Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Progress in

Lipid Research
Progress in Lipid Research 43 (2004) 350–380
www.elsevier.com/locate/plipres

Review

Contributions of domain structure and lipid interaction to


the functionality of exchangeable human apolipoproteins
Hiroyuki Saito, Sissel Lund-Katz, Michael C. Phillips *

Lipid Research Group, The Children’s Hospital of Philadelphia, Abramson Research Center, Suite 1102,
3615 Civic Center Boulevard, University of Pennsylvania School of Medicine, Philadelphia, PA 19104-4318, USA

Abstract

Exchangeable apolipoproteins function in lipid transport as structural components of lipoprotein par-


ticles, cofactors for enzymes and ligands for cell-surface receptors. Recent findings with apoA-I and apoE
suggest that the tertiary structures of these two members of the human exchangeable apolipoprotein gene
family are related. Characteristically, these proteins contain a series of proline-punctuated, 11- or 22-amino
acid, amphipathic a-helical repeats that can adopt a helix bundle conformation in the lipid-free state. The
amino- and carboxyl-terminal regions form separate domains with the latter being primarily responsible for
lipid binding. Interaction with lipid induces changes in the conformation of the amino-terminal domain
leading to alterations in function; for example, opening of the amino-terminal four-helix bundle in apo-
lipoprotein E upon lipid binding is associated with enhanced receptor-binding activity. The concept of a
two-domain structure for the larger exchangeable apolipoproteins is providing new molecular insights into
how these apolipoproteins interact with lipids and other proteins, such as receptors. The ways in which
structural changes induced by lipid interaction modulate the functionality of these apolipoproteins are
reviewed.
Ó 2004 Elsevier Ltd. All rights reserved.

Abbreviations: ABCA1, ATP-binding cassette transporter A1; ANS, 8-anilino-1-naphthalenesulfonic acid; apo,
apolipoprotein; CD, circular dichroism; CE, cholesteryl ester; DMPC, 1,2-dimyristoyl phosphatidylcholine; FRET,
fluorescence resonance energy transfer; FC, free cholesterol; HDL, high density lipoprotein; HSPG, heparan sulfate
proteoglycan; LCAT, lecithin: cholesterol acyltransferase; LDL, low density lipoprotein; LDLR, LDL receptor; LRP,
LDL receptor-related protein; PL, phospholipid; SR-BI, scavenger receptor class B type I; VLDL, very low density
lipoprotein.
*
Corresponding author. Tel.: +1-215-590-0587; fax: +1-215-590-0583.
E-mail address: phillipsmi@email.chop.edu (M.C. Phillips).

0163-7827/$ - see front matter Ó 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.plipres.2004.05.002
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 351

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351

2. Structures of apoA-I and apoE in the lipid-free state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 352


2.1. Primary and secondary structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 352
2.2. Tertiary structure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
2.3. Quaternary structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357

3. Interaction of apoA-I and apoE with lipids. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 358


3.1. Molecular mechanism of lipid-binding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 358
3.2. Lipid-bound conformation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359

4. Other members of the exchangeable apolipoprotein family . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 362


4.1. ApoA-II . . . . . . . . . . . . . . . . . . . . . . . . . . . ........................................ 363
4.2. ApoA-IV. . . . . . . . . . . . . . . . . . . . . . . . . . . ........................................ 363
4.3. ApoA-V . . . . . . . . . . . . . . . . . . . . . . . . . . . ........................................ 364
4.4. ApoC-I, C-II and C-III . . . . . . . . . . . . . . . . ........................................ 364

5. Structure–function relationships . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366


5.1. ApoE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
5.2. ApoA-I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 367

6. Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372

1. Introduction

The exchangeable apolipoproteins (apo) are part of a multigene family [1] and members of
the A, C and E classes of apolipoproteins play critical roles in lipoprotein metabolism. In
particular, they are structural components of lipoprotein particles and their ability to interact
with lipids is central to this function. ApoA, C and E regulate the transport of lipids into and
out of cells by acting as cofactors for plasma enzymes and ligands for cell-surface receptors.
The structures of human apoA-I and apoE have been studied the most extensively but many
aspects of their structures are poorly defined. This lack of molecular detail has limited un-
derstanding of the structure–function relationships of these proteins. Recently, we have shown
that apoA-I and apoE adopt similar conformations, as might be expected for members of the
same gene family. This new understanding of tertiary structure is providing important insights
into the molecular basis for the biological functions of apoA-I and apoE. The aim of this
review is to summarize recent advances in this area. Additional background information can
be obtained from some excellent reviews on apoA-I and apoE that have been published in the
last several years [2–7].
352 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

2. Structures of apoA-I and apoE in the lipid-free state

2.1. Primary and secondary structure

Exchangeable apolipoproteins (apoA, C and E) have the same genomic structure and are
members of multigene family that probably evolved from a common ancestral gene [1]. The last
exon codes for a primary structure of 11- and 22-amino acid tandem repeats that span residues
44–243 in apoA-I and 62–299 in apoE (Fig. 1(a)) [1,8]. Each of these repeats has the periodicity of
an amphipathic a-helix and these helices are often separated by a proline residue [8,9]. The am-
phipathic a-helices have been classified into several distinct classes according to the distribution of
charged residues around the axis of the helix [10]. The class A helix is a major lipid-binding motif
of exchangeable apolipoproteins and is characterized by the location of basic residues near the
hydrophilic/hydrophobic interface and acidic residues clustered at the center of the polar face
(Fig. 1(b)) [11]. Class G* and class Y helices have also been identified in the exchangeable apo-
lipoproteins and these types of amphipathic a-helix are proposed to have reduced lipid affinity
[12]. The class G* helix is distinguished by a random radial arrangement of positive and negative
amino acid residues on the polar face, while the class Y helix is characterized by the presence of
three clusters of basic amino acids in the polar face forming a Y pattern (Fig. 1(b)) [11].

Fig. 1. (a) Distribution of amphipathic a-helices in the human exchangeable apolipoproteins,apoA-I, apoA-IV and
apoE. The letter ‘‘P’’ below the rectangles indicates positions of all proline residues. (b) Amphipathic helix classes found
in the exchangeable apolipoproteins. Classification is based on the distribution of charged residues (see Section 2.1).
These figures were adapted from Segrest et al. [8].
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 353

Human apoA-I is a 243 amino acid protein (molecular mass ¼ 28.1 kDa) in which exon 3
encodes an N-terminal domain (residues 1–43) and the remaining region coded by exon 4 is
predicted to contain eight 22-mer and two 11-mer amphipathic a-helices with most of the helices
being punctuated by prolines (Fig. 1(a)) [1,3]. Comparison of sequences amongst mammals in-
dicates that the N-terminal region of apoA-I is highly conserved while the central and C-terminal
regions show conservative substitutions between species [5]. Despite changes in the primary
structure of apoA-I amongst different species, the secondary structure appears to be more con-
served [13]. Studies of synthetic peptides corresponding to each of the 22-residue amphipathic
segments of human apoA-I have shown that the first (residues 44–65) and last (residues 220–241)
repeat helices have the greatest lipid affinity [14]. The C-terminal region of human apoA-I is very
hydrophobic as revealed by hydropathy analysis of the amino acid sequence (Fig. 2(a)) [15].
Furthermore, recent studies using specific mutations in the C-terminal region [16] and site-directed
spin-label electron paramagnetic resonance spectroscopy [17] indicate that there is a stable helical
structure in the extreme C-terminus of apoA-I. Interestingly, the N-terminal region (residues 1–43)
also contain a very hydrophobic segment centered around residue 18 (Fig. 2(a)), suggesting that
this region also has significant lipid binding ability [18,19].
Human apoE is a single polypeptide chain of 299 amino acid residues with a molecular mass of
34.2 kDa [20]. There is a high degree of sequence conservation across species with exceptions at
the N- and C-termini: homology begins in the vicinity of residue 26 in the human sequence and
continues to approximately residue 288 [2]. Human apoE exists in three major isoforms, apoE2,
apoE3 and apoE4, each differing by cysteine and arginine at positions 112 and 158. ApoE3, the
most common form, contains cysteine and arginine at these positions, respectively, whereas
apoE2 contains cysteine and apoE4 contains arginine at both sites [21]. Physical and biochemical
analyses have demonstrated that there are two independently folded domains in apoE: a 22-kDa
N-terminal domain (residues 1–191) and a 10-kDa C-terminal domain (residues 216–299) [22,23]
(cf. Section 2.2). The N-terminal domain contains the low density lipoprotein (LDL) receptor-
binding region (cf. Section 5) and the C-terminal domain has a high affinity for lipid and is re-
sponsible for lipoprotein binding [2]. Secondary structure predictions indicate that both domains
are highly helical and the N-terminal and the C-terminal domains contain predominantly class G*
and class A amphipathic helices, respectively (Fig. 1(a)) [8]. The hydropathy plot (Fig. 2(b)) shows
that the extreme C-terminus as well as part of the linker region (residues 192–215) are relatively
hydrophobic.

2.2. Tertiary structure

High-resolution structures for intact human apolipoproteins in the lipid-free state are not
available to date. ApoE is perhaps the best defined because the crystal structure for the 22-kDa
N-terminal fragment was solved by X-ray crystallography over a decade ago (Fig. 3) [24]. This
domain forms a globular bundle of four elongated a-helices in which the helices are arranged in an
antiparallel manner with their hydrophobic faces oriented towards the interior of the bundle. This
structure shares the same basic architecture as the five-helix bundle of insect apolipophorin III
[25]. The N-terminal fragments of all three apoE isoforms adopt such a four-helix bundle motif,
but subtle differences in the side-chain conformations and salt-bridge arrangements of the iso-
forms affect their functions and characteristics [26–28]. Studies of guanidine [29] and thermal [30]
354 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

2
(a): apoA-I
1

-1

-2

-3
0 50 100 150 200 250
2
Nonpolar

(b): apoE3
1

-1

-2
Polar

-3
0 50 100 150 200 250 300
2
(c): apoA-IV
1

-1

-2

-3

0 50 100 150 200 250 300 350


Position

Fig. 2. Hydropathy plots of the amino acid sequences of human apoA-I, apoE3 and apoA-IV. Hydropathy was cal-
culated as described by Kyte and Doolittle [15] using a sliding window of nine residues.

denaturation demonstrated that the N-terminal fragments of the apoE isoforms differ in stability
(apoE4 < apoE3 < apoE2), indicating that replacing cysteine residues by arginine results in a
cumulative decrease in hydrophobicity and stability of the helix bundle (Table 1). The stabilities of
the N-terminal fragments are rather similar to that of other globular proteins [29]. The stability of
the C-terminal fragment is lower and similar to that of other apolipoproteins such as apoA-I
(Table 1): the structural organization of this domain is not well characterized. Recent 8-anilino-1-
naphthalenesulfonic acid (ANS) fluorescence measurements have suggested that the C-terminal
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 355

Fig. 3. Ribbon model of the structure of the 22-kDa fragment of human apoE3. Four of the five helices are arranged in
an antiparallel four-helix bundle. Reproduced with permission from [2].

domain of apoE forms a more solvent-exposed, less organized structure (Table 1) [31]. A model
has been proposed in which the helices in the C-terminal domain form an inter-molecular coiled-
coil helix structure [32].
A variety of studies using deletion and site-directed mutagenesis techniques have provided
important insights into the lipid-free structure of apoA-I (for reviews, see [4,5]). Proteolysis
analysis [33] and deletion mutagenesis studies [34,35] have suggested that the lipid-free apoA-I
molecule is organized into two structural domains; the N-terminal and central parts form a helix
bundle whereas the C-terminal helices form a separate, less organized structure (Fig. 4). The helix
bundle organization in the N-terminal domain is also supported by fluorescence studies of single
tryptophan mutants of human [36] and chicken apoA-I [37]. The crystal structure of N-terminal
truncated D1–43 apoA-I revealed a-helices arranged in a linear fashion rather than in a bundle but
this structure is postulated to represent more closely a lipid-bound conformation of apoA-I [38].
Such a helix bundle motif of the N-terminal domain in apoA-I is basically similar to apoE except
for the N-terminal helix bundle being less organized and less stable than that in apoE. As shown
356 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

Table 1
Parameters characterizing the domains of apoA-I and apoE isoforms
Apolipoprotein Average hydropathy per Tm b (°C) D1=2 c (M) ANS fluorescence
residuea ()kcal/mol) intensityd
apoE2 0.69 52 1.0, 2.7 (biphasic) 1.2
apoE3 0.71 46 1.0, 2.4 (biphasic) 1.2
apoE4 0.73 45 1.6 (biphasic but 1.4
overlapping)
apoE2 22 kDa 0.70 63 2.8 0.5
apoE3 22 kDa 0.74 57 2.5 0.4
apoE4 22 kDa 0.78 50 1.6 0.4
apoE 10 kDa 0.68 56 1.1 1.3
apoA-I 0.80 60 1.0 1.0
D190–243 apoA-I 0.93 56 1.0 0.6
a
Calculated as described by Kyte and Doolittle [15]. A more negative hydropathy is more polar.
b
Midpoint of thermal denaturation determined by far-UV CD measurements [29,35].
c
Midpoint of guanidine denaturation determined by far-UV CD measurements [29,39].
d
Values are ratios to wild-type apoA-I. Estimated error is within 0.1.

Fig. 4. Model of the two-step lipid binding mechanism of apoA-I on a spherical particle.In the lipid-free state, apoA-I
is organized into two structural domains in which the N-terminal domain forms a helix bundle whereas the C-terminal
domain forms a separate, less organized structure. Initial lipid binding occurs through amphipathic a-helices in the
C-terminal domain accompanied by an increase in a-helicity probably in the region including residues 187–220. Sub-
sequently, the helix bundle in the N-terminal domain undergoes a conformational opening, converting hydrophobic
helix–helix interactions to helix–lipid interactions. Reproduced with permission from [35].

in Table 1, guanidine-induced denaturation curves of apoE indicate that the N-terminal and C-
terminal domains unfold independently [29], whereas guanidine-induced denaturation of apoA-I
is monophasic [39]. In addition, a thermal unfolding study has demonstrated that the apoA-I
molecule exhibits a loosely folded, molten globular-like structure (i.e., the a-helices do not occupy
fixed positions with respect to one another and are not organized into a unique tertiary structure)
[40]. The fact that the N-terminal helix bundle domains in apoA-I and apoE contain different sorts
of amphipathic helix in which the helices are class G* in apoE and class A in apoA-I [8], may
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 357

underlie the differences in stability. The difference in charged group distribution (Fig. 1(b)) may
affect the burial of polar groups which is known to affect helix bundle stability [41]. In addition,
the proline punctuation is much more regular in apoA-I than in apoE (Fig. 1(a)), presumably
preventing the helices in the bundle from becoming as long and highly organized as they are in
apoE (Fig. 3).
The tertiary arrangement of the C-terminal domain in apoA-I is not understood well. An
electron paramagnetic resonance spectroscopic study indicated that the helices in the C-terminal
domain form a compact antiparallel alignment with residues 188–205 existing as a flexible loop
[17], whereas fluorescence resonance energy transfer (FRET) measurements suggest an extended
conformation [42,43]. Regardless of its exact conformation, the C-terminal domain appears to be
relatively disorganized because it contains an exposed hydrophobic surface that is accessible to
ANS binding (see the difference in ANS fluorescence between WT and D190–243 apoA-I in Table 1)
[35]. In addition, an interaction between the N- and C-terminal domains in the apoA-I molecule is
suggested in which the N- and C-termini interact with each other to maintain the overall structure
of the protein molecule [42,44].
The concept of domain interaction was introduced to account for the influence of the poly-
morphism at position 112 in the N-terminal domain of apoE on the lipid-binding properties of the
C-terminal domain [45]. In apoE4, the N- and C-terminal domains interact differently than they
do in the other isoforms: Arg-112 causes a rearrangement of the Arg-61 side chain in the
N-terminal domain of apoE4, allowing it to interact with Glu-255 in the C-terminal domain
[28,46]. This domain interaction in human apoE4 leads to a less organized structure in the
C-terminal domain and preferential association with very low density lipoprotein (VLDL) rather
than high density lipoprotein (HDL) which is contrary to the behavior of apoE3 [2]. The domain
interaction in apoE4 has been suggested to contribute to the accelerated catabolism of this iso-
form and, consequently, the increased cholesterol and LDL levels in the plasma of individuals
with this genotype [47]. A recent FRET study indicates that the N- and C-terminal domains of
lipid-free apoE3 are in a spatially proximate orientation with respect to each other, probably
through weak hydrophobic interaction [48].

2.3. Quaternary structure

Lipid-free apoA-I and apoE have a propensity to self-associate in aqueous solution because of
hydrophobic interaction between amphipathic a-helices in different molecules. Lipid-free apoA-I
tends to form oligomers at concentrations >0.1 mg/ml [49], but the oligomerization is less efficient
for deletion mutants lacking the C-terminal region [34], indicating that the self-association of
apoA-I is mediated by hydrophobic interaction between the C-terminal helices. Similarly, apoE
exists as a tetramer in the lipid-free state [50] and this tetramerization is thought to be mediated by
the C-terminal domain because the 22-kDa fragment is monomeric while the 10-kDa fragment is
tetrameric in aqueous solution [23]. Sedimentation equilibrium and C-terminal truncation anal-
yses indicate that the extreme C-terminal residues 267–299, which are shown to be very hydro-
phobic by hydropathy analysis (Fig. 2(b)), are responsible for the self-association of apoE [51].
Thus it can be inferred that self-association of apoA-I and apoE promotes stabilization of the
potential a-helical segments of the C-terminal domain that are less organized in the monomeric
358 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

form. Interestingly, apoE4 having a more solvent-exposed, less organized structure of the
C-terminal domain [31] shows a greater propensity to self-associate compared to apoE3 [52].

3. Interaction of apoA-I and apoE with lipids

3.1. Molecular mechanism of lipid-binding

It is well established that the C-terminal domain is critical for lipid-binding of both apoA-I and
apoE molecules. In apoE, this domain is the 10-kDa fragment corresponding to residues 216–299
[2], and it contains three of the proposed 22-mer repeats [1]. Assuming from the homology in the
two domain structure of both molecules that the equivalent three C-terminal 22- mer repeats also
comprise the lipid-binding domain in human apoA-I, it follows that this domain spans residues
187–243 [1]. This notion about the domain boundary in the apoA-I molecule is consistent with
previous observations that the C-terminal 187–243 region is necessary for the initial lipid-binding
of apoA-I [34,53,54] and, also, with previous studies employing limited proteolysis of lipid-free
apoA-I that revealed major cleavage sites in the C-terminus particularly near residue 192 [33,53].
Upon binding of apoE to lipids, it has been proposed that the four-helix bundle in the
N-terminal domain undergoes a conformational reorganization to expose the hydrophobic faces
of its amphipathic helices for interaction with lipid molecules [2]. Molecular area measurements at
an air–water interface indicated that the N-terminal domain occupies a larger surface area than
can be accounted for by its globular four-helix bundle conformation, suggesting adoption of an
open conformation by the helix bundle [55]. Subsequent studies using infrared spectroscopy [56],
FRET [57], and interhelical disulfide mutants of the apoE N-terminal domain [58] confirmed that
the four-helix bundle undergoes conformational opening when apoE is complexed with
phospholipid (PL). Kinetic studies of 1,2-dimyristoyl phosphatidylcholine (DMPC) multilamellar
vesicle solubilization by the isoforms of apoE and their N- and C-terminal domains revealed that
the C-terminal fragment is much more reactive than the N-terminal fragment [59]. Furthermore,
the reaction rates of the N-terminal fragments of the isoforms appear to vary inversely with the
stabilities of these fragments [59–61]. In addition, the two domains in full-length apoE molecule
were shown to be not fully available to interact with lipid probably due to conformational re-
striction. This is consistent with the finding that the N- and C-terminal domains of lipid-free
apoE3 are in a spatially proximate orientation with respect to each other [48].
The flexible molten globular-like structure of apoA-I explains its lipid-binding properties [4].
The loose arrangement of a-helices allows rapid interaction of exposed hydrophobic regions of the
protein with lipid, leading to rapid solubilization of DMPC vesicles to form discoidal HDL
particles [59,62]. Based on the finding that the first and last 22-mer a-helices have the greatest lipid
affinity (cf. Section 2.1), a model of association of apoA-I with lipid was proposed in which the
two helices nearest the ends of the molecule bind to lipid first, facilitating binding of the middle
helical domains that have relatively low lipid affinity [14]. Combining this information with
structural studies of lipid-free apoA-I showing that the protein can adopt both globular helical
bundle and elongated rod-like conformations, Rogers et al. [63] proposed a multiple-step mech-
anism for apoA-I lipid-binding to form discoidal complexes. In their model, the initial binding to
lipid occurs through the C-terminal region in the elongated hairpin structure, followed by a
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 359

conformational switch of residues 1–43 which unmasks a latent lipid-binding domain comprising
residues 44–65.
Using a series of deletion mutants that progressively lacked different regions along the mole-
cule, we have shown recently that binding of apoA-I to lipids is modulated by reorganization of
the N-terminal helix bundle structure [35]. Given that the two-domain structure in apoA-I is like
that in apoE (Section 2.2), we proposed a two-step mechanism for binding of apoA-I to spherical
lipid particles (Fig. 4) [35]. In this model, apoA-I initially binds to a lipid surface through am-
phipathic a-helices in the C-terminal domain; this process is accompanied by an increase in
a-helicity in the region including residues 187–220. Subsequently, the helix bundle in the N-ter-
minal domain undergoes a conformational opening, converting hydrophobic helix–helix inter-
actions to helix–lipid interactions. In this step, the extra helicity in the N-terminal domain appears
to come from the residues 123–142 [64] as well as 1–43 [18,65,66]. Given the structural similarity of
apoA-I and other exchangeable apolipoproteins such as apoE that are part of the apolipoprotein
multigene family [1], the two-step lipid binding mechanism could be a general feature for lipid
interaction of exchangeable apolipoproteins composed of two structural domains.
The conformational transition from random coil to a-helix upon lipid-binding is thought to
provide the energetic source to drive the lipid interaction of apolipoproteins [67]. Calorimetric
measurements of binding of apoA-I [35,68,69], apoE [70], and apoA-I model peptides [71] to lipid
particles indicate that lipid-binding of these proteins is accompanied by a large exothermic heat,
consistent with the lipid-binding of apolipoproteins being enthalpically driven. We have shown
recently that the contribution of a-helix formation to the enthalpy of binding of apoA-I to egg PC
small unilamellar vesicles is )1.1 kcal/a-helical residue [66], in agreement with an early estimate
()1.3 kcal/a-helical residue) for plasma apolipoproteins using different membrane systems [67].
a-Helix formation in apoA-I also contributes to an increase in the favorable free energy of binding
to lipid ()42 cal/a-helical residue), leading to an approximately two-order of magnitude increase
in the affinity of binding (Fig. 5). This indicates that the transition of random coil to a-helix plays
a critical role in driving apoA-I to interact with a lipid surface. This phenomenon probably ex-
plains why many exchangeable apolipoproteins in the lipid-free state contain random coil
structure especially in the lipid-binding domain, such as the C-terminal domain of apoA-I and
apoE [2,4].

3.2. Lipid-bound conformation

ApoA-I is the major protein component of both nascent discoidal and circulating spherical HDL
particles [72]. HDL discs comprise a segment of PL bilayer surrounded at the edge by apoA-I
molecules. Two general models have been proposed for the organization of apoA-I molecules in
such discoidal HDL particles: a ‘‘picket fence’’ model in which the a-helix repeats of apoA-I are
arranged parallel to the PL acyl chains [73] and a ‘‘belt’’ model where a continuous series of
a-helices of apoA-I are aligned perpendicular to the acyl chains [74]. Since the determination of the
X-ray crystal structure of the N-terminal truncated D1–43 apoA-I molecule that consists of con-
tinuous amphipathic a-helices punctuated by kinks at the regularly spaced proline residues [38],
several pieces of experimental evidence have supported the conclusion that lipid-bound apoA-I
adopts a ‘‘belt-like’’ arrangement around the edge of discoidal complexes. Polarized internal re-
flection infrared spectroscopy [75], fluorescence parallax analysis [76,77], and FRET [78] studies
360 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

Fig. 5. Schematic representation describing the thermodynamics of binding of apoA-I to lipids. A hypothetical free
energy of binding for apoA-I that does not undergo a-helix formation upon lipid binding has a value of )9.7 kcal/mol
that corresponds to a Kd value of 120 lg/ml (4.3 lM). The increase in a-helix content when apoA-I binds to lipid
contributes )2.5 kcal/mol to the free energy of binding (possible regions forming a-helices are depicted by hatched
cylinders). This additional free energy leads to a reduction in Kd to 2 g/ml (0.06 lM) that is actually measured in the
binding of apoA-I to small unilamellar vesicles [66].

have confirmed that the orientation of most of the apoA-I helices is perpendicular to the PL acyl
chains. Based on the ‘‘belt’’ model, Segrest et al. [79] have proposed a ‘‘double belt’’ model for a
discoidal HDL complex containing two molecules of apoA-I. In this model, two ring-shaped
molecules of apoA-I are stacked on top of each other with both molecules in an anti-parallel
orientation, allowing the helix registry to maximize intermolecular salt-bridge interactions
(Fig. 6(a)). In contrast to this extended belt-like conformation, a helical ‘‘hairpin’’ model for apoA-
I in discoidal HDL particles has been proposed [78,80]. The helical ‘‘hairpin’’ model offers an
explanation for the discrete particle size heterogeneity observed in reconstituted HDL particles; in
particular, it provides a model for discoidal reconstituted HDL particles containing three apoA-I
molecules [74]. A recent mass spectroscopic study indicated the presence of the salt-bridge inter-
actions between apoA-I molecules present in both the ‘‘double belt’’ and ‘‘hairpin’’ models [81].
Two models of the helix organization of apoE in discoidal particles have also been proposed:
the ‘‘picket fence’’ model [82] and the ‘‘belt’’ model [56]. A recent FRET analysis of the apoE
N-terminal domain in discoidal complexes suggests that the helix bundle opens to adopt a par-
tially extended conformation [83], consistent with the ‘‘belt’’ model. More recently, a fluorescence
parallax analysis study showed that the a-helices in both the N-terminal and the C-terminal
domains of apoE align perpendicular to the acyl chains of PL bilayers (Fig. 6(b)) [84]. It now
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 361

Fig. 6. (a) Double belt model for apoA-I structure at the edge of discoidal HDL complex. Two ring-shaped molecules
of apoA-I are stacked on top of each other with both molecules in an anti-parallel orientation, allowing the helix
registry to maximize intermolecular salt-bridge interactions. Only the charged residues at selected positions are ex-
plicitly displayed; basic residues are represented in blue, acidic residues in red, and prolines in green. Reproduced with
permisson from [79]. (b) Model of apoE in discoidal HDL complex depicting the locations of engineered tryptophan
residues on helix 4. Fluorescence from these amino acids was monitored to determine helix orientation. Two molecules
of apoE (blue and gray) of a total of about 4 molecules/particle are depicted in which the helical axes are orientated
perpendicular to the PL acyl chains. Reproduced with permission from [84].

seems clear that the perpendicular orientation of amphipathic helical domains with respect to PL
acyl chains in discoidal particles is a common feature of amphipathic exchangeable apolipopro-
teins.
Spherical HDL particles are quite heterogeneous in size and composition [72]. It is likely that
the apoA-I in spherical HDL is flexible with its conformation being governed by the size and core
lipid composition of the particles [85,86]. On the spherical surface, the amphipathic a-helices of
apolipoproteins are thought to be embedded between PL molecules with their hydrophobic faces
in contact with PL acyl chains [11]. In soluble apolipoproteins such as apoA-I and apoE con-
sisting of a number of different amphipathic helices, the conformational flexibility of the proteins
could allow some helices with low lipid affinity to be excluded from the particle surface [6,8,87].
Indeed, we demonstrated recently that the two domain structure in apoE leads to two different
lipid-bound conformations on lipoprotein-like spherical particles; the N-terminal four-helix
bundle can adopt either open or closed conformations, resulting from binding competition with
362 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

Fig. 7. Model of two possible conformations of apoE on spherical particle. This model proposes that at high apoE
surface concentration, the displacement of the N-terminal domain from the lipid surface by the C-terminal domain
causes the N-terminal four-helix bundle to adopt the closed conformation. At low surface coverage, the four-helix
bundle is open and all the a-helices are in contact with the lipid surface. Reproduced with permission from [70].

the C-terminal domain that has a high lipid affinity [70] (Fig. 7). Given the similar domain
structure of apoA-I and apoE [35], it is conceivable that the structural organizations of apoA-I on
discoidal and spherical particles are also different [66], such as with some helices in apoA-I being
out of contact with lipid on the spherical surface forming a so-called ‘‘hinge domain’’ [8,88]. It
seems likely that helices located around residue 100 are responsible for maintaining the plasticity
of the lipid-bound conformation of apoA-I, thereby allowing apoA-I to form discoidal HDL
particles of different size. Indeed, a recent FRET study suggested that apoA-I undergoes a con-
formational change as it adapts from a discoidal to a spherical surface [89]. The concept of
multiple lipid-bound conformations of apoA-I and apoE may provide new insights into how these
proteins modulate their functions (cf. Section 5).

4. Other members of the exchangeable apolipoprotein family

Compared to human apoA-I and apoE, the structure–function relationships of the other
members of the multigene family, apoA-II, apoA-IV, apoA-V and the apoCs have not been
examined as extensively. However, three-dimensional structures for some of these members of
the apolipoprotein family have been reported in recent years. The elucidation of these
structures has provided some insights into their biological functions and these are summarized
in this section.
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 363

4.1. ApoA-II

Human apoA-II is a major component of HDL particles and it is synthesized mainly in the
liver. Understanding of the contribution of apoA-II to lipid metabolism is much more limited
than that of apoA-I and the functions of apoA-II are still not entirely clear [90]. ApoA-II consists
of two 77 residue-long monomers that form a homodimer linked through a disulfide bridge and
the molecular mass of each monomer is 8.7 kDa [91]. Exon 4 of the apoA-II gene encodes for
residues 40–77 and this region of the protein is predicted to have only one 11-mer and one 22-mer
amphipathic a-helix: this can be compared to the two 11-mer and eight 22-mer helices encoded for
by exon 4 of apoA-I [1]. Three long a-helices make up most of the structure of each apoA-II
monomer [92]. The helices are largely punctuated at prolines, although not uniformly, unlike the
structure of apoA-I. Lipid-free apoA-II has been shown to have a hydrophobic core but is only
marginally stable under near-physiological conditions in plasma; this is consistent with a molten
globular-like state for lipid-free apoA-II [93]. A high-resolution crystal structure of lipid-free
human apoA-II has been published recently [92]. In this structure, each disulfide-bonded ho-
modimer contains three hydrophobic patches, including residues 6–29, 42–53 and 60–70 from
each chain. Two of these dimers associate to form a tetramer and these tetramers are held together
mostly by complementary interactions between their hydrophobic patches.
Interestingly, dimeric apoA-II binds well to PL and, in fact, with higher affinity than does
apoA-I probably because apoA-II is more hydrophobic [94]. Consistent with this enhanced
binding affinity, dimeric apoA-II can displace apoA-I from HDL particles [95–98]. This raises the
question of the role of the apoA-II cysteine 6-cysteine 6 disulfide bridge in human HDL structure
and function. It has been shown that reduction and carboxymethylation of the cysteine residues in
apoA-II to form monomeric apoA-II does not prevent binding to PL and formation of complexes
[99]. Experiments with synthetic peptides have suggested that there are at least two lipid-associ-
ating domains in apoA-II located at opposite ends of the molecule, between amino acids 12–31
and 40–77 [8]. The apoA-II dimer and monomer form discoidal complexes of similar size, with
twice as many of the latter molecules required per disc [100]. Removal of the disulfide bond in-
fluences the stability of the helical segments around the edge of a discoidal complex as seen by a
decrease in a-helix content of the monomeric protein. The discoidal particles containing the
monomeric form of apoA-II are somewhat more effective than particles containing either dimeric
apoA-II or apoA-I in removing cellular cholesterol [100]. Overall, reduction of the disulfide bridge
of apoA-II probably does not have a major effect in determination of HDL particle size in vivo.
The evolution of the cysteine 6–cysteine 6 disulfide bond in higher primates probably has not had
a major effect on the function of the apoA-II molecule.

4.2. ApoA-IV

ApoA-IV is a 46 kDa protein that was discovered in 1974 in rat plasma HDL [101]. In humans,
it is synthesized by the small intestine and secreted into the gastric lymphatics associated with
chylomicrons. It has been proposed to participate in numerous functions in vivo: these functions
include regulation of food intake [102], gastrointestinal motility [103], structural stabilization of
lipoprotein particles [104] and protection against lipid oxidation and atherosclerosis [105]. ApoA-
IV can also mimic certain roles of apoA-I such as in mediation of cholesterol efflux and activation
364 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

of lecithin: cholesterol acyltransferase (LCAT) [106,107]. The amino acid sequences of apoA-I,
apoA-IV and apoE are dominated by multiple 22-amino acid repeats that are predicted to form
amphipathic a-helices (Fig. 1) [1,108]. ApoA-IV contains at least 12 such domains, most of which
are punctuated by proline residues [1,108]. Lipid-free apoA-IV has been demonstrated to be less
stable than apoA-I in terms of its chemical denaturation characteristics [109]. Based on recent
studies in which our laboratory participated, it was concluded that apoA-IV exhibits one large
helix-bundle domain that may be similar to the N-terminal helical bundle domains of apoA-I and
apoE [110]. However, apoA-IV lacks the disordered and hydrophobic C-terminal lipid-binding
domain present in apoA-I and apoE. Indeed, apoA-IV contains a unique glutamine-rich domain
at the C-terminus that actually appears to inhibit the molecule from interacting with lipids and
promoting cholesterol efflux. Since there is no homolog of this sequence in any of the other ex-
changeable apolipoproteins, this domain may have evolved to allow apoA-IV to perform func-
tions that are distinct from other exchangeable apolipoproteins by suppressing its apoA-I-like
functions.

4.3. ApoA-V

Recently, a new family member of the exchangeable apolipoproteins has been identified [111].
Whereas genetic studies clearly demonstrate the important role of human apoA-V in plasma
triglyceride (triacylglycerol) metabolism [111–113], questions remain about the mechanism(s)
whereby this protein exerts its biological effects. The calculated molecular mass of the 366 amino
acid residues is 38.9 kDa and sequence analysis indicates that apoA-V is very hydrophobic [114].
Studies of the physicochemical properties of lipid-free apoA-V indicate that the protein is poorly
soluble between the pH limits of 3.5–9.0 at concentrations >0.1 mg/ml [115]. Like other apo-
lipoproteins such as apoA-I and apoA-IV, apoA-V contains a significant amount of a-helical
structure (32% a-helix, 33% b-sheet, 16% b-turn, and 18% random coil) [114,115]. However,
unlike apoA-I and apoA-IV, the a-helical regions in apoA-V are not composed of homologous
repeating 11- or 22-mer units but instead display considerable heterogeneity in both hydropho-
bicity and length. The observed differences in the hydrophobic properties and the unusual clus-
tering patterns of proline residues in apoA-V indicate that discrete independently folded domains
exerting distinct functional roles may exist within this protein. The association of apoA-V with
large HDL in mouse [111] suggests the presence of lipid-binding domains and, indeed, this has
been demonstrated by the ability of this protein to form discoidal complexes with DMPC [115].
As with other apolipoproteins, there is an increase in a-helix content when apoA-V interacts with
lipid [115]. It is not known which regions of apoA-V are responsible for lipid interaction and self-
association. Unlike apoA-I and apoA-IV that both have the ability to activate LCAT, apoA-V is
a poor activator of this enzyme. More studies will be required to understand the biological role of
this protein at the molecular level.

4.4. ApoC-I, C-II and C-III

The common ancestor of the apolipoprotein genes has been suggested to be very similar to the
apoC-I gene in structure and length. The present day apoC genes evolved through multiple du-
plications (for a review, see [1]). ApoC-I has both inhibitory and stimulatory effects on various
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 365

enzymes and lipid transfer proteins involved in lipoprotein metabolism. Thus, apoC-I deficiency
in HDL may modulate LCAT activity [116] and the protein can inhibit both phospholipase A2
[117] and hepatic lipase [118]. ApoC-I has also been shown to abolish cholesteryl ester transfer
protein activity in a concentration-dependent manner [119]. The majority of apoC-I is associated
with HDL in vivo with no detectable levels found in LDL. The apoE-mediated binding of
b-VLDL to the LDL receptor (LDLR) and LDL receptor-related protein (LRP) is inhibited by
apoC-I [120]. ApoC-I is the smallest exchangeable apolipoprotein (57 amino acid residues and a
molecular mass of 6.6 kDa) and it transfers among HDL, VLDL and chylomicron particles.
There appear to be at least two lipid-associating domains in apoC-I located between residues 1–31
and 32–57 [121]. Exon 4 encodes for only one 11-mer amphipathic a-helix in residues 40–57 and
the structure of apoC-I is believed to be comprised of two dynamic helices that are stabilized by
interhelical interactions and connected by a short linker region. The minimal folding unit in the
lipid-free state of this apolipoprotein comprises a helix-turn-helix motif formed of four 11-mer
sequence repeats [122,123]. The structures of two apoC-I fragments in ‘‘lipid-mimicking’’ sodium
dodecyl sulfate micelles have been solved by NMR [123]. A dramatic increase in a-helix content is
observed for both fragments and the three-dimensional structures correspond to the helix-turn-
helix motif with hydrophobic interactions as the stabilizing force.
ApoC-II is the major activator of lipoprotein lipase, a key enzyme in the regulation of tri-
glyceride levels in human serum [124]. It is a single polypeptide chain of 79 amino acid residues
and has a calculated molecular mass of 8.9 kDa. Exon 4 encodes for two 11-mer amphipathic
a-helices and the full-length apoC-II molecule contains amphipathic a-helical sequences that are
postulated to mediate binding to lipid surfaces and to account for the increase in a-helical content
on lipid binding [8]. The lipid-binding region of the apoC-II molecule is located towards the
N-terminus whereas the lipoprotein lipase-activating domain resides towards the C-terminus. It
seems that once apoC-II is bound to the surface of a lipid particle, formation of a high-affinity
complex with lipoprotein lipase is favored and activation occurs by interaction of the helical
apoC-II domain spanning residues 39–62 with the enzyme [125]. Lipid-free apoC-II is postulated
to exist as a compact but flexible intermediate in solution [126] and NMR spectroscopy has shown
that apoC-II in sodium dodecyl sulfate micelles contains three a-helical regions spanning residues
16–36, 50–56 and 63–77 [127].
ApoC-III is the most abundant C apolipoprotein in human plasma and it is synthesized mainly
in the liver and to a minor extent in the intestine. ApoC-III is the principal plasma inhibitor of
VLDL lipolysis; it acts both by direct inhibition of lipoprotein lipase activity and by interference
with lipoprotein association with cell-surface glycosaminoglycan matrix [128]. ApoC-III also
interferes with remnant lipoprotein clearance [128]. The mature apoC-III protein consists of 79
amino acid residues with a molecular mass of 8.8 kDa. Exon 4 encodes for one 11-mer and one
22-mer amphipathic a-helix [1]. No three-dimensional structure of apoC-III has been reported but
a model structure based on sequence alignment of apoC-III with apoA-II has been described [7].
The model structure exhibits a helix-turn-helix fold similar to that observed in apoC-I and apoC-
II. In addition, the apoC-III model reveals a cluster of C-terminal residues that are essential for
binding of apoC-III to PL. Recently, the lipid-interacting properties of the N-terminal domain of
apoC-III have been investigated [129]. Through the use of molecular modeling, it is predicted that
the fragment corresponding to residues 6–20 of apoC-III is obliquely oriented at the lipid-water
interface due to an asymmetric distribution of the hydrophobic residues along the a-helix axis. It
366 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

is hypothesized that this tilted peptide of apoC-III is responsible for its loose binding to lipo-
protein particles allowing it to exchange freely from the surface.

5. Structure–function relationships

Among the exchangeable apolipoproteins, the structure–function relationships of apoA-I and


apoE have been studied the most extensively. The high level of interest in these two proteins has
arisen because they both exhibit strong anti-atherogenic properties. The insights that under-
standing of the various structural domains in the apoA-I and apoE molecules provides into the
functionalities of these proteins are the focus of this section.

5.1. ApoE

The anti-atherogenic function of apoE arises, in large part, from its profound effect on plasma
lipoprotein levels. As a component of lipoprotein particles, apoE is a high-affinity ligand for the
LDLR and this interaction is followed by endocytosis of the lipoprotein particle [2,130]. This
process leads to hepatic clearance of apoE-containing lipoprotein particles and a reduction in
plasma cholesterol levels. A region in the center of the apoE molecule that is enriched in basic
residues is responsible for binding to acidic regions in the LDLR. Chemical modification and
mutagenesis studies with apoE have demonstrated that a cluster of arginine and lysine residues
located between residues 136–158 represents the binding site for the receptor [2]. The arginine
residue located in position 158 in the apoE3 isoform is mutated to cysteine in apoE2 and the loss
of a basic residue in the latter isoform leads to defective binding to the LDLR [2,130]. It seems
that maintenance of the appropriate positive charge at the C-terminal end of the receptor-binding
region is particularly critical for effective interaction with acidic residues on the LDLR [131].
Disruption of the amphipathic a-helix spanning residues 136–158 abolishes receptor binding,
indicating that this structural motif in apoE is critical for function [132]. As discussed in Section
2.2, the 22-kDa N-terminal domain of apoE adopts a four-helix bundle structure in the lipid-free
state (Fig. 3). The basic residues in the receptor-binding region are located on the polar face of the
fourth amphipathic a-helix in this bundle and they are exposed to the aqueous phase [2]. How-
ever, despite this apparent accessibility of the basic residues, apoE in the lipid-free state does not
bind to the LDLR. Interaction of apoE with PL is necessary to induce a conformational change
that promotes high affinity interaction with the receptor. Lipid-binding leads to opening of the
helix bundle (Section 3.1); this conformational change can increase exposure of lysines 143 and
146 to the aqueous phase [133] and induce helix formation by residues 165–179 [134]. These
changes increase the positive electrostatic potential in the receptor-binding domain of the apoE
molecule enhancing interaction with acidic elements of the LDLR. ApoE is also a ligand for
scavenger receptor class B type I (SR-BI) [135] and, as seen with the LDLR, lipidation of apoE
enhances binding to SR-BI [136]. Since the N-terminal helix bundle domain of apoE is recognized
by SR-BI [136,137], it is likely that opening of this bundle mediates the higher affinity binding of
lipidated apoE to SR-BI.
Since the conformation of an apoE molecule affects its ability to interact with a receptor such as
the LDLR, it is possible that not all apoE molecules on a given lipoprotein particle can function
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 367

as ligands for the receptor. This condition applies to VLDL isolated from hypertriglyceridemic
subjects where only a fraction of the apoE present on the surface mediates binding to the LDLR
[138]. This effect probably arises because apoE can exist in two conformations on the surface of
spherical lipid particles (cf. Section 3.2 and Fig. 7). When there is limited interfacial area available,
the N-terminal bundle is closed and not in contact with the particle surface. ApoE molecules in
this conformation are expected to bind poorly to the LDLR.
Besides being a ligand for the LDLR, apoE also interacts with other members of this receptor
family such as the VLDL receptor, apoE receptor 2, and LRP [139]. Interaction of apoE with the
latter receptor is required for hepatic catabolism of chylomicron remnants in the space of Disse
[130,140]. Heparan sulfate proteoglycans (HSPG) act in concert with LRP to complete the in-
teraction of chylomicron remnants with LRP [130]. ApoE in a lipid-poor state is bound to the
extracellular matrix of hepatocytes [141] where it is available for binding to lipids in chylomicron
remnant particles. This enrichment of the lipoprotein particle with apoE enhances binding to LRP
so that endocytosis can occur. It is likely that the two domains in the apoE molecule (Fig. 7) each
contribute to this process. Both domains of apoE contain a heparin-binding site [2], but the
N-terminal site plays a dominant role in the binding of apoE to heparin (Fig. 8) [142]. Thus, the
C-terminal lipid-binding domain of a lipid-free apoE molecule bound to HSPG on the surface of
hepatocytes is readily available for binding to chylomicron remnants. The cluster of positively
charged amino acid sidechains between residues 136–150 that is involved in binding to the LDLR
is involved in the heparin interaction. In the latter case, formation of the high affinity complex
involves juxtaposition of three arginine and two lysine residues of the apoE a-helix with several
complementary sulfate groups on the heparin molecule [143].
Besides modulating lipid transport, the ability of apoE to interact with members of the LDLR
family and with HSPGs can be significant for cell signalling events. Thus, apoE binding to LRP
activates cAMP-dependent protein kinase A and inhibits platelet-derived growth factor-stimu-
lated migration of smooth muscle cells [144]. On the other hand, inhibition of smooth muscle cell
proliferation by apoE is mediated by its binding to HSPG [145]. The anti-mitogenic effect of apoE
maps to its N-terminal domain and the Cox-2 (cyclooxgenase-2) gene is a target of the apoE
signaling [146].
ApoE is involved in the regulation of cholesterol transport in the central nervous system [139]
where it can contribute to neurodegenerative disorders [130,147]. For instance, apoE is associated
with neuritic amyloid plaques in brains of patients with Alzheimer’s disease. This plaque for-
mation seems to involve an interaction between apoE and b-amyloid peptide. The C-terminal
domain, but not the N-terminal domain, of apoE can complex efficiently with this peptide [148].
Interactions between the N- and C-terminal domains of human apoE (cf. Section 2.2) may
modulate the interaction with b-amyloid peptide because apoE3 and apoE4 behave differently in
this regard [147]. The acquisition of more insights into the biological consequences of the human
apoE4 domain interaction will be easier now that a mouse model for this has been created [149].

5.2. ApoA-I

The anti-atherogenic function of apoA-I arises primarily from its central role in reverse cho-
lesterol transport, the process whereby excess cholesterol is transported from cells in the periphery
back to the liver for eventual secretion from the body [150]. As the principal protein constituent of
368 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

(a)
0.6 lipid-free apoE
WT

K233E

ApoE bound (g/g dried weight of heparingel)


0.4

0.2 K146E

0.0

(b)
0.6 DMPC discs

0.4 K233E

0.2 WT

K146E
0.0
0 100 200 300 400
Free protein (µg/ml)

Fig. 8. Contributions of different apoE domains to heparin binding. The heparin binding site in the N-terminal domain
of human apoE is located between residues 142–147 while the C-terminal site involves basic residues around lysine 233
[142]. The separate 22-kDa N-terminal and 10- kDa C-terminal fragments in the lipid-free state each bind to heparin,
indicating that both sites are functional in this situation. However, the data in this figure show that only the N-terminal
site is functional in either the lipid-free or the lipid-associated intact apoE molecule. Thus, disruption of the cluster of
heparin-binding basic charges by mutation of a lysine residue to glutamic acid greatly reduces binding when the mu-
tation is in the N-terminal site (K146E) but not when it is in the C-terminal site (K233E) (a). Similarly, in discoidal
complexes with DMPC, only the K146E mutation reduces binding relative to wild-type apoE (b). Surprisingly, en-
hanced heparin binding was observed in apoE (K233E)–DMPC complexes and this is probably due to subtle differences
in the open conformation of the N-terminal domain in the DMPC disc. The results are replotted from data in [142].

HDL, apoA-I mediates the following three steps in reverse cholesterol transport. (1) The for-
mation of nascent HDL particles from cellular free cholesterol (FC) and PL via the interaction
with the cell surface transporter, ATP-binding cassette transporter A1 (ABCA1). (2) The con-
version of FC in HDL particles to cholesteryl ester (CE) by acting as a cofactor for the HDL-
associated enzyme, LCAT. (3) The delivery of CE and FC from HDL to the liver by acting a
ligand for SR-BI located on the surface of hepatocytes. The conformational plasticity of apoA-I
(cf. Section 2) underlies its ability to perform these three quite distinct functions. Thus, apoA-I
can bind various amounts of lipid and form HDL particles of different size and shape (cf. Section
3.2), and its functionality is modulated by its physical state. Current understanding of how the
physical state and domain structure of apoA-I contribute to its functionality in reverse cholesterol
transport is outlined below.
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 369

ApoA-I molecules in a lipid-free or lipid-poor state are the preferred acceptor of FC and PL
released from cell plasma membranes by the action of ABCA1 [150,151]. ApoA-I in plasma is able
to cycle between lipid-rich and lipid-poor states [152]. Thus, lipid-free/poor apoA-I molecules
dissociate from HDL during remodeling of the particles by lipid transfer proteins and lipases to
become available to accept lipids from ABCA1. The amphipathic a-helices in apoA-I interact with
ABCA1 and this interaction is not very specific. Thus, ABCA1 can export cellular PL and FC to
different apolipoproteins and peptides [151] and experiments with apoA-I mutants indicate that
different combinations of apoA-I helices can mediate binding to ABCA1 [153]. PL and FC are
released simultaneously to apoA-I [154] and more than one type of HDL particle is created [155].
The larger of these nascent HDL particles are almost certainly discoidal structures (cf. Fig. 6). The
molecular mechanism by which apoA-I acquires lipid from ABCA1 is not defined yet, although
the lipid-binding capability of apoA-I is clearly needed for formation of stable nascent HDL
particles. Consistent with this, deletion of the strongly lipid-binding C-terminal domain of apoA-I
(cf. Sections 2 and 3) leads to a major reduction in ABCA1-mediated FC efflux to apoA-I
(Fig. 9(a)). In comparison, deletions in the helix bundle domain of apoA-I have relatively minor
effects on FC efflux.
The FC in nascent apoA-I-containing HDL particles formed in vivo by the action of ABCA1 is
converted to CE by LCAT. ApoA-I is required for activation of LCAT and its conformation in
the belt-like arrangement in discoidal HDL particles (cf. Fig. 6) is apparently optimal for this
process. Thus, discoidal HDL particles are more reactive with LCAT than the spherical CE-
containing HDL particles that discoidal HDL are converted into by the enzyme [156]. LCAT
binds with higher affinity to discoidal HDL and the initial binding is to lipid surface, followed by
interaction with apoA-I [157]. The LCAT/apoA-I protein/protein interaction decreases the rate
constant of dissociation of LCAT thereby activating the enzyme; this activation may involve
conformational changes in LCAT and/or apoA-I [158]. There has been a lot of interest in es-
tablishing which regions of the apoA-I molecule are involved in this activation. Fig. 9(b) shows
that deletion of either the C-terminal domain (residues 190–243) or certain helices in the helix
bundle domain (residues 139–170) abolishes the ability of apoA-I to activate LCAT. The removal
of the C-terminal domain eliminates the ability of apoA-I to bind to lipid with high affinity
(cf. Section 3.1). Consequently, since apoA-I has to be retained in the surface of the substrate
HDL particle to interact with LCAT, it is possible that the deletion indirectly attenuates the
ability of apoA-I to activate LCAT by reducing the apoA-I surface concentration. In contrast,
removal of residues 139–170 does not alter the lipid affinity of the apoA-I molecule so it is likely
that this region is involved directly in activating LCAT. Consistent with this idea, detailed mu-
tagenesis studies have established that the class A amphipathic a-helix spanning residues 143–164
in human apoA-I is critical for activation of LCAT [159]. The appropriate spatial organization of
specific amino acids in this helix is apparently important because proper alignment and orien-
tation of the helix is required for optimal activation. The fact that three arginine residues (R149,
R153 and R160) located in this helix are critical for activation supports a model in which these
three residues interact electrostatically with acidic residues in the LCAT molecule [160]. This
interaction may be responsible for the ability of apoA-I to retain LCAT on the HDL particle
surface. Given the importance of LCAT in the maturation of HDL particles in plasma, it is to be
expected that this process will be dependent on the presence of apoA-I molecules that can ef-
fectively activate the enzyme. Consistent with the concept, expression in mice of apoA-I mutants
370 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

ABCA1 (a)

3
T

6
5

3
6

3
-4

-6

16

24

24
W

-1
∆1

44

0-
3-

3-
44

22

∆1

∆1


LCAT (b)
Relative activity

0
26

70

43
T
W

-1

2
9-

0-
44

9
∆1

∆1

SR-BI (c)

0
43

66

43

3
T

-6

24
W

-2
-

-1
44
∆1

3-
44

90

22

∆1

Fig. 9. Contributions of various segments of the human apoA-I molecule to its interactions with ABCA1,LCAT and
SR-BI. The activities of human apoA-I mutants with the indicated deletions of N-terminal, central and C-terminal
residues are plotted relative to the activity of the WT protein. (a) ABCA1-mediated efflux of cholesterol from human
fibroblasts incubated for 4 h with 2 lg/ml of the apoA-I variants (C. Vedhachalam, S. Lund-Katz and M.C. Phillips –
unpublished observations). (b) Relative Vmax for formation of cholesteryl ester when reconstituted discoidal HDL
particles (dipalmitoyl PC/cholesterol/variant apoA-I) were incubated. (c) Relative binding to SR-BI of reconstituted
discoidal HDL (palmitoyl-2-oleoyl PC/cholesteryl oleate/variant apoA-I) (S.T. Thuahnai, S. Lund-Katz and M.C.
Phillips – unpublished observations). 30 lg protein/ml of the HDL particles that were of similar size were incubated
with SR-BI-transfected COS-7 cells for 2 h at 37 °C and the binding was determined [135].
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 371

lacking central a-helices in the LCAT-activating domain leads to defective maturation of HDL
particles in vivo [161,162]. Other more C-terminally located a-helices in the helix bundle domain
also play a role in HDL maturation. Thus, the 22-mer repeats spanning residues 165–209 are
important in determining the size distribution of HDL particles [163].
There are several pathways by which the pool of LCAT-generated CE in mature HDL
particles can be delivered to the liver in the reverse cholesterol transport pathway. One such
route involves SR-BI-mediated uptake of the CE into hepatocytes [164,165]. The selective
uptake of CE from HDL occurs by a two-step process that involves (1) binding of HDL to
SR-BI and (2) diffusion of the CE molecules into the cell plasma membrane. ApoA-I is an
important ligand for SR-BI and the receptor can interact with multiple sites in the apoA-I
molecule [136,166]. A specific amino acid sequence in apoA-I is not responsible for the in-
teraction and the class A or class Y amphipathic a-helix (Fig. 1) is the recognition motif. The
data in Fig. 9(c) are consistent with the idea that the SR-BI/apoA-I interaction is not de-
pendent upon a specific site in apoA-I. Thus, deletions of a-helical segments in either the
helix-bundle domain or C-terminal lipid-binding domain have similar effects and decrease
discoidal HDL binding no more than 50%. This is in contrast to the situation with ABCA1
and LCAT where loss of certain helical segments in apoA-I essentially abolishes the activity
(cf. Fig. 9). Association with PL increases the amount of apoA-I binding to SR-BI [136] and
its conformation on discoidal HDL particles modulates the binding [167]. As a consequence,
larger discoidal HDL particles bind better than smaller particles and deliver more CE [168].
SR-BI also facilitates efflux of FC molecules from the cell plasma membrane to bound HDL
particles [150] and this process is sensitive to the conformation of apoA-I because large
discoidal HDL particles promote more efficient efflux than small particles [168]. This is in
contrast to efflux of cellular FC that occurs by simple aqueous diffusion [150] because, in this
case, changes in discoidal HDL particle size have little effect on efflux [169].
Naturally occurring mutations in apoA-I can affect its functionality [5]. As far as the reverse
cholesterol transport pathway is concerned, it is well established that point mutations in and
around the LCAT-activating region of the helix-bundle domain can reduce the ability of apoA-I
to activate the enzyme. Point mutations in the N-terminal helix bundle domain that interfere with
proper folding can also lead to a different dysfunctionality in the apoA-I molecule. Thus, small
N-terminal fragments of such apoA-I variants have been found in individuals with amyloidosis
[5]. The transition from a-helix to b-structure in these peptide fragments is associated with for-
mation of fibrils [170].

6. Summary and conclusions

It is established that the amphipathic a-helical repeats in the exchangeable apolipoproteins are
the key structural elements responsible for the functions of these molecules. However, the ways in
which these helices are organized to create tertiary structures are not well understood. Recently,
we have shown that two of the larger members of the family of exchangeable apolipoproteins,
apoA-I and apoE, share a common two-domain conformation. The relatively hydrophobic and
disordered C-terminal domain is responsible for binding to lipid. The multiple helices located
in the N-terminal and central regions adopt a helix-bundle conformation; this helix bundle is
372 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

relatively ordered and stable in apoE but disordered and unstable in apoA-I. The helix-helix
contacts in the bundle in the lipid-free state are replaced by helix-lipid interactions when binding
to lipid occurs. Thus, lipid binding causes the helix-bundle to open allowing specific a-helices
within this domain to be recognized. For example, the helix spanning residues 136–150 in human
apoE and the helix spanning residues 143–164 in apoA-I can bind to the LDLR and LCAT,
respectively. In comparison, not so much is known about the structures of apoA-IV and apoA-V
which are also larger members of this gene family. However, the lack of a hydrophobic C-terminal
domain in apoA-IV is apparently the reason that this protein binds weakly to lipoproteins.
Smaller members of the gene family, such as the apoC molecules, contain only one or two
amphipathic a-helices so they are unable to form an intramolecular helix-bundle domain.
It can be anticipated that in the future more will be learned about the nature of the helix-helix
interactions in the apolipoprotein molecules. It will be particularly important to generate such
information for these proteins in the lipid-bound state. Good progress is being made in defining
the helix organization in discoidal HDL particles where the ‘‘belt-model’’ is established. It remains
to be seen how the helices are arranged on spherical lipoprotein particles and what helix features
control the affinity for particles of different diameter. At this stage, it is difficult to predict how
changes of particular amino acids in an amphipathic a-helix affect the overall properties of the
helix. This information will be required to understand the effects of mutations on apolipoprotein
structure–function.

Acknowledgements

We are indebted to all our colleagues for their valuable contributions to the studies described
here from our laboratory. Our research reported here was supported by NIH Grants HL22633
and HL56083.

References

[1] Li WH, Tanimura M, Luo CC, Datta S, Chan L. The apolipoprotein multigene family: biosynthesis, structure,
structure–function relationships, and evolution. J Lipid Res 1988;29:245–71.
[2] Weisgraber KH. Apolipoprotein E: structure–function relationships. In: Schumaker VN, editor. Adv Protein
Chem. London: Academic Press; 1994. p. 249–302.
[3] Brouillette CG, Anantharamaiah GM. Structural models of human apolipoprotein A-I. Biochim Biophys Acta
1995;1256:103–29.
[4] Brouillette CG, Anantharamaiah GM, Engler JA, Borhani DW. Structural models of human apolipoprotein A-I:
a critical analysis and review. Biochim Biophys Acta 2001;1531:4–46.
[5] Frank PG, Marcel YL. Apolipoprotein A-I: structure–function relationships. J Lipid Res 2000;41:853–72.
[6] Narayanaswami V, Ryan RO. Molecular basis of exchangeable apolipoprotein function. Biochim Biophys Acta
2000;1483:15–36.
[7] Bolanos-Garcia VM, Miguel RN. On the structure and function of apolipoproteins: more than a family of lipid-
binding proteins. Prog Biophys Mol Biol 2003;83:47–68.
[8] Segrest JP, Jones MK, De Loof H, Brouillette CG, Venkatachalapathi YV, Anantharamaiah GM. The
amphipathic helix in the exchangeable apolipoproteins: a review of secondary structure and function. J Lipid Res
1992;33:141–66.
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 373

[9] Nolte RT, Atkinson D. Conformational analysis of apolipoprotein A-I and E-3 based on primary sequence and
circular dichroism. Biophys J 1992;63:1221–39.
[10] Segrest JP, De Loof H, Dohlman JG, Brouillette CG, Anantharamaiah GM. Amphipathic helix motif: classes and
properties. Proteins 1990;8:103–17.
[11] Segrest JP, Garber DW, Brouillette CG, Harvey SC, Anantharamaiah GM. The amphipathic alpha helix: a
multifunctional structural motif in plasma apolipoproteins. Adv Protein Chem 1994;45:303–69.
[12] Mishra VK, Palgunachari MN. Interaction of model class A1, class A2, and class Y amphipathic helical peptides
with membranes. Biochemistry 1996;35:11210–20.
[13] Collet X, Marcel YL, Tremblay N, Lazure C, Milne RW, Perret B, et al. Evolution of mammalian apolipoprotein
A-I and conservation of antigenicity: correlation with primary and secondary structure. J Lipid Res 1997;38:634–
44.
[14] Palgunachari MN, Mishra VK, Lund-Katz S, Phillips MC, Adeyeye SO, Alluri S, et al. Only the two end helixes
of eight tandem amphipathic helical domains of human apo A-I have significant lipid affinity. Implications for
HDL assembly. Arterioscler Thromb Vasc Biol 1996;16:328–38.
[15] Kyte J, Doolittle RF. A simple method for displaying the hydropathic character of a protein. J Mol Biol
1982;157:105–32.
[16] Gorshkova IN, Liadaki K, Gursky O, Atkinson D, Zannis VI. Probing the lipid-free structure and stability of
apolipoprotein A-I by mutation. Biochemistry 2000;39:15910–9.
[17] Oda MN, Forte TM, Ryan RO, Voss JC. The C-terminal domain of apolipoprotein A-I contains a lipid-sensitive
conformational trigger. Nat Struct Biol 2003;10:455–60.
[18] Mishra VK, Palgunachari MN, Datta G, Phillips MC, Lund-Katz S, Adeyeye SO, et al. Studies of synthetic
peptides of human apolipoprotein A-I containing tandem amphipathic a-helixes. Biochemistry 1998;37:10313–
24.
[19] Fang Y, Gursky O, Atkinson D. Lipid-binding studies of human apolipoprotein A-I and its terminally truncated
mutants. Biochemistry 2003;42:13260–8.
[20] Mahley RW, Innerarity TL, Rall Jr SC, Weisgraber KH. Plasma lipoproteins: apolipoprotein structure and
function. J Lipid Res 1984;25:1277–94.
[21] Weisgraber KH, Rall Jr SC, Mahley RW. Human E apoprotein heterogeneity. Cysteine-arginine interchanges in
the amino acid sequence of the apo-E isoforms. J Biol Chem 1981;256:9077–83.
[22] Wetterau JR, Aggerbeck LP, Rall Jr SC, Weisgraber KH. Human apolipoprotein E3 in aqueous solution. I.
Evidence for two structural domains. J Biol Chem 1988;263:6240–8.
[23] Aggerbeck LP, Wetterau JR, Weisgraber KH, Wu CS, Lindgren FT. Human apolipoprotein E3 in aqueous
solution. II. Properties of the amino- and carboxyl-terminal domains. J Biol Chem 1988;263:6249–58.
[24] Wilson C, Wardell MR, Weisgraber KH, Mahley RW, Agard DA. Three-dimensional structure of the LDL
receptor-binding domain of human apolipoprotein E. Science 1991;252:1817–22.
[25] Breiter DR, Kanost MR, Benning MM, Wesenberg G, Law JH, Wells MA, et al. Molecular structure of an
apolipoprotein determined at 2.5-A  resolution. Biochemistry 1991;30:603–8.
[26] Wilson C, Mau T, Weisgraber KH, Wardell MR, Mahley RW, Agard DA. Salt bridge relay triggers defective
LDL receptor binding by a mutant apolipoprotein. Structure 1994;2:713–8.
[27] Dong LM, Parkin S, Trakhanov SD, Rupp B, Simmons T, Arnold KS, et al. Novel mechanism for defective
receptor binding of apolipoprotein E2 in type III hyperlipoproteinemia. Nat Struct Biol 1996;3:718–22.
[28] Dong LM, Wilson C, Wardell MR, Simmons T, Mahley RW, Weisgraber KH, et al. Human apolipoprotein E.
Role of arginine 61 in mediating the lipoprotein preferences of the E3 and E4 isoforms. J Biol Chem
1994;269:22358–65.
[29] Morrow JA, Segall ML, Lund-Katz S, Phillips MC, Knapp M, Rupp B, et al. Differences in stability among the
human apolipoprotein E isoforms determined by theamino-terminal domain. Biochemistry 2000;39:11657–66.
[30] Acharya P, Segall ML, Zaiou M, Morrow J, Weisgraber KH, Phillips MC, et al. Comparison of the stabilities and
unfolding pathways of human apolipoprotein Eisoforms by differential scanning calorimetry and circular
dichroism. Biochim Biophys Acta 2002;1584:9–19.
[31] Saito H, Dhanasekaran P, Baldwin F, Weisgraber KH, Phillips MC, Lund-Katz S. Effects of polymorphism on
the lipid interaction of human apolipoprotein e. J Biol Chem 2003;278:40723–9.
374 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

[32] Choy N, Raussens V, Narayanaswami V. Inter-molecular coiled-coil formation in human apolipoprotein E C-


terminal domain. J Mol Biol 2003;334:527–39.
[33] Roberts LM, Ray MJ, Shih TW, Hayden E, Reader MM, Brouillette CG. Structural analysis of apolipoprotein
A-I: limited proteolysis of methionine-reduced and -oxidized lipid-free and lipid-bound human apo A-I.
Biochemistry 1997;36:7615–24.
[34] Davidson WS, Hazlett T, Mantulin WW, Jonas A. The role of apolipoprotein AI domains in lipid binding. Proc
Natl Acad Sci USA 1996;93:13605–10.
[35] Saito H, Dhanasekaran P, Nguyen D, Holvoet P, Lund-Katz S, Phillips MC. Domain structure and lipid
interaction in human apolipoproteins A-I and E, a general model. J Biol Chem 2003;278:23227–32.
[36] Davidson WS, Arnvig-McGuire K, Kennedy A, Kosman J, Hazlett TL, Jonas A. Structural organization of the
N-terminal domain of apolipoprotein A-I: studies of tryptophan mutants. Biochemistry 1999;38:14387–95.
[37] Kiss RS, Kay CM, Ryan RO. Amphipathic a-helix bundle organization of lipid-free chicken apolipoprotein A-I.
Biochemistry 1999;38:4327–34.
[38] Borhani DW, Rogers DP, Engler JA, Brouillette CG. Crystal structure of truncated human apolipoprotein A-I
suggests a lipid-bound conformation. Proc Natl Acad Sci USA 1997;94:12291–6.
[39] Reijngoud DJ, Phillips MC. Mechanism of dissociation of human apolipoprotein A-I from complexes with
dimyristoylphosphatidylcholine as studied by guanidine hydrochloride denaturation. Biochemistry 1982;21:2969–
76.
[40] Gursky O, Atkinson D. Thermal unfolding of human high-density apolipoprotein A-1: implications for a lipid-
free molten globular state. Proc Natl Acad Sci USA 1996;93:2991–5.
[41] Wang J, Sykes BD, Ryan RO. Structural basis for the conformational adaptability of apolipophorin III, a helix-
bundle exchangeable apolipoprotein. Proc Natl Acad Sci USA 2002;99:1188–93.
[42] Tricerri MA, Behling Agree AK, Sanchez SA, Jonas A. Characterization of apolipoprotein A-I structure using a
cysteine-specific fluorescence probe. Biochemistry 2000;39:14682–91.
[43] Behling Agree AK, Tricerri MA, Arnvig McGuire K, Tian SM, Jonas A. Folding and stability of the C-terminal
half of apolipoprotein A-I examined with a Cys-specific fluorescence probe. Biochim Biophys Acta
2002;1594:286–96.
[44] Fang Y, Gursky O, Atkinson D. Structural studies of N- and C-terminally truncated human apolipoprotein A-I.
Biochemistry 2003;42:6881–90.
[45] Weisgraber KH. Apolipoprotein E distribution among human plasma lipoproteins: role of the cysteine-arginine
interchange at residue 112. J Lipid Res 1990;31:1503–11.
[46] Dong LM, Weisgraber KH. Human apolipoprotein E4 domain interaction. Arginine 61 and glutamic acid 255
interact to direct the preference for very low density lipoproteins. J Biol Chem 1996;271:19053–7.
[47] Mahley RW, Huang Y, Rall Jr SC. Pathogenesis of type III hyperlipoproteinemia (dysbetalipoproteinemia).
Questions, quandaries, and paradoxes. J Lipid Res 1999;40:1933–49.
[48] Narayanaswami V, Szeto SS, Ryan RO. Lipid association-induced N- and C-terminal domain reorganization in
human apolipoprotein E3. J Biol Chem 2001;276:37853–60.
[49] Vitello LB, Scanu AM. Studies on human serum high density lipoproteins. Self-association of apolipoprotein A-I
in aqueous solutions. J Biol Chem 1976;251:1131–6.
[50] Yokoyama S, Kawai Y, Tajima S, Yamamoto A. Behavior of human apolipoprotein E in aqueous solutions and
at interfaces. J Biol Chem 1985;260:16375–82.
[51] Westerlund JA, Weisgraber KH. Discrete carboxyl-terminal segments of apolipoprotein E mediate lipoprotein
association and protein oligomerization. J Biol Chem 1993;268:15745–50.
[52] Perugini MA, Schuck P, Howlett GJ. Self-association of human apolipoprotein E3 and E4 in the presence and
absence of phospholipid. J Biol Chem 2000;275:36758–65.
[53] Ji Y, Jonas A. Properties of an N-terminal proteolytic fragment of apolipoprotein AI in solution and in
reconstituted high density lipoproteins. J Biol Chem 1995;270:11290–7.
[54] Holvoet P, Zhao Z, Vanloo B, Vos R, Deridder E, Dhoest A, et al. Phospholipid binding and lecithin-cholesterol
acyltransferase activation properties of apolipoprotein A-I mutants. Biochemistry 1995;34:13334–42.
[55] Weisgraber KH, Lund-Katz S, Phillips MC. Apolipoprotein E: structure–function correlations. In: Mill NE, Tall
AR, editors. High density lipoproteins and atherosclerosis III. Amsterdam: Elsevier; 1992. p. 175–81.
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 375

[56] Raussens V, Fisher CA, Goormaghtigh E, Ryan RO, Ruysschaert JM. The low density lipoprotein receptor active
conformation of apolipoprotein E. Helix organization in n-terminal domain-phospholipid disc particles. J Biol
Chem 1998;273:25825–30.
[57] Fisher CA, Ryan RO. Lipid binding-induced conformational changes in the N-terminal domain of human
apolipoprotein E. J Lipid Res 1999;40:93–9.
[58] Lu B, Morrow JA, Weisgraber KH. Conformational reorganization of the four-helix bundle of human
apolipoprotein E in binding to phospholipid. J Biol Chem 2000;275:20775–81.
[59] Segall ML, Dhanasekaran P, Baldwin F, Anantharamaiah GM, Weisgraber KH, Phillips MC, et al. Influence of
apoE domain structure and polymorphism on the kinetics of phospholipid vesicle solubilization. J Lipid Res
2002;43:1688–700.
[60] Weers PM, Narayanaswami V, Choy N, Luty R, Hicks L, Kay CM, et al. Lipid binding ability of human
apolipoprotein E N-terminal domain isoforms: correlation with protein stability. Biophys Chem 2003;100:481–92.
[61] Morrow JA, Hatters DM, Lu B, Hochtl P, Oberg KA, Rupp B, et al. Apolipoprotein E4 forms a molten globule.
A potential basis for its association with disease. J Biol Chem 2002;277:50380–5.
[62] Pownall HJ, Massey JB, Kusserow SK, Gotto Jr AM. Kinetics of lipid–protein interactions: interaction of
apolipoprotein A-I from human plasma high density lipoproteins with phosphatidylcholines. Biochemistry
1978;17:1183–8.
[63] Rogers DP, Roberts LM, Lebowitz J, Datta G, Anantharamaiah GM, Engler JA, et al. The lipid-free structure of
apolipoprotein A-I: effects of amino-terminal deletions. Biochemistry 1998;37:11714–25.
[64] Gorshkova IN, Liu T, Zannis VI, Atkinson D. Lipid-free structure and stability of apolipoprotein A-I: probing
the central region by mutation. Biochemistry 2002;41:10529–39.
[65] Okon M, Frank PG, Marcel YL, Cushley RJ. Heteronuclear NMR studies of human serum apolipoprotein A-I.
Part I. Secondary structure in lipid-mimetic solution. FEBS Lett 2002;517:139–43.
[66] Saito H, Dhanasekaran P, Nguyen D, Deridder E, Holvoet P, Lund-Katz S, et al. a-Helix formation is required
for high affinity binding of human apolipoprotein A-I to lipids. J Biol Chem 2004;279:20974–81.
[67] Massey JB, Gotto Jr AM, Pownall HJ. Contribution of a helix formation in human plasma apolipoproteins to
their enthalpy of association with phospholipids. J 1979;254:9559–61.
[68] Tricerri MA, Sanchez SA, Arnulphi C, Durbin DM, Gratton E, Jonas A. Interaction of apolipoprotein
A-I in three different conformations with palmitoyl oleoyl phosphatidylcholine vesicles. J Lipid Res 2002;43:187–
97.
[69] Derksen A, Gantz D, Small DM. Calorimetry of apolipoprotein-A1 binding to phosphatidylcholine–triolein–
cholesterol emulsions. Biophys J 1996;70:330–8.
[70] Saito H, Dhanasekaran P, Baldwin F, Weisgraber KH, Lund-Katz S, Phillips MC. Lipid binding-induced
conformational change in human apolipoprotein E. Evidence for two lipid-bound states on spherical particles. J
Biol Chem 2001;276:40949–54.
[71] Gazzara JA, Phillips MC, Lund-Katz S, Palgunachari MN, Segrest JP, Anantharamaiah GM, et al. Interaction of
class A amphipathic helical peptides with phospholipid unilamellar vesicles. J Lipid Res 1997;38:2134–46.
[72] Lund-Katz S, Liu L, Thuahnai ST, Phillips MC. High density lipoprotein structure. Front Biosci 2003;8:d1044–
1054.
[73] Phillips JC, Wriggers W, Li Z, Jonas A, Schulten K. Predicting the structure of apolipoprotein A-I in
reconstituted high-density lipoprotein disks. Biophys J 1997;73:2337–46.
[74] Klon AE, Segrest JP, Harvey SC. Comparative models for human apolipoprotein A-I bound to lipid in discoidal
high-density lipoprotein particles. Biochemistry 2002;41:10895–905.
[75] Koppaka V, Silvestro L, Engler JA, Brouillette CG, Axelsen PH. The structure of human lipoprotein A-I.
Evidence for the ‘‘belt model. J Biol Chem 1999;274:14541–4.
[76] Maiorano JN, Davidson WS. The orientation of helix 4 in apolipoprotein A-I-containing reconstituted high
density lipoproteins. J Biol Chem 2000;275:17374–80.
[77] Panagotopulos SE, Horace EM, Maiorano JN, Davidson WS. Apolipoprotein A-I adopts a belt-like orientation
in reconstituted high density lipoproteins. J Biol Chem 2001;276:42965–70.
[78] Li H, Lyles DS, Thomas MJ, Pan W, Sorci-Thomas MG. Structural determination of lipid-bound ApoA-I using
fluorescence resonance energy transfer. J Biol Chem 2000;275:37048–54.
376 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

[79] Segrest JP, Jones MK, Klon AE, Sheldahl CJ, Hellinger M, De Loof H, et al. A detailed molecular belt model for
apolipoprotein A-I in discoidal high density lipoprotein. J Biol Chem 1999;274:31755–8.
[80] Tricerri MA, Behling Agree AK, Sanchez SA, Bronski J, Jonas A. Arrangement of apolipoprotein A-I in
reconstituted high-density lipoprotein disks: an alternative model based on fluorescence resonance energy transfer
experiments. Biochemistry 2001;40:5065–74.
[81] Davidson WS, Hilliard GM. The spatial organization of apolipoprotein A-I on the edge of discoidal high density
lipoprotein particles: a mass specrometry study. J Biol Chem 2003;278:27199–207.
[82] De Pauw M, Vanloo B, Weisgraber K, Rosseneu M. Comparison of lipid-binding and lecithin:cholesterol
acyltransferase activation of the amino- and carboxyl-terminal domains of human apolipoprotein E3.
Biochemistry 1995;34:10953–66.
[83] Fisher CA, Narayanaswami V, Ryan RO. The lipid-associated conformation of the low density lipoprotein
receptor binding domain of human apolipoprotein E. J Biol Chem 2000;275:33601–6.
[84] Narayanaswami V, Maiorano JN, Dhanasekaran P, Ryan RO, Phillips MC, Lund-Katz S, et al. Helix orientation
of the functional domains in apolipoprotein E in discoidal high density lipoprotein particles. J Biol Chem
2004;279:14273–9.
[85] Curtiss LK, Bonnet DJ, Rye KA. The conformation of apolipoprotein A-I in high-density lipoproteins is
influenced by core lipid composition and particle size: a surface plasmon resonance study. Biochemistry
2000;39:5712–21.
[86] Sparks DL, Phillips MC, Lund-Katz S. The conformation of apolipoprotein A-I in discoidal and spherical
recombinant high density lipoprotein particles. 13 C NMR studies of lysine ionization behavior. J Biol Chem
1992;267:25830–8.
[87] Wang L, Atkinson D, Small DM. Interfacial properties of an amphipathic a-helix consensus peptide of
exchangeable apolipoproteins at air/water and oil/water interfaces. J Biol Chem 2003;278:37480–91.
[88] Corsico B, Toledo JD, Garda HA. Evidence for a central apolipoprotein A-I domain loosely bound to lipids in
discoidal lipoproteins that is capable of penetrating the bilayer of phospholipid vesicles. J Biol Chem
2001;276:16978–85.
[89] Li HH, Lyles DS, Pan W, Alexander E, Thomas MJ, Sorci-Thomas MG. ApoA-I structure on discs and spheres.
Variable helix registry and conformational states. J Biol Chem 2002;277:39093–101.
[90] Blanco-Vaca F, Escola-Gil JC, Martin-Campos JM, Julve J. Role of apoA-II in lipid metabolism and
atherosclerosis: advances in the study of an enigmatic protein. J Lipid Res 2001;42:1727–39.
[91] Brewer HB, Jr SE, Lux R, Ronan KM. Amino acid sequence of human apoLp-Gln-II (apoA-II), an
apolipoprotein isolated from the high-density lipoprotein complex. Proc Natl Acad Sci USA 1972;69:1304–
8.
[92] Kumar MS, Carson M, Hussain MM, Murthy HM. Structures of apolipoprotein A-II and a lipid-surrogate
complex provide insights into apolipoprotein-lipid interactions. Biochemistry 2002;41:11681–91.
[93] Gursky O, Atkinson D. High- and low-temperature unfolding of human high-density apolipoprotein A-2. Protein
Sci 1996;5:1874–82.
[94] Ibdah JA, Krebs KE, Phillips MC. The surface properties of apolipoproteins A-I and A-II at the lipid/water
interface. Biochim Biophys Acta 1989;1004:300–8.
[95] Lagocki PA, Scanu AM. 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
1980;255:3701–6.
[96] van Tornout P, Caster H, Lievens MJ, Rosseneu M, Assmann G. In vitro interaction of human HDL with human
apolipoprotein A-II. Synthesis of apolipoprotein A-II-rich HDL. Biochim Biophys Acta 1981;663:630–6.
[97] Rye KA. Interaction of apolipoprotein A-II with recombinant HDL containing egg phosphatidylcholine,
unesterified cholesterol and apolipoprotein A-I. Biochim Biophys Acta 1990;1042:227–36.
[98] Rye KA, Barter PJ. The influence of apolipoproteins on the structure and function of spheroidal, reconstituted
high density lipoproteins. J Biol Chem 1994;269:10298–303.
[99] Pownall HJ, Hickson D, Gotto Jr AM. Thermodynamics of lipid-protein association. The free energy of
association of lecithin with reduced and carboxymethylated apolipoprotein A-II from human plasma high density
lipoprotein. J Biol Chem 1981;256:9849–54.
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 377

[100] Lund-Katz S, Murley YM, Yon E, Gillotte KL, Davidson WS. Comparison of the structural and functional
effects of monomeric and dimeric human apolipoprotein A-II in high density lipoprotein particles. Lipids
1996;31:1107–13.
[101] Swaney JB, Reese H, Eder HA. Polypeptide composition of rat high density lipoprotein: characterization by SDS-
gel electrophoresis. Biochem Biophys Res Commun 1974;59:513–9.
[102] Fujimoto K, Fukagawa K, Sakata T, Tso P. Suppression of food intake by apolipoprotein A-IV is mediated
through the central nervous system in rats. J Clin Invest 1993;91:1830–3.
[103] Okumura T, Fukagawa K, Tso P, Taylor IL, Pappas TN. Apolipoprotein A-IV acts in the brain to inhibit gastric
emptying in the rat. Am J Physiol 1996;270:G49–53.
[104] Lefevre M, Roheim PS. Metabolism of apolipoprotein A-IV. J Lipid Res 1984;25:1603–10.
[105] Ostos MA, Conconi M, Vergnes L, Baroukh N, Ribalta J, Girona J, et al. Antioxidative and antiatherosclerotic
effects of human apolipoprotein A-IV in apolipoprotein E-deficient mice. Arterioscler Thromb Vasc Biol
2001;21:1023–8.
[106] Remaley AT, Stonik JA, Demosky SJ, Neufeld EB, Bocharov AV, Vishnyakova TG, et al. Apolipoprotein
specificity for lipid efflux by the human ABCAI transporter. Biochem Biophys Res Commun 2001;280:818–23.
[107] Steinmetz A, Utermann G. Activation of lecithin: cholesterol acyltransferase by human apolipoprotein A-IV. J
Biol Chem 1985;260:2258–64.
[108] Boguski MS, Elshourbagy N, Taylor JM, Gordon JI. Comparative analysis of repeated sequences in rat
apolipoproteins A-I, A-IV, and E. Proc Natl Acad Sci USA 1985;82:992–6.
[109] Weinberg RB, Spector MS. Structural properties and lipid binding of human apolipoprotein A-IV. J Biol Chem
1985;260:4914–21.
[110] Pearson K, Saito H, Woods SC, Lund-Katz S, Tso P, Phillips MC, et al. The structure of human apolipoprotein
A-IV: a distinct domain architecture with potential functional implications; submitted.
[111] Pennacchio LA, Olivier M, Hubacek JA, Cohen JC, Cox DR, Fruchart JC, et al. An apolipoprotein influencing
triglycerides in humans and mice revealed by comparative sequencing. Science 2001;294:169–73.
[112] Talmud PJ, Hawe E, Martin S, Olivier M, Miller GJ, Rubin EM, et al. Relative contribution of variation within
the APOC3/A4/A5 gene cluster in determining plasma triglycerides. Hum Mol Genet 2002;11:3039–46.
[113] Nabika T, Nasreen S, Kobayashi S, Masuda J. The genetic effect of the apoprotein AV gene on the serum
triglyceride level in Japanese. Atherosclerosis 2002;165:201–4.
[114] Weinberg RB, Cook VR, Beckstead JA, Martin DD, Gallagher JW, Shelness GS, et al. Structure and interfacial
properties of human apolipoprotein A-V. J Biol Chem 2003;278:34438–44.
[115] Beckstead JA, Oda MN, Martin DD, Forte TM, Bielicki JK, Berger T, et al. Structure–function studies of human
apolipoprotein A-V: a regulator of plasma lipid homeostasis. Biochemistry 2003;42:9416–23.
[116] Dumon MF, Clerc M. Preliminary report on a case of apolipoproteins CI and CII deficiency. Clin Chim Acta
1986;157:239–48.
[117] Conde-Knape K, Bensadoun A, Sobel JH, Cohn JS, Shachter NS. Overexpression of apoC-I in apoE-null mice:
severe hypertriglyceridemia due to inhibition of hepatic lipase. J Lipid Res 2002;43:2136–45.
[118] Kinnunen PK, Ehnolm C. Effect of serum and C-apoproteins from very low density lipoproteins on human
postheparin plasma hepatic lipase. FEBS Lett 1976;65:354–7.
[119] Gautier T, Masson D, de Barros JP, Athias A, Gambert P, Aunis D, et al. Human apolipoprotein C-I accounts
for the ability of plasma high density lipoproteins to inhibit the cholesteryl ester transfer protein activity. J Biol
Chem 2000;275:37504–9.
[120] Kowal RC, Herz J, Weisgraber KH, Mahley RW, Brown MS, Goldstein JL. Opposing effects of apolipoproteins
E and C on lipoprotein binding to low density lipoprotein receptor-related protein. J Biol Chem 1990;265:10771–
9.
[121] Jackson RL, Morrisett JD, Sparrow JT, Segrest JP, Pownall HJ, Smith LC, et al. The interaction of
apolipoprotein-serine with phosphatidylcholine. J Biol Chem 1974;249:5314–20.
[122] Gursky O. Solution conformation of human apolipoprotein C-1 inferred from proline mutagenesis: far- and near-
UV CD study. Biochemistry 2001;40:12178–85.
[123] Rozek A, Sparrow JT, Weisgraber KH, Cushley RJ. Conformation of human apolipoprotein C-I in a lipid-
mimetic environment determined by CD and NMR spectroscopy. Biochemistry 1999;38:14475–84.
378 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

[124] Semenkovich CF, Luo CC, Nakanishi MK, Chen SH, Smith LC, Chan L. In vitro expression and site-specific
mutagenesis of the cloned human lipoprotein lipase gene. Potential N-linked glycosylation site asparagine 43 is
important for both enzyme activity and secretion. J Biol Chem 1990;265:5429–33.
[125] MacPhee CE, Hatters DM, Sawyer WH, Howlett GJ. Apolipoprotein C-II39-62 activates lipoprotein lipase by
direct lipid-independent binding. Biochemistry 2000;39:3433–40.
[126] Hatters DM, MacPhee CE, Lawrence LJ, Sawyer WH, Howlett GJ. Human apolipoprotein C-II forms twisted
amyloid ribbons and closed loops. Biochemistry 2000;39:8276–83.
[127] MacRaild CA, Hatters DM, Howlett GJ, Gooley PR. NMR structure of human apolipoprotein C-II in the
presence of sodium dodecyl sulfate. Biochemistry 2001;40:5414–21.
[128] Shachter NS. Apolipoproteins C-I and C-III as important modulators of lipoprotein metabolism. Curr Opin
Lipidol 2001;12:297–304.
[129] Lins L, Flore C, Chapelle L, Talmud PJ, Thomas A, Brasseur R. Lipid-interacting properties of the N-terminal
domain of human apolipoprotein C-III. Protein Eng 2002;15:513–20.
[130] Mahley RW, Huang Y. Apolipoprotein E: from atherosclerosis to Alzheimer’s disease and beyond. Curr Opin
Lipidol 1999;10:207–17.
[131] Lund-Katz S, Wehrli S, Zaiou M, Newhouse Y, Weisgraber KH, Phillips MC. Effects of polymorphism on the
microenvironment of the LDL receptor-binding region of human apoE. J Lipid Res 2001;42:894–901.
[132] Zaiou M, Arnold KS, Newhouse YM, Innerarity TL, Weisgraber KH, Segall ML, et al. Apolipoprotein E-low
density lipoprotein receptor interaction: Influences of basic residue and amphipathic a-helix organization in the
ligand. J Lipid Res 2000;41:1087–95.
[133] Lund-Katz S, Zaiou M, Wehrli S, Dhanasekaran P, Baldwin F, Weisgraber KH, et al. Effects of lipid interaction
on the lysine microenvironments in apolipoprotein E. J Biol Chem 2000;275:34459–64.
[134] Raussens V, Slupsky CM, Sykes BD, Ryan RO. Lipid-bound structure of an apolipoprotein E-derived peptide. J
Biol Chem 2003;278:25998–6006.
[135] Thuahnai ST, Lund-Katz S, Williams DL, Phillips MC. Scavenger receptor class B, type I-mediated uptake of
various lipids into cells. Influence of the nature of the donor particle interaction with the receptor. J Biol Chem
2001;276:43801–8.
[136] Thuahnai ST, Lund-Katz S, Anantharamaiah GM, Williams DL, Phillips MC. A quantitative analysis of
apolipoprotein binding to SR-BI: multiple binding sites for lipid-free and lipid-associated apolipoproteins. J Lipid
Res 2003;44:1132–42.
[137] Li X, Kan HY, Lavrentiadou S, Krieger M, Zannis V. Reconstituted discoidal ApoE-phospholipid particles are
ligands for the scavenger receptor BI. The amino-terminal 1–165 domain of ApoE suffices for receptor binding. J
Biol Chem 2002;277:21149–57.
[138] Gianturco SH, Gotto Jr AM, Hwang SL, Karlin JB, Lin AH, Prasad SC, et al. Apolipoprotein E mediates uptake
of Sf 100-400 hypertriglyceridemic very low density lipoproteins by the low density lipoprotein receptor pathway
in normal human fibroblasts. J Biol Chem 1983;258:4526–33.
[139] Herz J, Bock HH. Lipoprotein receptors in the nervous system. Annu Rev Biochem 2002;71:405–34.
[140] Havel RJ, Hamilton RL. Hepatic catabolism of remnant lipoproteins: where the action is. Arterioscler Thromb
Vasc Biol 2004;24:213–5.
[141] Burgess JW, Gould DR, Marcel YL. The HepG2 extracellular matrix contains separate heparinase- and lipid-
releasable pools of ApoE. Implications for hepatic lipoprotein metabolism. J Biol Chem 1998;273:5645–54.
[142] Saito H, Dhanasekaran P, Nguyen D, Baldwin F, Weisgraber KH, Wehrli S, et al. Characterization of the heparin
binding sites in human apolipoprotein E. J Biol Chem 2003;278:14782–7.
[143] Libeu CP, Lund-Katz S, Phillips MC, Wehrli S, Hernaiz MJ, Capila I, et al. New insights into heparin sulfate
proteoglycan-binding activity of apolipoprotein E. J Biol Chem 2001;276:39138–44.
[144] Zhu Y, Hui DY. Apolipoprotein E binding to low density lipoprotein receptor-related protein-1 inhibits cell
migration via activation of cAMP-dependent protein kinase A. J Biol Chem 2003;278:36257–63.
[145] Swertfeger DK, Hui DY. Apolipoprotein E: receptor versus heparan sulfate proteoglycan binding in its regulation
of smooth muscle cell migration and proliferation. J Biol Chem 2001;276:25043–8.
[146] Kothapalli D, Fuki I, Ali K, Stewart SA, Zhao L, Yahil R, et al. Antimitogenic effects of HDL and APOE
mediated by Cox-2-dependent IP activation. J Clin Invest 2004;113:609–18.
H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380 379

[147] Weisgraber KH, Mahley RW. Human apolipoprotein E: the Alzheimer’s disease connection. FASEB J
1996;10:1485–94.
[148] Pillot T, Goethals M, Najib J, Labeur C, Lins L, Chambaz J, et al. Beta-amyloid peptide interacts specifically with
the carboxy-terminal domain of human apolipoprotein E: relevance to Alzheimer’s disease. J Neurochem
1999;72:230–7.
[149] Raffai RL, Dong LM, Farese Jr RV, Weisgraber KH. Introduction of human apolipoprotein E4 ‘‘domain
interaction’’ into mouse apolipoprotein E. Proc Natl Acad Sci USA 2001;98:11587–91.
[150] Yancey PG, Bortnick AE, Kellner-Weibel G, de la Llera-Moya M, Phillips MC, Rothblat GH. Importance of
different pathways of cellular cholesterol efflux. Arterioscler Thromb Vasc Biol 2003;23:712–9.
[151] Oram JF. HDL apolipoproteins and ABCA1: partners in the removal of excess cellular cholesterol. Arterioscler
Thromb Vasc Biol 2003;23:720–7.
[152] Rye KA, Barter PJ. Formation and metabolism of prebeta-migrating, lipid-poor apolipoprotein A-I. Arterioscler
Thromb Vasc Biol 2004;24:421–8.
[153] Chroni A, Liu T, Fitzgerald ML, Freeman MW, Zannis VI. Cross-linking and lipid efflux properties of
ApoA-I mutants suggest direct association between ApoA-I helices and ABCA1. Biochemistry 2004;43:
2126–39.
[154] Gillotte-Taylor K, Nickel M, Johnson WJ, Francone OL, Holvoet P, Lund-Katz S, et al. Effects of enrichment of
fibroblasts with unesterified cholesterol on the efflux of cellular lipids to apolipoprotein A-I. J Biol Chem
2002;277:11811–20.
[155] Liu L, Bortnick AE, Nickel M, Dhanasekaran P, Subbaiah PV, Lund-Katz S, et al. Effects of apolipoprotein A-I
on ATP-binding cassette transporter A1- mediated efflux of macrophage phospholipid and cholesterol: formation
of nascent high density lipoprotein particles. J Biol Chem 2003;278:42976–84.
[156] Jonas A. Lecithin-cholesterol acyltransferase in the metabolism of high-density lipoproteins. Biochim Biophys
Acta 1991;1084:205–20.
[157] Jonas A. Lecithin cholesterol acyltransferase. Biochim Biophys Acta 2000;1529:245–56.
[158] Jin L, Shieh JJ, Grabbe E, Adimoolam S, Durbin D, Jonas A. Surface plasmon resonance biosensor studies of
human wild-type and mutant lecithin cholesterol acyltransferase interactions with lipoproteins. Biochemistry
1999;38:15659–65.
[159] Sorci-Thomas MG, Curtiss L, Parks JS, Thomas MJ, Kearns MW, Landrum M. The hydrophobic face
orientation of apolipoprotein A-I amphipathic helix domain 143–164 regulates lecithin:cholesterol acyltransferase
activation. J Biol Chem 1998;273:11776–82.
[160] Roosbeek S, Vanloo B, Duverger N, Caster H, Breyne J, De Beun I, et al. Three arginine residues in
apolipoprotein A-I are critical for activation of lecithin:cholesterol acyltransferase. J Lipid Res 2001;42:31–
40.
[161] McManus DC, Scott BR, Frank PG, Franklin V, Schultz JR, Marcel YL. Distinct central amphipathic a-helices
in apolipoprotein A-I contribute to the in vivo maturation of high density lipoprotein by either activating lecithin-
cholesterol acyltransferase or binding lipids. J Biol Chem 2000;275:5043–51.
[162] Sorci-Thomas MG, Thomas M, Curtiss L, Landrum M. 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
2000;275:12156–63.
[163] Reschly EJ, Sorci-Thomas MG, Davidson WS, Meredith SC, Reardon CA, Getz GS. Apolipoprotein A-I a-
helices 7 and 8 modulate high density lipoprotein subclass distribution. J Biol Chem 2002;277:9645–54.
[164] Rigotti A, Miettinen HE, Krieger M. The role of the high-density lipoprotein receptor SR-BI in the lipid
metabolism of endocrine and other tissues. Endocr Rev 2003;24:357–87.
[165] Williams DL, Connelly MA, Temel RE, Swarnakar S, Phillips MC, de la Llera-Moya M, et al. Scavenger receptor
BI and cholesterol trafficking. Curr Opin Lipidol 1999;10:329–39.
[166] Williams DL, de La Llera-Moya M, Thuahnai ST, Lund-Katz S, Connelly MA, Azhar S, et al. Binding and cross-
linking studies show that scavenger receptor BI interacts with multiple sites in apolipoprotein A-I and identify the
class A amphipathic a-helix as a recognition motif. J Biol Chem 2000;275:18897–904.
[167] de Beer MC, Durbin DM, Cai L, Jonas A, de Beer FC, van der Westhuyzen DR. Apolipoprotein
A-I conformation markedly influences HDL interaction with scavenger receptor BI. J Lipid Res 2001;42:309–13.
380 H. Saito et al. / Progress in Lipid Research 43 (2004) 350–380

[168] Thuahnai ST, Lund-Katz S, Dhanasekaran P, De La Llera-Moya M, Connelly MA, Williams DL, et al. SR-BI-
mediated cholesteryl ester selective uptake and efflux of unesterified cholesterol: influence of HDL size and
structure. J Biol Chem 2004:12448–55.
[169] Gillotte KL, Davidson WS, Lund-Katz S, Rothblat GH, Phillips MC. Apolipoprotein A-I structural modification
and the functionality of reconstituted high density lipoprotein particles in cellular cholesterol efflux. J Biol Chem
1996;271:23792–8.
[170] Andreola A, Bellotti V, Giorgetti S, Mangione P, Obici L, Stoppini M, et al. Conformational switching and
fibrillogenesis in the amyloidogenic fragment of apolipoprotein A-I. J Biol Chem 2003;278:2444–51.

You might also like