic target: a systematic review. J Am Med Assoc, 298(7), 786e798. Singh-Manoux, A., Gimeno, D., Kivimaki, M., Brunner, E., & Marmot, M. G. (2008). Low HDL cholesterol is a risk factor for deficit and decline in memory in midlife: the Whitehall II study. Arterioscler Thromb Vasc Biol, 28(8), 1556e1562. Spieker, L. E., Sudano, I., Hurlimann, D., et al. (2002). High-density lipoprotein restores endothelial function in hypercholesterolemic men. Circulation, 105(12), 1399e1402. Tardif, J. C., Gregoire, J., L’Allier, P. L., et al. (2007). Effects of reconstituted high-density lipoprotein infusions on coronary atherosclerosis: a randomized controlled trial. J Am Med Assoc, 297(15), 1675e1682.
Chapter j 9
HDL Mimetic Peptides
197
Vaisar, T., Pennathur, S., Green, P. S., et al. (2007). Shotgun proteomics implicates protease inhibition and complement activation in the antiinflammatory properties of HDL. J Clin Invest, 117(3), 746e756. Van Lenten, B. J., Hama, S. Y., de Beer, F. C., et al. (1995). Anti-inflammatory HDL becomes proinflammatory during the acute phase response. Loss of protective effect of HDL against LDL oxidation in aortic wall cell cocultures. J Clin Invest, 96(6), 2758e2767. Van Lenten, B. J., Wagner, A. C., Jung, C. L., et al. (2008). Anti-inflammatory apoA-I-mimetic peptides bind oxidized lipids with much higher affinity than human apoA-I. J Lipid Res, 49(11), 2302e2311. Venkatachalapathi, Y. V., Phillips, M. C., Epand, R. M., et al. (1993). Effect of end group blockage on the properties of a class A amphipathic helical peptide. Proteins, 15(4), 349e359. Weihrauch, D., Xu, H., Shi, Y., et al. (2007). Effects of D-4F on vasodilation, oxidative stress, angiostatin, myocardial inflammation, and angiogenic potential in tight-skin mice. Am J Physiol Heart Circ Physiol, 293(3), H1432eH1441. Weis, M., & von Scheidt, W. (1997). Cardiac allograft vasculopathy: a review. Circulation, 96(6), 2069e2077. White, C. R., Darley-Usmar, V., Berrington, W. R., et al. (1996). Circulating plasma xanthine oxidase contributes to vascular dysfunction in hypercholesterolemic rabbits. Proc Natl Acad Sci USA, 93(16), 8745e8749. White, C. R., Datta, G., Zhang, Z., et al. (2008). HDL therapy for cardiovascular diseases: the road to HDL mimetics. Curr Atheroscler Rep, 10(5), 405e412. Wu, Z., Wagner, M. A., Zheng, L., et al. (2007). The refined structure of nascent HDL reveals a key functional domain for particle maturation and dysfunction. Nat Struct Mol Biol, 14(9), 861e868. Zhang, Z., Datta, G., Zhang, Y., et al. (2009). Apolipoprotein A-I mimetic peptide treatment inhibits inflammatory responses and improves survival in septic rats. Am J Physiol Heart Circ Physiol, 297(2), H866eH873.
Chapter 10
Sterol Efflux by ABCA1 and ABCG1 Naoki Terasaka Biological Research Laboratories, Daiichi Sankyo Co., Ltd, Tokyo, Japan
INTRODUCTION Plasma high density lipid (HDL) levels are inversely related to the risk of atherosclerotic cardiovascular disease (Rhoads et al., 1976; Castelli et al., 1986). HDL has anti-inflammatory, antioxidant, antithrombotic and vasodilatory properties that may be relevant to this relationship (Assmann and Gotto, 2004; Barter et al., 2004). One of the most important atheroprotective roles of HDL is reverse cholesterol transport (RCT), where excess cholesterol in macrophage foam cells undergoes efflux and then is transported to the liver for excretion in the bile (Rader, 2006). In recent years, there has been substantial progress in understanding the biology of HDL and its interaction with cells, and the RCT hypothesis has been modified and amplified. Progress has been emphasized by the discovery of new molecules involved in the different steps of RCT or other aspects of HDL biology, development of transgenic mouse models and the beginning of human clinical trials targeting HDL. A remarkable feature of the RCT pathway is that each of the individual steps involves molecules that are directly regulated by liver X receptors (LXRs), including ATP-binding cassette transporters ABCA1 and ABCG1 in macrophages (Tall et al., 2002; Tontonoz and Mangelsdorf, 2003). In the absence of ABCA1 and ABCG1, macrophages massively accumulate cholesteryl esters (CEs), highlighting the physiological importance of cholesterol efflux in the pathogenesis of atherosclerosis (Out et al., 2007a; Yvan-Charvet et al., 2007a). This chapter will summarize recent advances in the basic science of HDL and ABCA1 and ABCG1.
ABCA1-MEDIATED CHOLESTEROL EFFLUX The ability of HDL and its major apolipoprotein, apoA-I, to stimulate efflux of cholesterol from macrophage foam cells in atherosclerotic lesions is thought to be central to its anti-atherogenic mechanism, and to represent the first step in an The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
199
200
The HDL Handbook
overall process of RCT. A major breakthrough in the HDL field was the discovery that mutations in the ABCA1 are responsible for Tangier disease (Bodzioch et al., 1999; Brooks-Wilson et al., 1999; Rust et al., 1999), a condition characterized by an almost complete deficiency of plasma HDL, macrophage foam cell accumulation in various tissues, peripheral neuropathy and an apparent moderate increase in atherosclerosis. More than twenty different mutations in ABCA1 have been shown to cause Tangier disease. Moreover, about 10% of individuals with very low HDL have a variety of different mutations in ABCA1 (Cohen et al., 2004; Frikke-Schmidt et al., 2004). The human ABCA1 gene has been mapped to chromosome 9q31 and is composed of 50 exons, which encode 2261-amino acid residues (SantamarinaFojo et al., 2000). The ABCA1 protein is a full-size ABC transporter containing two transmembrane domains of six alpha helices and two intracellular nucleotide binding domains (Santamarina-Fojo et al., 2000). Like Tangier disease patients, ABCA1 knockout mice exhibit HDL deficiency and reduced cellular cholesterol efflux activity (Fitzgerald et al., 2001). Both systemic and selective hepatic overexpression of ABCA1 in mice results in an increase of HDL (Vaisman et al., 2001; Singaraja et al., 2002). In mice with a liverspecific deletion of ABCA1, apoA-I and HDL plasma levels are dramatically reduced (Timmins et al., 2005). By contrast, the selective inactivation or expression of ABCA1 in macrophages had little or no effect on the plasma concentration of HDL (Haghpassand et al., 2001). Therefore, hepatic ABCA1 expression is a rate-limiting step for plasma HDL production, whereas macrophages do not contribute significantly to the formation of HDL. However, the selective knockout of ABCA1 in macrophages of either ApoE/ or Ldlr/ mice significantly enhanced the development of atherosclerosis (Francis et al., 1995). Thus, although ABCA1 in macrophages has little influence on HDL plasma levels, it is a crucial factor in the prevention of excessive cholesterol accumulation in macrophages of the arterial wall, independently of plasma HDL levels. ABCA1 gene expression is induced by LXRa or LXRb and the retinoid-Xreceptor (RXR) which form heterodimers and are activated by oxysterols and retinoids, respectively (Costet et al., 2000). The most likely physiological ligand for LXRa and LXRb is 27-hydroxycholesterol, which is produced by the cytochrome P450 enzyme 27-cholesterol hydroxylase (CYP27). The levels of 27-hydroxycholesterol correlate with cellular cholesterol levels highlighting the regulation of ABCA1-mediated cholesterol efflux by oxysterols. Indeed, increasing levels of cellular cholesterol lead to the formation of 27hydroxycholesterol, which binds to LXRs inducing ABCA1 and thereby cholesterol efflux (Fu et al., 2001). Conversely, unsaturated fatty acids inhibit ABCA1 expression presumably because they compete with oxysterols for the binding to LXRa and LXRb (Ou et al., 2001). cAMP not only stimulates the expression of ABCA1 in macrophages but also its phosphorylation (Oram et al., 2000; Haidar et al., 2002, 2004). This involves protein kinase A, occurs at serine residues 1042 and 2054 and
Chapter j 10 Sterol Efflux by ABCA1 and ABCG1
201
increases the cholesterol efflux activity of ABCA1. Interestingly, apoA-I activates cAMP signaling and thereby the phosphorylation of ABCA1, its abundance on the cell surface, and cholesterol efflux (Haidar et al., 2004). Janus kinase 2 (JAK2) also phosphorylates ABCA1, a process that increases apoA-I binding and cholesterol efflux (Roosbeek et al., 2004; Tang et al., 2004). However, in the absence of apolipoproteins, phosphorylation of threonine residues 1286 and 1305 within the ABCA1 PEST motif is a signal for calpainmediated degradation of ABCA1 (Martinez et al., 2003). Other signals leading to ABCA1 degradation are exerted by unsaturated fatty acids via a phospholipase D2-dependent pathway (Wang and Oram, 2002, 2005) and by excess free cholesterol (FC), which induces proteasomal degradation. The intracellular targeting of ABCA1 is also regulated by various adapter proteins. a1syntrophin (Munehira et al., 2004), b1-syntrophin (Okuhira et al., 2005), b2syntrophin (Buechler et al., 2002), and Cdc42 (Tsukamoto et al., 2001), which interact with the carboxy-terminal domain of ABCA1.
ABCG1-MEDIATED CHOLESTEROL EFFLUX Whereas ABCA1 promotes cholesterol efflux to lipid-poor apoA-I, it only modestly stimulates lipid efflux to smaller HDL3 particles and does not promote cholesterol efflux to the larger HDL2 fraction (Francis et al., 1995; Wang et al., 2000). This suggested the possibility that there might be another cholesterol efflux pathway promoting lipid efflux to HDL particles which led to the discovery that ABCG1 promotes cholesterol efflux from transfected cells to HDL particles but not to lipid-poor apoA-I (Figure 10.1) (Wang et al., 2004). ApoE and lecithin:cholesterol acyltransferase (LCAT) in the HDL particles have an important role in promoting cholesterol efflux via ABCG1, especially in CETP deficiency states (Matsuura et al., 2006). ABCG1 is a half-transporter that is likely to act as homo-dimers. ABCG1 is highly expressed in macrophages and promotes cholesterol efflux from macrophage foam cells to HDL particles (Wang et al., 2006). The human ABCG1 gene has been mapped to chromosome 21q22.3. It is composed of 23 exons and has multiple transcripts (Kennedy et al., 2001). ABCG1 mRNA levels are highly increased when macrophages are incubated with LXR agonists (Nakamura et al., 2004). Its silencing by RNA interference results in reduced efflux of cholesterol and phospholipid to HDL. Conversely, its overexpression causes an increase in HDL-mediated cholesterol efflux and a reduction in cellular cholesterol mass (Nakamura et al., 2004; Wang et al., 2004; Vaughan and Oram, 2005). These results were confirmed in knockout and transgenic mice. The disruption of ABCG1 in mice on a high-fat and cholesterol-rich diet causes the accumulation of neutral lipids and phospholipids in hepatocytes and macrophages, although not affecting plasma lipids (Kennedy et al., 2005). In contrast, overexpression of ABCG1 is protective against lipid accumulation (Kennedy et al., 2005).
202
The HDL Handbook
HDL ABCG1
FC
LCAT ABCA1
HL, EL PLTP Free ApoAI
FIGURE 10.1 Role of ABCA1 and ABCG1 transporters in cholesterol efflux in macrophages. Lipid-free or lipid-poor apoA-I can interact with ABCA1 in macrophage foam cells in atherosclerotic lesions, promoting efflux of free cholesterol (FC). This results in the formation of nascent HDL particles that are further modified by lecithin:cholesterol acyltransferase (LCAT) generating cholesterol esters (CEs) and forming mature HDL. HDL is also formed by ABCA1 in the liver and intestine (not shown). Mature HDL particles can serve as acceptors for ABCG1-mediated cholesterol efflux. In the sub-endothelial space, mature HDL particles are remodeled in the arterial wall by hepatic lipases (HL), endothelial lipase (EL) and phospholipid transfer protein (PLTP) secreted by macrophages. In addition, large HDL may promote efflux of cholesterol and oxysterols from endothelial cells, which express abundant ABCG1.
CELLULAR MECHANISMS OF CHOLESTEROL EFFLUX VIA ABCA1 AND ABCG1 ABCA1 and ABCG1 are ATPases that promote unidirectional, net cholesterol efflux to lipid-poor helical apolipoproteins and lipoprotein particles, respectively (Wang et al., 2006). In transfected cells and in macrophages, ABCA1 and ABCG1 can act in a sequential fashion, with ABCA1 generating nascent HDL particles which then promote cholesterol efflux via ABCG1 (Gelissen et al., 2006). Genetic knock-down studies suggest that ABCA1 and ABCG1 together account for about 60e70% of the net cholesterol efflux to HDL or serum from cholesterol-loaded LXR-activated macrophages (Yvan-Charvet et al., 2007b). Even though scavenger receptor class BI (SR-BI) can promote the bidirectional exchange of FC between cells and HDL, SR-BI knockout macrophages have no change in net cholesterol efflux to HDL, suggesting it does not make a significant contribution to this process. ABCA1 promotes efflux of phospholipids and cholesterol to lipid-poor apoA-I in a process that involves the direct binding of apoA-I to the transporter (Oram et al., 2000; Wang et al., 2000). Most likely ABCA1 translocates phospholipids from the inner to the outer membrane leaflet of the plasma membrane, perhaps creating outward curvature and packing defects in the membrane; this may allow interpolation of amphipathic helices of apoA-I into the membrane and formation of nascent HDL particles (Vedhachalam et al.,
203
Chapter j 10 Sterol Efflux by ABCA1 and ABCG1
A
B HDL
HDL
FC, 7-KC
PL, FC, 7-KC III
ABCG1
ABCG1
I
II
IV
FC, 7-KC
ER
unsaturated phospholipid
FC, 7-KC
ER
saturated phospholipid/sphingomyelin
FIGURE 10.2 Potential mechanisms of sterol efflux mediated by ABCG1. Two different models of ABCG1- mediated sterol efflux have been proposed. (A) One model suggests that ABCG transporters accomplish free cholesterol (FC) and 7-ketocholesterol (7-KC) transfer by helping sterol molecules to overcome the energy barrier for entry into the hydrophilic water layer perhaps by utilizing ATP to promote protrusion of the cholesterol molecule into water, followed by a transient collision with acceptor, in this case HDL. (B) The ability of ABCG1 to promote phospholipid (PL) efflux points to a function as a phospholipid flippase mediating the transfer of PL from the inner- to the outer-leaflet (I). This could lead to such an extensive change in the equilibrium of membrane components that the outer-leaflet becomes more attractive to sterol (e.g., an increased content of sphingomyelin (SM) or saturated phosphatidyl choline (PC)), followed by transbilayer diffuson of cholesterol molecules towards the outer leaflet (II) where they can dissociate onto HDL particles perhaps following non-specific binding of HDL to the plasma membrane (III). In both models, the movement of cholesterol from the inner to the outer membrane is followed by carrier-facilitated diffusion from cellular organelles notably the endoplasmic reticulum (IV) leading to altered sterol-mediated ER regulation of cholesterol homeostasis.
2007). Cholesterol and phospholipid efflux via ABCA1 appear to occur simultaneously (Smith et al., 2004) and it is possible that the transporter also translocates cholesterol onto the forming HDL particle (Gillotte-Taylor et al., 2002). In contrast, ABCG1 promotes efflux of cholesterol onto a variety of lipoprotein particles, including HDL, low density lipoprotein (LDL), phospholipid vesicles and cyclodextrin, but ABCG1 does not appear to bind lipoprotein particles (Wang et al., 2004, 2006). Overexpression of ABCG1 also promotes efflux of choline-containing phospholipids onto HDL (Wang et al., 2006), likely including both sphingomyelin (SM) and phosphatidyl choline (PC) (Kobayashi et al., 2006). Two potential mechanisms of sterol efflux by ABCG transporters have been proposed (Figure 10.2): 1. Small (2003) has suggested that ABCG5/8, heterodimeric transporters that promote secretion of cholesterol and plant sterols into bile, mediate protrusion of the hydrophobic sterol molecule into the aqueous phase, followed by
204
The HDL Handbook
collision with a micelle. By analogy ABCG1 could promote protrusion of cholesterol from the plasma membrane followed by transient collision with an HDL particle (see Figure 10.2A) 2. Alternatively, ABCG1 could change the organization of phospholipids in the plasma membrane such that the membrane more readily releases sterol to lipoprotein acceptors (see Figure 10.2B). Cholesterol efflux would then be facilitated by diffusion or collision of lipoprotein particles with the plasma membrane. This model is consistent with the non-specific nature and lack of binding of lipoprotein acceptors by ABCG1. It is notable that ABCG1 appears to promote movement of cholesterol from the endoplasmic reticulum (ER) to the plasma membrane, manifested as decreased cellular ACAT activity and decreased activity of SREBP-2 target genes, even in the absence of extracellular lipoprotein acceptors (Wang et al., 2006). Thus, ABCG1 appears to increase the affinity of the plasma membrane for ERderived sterols even while releasing sterols more readily to extracellular lipoprotein acceptors (Wang et al., 2006). This could be achieved by promoting the movement of sterol from the inner to the outer leaflet of the plasma membrane. Since sterol can readily flip-flop across bilayer membranes (Lange et al., 1981), these findings imply a change in phospholipid organization that results in a change in the bilayer distribution of sterol, such that the sterol content of the outer leaflet of the plasma membrane is increased.
THE ROLE OF ABCA1- AND ABCG1-MEDIATED CHOLESTEROL EFFLUX PATHWAYS IN ATHEROSCLEROSIS Several different potential cholesterol efflux pathways from macrophages to HDL have been documented: passive or diffusional efflux; efflux associated with macrophage apoE secretion; SR-BI-mediated cholesterol efflux; and active cholesterol efflux mediated by ABCA1 and ABCG1 (Yancey et al., 2003; Tall, 2003). Recent studies indicate that, together, ABCA1 and ABCG1 have the major role in mediating net cholesterol efflux from macrophages to HDL or serum (Adorni et al., 2007; Out et al., 2007a; Yvan-Charvet et al., 2007a, 2008). When one transporter is deficient, the other is induced as a result of sterol accumulation and LXR activation, resulting in mutual compensation in the activities of the two transporters (Ranalletta et al., 2006; Yvan-Charvet et al., 2007b). Macrophages with combined deficiency of ABCA1 and ABCG1 have major defects in cholesterol efflux to apoA-I, HDL and serum (Out et al., 2007a; Yvan-Charvet et al., 2007a). In one study, combined deficiency of ABCA1 and ABCG1 resulted in a 60% decrease in macrophage net cholesterol efflux to HDL (Yvan-Charvet et al., 2007a), while in another study, the decrease in cholesterol efflux was reported as 100% (Out et al., 2007a). A third set of studies suggested major roles of ABCA1 and ABCG1 in cholesterolloaded macrophages and a relatively larger role of passive cholesterol efflux
Chapter j 10 Sterol Efflux by ABCA1 and ABCG1
205
from cells to HDL especially when macrophages are not loaded with cholesterol (Adorni et al., 2007). Abca1/Abcg1/ mice showed much greater accumulation of cholesterol in peritoneal macrophages on either chow or high cholesterol diets compared to either Abca1/ or Abcg1/ mice (Out et al., 2007a,b; Yvan-Charvet et al., 2007a). Together these studies have confirmed that ABCA1 and ABCG1 account for the major portion of the net cholesterol efflux from cholesterol-loaded macrophages to plasma lipoproteins in vivo. The atherosclerosis studies provide a close parallel with the cholesterol efflux data and indicate that combined deficiency of ABCA1 and ABCG1 in bone marrow-derived cells results in a dramatic deterioration of atherosclerosis. Transplantation of Abca1/ bone marrow into Ldlr/ or ApoE/ mice results in a modest increase in atherosclerosis (Aiello et al., 2002; van Eck et al., 2002). In contrast, transplantation of Abcg1/ bone marrow into Ldlr/ or apoE/ mice resulted in either no change or a decrease in atherosclerosis (Baldan et al., 2006; Ranalletta et al., 2006; Out et al., 2007b). The decrease in atherosclerosis was attributed to either compensatory upregulation of ABCA1 in Abcg1/ macrophages (Ranalletta et al., 2006), or to enhanced apoptosis of Abcg1/ macrophage (Baldan et al., 2006). Importantly, transplantation of Abca1/Abcg1/ bone marrow into Ldlrþ/ mice, followed by feeding a high cholesterol diet led to markedly increased atherosclerosis compared to mice receiving bone marrow with single deficiency of ABCA1 or ABCG1 (Yvan-Charvet et al., 2007a). Moreover, pharmacological studies with synthetic LXR agonists demonstrated the increased expression of ABCA1 and ABCG1 in the arterial wall, and consequent reduction of atherosclerosis in both ApoE/ and Ldlr/ mice (Joseph et al., 2002; Terasaka et al., 2003).
A POTENTIAL ROLE OF HDL, ABCA1 AND ABCG1 IN MACROPHAGE APOPTOSIS AND INFLAMMATION IN PLAQUES The death of macrophages in atherosclerotic lesions by apoptosis or postapoptotic necrosis is thought to contribute to inflammation, necrotic core formation and destabilization of plaques (Tabas, 2005). HDL was found to protect macrophages from apoptosis induced by oxidized LDL, or by loading with FC (Cui et al., 2007; Terasaka et al., 2007). In the case of oxidized LDLinduced apoptosis, the protective effect of HDL was abolished in Abcg1/ macrophages (Terasaka et al., 2007). In contrast, for FC-induced apoptosis, deficiency of both ABCG1 and ABCA1 was required to see an abolition of the protective effect of HDL (Yvan-Charvet et al., 2007b). ABCG1 has a specific role in promoting cellular efflux of sterols modified at the 7-position such as 7-ketocholesterol (7-KC) (Terasaka et al., 2007). 7-KC is a spontaneously formed cholesterol oxidation product that is present in high cholesterol diets, and is the most abundant oxysterol in oxidized LDL and in human atherosclerotic plaques (Brown et al., 1996; Vine et al., 1998; Brown and Jessup,
206
The HDL Handbook
1999). Dietary 7-KC is normally absorbed on chylomicrons, rapidly cleared from the circulation in remnants and converted into bile salts in the liver. However, at concentrations apparently found in vivo 7-KC is cytotoxic, inducing apoptosis and necrosis of endothelial cells and macrophages. Previous studies found that HDL could promote efflux of 7-KC from macrophages, while apoA-I had only limited ability to do so (Terasaka et al., 2007). HDL and ABCG1 have a specific role in promoting efflux of 7-KC and 7b-OH cholesterol in transfected 293 cells and macrophages, while ABCA1 and apoA-I have no ability to stimulate efflux of these oxysterols. Other oxysterols such as 25OH cholesterol can undergo efflux by both ABCA1 and ABCG1 pathways. Moreover, Abcg1/ mice fed the Western diet showed prominent accumulation of 7-KC in macrophages. ABCG1 is also highly expressed in endothelial cells (ECs) (O’Connell et al., 2004). The role of ABCG1 in promoting efflux of 7-KC suggests that large HDL2 particles that promote sterol efflux via ABCG1 may have a particular role in protecting endothelial cells and macrophages from the deleterious effects of oxysterols consumed in the diet or formed on LDL (Terasaka et al., 2007). This could be important in maintaining normal endothelial functions and in atherosclerotic plaque stabilization (Libby et al., 2002). In addition to the severe defect in cholesterol efflux, Abca1/Abcg1/ macrophages showed increased mRNA and secretion of chemokines and inflammatory cytokines (Yvan-Charvet et al., 2007a). The single deficiency of ABCG1 produced a similar though milder defect in secretion of inflammatory cytokines and chemokines, while the deficiency of ABCA1 had smaller effects (Yvan-Charvet et al., 2007a). Abca1/ macrophages show increased tumor necrosis factor (TNF) secretion following treatment with lipopolysaccharide (LPS) (Koseki et al., 2007) and it is possible that macrophage inflammatory responses were mediated by LPS (Yvan-Charvet et al., 2007a). The increased inflammatory gene expression was abolished in macrophages deficient in Toll-like receptor 4 (TLR4) or MyD88/TRIF (Yvan-Charvet et al., 2008). TLR4 cell surface concentration was increased in the order Abca1/ Abcg1/>Abcg1/>Abca1/>wild-type macrophages (Yvan-Charvet et al., 2008). Treatment of transporter-deficient cells with cyclodextrin reduced and cholesterol-cyclodextrin loading increased inflammatory gene expression. Abca1/Abcg1/ bone marrow-derived macrophages showed enhanced inflammatory gene responses to TLR2, TLR3, and TLR4 ligands (YvanCharvet et al., 2008). In Abcg1/ bone marrow-transplanted, Western dietfed Ldlr-deficient mice, there was a profound inflammatory infiltrate in the adventitia and necrotic core region of atherosclerotic lesions, consisting primarily of neutrophils after intraperitoneal injection with thioglycollate (Yvan-Charvet et al., 2008). Other studies have shown that TLR4 mutant mice have decreased atherosclerosis (Michelsen et al., 2004), and that repeated injections of LPS worsen atherosclerosis in mice and rabbits (Lehr et al., 2001; Westerterp et al., 2007). The results suggest that HDL and apoA-I exert anti-inflammatory effects by promoting cholesterol efflux via
Chapter j 10 Sterol Efflux by ABCA1 and ABCG1
207
ABCG1 and ABCA1 with consequent attenuation of signaling via Toll-like receptors. Thus, HDL and ABC transporters may be important in protecting advanced plaques from macrophage apoptosis and inflammatory responses, suggesting a role in plaque stabilization and acute coronary syndromes.
A POTENTIAL ROLE OF HDL AND ABCG1 IN ENDOTHELIAL FUNCTION A part of the atheroprotective effect of HDL may be related to its role in preserving endothelial function (Li et al., 2000; Kuvin et al., 2003). In humans, HDL levels are correlated with flow-mediated vasodilation responses of the brachial artery (Li et al., 2000; Kuvin et al., 2003) and with decreased coronary vasoconstrictor responses (Zeiher et al., 1994). Importantly, infusion of recombinant phospholipid/apoA-I particles into Tangier disease with isolated low HDL levels reversed defective forearm blood flow measurements (Bisoendial et al., 2003). Niacin therapy has been shown to improve NO-mediated vascular relaxation in humans (Kuvin et al., 2002). The beneficial effects of HDL on ECs may include stimulation of proliferation, cell survival, migration, inhibition of the expression of VCAM-1 and ICAM-1 and NO synthesis (Collins and Cybulsky, 2001; O’Connell et al., 2004; Rohrer et al., 2004; Mineo et al., 2006). HDL may have a specific role in reversing decreased eNOS activity in human ECs treated with oxidized LDL (Uittenbogaard et al., 2000) or in reversing the decrease in eNOS dependent vascular relaxation induced by high cholesterol diets (Deckert et al., 1999). Arteries from Abcg1/ and Abca1/Abcg1/ mice fed a high cholesterol diet exhibited a marked decrease in endothelium-dependent vasorelaxation, while Abca1/ mice had a relative milder defect (Terasaka et al., 2008). In addition, eNOS activity was reduced in aortic homogenates of Abcg1/ mice fed either the high cholesterol diet or a Western diet, and this correlated with decreased levels of the active dimeric form of eNOS (Terasaka et al., 2008). More detailed analysis indicated that ABCG1 was expressed primarily in ECs, and that these cells accumulated 7-KC when Abcg1/ mice were fed the Western diet (Terasaka et al., 2008). Consistent with these data, ABCG1 had a major role in promoting efflux of cholesterol and 7-KC to HDL in cultured human aortic ECs (HAECs) (Terasaka et al., 2008). Furthermore, HDL treatment of HAECs prevented 7-KC-induced reactive oxygen species production and active eNOS dimer disruption in an ABCG1-dependent manner (Terasaka et al., 2008). In addition, apoA-I transgene expression increased HDL, improved endothelial function, decreased aortic 7-KC content and increased eNOS activity in high cholesterol-fed Ldlrþ/ mice (Terasaka et al., 2008). The ability of ABCG1 to preserve endothelial function appears to be at least partly related to its role in promoting efflux of 7-oxysterols such as 7-KC to HDL.
208
The HDL Handbook
CONCLUSION AND PERSPECTIVE ABCA1 and ABCG1 sterol efflux pathways show mutual compensation, and deficiency of both transporters results in a dramatic increase in macrophage inflammatory and apoptotic responses and in atherosclerosis (Yvan-Charvet et al., 2007a). Treatments that increase cholesterol efflux via ABCA1 and/or ABCG1 are likely to be beneficial for atherosclerosis, though perhaps with a different spectrum of clinical effects: the ABCG1 pathway might have a particular role in protecting vulnerable plaques as a result of its anti-apoptotic, anti-inflammatory functions and improvement of endothelial function. It is possible that in vivo CETP inhibitors and niacin increase cholesterol efflux via ABCG1, but also, by decreasing catabolism of HDL in the kidney and liver, lead to an increase in apoA-I levels which may be made available to foam cell ABCA1 in the arterial wall. A variety of different approaches to raising HDL are likely to be further developed. Drugs that increase expression of apoA-I in the liver, or infusion of apoA-I or peptides derived from apoA-I that have been optimized for interaction with ABCA1 or LCAT are also likely to be beneficial (Belalcazar et al., 2003; Nissen et al., 2003; Linsel-Nitschke and Tall, 2005; Navab et al., 2006; Rader, 2007). These approaches may also lead to increased formation and levels of HDL particles that can in turn promote efflux of sterols via ABCG1. Therapeutic strategies that upregulate both ABCA1 and ABCG1 transporters themselves may also be beneficial as low levels of the transporters themselves may be rate-limiting for cellular cholesterol efflux in some subjects with coronary heat disease (Trogan et al., 2006). Thus, assuming problems of increased triglycerides can be solved (Repa et al., 2000; Schultz et al., 2000), LXR agonists might be ideal agents for increasing a variety of molecules involved in sterol efflux (Bradley et al., 2007; Scott, 2007).
REFERENCES Adorni, M. P., Zimetti, F., Billheimer, J. T., et al. (2007). The roles of different pathways in the release of cholesterol from macrophages. J Lipid Res, 48, 2453e2462. Aiello, R. J., Brees, D., Bourassa, P. A., et al. (2002). Increased atherosclerosis in hyperlipidemic mice with inactivation of ABCA1 in macrophages. Arterioscler Thromb Vasc Biol, 22, 630e637. Assmann, G., & Gotto, A. M., Jr. (2004). HDL cholesterol and protective factors in atherosclerosis. Circulation, 109(suppl. III), III-8eIII-14. Baldan, A., Pei, L., Lee, R., et al. (2006). Impaired development of atherosclerosis in hyperlipidemic Ldlr/ and ApoE/ mice transplanted with Abcg1/ bone marrow. Arterioscler Thromb Vasc Biol, 26, 2301e2307. Barter, P. J., Nicholls, S., Rye, K. A., Anantharamaiah, G. M., Navab, M., & Fogelman, A. M. (2004). Antiinflammatory properties of HDL. Circ Res, 95, 764e772. Belalcazar, L. M., Merched, A., Carr, B., et al. (2003). Long-term stable expression of human apolipoprotein A-I mediated by helper-dependent adenovirus gene transfer inhibits
Chapter j 10 Sterol Efflux by ABCA1 and ABCG1
209
atherosclerosis progression and remodels atherosclerotic plaques in a mouse model of familial hypercholesterolemia. Circulation, 107, 2726e2732. Bisoendial, R. J., Hovingh, G. K., Levels, J. H., et al. (2003). Restoration of endothelial function by increasing high-density lipoprotein in subjects with isolated low high-density lipoprotein. Circulation, 107, 2944e2948. Bodzioch, M., Orso, E., Klucken, J., et al. (1999). The gene encoding ATP-binding cassette transporter 1 is mutated in Tangier disease. Nat Genet, 22, 347e351. Bradley, M. N., Hong, C., Chen, M., et al. (2007). Ligand activation of LXR beta reverses atherosclerosis and cellular cholesterol overload in mice lacking LXR alpha and apoE. J Clin Invest, 117, 2337e2346. Brooks-Wilson, A., Marcil, M., Clee, S. M., et al. (1999). Mutations in ABC1 in Tangier disease and familial high-density lipoprotein deficiency. Nat Genet, 22, 336e345. Brown, A. J., & Jessup, W. (1999). Oxysterols and atherosclerosis. Atherosclerosis, 142, 1e28. Brown, A. J., Dean, R. T., & Jessup, W. (1996). Free and esterified oxysterol: formation during copper-oxidation of low density lipoprotein and uptake by macrophages. J Lipid Res, 37, 320e335. Buechler, C., Boettcher, A., Bared, S. M., Probst, M. C., & Schmitz, G. (2002). The carboxyterminus of the ATP-binding cassette transporter A1 interacts with a beta2-syntrophin/ utrophin complex. Biochem Biophys Res Commun, 293, 759e765. Castelli, W. P., Garrison, R. J., Wilson, P. W., Abbott, R. D., Kalousdian, S., & Kannel, W. B. (1986). Incidence of coronary heart disease and lipoprotein cholesterol levels. The Framingham Study. J Am Med Assoc, 256, 2835e2838. Cohen, J. C., Kiss, R. S., Pertsemlidis, A., Marcel, Y. L., McPherson, R., & Hobbs, H. H. (2004). Multiple rare alleles contribute to low plasma levels of HDL cholesterol. Science, 305, 869e872. Collins, T., & Cybulsky, M. I. (2001). NF-kappaB: pivotal mediator or innocent bystander in atherogenesis? J Clin Invest, 107, 255e264. Costet, P., Luo, Y., Wang, N., & Tall, A. R. (2000). Sterol-dependent transactivation of the ABC1 promoter by the liver X receptor/retinoid X receptor. J Biol Chem, 275, 28240e28245. Cui, D., Thorp, E., Li, Y., et al. (2007). Pivotal advance: macrophages become resistant to cholesterol-induced death after phagocytosis of apoptotic cells. J Leukoc Biol, 82, 1040e1050. Deckert, V., Lizard, G., Duverger, N., et al. (1999). Impairment of endothelium-dependent arterial relaxation by high-fat feeding in ApoE-deficient mice. Circulation, 100, 1230e1235. Fitzgerald, M. L., Mendez, A. J., Moore, A. J., Andersson, L. P., Panjeton, H. A., & Freeman, M. W. (2001). ATP-binding cassette transporter A1 contains an NH2-terminal signal anchor sequence that translocates the protein’s first hydrophilic domain to the exoplasmic space. J Biol Chem, 276, 15137e15145. Francis, G. A., Knopp, R. H., & Oram, J. F. (1995). Defective removal of cellular cholesterol and phospholipids by apolipoprotein A-I in Tangier disease. J Clin Invest, 96, 78e87. Frikke-Schmidt, R., Nordestgaard, B. G., Jensen, G. B., & Tybjaerg-Hansen, A. (2004). Genetic variation in ABC transporter A1 contributes to HDL cholesterol in the general population. J Clin Invest, 114, 1343e1353. Fu, X., Menke, J. G., Chen, Y., et al. (2001). 27-Hydroxycholesterol is an endogenous ligand for liver X receptor in cholesterol-loaded cells. J Biol Chem, 276, 38378e38387. Gelissen, I. C., Harris, M., Rye, K. A., et al. (2006). ABCA1 and ABCG1 synergize to mediate cholesterol export to apoA-I. Arterioscler Thromb Vasc Biol, 26, 534e540.
210
The HDL Handbook
Gillotte-Taylor, K., Nickel, M., Johnson, W. J., et al. (2002). Effects of enrichment of fibroblasts with unesterified cholesterol on the efflux of cellular lipids to apolipoprotein A-I. J Biol Chem, 277, 11811e11820. Haghpassand, M., Bourassa, P. A., Fancone, O. L., & Aiello, R. J. (2001). Monocyte/macrophage expression of ABCA1 has minimal contribution to plasma HDL levels. J Clin Invest, 108, 1315e1320. Haidar, B., Denis, M., Krimbou, L., Marcil, M., & Genest, J., Jr. (2002). cAMP induces ABCA1 phosphorylation activity and promotes cholesterol efflux from fibroblasts. J Lipid Res, 43, 2087e2094. Haidar, B., Denis, M., Marcil, M., Krimbou, L., & Genest, J., Jr. (2004). ApolipoproteinA-I activates cellular cAMP signaling through the ABCA1 transporter. J Biol Chem, 279, 9963e9969. Joseph, S. B., McKilligin, E., Pei, L., et al. (2002). Synthetic LXR ligand inhibits the development of atherosclerosis in mice. Proc Natl Acad Sci USA, 99, 7604e7609. Kennedy, M. A., Venkateswaran, A., Tarr, P. T., et al. (2001). Characterization of the human ABCG1 gene: liver X receptor activates an internal promoter that produces a novel transcript encoding an alternative form of the protein. J Biol Chem, 276, 39438e39447. Kennedy, M. A., Barrera, G. C., Nakamura, K., et al. (2005). ABCG1 has a critical role in mediating cholesterol efflux to HDL and preventing cellular lipid accumulation. Cell Metab, 1, 121e131. Kobayashi, A., Takanezawa, Y., Hirata, T., et al. (2006). Efflux of sphingomyelin, cholesterol, and phosphatidylcholine by ABCG1. J Lipid Res, 47, 1791e1802. Koseki, M., Hirano, K., Masuda, D., et al. (2007). Increased lipid rafts and accelerated lipopolysaccharide-induced tumor necrosis factor-alpha secretion in Abca1-deficient macrophages. J Lipid Res, 48, 299e306. Kuvin, J. T., Ramet, M. E., Patel, A. R., Pandian, N. G., Mendelsohn, M. E., & Karas, R. H. (2002). A novel mechanism for the beneficial vascular effects of highdensity lipoprotein cholesterol: enhanced vasorelaxation and increased endothelial nitric oxide synthase expression. Am Heart J, 144, 165e172. Kuvin, J. T., Patel, A. R., Sidhu, M., et al. (2003). Relation between high-density lipoprotein cholesterol and peripheral vasomotor function. Am J Cardiol, 92, 275e279. Lange, Y., Dolde, J., & Steck, T. L. (1981). The rate of transmembrane movement of cholesterol in the human erythrocyte. J Biol Chem, 256, 5321e5323. Lehr, H. A., Sagban, T. A., Ihling, C., et al. (2001). Immunopathogenesis of atherosclerosis: endotoxin accelerates atherosclerosis in rabbits on hypercholesterolemic diet. Circulation, 104, 914e920. Li, X. P., Zhao, S. P., Zhang, X. Y., Liu, L., Gao, M., & Zhou, Q. C. (2000). Protective effect of high density lipoprotein on endothelium-dependent vasodilatation. Int J Cardiol, 73, 231e236. Libby, P., Ridker, P. M., & Maseri, A. (2002). Inflammation and atherosclerosis. Circulation, 105, 1135e1143. Linsel-Nitschke, P., & Tall, A. R. (2005). HDL as a target in the treatment of atherosclerotic cardiovascular disease. Nat Rev Drug Discov, 4, 193e205. Martinez, L. O., Agerholm-Larsen, B., Wang, N., Chen, W., & Tall, A. R. (2003). Phosphorylation of a pest sequence in ABCA1 promotes calpain degradation and is reversed by ApoA-I. J Biol Chem, 278, 37368e37374.
Chapter j 10 Sterol Efflux by ABCA1 and ABCG1
211
Matsuura, F., Wang, N., Chen, W., Jiang, X. C., & Tall, A. R. (2006). HDL from CETP-deficient subjects shows enhanced ability to promote cholesterol efflux from macrophages in an apoEand ABCG1-dependent pathway. J Clin Invest, 116, 1435e1442. Michelsen, K. S., Wong, M. H., Shah, P. K., et al. (2004). Lack of Toll-like receptor 4 or myeloid differentiation factor 88 reduces atherosclerosis and alters plaque phenotype in mice deficient in apolipoprotein E. Proc Natl Acad Sci USA, 101, 10679e10684. Mineo, C., Deguchi, H., Griffin, J. H., & Shaul, P. W. (2006). Endothelial and antithrombotic actions of HDL. Circ, Res, 98, 1352e1364. Munehira, Y., Ohnishi, T., Kawamoto, S., et al. (2004). Alpha1-syntrophin modulates turnover of ABCA1. J Biol Chem, 279, 15091e15095. Nakamura, K., Kennedy, M. A., Baldan, A., Bojanic, D. D., Lyons, K., & Edwards, P. A. (2004). Expression and regulation of multiple murine ATP-binding cassette transporter G1 mRNAs/ isoforms that stimulate cellular cholesterol efflux to high density lipoprotein. J Biol Chem, 279, 45980e45989. Navab, M., Anantharamaiah, G. M., Reddy, S. T., & Fogelman, A. M. (2006). Apolipoprotein A-I mimetic peptides and their role in atherosclerosis prevention. Nat Clin Pract Cardiovasc Med, 3, 540e547. Nissen, S. E., Tsunoda, T., Tuzcu, E. M., et al. (2003). Effect of recombinant ApoA-I Milano on coronary atherosclerosis in patients with acute coronary syndromes: a randomized controlled trial. J Am Med Assoc, 290, 2292e2300. O’Connell, B. J., Denis, M., & Genest, J. (2004). Cellular physiology of cholesterol efflux in vascular endothelial cells. Circulation, 110, 2881e2888. Okuhira, K., Fitzgerald, M. L., Sarracino, D. A., et al. (2005). Purification of ATP-binding cassette transporter A1 and associated binding proteins reveals the importance of beta1-syntrophin in cholesterol efflux. J Biol Chem, 280, 39653e39664. Oram, J. F., Lawn, R. M., Garvin, M. R., & Wade, D. P. (2000). ABCA1 is the cAMP inducible apolipoprotein receptor that mediates cholesterol secretion from macrophages. J Biol Chem, 275, 34508e34511. Ou, J., Tu, H., Shan, B., et al. (2001). Unsaturated fatty acids inhibit transcription of the sterol regulatory element-binding protein-1c (SREBP-1c) gene by antagonizing ligand-dependent activation of the LXR. Proc Natl Acad Sci USA, 98, 6027e6032. Out, R., Hoekstra, M., Habets, K., et al. (2007). Combined deletion of macrophage ABCA1 and ABCG1 leads to massive lipid accumulation in tissue macrophages and distinct atherosclerosis at relatively low plasma cholesterol levels. Arterioscler Thromb Vasc Biol, 28, 258e264. Out, R., Jessup, W., Le Goff, W., et al. (2007a). Coexistence of foam cells and hypocholesterolemia in mice lacking the ABC transporters A1 and G1. Circ Res, 102, 113e120. Rader, D. J. (2006). Molecular regulation of HDL metabolism and function: implications for novel therapies. J Clin Invest, 116, 3090e3100. Rader, D. J. (2007). Mechanisms of disease: HDL metabolism as a target for novel therapies. Nat Clin Pract, 4, 102e109. Ranalletta, M., Wang, N., Han, S., Yvan-Charvet, L., Welch, C., & Tall, A. R. (2006). Decreased atherosclerosis in low-density lipoprotein receptor knockout mice transplanted with Abcg1/ bone marrow. Arterioscler Thromb Vasc Biol, 26, 2308e2315. Repa, J. J., Liang, G., Ou, J., et al. (2000). Regulation of mouse sterol regulatory element-binding protein-1c gene (SREBP-1c) by oxysterol receptors, LXRalpha and LXRbeta. Genes Dev, 14, 2819e2830.
212
The HDL Handbook
Rhoads, G. G., Gulbrandsen, C. L., & Kagan, A. (1976). Serum lipoproteins and coronary heart disease in a population study of Hawaii Japanese men. N Engl J Med, 294, 293e298. Rohrer, L., Hersberger, M., & von Eckardstein, A. (2004). High density lipoproteins in the intersection of diabetes mellitus, inflammation and cardiovascular disease. Curr Opin Lipidol, 15, 269e278. Roosbeek, S., Peelman, F., Verhee, A., et al. (2004). Phosphorylation by protein kinase CK2 modulates the activity of the ATP binding cassette A1 transporter. J Biol Chem, 279, 37779e37788. Rust, S., Rosier, M., Funke, H., et al. (1999). Tangier disease is caused by mutations in the gene encoding ATP-binding cassette transporter. Nat Genet, 22, 352e355. Santamarina-Fojo, S., Peterson, K., Knapper, C., et al. (2000). Complete genomic sequence of the human ABCA1 gene: analysis of the human and mouse ATP-binding cassette A promoter. Proc Natl Acad Sci USA, 97, 7987e7992. Schultz, J. R., Tu, H., Luk, A., et al. (2000). Role of LXRs in control of lipogenesis. Genes Dev, 14, 2831e2838. Scott, J. (2007). The liver X receptor and atherosclerosis. N Engl J Med, 357, 2195e2197. Singaraja, R. R., Fievet, C., Castro, G., et al. (2002). Increased ABCA1 activity protects against atherosclerosis. J Clin Invest, 110, 35e42. Small, D. M. (2003). Role of ABC transporters in secretion of cholesterol from liver into bile. Proc Natl Acad Sci USA, 100, 4e6. Smith, J. D., Le Goff, W., Settle, M., et al. (2004). ABCA1 mediates concurrent cholesterol and phospholipid efflux to apolipoprotein A-I. J Lipid Res, 45, 635e644. Tabas, I. (2005). Consequences and therapeutic implications of macrophage apoptosis in atherosclerosis: the importance of lesion stage and phagocytic efficiency. Arterioscler Thromb Vasc Biol, 25, 2255e2264. Tall, A. R. (2003). Role of ABCA1 in cellular cholesterol efflux and reverse cholesterol transport. Arterioscler Thromb Vasc Biol, 23, 710e711. Tall, A. R., Costet, P., & Wang, N. (2002). Regulation and mechanisms of macrophage cholesterol efflux. J Clin Invest, 110, 899e904. Tang, C., Vaughan, A. M., & Oram, J. F. (2004). Janus kinase 2 modulates the apolipoprotein interactions with ABCA1 required for removing cellular cholesterol. J Biol Chem, 279, 7622e7628. Terasaka, N., Wang, N., Yvan-Charvet, L., & Tall, A. R. (2007). High-density lipoprotein protects macrophages from oxidized low-density lipoprotein-induced apoptosis by promoting efflux of 7-ketocholesterol via ABCG1. Proc Natl Acad Sci USA, 104, 15093e15098. Terasaka, N., Hiroshima, A., Koieyama, T., et al. (2003). T-0901317, a synthetic liver X receptor ligand, inhibits development of atherosclerosis in LDL receptor-deficient mice. FEBS Lett, 536, 6e11. Terasaka, N., Yu, S., Yvan-Charvet, L., et al. (2008). ABCG1 and HDL protect against endothelial dysfunction in mice fed a high-cholesterol diet. J Clin Invest, 118, 3701e3713. Timmins, J. M., Lee, J. Y., Boudyguina, E., et al. (2005). Targeted inactivation of hepatic Abca1 causes profound hypoalphalipoproteinemia and kidney hypercatabolism of apoA-I. J Clin Invest, 115, 1333e1342. Tontonoz, P., & Mangelsdorf, D. J. (2003). Liver X receptor signaling pathways in cardiovascular disease. Mol Endocrinol, 17, 985e993. Trogan, E., Feig, J. E., Dogan, S., et al. (2006). Gene expression changes in foam cells and the role of chemokine receptor CCR7 during atherosclerosis regression in ApoE-deficient mice. Proc Natl Acad Sci USA, 103, 3781e3786.
Chapter j 10 Sterol Efflux by ABCA1 and ABCG1
213
Tsukamoto, T., Hirano, K., Tsujii, K., et al. (2001). ATPbinding cassette transporter-1 induces rearrangement of actin cytoskeletons possibly through Cdc42/N-WASP. Biochem Biophys Res Commun, 287, 757e765. Uittenbogaard, A., Shaul, P. W., Yuhanna, I. S., Blair, A., & Smart, E. J. (2000). High density lipoprotein prevents oxidized low density lipoprotein-induced inhibition of endothelial nitricoxide synthase localization and activation in caveolae. J Biol Chem, 275, 11278e11283. Vaisman, B. L., Lambert, G., Amar, M., et al. (2001). ABCA1 overexpression leads to hyperalphalipoproteinemia and increased biliary cholesterol excretion in transgenic mice. J Clin Invest, 108, 303e309. van Eck, M., Bos, I. S., Kaminski, W. E., et al. (2002). Leukocyte ABCA1 controls susceptibility to atherosclerosis and macrophage recruitment into tissues. Proc Natl Acad Sci USA, 99, 6298e6303. Vaughan, A. M., & Oram, J. F. (2005). ABCG1 redistributes cell cholesterol to domains removable by HDL but not by lipid-depleted apolipoproteins. J Biol Chem, 280, 30150e30157. Vedhachalam, C., Duong, P. T., Nickel, M., et al. (2007). Mechanism of ATP-binding cassette transporter A1-mediated cellular lipid efflux to apolipoprotein A-I and formation of high density lipoprotein particles. J Biol Chem, 282, 25123e25130. Vine, D. F., Mamo, C. L., Beilin, L. J., Mori, T. A., & Croft, K. D. (1998). Dietary oxysterols are incorporated in plasma triglyceride-rich lipoproteins, increase their susceptibility to oxidation and increase aortic cholesterol concentration of rabbits. J Lipid Res, 39, 1995e2004. Wang, Y., & Oram, J. F. (2002). Unsaturated fatty acids inhibit cholesterol efflux from macrophages by increasing degradation of ATP-binding cassette transporter A1. J Biol Chem, 277, 5692e5697. Wang, Y., & Oram, J. F. (2005). Unsaturated fatty acids phosphorylate and destabilize ABCA1 through a phospholipase D2 pathway. J Biol Chem, 280, 35896e35903. Wang, N., Silver, D. L., Costet, P., & Tall, A. R. (2000). Specific binding of ApoA-I, enhanced cholesterol efflux, and altered plasma membrane morphology in cells expressing ABC1. J Biol Chem, 275, 33053e33058. Wang, N., Lan, D., Chen, W., Matsuura, F., & Tall, A. R. (2004). ATP-binding cassette transporters G1 and G4 mediate cellular cholesterol efflux to high-density lipoproteins. Proc Natl Acad Sci USA, 101, 9774e9779. Wang, N., Ranalletta, M., Matsuura, F., Peng, F., & Tall, A. R. (2006). LXR induced redistribution of ABCG1 to plasma membrane in macrophages enhances cholesterol mass efflux to HDL. Arterioscler Thromb Vasc Biol, 26, 1310e1316. Westerterp, M., Berbee, J. F., Pires, N. M., et al. (2007). Apolipoprotein C-I is crucially involved in lipopolysaccharide-induced atherosclerosis development in apolipoprotein E-knockout mice. Circulation, 116, 2173e2181. Yancey, P. G., Bortnick, A. E., Kellner-Weibel, G., de la Llera-Moya, M., Phillips, M. C., & Rothblat, G. H. (2003). Importance of different pathways of cellular cholesterol efflux. Arterioscler Thromb Vasc Biol, 23, 712e719. Yvan-Charvet, L., Ranalletta, M., Wang, N., et al. (2007a). Combined deficiency of ABCA1 and ABCG1 promotes foam cell accumulation and accelerates atherosclerosis in mice. J Clin Invest, 117, 3900e3908. Yvan-Charvet, L., Matsuura, F., Wang, N., et al. (2007b). Inhibition of cholesteryl ester transfer protein by torcetrapib modestly increases macrophage cholesterol efflux to HDL. Arterioscler Thromb Vasc Biol, 27, 1132e1138. Yvan-Charvet, L., Pagler, T. A., Wang, N., et al. (2008). SR-BI inhibits ABCG1-stimulated net cholesterol efflux from cells to plasma HDL. J Lipid Res, 49, 107e114.
214
The HDL Handbook
Yvan-Charvet, L., Welch, C., Pagler, T. A., et al. (2008). Increased inflammatory gene expression in ABC transporter-deficient macrophages: free cholesterol accumulation, increased signaling via toll-like receptors, and neutrophil infiltration of atherosclerotic lesions. Circulation, 118, 1837e1847. Zeiher, A. M., Schachlinger, V., Hohnloser, S. H., Saurbier, B., & Just, H. (1994). Coronary atherosclerotic wall thickening and vascular reactivity in humans. Elevated high-density lipoprotein levels ameliorate abnormal vasoconstriction in early atherosclerosis. Circulation, 89, 2525e2532.
Chapter 11
Functional Change in the HDL Particle by Oxidative Modification and its Contribution to Atherogenesis Toshiyuki Matsunaga 1, Akira Hara 1 and Tsugikazu Komoda 2 1 Laboratory of Biochemistry, Gifu Pharmaceutical University, 2 Department of Biochemistry, Saitama Medical University, Japan
INTRODUCTION In epidemiological studies, low concentrations of high density lipoprotein (HDL) have been shown to inversely correlate with the risk of atherosclerosis and the associated cardiovascular diseases (Reichl and Miller, 1989; Rader, 2003; Assmann and Gotto, 2004). HDL is involved in the transportation of cholesterol from extrahepatic tissues to the liver. In the system of so-called reverse cholesterol transport, HDL can also enhance the cholesterol efflux from macrophage-derived foam cells (Brown et al., 1980). Therefore, the system is considered to be a major anti-atherogenic function of HDL (Steinberg, 1978; Tall, 1990; Johnson et al., 1991; Rader, 2003; Assmann and Gotto, 2004). Recent studies have, in addition, found that HDL has other abilities to protect from atherogenic events. Those beneficial abilities of HDL are to inhibit low density lipoprotein (LDL) oxidation, vascular cell apoptosis and endothelial dysfunction (O’Connell and Genest, 2001; Rohrer et al., 2004). In order to protect from the endothelial dysfunction observed at the onset and development of atherosclerosis, HDL appears to exert a variety of actions including enhanced production of nitric oxide (NO) and downregulation of adhesion molecules (Collins and Cybulsky, 2001; O’Connell and Genest, 2001; Rohrer et al., 2004). Based on multiple investigations, HDL has newly been reported to possess antiviral activity (Singh and Baron, 2000; Van Lenten et al., 2001), anti-thrombotic effects (Saku et al., 1985) and prostacyclin stabilizing activity (Fleisher et al., 1982; Yui et al., 1988). The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
215
216
The HDL Handbook
HDL contains antioxidant molecules and enzymes in the particle; nevertheless, it is oxidized more rapidly than LDL during in vitro oxidation (Nakajima et al., 1995; Hurtado et al., 1996). Ohmura et al. (1999) demonstrated that HDL from patients with coronary artery spasm has a higher susceptibility to lipid peroxidative modification than LDL. This suggests a possibility that oxidized lipoproteins, including both HDL and LDL, are involved in the genesis of coronary artery diseases and reflect an oxidative stress state of the diseases. An immunohistochemical study by Vollmer et al. (1991) also revealed that apoprotein A-I (apoA-I) is present in lumen-adjacent layers of the intima in the earlier stage and in deeper layers of vascular wall of patients with the advanced stage of the diseases. Within the intimal layer, apoproteins in HDL are detected either in an intracellular (mainly in foam cells) or extracellular location, depending on the stage of atherosclerosis. The altered localization of apoproteins in the lesion suggests the contribution of oxidized HDL (ox-HDL) to each step in the atherogenic processes. Greilberger and Ju¨rgens (1998) showed that mild oxidation enhances the ability of HDL to strongly compete with binding of ox-LDL to type I and III collagens, whereas native and highly oxidized HDL are weak competitors. This may imply that longer retention in the atheromatous lesion enables ox-HDL to participate in atherogenic events, because an increased binding of ox-LDL to type I, III and IV collagens is thought to lead to the progression of atherosclerosis (Jimi et al., 1994). Our recent analyses using ox-HDL-specific antibody have shown that ox-HDL is present in the intima of atheromatous plaques in the human abdominal aorta (Nakajima et al., 2000), and its level is increased in patients with coronary artery diseases, type-II diabetes mellitus and chronic renal failure (Tsumura et al., 2001; Nakajima et al., 2004). Thus, the oxidative modification of HDL could occur in patients with a variety of vascular diseases. In addition to verification of the in vivo presence of ox-HDL, recent studies using different oxidants have proposed that oxidative modification of HDL raises multiple alterations of its components and conformation (Nagano et al., 1991; Salmon et al., 1992a; Morel, 1994; Bonnefont-Rousselot et al., 1995). In this chapter, we review the current literature on the structural and functional alterations in HDL by oxidation and discuss the clinical relevance of ox-HDL for atherogenic events.
STRUCTURAL ALTERATIONS IN HDL COMPONENTS BY OXIDATION Oxidation of HDL HDL and its lipidic components are oxidized rapidly and preferably, compared with LDL during in vitro oxidation (Bowry et al., 1992; Nakajima et al., 1995; Hurtado et al., 1996). In HDL subclasses, HDL2 is more susceptible to Cu2þ-mediated oxidation than HDL3 (Stojanovic et al., 2006). Accordingly,
Chapter j 11 Function of oxidized high-density lipoprotein
217
determination of oxidized lipids in HDL2 particles might be a useful marker for the earlier stages of atherogenetic events, while in vivo oxidation of HDL2 lipids has been little investigated. HDL can readily diffuse into the subendothelial fraction because its particle size is smaller than that of LDL (Sloop et al., 1987; Bjornheden et al., 1996). The long retention in the arterial intima might result in an oxidation of HDL, dependent on enhanced expression of oxidant enzymes, and lack of free radical scavengers and reducing agents. Immunohistochemical study using a monoclonal antibody against ox-HDL has shown that ox-HDL is localized in the surface of the endothelial layer and arterial intima of atheromatous plaques of the human abdominal aorta, in contrast to the localization of ox-LDL in the whole area of the plaques (Nakajima et al., 2000). This different localization between ox-HDL and oxLDL may be of crucial importance in order to clarify the role of oxidized lipoprotein in atherogenic events. In previous investigations, numerous potential mechanisms for in vivo oxidation of lipoproteins were proposed (Berliner and Heinecke, 1996). Possible in vivo mechanisms of LDL oxidation are well-documented, and include the myeloperoxidase pathway (Carr et al., 2000; Gaut et al., 2002; Zhang et al., 2002; Brennan and Hazen, 2003; Brennan et al., 2003), the 12/15-lipoxygenase pathway (Harats et al., 2000; Cyrus et al., 2001; George et al., 2001; Zhao et al., 2002), the 5-lipoxygenase pathway (Mehrabian et al., 2002; Mehrabian and Allayee, 2003; Dwyer et al., 2005), the cyclooxygenase pathway (Linton and Fazio, 2004; Natarajan and Nadler, 2004; Vila, 2004), and the NADPH oxidase pathway (Sorescu et al., 2001; Cathcart, 2004). Although there is little evidence for how ox-HDL is formed, Nagano et al. (1991) suggested that HDL is oxidized in the same manner as LDL. Myeloperoxidase, an enzyme in macrophages and neutrophils, may be one of the most potential enzymes concerning in vivo oxidant generation. This enzyme system generates reactive species including hypochlorous acid/ hypochlorite (HOCl/OCl), tyrosyl radicals, chloramines, and nitrogen dioxides (NO2) (Weiss et al., 1982; Kettle et al., 1997). Recent investigations have provided direct evidence for activation of the enzyme in human atherosclerotic lesions (Daugherty et al., 1994), and increased levels of hypochlorite-modified proteins and dityrosines in early and advanced lesions of atherosclerosis (Heinecke et al., 1993; Hazell et al., 1996; Hazen and Heinecke, 1997; Leeuwenburgh et al., 1997a). Lipoxygenase, which is produced by endothelial cells and macrophages, may also participate in the in vivo oxidation of lipoproteins, because treatment with the enzyme inhibitors prevents cell-mediated oxidation of LDL (Parthasarathy et al., 1989; Rankin et al., 1991). The oxidative mechanism of lipoproteins by lipoxygenase is mediated by conversion of polyunsaturated fatty acids into lipid peroxides including 13S-hydroxy-9,11-octadecadienoic acid, and the metabolites are detected in the early stage of atherosclerosis, accordingly suggesting the positive relationship of the enzyme with atherogenesis
218
The HDL Handbook
(Kuhn et al., 1997). Cyrus et al. (1999) have also reported that the inhibition of 12/15-lipoxygenase diminishes atherosclerotic lesions in apoE-deficient mice. Peroxynitrite, a potent oxidant biologically produced by the interaction of nitric oxide with superoxide anion, is thought to be relevant to cardiovascular diseases. Peroxynitrite can directly oxidize polyunsaturated fatty acids, tocopherols, carbohydrates, DNA and proteins (Christen et al., 1997; Pannala et al., 1998; Ducrocq et al., 1999). Peroxynitrite-induced nitration generates high amounts of 3-nitrotyrosine in proteins (Ischiropoulos et al., 1992; Matsunaga et al., 2001). A high level of 3-nitrotyrosine in the LDL isolated from human atherosclerotic intima suggests that 3-nitrotyrosine-containing LDL promotes atherogenesis by counteracting the anti-atherogenic effect of NO (Leeuwenburgh et al., 1997b). Interestingly, the content of 3-nitrotyrosine in HDL isolated from human aortic atherosclerotic intima is six-fold higher than that in circulating HDL (Pennathur et al., 2004). Thus, the formation of peroxynitrite during atherogenic events may also promote oxidative modification of HDL. Tocopheroxy radical (Neuzil et al., 1997), metal ions (Parthasarathy et al., 1990; Reyftmann et al., 1991; Nagano et al., 1991) and ultraviolet irradiation (Alomar et al., 1992; Salmon et al., 1992b) are also proposed as oxidants for HDL, although in vivo oxidation mechanisms have not been established. It is recognized that LDL from patients with hyperglycemia is more susceptible to oxidation than that from healthy volunteers, and that, in the LDL particles, the oxidative rate is positively related with the rate of glycation (Bowie et al., 1993). This oxidative modification of LDL is presumably due to oxygen radicals derived from amadori products, intermediates of non-enzymatic glycosylation (i.e., Maillard reaction) (Yim et al., 2000). Previous clinical investigation detected high amounts of ox-HDL in sera from subjects with type II diabetes as described above (Nakajima et al., 2004). In addition, other studies demonstrated that exposure to glucose makes HDL more susceptible to oxidation, and the oxidative inhibition of paraoxonase (PON) is responsible for the resultant HDL dysfunction such as a loss of antioxidative capacity (Hedrick et al., 2000; Matsunaga et al., 2003a; Zhou et al., 2008). These findings indicate a possibility that ox-HDL is involved in accelerating diabetic complications, besides ox-LDL, although mechanisms underlying the HDL oxidation and PON inactivation during the incubation with glucose are not fully understood. Molecular mechanisms of oxidation of the HDL particle appear to be different among oxidants. In the system including metal ions and lipoxygenase, oxidants primarily attack lipidic components, especially phospholipids and cholesterols, in the HDL particle and subsequent oxidation of apoproteins is mediated by the oxidized lipids (Marcel et al., 1989; Morel, 1994; Garner et al., 1998a). In contrast, a low dose of tyrosyl radical, which is generated by the myeloperoxidase system, oxidizes lipid-free apoproteins in HDL (Wang et al., 1998), and the apoproteins in HDL are more highly oxidized than the lipidic components (Panzenboeck et al., 1997; Wang et al., 1998). Possible
Chapter j 11 Function of oxidized high-density lipoprotein
219
HDL Oxidation myeloperoxidase, lipoxygenase, cyclooxygenase, peroxynitrite, tocopheroxy radical, metals and ultraviolet irradiation
ox-HDL
Structural alterations in ox-HDL Oxidation of amino acids (Lys, Trp and Met) in apoproteins Cross-linking of apoproteins Conversion of phosphatidylcholine (PC) into the lyso form Reduction in polyunsaturated fatty acids in PC Increase in hydroperoxide forms of PC and cholesterol Alteration of fluidity at surface of particle Inactivation of antioxidant enzymes (paraoxonase, lecithin: cholesterol acyltransferase and so on) Depletion of antioxidant molecules
SCHEME 11.1 Possible mechanisms for in vivo oxidation of HDL and structural alterations in ox-HDL.
mechanisms for in vivo oxidation of HDL and structural alterations previously found in ox-HDL are shown in Scheme 11.1.
Alteration in apoproteins by oxidation Cu2þ-mediated oxidation increases the net negative charge in HDL particles. This is speculated to be due to modification of positively charged residues such as Lys in apoproteins (Nagano et al., 1991). The modification in residues of Lys and Trp during oxidation appears to be involved in the reduction of the binding affinity of HDL for the receptor and in the reduced activity of the particle to remove cholesterol from the cells (Nagano et al., 1991; Reyftmann et al., 1991). Indeed, Duell et al. (1991) demonstrated the importance of the Lys residues in the HDL receptor-dependent efflux of intracellular cholesterol. In addition, mild oxidation of HDL by 2,2’-azo-bis(2-amidinopropane) dihydrochloride (AAPH), a generator of aqueous peroxyl radical, is capable of converting Met residues of HDL apoproteins into Met sulfoxides (MetO) (Garner et al.,
220
The HDL Handbook
1998a,b; Pankhurst et al., 2003). Lipid hydroperoxides formed during the AAPH-induced oxidation of HDL convert two Met residues (Met86 and Met112) in apoA-I molecules to MetO in a step-wise manner, while Met148 is not oxidized. In the case of homodimeric apoA-II, two Met26 residues of the subunits are oxidized to MetO. These studies also suggest the clinical importance of apoA-I containing MetO as a new marker for coronary artery diseases because there is a relationship between its circulating concentration and the risk for the vascular diseases. It should be noted that the high molar ratio of hypochlorite to HDL induces oxidative modification of amino acids other than Met (Panzenboeck et al., 1997), although at low ratios of the oxidant/HDL (<6:1) the oxidation is limited to MetO formation (Bergt et al., 2000). Apoproteinelipid association in Cu2þ-oxidized HDL is more stable than that in native HDL, which may be due to oxidant-induced cross-linking of apoproteins (Shoukry et al., 1994). Cross-linking between apoA-I and apoA-II was previously detected in HDL modified by other oxidants such as hypochlorite (Marsche et al., 2002), peroxynitrite (Ahmed et al., 2001; Matsunaga et al., 2001), and aldehydes (e.g., malondialdehyde, acrolein and 4-hydroxynonenal) (Salmon et al., 1992a; McCall et al., 1995). On SDS-PAGE analysis of apoproteins in HDL oxidized by these oxidants, typical bands of apoA-I dimer, apoA-I trimer and apoAI-(apoA-II)2 heterotrimer are detectable in addition to unoxidized monomeric apoA-I. As shown in Figure 11.1, chemically synthesized peroxynitrite also shows a similar pattern of heterogeneous
A
CuSO4 (µ µM) M
0
0.1 0.3 1
3 10 30
B
Peroxynitrite (mM) M
0 0.1 0.3 1
3
(kDa) 97 66
(A-I)3 (A-I)2
45 (A-I)-(AII)2 31 A-I 21 FIGURE 11.1 Cross-linking of apoproteins in Cu2D- and peroxynitrite-oxidized HDLs. (A) Cu2þ- and (B) peroxynitrite-oxidized HDLs were prepared by incubation at 37 C at the indicated concentrations of CuSO4 and chemically synthesized peroxynitrite, respectively. Twenty micrograms of each sample was applied to SDS-polyamide gel electrophoresis (12.5% slab gel) under reducing conditions, and then stained with Coomassie brilliant blue R-250. Molecular weight markers (M) are indicated in kDa on the left of the panels. The expected forms of apoproteins are confirmed by Western blotting using monoclonal antibodies against apoA-I and apoA-II, and indicated on the right of the panels.
Chapter j 11 Function of oxidized high-density lipoprotein
221
cross-linking of apoA-I and apoA-II. Lipid peroxides and their aldehydic metabolites are considered to be responsible for the formation of these aggregates of the apoproteins in the oxidation system of AAPH (Garner et al., 1998a). This implies that the active metabolites of lipid peroxides act as crosslinkers of apoproteins. Compared with other oxidants, oxidation induced by the tyrosyl radical can form oligomeric apoproteins without the involvement of lipid components (Wang et al., 1998). Francis (2000) suggests that the cross-linking by the tyrosyl radical is mediated by dityrosine formation that is a covalent interaction of the free tyrosyl radical with the protein tyrosyl radical. It has been postulated that hinged regions of the amphipathic helix of apoproteins in HDL are involved in the interaction with the cell membrane (Rothblat et al., 1992). Sakai et al. (1992) demonstrated that the significant reduction of the ligand activity of HDL toward the HDL receptor upon Cu2þmediated oxidation results from cross-linking of HDL apoproteins, particularly apoA-I. Other studies also indicated that the cross-linking of HDL apoproteins by treatment with tetranitromethane (Chacko, 1985; Brinton et al., 1986) and chemical cross-linkers (Chacko et al., 1988; Miyazaki et al., 1991) leads to the loss of ligand activity of HDL. Moreover, it must be noted that the affinity of HDL for its binding site has been shown to be markedly decreased by crosslinking of apoA-I (Zhang et al., 1989). Thus, it is likely that cross-linking of apoproteins alters such segments of the amphipathic helix and consequently affects the affinity of HDL for its receptor on the cell surface. Cu2þ-mediated oxidation of HDL reduces its capacity to enhance cholesterol efflux from macrophage foam cells (Nagano et al., 1991) and fibroblasts (Salmon et al., 1992a), and its affinity for a specific receptor (Sakai et al., 1992). Ox-HDL are recognized and endocytosed by scavenger receptors for oxLDL on the macrophages (Steinberg et al., 1989; La Ville et al., 1994). On the other hand, it is reported that ox-HDL does not apparently behave as a ligand for the scavenger receptor (Parthasarathy et al., 1990). Recently, CD36 has been reported to be a receptor for ox-HDL, but not for native HDL and LDL (Thorne et al., 2007). We showed that ox-HDL reduces binding affinity for the HDL receptor and instead enhances the affinity for a 130-kDa protein, distinct from the native HDL receptor on the surface of human endothelial cells, suggesting that a novel receptor for ox-HDL is present on human endothelial cells (Nakajima et al., 2000). In addition, the lethal effects of ox-HDL on cultured human endothelial cells were markedly suppressed by concomitant incubation with an antibody against lectin-like ox-LDL receptor 1, which is involved in initiation and progression of atherosclerotic plaques (Matsunaga et al., 2003b; Vohra et al., 2006). These apparently conflicting observations concerning oxHDL receptor(s) may depend on the type of oxidants used experimentally and the degree of oxidative modification of HDL. Thus, the receptor specific for ox-HDL and its underlying molecular mechanism for binding to the ligand have not been clarified yet.
222
The HDL Handbook
Alteration in lipids by oxidation Oxidation of lipidic components has been detected in each particle of HDL and LDL separated from human atherosclerotic lesions, and the amount of the oxidized lipids is positively correlated with the severity of the disease (Frei et al., 1988). Exposure to aqueous peroxyl radicals preferably oxidizes HDL lipids rather than those in LDL (Bowry et al., 1992; Hurtado et al., 1996). Characterization of the lipid moiety in native HDL revealed that phosphatidylcholine (PC, 1,2-diacyl-sn-glycerol-3-phosphorylcholine) is a major phospholipid occupying up to 90% of total phospholipids, whereas the remaining species are composed of phosphatidylinositol, phosphatidic acid, phosphatidylserine, and phosphatidylethanolamine. In general, PC plays an important role in the functional properties of the cell membrane of all living organisms, and is metabolized into lysophosphatidylcholine (lyso-PC) and polyunsaturated fatty acid by phospholipase A2. HDL oxidation decreases the PC content and, in parallel, increases that of lyso-PC in the lipidic component of the particle, in addition to the disappearance of phosphatidylethanolamine (Bradamante et al., 1992; Morel, 1994; Nakajima et al., 1995; Ahmed et al., 2001). The similar enhanced formation of lyso-PC is also observed in HDL treated with peroxynitrite donor and AAPH (Garner et al., 1998a; Ahmed et al., 2001). In contrast, it seems that the oxidation of HDL by increasing HOCl concentrations solely results in a slight decrease of PC (Marsche et al., 2002). These pieces of evidence suggest that increased formation of lyso-PC is one of the most reasonable indicators observed in HDL particles during oxidation. It has recently been detected that Cu2þ-mediated oxidation of the HDL particle markedly reduces the amounts of polyunsaturated fatty acids such as linoleic acid (18:2), arachidonic acid (20:4), docosahexaenoic acid (22:6) and linolenic acid (20:3) in esterified fatty acid, whereas the oxidation is unable to alter the contents of saturated and monounsaturated fatty acids (Nakajima, 1998). These data may imply that two-electron oxidation by incubation with Cu2þ specifically attacks polyunsaturated fatty acids at the site of sn-2 in PC. It should be noted that PC hydroperoxides are formed as reactive metabolites of PC during oxidation of HDL (Mashima et al., 1998). Compositional study showed that HDL oxidation alters the moiety of cholesteryl ester and its unesterified form through reduction of the ability to esterify cholesterols, and these alterations might affect its role in reverse cholesterol transport (Musanti and Ghiselli, 1993; Wang et al., 1998). The cholesterols in the HDL particle are readily oxidized by incubation with oxidant molecules, and then converted into cholesterol hydroperoxides and its hydroxides, like phospholipids (Garner et al., 1998a,b). Nuclear magnetic resonance spectroscopy revealed that several kinds of oxidized derivatives of cholesterol, which is located in an inner fraction of HDL particle, are identified in the Cu2þ-oxidized HDL (Bradamante et al., 1992).
Chapter j 11 Function of oxidized high-density lipoprotein
223
Although the formation of reactive metabolites containing lipid aldehydes in the HDL particle is not understood yet, the metabolites are formed from different lipid hydroperoxides in LDL by action of trace metals or vitamin C (Requena et al., 1997; Oe et al., 2003). Increased levels of proteins modified by the reactive lipid metabolites are detected in the plasma of patients with atherosclerosis (Palinski et al., 1990; Holvoet et al., 1995). One of the reactive metabolites, 4-hydroxy-2-noneal (HNE) is a bifunctional electrophile and is generated by undergoing alkoxyl radical formation and its b-scission (Esterbauer et al., 1991). HNE can form Michael adducts with side chains of Lys, His and Cys, and pyrrole adducts with Lys in proteins (Uchida et al., 1994; O’Neil et al., 1997). The modification of LDL with HNE appears to be responsible for enhanced recognition by scavenger receptors on macrophages, because the binding affinities for the LDL receptor and the scavenger receptor are critically dependent on free and modified forms of Lys residues in LDL, respectively (Hoff et al., 1989; Steinbrecher et al., 1989; Esterbauer et al., 1992). In addition, it is suggested that HNE participates in the mechanism of structural modification including cross-linking in ox-LDL (Requena et al., 1997). Therefore, HNE might play an important role in oxidant-induced molecular alterations in the HDL particle, if HNE formation can occur in vivo during the oxidative process of HDL. It is well known that each lipoprotein possesses a characteristic fluid state, which is settled by the ratio of various constituents, i.e., apoproteins, cholesterol, phospholipids, and triacylglycerol, in the particle (Jonas, 1976). It is therefore suggested that the fluid state of the particle is a valuable indicator for plasma lipoprotein metabolism (Schroeder and Goh, 1979). Indeed, Ferretti et al. (1993) showed that Cu2þ-mediated oxidation decreases the fluidity at the surface of the HDL particle. In addition, it is suggested that the HDL oxidation leads to a loss of cholesterol-effluxing capacity, resulting from the rigidification through disturbance of optimal fluidity (Sola et al., 1993; Bonnefont-Rousselot et al., 1995). Thus, peroxidative incidents seem to cause physicochemical modifications in the lipidic structure of HDL.
Alteration in other components by oxidation It was well accepted that there is a negative relationship between the amounts of antioxidant molecules, such as a-tocopherol and ubiquinol-10, in HDL particles, and the degree of oxidation. A recent report has, however, shown that a-tocopherol acts as a pro-oxidant molecule of apoproteins in the early step of HDL oxidation under conditions without other antioxidants (Garner et al., 1998b). The oxidative mechanism of a-tocopherol appears to include formation of an a-tocopheroxy radical by one electron oxidants and the radical-induced promotion of continuous lipid peroxidation. It is also suggested that ubiquinol10 is unlikely to be a major protective antioxidant against HDL oxidation,
224
The HDL Handbook
because of a traced ratio (less than 1%) of HDL bearing ubiquinol-10 per total circulating HDL (Garner et al., 1998b).
FUNCTIONAL ALTERATIONS IN HDL BY OXIDATION Functional alterations detected in ox-HDL compared with HDL are summarized in Table 11.1.
Reverse cholesterol transport HDL is reported to suppress the progression of established lesions of atherosclerosis (Badimon et al., 1990; Rubin et al., 1991). The anti-atherogenic effect of HDL is mainly attributed to transport of excess cholesterol from peripheral tissues to the liver as described above. In contrast, oxidative modification of HDL is reported to reduce cholesterol efflux from cells and to induce the
TABLE 11.1
Functional alterations in HDL by oxidation
Alteration
Reference
Reduction of cholesterol transport Reduction of cholesterol efflux
Duell et al. (1991); Nagano et al. (1991); Salmon et al. (1992a); Mazie`re et al. (1993); Morel, (1994); Bonnefont-Rousselot et al. (1995); Marsche et al. (2002); Girona et al. (2003)
Reduction of cholesterol esterification
Francis et al. (1993, 1996); Musanti and Ghiselli (1993)
Disruption of cholesterol homeostasis
Ghiselli et al. (1992); Francis et al. (1993)
Reduction of antioxidant activity
Hahn and Subbiah, (1994); Jaouad et al. (2003); Nguyen and Sok (2003); Nguyen et al. (2004); Deakin et al. (2007)
Decrease in nitric oxide production and its Chin et al. (1992); Matsunaga et al. (2003a) function Induction of pro-inflammatory effect
Sharma et al. (1999); Matsunaga et al. (2003b)
Increase in expression of plasminogen activator inhibitor
Takahashi et al. (1996); Norata et al. (2004)
Induction of cytotoxic effect
Alomar et al. (1992); Hurtado et al. (1996); Sharma et al. (1999); Matsunaga et al. (2001, 2003a)
Chapter j 11 Function of oxidized high-density lipoprotein
225
accumulation of unesterified cholesterol in macrophages and fibroblasts (Ghiselli et al., 1992; Musanti and Ghiselli, 1993). Accordingly, ox-HDL, like ox-LDL, is believed to participate in the progression of atherosclerosis by the accumulation of free cholesterol in the peripheral tissues. Many results concerning the effects of HDL oxidation on its capacity to mediate cellular cholesterol efflux have been documented so far. Morel (1994) showed that a short incubation time (<6h) of human HDL with Cu2þ decreases its phospholipid content and the ability to mediate cholesterol efflux from cells, and instead increases free cholesterol, whereas a long exposure (>12h) diminishes efflux of cell cholesterol without alteration of apoproteins. In contrast, tyrosylation is unable to interfere with the ability of intrinsic HDL to take up cholesterol efflux, because there is no alteration in the lipidic components (Francis, 2000). To our knowledge, many experiments suggest that lipid peroxidation plays a key role in reduction of cholesterol efflux (Salmon et al., 1992a; Morel, 1994; Bonnefont-Rousselot et al., 1995; Sharma et al., 1999; Marsche et al., 2002; Girona et al., 2003). Accordingly, reduced cholesterol efflux to mildly oxidized HDL may be based on changes in its lipids rather than its apoproteins. Inactivation of lecithin:cholesterol acyltransferase (LCAT) observed during HDL oxidation is thought to be responsible for the reduction in the cholesterol efflux from the cells. McCall et al. (1995) and Mazie`re et al. (1993) independently found that a short exposure time of aldehyde with HDL particles induces both inactivation and reduction of LCAT prior to initiation of cross-linkage of apoproteins. These results imply that inactivation of LCAT is more dependent on the lipid peroxidation than on protein oxidation induced during HDL oxidation. Experiments using foam cells found that Cu2þ-mediated oxidation of HDL reduces cholesterol efflux from the cells, leading to a loss of macrophage membrane fluidity (Girona et al., 2003), and the decrease results from both a cross-linking of apoproteins and the resulting reduced binding affinity of HDL for the cell surface (Nagano et al., 1991; Marsche et al., 2002). In addition, ox-HDL disrupts cholesterol synthesis through inhibition of hydroxymethylglutaryl-CoA reductase by lipid peroxides (Ghiselli et al., 1992). Moreover, oxidation of HDL influences cholesterol transport mediated by both scavenger receptor-B1 and ATP-binding cassette (ABC)-A1 (Marsche et al., 2002; Shao et al., 2005). Thus, ox-HDL is suggested to act as a less effective transporter for cholesterol than native HDL and to induce disruption of cholesterol homeostasis. Reciprocal conversions between free cholesterol and cholesterol ester are mediated by intracellular enzymes such as acyl-CoA:cholesterol acyltransferase (ACAT) and neutral cholesterol esterase. Lipid-free HDL apoproteins take up the unesterified cholesterol, together with phospholipids, after hydrolysis of cholesterol ester by the activated ACAT, leading to pre-b-HDL formation. Oxidative tyrosylation of HDL enhanced the ability of ACAT to inhibit cholesterol esterification and to deplete cholesteryl ester stores in human skin fibroblasts (Francis et al., 1993). While these findings suggest ACAT inhibition
226
The HDL Handbook
is mainly responsible for the accumulation of unesterified cholesterol observed with ox-HDL, the oxidized lipids present in ox-HDL appear unable to modulate the enzymatic activity (Musanti and Ghiiiselli, 1993; Francis et al., 1996). Therefore, Francis et al. (1996) proposed the existence of a novel pathway, independent of ACAT inhibition, for depletion of cellular cholesteryl esters prior to the removal of cholesterol from the plasma membrane. In addition, it is suggested that internalized ox-HDL, like ox-LDL, is delivered to lysosomes rather than to the Golgi apparatus, thereby leading to the accumulation of unesterified cholesterol, distinct from transportation of native HDL to the Golgi compartment (Musanti and Ghiselli, 1993).
Antioxidant activity HDL plays an important role in inhibiting the generation of ox-LDL (Parthasarathy et al., 1990; Mackness et al., 1993; Hahn and Subbiah, 1994; Maier et al., 1994; Nofer et al., 2002). The inhibitory mechanism is thought to be coordinately mediated by apoA-I and several enzymes, such as paraoxonase (PON), LCAT, platelet activating factor acetylhydrolase and glutathione peroxidase, in addition to antioxidant molecules in the HDL particle. ApoA-I is solely capable of reducing peroxides of both phospholipids and cholesteryl esters, and removing eicosanoid hydroperoxides that cause the non-enzymatic oxidation of lipoprotein phospholipids from LDL particles (Ohta et al., 1989; Garner et al., 1998a; Navab et al., 2000a; Assmann and Nofer, 2003). On the other hand, it is suggested that a calcium-dependent HDL-associated ester hydrolase, PON, catalyzes the hydrolysis of organic phosphates, aromatic carboxylic acid ester, and carbamates, and mainly participates in the antioxidant function of HDL toward LDL oxidation (Mackness and Durrington, 1995). The antioxidant mechanism of PON appears to be mediated by hydrolyzing long-chain oxidized phospholipids, formed during the oxidation of lipoproteins (Watson et al., 1995). In animal models, downregulation of PON raises the susceptibility for the formation of atherosclerotic lesions (Shih et al., 1998), and its genetic overexpression diminishes the lesions (Tward et al., 2002). Thus, it is well accepted that both apoA-I and PON play key roles in prevention of lipoprotein oxidation by degrading the reactive the lipid products formed (Mackness et al., 1991, 1993, 2000; Bowry et al., 1992; Watson et al., 1995; Aviram et al., 1998; Garner et al., 1998a,b; Laplaud et al., 1998; Shih et al., 1998). It has been shown that HDL isolated from Cu2þ-treated plasma significantly loses its ability to inhibit LDL oxidation, suggesting that the antioxidant ability of HDL is quenched when HDL itself is oxidized (Hahn and Subbiah, 1994). The loss of antioxidant function of HDL by oxidation is inferred to be due to a decrease in HDL-associated PON activity, but not free PON activity (Nguyen and Sok, 2003; Nguyen et al., 2004). Although the extent of PON inactivation depends on both the oxidation extent of HDL and the oxidation system used,
Chapter j 11 Function of oxidized high-density lipoprotein
227
a significant correlation between the rates of HDL oxidation and PON inactivation is found (Aviram et al., 1998, 1999; Jaouad et al., 2003; Nguyen et al., 2004; Ferretti et al., 2005). Recent studies have found that the arylesterase activity of PON is markedly inhibited during HDL oxidation, whereas its phospholipase A2 activity remains unaffected (Ahmed et al., 2001; Karabina et al., 2005). These findings suggest that the unaltered phospholipase-A2 activity of PON under oxidative conditions participates in the cleavage of peroxidized PC into lyso-PC and the peroxidized fatty acid detected in oxidized lipoproteins, because incubation of phospholipid with PON in the presence of apoA-I generates lyso-PC (Ahmed et al., 2002). Thus, the formation of lyso-PC appears to be mainly responsible for the interaction of apoA-I and PON in the HDL particle. PON translocates from cell membrane to lipoproteins during association of HDL with hepatocytes. In the ox-HDL particle, the ability of HDL to promote the release of PON and stabilize its activity is markedly decreased, indicating that the depletion of active PON in ox-HDL particles leads to impairment of the antioxidant property of intrinsic HDL (Deakin et al., 2007). It has been demonstrated that HDL, especially apoA-I, is involved in accepting cholesteryl ester hydroperoxides via cholesteryl ester transfer protein (CETP) (Christison et al., 1995) and selectively transferring them to liver cells after rapid transportation (Sattler and Stocker, 1993; Christison et al., 1996). One of the most significant antioxidant functions of HDL is, therefore, regarded to be the apoA-I-mediated transport system that accepts and carries away oxidant molecules. While oxidative modification of HDL probably impairs the capacity for transporting cholesterol hydroperoxides in the manner described above, the exact effect of oxidative modification on the structure and function of CETP remains unclear. Other enzymes implicated in the antioxidant properties of HDL are LCAT and platelet-activating factor acyl hydrolase. These enzymes have also been reported to be highly inhibited by oxidative modification (Mazie`re et al., 1993; McCall et al., 1995; Ahmed et al., 2001). It should be mentioned that oxidation not only removes the antioxidant activity of HDL but also converts HDL into a cytotoxic particle such as ox-LDL (Alomar et al., 1992; Hurtado et al., 1996). In cultured macrophages, the lethal effect of ox-HDL is equally or more cytotoxic than ox-LDL when both lipoproteins were compared at the same level of lipids (Hurtado et al., 1996). We have also shown that Cu2þ-mediated oxidation of HDL induced endothelial apoptosis, and suggested that ox-HDL participates in the onset and the progression of atherosclerosis (Matsunaga et al., 2003a). As a possible lethal pathway of ox-HDL, it is proposed that ox-HDL increases the intracellular level of reactive oxygen species, and the oxidative stress is associated with activation of NADPH oxidase and downregulation of antioxidant enzymes (superoxide dismutase, catalase and glutathione-peroxidase), similar to ox-LDL (Sharma et al., 1999; Matsunaga et al., 2001, 2003a).
228
The HDL Handbook
Nitric oxide generation NO, a potent vasodilator produced in macrophages and endothelial cells, is considered to act as a mediator for both atherogenic and anti-atherogenic events, dependent on its source of production. The literature shows that HDL enhances the release of endothelial constitutive NO synthase (ecNOS)-derived NO through the pathway including scavenger receptor-B1 and protein kinase, Akt, leading to vasorelaxation (Tauber et al., 1980; Kuhn et al., 1991; Murugesan et al., 1994; Zeiher et al., 1994; Uittenbogaard et al., 2000; Yuhanna et al., 2001; Li et al., 2002; Mineo et al., 2003). These studies suggest that the increased level of the NO derived from ecNOS plays an important role in the anti-atherogenic ability of HDL to preserve normal endothelial function. Increased level of ox-LDL and lyso-PC in the ox-LDL particle are suggested to play a crucial role in the alteration of endothelium-derived arterial relaxation via reduced expression of ecNOS and the following decrease in NO level (Simon et al., 1990; Chin et al., 1992; Mangin et al., 1993; Nuszkowski et al., 2001). We have shown that, like ox-LDL, HDL highly oxidized by Cu2þ reduces the expression of ecNOS and the release of NO in human endothelial cells (Matsunaga et al., 2003a), whereas such an alteration is not observed in the cells treated with peroxynitrite-modified HDL (Matsunaga et al., 2001). Stimulation with low concentration of ox-HDL3 or its lipidic fraction almost blocks NO-stimulated cGMP accumulation, suggesting that the lipidic component of ox-HDL quenches NO-mediated function after its release from endothelial cells (Chin et al., 1992). These findings speculate that reduced generation of NO by HDL oxidation removes many of its beneficial properties against the development of atherosclerosis and its complications because NO inhibits platelet aggregation, leukocyte adhesion and vascular smooth muscle cell proliferation, in addition to maintaining vasodilation. HDL and the increased level of intracellular NO inhibit oxidative stress and apoptosis in endothelial cells (Kimura et al., 2001; Nofer et al., 2001; Kotamraju et al., 2006). The inhibitory mechanism of NO includes inhibition of typical apoptotic pathways such as caspase activation (Kimura et al., 2001; Nofer et al., 2001) and enhancement of proteasomal activity to degrade oxidized proteins (Kotamraju et al., 2006). In addition, increases in the intracellular reactive oxygen species level and apoptotic cells by ox-HDL are markedly inhibited by an exogenous NO donor (Matsunaga et al. unpublished data). These findings, together with the data reported by Kotamraju et al. (2001, 2006), indicate that NO generation is responsible for protecting against oxidized lipoprotein-induced cytotoxicity. Recent studies have demonstrated that HDL serves as a carrier of bioactive lysosphingolipids such as sphingosine-1-phosphate, sphingosylphosphorylcholine and lysosulfatide, and mimic HDL bearing these lipids has the ability to induce vasorelaxation and to inhibit apoptosis (Kimura et al., 2001; Nofer
Chapter j 11 Function of oxidized high-density lipoprotein
229
et al., 2001, 2004). Therefore, the phenomenon observed during HDL oxidation may be explained by alterations in the structure and the content of these lipids.
Anti-inflammatory effect Increased binding of circulating leukocytes on the endothelial surface is often encountered in the early stages of atherosclerosis. In experiments using various animal models, it has been shown that HDL reduces expression of adhesion molecules, such as vascular cell adhesion molecule (VCAM)-1, intracellular adhesion molecule-1 and E-selectin as well as production of pro-inflammatory cytokines including interleukin (IL)-1 and IL-8 (Dimayuga et al., 1999; Cockerill et al., 2001a,b). HDL also appears to elicit LDL resistance to oxidation in the arterial wall and to reduce the ability of ox-LDL to enhance monocyte chemotactic activity (Navab et al., 2000b). In addition, these anti-inflammatory effects are though to be mediated by lysosphingolipids that are localized in the HDL particle (Mehrabian and Allayee, 2003; Nofer et al., 2003). HDL can completely inhibit ox-LDL-induced adhesion of macrophages to the endothelial monolayer, whereas ox-HDL is unable to alter ox-LDL-induced adhesion (Maier et al., 1994). Sharma et al. (1999) found that ox-HDL increases the number of monocytes adhered to aortic endothelium, indicating that oxidation makes HDL more pro-inflammatory. Our recent study has shown that Cu2þ-oxidized HDL activates signaling, including nuclear factor-kappa B (NF-kB), and this signaling is triggered by enhanced generation of reactive oxygen species and subsequent I-kB ubiquitination, as observed in ox-LDL (Matsunaga et al., 20003b). This fact permits us to speculate that ox-HDL particles include a certain initiating factor for transcriptional activation of adhesion molecules via the NF-kB signaling pathway, because genes for adhesion molecules, including VCAM-1, are conserved downstream of the transcriptional region activated by NF-kB. Other literature shows that ox-HDL inhibits the secretion of tumor necrosis factor-a by macrophages and the inhibition is mainly due to reactive low-molecular weight aldehydes (2,4heptadienal, hexanal, 2-nonenal, 2-octenal, 2,4-decadienal), but not hydroperoxides of fatty acids (Girona et al., 1997). Thus, ox-HDL, in particular some reactive products derived from its lipid peroxidation, is likely to play a crucial role in the modulation of the inflammatory response by macrophages in the early process of atherogenesis. However, at present, there are a lot of ambiguous points about the effect of ox-HDL on the expression of adhesion molecules and the production of cytokines.
Other functions In addition to the protective functions against atherogenesis described above, HDL is known to have an anti-thrombotic effect (Saku et al., 1985) and prostacyclin-stabilizing activity (Fleischer et al., 1982; Yui et al., 1988). Norata
230
The HDL Handbook
et al. (2004) have recently found that ox-HDL3, but not native HDL3, increases the expression and release of plasminogen activator inhibitor-1 in endothelial cells through a molecular mechanism involving MAPK activation and mRNA stabilization. While native HDL inhibits oxidized lipoprotein-induced platelet activation, ox-HDL, like ox-LDL, has the ability to activate platelets (Ardlie et al., 1989; Takahashi et al., 1996). HDL also has antiviral activity (Singh and Baron, 2000). Van Lenten et al. (2001) found that HDL isolated from influenza-infected mice had the ability to inhibit both LDL oxidation and LDL-induced monocyte chemotactic activity in human artery wall cell cocultures. It has also been suggested that loss of the beneficial effect of HDL during acute influenza infection is mediated by alterations in PON, platelet-activating factor acetylhydrolase, ceruloplasmin and apoJ in HDL particles. In addition, oxidative stress is believed to be tightly associated with inflammatory disorders observed after influenza virus infection (Cai et al., 2003; Kumar et al., 2005). Therefore, those alterations in HDL components might be attributable to oxidative stress, whereas the exact molecular mechanism of modification of HDL particles has not yet been clarified.
CONCLUSION Collectively, localization of ox-HDL has been a gradual process in plasma from patients with atherosclerosis and its complications. In addition, oxidation of the HDL particle is likely to have important consequences not only for cholesterol homeostasis in peripheral tissue but also development of oxidative stressassociated vascular diseases including inflammatory diseases. In previous experiments using oxidant molecules, functional alterations in the oxidized components in HDL particles have also become clear from continuous investigations by numerous researchers. However, little has reported about the direct pathophysiological role of HDL oxidation in vascular diseases. Therefore, further clinical investigation will be needed in order to establish ox-HDL as a useful indicator and to exploit new methods of medication for these diseases.
REFERENCES Ahmed, Z., Ravandi, A., Maguire, G. F., et al. (2001). Apolipoprotein A-I promotes the formation of phosphatidylcholine core aldehydes that are hydrolyzed by paraoxonase (PON-1) during high density lipoprotein oxidation with a peroxynitrite donor. J Biol Chem, 276, 24473e24481. Ahmed, Z., Ravandi, A., Maguire, G. F., et al. (2002). Multiple substrates for paraoxonase-1 during oxidation of phosphatidylcholine by peroxynitrite. Biochem Biophys Res Commun, 290, 391e396. Alomar, Y., Ne`gre-Salvayre, A., Levade, T., Valdiguie´, P., & Salvayre, R. (1992). Oxidized HDL are much less cytotoxic to lymphoblastoid cells than oxidized LDL. Biochim Biophys Acta, 1128, 163e166. Ardlie, N. G., Selley, M. L., & Simons, L. A. (1989). Platelet activation by oxidatively modified low density lipoproteins. Atherosclerosis, 76, 117e124.
Chapter j 11 Function of oxidized high-density lipoprotein
231
Assmann, G., & Gotto, A. M., Jr. (2004). HDL cholesterol and protective factors in atherosclerosis. Circulation, 109(Suppl 1), III8e14. Assmann, G., & Nofer, J.-R. (2003). Atheroprotective effects of high-density lipoproteins. Annu Rev Med, 54, 321e341. Aviram, M., Rosenblat, M., Bisgaier, C. L., Newton, R. S., Primo-Parmo, S. L., & La Du, B. N. (1998). Paraoxonase inhibits high-density lipoprotein oxidation and preserves its functions. A possible peroxidative role for paraoxonase. J Clin Invest, 101, 1581e1590. Aviram, M., Rosenblat, M., Billecke, S., et al. (1999). Human serum paraoxonase (PON 1) is inactivated by oxidized low density lipoprotein and preserved by antioxidants. Free Radic Biol Med, 26, 892e904. Badimon, J. J., Badimon, L., & Fuster, V. (1990). Regression of atherosclerotic lesions by high density lipoprotein plasma fraction in the cholesterol-fed rabbit. J Clin Invest, 85, 1234e1241. Bergt, C., Oettl, K., Keller, W., et al. (2000). Reagent or myeloperoxidase-generated hypochlorite affects discrete regions in lipid-free and lipid-associated human apolipoprotein A-I. Biochem J, 346, 345e354. Berliner, J. A., & Heinecke, J. W. (1996). The role of oxidized lipoproteins in atherogenesis. Free Radic Biol Med, 20, 707e727. Bjornheden, T., Babyi, A., Bondjers, G., & Wiklund, O. (1996). Accumulation of lipoprotein fractions and subfractions in the arterial wall, determined in an in vitro perfusion system. Atherosclerosis, 123, 43e56. Bonnefont-Rousselot, D., Motta, C., Khalil, A. O., et al. (1995). Physicochemical changes in human high-density lipoproteins (HDL) oxidized by gamma radiolysis-generated oxyradicals. Effect on their cholesterol effluxing capacity. Biochim Biophys Acta, 1255, 23e30. Bowie, A., Owens, D., Collins, P., Johnson, A., & Tomkin, G. H. (1993). Glycosylated low density lipoprotein is more sensitive to oxidation: implications for the diabetic patient? Atherosclerosis, 102, 63e67. Bowry, V. W., Stanley, K. K., & Stocker, R. (1992). High density lipoprotein is the major carrier of lipid hydroperoxides in human blood plasma from fasting donors. Proc Natl Acad Sci USA, 89, 10316e10320. Bradamante, S., Barenghi, L., Giudici, G. A., & Vergani, C. (1992). Free radicals promote modifications in plasma high-density lipoprotein: nuclear magnetic resonance analysis. Free Radic Biol Med, 12, 193e203. Brennan, M.-L., & Hazen, S. L. (2003). Emerging role of myeloperoxidase and oxidant stress markers in cardiovascular risk assessment. Curr Opin Lipidol, 14, 353e359. Brennan, M. L., Penn, M. S., Van Lente, F., et al. (2003). Prognostic value of myeloperoxidase in patients with chest pain. N Engl J Med, 349, 1595e1604. Brinton, E. A., Oram, J. F., Chen, C. H., Albers, J. J., & Bierman, E. L. (1986). Binding of high density lipoprotein to cultured fibroblasts after chemical alteration of apolipoprotein amino acid residues. J Biol Chem, 261, 495e503. Brown, M. S., Ho, Y. K., & Goldstein, J. L. (1980). The cholesteryl ester cycle in macrophage foam cells. Continual hydrolysis and re-esterification of cytoplasmic cholesteryl esters. J Biol Chem, 255, 9344e9352. Cai, J., Chen, Y., Seth, S., Furukawa, S., Compans, R. W., & Jones, D. P. (2003). Inhibition of influenza infection by glutathione. Free Radic Biol Med, 34, 928e936. Carr, A. C., McCall, M. R., & Frei, B. (2000). Oxidation of LDL by myeloperoxidase and reactive nitrogen species oxidation of LDL by myeloperoxidase and reactive nitrogen species. Arterioscler Thromb Vasc Biol, 20, 1716e1723.
232
The HDL Handbook
Cathcart, M. K. (2004). Regulation of superoxide anion production by NADPH oxidase in monocytes/macrophages. Contributions to atherosclerosis. Arterioscler Thromb Vasc Biol, 24, 23e28. Chacko, G. K. (1985). Modification of human high density lipoprotein (HDL3) with tetranitromethane and the effect on its binding to isolated rat liver plasma membranes. J Lipid Res, 26, 745e754. Chacko, G. K., Mahlberg, F. H., & Johnson, W. J. (1988). Cross-linking of apolipoproteins in high density lipoprotein by dimethylsuberimidate inhibits specific lipoprotein binding to membranes. J Lipid Res, 29, 319e324. Chin, J. H., Azhar, S., & Hoffman, B. B. (1992). Inactivation of endothelial derived relaxing factor by oxidized lipoproteins. J Clin Invest, 89, 10e18. Christen, S., Woodall, A. A., Shigenaga, M. K., Southwell-Keely, P. T., Duncan, M. W., & Ames, B. N. (1997). Gamma-tocopherol traps mutagenic electrophiles such as NO(X) and complements alpha-tocopherol: physiological implications. Proc Natl Acad Sci USA, 94, 3217e3222. Christison, J. K., Rye, K. A., & Stocker, R. (1995). Exchange of oxidized cholesteryl linoleate between LDL and HDL mediated by cholesteryl ester transfer protein. J Lipid Res, 36, 2017e2026. Christison, J., Karjalainen, A., Brauman, J., Bygrave, F., & Stocker, R. (1996). Rapid reduction and removal of HDL- but not LDL-associated cholesteryl ester hydroperoxides by rat liver perfused in situ. Biochem J, 314, 739e742. Cockerill, G. W., Huehns, T. Y., Weerasinghe, A., et al. (2001a). Elevation of plasma high-density lipoprotein concentration reduces interleukin-1-induced expression of E-selectin in an in vivo model of acute inflammation. Circulation, 103, 108e112. Cockerill, G. W., McDonald, M. C., Mota-Filipe, H., Cuzzocrea, S., Miller, N. E., & Thiemermann, C. (2001b). High density lipoproteins reduce organ injury and organ dysfunction in a rat model of hemorrhagic shock. FASEB J, 15, 1941e1952. Collins, T., & Cybulsky, M. I. (2001). NF-kappaB: pivotal mediator or innocent bystander in atherogenesis? J Clin Invest, 107, 255e264. Cyrus, T., Witztum, J. L., Rader, D. J., et al. (1999). Disruption of the 12/15-lipoxygenase gene diminishes atherosclerosis in apo E-deficient mice. J Clin Invest, 103, 1597e1604. Cyrus, T., Pratico, D., Zhao, L., et al. (2001). Absence of 12/15-lipoxygenase expression decreases lipid peroxidation and atherogenesis in apolipoprotein E-deficient mice. Circulation, 103, 2277e2282. Daugherty, A., Dunn, J. L., Rateri, D. L., & Heinecke, J. W. (1994). Myeloperoxidase, a catalyst for lipoprotein oxidation, is expressed in human atherosclerotic lesions. J Clin Invest, 94, 437e444. Deakin, S., Moren, X., & James, R. W. (2007). HDL oxidation compromises its influence on paraoxonase-1 secretion and its capacity to modulate enzyme activity. Arterioscler Thromb Vasc Biol, 27, 1146e1152. Dimayuga, P., Zhu, J., Oguchi, S., et al. (1999). Reconstituted HDL containing human apolipoprotein A-1 reduces VCAM-1 expression and neointima formation following periadventitial cuff-induced carotid injury in apoE null mice. Biochem Biophys Res Commun, 264, 465e468. Ducrocq, C., Blanchard, B., Pignatelli, B., & Ohshima, H. (1999). Peroxynitrite: an endogenous oxidizing and nitrating agent. Cell Mol Life Sci, 55, 1068e1077. Duell, P. B., Oram, J. F., & Bierman, E. L. (1991). Nonenzymatic glycosylation of HDL and impaired HDL-receptor-mediated cholesterol efflux. Diabetes, 40, 377e384.
Chapter j 11 Function of oxidized high-density lipoprotein
233
Dwyer, J. H., Allayee, H., Dwyer, K. M., et al. (2004). Arachidonate 5-lipoxygenase promoter genotype, dietary arachidonic acid, and atherosclerosis. N Engl J Med, 350, 29e37. Esterbauer, H., Gebicki, J., Puhl, H., & Jurgens, G. (1992). The role of lipid peroxidation and antioxidants in oxidative modification of LDL. Free Radic Biol Med, 13, 341e390. Esterbauer, H., Schaur, R. J., & Zollner, H. (1991). Chemistry and biochemistry of 4-hydroxynonenal, malonaldehyde and related aldehydes. Free Radic Biol Med, 11, 81e128. Ferretti, G., Bacchetti, T., Moroni, C., et al. (2005). Paraoxonase activity in high-density lipoproteins: a comparison between healthy and obese females. J Clin Endocrinol Metab, 90, 1728e1733. Ferretti, G., Taus, M., Dousset, N., Solera, M. L., Valdiguie´, P., & Curatola, G. (1993). Physicochemical properties of copper-oxidized high density lipoprotein: a fluorescence study. Biochem Mol Biol Int, 30, 713e719. Fleisher, L. N., Tall, A. R., Witte, L. D., Miller, R. W., & Cannon, P. J. (1982). Stimulation of arterial endothelial cell prostacyclin synthesis by high density lipoproteins. J Biol Chem, 257, 6653e6655. Francis, G. A. (2000). High density lipoprotein oxidation: in vitro susceptibility and potential in vivo consequences. Biochim Biophys Acta, 1483, 217e235. Francis, G. A., Mendez, A. J., Bierman, E. L., & Heinecke, J. W. (1993). Oxidative tyrosylation of high density lipoprotein by peroxidase enhances cholesterol removal from cultured fibroblasts and macrophage foam cells. Proc Natl Acad Sci USA, 90, 6631e6635. Francis, G. A., Oram, J. F., Heinecke, J. W., & Bierman, E. L. (1996). Oxidative tyrosylation of HDL enhances the depletion of cellular cholesteryl esters by a mechanism independent of passive sterol desorption. Biochemistry, 35, 15188e15197. Frei, B., Yamamoto, Y., Niclas, D., & Ames, B. N. (1988). Evaluation of an isoluminol chemiluminescence assay for the detection of hydroperoxides in human blood plasma. Anal Biochem, 175, 120e130. Garner, B., Waldeck, A. R., Witting, P. K., Rye, K. A., & Stocker, R. (1998a). Oxidation of high density lipoproteins. II. Evidence for direct reduction of lipid hydroperoxides by methionine residues of apolipoproteins AI and AII. J Biol Chem, 273, 6088e6095. Garner, B., Witting, P. K., Waldeck, A. R., Christison, J. K., Raftery, M., & Stocker, R. (1998b). Oxidation of high density lipoproteins. I. Formation of methionine sulfoxide in apolipoproteins AI and AII is an early event that accompanies lipid peroxidation and can be enhanced by alpha-tocopherol. J Biol Chem, 273, 6080e6087. Gaut, J. P., Byun, J., Tran, H. D., et al. (2002). Myeloperoxidase produces nitrating oxidants in vivo. J Clin Invest, 109, 1311e1319. George, J., Afek, A., Shaish, A., et al. (2001). 12/15-lipoxygenase gene disruption attenuates atherogenesis in LDL receptor-deficient mice. Circulation, 104, 1646e1650. Ghiselli, G., Giorgini, L., Gelati, M., & Musanti, R. (1992). Oxidatively modified HDLs are potent inhibitors of cholesterol biosynthesis in human skin fibroblasts. Arterioscler Thromb, 12, 929e935. Girona, J., La Ville, A. E., Heras, M., Olive´, S., & Masana, L. (1997). Oxidized lipoproteins including HDL and their lipid peroxidation products inhibit TNF-alpha secretion by THP-1 human macrophages. Free Radic Biol Med, 23, 658e667. Girona, J., LaVille, A. E., Sola`, R., Motta, C., & Masana, L. (2003). HDL derived from the different phases of conjugated diene formation reduces membrane fluidity and contributes to a decrease in free cholesterol efflux from human THP-1 macrophages. Biochim Biophys Acta, 1633, 143e148.
234
The HDL Handbook
Greilberger, J., & Ju¨rgens, G. (1998). Oxidation of high-density lipoprotein HDL3 leads to exposure of ApoA-I and ApoA-II epitopes and to formation of aldehyde protein adducts, and influences binding of oxidized low-density lipoprotein to type I and type III collagen in vitro1. Biochem J, 331, 185e191. Hahn, M., & Subbiah, M. T. (1994). Significant association of lipid peroxidation products with high density lipoproteins. Biochem Mol Biol Int, 33, 699e704. Harats, D., Shaish, A., George, J., et al. (2000). Overexpression of 15-lipoxygenase in vascular endothelium accelerates early atherosclerosis in LDL receptor-deficient mice. Arterioscler Thromb Vasc Biol, 20, 2100e2105. Hazell, L. J., Arnold, L., Flowers, D., Waeg, G., Malle, E., & Stocker, R. (1996). Presence of hypochlorite-modified proteins in human atherosclerotic lesions. J Clin Invest, 97, 1535e1544. Hazen, S. L., & Heinecke, J. W. (1997). 3-Chlorotyrosine, a specific marker of myeloperoxidasecatalyzed oxidation, is markedly elevated in low density lipoprotein isolated from human atherosclerotic intima. J Clin Invest, 99, 2075e2081. Hedrick, C. C., Thorpe, S. R., Fu, M. X., et al. (2000). Glycation impairs high-density lipoprotein function. Diabetologia, 43, 312e320. Heinecke, J. W., Li, W., Francis, G. A., & Goldstein, J. A. (1993). Tyrosyl radical generated by myeloperoxidase catalyzes the oxidative cross-linking of proteins. J Clin Invest, 91, 2866e2872. Hoff, H. F., O’Neil, J., Chisolm, G. M., 3rd, et al. (1989). Modification of low density lipoprotein with 4-hydroxynonenal induces uptake by macrophages. Arteriosclerosis, 9, 538e549. Holvoet, P., Perez, G., Zhao, Z., Brouwers, E., Bernar, H., & Collen, D. (1995). Malondialdehydemodified low density lipoproteins in patients with atherosclerotic disease. J Clin Invest, 95, 2611e2619. Hurtado, I., Fiol, C., Gracia, V., & Caldu, P. (1996). In vitro oxidised HDL exerts a cytotoxic effect on macrophages. Atherosclerosis, 125, 39e46. Ischiropoulos, H., Zhu, L., Chen, J., et al. (1992). Peroxynitrite-mediated tyrosine nitration catalyzed by superoxide dismutase. Arch Biochem Biophys, 298, 431e437. Jaouad, L., Milochevitch, C., & Khalil, A. (2003). PON1 paraoxonase activity is reduced during HDL oxidation and is an indicator of HDL antioxidant capacity. Free Radic Res, 37, 77e83. Jimi, S., Sakata, N., Matunaga, A., & Takebayashi, S. (1994). Low density lipoproteins bind more to type I and III collagens by negative charge-dependent mechanisms than to type IV and V collagens. Atherosclerosis, 107, 109e116. Johnson, W. J., Mahlberg, F. H., Rothblat, G. H., & Phillips, M. C. (1991). Cholesterol transport between cells and high-density lipoproteins. Biochim Biophys Acta, 1085, 273e298. Jonas, A. (1976). Microviscosity of lipid domains in human serum lipoproteins. Biochim Biophys Acta, 486, 10e22. Karabina, S. A., Lehner, A. N., Frank, E., Parthasarathy, S., & Santanam, N. (2005). Oxidative inactivation of paraoxonase e implications in diabetes mellitus and atherosclerosis. Biochim Biophys Acta, 1725, 213e221. Kettle, A. J., van Dalen, C. J., & Winterbourn, C. C. (1997). Peroxynitrite and myeloperoxidase leave the same footprint in protein nitration. Redox Rep, 3, 257e258. Kimura, T., Sato, K., Kuwabara, A., et al. (2001). Sphingosine 1-phosphate may be a major component of plasma lipoproteins responsible for the cytoprotective actions in human umbilical vein endothelial cells. J Biol Chem, 276, 31780e31785. Kotamraju, S., Hogg, N., Joseph, J., Keefer, L. K., & Kalyanaraman, B. (2001). Inhibition of oxidized low-density lipoprotein-induced apoptosis in endothelial cells by nitric oxide. Peroxyl radical scavenging as an antiapoptotic mechanism. J Biol Chem, 276, 17316e17323.
Chapter j 11 Function of oxidized high-density lipoprotein
235
Kotamraju, S., Matalon, S., Matsunaga, T., Shang, T., Hickman-Davis, J. M., & Kalyanaraman, B. (2006). Upregulation of immunoproteasomes by nitric oxide: potential antioxidative mechanism in endothelial cells. Free Radic Biol Med, 40, 1034e1044. Kuhn, F. E., Mohler, E. R., Satler, L. F., Reagan, K., Lu, D. Y., & Rackley, C. E. (1991). Effects of high-density lipoprotein on acetylcholine-induced coronary vasoreactivity. Am J Cardiol, 68, 1425e1430. Kuhn, H., Heydeck, D., Hugou, I., & Gniwotta, C. (1997). In vivo action of 15-lipoxygenase in early stages of human atherogenesis. J Clin Invest, 99, 888e893. Kumar, P., Khanna, M., Srivastava, V., Tyagi, Y. K., Raj, H. G., & Ravi, K. (2005). Effect of quercetin supplementation on lung antioxidants after experimental influenza virus infection. Exp Lung Res, 31, 449e459. La Ville, A. E., Sola, R., Balanya, J., Turner, P. R., & Masana, L. (1994). In vitro oxidised HDL is recognized by the scavenger receptor of macrophages: implications for its protective role in vivo. Atherosclerosis, 105, 179e189. Laplaud, P. M., Dantoine, T., & Chapman, M. J. (1998). Paraoxonase as a risk marker for cardiovascular disease: facts and hypotheses. Clin Chem Lab Med, 36, 431e441. Leeuwenburgh, C., Rasmussen, J. E., Hsu, F. F., Mueller, D. M., Pennathur, S., & Heinecke, J. W. (1997a). Mass spectrometric quantification of markers for protein oxidation by tyrosyl radical, copper, and hydroxyl radical in low density lipoprotein isolated from human atherosclerotic plaques. J Biol Chem, 272, 3520e3526. Leeuwenburgh, C., Hardy, M. M., Hazen, S. L., et al. (1997b). Reactive nitrogen intermediates promote low density lipoprotein oxidation in human atherosclerotic intima. J Biol Chem, 272, 1433e1436. Li, X. A., Titlow, W. B., Jackson, B. A., et al. (2002). High density lipoprotein binding to scavenger receptor, Class B, type I activates endothelial nitric-oxide synthase in a ceramidedependent manner. J Biol Chem, 277, 11058e11063. Linton, M. F., & Fazio, S. (2004). Cyclooxygenase-2 and inflammation in atherosclerosis. Curr Opin Pharmacol, 4, 116e123. Mackness, M. I., & Durrington, P. N. (1995). HDL, its enzymes and its potential to influence lipid peroxidation. Atherosclerosis, 115, 243e253. Mackness, M. I., Arrol, S., & Durrington, P. N. (1991). Paraoxonase prevents accumulation of lipoperoxides in low-density lipoprotein. FEBS Lett, 286, 152e154. Mackness, M. I., Durrington, P. N., & Mackness, B. (2000). How high-density lipoprotein protects against the effects of lipid peroxidation. Curr Opin Lipidol, 11, 383e388. Mackness, M. I., Abbott, C., Arrol, S., & Durrington, P. N. (1993). The role of high-density lipoprotein and lipid-soluble antioxidant vitamins in inhibiting low-density lipoprotein oxidation. Biochem J, 294, 829e834. Mackness, M. I., Arrol, S., Abbott, C., & Durrington, P. N. (1993). Protection of low-density lipoprotein against oxidative modification by high-density lipoprotein associated paraoxonase. Atherosclerosis, 104, 129e135. Maier, J. A., Barenghi, L., Pagani, F., Bradamante, S., Comi, P., & Ragnotti, G. (1994). The protective role of high-density lipoprotein on oxidized-low-density-lipoprotein-induced U937/ endothelial cell interactions. Eur J Biochem, 221, 35e41. Mangin, E. L., Jr., Kugiyama, K., Nguy, J. H., Kerns, S. A., & Henry, P. D. (1993). Effects of lysolipids and oxidatively modified low density lipoprotein on endothelium-dependent relaxation of rabbit aorta. Circ Res, 72, 161e166. Marcel, Y. L., Jewer, D., Leblond, L., Weech, P. K., & Milne, R. W. (1989). Lipid peroxidation changes the expression of specific epitopes of apolipoprotein A-I. J Biol Chem, 264, 19942e19950.
236
The HDL Handbook
Marsche, G., Hammer, A., Oskolkova, O., Kozarsky, K. F., Sattler, W., & Malle, E. (2002). Hypochlorite-modified high density lipoprotein, a high affinity ligand to scavenger receptor class B, type I, impairs high density lipoprotein-dependent selective lipid uptake and reverse cholesterol transport. J Biol Chem, 277, 32172e32179. Mashima, R., Yamamoto, Y., & Yoshimura, S. (1998). Reduction of phosphatidylcholine hydroperoxide by apolipoprotein A-I: purification of the hydroperoxide-reducing proteins from human blood plasma. J Lipid Res, 39, 1133e1140. Matsunaga, T., Nakajima, T., Sonoda, M., et al. (2001). Modulation of reactive oxygen species in endothelial cells by peroxynitrite-treated lipoproteins. J Biochem (Tokyo), 130, 285e293. Matsunaga, T., Nakajima, T., Miyazaki, T., et al. (2003a). Glycated high-density lipoprotein regulates reactive oxygen species and reactive nitrogen species in endothelial cells. Metabolism, 52, 42e49. Matsunaga, T., Hokari, S., Koyama, I., Harada, T., & Komoda, T. (2003b). NF-kappa B activation in endothelial cells treated with oxidized high-density lipoprotein. Biochem Biophys Res Commun, 303, 313e319. Mazie`re, J. C., Myara, I., Salmon, S., et al. (1993). Copper- and malondialdehyde-induced modification of high density lipoprotein and parallel loss of lecithin cholesterol acyltransferase activation. Atherosclerosis, 104, 213e219. McCall, M. R., Tang, J. Y., Bielicki, J. K., & Forte, T. M. (1995). Inhibition of lecithin-cholesterol acyltransferase and modification of HDL apolipoproteins by aldehydes. Arterioscler Thromb Vasc Biol, 15, 1599e1606. Mehrabian, M., & Allayee, H. (2003). 5-lipoxygenase and atherosclerosis. Curr Opin Lipidol, 14, 447e457. Mehrabian, M., Allayee, H., Wong, J., et al. (2002). Identification of 5-lipoxygenase as a major gene contributing to atherosclerosis susceptibility in mice. Circ Res, 91, 120e126. Mineo, C., Yuhanna, I. S., Quon, M. J., & Shaul, P. W. (2003). High density lipoprotein-induced endothelial nitric-oxide synthase activation is mediated by Akt and MAP kinases. J Biol Chem, 278, 9142e9149. Miyazaki, A., Rahim, A. T., Araki, S., Morino, Y., & Horiuchi, S. (1991). Chemical cross-linking alters high-density lipoprotein to be recognized by a scavenger receptor in rat peritoneal macrophages. Biochim Biophys Acta, 1082, 143e151. Morel, D. W. (1994). Reduced cholesterol efflux to mildly oxidized high density lipoprotein. Biochem Biophys Res Commun, 200, 408e416. Murugesan, G., Sa, G., & Fox, P. L. (1994). High-density lipoprotein stimulates endothelial cell movement by a mechanism distinct from basic fibroblast growth factor. Circ Res, 74, 1149e1156. Musanti, R., & Ghiselli, G. (1993). Interaction of oxidized HDLs with J774-A1 macrophages causes intracellular accumulation of unesterified cholesterol. Arterioscler Thromb, 13, 1334e1345. Nagano, Y., Arai, H., & Kita, T. (1991). High density lipoprotein loses its effect to stimulate efflux of cholesterol from foam cells after oxidative modification. Proc Natl Acad Sci USA, 88, 6457e6461. Nakajima, T. (1998). Evaluation of the oxidized HDL-specific monoclonal antibody 9F5-3a epitope and identification of possible oxidized HDL receptor in atherosclerotic lesions. J Saitama Med School, 25, 255e266. Nakajima, T., Matsunaga, T., Kawai, S., et al. (2004). Characterization of the epitopes specific for the monoclonal antibody 9F5-3a and quantification of oxidized HDL in human plasma. Ann Clin Biochem, 41, 309e315.
Chapter j 11 Function of oxidized high-density lipoprotein
237
Nakajima, T., Sakagishi, Y., Katahira, T., et al. (1995). Characterization of a specific monoclonal antibody 9F5-3a and the development of assay system for oxidized HDL. Biochem Biophys Res Commun, 217, 407e411. Nakajima, T., Origuchi, N., Matsunaga, T., et al. (2000). Localization of oxidized HDL in atheromatous plaques and oxidized HDL binding sites on human aortic endothelial cells. Ann Clin Biochem, 37, 179e186. Natarajan, R., & Nadler, J. L. (2004). Lipid inflammatory mediators in diabetic vascular disease. Arterioscler Thromb Vasc Biol, 24, 1542e1548. Navab, M., Berliner, J. A., Subbanagounder, G., et al. (2001). HDL and the inflammatory response induced by LDL-derived oxidized phospholipids. Arterioscler Thromb Vasc Biol, 21, 481e488. Navab, M., Hama, S. Y., Anantharamaiah, G. M., et al. (2000a). Normal high density lipoprotein inhibits three steps in the formation of mildly oxidized low density lipoprotein: steps 2 and 3. J Lipid Res, 41, 1495e1508. Navab, M., Hama, S. Y., Cooke, C. J., et al. (2000b). Normal high density lipoprotein inhibits three steps in the formation of mildly oxidized low density lipoprotein: step 1. J Lipid Res, 41, 1481e1494. Neuzil, J., Thomas, S. R., & Stocker, R. (1997). Requirement for, promotion, or inhibition by alpha-tocopherol of radical-induced initiation of plasma lipoprotein lipid peroxidation. Free Radic Biol Med, 22, 57e71. Nguyen, S. D., & Sok, D. E. (2003). Oxidative inactivation of paraoxonase1, an antioxidant protein and its effect on antioxidant action. Free Radic Res, 37, 1319e1330. Nguyen, S. D., Kim, J. R., Kim, M. R., Jung, T. S., & Soka, D. E. (2004). Copper ions and hypochlorite are mainly responsible for oxidative inactivation of paraoxon-hydrolyzing activity in human high density lipoprotein. Toxicol Lett, 147, 201e208. Nofer, J. R., Geigenmuller, S., Gopfert, C., Assmann, G., Buddecke, E., & Schmidt, A. (2003). High density lipoprotein-associated lysosphingolipids reduce E-selectin expression in human endothelial cells. Biochem Biophys Res Commun, 310, 98e103. Nofer, J. R., Kehrel, B., Fobker, M., Levkau, B., Assmann, G., & von Eckardstein, A. (2002). HDL and arteriosclerosis: beyond reverse cholesterol transport. Atherosclerosis, 161, 1e16. Nofer, J. R., Levkau, B., Wolinska, I., et al. (2001). Suppression of endothelial cell apoptosis by high density lipoproteins (HDL) and HDL-associated lysosphingolipids. J Biol Chem, 276, 34480e34485. Nofer, J. R., van der Giet, M., Tolle, M., et al. (2004). HDL induces NO-dependent vasorelaxation via the lysophospholipid receptor S1P3. J Clin Invest, 113, 569e581. Norata, G. D., Banfi, C., Pirillo, A., et al. (2004). Oxidised-HDL3 induces the expression of PAI-1 in human endothelial cells. Role of p38MAPK activation and mRNA stabilization. Br J Haematol, 127, 97e104. Nuszkowski, A., Gra¨bner, R., Marsche, G., Unbehaun, A., Malle, E., & Heller, R. (2001). Hypochlorite-modified low density lipoprotein inhibits nitric oxide synthesis in endothelial cells via an intracellular dislocalization of endothelial nitric-oxide synthase. J Biol Chem, 276, 14212e14221. O’Connell, B. J., & Genest, J., Jr. (2001). High-density lipoproteins and endothelial function. Circulation, 104, 1978e1983. O’Neil, J., Hoppe, G., Sayre, L. M., & Hoff, H. F. (1997). Inactivation of cathepsin B by oxidized LDL involves complex formation induced by binding of putative reactive sites exposed at low pH to thiols on the enzyme. Free Radic Biol Med, 23, 215e225.
238
The HDL Handbook
Oe, T., Lee, S. H., Silva Elipe, M. V., Arison, B. H., & Blair, I. A. (2003). A novel lipid hydroperoxide-derived modification to arginine. Chem Res Toxicol, 16, 1598e1605. Ohmura, H., Watanabe, Y., Hatsumi, C., et al. (1999). Possible role of high susceptibility of highdensity lipoprotein to lipid peroxidative modification and oxidized high-density lipoprotein in genesis of coronary artery spasm. Atherosclerosis, 142, 179e184. Ohta, T., Takata, K., Horiuchi, S., Morino, Y., & Matsuda, I. (1989). Protective effect of lipoproteins containing apolipoprotein A-I on Cu2þ-catalyzed oxidation of human low density lipoprotein. FEBS Lett, 257, 435e438. Palinski, W., Yla-Herttuala, S., Rosenfeld, M. E., et al. (1990). Antisera and monoclonal antibodies specific for epitopes generated during oxidative modification of low density lipoprotein. Arteriosclerosis, 10, 325e335. Pankhurst, G., Wang, X. L., Wilcken, D. E., et al. (2003). Characterization of specifically oxidized apolipoproteins in mildly oxidized high density lipoprotein. J Lipid Res, 44, 349e355. Pannala, A. S., Rice-Evans, C., Sampson, J., & Singh, S. (1998). Interaction of peroxynitrite with carotenoids and tocopherols within low density lipoprotein. FEBS Lett, 423, 297e301. Panzenboeck, U., Raitmayer, S., Reicher, H., et al. (1997). Effects of reagent and enzymatically generated hypochlorite on physicochemical and metabolic properties of high density lipoproteins. J Biol Chem, 272, 29711e29720. Parthasarathy, S., Wieland, E., & Steinberg, D. (1989). A role for endothelial cell lipoxygenase in the oxidative modification of low density lipoprotein. Proc Natl Acad Sci USA, 86, 1046e1050. Parthasarathy, S., Barnett, J., & Fong, L. G. (1990). High-density lipoprotein inhibits the oxidative modification of low-density lipoprotein. Biochim Biophys Acta, 1044, 275e283. Pennathur, S., Bergt, C., Shao, B., et al. (2004). Human atherosclerotic intima and blood of patients with established coronary artery disease contain high density lipoprotein damaged by reactive nitrogen species. J Biol Chem, 279, 42977e42983. Rader, D. J. (2003). Regulation of reverse cholesterol transport and clinical implications. Am J Cardiol, 92, 42Je49J. Rankin, S. M., Parthasarathy, S., & Steinberg, D. (1991). Evidence for a dominant role of lipoxygenase(s) in the oxidation of LDL by mouse peritoneal macrophages. J Lipid Res, 32, 449e456. Reichl, D., & Miller, N. E. (1989). Pathophysiology of reverse cholesterol transport. Insights from inherited disorders of lipoprotein metabolism. Arteriosclerosis, 9, 785e797. Requena, J. R., Fu, M. X., Ahmed, M. U., et al. (1997). Quantification of malondialdehyde and 4-hydroxynonenal adducts to lysine residues in native and oxidized human low-density lipoprotein. Biochem J, 322, 317e325. Reyftmann, J. P., Santus, R., Mazie`re, J. C., et al. (1990). Sensitivity of tryptophan and related compounds to oxidation induced by lipid autoperoxidation. Application to human serum lowand high-density lipoproteins. Biochim Biophys Acta, 1042, 159e167. Rohrer, L., Hersberger, M., & von Eckardstein, A. (2004). High density lipoproteins in the intersection of diabetes mellitus, inflammation and cardiovascular disease. Curr Opin Lipidol, 15, 269e278. Rothblat, G. H., Mahlberg, F. H., Johnson, W. J., & Phillips, M. C. (1992). Apolipoproteins, membrane cholesterol domains, and the regulation of cholesterol efflux. J Lipid Res, 33, 1091e1097. Rubin, E. M., Krauss, R. M., Spangler, E. A., Verstuyft, J. G., & Clift, S. M. (1991). Inhibition of early atherogenesis in transgenic mice by human apolipoprotein AI. Nature, 353, 265e267.
Chapter j 11 Function of oxidized high-density lipoprotein
239
Sakai, M., Miyazaki, A., Sakamoto, Y., Shichiri, M., & Horiuchi, S. (1992). Cross-linking of apolipoproteins is involved in a loss of the ligand activity of high density lipoprotein upon Cu (2þ)-mediated oxidation. FEBS Lett, 314, 199e202. Saku, K., Ahmad, M., Glas-Greenwalt, P., & Kashyap, M. L. (1985). Activation of fibrinolysis by apolipoproteins of high density lipoproteins in man. Thromb Res, 39, 1e8. Salmon, S., Mazie`re, C., Auclair, M., Theron, L., Santus, R., & Mazie`re, J. C. (1992a). Malondialdehyde modification and copper-induced autooxidation of high-density lipoprotein decrease cholesterol efflux from human cultured fibroblasts. Biochim Biophys Acta, 1125, 230e235. Salmon, S., Santus, R., Mazie`re, J. C., et al. (1992b). Modified apolipoprotein pattern after irradiation of human high-density lipoproteins by ultraviolet B. Biochim Biophys Acta, 1128, 167e173. Sattler, W., & Stocker, R. (1993). Greater selective uptake by Hep G2 cells of high-density lipoprotein cholesteryl ester hydroperoxides than of unoxidized cholesteryl esters. Biochem J, 294, 771e778. Schroeder, F., & Goh, E. H. (1979). Regulation of very low density lipoprotein interior core lipid physicochemical properties. J Biol Chem, 254, 2464e2470. Shao, B., Bergt, C., Fu, X., et al. (2005). Tyrosine 192 in apolipoprotein A-I is the major site of nitration and chlorination by myeloperoxidase, but only chlorination markedly impairs ABCA1-dependent cholesterol transport. J Biol Chem, 280, 5983e5993. Sharma, N., Desigan, B., Ghosh, S., Sanyal, S. N., Ganguly, N. K., & Majumdar, S. (1999). The role of oxidized HDL in monocyte/macrophage functions in the pathogenesis of atherosclerosis in Rhesus monkeys. Scand J Clin Lab Invest, 59, 215e225. Shih, D. M., Gu, L., Xia, Y. R., Navab, M., et al. (1998). Mice lacking serum paraoxonase are susceptible to organophosphate toxicity and atherosclerosis. Nature, 394, 284e287. Shoukry, M. I., Gong, E. L., & Nichols, A. V. (1994). Apolipoprotein-lipid association in oxidatively modified HDL and LDL. Biochim Biophys Acta, 1210, 355e360. Simon, B. C., Cunningham, L. D., & Cohen, R. A. (1990). Oxidized low density lipoproteins cause contraction and inhibit endothelium-dependent relaxation in the pig coronary artery. J Clin Invest, 86, 75e79. Singh, I. P., & Baron, S. (2000). Innate defences against viremia. Rev Med Virol, 10, 395e403. Sloop, C. H., Dory, L., & Roheim, P. S. (1987). Interstitial fluid lipoproteins. J Lipid Res, 28, 225e237. Sola, R., Motta, C., Maille, M., et al. (1993). Dietary monounsaturated fatty acids enhance cholesterol efflux from human fibroblasts. Relation to fluidity, phospholipid fatty acid composition, overall composition, and size of HDL3. Arterioscler Thromb, 13, 958e966. Sorescu, D., Szocs, K., & Griendling, K. K. (2001). NAD(P)H oxidases and their relevance to atherosclerosis. Trends Cardiovasc Med, 11, 124e131. Steinberg, D. (1978). The rediscovery of high density lipoprotein: a negative risk factor in atherosclerosis. Eur J Clin Invest, 8, 107e109. Steinberg, D., Parthasarathy, S., Carew, T. E., Khoo, J. C., & Witztum, J. L. (1989). Beyond cholesterol. Modifications of low-density lipoprotein that increase its atherogenicity. N Engl J Med, 320, 915e924. Steinbrecher, U. P., Lougheed, M., Kwan, W. C., & Dirks, M. (1989). Recognition of oxidized low density lipoprotein by the scavenger receptor of macrophages results from derivatization of apolipoprotein B by products of fatty acid peroxidation. J Biol Chem, 264, 15216e15223. Stojanovic, N., Krilov, D., & Herak, J. N. (2006). Slow oxidation of high density lipoproteins as studied by EPR spectroscopy. Free Radic Res, 40, 135e140.
240
The HDL Handbook
Takahashi, Y., Chiba, H., Matsuno, K., et al. (1996). Native lipoproteins inhibit platelet activation induced by oxidized lipoproteins. Biochem Biophys Res Commun, 222, 453e459. Tall, A. R. (1990). Plasma high density lipoproteins. Metabolism and relationship to atherogenesis. J Clin Invest, 86, 379e384. Tauber, J. P., Cheng, J., & Gospodarowicz, D. (1980). Effect of high and low density lipoproteins on proliferation of cultured bovine vascular endothelial cells. J Clin Invest, 66, 696e708. Thorne, R. F., Mhaidat, N. M., Ralston, K. J., & Burns, G. F. (2007). CD36 is a receptor for oxidized high density lipoprotein: implications for the development of atherosclerosis. FEBS Lett, 581, 1227e1232. Tsumura, M., Kinouchi, T., Ono, S., Nakajima, T., & Komoda, T. (2001). Serum lipid metabolism abnormalities and change in lipoprotein contents in patients with advanced-stage renal disease. Clin Chim Acta, 314, 27e37. Tward, A., Xia, Y. R., Wang, X. P., et al. (2002). Decreased atherosclerotic lesion formation in human serum paraoxonase transgenic mice. Circulation, 106, 484e490. Uchida, K., Toyokuni, S., Nishikawa, K., et al. (1994). Michael addition-type 4-hydroxy-2-nonenal adducts in modified low-density lipoproteins: markers for atherosclerosis. Biochemistry, 33, 12487e12494. Uittenbogaard, A., Shaul, P. W., Yuhanna, I. S., Blair, A., & Smart, E. J. (2000). High density lipoprotein prevents oxidized low density lipoprotein-induced inhibition of endothelial nitricoxide synthase localization and activation in caveolae. J Biol Chem, 275, 11278e11283. Van Lenten, B. J., Wagner, A. C., Nayak, D. P., Hama, S., Navab, M., & Fogelman, A. M. (2001). High-density lipoprotein loses its anti-inflammatory properties during acute influenza A infection. Circulation, 103, 2283e2288. Vila, L. (2004). Cyclooxygenase and 5-lipoxygenase pathways in the vessel wall: Role in atherosclerosis. Med Res Rev, 24, 399e424. Vohra, R. S., Murphy, J. E., Walker, J. H., Ponnambalam, S., & Homer-Vanniasinkam, S. (2006). Atherosclerosis and the Lectin-like OXidized low-density lipoprotein scavenger receptor. Trends Cardiovasc Med, 16, 60e64. Vollmer, E., Brust, J., Roessner, A., et al. (1991). Distribution patterns of apolipoproteins A1, A2, and B in the wall of atherosclerotic vessels. Virchows Arch A Pathol Anat Histopathol, 419, 79e88. Wang, W. Q., Merriam, D. L., Moses, A. S., & Francis, G. A. (1998). Enhanced cholesterol efflux by tyrosyl radical-oxidized high density lipoprotein is mediated by apolipoprotein AI-AII heterodimers. J Biol Chem, 273, 17391e17398. Watson, A. D., Berliner, J. A., Hama, S. Y., et al. (1995). Protective effect of high density lipoprotein associated paraoxonase. Inhibition of the biological activity of minimally oxidized low density lipoprotein. J Clin Invest, 96, 2882e2891. Weiss, S. J., Klein, R., Slivka, A., & Wei, M. (1982). Chlorination of taurine by human neutrophils. Evidence for hypochlorous acid generation. J Clin Invest, 70, 598e607. Yim, M. B., Kang, S. O., & Chock, P. B. (2000). Enzyme-like activity of glycated cross-linked proteins in free radical generation. Ann NY Acad Sci, 899, 168e181. Yuhanna, I. S., Zhu, Y., Cox, B. E., et al. (2001). High-density lipoprotein binding to scavenger receptor-BI activates endothelial nitric oxide synthase. Nat Med, 7, 853e857. Yui, Y., Aoyama, T., Morishita, H., Takahashi, M., Takatsu, Y., & Kawai, C. (1988). Serum prostacyclin stabilizing factor is identical to apolipoprotein A-I (ApoAI). A novel function of ApoAI. J Clin Invest, 82, 803e807. Zeiher, A. M., Scha¨chlinger, V., Hohnloser, S. H., Saurbier, B., & Just, H. (1994). Coronary atherosclerotic wall thickening and vascular reactivity in humans. Elevated high-density
Chapter j 11 Function of oxidized high-density lipoprotein
241
lipoprotein levels ameliorate abnormal vasoconstriction in early atherosclerosis. Circulation, 89, 2525e2532. Zhang, H. F., Davis, W. B., Chen, X. S., Whisler, R. L., & Cornwell, D. G. (1989). Studies on oxidized low density lipoproteins. Controlled oxidation and a prostaglandin artifact. J Lipid Res, 30, 141e148. Zhang, R., Brennan, M. L., Shen, Z., et al. (2002). Myeloperoxidase functions as a major enzymatic catalyst for initiation of lipid peroxidation at sites of inflammation. J Biol Chem, 277, 46116e46122. Zhao, L., Cuff, C. A., Moss, E., et al. (2002). Selective interleukin-12 synthesis defect in 12/15lipoxygenase deficient macrophages associated with reduced atherosclerosis in a mouse model of familial hypercholesterolemia. J Biol Chem, 277, 35350e35356. Zhou, H., Tan, K. C., Shiu, S. W., & Wong, Y. (2008). Increased serum advanced glycation end products are associated with impairment in HDL antioxidative capacity in diabetic nephropathy. Nephrol Dial Transplant, 23, 927e933.
Chapter 12
Preb1-HDL, a Native Lipidpoor HDL, and its Potential as a New Marker for HDL Metabolism Takashi Miida and Satoshi Hirayama Department of Laboratory Medicine, Juntendo University School of Medicine, Tokyo, Japan
INTRODUCTION High density lipoprotein (HDL) has a density of 1.063 to 1.210, as determined by ultracentrifugation. Numerous epidemiological studies have shown that low HDL-cholesterol concentration is a strong predictor of coronary heart disease (CHD) (Miller et al., 1977; Gordon et al., 1989; Kitamura et al., 1994; Gordon, 1999; Foody et al., 2000). Although reducing the plasma concentration of low density lipoprotein-cholesterol (LDL-C) is the main therapeutic goal in the management of dyslipidemia (Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults., 2001; De Backer et al., 2003; Teramoto et al., 2007), recent meta-analyses have clearly shown that statins (the strongest LDL-C-lowering agents) reduce the risk of cardiovascular events by only 20e30% (Zhou et al., 2006). Moreover, the risk of CHD was shown to be negatively correlated with HDL-cholesterol concentration even in patients treated with statins (Mabuchi et al., 2002). Experimental studies strongly suggest that the anti-atherogenic activity of HDL centers mainly on the removal of excess cholesterol from atherosclerotic lesions. The removed cholesterol is then transported to the liver and excreted into the bile by a process known as reverse cholesterol transport (Gomaraschi et al., 2006). In addition, HDL particles have anti-inflammatory, antioxidant, and anti-thrombotic effects. The compositions of lipids and apolipoproteins, particle size distribution, and electrophoretic mobility vary among HDL particles, and HDL particles exhibit different functions. Thus, HDL particles are classified into several The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
243
244
The HDL Handbook
subfractions according to their physical and functional properties (Barbaras et al., 1987; Lewis and Cabana, 1996; Miyazaki et al., 2000; Okazaki et al., 2005). Preb1-HDL is a unique lipid-poor HDL subfraction that is closely associated with cholesterol efflux from cellular membranes (Miida et al., 1990, 1992; Fielding et al., 1991; Kawano et al., 1993). Over the past two decades, much has been learned about the physiological functions and clinical significance of preb1-HDL (Miida et al., 1996, 1997, 1998, 2000a,b, 2003a, 2004a,b; Hirayama et al., 2007). In this chapter, we provide a brief overview of preb1HDL and discuss recent advances in the field.
CHARACTERISTICS OF HDL HDL is a technical term referring to lipoproteins exhibiting a density of 1.063e1.210 with ultracentrifugation. Given that all other known lipoproteins, including chylomicrons, very low density lipoprotein (VLDL), intermediate density lipoprotein (IDL), and LDL, contain apolipoprotein B (apoB: that is, apoB48 or B100), HDL may be defined as a lipoprotein that does not contain apoB. In patients with cholesteryl ester transfer protein (CETP) deficiency or cholestatic liver disease, large, apoE-rich lipoproteins without apoB accumulate. They are called apoE-rich HDL, even though their density range overlaps those of LDL and apoE-poor HDL. In this chapter, we describe the nature of HDL in detail.
Lipid and protein constituents As in the other classes of lipoproteins, HDL has a hydrophobic lipid core containing a cholesteryl ester (CE) and a triglyceride (TG) and is covered with a phospholipid monolayer. A small amount of free cholesterol also exists on the surface of HDL. Clinically, the cholesterol concentration associated with HDL particles is used as a marker for HDL and is referred to as HDL-cholesterol (HDL-C). However, cholesterol accounts for only 15e20% of the total weight of an HDL particle, while phospholipids and HDL-associated proteins account for 25e30% and 50%, respectively. Nascent HDL particles, which are secreted into the blood from the liver, consist mainly of phospholipids, apoA-I, and a little cholesterol, but no core lipids. Thus, the measurement of HDL-C concentration is unlikely to reflect the amount of lipid-poor HDL. HDL particles carry enzymes and transfer proteins as well as apolipoproteins (Table 12.1). The proteins associated with HDL particles reflect a heterogeneous distribution (Francone et al., 1989). For example, some HDL particles are associated with lecithin-cholesterol acyltransferase (LCAT), and others are not (Francone et al., 1989; Miida et al., 2004b). Unlike apoBcontaining particles such as LDL, some HDL particles contain weakly bound proteins that move to or from other lipoproteins during HDL metabolism. In
Chapter j 12 Preb1-HDL
TABLE 12.1
245
Lipid and protein constituents of HDL particles
HDL component Constituent
Function
Protein moiety (a) Apolipoproteins ApoA-I* ApoA-II*
Activation of LCAT Inhibition of LCAT? Activation of HL? ApoA-IV Activation of LCAT? ApoC-I Activation of LCAT? Inhibition of CETP? Inhibition of HDL binding mediated by apoE-recognizing receptors? ApoC-II Inhibition of LCAT by dissociating apoA-I? ApoC-III Inhibition of LCAT by dissociating apoA-I? ApoD Binding to lipids and their transport? ApoE Binding to lipoprotein receptors? ApoJ Binding to PON1 ApoM Preb1-HDL generation from b-HDL Serum amyloid A protein (SAA) Dissociation of preb1-HDL
(b) Enzymes
Lecithin-cholesterol acyltransferase (LCAT) Paraoxonase-1 (PON1)
Conversion of FC to CE
Cholesteryl ester transfer protein (CETP)
Exchange of CE on HDL for TG on apoB-containing lipoproteins Transfer of PL on TG-rich lipoproteins Fusion of HDL particles and simultaneous dissociation of preb1-HDL
Inhibition of LDL and HDL oxidation Platelet activating Degradation of PAF and factor acetyl hydrolase (PAF-AH) oxidized PL
(c) Transfer proteins
Phospholipid transfer protein (PLTP) Lipid moiety Cholesterol (free and esterified forms) Triglyceride (TG) Phospholipids (PL) HL, hepatic lipase; FC, free cholesterol; CE, cholesteryl ester * major apolipoproteins of HDL
246
The HDL Handbook
some cases, exogenous proteins dissociate the protein components from HDL particles (Lagocki and Scanu, 1980; Miida et al., 1999a, 2006).
Particle size On average, HDL particles are 7e10 nm in diameter, which is nearly equal to the thickness of a cell membrane and is about 0.1 times the diameter of an influenza virus. As lipoproteins less than 70 nm in diameter can penetrate vascular walls (Mamo et al., 1998), HDL particles can easily access atherosclerotic lesions in the subendothelial space. It is of great interest that HDL particles are detectable not only in blood but also in other body fluids such as lymph (Nanjee et al., 2001), follicular fluid (Jaspard et al., 1997), aqueous humor (Cenedella, 1984), and cerebrospinal fluid (Roheim et al., 1979; Miida et al., 1999b, 2006; Kay et al., 2003). The HDL in cerebrospinal fluid has been extensively investigated by many researchers, and cerebrospinal HDL is believed to play a crucial role in cholesterol recycling (Dietschy and Turley, 2001). Surprisingly, brain cholesterol has an extremely long half-life of about 5 years (Bjo¨rkhem et al., 1998).
HDL SUBFRACTIONS HDL is composed of several different particles, known as HDL subfractions; however, the clinical significance of these subfractions as predictors of atherosclerotic disorders is controversial.
Analytical and separation methods HDL particles vary in density, electrical charge, apolipoprotein composition, lipid composition, particle size, and physiological function. According to these differences, HDL particles can be separated into several subfractions using various analytical and separation methods, including ultracentrifugation (Lewis and Cabana, 1996), gel filtration chromatography (Kay et al., 2003), high performance liquid chromatography (Okazaki et al., 2005), agarose gel electrophoresis (Peynet et al., 1986), gradient polyacrylamide gel electrophoresis, native two-dimensional gel (2-D gel) electrophoresis (Francone et al., 1989; Miida et al., 1990, 1992, 1996, 1997, 1998, 1999a, 2000a,b, 2003a, 2004b, 2006; Fielding et al., 1991; Kawano et al., 1993; Jaspard et al., 1997), capillary electrophoresis (Inano et al., 2000; Zhang et al., 2006), immunoaffinity chromatography (Barbaras et al., 1987), enzyme-linked immunoassays (ELISAs) using monoclonal antibodies (Miyazaki et al., 2000; Miida et al., 2003a,b, 2004a, 2006; Hirayama et al., 2007), and crossed immunoelectrophoresis (Neary et al., 1991). However, these methods present several problems. For example, ultracentrifugation exposes HDL subfractions to harsh conditions, including extreme gravity and salinity (Lewis and Cabana, 1996). Furthermore,
Chapter j 12 Preb1-HDL
247
some HDL subfractions are rendered unstable by storage at 4 C or even 80 C (Miida et al., 2003b). We have analyzed the HDL subfractions in fresh plasma by native 2-D gel electrophoresis (Miida et al., 1990, 1992, 1996, 1997, 1998, 1999a; 2000a,b, 2003a, 2004a,b, 2006; Fielding et al., 1991; Kawano et al., 1993). Although this method minimizes changes in the distribution of the subfractions during separation, it is time-consuming and costly, and considerable technical proficiency is required to obtain reproducible results. In addition, it is difficult to run several samples simultaneously. Our system uses a 0.75% agarose gel in the first dimension and a 2e15% gradient polyacrylamide gel in the second dimension. Unlike most 2-D gel systems, ours does not use sodium dodecyl sulfate in either the polyacrylamide gel or the electrophoresis buffer solution.
What is preb1-HDL? In agarose gel electrophoresis, plasma lipoproteins are separated by electric charge. Lipoproteins with greater negative charge will migrate further from the point of origin. LDL, VLDL, and HDL migrate with b-, preb-, and a-mobility, respectively. In agarose gels stained with lipophilic dyes (e.g., Sudan Black B or Fat Red O), HDL is detectable as an a-lipoprotein. To detect lipid-poor HDL subfractions, the separated lipoproteins are transferred to a nitrocellulose membrane and visualized with anti-apoA-I antibodies. In addition to the a-migrating HDL, there is a minor apoA-I band with preb-mobility, which is known as preb-HDL (Kunitake et al., 1985). Preb-HDL is an LpA-I particle containing apoA-I but not apoA-II (Castro and Fielding, 1988). Preb-HDL may be separated further into three subspecies (preb1-HDL, preb2-HDL, and preb3HDL) by native 2-D gel electrophoresis (Figure 12.1) (Francone et al., 1989). The estimated molecular weight of preb1-HDL is 60e70 kDa (Castro and Fielding, 1988; Miida et al., 2000a). Preb1-HDL consists mainly of apoAI and phospholipids (Castro and Fielding, 1988). The sphingomyelin/phosphatidyl choline ratio of preb1-HDL is very similar to that of the outer leaflet of cellular membranes (Miida T and Fielding PE. unpublished data). Preb1-HDL has almost no core lipids and little free cholesterol. The lipoproteins in fresh plasma were first separated on a 0.75% agarose gel (first dimension). Then, a longitudinal piece of the gel was run perpendicularly on a 2e15% gradient polyacrylamide gel without detergent (second dimension). After electrical transfer to a nitrocellulose sheet, the HDL subfractions were visualized by Western blotting with 125I-labeled anti-human apoAI antibody.
Physiological role and metabolism of preb1-HDL Studies have shown that preb1-HDL is closely associated with the release of cellular cholesterol into plasma (i.e., the initial step of reverse cholesterol
248
FIGURE 12.1
The HDL Handbook
HDL subfractions separated by native two-dimensional gel electrophoresis.
transport). When plasma is incubated with 3H-cholesterol-labeled cells, the radiolabel moves from the membranes to the plasma as a function of time. This is called cholesterol efflux and occurs via two independent pathways (Kawano et al., 1993; Yokoyama, 2006): a specific lipid-free, or lipid-poor, apolipoproteinmediated pathway; and a non-specific diffusion-mediated pathway. The former requires the specific interaction of free apoA-I or lipid-poor HDL with cellular membranes, and the latter involves various extracellular acceptors such as albumin and lipoproteins in the aqueous phase. In cultured fibroblasts, the magnitude of cholesterol efflux is proportional to the plasma preb1-HDL concentration. Preb1-HDL-dependent cholesterol efflux accounts for about 60% of the total cholesterol efflux (Kawano et al., 1993). Cholesterol efflux can be reduced to half the original value by the addition of monoclonal anti-preb1-HDL antibodies or by the proteolytic treatment of cultured cells (Kawano et al., 1993; Fielding et al., 1994). Among the many HDL subfractions, cellular cholesterol is first transferred preferentially to preb1-HDL and then to preb2-HDL (Castro and Fielding, 1988). Ultimately, cellular cholesterol is moved to preb3-HDL, the largest preb-HDL, where LCAT is located (Francone et al., 1989). LCAT converts free cholesterol to CE, a hydrophobic compound that enters the core of HDL particles. Finally, preb-HDL is converted into large, CE-rich, spherical HDL particles with a-mobility upon agarose gel electrophoresis; this process is referred to as the maturation of preb1-HDL. Although CETP promotes the
Chapter j 12 Preb1-HDL
249
bidirectional transfer of CEs and TGs between lipoproteins (Moulin et al., 1992), CETP activity results in the net transfer of CEs from a-HDL to VLDL and LDL and of TGs in the reverse direction, owing to the CE and TG concentration gradients between lipoproteins. CEs carried by VLDL and LDL are transported to the liver where they are taken up via LDL receptors for further metabolism. Meanwhile, the TG-enriched a-migrating HDL is susceptible to hydrolysis by hepatic lipase, thereby regenerating preb1-HDL.
Preb1-HDL production Recent studies have revealed that preb1-HDL is generated by ATP-binding cassette transporter A-1 (ABCA1)-dependent and ABCA1-independent mechanisms.
ABCA1-dependent mechanism The main source of plasma preb1-HDL is probably direct secretion from the liver. ABCA1 plays a crucial role in the lipidation of apoA-I with phospholipids and in the subsequent efflux of cholesterol from cellular membranes; however, there is considerable debate over its role in preb1-HDL formation. Preb1-HDL is not free apoA-I but apoA-I coupled with phospholipids and a little free cholesterol (Castro and Fielding, 1988). In vitro studies suggest that phospholipids are added to the apoA-I secreted from hepatocytes via ABCA1dependent and -independent mechanisms (Kiss et al., 2003). This is consistent with the fact that patients with Tangier disease have preb1-HDL despite a lack of ABCA1 activity (Huang et al., 1995; Asztalos et al., 2001). Recent data strongly suggest that the production of biologically active preb1-HDL involves at least two steps and that ABCA1 activity is closely related to preb1-HDL synthesis. In J774 macrophages, ABCA1 activity creates two types of high-affinity apoA-I binding sites at the cell surface (Vedhachalam et al., 2007). One is a low-capacity site formed by the direct interaction of apoA-I with ABCA1, and the other is a high-capacity site secondarily induced by the interaction of apoA-I with membrane lipids. In support of the multistep hypothesis, two C-terminal deletion mutants of apoA-I (D190-243 and D223243) can be cross-linked to ABCA1, although this significantly decreases their binding affinity for the cell surface (Vedhachalam et al., 2007). It has also been shown that C-terminal deletions in apoA-I (D185-243 and D220-243) do not prevent preb1-HDL formation but greatly reduce ABCA1-mediated cholesterol efflux (Chroni et al., 2007). A separate study examined the apoA-I secreted from HepG2 cells, Caco cells, and CHO cells expressing human apoA-I and found that apoA-I was first secreted as particles with preb-mobility and a Stokes radius of 2.6 nm (Chau et al., 2006). With additional incubation, however, 3.6-nm particles with preb-mobility became predominant. The conversion from 2.6-nm to 3.6-nm particles required ABCA1 activity, and only
250
The HDL Handbook
the 3.6-nm particles were able to bind phospholipids. When the 3.6-nm particles were incubated with smooth muscle cells, cholesterol efflux was increased to 28-fold that produced by phospholipids. These data strongly support the hypothesis that biologically active preb1-HDL is generated by ABCA1 activity and that preb1-HDL promotes cholesterol efflux from cells.
ABCA1-independent mechanisms In addition to the ABCA1-dependent mechanism described above, preb1-HDL is also generated from an a-migrating spherical HDL by ABCA1-independent mechanisms. For example, a 2-h incubation of TG-enriched HDL2 with rat hepatic lipase yielded preb1-HDL, as determined by native 2-D gel analysis (58). Preb1-HDL can be also formed from a-HDL during its hydrolysis by bacterial TG lipase (Miida et al., 2000a). Lipase-dependent shedding of preb1HDL from a-HDL is likely to occur in vivo, because preb1-HDL was formed during recycling rat liver perfusion in the presence of heparin (Barrans et al., 1994). Other studies suggest that preb1-HDL is also formed by lipoprotein lipase (LPL). Neary et al. (1991) incubated plasma samples, obtained before and after the injection of heparin, at 37 C and determined the total preb-HDL concentration by crossed immunoelectrophoresis. When LCAT was inactivated by p-chloromercuriphenylsulfonic acid, total preb-HDL was increased by a much greater amount in the post-heparin plasma than in the pre-heparin plasma. This increase was especially marked in four hypertriglyceridemic subjects following a fat load, suggesting that preb1-HDL is generated from TG-rich lipoproteins during lipolysis. Furthermore, Sviridov et al. (2003) examined preb1-HDL formation in skeletal muscle during physical exercise by simultaneously measuring the concentration of preb1-HDL in the femoral artery and vein before and after 25 min of cycling. They found that the preb1-HDL concentration was higher in venous blood than in arterial blood, independent of exercise, and that preb1-HDL production (calculated as the difference in preb1-HDL concentration between venous and arterial blood) increased 6.6-fold after exercise. Another report showed a 30% increase in the plasma preb1-HDL concentration after cardiopulmonary exercise followed by 4 km of jogging (Jafari et al., 2003). These results agree with those of another study showing that the plasma preb1-HDL concentration was 46% greater in athletes than in control subjects (Olchawa et al., 2004). Although it is unclear whether preb1-HDL is generated from TG-rich lipoproteins or a-migrating HDL, LPL is likely to enhance preb1-HDL formation during the passage of blood through skeletal muscle. Even without lipase activity, preb1-HDL can be generated from a-HDL by exogenous apolipoproteins, which have greater affinity for HDL particles than does apoA-I. When serum amyloid A protein (SAA) was added to plasma, the preb1-HDL concentration increased in a dose-dependent manner (Miida et al., 1999a). Dose-dependent preb1-HDL formation was also observed when SAA
Chapter j 12 Preb1-HDL
251
was added to cerebrospinal fluid (Miida et al., 2006). This phenomenon results from the displacement of apoA-I and phospholipids on HDL particles by SAA. In fact, preb1-HDL production occurred immediately when SAA was added to ultracentrifugally isolated HDL coupled to agarose beads (Miida et al., 1999a). Recent studies suggest that preb1-HDL is also formed by a mechanism involving apolipoprotein M (apoM) (Wolfrum et al., 2005). ApoM was recently identified in human plasma and is also found in the liver and kidney. ApoM is synthesized by hepatocytes and proximal tubule cells and is secreted with a hydrophobic signal peptide. Using this signal peptide as an anchor, apoM binds to the phospholipid bilayers of plasma lipoproteins such as HDL (Christoffersen et al., 2006). In apoM knockout mice, large HDL particles accumulated, and no preb-HDL could be detected (Wolfrum et al., 2005). ApoM promotes the formation of preb1-HDL, resulting in increased cholesterol efflux in cultured cells.
PLASMA PREb1-HDL CONCENTRATION IN VARIOUS DISORDERS Several research groups have already reported the plasma preb1-HDL concentration in healthy subjects based on native 2-D gel analysis followed by immunoblotting (Miida et al., 1996, 1998, 2000a,b; Sasahara et al., 1997; Asztalos et al., 2000, 2001, 2004), ELISA (Miyazaki et al., 2000; Miida et al., 2003a, 2004a; Sviridov et al., 2003, 2006; Olchawa et al., 2004; Hirayama et al., 2007), and an ultrafiltration-isotope dilution method (O’Connor et al., 1998; Jafari et al., 2003). However, there is significant variation in the measured values between the different methods, and the same research group reported considerably diverse results (Table 12.2). Such discrepancies may be the result of preb1-HDL instability during storage (Miida et al., 2003b) or differences in the methods used to separate the HDL subfractions. To avoid these complications when analyzing clinical samples, we produced a monoclonal antibody specific for preb1-HDL (Mab55201) and established an ELISA system for preb1-HDL measurement (Miyazaki et al., 2000). Mab55201 reacted with LpA-I but not LpA-II. The gel filtration analysis revealed that Mab55201 had strong reactivity against the lipoprotein fraction with a molecular weight of less than 67 kDa. Preincubation of human plasma with Mab55201 resulted in disappearance of preb1-HDL spots in native 2D-gel analysis (Miyazaki et al., 2000). When plasma was diluted 21-fold with 50% sucrose, the preb1-HDL concentration did not change significantly for at least 5 days at 4 C or for 30 days at 80 C (Miida et al., 2003b). There was no significant difference in the plasma preb1-HDL concentration between men and postmenopausal women (Figure 12.2), although premenopausal women had less preb1-HDL than either of the other two groups (Miida et al., 2004a). It is interesting that the plasma preb1-HDL concentration does not always parallel the HDL-C, apoA-I, and LpA-I concentrations (Miida et al., 1997,
252
The HDL Handbook
TABLE 12.2
Reference ranges of plasma preb1-HDL concentration
Method
Number (M/F)
Relative conc. Absolute conc. (% apoA-I) (mg/L apoA-I)1 Reference
20 (11/9) 12 (3/9) 20 (18/2) 58 (22/36)
4.6 2.3 4.2 1.6
Native 2-D gels
63 28 59 23 73 34 78 39 (M) 88 45 (F) 92 31 79 49 (M) 129 60 (M) 120 85 (F) 120 14 (M)
Miida et al. (1996) Miida et al. (1998) Miida et al. (2000a) Miida et al. (2000b)
Miyazaki et al. (2000) Miida et al. (2003a) Miida et al. (2004a)
1.5 0.4
23 8 20 7 22 7 (M) 19 7(F) 150 33(M) 37 32(M) 25 43 20 6
6.1 3.6 7.2 4.0 (M) 5.5 3.3 (F) 7.9 3.8
73 44 (Total) O’Connor et al. (1997) 68 40 (M) 84 49 (F) 100 60 Jafari et al. (2003)
18 (10/8) 79 (79/0) 71 (30/41) 1277 (1277/0) EIA
25 (15/10) 45 (24/21) 100 (47/53) 7 (7/0) 33 (33/0) 26 (21/6) 30 (13/17)
Ultrafiltration- 136 (46/90) isotope dilution technique 19 (11/8)
Sasahara et al. (1997) Asztalos et al. (2000) Asztalos et al. (2001) Asztalos et al. (2004)
Sviridov et al. (2003) Olchawa et al. (2004) Sviridov et al. (2006) Hirayama et al. (2007)
Absolute concentration is presented as the amount of apoA-I associated with preb1-HDL. Healthy subjects with TG <150 mg/dL and HDL-C of 35e80 mg/dL. 3 Values are expressed as the mean SE. 1 2
2000a). Instead, the preb1-HDL concentration is positively correlated with the concentrations of LDL-cholesterol and TGs (Miida et al., 2000a,b). Previously, we compared the plasma preb1-HDL concentration between patients with angiographically defined coronary artery disease (CAD) and age- and sexmatched controls. We excluded those patients taking lipid-lowering agents. The absolute and relative preb1-HDL concentrations were 38% and 40% greater, respectively, in the CAD group than in the control group (P < 0.05 and P < 0.01, respectively) (Miida et al., 1996). The increase was dominant among those CAD patients exhibiting low LCAT activity. Asztalos et al. (2000) also reported that CAD patients with low HDL-cholesterol levels had high preb1-HDL concentrations. They further examined the clinical significance of preb1-HDL using frozen samples collected for the Veterans Affairs HDL Intervention Trial (VA-HIT) (Asztalos et al., 2005). Those eligible for the VA-HIT were men below 74 years of age with a documented history of CHD, HDL-C 40 mg/dL,
Chapter j 12 Preb1-HDL
253
FIGURE 12.2 Plasma preb1-HDL concentration according to age and sex. Fasting plasma was obtained from 100 healthy Japanese subjects. The preb1-HDL concentration was measured by immunoassay using a specific monoclonal antibody.
LDL-C 140 mg/dL, and a TG of 300 mg/dL. The mean preb1-HDL concentration, measured by native 2-D gel electrophoresis, was significantly higher in the VA-HIT subjects than in those male controls from the Framingham Offspring Study who had a low HDL-C level (<40 mg/dL). Univariate analysis revealed that the preb1-HDL concentration is a significant risk factor for new cerebrovascular events, although the trend disappeared in a multivariate analysis. In patients undergoing hemodialysis (HD), the preb1-HDL concentration markedly increased independent of the HDL-C concentration (Miida et al., 2003a). In HD patients, the LCAT-dependent conversion of preb1-HDL to a-HDL is severely delayed. When HD patients and control subjects were analyzed together, LCAT activity was negatively correlated not with the baseline preb1-HDL concentration but with a decrease in the preb1-HDL concentration during 120 min of incubation at 37 C (Miida et al., 2003a). When the HD patients were considered alone, however, LCAT activity was not correlated with the drop in preb1-HDL. LCAT esterifies free cholesterol on LDL and HDL via its b- and a-LCAT activities, respectively (Jonas, 1998). We previously examined preb1-HDL metabolism in a patient with fish-eye disease (FED), which is characterized by a lack of a-LCAT activity as a result of a mutation in LCAT (Miida et al., 2004b). Although patients with FED usually show subnormal total plasma LCAT activity owing to the remaining b-LCAT activity, they have very low levels of HDL-C (<10 mg/dL). In our patient, the plasma preb1-HDL concentration was nearly normal, but little preb1-HDL was converted to a-HDL at 37 C (Figure 12.3). These observations strongly suggest
254
The HDL Handbook
FIGURE 12.3 Impaired conversion of preb1-HDL into a-HDL in fish-eye disease.
that plasma preb1-HDL metabolism is closely related to a-LCAT activity but not to b-LCAT activity. Blood samples were obtained from a healthy normolipidemic volunteer and from a patient with (FED). The fresh plasma was incubated at 37 C for 90 min. The amount of preb1-HDL in the normal plasma was decreased markedly, whereas the amount in the FED plasma was increased. Patients with poorly controlled type 2 diabetes, another group at high risk for CAD, have elevated fasting preb1-HDL concentrations, which are associated with the severity of carotid artery atherosclerosis (Hirayama et al., 2007). Those patients with type 2 diabetes had mean absolute and relative preb1-HDL concentrations that were 21 and 34% higher, respectively, than control levels. Stepwise regression analysis showed that the absolute and relative concentrations of preb1-HDL were significant independent determinants for maximum intima-media thickness and plaque score. The mechanism underlying these findings should be clarified in future studies.
TURNOVER OF PREb1-HDL IN HUMAN PLASMA When plasma is incubated at 37 C, the preb1-HDL concentration is decreased exponentially as preb1-HDL is converted to a-HDL by LCAT. Therefore, we
Chapter j 12 Preb1-HDL
255
FIGURE 12.4 Determination of the conversion half-time of preb1-HDL (CHTpreb1). Fresh plasma was incubated with or without DTNB, and paired samples were taken at baseline (C0) and every 30 min. The preb1-HDL concentration was measured by immunoassay. The difference in the preb1-HDL concentration between the paired samples was assumed to reflect the conversion of preb1-HDL to a-HDL by LCAT. The CHTpreb1 is defined as the time required for the preb1-HDL concentration to reach 50% of C0.
determined the time required for a 50% reduction in the preb1-HDL concentration from the baseline level and defined this as the conversion half-time (CHTpreb1) (Figure 12.4). The CHTpreb1 probably reflects the turnover rate of preb1-HDL in vitro (Miida et al., 2004a). In 100 healthy Japanese subjects, the mean CHTpreb1 was 47.4 13.0 min, with no significant difference according to sex or age. Therefore, CHTpreb1 may be used as a new marker for the turnover of HDL. At least in healthy subjects, CHTpreb1 is the strongest determinant of the plasma preb1-HDL concentration (Miida et al., 2004a).
CONCLUSIONS It has been nearly 20 years since preb1-HDL was first detected using native 2-D gel electrophoresis; however, it is still controversial whether plasma preb1HDL is the initial acceptor of cellular cholesterol or is the first product of the interaction between free apoA-I and cellular lipids via the ABCA1-dependent pathway. Thus, the plasma preb1-HDL concentration may be useful as a marker for atherosclerosis or for the metabolic turnover of HDL. Additional studies are needed to distinguish between these possibilities.
ACKNOWLEDGMENTS We thank Osamu Miyazaki (Sekisui Medical) and Ms Utako Seino (Niigata University) for their excellent technical assistance. We are most grateful to Ms Mayumi Okada for her invaluable support to our laboratory. This work was funded in part by Grants-in-Aid for Science Research from the Ministry of Education, Science and Culture of Japan (No. 12671102, 16590815, and 20590558).
256
The HDL Handbook
REFERENCES Asztalos, B. F., Brousseau, M. E., McNamara, J. R., Horvath, K. V., Roheim, P. S., & Schaefer, E. J. (2001). Subpopulations of high density lipoproteins in homozygous and heterozygous Tangier disease. Atherosclerosis, 156, 217e225. Asztalos, B. F., Collins, D., Cupples, L. A., et al. (2005). Value of high-density lipoprotein (HDL) subpopulations in predicting recurrent cardiovascular events in the Veterans Affairs HDL Intervention Trial. Arterioscler Thromb Vasc Biol, 25, 2185e2191. Asztalos, B. F., Cupples, L. A., Demissie, S., et al. (2004). High-density lipoprotein subpopulation profile and coronary heart disease prevalence in male participants of the Framingham Offspring Study. Arterioscler Thromb Vasc Biol, 24, 2181e2187. Asztalos, B. F., Roheim, P. S., Milani, R. L., et al. (2000). Distribution of apoA-I-containing HDL subpopulations in patients with coronary heart disease. Arterioscler Thromb Vasc Biol, 20, 2670e2676. Barbaras, R., Puchois, P., Fruchart, J. C., & Ailhaud, G. (1987). Cholesterol efflux from cultured adipose cells is mediated by LpAI particles but not by LpAI:AII particles. Biochem Biophys Res Commun, 142, 63e69. Barrans, A., Collet, X., Barbaras, R., et al. (1994). Hepatic lipase induces the formation of pre-b 1 high density lipoprotein (HDL) from triacylglycerol-rich HDL2. A study comparing liver perfusion to in vitro incubation with lipases. J Biol Chem, 269, 11572e11577. Bjo¨rkhem, I., Lu¨tjohann, D., Diczfalusy, U., Sta¨hle, L., Ahlborg, G., & Wahren, J. (1998). Cholesterol homeostasis in human brain: turnover of 24S-hydroxycholesterol and evidence for a cerebral origin of most of this oxysterol in the circulation. J Lipid Res, 39, 1594e1600. Castro, G. R., & Fielding, C. (1988). Early incorporation of cell-derived cholesterol into pre-betamigrating high-density lipoprotein. Biochemistry, 27, 25e29. Cenedella, R. J. (1984). Lipoproteins and lipids in cow and human aqueous humor. Biochim Biophys Acta, 793, 448e454. Chau, P., Nakamura, Y., Fielding, C. J., & Fielding, P. E. (2006). Mechanism of prebeta-HDL formation and activation. Biochemistry, 45, 3981e3987. Christoffersen, C., Nielsen, L. B., Axler, O., Andersson, A., Johnsen, A. H., & Dahlback, B. (2006). Isolation and characterization of human apolipoprotein M-containing lipoproteins. J Lipid Res, 47, 1833e1843. Chroni, A., Koukos, G., Duka, A., & Zannis, V. I. (2007). The carboxy-terminal region of apoA-I is required for the ABCA1-dependent formation of a-HDL but not preb-HDL particles in vivo. Biochemistry, 46, 5697e5708. De Backer, G., Ambrosioni, E., Borch-Johnsen, K., et al., Third Joint Task Force of European and Other Societies on Cardiovascular Disease Prevention in Clinical Practice. (2003). European guidelines on cardiovascular disease prevention in clinical practice. Third joint task force of European and other societies on cardiovascular disease prevention in clinical practice. Eur Heart J, 24, 1601e1610. Dietschy, J. M., & Turley, S. D. (2001). Cholesterol metabolism in the brain. Curr Opin Lipidol, 12, 105e112. Expert Panel on Detection, Evaluation, and Treatment of High Blood Cholesterol in Adults. (2001). Executive summary of the third report of the National Cholesterol Education Program (NCEP) expert panel on detection, evaluation, and treatment of high blood cholesterol in adults (Adult Treatment Panel III). J Am Med Assoc, 285, 2486e2497. Fielding, P. E., Miida, T., & Fielding, C. J. (1991). Metabolism of low-density lipoprotein free cholesterol by human plasma lecithin-cholesterol acyltransferase. Biochemistry, 30, 8551e8557.
Chapter j 12 Preb1-HDL
257
Fielding, P. E., Kawano, M., Catapano, A. L., Zoppo, A., Marcovina, S., & Fielding, C. J. (1994). Unique epitope of apolipoprotein A-I expressed in pre-b-1 high-density lipoprotein and its role in the catalyzed efflux of cellular cholesterol. Biochemistry, 33, 6981e6985. Foody, J. M., Ferdinand, F. D., Pearce, G. L., Lytle, B. W., Cosgrove, D. M., & Sprecher, D. L. (2000). HDL cholesterol level predicts survival in men after coronary artery bypass graft surgery: 20-year experience from the Cleveland Clinic Foundation. Circulation, 102, 90e94. Francone, O. L., Gurakar, A., & Fielding, C. (1989). Distribution and functions of lecithin: cholesterol acyltransferase and cholesteryl ester transfer protein in plasma lipoproteins. Evidence for a functional unit containing these activities together with apolipoproteins A-I and D that catalyzes the esterification and transfer of cell-derived cholesterol. J Biol Chem, 264, 7066e7072. Gomaraschi, M., Calabresi, L., & Franceschini, G. (2006). High-density lipoproteins: a therapeutic target for atherosclerotic cardiovascular disease. Expert Opin Ther Targets, 10, 561e572. Gordon, D. J. (1999). Epidemiology of lipoproteins. In D. J. Betteridge, D. R. Illingworth, & J. Shepherd (Eds.), Lipoproteins in health and disease (pp. 587e595). London: Arnold. Gordon, D. J., Probstfield, J. L., Garrison, R. J., et al. (1989). High-density lipoprotein cholesterol and cardiovascular disease: four prospective American studies. Circulation, 79, 8e15. Hirayama, S., Miida, T., Miyazaki, O., & Aizawa, Y. (2007). Preb1-HDL concentration is a predictor of carotid atherosclerosis in type 2 diabetic patients. Diabetes Care, 30, 1289e1291. Huang, Y., von Eckardstein, A., Wu, S., Langer, C., & Assmann, G. (1995). Generation of pre-beta 1-HDL and conversion into alpha-HDL. Evidence for disturbed HDL conversion in Tangier disease. Arterioscler Thromb Vasc Biol, 15, 1746e1754. Inano, K., Tezuka, S., Miida, T., & Okada, M. (2000). Capillary isotachophoretic analysis of serum lipoproteins using a carrier ampholyte as spacer ion. Ann Clin Biochem, 37, 708e716. Jafari, M., Leaf, D. A., Macrae, H., et al. (2003). The effects of physical exercise on plasma prebeta-1 high-density lipoprotein. Metabolism, 52, 437e442, Erratum in Metabolism, 52, 1372. Jaspard, B., Fournier, N., Vieitez, G., et al. (1997). Structural and functional comparison of HDL from homologous human plasma and follicular fluid. A model for extravascular fluid. Arterioscler Thromb Vasc Biol, 17, 1605e1613. Jonas, A. (1998). Regulation of lecithin cholesterol acyltransferase activity. Prog Lipid Res, 37, 209e234. Kawano, M., Miida, T., Fielding, C. J., & Fielding, P. E. (1993). Quantitation of prebeta-HDL dependent and specific components of the total efflux of cellular cholesterol and phospholipid. Biochemistry, 32, 5025e5028. Kay, A. D., Day, S. P., Nicoll, J. A., Packard, C. J., & Caslake, M. J. (2003). Remodelling of cerebrospinal fluid lipoproteins after subarachnoid hemorrhage. Atherosclerosis, 170, 141e146. Kiss, R. S., McManus, D. C., Franklin, V., et al. (2003). The lipidation by hepatocytes of human apolipoprotein A-I occurs by both ABCA1-dependent and -independent pathways. J Biol Chem, 278, 10119e10127. Kitamura, A., Iso, H., Naito, Y., et al. (1994). High-density lipoprotein cholesterol and premature coronary heart disease in urban Japanese men. Circulation, 89, 2533e2539. Kunitake, S. T., La Sala, K. J., & Kane, J. P. (1985). Apolipoprotein A-I-containing lipoproteins with pre-beta electrophoretic mobility. J Lipid Res, 26, 549e555. Lagocki, P. A., & Scanu, A. M. (1980). In vitro modulation of the apolipoprotein composition of high density lipoprotein. Displacement of apolipoprotein A-I from high density lipoprotein by apolipoprotein A-II. J Biol Chem, 255, 3701e3706.
258
The HDL Handbook
Lewis, G. F., & Cabana, V. G. (1996). Postprandial changes in high-density lipoprotein composition and subfraction distribution are not altered in patients with insulin-dependent diabetes mellitus. Metabolism, 45, 1034e1041. Mabuchi, H., Kita, T., Matsuzaki, M., et al., and the J-LIT Study Group. (2002). Large scale cohort study of the relationship between serum cholesterol concentration and coronary events with low-dose simvastatin therapy in Japanese patients with hypercholesterolemia and coronary heart disease: secondary prevention cohort study of the Japan Lipid Intervention Trial (J-LIT). Circ J, 66, 1096e1100. Mamo, J. C. L., Proctor, S. D., & Smith, D. (1998). Retention of chylomicron remnants by arterial tissue; importance of an efficient clearance mechanism from plasma. Atherosclerosis, 141 (Suppl. 1), S63eS69. Miida, T., Fielding, C. J., & Fielding, P. E. (1990). Mechanism of transfer of LDL-derived free cholesterol to HDL subfractions in human plasma. Biochemistry, 29, 10469e10474. Miida, T., Kawano, M., Fielding, P. E., & Fielding, C. J. (1992). Regulation of the concentration of prebeta high density lipoprotein in human plasma by cell membrane and lecithin: cholesterol acyltransferase activity. Biochemistry, 31, 11112e11117. Miida, T., Nakamura, Y., Inano, K., et al. (1996). Preb1-high-density lipoprotein increases in coronary artery disease. Clin Chem, 42, 1992e1995. Miida, T., Inano, K., Yamaguchi, T., Tsuda, T., & Okada, M. (1997). LpA-I levels do not reflect preb1-HDL levels in human plasma. Atherosclerosis, 133, 221e226. Miida, T., Yamaguchi, T., Tsuda, T., & Okada, M. (1998). High preb 1-HDL levels in hypercholesterolemia are maintained by probucol but reduced by a low-cholesterol diet. Atherosclerosis, 138, 129e134. Miida, T., Yamada, T., Yamadera, T., Ozaki, K., Inano, K., & Okada, M. (1999a). Serum amyloid A protein generates preb1 high-density lipoprotein from b-migrating high-density lipoprotein. Biochemistry, 38, 16958e16962. Miida, T., Yamazaki, F., Sakurai, M., et al. (1999b). The apolipoprotein E content of HDL in cerebrospinal fluid is higher in children than in adults. Clin Chem, 45, 1294e1296. Miida, T., Sakai, K., Ozaki, K., et al. (2000a). Bezafibrate increases preb1-HDL at the expense of HDL2b in hypertriglyceridemia. Arterioscler Thromb Vasc Biol, 20, 2428e2433. Miida, T., Ozaki, K., Murakami, T., et al. (2000b). Preb1-high-density lipoprotein (preb1-HDL) concentration can change with low-density lipoprotein-cholesterol (LDL-C) concentration independent of cholesteryl ester transfer protein (CETP). Clin Chim Acta, 292, 69e80. Miida, T., Miyazaki, O., Hanyu, O., et al. (2003a). LCAT-dependent conversion of preb1-HDL into b-migrating HDL is severely delayed in hemodialysis patients. J Am Soc Nephrol, 14, 732e738. Miida, T., Miyazaki, O., Nakamura, Y., et al. (2003b). Analytical performance of a sandwich enzyme immunoassay for preb 1-HDL in stabilized plasma. J Lipid Res, 44, 645e650. Miida, T., Obayashi, K., Seino, U., et al. (2004a). LCAT-dependent conversion rate is a determinant of plasma preb1-HDL concentration in healthy Japanese. Clin Chim Acta, 350, 107e114. Miida, T., Zhang, B., Obayashi, K., et al. (2004b). T13M mutation of lecithin-cholesterol acyltransferase gene causes fish-eye disease. Clin Chim Acta, 343, 201e208. Miida, T., Yamada, T., Seino, U., et al. (2006). Serum amyloid A (SAA)-induced remodeling of CSF-HDL. Biochim Biophys Acta, 1761, 424e433. Miller, N. E., Thelle, D. S., Førde, O. H., & Mjøs, O. D. (1977). The Tromsø heart-study. Highdensity lipoprotein and coronary heart disease: a prospective case-control study. Lancet, 309, 965e968.
Chapter j 12 Preb1-HDL
259
Miyazaki, O., Kobayashi, J., Fukamachi, I., Miida, T., Bujo, H., & Saito, Y. (2000). A new sandwich enzyme immunoassay for measurement of plasma pre-beta1-HDL levels. J Lipid Res, 41, 2083e2088. Moulin, P., Appel, G. B., Ginsberg, H. N., & Tall, A. R. (1992). Increased concentration of plasma cholesteryl ester transfer protein in nephritic syndrome: role in dyslipidemia. J Lipid Res, 33, 1817e1822. Nanjee, M. N., Cooke, C. J., Wong, J. S., Hamilton, R. L., Olszewski, W. L., & Miller, N. E. (2001). Composition and ultrastructure of size subclasses of normal human peripheral lymph lipoproteins: quantification of cholesterol uptake by HDL in tissue fluids. J Lipid Res, 42, 639e648. Neary, R., Bhatnagar, D., Durrington, P., Ishola, M., Arrol, S., & Mackness, M. (1991). An investigation of the role of lecithin:cholesterol acyltransferase and triglyceride-rich lipoproteins in the metabolism of pre-beta high density lipoproteins. Atherosclerosis, 89, 35e48. O’Connor, P. M., Zysow, B. R., Schoenhaus, S. A., et al. (1998). Prebeta-1 HDL in plasma of normolipidemic individuals: influences of plasma lipoproteins, age, and gender. J Lipid Res, 39, 670e678. Okazaki, M., Usui, S., Ishigami, M., et al. (2005). Identification of unique lipoprotein subclasses for visceral obesity by component analysis of cholesterol profile in high-performance liquid chromatography. Arterioscler Thromb Vasc Biol, 25, 578e584. Olchawa, B., Kingwell, B. A., Hoang, A., et al. (2004). Physical fitness and reverse cholesterol transport. Arterioscler Thromb Vasc Biol, 24, 1087e1091. Peynet, J., Feneant-Thibault, M., Legrand, A., Marot, D., Rousselet, F., & Lemonnier, A. (1986). Isolation and characterization of an abnormal alpha slow-moving high-density lipoprotein subfraction in serum from children with long-standing cholestasis. Clin Chem, 32, 646e651. Roheim, P. S., Carey, M., Forte, T., & Vega, G. L. (1979). Apolipoproteins in human cerebrospinal fluid. Proc Natl Acad Sci USA, 76, 4646e4649. Sasahara, T., Yamashita, T., Sviridov, D., Fidge, N., & Nestel, P. (1997). Altered properties of high density lipoprotein subfractions in obese subjects. J Lipid Res, 38, 600e611. Sviridov, D., Kingwell, B., Hoang, A., Dart, A., & Nestel, P. (2003). Single session exercise stimulates formation of pre beta 1-HDL in leg muscle. J Lipid Res, 44, 522e526. Sviridov, D., Chin-Dusting, J., Nestel, P., et al. (2006). Elevated HDL cholesterol is functionally ineffective in cardiac transplant recipients: evidence for impaired reverse cholesterol transport. Transplantation, 81, 361e366. Teramoto, T., Sasaki, J., Ueshima, H., et al. (2007). Executive summary of Japan Atherosclerosis Society (JAS) guideline for diagnosis and prevention of atherosclerotic cardiovascular diseases for Japanese. J Atheroscler Thromb, 14, 45e50. Vedhachalam, C., Ghering, A. B., Davidson, W. S., Lund-Katz, S., Rothblat, G. H., & Phillips, M. C. (2007). ABCA1-induced cell surface binding sites for apoA-I. Arterioscler Thromb Vasc Biol, 27, 1603e1609. Wolfrum, C., Poy, M. N., & Stoffel, M. (2005). Apolipoprotein M is required for prebeta-HDL formation and cholesterol efflux to HDL and protects against atherosclerosis. Nat Med, 11, 418e422. Yokoyama, S. (2006). ABCA1 and biogenesis of HDL. J Atheroscler Thromb, 13, 1e15. Zhang, B., Miura, S., Fan, P., et al. (2006). ApoA-I/phosphatidylcholine discs remodels fastmigrating HDL into slow-migrating HDL as characterized by capillary isotachophoresis. Atherosclerosis, 188, 95e101. Zhou, Z., Rahme, E., & Pilote, L. (2006). Are statins created equal? Evidence from randomized trials of pravastatin, simvastatin, and atorvastatin for cardiovascular disease prevention. Am Heart J, 151, 273e281.
Chapter 13
Determination of Circulating Native and Denaturated HDL Concentrations and its Clinical Implications Takanari Nakano 1, 2, Makoto Seo 1 and Tsugikazu Komoda 3 1
Department of Biochemistry, Faculty of Medicine, Saitama Medical University, Saitama, Japan, Brentwood Biomedical Research Institute, Department of Medicine, School of Medicine, University of California Los Angeles, Los Angeles, CA, USA, 3 Nihon Medical Science Institute, 11-4 Minami-Tohrimachi, Kawagoe, Saitama, Japan 2
INTRODUCTION Atherosclerosis and resulting cardiovascular diseases are the primary cause of mortality. Several large epidemiologic studies have demonstrated appropriate laboratory parameters that are associated with and can predict coronary artery disease (CAD) and/or cardiovascular events. One of the major independent risk factors is total cholesterol. In particular, elevated low density lipoprotein cholesterol (LDL-C) is strongly and independently associated with the increased risk. Large CAD prevention trials have shown that, despite intensive statin treatment regimens, the reduction of LDL-C did not prevent the majority of CAD events (Mudd et al., 2007), which could be explained by modifications of LDLs, because the atherogeneicity is not only dependent on their concentration, but is also associated to their size, density, and/or other physicochemical aspects. Among them, oxidation of low density lipoprotein (LDL) has been proposed to play a critical role in converting macrophages into cholesteryl ester-laden foam cells and vigorously examined (Liao et al., 1995; Mertens and Holvoet, 2001). With accumulating evidence, circulating oxidized LDL has been proposed as a clinical marker and the data obtained so far are consistent in stating that oxidized LDL concentrations are increased in patients with CAD (Holvoet et al., 2001). The mean concentrations of oxidized LDL in the circulation The HDL Handbook. ISBN: 978-0-12-382171-3 Copyright Ó 2010 Elsevier Inc. All rights of reproduction in any form reserved.
261
262
The HDL Handbook
significantly differ between patients and healthy subjects, but the measurement itself discriminates only the part of the population examined, suggesting that the concentration of oxidized LDL may not be sufficiently useful in diagnosis of individual patients if it is used alone. On the other hand, high density lipoprotein (HDL) appears to be preferentially oxidized among lipoproteins in the circulation because HDL is highly susceptible to oxidation rather than LDL (Garner et al., 1998a,b). Recent findings also indicate that myeloperoxidase, one of the enzymes that produces reactive oxygen species, oxidizes HDL preferentially in the circulation and the artery wall and thereby impairs its function for reverse cholesterol transport (Pieters et al., 1994; von Eckardstein et al., 2001; Zhang et al., 2003). These suggest that oxidation of HDL is a prospective marker for atherosclerosis; however, the possibility has not fully been explored. In this chapter, we focus on the detection of oxidized HDL and its clinical usefulness.
EVIDENCE OF OXIDIZED HDL IN TISSUE AND CIRCULATION Oxidized LDL was first found in human and rabbit atherosclerotic lesions. The extracted LDL resembled the LDL oxidized in vitro (Yla-Herttuala et al., 1989). The concept of “LDL oxidation” was well accepted and studies have demonstrated the involvement of oxidized LDL in the pathogenesis of atheroscrelosis. In contrast to LDL, increased plasma levels of HDL are inversely correlated with the risk of CAD (Stampfer et al., 1991). The reverse cholesterol transport by HDL plays a major role in the favorable effect (Pieters et al., 1994). On the other hand, studies indicated that the high susceptibility of HDL to oxidation is also associated with its anti-atherogenic effects. By acting as a sacrificial target for oxidation, HDL can prevent LDL from oxidation, which plays a key role in the progression of atherosclerosis as described above (Gerrity, 1981; Parthasarathy et al., 1991; Yla-Herttuala, 1999; Francis, 2000). Moreover, HDL can also detoxify oxidized LDL by receiving the lipid hydroperoxides in oxidized LDL (Christison et al., 1996). These findings indicate that HDL is closely associated with the oxidation of LDL. Further, oxidized HDL may increase in the early stage of the development of atherosclerosis, suggesting that increased levels of oxidized HDL provide an early marker and predictor for it. However, in contrast to the concept of “oxidized LDL”, we have experienced resistance to the concept of “oxidized HDL”. In fact, evidence for oxidized HDL in humans has not been as hotly pursued. Here, we summarize evidence reported so far concerning the existence of “oxidized” HDL in vivo. The flux of HDL into the artery wall can be greater than LDL in normal aortic intima and in arterial plaques (Bjornheden et al., 1996). This difference in the flux is presumably due to the smaller particle size of HDL (Smith, 1990). In fact, about half of the total body mass of HDL apoproteins is found in
Chapter j 13 Determination of Circulating Native
263
interstitial fluid (Reichl, 1990). These findings suggest a possibility that HDL is oxidized in interstitial fluid at sites of inflammation, such as the atherosclerotic lesion. While important insights have been obtained with regard to the cardioprotective nature of HDL as a scavenger for oxidation (Francis, 2000), studies on the determination of circulating oxidized HDL have been limited. Bowry et al. (1992) detected oxidized lipids in HDL, and suggested a role for HDL as a carrier of hydroperoxides in the circulation. HDL is a complex mixture of cholesterol carrying heterogeneous lipoprotein particles with a predominant tetrametric apolipoprotein A-I (apoA-I) backbone (Brouillette et al., 2001; Undurti et al., 2009). von Eckardstein et al. (1991) reported that the oxidized form of apoA-I, the major protein component of the HDL molecule, was found in freshly prepared HDL particles from healthy human plasma. Artola et al. (1997) reported that cross-linked apoproteins existed in HDL isolated from hypercholesterolemic chickens without lipid oxidation products. On the other hand, Niu et al. (1999) found that HDL and LDL isolated from human endarterectomy specimens contained similar levels of oxidized lipids. We previously developed an enzyme-linked immunosorbent assay (ELISA) with an antibody against an oxidized form of apoA-I (Nakano and Nagata, 2003). Moreover, the antibody showed a reactivity to HOCl-modified HDL (Bergt et al., 2006). By using the ELISA, we found that the association of serum oxidized HDL concentrations with the polymorphism in the gene encoding p22phox of NADPH oxidase was greater than that with oxidized LDL (Nakano et al., 2003). Heinecke and his colleagues have demonstrated the presence of a number of oxidized amino acids in apoA-I using liquid chromatographye electrospray ionization mass spectrometry (LCeESIeMS and MS/MS) (Bergt et al., 2004a,b; Pennathur et al., 2004; Shao et al., 2006; Shao and Heinecke, 2008). Analysis of separated HDL from sera by ultracentrifugation showed that tyrosine 192 is preferentially chlorinated or nitrated. Moreover, they found that three methionine residues in an apoA-I molecule could be converted to methionine sulfoxide. These findings indicate the presence of oxidized HDL in the circulation and/or interstitial fluid in humans.
MEASUREMENT OF OXIDIZED HDL IN CLINICAL SAMPLES Immunoassays The high oxidative susceptibility of HDL not only contributes to its antiatherogenic effects, but also may serve as an early marker for oxidative stress as HDL acts as a sacrificial target for oxidation (Francis, 2000). In fact, HDL can reduce the level of oxidized LDL by accepting the lipid hydroperoxides of oxidized LDL (Christison et al., 1996). Thus, HDL is closely involved in the oxidation of LDL, suggesting that increased levels of oxidized HDL provide an early marker and predictor of atherosclerosis. A couple of immunoassays have
264
The HDL Handbook
been developed for the detection of oxidized LDL (Itabe et al., 1996; Holvoet et al., 1999, 2006) and their clinical usefulness has been demonstrated in many large-scale clinical studies (Holvoet et al., 1999, 2000, 2001, 2004; Toshima et al., 2000; Ehara et al., 2001; Uno et al., 2003). On the other hand, there was no method available to measure circulating levels of oxidized HDL when we started studying the oxidation of HDL. Thus, our first attempt was to develop an assay method for oxidized HDL. To assess the concentration of oxidized HDL using antioxidized apoA-I as a target seemed reasonable because apoA-I is highly susceptible to oxidation and is itself oxidized in the early phase of HDL oxidation, providing a specific antigenic determinant for oxidized HDL (Francis, 2000; von Eckardstein et al., 2001). Also, an HDL particle is built on a tetrametric apoA-I backbone. We produced antioxidized apoA-I monoclonal antibody 3C11 and developed an enzyme-linked immunosorbent assay (ELISA) (Nakano and Nagata, 2003). The ELISA with mAb 3C11 and an anti-apoA-I mAb showed high sensitivity and specificity for oxidized HDL. With this assay, we observed detectable oxidized HDL in human serum samples. Oxidized apoA-I would be more suitable than oxidized lipids for an antigenic determinant for oxidized HDL, because each human lipoprotein, such as LDL, contains the same lipid components. Thus, the detection of oxidized lipids results in substantial crossreactivity if other lipoproteins are oxidized. Using homogenates of human atheromatous plaque as immunogen, Itabe et al. (1994) established an mAb, named FOH1a/DLH3, that reacts with oxidized phosphatidylcholine. The FOH1a/DLH3 mAb was reported to react with oxidized HDL (Itabe et al., 1994), because HDL contains phosphatidylcholine in its particle. This crossreactivity can impair the assay reliability of FOH1a/DLH3-based assays. MAb 3C11 and the ELISA with the mAb had high specificity for oxidized forms of HDL and apoA-I, but they had no substantial reactivity against the other oxidized forms of lipoproteins and apolipoproteins. Native apoA-I showed relatively high cross-reactivity in the assay; however, this may have been because some of the apoA-I molecules, which were purified from human plasma, were already oxidized, as reported by von Eckardstein et al. (1991). Using H2O2- or chloramine T-oxidized apoA-I as immunogen, Ueda et al. (2007) prepared two monoclonal antibodies, 7D3 or 98A3, and established ELISAs for the measurement of oxidized apoA-I. They also measured oxidized HDL in serum from healthy subjects and patients with type 2 diabetes and inflammatory conditions and found detectable oxidized HDL in them. Immunoassay has advantages in its clinical application, because: 1. it usually does not require the separation of HDL before analysis 2. immunoassay can be done manually or can be applied instrumentally in clinical laboratories 3. immunoassay can be standardized among laboratories that allows us to share and trace data to conduct clinical studies.
Chapter j 13 Determination of Circulating Native
265
One can generate anti-oxidized apoA-I specific mAbs as above; however, they would be antigenically different. Oxidized forms of HDL can differ with the method of oxidation employed in each study. In fact, many changes can occur in the apoA-I molecule with oxidative modification (Maziere et al., 1993; Shao and Heinecke, 2008). Thus, each oxidized HDL or apoA-I immunoassay should be examined for its determinant and clinical relevancy. Also, comparisons should be made between assays. Wang et al. recently developed an immunoassay, which detects methionine sulfoxide residues in the apoA-I molecule (Wang et al., 2009). The antigenic deteriminant of the antibody used in the study is well identified. Using such immunoassays, further studies will give chemical insight into site-specific HDL modifications in relation to the risk of atheroscrelosis in humans.
OTHER METHODS Heinecke and his colleagues have revealed the mechanisms of HDL oxidation (Bergt et al., 2004a,b; Pennathur et al., 2004; Shao et al., 2006; Shao and Heinecke, 2008) using a mass spectrometry approach. Possible pathways for the oxidation of HDL are reviewed in Shao et al., 2006. They also showed that the nitrotyrosine content in separated HDL from patients with CAD was greater than that from healthy subjects (Pennathur et al., 2004). A series of studies by the research group has shown that there is a variety of oxidized forms in the apoA-I molecule and that the limited amino acids are oxidized by specific enzymes and reactive oxygen species (Shao and Heinecke, 2008). For example, HOCl and myeloperoxidase preferentially chlorinate tyrosine 192 in lipid-free apoA-I. These data may identify oxidative sites in apoA-I that are associated with the deficiency of apoA-I functions and disease progression. Such determinants would provide information to develop meaningful antibodies for the detection of oxidized HDL or apoA-I. Wang and Stocker (2008) demonstrated the measurement of oxidized apoAI and apoA-II, a minor protein constituent of HDL, by a high performance liquid chromatography (HPLC) method, which may be useful to assess oxidized HDL in biological samples.
CONSIDERATIONS FOR HDL OXIDATION IN VITRO Accumulating evidence shows that HDL is rapidly modified by a variety of oxidants in vitro (Banka, 1996; Francis, 2000). Transition metal ions (Nagano et al., 1991), aqueous peroxyl and hydroxyl radicals (Bonnefont-Rousselot et al., 1995), peroxidase-generated tyrosyl radical (Francis et al., 1993), and hypochlorous acid (Panzenboeck et al., 1997) can oxidize HDL. Moreover, HDL is more susceptible than LDL. HDL oxidation starts readily without any ‘lag phase’, which is seen in LDL oxidation (Bowry et al., 1992). HDL contains
266
The HDL Handbook
relatively lower antioxidants than LDL, with an average of 0.3 vs. 8e12 a-tocopherol and 0.015 vs. 0.5-1.0 ubiquinol-10 molecules per particle on HDL and LDL, respectively (Bowry et al., 1992). This fact is likely to allow HDL to be oxidized more rapidly than LDL. Further, apoA-I is a chromophore and can be photo-oxidized under light (Salmon et al., 1992; Nakano and Nagata, 2005). Indeed, four tryptophan residues in an apoA-I molecule are oxidized by absorbing wavelengths of 295 nm by generating superoxide radical anions and the tryptophanyl radical. We observed that fluorescent lights can easily oxidize apoA-I. The exposure to light strongly oxidized apoA-I in serum samples, the oxidized HDL values measured by an ELISA increased 58.8-fold under UV light and 18.4-fold under fluorescent light (Nakano and Nagata, 2005). Moreover, apoA-I in serum is likely to be oxidized more easily than purified HDL with exposure to fluorescent light, because serum contains various chromophores. Considering these results, researchers should keep samples containing apoA-I in the dark to prevent exposure to light; otherwise the amount of oxidized and total apoA-I may be overestimated and such ‘in vitro’ oxidation would be mistakenly seen as ‘in vivo’ oxidation. We found that quercetin, a flavonoid, prevents apoA-I from being oxidized in light by absorbing it, suggesting quercetin is useful as a preservative for apoA-I and HDL.
CLINICAL SIGNIFICANCE OF OXIDIZED HDL MEASUREMENT Although important insights have been obtained with regard to the cardioprotective nature of HDL as a scavenger for oxidation, studies involving the determination of circulating oxidized HDL concentrations have been limited. Pennathur et al. (2004) analyzed the nitrotyrosine content in purified HDL fractions from patients and found that the content was greater in patients with CAD than healthy control subjects. The method that Pennathur et al. employed is reliable and accurate, but it is technically- and equipment-demanding; thus, it is difficult to apply to many specimens for clinical studies. We first developed an ELISA for oxidized HDL and observed detectable amounts of oxidized HDL in serum specimens (Nakano et al., 2003). Ueda et al. (2007) also reported ELISAs for oxidized HDL. Here we summarize available clinical data conducted with these immunoassay kits.
NAD(P)H oxidase p22phox gene C242T polymorphism and HDL oxidation The NAD(P)H oxidase system plays a major role in superoxide anion (O 2) production in human vessel walls (Griendling et al., 2000). p22phox, a protein component of NAD(P)H oxidase, is essential for the assembly and
Chapter j 13 Determination of Circulating Native
267
activation of the system (Dinauer et al.,1990). The gene coding for p22phox (CYBA) has four types of allelic variant including a C242T mutation that substitutes Tyr for His at codon 72 of p22phox involving a potential hemebinding site (Dinauer et al., 1990), suggesting that consequences of this mutation include changes in O 2 production due to impairment of heme binding (Wolin, 2000). Such changes in O 2 production levels could alter the redox state in human vessel walls. Although the influence of the mutation on oxidative stress has been implicated, the clinical consequences of the polymorphism are still controversial. Guzik et al. (2000a,b) have shown that human saphenous veins with the CYBA 242T allele (CT/TT genotypes) had less O 2 -producing activity than those with CC genotypes. Inoue et al. (1998) showed that the CYBA 242T allele reduced the risk for CAD in the Japanese population. HDL can freely access the active site of NAD(P)H oxidase because HDL diffuses rapidly into the interstitial fluid of human arterial intima for cholesterol transportation (Francis, 2000). We hypothesized that the oxidation of HDL may relate closely to the NAD(P)H oxidase O 2 -producing activity and therefore evaluated the effect of the CYBA C242T polymorphism on serum concentrations of oxidized HDL in patients with type 2 diabetes mellitus. Results of this study showed that the 242 T genotypes were associated with lowering serum concentrations of oxidized lipoproteins in type 2 diabetic patients (Nakano et al., 2003). The mean concentration of oxidized HDL was 1.5-fold lower in sera from patients with the 242T allele, consistent with the above findings. On the other hand, serum concentration of oxidized LDL in patients with the 242T allele was slightly greater than those with the CC genotype. This controversy may be due to the difference in the oxidative susceptibility between HDL and LDL.
Oxidized HDL in patients with Alzheimer’s disease Inflammation and oxidative stress are associated with the pathogenesis of Alzheimer’s disease (AD). Several biomarkers of oxidative stress (Sayre et al., 1997; Smith et al., 1997, 1998; Pratico et al., 1998) are increased in the brain of AD patients. However, little is known about the role of lipoprotein oxidation in AD. To examine the role of plasma lipoprotein oxidation in Alzheimer’s pathology, we quantified plasma levels of oxidized HDL and apo A-I in AD patients and cognitive normal individuals. Interestingly, levels of oxidized HDL were significantly lower in AD patients compared to the controls, whereas apo A-I levels were comparable (Bergt et al., 2006). These results were quite unexpected considering a series of evidence showing increased oxidative stress in patients with AD. We observed that statins and/ or COX-2 inhibitors contribute to reduced oxidized HDL plasma levels in controls and in AD patients. However, AD patients always had lower oxidized HDL levels.
268
The HDL Handbook
Oxidized HDL concentrations do not increase in patients undergoing carotid endarterectomy The observations stated above suggest that oxidation of HDL is an early marker for atherosclerosis; however, the possibility has not fully been explored. To address this, we assayed plasma concentrations of oxidized HDL in patients with more than 70% stenosis, a complicated form of atherosclerosis. The significant increase of circulating oxidized LDL in the population was determined in a previous study (von Eckardstein et al., 2001). We recruited 29 (24 men; aged 67.6 (7.2 SD) years) patients with more than 70% stenosis who underwent carotid endarterectomy (von Eckardstein et al., 2001). The results of the present study showed that there was no significant difference in circulating oxidized HDL/apoA-I protein values between patients with stenosis and controls, although we previously found that oxidized LDL/apoB protein values were greater in patients. Furthermore, there was no relationship between the values of oxidized HDL/apoA-I and oxidized LDL/apoB.
Oxidized HDL in patients with diabetes and inflammatory conditions Ueda et al. (2007) also developed oxidized HDL ELISAs and applied a preliminary clinical study with specimens obtained from patients with diabetes and inflammatory conditions. In inflammatory disease, the activation of neutrophils may involve activation of NADPH oxidase and myeloperoxidase, followed by formation of oxidants. As a result, the level of oxidized apoA-I was greater in patients with inflammatory disease than in healthy individuals. They suggested the oxidation of apoA-I by increased oxidative stress in those patients. They also observed increased oxidized HDL levels in patients with diabetes. In diabetic conditions, it was suggested that patients have increased oxidant stress with decreased antioxidant capacity (Seghrouchni et al., 2002), and activation of NADPH oxidase by advanced glycation end products (Wautier et al., 2001). These findings support the hypothesis that the increased oxidized HDL was caused by the elevated oxidative stress in those patients.
Cellular uptake and clearance of oxidized HDL from the circulation The results of the preliminarily clinical studies above are inconsistent. Patients with AD or stenosis should be under increased oxidative stress, but oxidized HDL was lower than control subjects. Rapid clearance and detoxification of oxidized HDL from the circulation in these patients may be associated with the lack of increase in oxidized HDL. In vivo kinetic studies
Chapter j 13 Determination of Circulating Native
269
of native and oxidized HDL in Watanabe hereditable hyperlipidemic (WHHL) and control rabbits have demonstrated that oxidized HDL is cleared more rapidly than native HDL (Liu et al., 1993). Christison et al. (1996) also demonstrated that lipid hydroperoxides in HDL are readily and selectively removed by liver cells. Moreover, Braschi et al. (1999) showed that lipid-poor HDL is cleared rapidly from the circulation. These findings suggest a possibility that oxidized HDL is cleared rapidly from the circulation, and maybe the clearance is accelerated under increased oxidative stress. In fact, WHHL rabbits cleared native and oxHDL at a higher fraction than control rabbits (Liu et al., 1993). We demonstrated the rapid clearance in mice by administering human oxidized HDL or native HDL via i.v. injection and measuring the remaining human apoA-I in the circulation. Oxidized HDL was cleared faster than native HDL or oxidized LDL from the circulation (Figure 13.1), decreasing by half in the circulating concentration within 10 min. On the other hand, there was no apparent decrease in native HDL within 15 min. Oxidized LDL showed time-dependent clearance over the tested period, suggesting there is no accelerated absorption mechanism for oxidized LDL. The cellular uptake of oxidized HDL was also examined with THP-1 cells, a monocyte lineage. The results showed that
FIGURE 13.1 Uptake of native or oxidized lipoproteins by THP-1 cells. Lipoproteins were purified by ultracentrifugation. Oxidized HDL or LDL were prepared by adding HOCl. Lipoproteins were added to THP-1 cells at a final concentration of 10 mg/mL and incubated for 4 h at 37 C. After the incubation, the cells were washed with cold PBS three times and lyzed with 0.1% Triton X-100-containing PBS. The cell lysates were assayed for apoA-I or apoB to determine the contents of lipoproteins taken up. Plots indicate mean of triplicate assays.
270
The HDL Handbook
FIGURE 13.2 Clearance of oxidized lipoproteins prepared form human plasma in mice. Male Balb/c mice were fasted overnight and injected each lipoproteins (150 mg for native or oxidized HDL, 30 mg for native or oxidized LDL). After the injection, blood samples were taken from the tails periodically over 60 min. Lipoproteins were oxidized by HOCl with the different concentration of HOCl as indicated in legends. Bar indicates mean and SEM of triplicate assays.
oxidation of HDL increased its uptake by three- to four-fold, while oxidation of LDL increased its uptake by 30% (Figure 13.2). These results support the hypothesis that oxidation of HDL accelerates its clearance from the circulation. Elucidation of such mechanisms with oxidized HDL may provide a better understanding for the role of oxidation of HDL in the pathogenesis of atherosclerosis.
CONCLUSION Biochemical analysis of HDL or apoA-I have revealed that apoA-I, the major protein constitute of HDL particles, is readily oxidized in vivo. The oxidation is mediated by a variety of radicals or enzymes. Several pathways for the oxidation have been shown and target amino acids have been identified. The high susceptibility of HDL to oxidation and its association with loss of antiatherogenic ability have attracted several groups as a mechanism of atherosclerosis. We and some other groups have been focusing on the clinical usefulness of its measurement. Assays for oxidized HDL have been developed and applied for the measurement of clinical specimens. Unexpectedly, some of the results are paradoxical. Oxidized HDL was decreased in patients who were supposed to have increased oxidative stress. Oxidized HDL may be removed from the circulation rapidly considering the more rapid clearance of oxidized HDL than native HDL. Moreover, decreased oxidized HDL in patients suggests
Chapter j 13 Determination of Circulating Native
271
that the pathway to clear or scavenge oxidized HDL is accelerated in some disease conditions. Thus, the quantitative measurement of oxidized HDL may be not useful as a clinical marker. A number of amino acids in apoA-I can be oxidized. Some may be related to its dysfunction for reverse cholesterol transport, suggesting that such determinants are candidates to be examined as antibody targets for further development of immunoassays.
REFERENCES Artola, R. L., Conde, C. B., Bagatolli, L., Pecora, R. P., Fidelio, G. D., & Kivatinitz, S. C. (1997). High-density lipoprotein from hypercholesterolemic animals has peroxidized lipids and oligomeric apolipoprotein A-I: its putative role in atherogenesis. Biochem Biophys Res Commun, 239(2), 570e574. Banka, C. L. (1996). High density lipoprotein and lipoprotein oxidation. Curr Opin Lipidol, 7(3), 139e142. Bergt, C., Fu, X., Huq, N. P., Kao, J., & Heinecke, J. W. (2004a). Lysine residues direct the chlorination of tyrosines in YXXK motifs of apolipoprotein A-I when hypochlorous acid oxidizes high density lipoprotein. J Biol Chem, 279(9), 7856e7866. Bergt, C., Pennathur, S., Fu, X., et al. (2004b). The myeloperoxidase product hypochlorous acid oxidizes HDL in the human artery wall and impairs ABCA1-dependent cholesterol transport. Proc Natl Acad Sci USA, 101(35), 13032e13037. Bergt, C., Nakano, T., Ditterich, J., DeCarl, C., & Eiserich, J. (2006). Oxidized plasma high density lipoprotein is decreased in Alzheimer’s disease. Free Radic Biol Med, 41(10), 1542e1547. Bjornheden, T., Babyi, A., Bondjers, G., & Wiklund, O. (1996). Accumulation of lipoprotein fractions and subfractions in the arterial wall, determined in an in vitro perfusion system. Atherosclerosis, 123, 43e56. Bonnefont-Rousselot, D., Motta, C., Khalil, A. O., et al. (1995). Physicochemical changes in human high-density lipoproteins (HDL) oxidized by gamma radiolysis-generated oxyradicals. Effect on their cholesterol effluxing capacity. Biochim Biophys Acta, 1255(1), 23e30. Bowry, V. W., Stanley, K. K., & Stocker, R. (1992). High density lipoprotein is the major carrier of lipid hydroperoxides in human blood plasma from fasting donors. Proc Natl Acad Sci USA, 89 (21), 10316e10320. Braschi, S., Neville, T. A., Vohl, M. C., & Sparks, D. L. (1999). Apolipoprotein A-I charge and conformation regulate the clearance of reconstituted high density lipoprotein in vivo. J Lipid Res, 40(3), 522e532. Brouillette, C. G., Anantharamaiah, G. M., Engler, J. A., & Borhani, D. W. (2001). Structural models of human apolipoprotein A-I: a critical analysis and review. Biochim Biophys Acta, 1531(1-2), 4e46. Christison, J., Karjalainen, A., Brauman, J., Bygrave, F., & Stocker, R. (1996). Rapid reduction and removal of HDL- but not LDL-associated cholesteryl ester hydroperoxides by rat liver perfused in situ. Biochem J, 314(Pt 3), 739e742. Dinauer, M. C., Pierce, E. A., Bruns, G. A., Curnutte, J. T., & Orkin, S. H. (1990). Human neutrophil cytochrome b light chain (p22-phox). Gene structure, chromosomal location, and mutations in cytochrome-negative autosomal recessive chronic granulomatous disease. J Clin Invest, 86(5), 1729e1737.
272
The HDL Handbook
Ehara, S., Ueda, M., Naruko, T., et al. (2001). Elevated levels of oxidized low density lipoprotein show a positive relationship with the severity of acute coronary syndromes. Circulation, 103(15), 1955e1960. Francis, G. A. (2000). High density lipoprotein oxidation: in vitro susceptibility and potential in vivo consequences. Biochim Biophys Acta, 1483(2), 217e235. Francis, G. A., Mendez, A. J., Bierman, E. L., & Heinecke, J. W. (1993). Oxidative tyrosylation of high density lipoprotein by peroxidase enhances cholesterol removal from cultured fibroblasts and macrophage foam cells. Proc Natl Acad Sci USA, 90(14), 6631e6635. Garner, B., Waldeck, A. R., Witting, P. K., Rye, K. A., & Stocker, R. (1998a). Oxidation of high density lipoproteins. II. Evidence for direct reduction of lipid hydroperoxides by methionine residues of apolipoproteins AI and AII. J Biol Chem, 273(11), 6088e6095. Garner, B., Witting, P. K., Waldeck, A. R., Christison, J. K., Raftery, M., & Stocker, R. (1998b). Oxidation of high density lipoproteins. I. Formation of methionine sulfoxide in apolipoproteins AI and AII is an early event that accompanies lipid peroxidation and can be enhanced by alpha-tocopherol. J Biol Chem, 273(11), 6080e6087. Gerrity, R. G. (1981). The role of the monocyte in atherogenesis: I. Transition of blood-borne monocytes into foam cells in fatty lesions. Am J Pathol, 103(2), 181e190. Griendling, K. K., Sorescu, D., & Ushio-Fukai, M. (2000). NAD(P)H oxidase: role in cardiovascular biology and disease. Circ Res, 86(5), 494e501. Guzik, T. J., West, N. E., Black, E., et al. (2000b). Vascular superoxide production by NAD(P)H oxidase: association with endothelial dysfunction and clinical risk factors. Circ Res, 86(9), E85e90. Guzik, T. J., West, N. E., Black, E., et al. (2000a). Functional effect of the C242T polymorphism in the NAD(P)H oxidase p22phox gene on vascular superoxide production in atherosclerosis. Circulation, 102(15), 1744e1747. Holvoet, P., Collen, D., & Van de Werf, F. (1999). Malondialdehyde-modified LDL as a marker of acute coronary syndromes. J Am Med Assoc, 281(18), 1718e1721. Holvoet, P., Van Cleemput, J., Collen, D., & Vanhaeck, J. (2000). Oxidized low density lipoprotein is a prognostic marker of transplant-associated coronary artery disease. Arterioscler Thromb Vasc Biol, 20(3), 698e702. Holvoet, P., Kritchevsky, S. B., Tracy, R. P., et al. (2004). The metabolic syndrome, circulating oxidized LDL, and risk of myocardial infarction in well-functioning elderly people in the health, aging, and body composition cohort. Diabetes, 53(4), 1068e1073. Holvoet, P., Macy, E., Landeloos, M., et al. (2006). Analytical performance and diagnostic accuracy of immunometric assays for the measurement of circulating oxidized LDL. Clin Chem, 52(4), 760e764. Holvoet, P., Mertens, A., Verhamme, P., et al. (2001). Circulating oxidized LDL is a useful marker for identifying patients with coronary artery disease. Arterioscler Thromb Vasc Biol, 21(5), 844e848. Inoue, N., Kawashima, S., Kanazawa, K., Yamada, S., Akita, H., & Yokoyama, M. (1998). Polymorphism of the NADH/NADPH oxidase p22 phox gene in patients with coronary artery disease. Circulation, 97(2), 135e137. Itabe, H., Takeshima, E., Iwasaki, H., et al. (1994). A monoclonal antibody against oxidized lipoprotein recognizes foam cells in atherosclerotic lesions. Complex formation of oxidized phosphatidylcholines and polypeptides. J Biol Chem, 269(21), 15274e15279. Itabe, H., Yamamoto, H., Imanaka, T., et al. (1996). Sensitive detection of oxidatively modified low density lipoprotein using a monoclonal antibody. J Lipid Res, 37(1), 45e53.
Chapter j 13 Determination of Circulating Native
273
Liao, L., Aw, T. Y., Kvietys, P. R., & Granger, D. N. (1995). Oxidized LDL-induced microvascular dysfunction. Dependence on oxidation procedure. Arterioscler Thromb Vasc Biol, 15(12), 2305e2311. Liu, R., Saku, K., Zhang, B., Hirata, K., Shiomi, M., & Arakawa, K. (1993). In vivo kinetics of oxidatively modified HDL. Biochem Med Metab Biol, 49(3), 392e397. Maziere, J. C., Myara, I., Salmon, S., et al. (1993). Copper- and malondialdehyde-induced modification of high density lipoprotein and parallel loss of lecithin cholesterol acyltransferase activation. Atherosclerosis, 104(1-2), 213e219. Mertens, A., & Holvoet, P. (2001). Oxidized LDL and HDL: antagonists in atherothrombosis. Faseb J, 15(12), 2073e2084. Mudd, J. O., Borlaug, B. A., Johnston, P. V., et al. (2007). Beyond low-density lipoprotein cholesterol: defining the role of low-density lipoprotein heterogeneity in coronary artery disease. J Am Coll Cardiol, 50(18), 1735e1741. Nagano, Y., Arai, H., & Kita, T. (1991). High density lipoprotein loses its effect to stimulate efflux of cholesterol from foam cells after oxidative modification. Proc Natl Acad Sci USA, 88(15), 6457e6461. Nakano, T., & Nagata, A. (2003). Immunochemical detection of circulating oxidized high-density lipoprotein with antioxidized apolipoprotein A-I monoclonal antibody. J Lab Clin Med, 141(6), 378e384. Nakano, T., & Nagata, A. (2005). Oxidative susceptibility of apolipoprotein AI in serum. Clin Chim Acta, 362(1-2), 119e124. Nakano, T., Matsunaga, S., Nagata, A., & Maruyama, T. (2003). NAD(P)H oxidase p22phox Gene C242T polymorphism and lipoprotein oxidation. Clin Chim Acta, 335(1-2), 101e107. Niu, X., Zammit, V., Upston, J. M., Dean, R. T., & Stocker, R. (1999). Coexistence of oxidized lipids and alpha-tocopherol in all lipoprotein density fractions isolated from advanced human atherosclerotic plaques. Arterioscler Thromb Vasc Biol, 19(7), 1708e1718. Panzenboeck, U., Raitmayer, S., Reicher, H., et al. (1997). Effects of reagent and enzymatically generated hypochlorite on physicochemical and metabolic properties of high density lipoproteins. J Biol Chem, 272(47), 29711e29720. Parthasarathy, S., Barnett, J., & Fong, L. G. (1990). High-density lipoprotein inhibits the oxidative modification of low-density lipoprotein. Biochim Biophys Acta, 1044(2), 275e283. Pennathur, S., Bergt, C., Shao, B., et al. (2004). Human atherosclerotic intima and blood of patients with established coronary artery disease contain high density lipoprotein damaged by reactive nitrogen species. J Biol Chem, 279(41), 42977e42983. Pieters, M. N., Schouten, D., & Van Berkel, T. J. (1994). In vitro and in vivo evidence for the role of HDL in reverse cholesterol transport. Biochim Biophys Acta, 1225(2), 125e134. Pratico, D., Viginia, M. Y. L., Trojanowski, J. Q., Rokach, J., & Fitzgerald, G. A. (1998). Increased F2-isoprostanes in Alzheimer’s disease: evidence for enhanced lipid peroxidation in vivo. FASEB J, 12(15), 1777e1783. Reichl, D. (1990). Lipoproteins of human peripheral lymph. Eur Heart J, 11(Suppl E), 230e236. Salmon, S., Santus, R., Maziere, J. C., Aubailly, M., & Haigle, J. (1992). Modified apolipoprotein pattern after irradiation of human high-density lipoproteins by ultraviolet B. Biochim Biophys Acta, 1128(2-3), 167e173. Sayre, L. M., Zelasko, D. A., Harris, P. L., Perry, G., Salomon, R. G., & Smith, M. A. (1997). 4-Hydroxynonenal-derived advanced lipid peroxidation end products are increased in Alzheimer’s disease. J Neurochem, 68(5), 2092e2097.
274
The HDL Handbook
Seghrouchni, I., Drai, J., Bannier, E., et al. (2002). Oxidative stress parameters in type I, type II and insulin-treated type 2 diabetes mellitus; insulin treatment efficiency. Clin Chim Acta, 321(1-2), 89e96. Shao, B., & Heinecke, J. W. (2008). Using tandem mass spectrometry to quantify site-specific chlorination and nitration of proteins: model system studies with high-density lipoprotein oxidized by myeloperoxidase. Methods Enzymol, 440, 33e63. Shao, B., Oda, M. N., Vaisar, T., Oram, J. F., & Heinecke, J. W. (2006). Pathways for oxidation of high-density lipoprotein in human cardiovascular disease. Curr Opin Mol Ther, 8(3), 198e205. Smith, E. B. (1990). Transport, interactions and retention of plasma proteins in the intima: the barrier function of the internal elastic lamina. Eur Heart J, 11(Suppl E), 72e81. Smith, M. A., Richey Harris, P. L., Sayre, L. M., Beckman, J. S., & Perry, G. (1997). Widespread peroxynitrite-mediated damage in Alzheimer’s disease. J Neurosci, 17(8), 2653e2657. Smith, M. A., Sayre, L. M., Anderson, V. E., et al. (1998). Cytochemical demonstration of oxidative damage in Alzheimer disease by immunochemical enhancement of the carbonyl reaction with 2,4-dinitrophenylhydrazine. J Histochem Cytochem, 46(6), 731e735. Stampfer, M. J., Sacks, F. M., Salvini, S., Willett, W. C., & Hennekens, C. H. (1991). A prospective study of cholesterol, apolipoproteins, and the risk of myocardial infarction. N Engl J Med, 325(6), 373e381. Toshima, S., Hasegawa, A., Kurabayashi, M., et al. (2000). Circulating oxidized low density lipoprotein levels. A biochemical risk marker for coronary heart disease. Arterioscler Thromb Vasc Biol, 20(10), 2243e2247. Ueda, M., Hayase, Y., & Mashiba, S. (2007). Establishment and evaluation of 2 monoclonal antibodies against oxidized apolipoprotein A-I (apoA-I) and its application to determine blood oxidized apoA-I levels. Clin Chim Acta, 378(1-2), 105e111. Undurti, A., Huang, Y., Lupica, J. A., Smith, J. D., Didonato, J. A., & Hazen, S. L. (2009). Modification of high density lipoprotein by myeloperoxidase generates a pro-inflammatory particle. J Biol Chem, 284, 30825e30835. Uno, M., Kitazato, K. T., Nishi, K., Itabe, H., & Nagahiro, S. (2003). Raised plasma oxidised LDL in acute cerebral infarction. J Neurol Neurosurg Psychiatry, 74(3), 312e316. von Eckardstein, A., Nofer, J. R., & Assmann, G. (2001). High density lipoproteins and arteriosclerosis. Role of cholesterol efflux and reverse cholesterol transport. Arterioscler Thromb Vasc Biol, 21(1), 13e27. von Eckardstein, A., Walter, M., Holz, H., Benninghoven, A., & Assmann, G. (1991). Site-specific methionine sulfoxide formation is the structural basis of chromatographic heterogeneity of apolipoproteins A-I, C-II, and C-III. J Lipid Res, 32(9), 1465e1476. Wang, X. S., & Stocker, R. (2008). Detection of specifically oxidized apolipoproteins in oxidized HDL. Methods Mol Biol, 477, 49e63. Wang, X. S., Shao, B., Oda, M. N., Heinecke, J. W., Mahler, S., & Stocker, R. (2009). A sensitive and specific ELISA detects methionine sulfoxide-containing apolipoprotein A-I in HDL. J Lipid Res, 50(3), 586e594. Wautier, M. P., Chappey, O., Corda, S., Stern, D. M., Schmidt, A. M., & Wautier, J. L. (2001). Activation of NADPH oxidase by AGE links oxidant stress to altered gene expression via RAGE. Am J Physiol Endocrinol Metab, 280(5), E685e694. Wolin, M. S. (2000). How could a genetic variant of the p22(phox) component of NAD(P)H oxidases contribute to the progression of coronary atherosclerosis? Circ Res, 86(4), 365e366.
Chapter j 13 Determination of Circulating Native
275
Yla-Herttuala, S., Palinski, W., Rosenfeld, M. E., et al. (1989). Evidence for the presence of oxidatively modified low density lipoprotein in atherosclerotic lesions of rabbit and man. J Clin Invest, 84(4), 1086e1095. Yla-Herttuala, S. (1999). Oxidized LDL and atherogenesis. Ann NY Acad Sci, 874, 134e137. Zhang, Y., Zanotti, I., Reilly, M. P., Glick, J. M., Rothblat, G. H., & Rader, D. J. (2003). Overexpression of apolipoprotein A-I promotes reverse transport of cholesterol from macrophages to feces in vivo. Circulation, 108(6), 661e663.
Index
ABCA1 atherosclerosis role, 204e205 cholesterol efflux mediation and mechanisms, 199e204 expression regulation, 200e201 functional overview, 115 macrophage apoptosis and inflammation role in plaques, 205e207 preb1-high-density lipoprotein production mechanism, 249e250 Tangier disease mutation, 133, 200 ABCG1 atherosclerosis role, 204e205 cholesterol efflux mediation and mechanisms, 201e204 endothelial function, 207 gene, 201 macrophage apoptosis and inflammation role in plaques, 205e207 AD, see Alzheimer’s disease Adiponectin, paraoxonase-1 induction, 107 Aging cholesteryl ester transfer protein role, 48 paraoxonase-1 antioxidant activity effects, 113 reverse cholesterol transport effects, 61 AKT, high-density lipoprotein signaling via scavenger receptor-B1, 165e168 Alcohol, reverse cholesterol transport effects, 67e68 Alzheimer’s disease (AD), cholesteryl ester transfer protein deficiency and protection, 48 apolipoprotein A-I mimetic peptide studies, 189 oxidized high-density lipoprotein levels, 267 Amyloidosis, apolipoprotein A-I variants, 141e143 Anacetrapib, cholesteryl ester transfer protein inhibition, 46
Angiotensin-II, scavenger receptor-B1 response, 166 Antioxidant activity, see Oxidative stress Apolipoprotein A-I amyloidosis variants, 141e143 ApoA-IMilano, 144 atherogenesis prevention, 144 cholesteryl ester transfer protein and release from high-density lipoprotein, 22, 24, 47 cysteine mutants, 143 deficiency coronary artery disease risks, 134, 138 deletion, nonsense, or frameshift mutations, 135e139 variants and pathological phenotypes, 139e143 estrogen effects, 62 functional overview, 134e135 gene, 134 high-density lipoprotein composition, 134 quantification, 35e36 mimetic peptides Alzheimer’s disease studies, 189 aortic lesion formation inhibition, 186e187 diabetes complication reduction, 188e189 lecithin:cholesterol acyltransferase activation, 185e186 nitric oxide synthase and endothelial function improvement, 188 organ transplantation rejection prevention, 189 overview, 185e186 oxidation, 266 structure, 134, 185 Apolipoprotein B, cholesteryl ester transfer protein inhibition effects on lipoproteins, 47e48
277
278
Apolipoprotein E apolipoprotein E-rich high-density lipoprotein cholesteryl ester transfer protein in formation, 41e42 function, 41 measurement, 44 structure, 41 mimetic peptides, 190e192 paraoxonase-1 relationship, 106e107 Apolipoprotein F, lipid transfer inhibition activity, 39 Apolipoprotein J mimetic peptides, 191e192 paraoxonase-1 relationship, 106 Apolipoprotein M, preb1-high-density lipoprotein production role, 251 Apoptosis macrophage ABC transporter roles in plaque, 205e207 overview, 167 phosphatidylinositol 3-kinase/AKT protection, 167 Atherosclerosis cholesteryl ester transfer protein activity, 20e22 high-density lipoprotein prevention mechanisms anti-inflammatory effects, 81, 182 antioxidant effects, 81e82, 182 overview, 22, 79e80, 101 reverse cholesterol transport, 80e81, 181 vasoprotective effects, 81, 181e182 paraoxonase-1 relationship, 114e115 lipoprotein-associated phospholipase A2 inhibition and prevention, 48 macrophage apoptosis ABC transporter roles in plaques, 205e207 overview, 167 phospholipid transfer protein activity, 9e10 reverse cholesterol transport relationship, 22e23, 181 scavenger receptor-B1 protection, 155 Brain, cholesterol half-life, 246 CAD, see Coronary artery disease CETP, see Cholesteryl ester transfer protein
Index
Cholesterol, see also Reverse cholesterol transport; specific lipoproteins excretion, 35 half-life in brain, 246 Cholesteryl ester transfer protein (CETP) aging and longevity role Alzheimer’s disease, 48 infection susceptibility, 49e50 overview, 48 apolipoprotein E-rich high-density lipoprotein formation role, 41e42 function, 41 structure, 41 atherosclerosis role, 20e22 deficiency in humans Alzheimer’s disease protection, 48 effects cholesterol efflux and preb-HDL formation, 40e41 high-density lipoprotein-cholesterol levels, 43e45 lipoprotein metabolism, 41 low-density lipoprotein-cholesterol levels, 43 genetics and ethnic differences, 42e43 malignancy risks, 46 polymorphisms and heart disease protection, 46 environmental effects on levels, 51 expression regulation, 37e38 functional overview, 17e18, 38 gene polymorphisms and environmental interactions, 26e27 high-density lipoprotein remodeling, 4e5, 20e26, 40e41 lipid binding and exchange activity, 18e19, 36 lipid transfer modulators, 39 low-density lipoprotein subclass remodeling role, 40 metabolic syndrome role, 50e51 plasma activity in young women, 38e39 reverse cholesterol transport effects, 40 sex differences, 63 species differences in activity, 18, 21e22 structure, 19, 36 therapeutic targeting, 18, 27e28, 46e48 transgenic mouse studies, 25 Chylomicron (CM), postprandial lipoprotein metabolism, 69e76 CM, see Chylomicron
279
Index
Coronary artery disease (CAD), see also Atherosclerosis apolipoprotein A-I deficiency, 134, 138 cholesteryl ester transfer protein polymorphisms and protection, 46 preb1-high-density lipoprotein levels, 252e253 Dalcetrapib, cholesteryl ester transfer protein inhibition, 46 Diabetes apolipoprotein A-I mimetic peptides complication reduction, 188e189 cholesteryl ester transfer protein role, 50e51 high-density lipoprotein oxidation susceptibility, 218 oxidized high-density lipoprotein levels, 268 paraoxonase-1 and high-density lipoprotein relationship, 90e92, 115 preb1-high-density lipoprotein levels, 254 Diet, fatty acid intake and reverse cholesterol transport effects, 66e67 Endotoxin, cholesteryl ester transfer protein and protection, 49e50 ERASE study, recombinant high-density lipoprotein therapy, 184 Estrogen, reverse cholesterol transport effects, 62e64 Exercise, reverse cholesterol transport effects, 65e66 Ghrelin, high-density lipoprotein binding, 107 HDL, see High-density lipoprotein Hemodialysis paraoxonase-1 as mortality marker, 121 preb1-high-density lipoprotein levels, 253e254 Hepatic lipase, cholesteryl ester transfer protein in substrate generation, 24 High-density lipoprotein (HDL), see also Preb1-high-density lipoprotein apolipoprotein A-1 for quantification, 35e36 apolipoprotein E-rich high-density lipoprotein
cholesteryl ester transfer protein in formation, 41e42 function, 41 measurement, 44 structure, 41 atherosclerosis prevention mechanisms anti-inflammatory effects, 81, 182 antioxidant effects, 81e82, 182 overview, 22, 79e80, 101 reverse cholesterol transport, 80e81, 181 vasoprotective effects, 81, 181e182 cholesteryl ester transfer protein high-density lipoprotein-cholesterol levels in deficiency, 43e45 remodeling role, 4e5, 20e26, 40e41 estrogen effects, 62 heterogeneity, 181 lipid composition, 244e246 oxidation, see Oxidized high-density lipoprotein paraoxonase-1 relationship anti-atherogenic activity, 114e115 cardiovascular disease, 85e90 diabetes, 90e92 diseases altering high-density lipoprotein levels, 115e118 healthy subjects, 83e85, 112e113 neighbor proteins, 106e107 particle size, 246 phospholipid transfer protein in remodeling, 1e2, 4e10 pro-inflammatory effects, 183 protein composition, 78 proteomics and protein composition, 111e112, 244e246 subclasses, 78e79, 102 subfraction separation and analysis, 246e247 therapy, 183e185 HNE, see 4-Hydroxy-2-noneal HPBP, see Human phosphate binding protein Human phosphate binding protein (HPBP) function, 114e115 paraoxonase-1 interactions, 99, 107e108 structure, 108e109 4-Hydroxy-2-noneal (HNE), high-density lipoprotein oxidation, 223 Inflammation ABC transporter roles in plaques, 205e207 high-density lipoprotein
280
Inflammation (Continued ) anti-inflammatory effects, 81, 182 pro-inflammatory effects, 183 oxidixed high-density lipoprotein effects, 229 paraoxonase-1 levels, 116 Insulin, reverse cholesterol transport effects, 64e65 Lactone, paraoxonase-1 hydrolysis in infection, 116 Lapropiprant, nicotinic acid side effect control, 38 LCAT, see Lecithin:cholesterol acyltransferase LDL, see Low-density lipoprotein Lecithin:cholesterol acyltransferase (LCAT) apolipoprotein A-I mimetic peptide activation, 185e186 deficiency, 133 functional overview, 106 plasma activity in young women, 38e39 Lipoprotein lipase (LPL), preb1-high-density lipoprotein production role, 250 Lipoprotein-associated phospholipase A2 (PAF-AH), inhibition and atherosclerosis prevention, 48 Lipoxygenase, high-density lipoprotein oxidation role, 217e218 Low-density lipoprotein (LDL) cholesteryl ester transfer protein effects low-density lipoprotein-cholesterol levels, 43 subclass remodeling, 40 oxidation, 262 receptor knockout mice and phospholipid transfer protein effects, 6e8 LPL, see Lipoprotein lipase Macrophage apoptosis in atherosclerosis ABC transporter roles in plaques, 205e207 overview, 167 cholesterol transport, see Reverse cholesterol transport MAPK, see Mitogen-activated protein kinase Metabolic syndrome, cholesteryl ester transfer protein role, 50e51
Index
Mitogen-activated protein kinase (MAPK), high-density lipoprotein signaling via scavenger receptor-B1, 163e165 Myeloperoxidase, high-density lipoprotein oxidation role, 217 NAD(P)H oxidase p22phox, polymorphism and high-density lipoprotein oxidation, 266e267 Nicotinic acid, lipid-lowering activity, 38 Nitric oxide synthase (NOS) ABCG1 and endothelial function, 207 apolipoprotein A-I mimetic peptide and endothelial function improvement, 188 high-density lipoprotein modulation, 81, 166, 181e182 oxidixed high-density lipoprotein effects, 228e229 NOS, see Nitric oxide synthase Organ transplantation, apolipoprotein A-I mimetic peptides and rejection prevention, 189 Oxidative stress high-density lipoprotein antioxidant effects, 81e82, 182 high-density lipoprotein oxidation, see Oxidized high-density lipoprotein Oxidized high-density lipoprotein Alzheimer’s disease, 267 apolipoprotein A-I, 266 apolipoprotein alterations, 219e221 carotid endarterectomy studies, 268 cellular uptake and clearance, 268e270 diabetes, 268 functional effects anti-inflammatory effect, 229 antioxidant activity, 226e228 nitric oxide synthase, 228e229 reverse cholesterol transport, 224e226 vascular function, 229e230 lipid oxidation, 222e223 measurement high-performance liquid chromatography, 265 immunoassay, 263e265 NAD(P)H oxidase p22phox polymorphism, 266e267
281
Index
oxidation mechanisms, 217e219 susceptibility, 216e217 structural alterations, 219 p22phox, see NAD(P)H oxidase p22phox PAF, see Platelet-activating factor PAF-AH, see Lipoprotein-associated phospholipase A2 Paraoxonase-1 (PON-1) alcohol effects on activity, 67 antioxidant activity and aging, 113 developmental changes in activity, 104 functional overview, 82, 102 gene, 104 gene therapy, 128 high-density lipoprotein relationship anti-atherogenic activity, 114e115 cardiovascular disease, 85e90 diabetes, 90e92, 115 diseases altering high-density lipoprotein levels, 115e118 healthy subjects, 83e85, 112e113 neighbor proteins, 106e107 human phosphate binding protein interactions, 99, 107e108 inducers, 119e120 lifestyle modification effects on levels, 118 organophosphate detoxification, 99e101 polymorphisms, 82e84, 104, 111 purification, 110e111 structure, 104e106 substrate specificity, 99, 102e104 PC, see Phospatidylcholine PDZK1, scavenger receptor-B1 interactions, 160e162 Peroxynitrite, high-density lipoprotein oxidation role, 218 Phospatidylcholine (PC), oxidation in high-density lipoprotein, 222 Phosphatidylinositol 3-kinase (PI3K), high-density lipoprotein signaling via scavenger receptor-B1, 165e168 Phospholipase A2, see Lipoproteinassociated phospholipase A2 Phospholipid transfer protein (PLTP) atherosclerosis role, 9e10 functional overview, 1e2 gene, 2e3 high-density lipoprotein remodeling, 1e2, 4e5
isoforms, 3e4 knockout mouse studies, 6e8 lipid transfer activity, 4 structure, 3 transgenic mouse studies, 5e6 triglyceride-rich lipoprotein remodeling, 5 Phospholipid transfer protein (PLTP), functional overview, 39 PI3K, see Phosphatidylinositol 3-kinase PKC, see Protein kinase C Platelet-activating factor (PAF), acetylhydrolase, 106 PLTP, see Phospholipid transfer protein PON-1, see Paraoxonase-1 Preb1-high-density lipoprotein assays, 251e252 cholesteryl ester transfer protein deficiency and formation, 40e41 fractionation, 247 functional overview, 247e249 levels in disorders, 252e254 metabolism, 247e249 production ABCA1-dependent mechanism, 249e250 ABCA1-independent mechanism, 250e251 turnover, 254e255 Protein kinase C (PKC), high-density lipoprotein signaling via scavenger receptor-B1, 162e163 RCT, see Reverse cholesterol transport Recombinant high-density lipoprotein, therapy, 184 Renal failure, paraoxonase-1 levels, 116 Reverse cholesterol transport (RCT) aging effects, 61 alcohol effects, 67e68 atherosclerosis relationship, 22e23, 181 cholesteryl ester transfer protein effects, 40 dietary fatty acid effects, 66e67 exercise effects, 65e66 high-density lipoprotein effects, 80e81 hormone effects estrogen, 62e64 insulin, 64e65 testosterone, 64 thyroid hormone, 64
282
Index
Reverse cholesterol transport (RCT) (Continued ) oxidixed high-density lipoprotein effects, 224e226 postprandial state effects, 69e71 smoking effects, 68e69
structure, 156 subcellular localization, 156e159 testosterone effects on expression, 64 Smoking, reverse cholesterol transport effects, 68e69 SR-B1, see Scavenger receptor-B1
Scavenger receptor-B1 (SR-B1) atherosclerosis prevention, 155 estrogen effects on expression, 63 expression regulation, 161 high-density lipoprotein signaling mechanisms mitogen-activated protein kinase, 163e165 overview, 161e162 phosphatidylinositol 3-kinase/AKT, 165e168 protein kinase C, 162e163 uptake and efflux mediation, 155e156, 159 knockout mouse studies, 153e155 multimerization, 160 PDZK1 interactions, 160e162
Tangier disease, features, 133, 200 Testosterone, reverse cholesterol transport effects, 64 TGRLP, see Triglyceride-rich lipoprotein Thyroid hormone, reverse cholesterol transport effects, 64 a-Tocopherol, oxidation in high-density lipoprotein, 223 Torcetrapib, cholesteryl ester transfer protein inhibition, 46 Triglyceride-rich lipoprotein (TGRLP), phospholipid transfer protein in remodeling, 5 Ubiquinol-10, oxidation in high-density lipoprotein, 223e224