Download View Full Page PDF

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Membrane potential wikipedia , lookup

Cell growth wikipedia , lookup

Cell culture wikipedia , lookup

Cell nucleus wikipedia , lookup

Cell cycle wikipedia , lookup

Extracellular matrix wikipedia , lookup

Cellular differentiation wikipedia , lookup

Cytosol wikipedia , lookup

Theories of general anaesthetic action wikipedia , lookup

Magnesium transporter wikipedia , lookup

Mitosis wikipedia , lookup

Cell encapsulation wikipedia , lookup

Ethanol-induced non-lamellar phases in phospholipids wikipedia , lookup

Thylakoid wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Lipid bilayer wikipedia , lookup

Amitosis wikipedia , lookup

SNARE (protein) wikipedia , lookup

Cytokinesis wikipedia , lookup

Signal transduction wikipedia , lookup

Model lipid bilayer wikipedia , lookup

JADE1 wikipedia , lookup

Lipid raft wikipedia , lookup

Lipid signaling wikipedia , lookup

Cell membrane wikipedia , lookup

List of types of proteins wikipedia , lookup

Endomembrane system wikipedia , lookup

Transcript
PHYSIOLOGICAL REVIEWS
Vol. 81, No. 4, October 2001
Printed in U.S.A.
The Organizing Potential of Sphingolipids
in Intracellular Membrane Transport
JOOST C. M. HOLTHUIS, THOMAS POMORSKI, RENÉ J. RAGGERS, HEIN SPRONG,
AND GERRIT VAN MEER
Department of Cell Biology and Histology, Academic Medical Center, University of Amsterdam, Amsterdam;
and Department of Cell Biology, Utrecht University School of Medicine, Utrecht, The Netherlands
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
I. Introduction
II. Why Are Sphingolipids Essential for Eukaryotic Life?
A. Sphingolipid requirements at the cellular level
B. Sphingolipid requirements at the organismal level
C. Do sphingolipids exert a vital signaling function?
D. Sphingolipids and the spatial organization of cells
III. Sphingolipid Structure and Biophysical Properties
A. A concise inventory of sphingolipid structure in distinct eukaryotic life forms
B. Biophysical differences between sphingolipids and glycerolipids
IV. Sphingolipids and the Lateral Organization of Biological Membranes
A. Ordered sphingolipid domains in model membranes
B. Evidence for the existence of ordered sphingolipid domains in cellular membranes
V. Sphingolipid Assembly and Transport
A. Enzymes of sphingolipid metabolism and their topology
B. Subcellular distribution and topology of sphingolipids
C. Sphingolipid transport and sorting
VI. Sphingolipids and the Creation of Selectivity in Intracellular Membrane Transport
A. The Golgi: a central sorting device on the exocytic and endocytic pathways
B. A maturing view on Golgi organization
C. Sphingolipids: integral parts of the Golgi-based sorting machinery?
VII. Concluding Remarks
A. Sphingolipids and Golgi maturation
B. Domains of saturated lipids on the cytosolic surface
C. Sphingolipid domains as a general theme in cellular membrane traffic
1690
1691
1691
1693
1694
1694
1695
1695
1696
1696
1696
1697
1701
1701
1703
1704
1706
1706
1707
1709
1712
1712
1713
1713
Holthuis, Joost C. M., Thomas Pomorski, René J. Raggers, Hein Sprong, and Gerrit van Meer. The
Organizing Potential of Sphingolipids in Intracellular Membrane Transport. Physiol Rev 81: 1689 –1723, 2001.—
Eukaryotes are characterized by endomembranes that are connected by vesicular transport along secretory and
endocytic pathways. The compositional differences between the various cellular membranes are maintained by
sorting events, and it has long been believed that sorting is based solely on protein-protein interactions. However,
the central sorting station along the secretory pathway is the Golgi apparatus, and this is the site of synthesis of the
sphingolipids. Sphingolipids are essential for eukaryotic life, and this review ascribes the sorting power of the Golgi
to its capability to act as a distillation apparatus for sphingolipids and cholesterol. As Golgi cisternae mature,
ongoing sphingolipid synthesis attracts endoplasmic reticulum-derived cholesterol and drives a fluid-fluid lipid phase
separation that segregates sphingolipids and sterols from unsaturated glycerolipids into lateral domains. While
sphingolipid domains move forward, unsaturated glycerolipids are retrieved by recycling vesicles budding from the
sphingolipid-poor environment. We hypothesize that by this mechanism, the composition of the sphingolipid
domains, and the surrounding membrane changes along the cis-trans axis. At the same time the membrane thickens.
These features are recognized by a number of membrane proteins that as a consequence of partitioning between
domain and environment follow the domains but can enter recycling vesicles at any stage of the pathway. The
interplay between protein- and lipid-mediated sorting is discussed.
www.prv.org
0031-9333/01 $15.00 Copyright © 2001 the American Physiological Society
1689
1690
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
I. INTRODUCTION
Sphingolipids are typically found in eukaryotic cells
where they comprise a small but vital fraction (10 –20%)
of the membrane lipids. Their lipidic part consists of a
sphingoid base, a straight-chain amino alcohol of 18 –20
carbon atoms, which normally carries a long saturated
fatty acid amide bonded to the amino group at the C2
position (Fig. 1). Based on the type of headgroup attached
to the C1, sphingolipids are classified as phosphosphingolipids or glycosphingolipids. The phosphosphingolipids
sphingomyelin (SM) in animals and inositol phosphoceramide (IPC) in plants and fungi carry the polar headgroups phosphocholine and phosphoinositol, just like the
major glycerolipids phosphatidylcholine (PC) and phosphatidylinositol (PI). In glycosphingolipids, the headgroup can contain a variety of monosaccharides linked by
various types of glycosidic bonds. Since the discovery of
the sphingolipids more than 100 years ago by Thudichum
(376), their special lipid backbone and their bewildering
structural heterogeneity have fascinated biologists, biochemists, and biophysicists alike.
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
FIG. 1. Common lipids in fungi and various animals. a, Sphingomyelin (SM); b, glucosylceramide (GlcCer), and a
derived glycosphingolipid, e.g., GM1; c, glycerolipids; d, cholesterol; e, SM; f, glycosphingolipid of the arthro series; g, see
b; h, sterol; i and j, see f; k, see b; l, sterol; m, M(IP)2C; n, see c; o, ergosterol. Fat print indicates the sphingoid bases in
the sphingolipids a, b, e, f, i, j, and m; glycerol in the glycerolipids c, g, k, and n; and the sterols d, h, l, and o. A, head
group: choline, ethanolamine, serine, or inositol; C, choline; E, ethanolamine; G, glucose; Ino, inositol; M, mannose; P,
phosphate (note that insects and nematodes are sterol auxotrophs).
Physiol Rev • VOL
81 • OCTOBER 2001 •
www.prv.org
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
1
It should be noted that the name raft implies that this would be
a small domain floating in a sea of other lipids. However, in the many
cases where the relative sizes of the sphingolipid-rich and glycerophospholipid-rich environments are not known the term domain would be
less suggestive.
Physiol Rev • VOL
synthesize a novel set of glycerolipids whose structural
appearence and predicted biophysical properties closely
mimic those of the sphingolipids (74, 193, 281). As discussed in section II, this would indicate that some structural function, rather than a signaling one, accounts for
the sphingolipid requirement in cell growth. A comparison of sphingolipid structures from evolutionary distinct
organisms (see sect. III) indicates that, despite considerable chemical differences, the biophysical properties held
responsible for microdomain formation have been well
preserved. Indeed, counterparts of the microdomains analyzed in mammalian cell membranes appear to exist in
flies, worms, and even in yeast (see sect. IV).
A further striking aspect of sphingolipids is that their
assembly in eukaryotic cells is spatially separated from that
of glycerolipids and sterols. Whereas the latter are produced
in the endoplasmic reticulum (ER), sphingolipid synthesis
occurs primarily in the Golgi complex (see sect. V). The
Golgi has been well established as a polarized sorting and
processing station that is situated right at the intersection of
two major circuits of intracellular membrane trafficking.
One circuit interconnects the cis-side of the Golgi with the
ER; the other one the trans-side with the plasma membrane.
The ER and plasma membrane are engaged in fundamentally distinct cellular processes, as reflected by the dramatic
differences found in the molecular composition and biophysical properties of their bilayers. In section VI we consider the possibility that sphingolipids evolved as an integral
and essential part of the Golgi machinery responsible for
establishing the compositional and functional differences
between the ER and the plasma membrane.
II. WHY ARE SPHINGOLIPIDS ESSENTIAL
FOR EUKARYOTIC LIFE?
A. Sphingolipid Requirements at the Cellular Level
The early steps in sphingolipid synthesis, up to the
formation of sphingoid bases and ceramides, are essentially the same in all eukaryotes. It is in the subsequent
reactions, when ceramide is converted to complex sphingolipids, that major species- and cell type-specific differences become apparent. Yeast, for example, is equipped
with a single biosynthetic pathway that serves to convert
ceramide into only a handful of inositol-containing phosphosphingolipids. Animals, on the other hand, evolved
three independent pathways that can operate simultaneously to produce both phosphosphingolipids and hundreds of different types of glycosphingolipids.2 The divergence in headgroup structures found in glycosphingolipids
2
Glycosphingolipid designation is according to the recommendations of the IUPAC-IUB Joint Commission on Biochemical Nomenclature (53), using the Svennerholm abbreviations for gangliosides and
using sulfatide to indicate HSO3 -3 Gal ␤1–1 Cer.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
The major mission of the sphingolipid field is to
understand the specific functions of these lipids in eukaryotic organisms and to evaluate their significance both
for the functioning of individual cells and the organism as
a whole. This is an arduous task, not in the least because,
unlike glycerolipids and sterols, sphingolipids display a
striking structural variation between distinct organisms,
and even between different tissues and cell types within
one organism. Moreover, sphingolipids are no longer
viewed as primarily inert structural components of cellular membranes, but also increasingly recognized as diverse and dynamic regulators of a multitude of cellular
processes. An important development in this respect has
been the realization that sphingoid bases, ceramides, and
other intermediates of sphingolipid metabolism act as
signaling molecules in mediating cell cycle control, stress
responses, and apoptosis (75, 127, 356). In addition, considerable attention has been drawn to the concept that
sphingolipids drive the lateral differentiation of cellular
membranes into a mosaic of areas with unique molecular
compositions. In essence, it has been postulated that a
differential miscibility of sphingolipids, glycerolipids, and
sterols triggers the formation of lateral lipid assemblies,
termed microdomains or rafts, that acquire specific functions by concentrating or excluding specific membrane
proteins (341).1 Rafts are now believed to serve as platforms for various cellular events including polarized protein sorting, signal transduction, and cell adhesion (35,
118, 178, 305, 340).
Sphingolipids are essential to sustain eukaryotic life.
Whereas glycosphingolipids are indispensable for the development of complex multicellular organisms, phosphosphingolipids fulfill a vital function at a more fundamental
level, namely, in the growth and survival of individual
cells. This requirement for phosphosphingolipids appears
to be a conserved feature of eukaryotic cells, as it is found
both in animal cells and in yeast. Hence, it is not unlikely
that the different sphingolipids synthesized in the various
eukaryotic cell types serve a common function. If so, one
would expect sphingolipids to share essential features
that determine their interactions with other cellular components. If there is such a unifying principle, what could
it be?
This review serves to highlight some remarkable aspects of sphingolipids that we believe represent important guidelines for unraveling their vital function in eukaryotic cells. One of these is the striking observation that
in yeast the requirement for sphingolipids can be bypassed by a suppressor mutation that enables yeast to
1691
1692
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
accompanies functional differences between the various
cell types of an animal. Whereas the synthesis of at least
some glycosphingolipid species appears essential for the
viability of animal organisms (see sect. IIB), as a lipid
class they are dispensable for the growth and survival of
animal cells in culture. This is not the case for phosphosphingolipids. Phosphosphingolipid synthesis is required
for cell growth and survival, in yeast as well as in mammals (the only organisms in which this has been studied
so far). Yeast and Chinese hamster ovary cell mutants
lacking the first committed enzyme in sphingolipid biosynthesis, serine palmitoyl-transferase (Fig. 2), die in the
absence of externally added sphingoid base (123, 411).
The mammalian mutant could be rescued by exogenous
SM, but not by glucosylceramide (GlcCer), the precursor
of higher glycosphingolipids (121, 123). It therefore appears that SM, and not GlcCer (or higher glycosphingolipids), fulfills a vital role in mammalian cells. Indeed,
Physiol Rev • VOL
mammalian cell lines unable to synthesize GlcCer are
viable and proliferate (145). The alternative monoglycosylceramide, galactosylceramide (GalCer), is not required
for cell survival either, since it is absent from most mammalian cell types including the GlcCer-negative cells
(145).
Collectively, these observations raise two major
questions. What vital function do phosphosphingolipids
serve in the cell? And for what purpose do animal cells
create so many different types of glycosphingolipids in
addition, given that, in principle, they can live without
them? Complete answers to these questions are still
lacking. However, experimental work over the last decade has indicated that cells exploit both chemical and
biophysical characteristics of sphingolipids to accomplish some of their most fundamental tasks. Before
exploring which of these characteristics account for
the sphingolipid requirement in cell growth and sur-
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
FIG. 2. Pathways of ceramide synthesis. Known metabolic intermediates, substrates, and enzymes are indicated.
Names of genes whose products are required for specific steps in yeast ceramide synthesis are shown in italics and
discussed in section VA. Note that it remains to be established whether C4-hydroxylation occurs on the free sphingoid
base (sphinganine) or on ceramide.
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
1693
vival, we will first turn our attention to what is known
about the sphingolipid requirements of complex multicellular organisms.
B. Sphingolipid Requirements at the
Organismal Level
Physiol Rev • VOL
FIG. 3. Major routes of sphingolipid synthesis in eukaryotic cells.
Ceramide serves as the immediate precursor for the phosphosphingolipids inositol phosphoceramide (IPC), ethanolamine phosphoceramide
(EPC), and sphingomyelin (SM), which obtain their phosphoinositol,
phosphoethanolamine, and phosphocholine headgroups by transfer
from the corresponding glycerophospholipids. In addition, ceramide can
be converted to the glycosphingolipids galactosylceramide (GalCer) and
glucosylceramide (GlcCer) by transfer of galactose or glucose from
UDP-Gal or UDP-Glc. GalCer can be further processed by the addition of
galactose (Ga2Cer), sialic acid (GM4), or sulfate (SGalCer). In contrast,
GlcCer can only be converted to LacCer, which in turn serves as the
precursor for a large number of glycosphingolipid series. The ganglioseries (GgCer) are conveniently described by the Svennerholm nomenclature (GA2, GM3, etc.). Although the term sulfatide is generally used
for SGalCer, often it is meant to include SLacCer.
their specific carbohydrate moiety, and in the modulation of plasma membrane signal processing. The latter
may occur via specific glycosphingolipid-protein interactions (119) or via organizing functions of the glycosphingolipids in signaling domains (118, 202).
It should be noted that not only sphingolipid production, but also the proper removal of sphingolipids, is of
physiological significance for an organism. The absence
of hydrolytic enzymes down to ceramidase or cofactors
like sphingolipid activator proteins from the lysosomes
results in lysosomal storage of sphingolipids with characteristic and often severe pathologies in humans (167, 171,
186, 260). Strategies devised to cure the symptoms of
these diseases now include enzyme replacement and gene
therapy (22, 29, 323, 435) and prevention of accumulation
by administrating an inhibitor of glycosphingolipid synthesis (153, 287).
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
Gene knock-out studies in mice are starting to reveal
the physiological significance of sphingolipids in mammals. First of all, in addition to phosphosphingolipids
being indispensable for cell proliferation (see sect. IIA),
glycosphingolipid production is essential for development
and differentiation. Mice lacking GlcCer and all complex
glycosphingolipids due to disruption of the gene encoding
ceramide:glucosyltransferase were found to be embryonically lethal at the gastrulation stage just after formation
of the primitive germ layers (421). After ectopic transplantation, glycosphingolipid-deficient embryonic stem cells
were able to differentiate into endodermal, mesodermal,
and ectodermal derivatives but failed to give rise to mature, well-differentiated tissues.
Knock-out mice have also been generated without
a functional ceramide:galactosyltransferase (27, 58),
the enzyme that synthesizes GalCer, which is the precursor of only a few other lipids like sulfatide (HSO3 -3
GalCer). These mice live, but male mice were unable to
breed, which reflects a function of galactolipids in spermatogenesis (95). In addition, such mice displayed
compromised nerve fuction (27, 58), a finding consistent with the putative insulatory function of the bulk
quantities of GalCer and sulfatide found in myelin, the
plasma membrane of oligodendrocytes, and Schwann
cells. High concentrations of GlcCer (murine) and/or
GalCer (bovine, human) and derivatives have also been
found in the apical plasma membrane domain of epithelial cells lining the gastrointestinal and part of the
urogenital tract (341). Here, the glycosphingolipids are
thought to provide mechanical stability and, especially
in the gut, protection against harmful, hydrolytic enzymes such as phospholipases.
In contrast to GlcCer and GalCer, complex glycosphingolipids are mostly present on cell surfaces in low
quantities only. Their importance for development and
differentiation is supported first of all by the observation that knock-out mice being unable to synthesize the
complex sialoglycosphingolipids. Of the double knockout mice being unable to synthesize glycosphingolipids
beyond the simple ganglioside GM3 (Fig. 3), 50% died
within 13 wk (158). Mice unable to synthesize glycosphingolipids beyond GM3 and GD3 displayed axonal
degeneration and decreased myelination (332, 371). As
a possible explanation, it has been proposed that glycosphingolipids are involved in specific recognition
events between cells and between cells and matrix via
1694
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
C. Do Sphingolipids Exert a Vital
Signaling Function?
Physiol Rev • VOL
D. Sphingolipids and the Spatial Organization
of Cells
Clearly, sphingolipids are not just a reservoir of signaling molecules; they also contribute to vital properties
of cellular membranes. Studies of their physical behavior
(see sect. III) have provided thorough insights in the basis
of how sphingolipids and cholesterol induce lateral segregation of membrane components (see sect. IV). However, to understand the functional implications, we will
have to define the consequences of this lateral organization for activities in and on the membrane. With what
other molecules do sphingolipids interact, and for what
processes are these interactions relevant? If we want to
learn how sphingolipid-mediated processes are integrated
in the physiology of the cell, we will also need to know
how these processes are regulated at the level of the
sphingolipids. What rules govern their interactions at the
biophysical level, and what determines their concentration in the various cellular membranes in time? First
insights have been obtained from the localization of the
subcellular sites of sphingolipid synthesis and hydrolysis,
and from studying their mechanisms of transport (see
sect. V). Because metabolism and transport are mediated
by enzymes and transporters, regulation of these processes must be exerted at the level of the proteins and the
genes by which they are encoded. The available data
suggest a pivotal role for sphingolipids in the operation of
the Golgi complex, the central sorting station in the delivery of cargo, and membrane components to their
proper destinations (see sect. VI). So far, it was believed
that sorting processes were governed exclusively by information in the molecular structure of proteins. We now
start to realize that sphingolipids produced in the Golgi
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
An impressive number of studies have implicated
sphingolipids in virtually all aspects of cellular signaling
(reviewed in Refs. 75, 118, 127, 356). First, sphingolipids
serve as ligands for receptors present on neighboring cells
(or in the matrix) to trigger various types of cell behavior
(growth, adhesion, differentiation, migration). Second,
sphingolipids influence properties of receptors on the
same cell via specific lipid-protein interactions, thereby
changing the cellular responsiveness to external stimuli
(119). Third, sphingolipids modulate signaling by their
ability to assemble both receptors and their downstream
effectors (e.g., Src family kinases, G proteins) in specialized plasma membrane microdomains, known as rafts and
caveolae (7, 36, 118, 147, 202, 340). Finally, sphingoid
bases, ceramides, and their phosphorylated derivatives
act as signaling molecules in the regulation of membrane
trafficking, cell growth, cell death, and the ability of cells
to cope with environmental stress (13, 152, 346, 356).
Especially this last paradigm has attracted much attention
in the recent literature. Although ceramide-activated protein kinases and phosphatases have been implicated in
transmitting sphingolipid-derived signals (126, 431), the
mechanisms by which ceramide pathways operate have
not been elucidated (138, 139, 170, 409). Recent work has
shown that ongoing synthesis of sphingoid bases forms a
prerequisite for the internalization step of endocytosis in
yeast (429). It appears that sphingoid base levels help
control the relative activities of specific protein kinases
and phosphatases whose downstream targets are elements of the endocytic machinery and/or actin cytoskeleton (93). Another exciting development in the field is the
emergence of sphingosine-1-phosphate as a prototype of a
new class of lipid signaling molecules that function not
only as intracellular second messengers, but also as extracellular ligands for cell surface receptors (356). In
support of the extracellular ligand function, several
closely related transmembrane receptors have recently
been identified as putative sphingosine-1-phosphate receptors in mammals (6, 191).
Sphingolipid signaling pathways have been found to
operate in many different cell types, from mammals down
to yeast (74). The impressive array of cellular processes
that appears to be regulated by these pathways would
provide a logical explanation for the observed lethality of
sphingolipid-deficient mutant cells and organisms. However, studies in yeast have demonstrated that the putative
signaling function of its sphingolipids is dispensable for
cell growth and survival, although only under nonstressed
conditions. A mutant strain lacking sphingolipids has
been isolated upon suppression of a genetic defect in
sphingoid base synthesis. This suppression is due to a
mutation in the SLC1 gene, believed to encode a fatty
acyltransferase (261). The suppressor mutation enables
cells to produce a novel set of glycerolipids that mimic
sphingolipid structures, both with respect to their headgroup and fatty acyl chain composition (193). The novel
lipids identified were phosphatidylinositol (PI), mannosylPI, and inositol-P-(mannosyl-PI), all containing a C26 fatty
acid in the sn-2 position of the glycerol moiety. Normally
the C26 fatty acid is not found in yeast glycerolipids, but
only in the sphingolipids (Fig. 1) and in the lipid backbone
of some glycosylphosphatidylinositol (GPI)-anchored
proteins. When exposed to extremes of pH or temperature, the suppressor mutant fails to grow unless provided
with externally added phytosphingosine (74, 193, 281).
These and other observations (76) show that yeast requires sphingolipids to build up a proper stress response.
In contrast, the essential function of sphingolipids in
growth and survival under normal conditions can be
taken over by the novel glycerolipids, and is, apparently,
structural.
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
may well form indispensable parts of the organelle’s sorting machinery. Our challenge is to find out how the machine works.
III. SPHINGOLIPID STRUCTURE AND
BIOPHYSICAL PROPERTIES
A. A Concise Inventory of Sphingolipid Structure
in Distinct Eukaryotic Life Forms
Physiol Rev • VOL
of ethanolaminephosphorylceramide (213). The latter is
the principal phosphosphingolipid found in insects (304,
414). Tremendous diversity exists among the carbohydrate head groups of glycosphingolipids (Fig. 3), which
may contain as many as 30 glycosyl residues per lipid. A
glycosphingolipid is termed ganglioside if one or more of
its sugar residues is a sialic acid. Gangliosides are particularly abundant in the central nervous system of higher
organisms (240). Whereas sphingolipids display a striking
organism- and cell type-dependent variation in head
group composition, this is generally not the case for the
glycerolipids. In most eukaryotic cells, glycerolipids utilize phosphate ester-linked choline, serine, ethanolamine,
and inositol to achieve diversity.
One additional feature that makes sphingolipids different from glycerolipids is their fatty acyl chain. Typically, the amide-linked fatty acyl chains in sphingolipids
are long and saturated. Often they are hydroxylated at the
␣-position.3 In mammals, SM typically consists of two
types, roughly half containing C16:0 and C18:0 and the
other half C22:0, C24:0, and C24:1 (12, 156). The glycosphingolipids have similar fatty acids, but, depending on
the tissue, a large fraction (up to 70%) may consist of the
␣-hydroxylated form of each of these fatty acids (28, 268).
The acyl chains in the sn-2 position of mammalian glycerophospholipids, on the other hand, are mostly shorter,
(poly)unsaturated (e.g., C18:1, C18:2, or C20:4) and not
hydroxylated. As illustrated in Figure 1, this striking difference in length, saturation, and hydroxylation status
between fatty acyl chains of sphingolipids and glycerolipids has also been observed in Caenorhabditis elegans
(55), Drosophila (304), and yeast. In yeast sphingolipids,
C26:0 and hydroxylated C26:0 are the major fatty acids
(192, 411). Interestingly, the prevalence of monounsaturated species of aminophospholipids at the expense of
diunsaturated species in the plasma membrane of wildtype yeast is reversed in elo3 mutant cells that accumulate
C22:0 fatty acid-containing sphingolipids instead of the
normal C26:0 fatty acid-containing ones (269, 325). This
observation indicates that the structural differences between the fatty acid chains of sphingolipids and glycerolipids serve important biological functions.
We conclude that, despite the considerable variation
in chemical composition between lipids of distinct organisms, the structural features by which sphingolipids can
be discriminated from glycerolipids have been preserved
from vertebrates down to flies, worms, and fungi. Moreover, when these differences are undermined by muta-
3
Skin contains sphingolipids carrying very-long-chain (C28-C34)
fatty acids that are hydroxylated at the ␻-position. After having been
esterified to linoleic acid and hydrolyzed, the ␻-hydroxyl serves to
couple the sphingolipid, first glucosylceramide and later ceramide, to
proteins on the cell surface (79, 412) where they are essential for the
permeability barrier of the skin (17).
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
The key structural features by which sphingolipids
can be distinguished from glycerolipids and sterols are
outlined in Figure 1. First of all, the backbone of sphingolipids is a sphingoid base. With respect to molecular
conformation, the sphingoid base is structurally equivalent to the glycerol and the sn-1 acyl chain in a typical
glycerolipid. Saturated sn-1 chains of 16 or 18 carbons
(C16:0, C18:0) predominate in most eukaryotic glycerolipids. The prevalent sphingoid base found in mammalian
sphingolipids is sphingosine with a chain length of 18
carbon atoms and a trans-double bond between carbons
4 and 5. However, over 60 different species of sphingoid
bases have been described with alkyl chain lengths varying from 14 to 22 carbon atoms, different degrees of
saturation [the saturated form of sphingosine is called
sphinganine (53)] and hydroxylation (154). The saturated
and 4-hydroxylated form of sphingosine is generally referred to as phytosphingosine because it is the common
sphingoid base found in sphingolipids of plants and fungi.
In vertebrates, high levels (even up to 70%) of phytosphingosine-based sphingolipids have been found in kidney and
intestine (8, 28, 113, 155, 266). Sphingolipids in insects are
based on a C14 sphingoid base rather than the C18 backbone generally found in mammals, plants, and fungi (192,
304, 414). Sphingolipids of nematodes are unusual in that
they are based entirely on a C16 sphingosine methylated
at C15 (55, 105).
The amino group at the second carbon of the sphingoid base serves to attach a fatty acid in amide linkage.
The N-acylated sphingoid base ceramide is the precursor
of all membrane sphingolipids. The trivial name of ceramides based on sphinganine is dihydroceramide. The
sphingoid C1 hydroxyl group in ceramide is used as the
attachment site for a polar head group. Based on their
head groups, sphingolipids are often grouped into phosphosphingolipids and glycosphingolipids. However, these
categories are not mutually exclusive; plants and fungi,
for example, add phosphoinositol to phytoceramide to
generate IPC, and this head group can be further decorated with one or more monosaccharides. Sphingolipids
with inositol phosphate-containing head groups are also
common in protozoans (192). The main phosphosphingolipid in mammals and nematodes, SM, carries a phosphocholine (12, 55). Mammals also produce small quantities
1695
1696
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
tion, organisms will exploit lipid-remodeling mechanisms
to recreate structural diversity (193, 325). This presents us
with the question, What would be the functional significance of these structural differences for eukaryotic life?
B. Biophysical Differences Between Sphingolipids
and Glycerolipids
4
The distance between the head groups in a double layer of C18:0
SM in the presence of cholesterol was reported to be 46 – 48 Å (232, 300),
whereas this was 40 Å for C16:0-C18:1 PC in the presence and 35 Å in the
absence of cholesterol (262, 300).
Physiol Rev • VOL
IV. SPHINGOLIPIDS AND THE LATERAL
ORGANIZATION OF BIOLOGICAL
MEMBRANES
A. Ordered Sphingolipid Domains
in Model Membranes
At low temperatures, bilayers of a single phospholipid or sphingolipid exist in a frozen state. Above a
melting temperature (Tm) characteristic of each lipid, the
bilayer enters a phase in which the lipid acyl chains are
fluid and disordered (50). The frozen state is referred to as
the gel, solid-ordered (so), or lß phase. The fluid phase is
5
The cis double bond in C24:1, which in its 2-hydroxylated form or
its nonhydroxylated form is found in a significant fraction (up to 40%) of
the sphingolipids in mammalian myelin (89) and intestine (28), is located
at C15, deep in the bilayer, where it does not disrupt the close packing
arrangement of the sphingolipids (204) and the sphingolipid-cholesterol
interaction (296). This contrasts with the C9 cis double bond in the C2
fatty acids of the glycerolipids.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
The interfacial region that links the nonpolar hydrocarbon region to the polar headgroup provides sphingolipids and glycerolipids with different biophysical properties. In sphingolipids, this region contains the 2-amide and
3-hydroxyl groups, often supplemented with additional
hydroxyls at the sphingoid C4 and fatty acid C2, and with
further hydroxyls on the carbohydrates. These groups can
function both as hydrogen bond donors and acceptors
and are thought to participate in extensive inter- and
intramolecular hydrogen bonding of the sphingolipids,
whereas glycerolipids have only hydrogen bond-accepting
properties in this part of the molecule (278). This property
allows sphingolipids but not glycerolipids to associate
with themselves in the plane of the membrane into a
highly flexible hydrogen-bonded network (23; see sect.
IIIB). In membranes exposed to physical and chemical
stress as in the apical membranes in kidney and intestine
and in microorganisms, the sphingolipids exhibit an increased number of hydroxyl groups. This property is believed to improve the stability and decrease the permeability of those surface membranes (278). Although this
may be generally true, under certain conditions the loss of
a specific hydroxyl was found to increase membrane stability in yeast (82).
The lipid tails of natural sphingolipids are more
tightly packed than those of the most abundant phosphoglycerolipid, monounsaturated PC. The surface area of
GlcCer and GalCer at a surface pressure of 30 mN/m,
typical of plasma membranes, is 0.40 nm2 (212), versus
0.63 nm2 for stearoyl-oleoyl PC (71). The fatty chains of
sphingolipids are therefore more extended, and consequently the thinner and taller sphingolipids form up to
30% thicker membranes than unsaturated phospholipids.4
Second, sphingolipids contact their neighbors along a
greater and flatter surface, which results in a dramatic
increase in the van der Waals attraction between neighboring sphingolipid molecules. Van der Waals interactions
have also been held responsible for the strong binding
between the sphingolipids and cholesterol with its rigid
and flat-cylindrical steroid backbone (235). A preferential
binding of cholesterol to sphingolipids, in particular to
SM, has been observed in many biophysical studies (23).5
Interestingly, the surface areas of disaturated PC (0.41–
0.44 nm2; Refs. 71, 212) and PS (0.40 – 0.44 nm2; Ref. 72)
are close to those of sphingolipids, and a preferential
interaction of cholesterol with diC16:0 PC (less than or
equal to that with SM; Refs. 181, 208) over mono-unsaturated PC has been observed (208). This illustrates the
possibility that disaturated phospholipids may resemble
sphingolipids in their hydrophobic interactions with other
membrane components. This is especially relevant for the
cholesterol-rich plasma membrane, where 40% of the PC
and 75% of the PS are disaturated, while disaturated species, sphingolipids, and cholesterol are virtually absent
from the ER (160).
The sphingoid chain extends less far into the membrane core than the N-linked fatty acid, a disparity not
found in glycerolipids (Fig. 1). It has been argued that the
difference in length between the two chains leads to
packing defects (voids) in the membrane interior and that
these defects may be complemented via interdigitation of
two sphingolipids in the opposed bilayer leaflets (23, 44,
324). This may be an example of how lipids in one bilayer
leaflet may couple to lipids in the opposite leaflet (see
sect. VIIB). Interestingly, also cholesterol molecules in the
two opposed bilayer leaflets may interact by forming
transmembrane cholesterol dimers (129, 205, 250).
Whether these transmembrane lipid interactions are of
physiological relevance depends first of all on whether
sphingolipids and cholesterol are present in both bilayer
leaflets, and second on the relative affinities involved in
transmembrane interactions and in lateral interactions of
the lipids.
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
Physiol Rev • VOL
glycerolipids into an lo phase (337).6 Ceramides at low
concentrations occur as single monomers in a membrane,
whereas at high concentrations they partition into hydrophobic ceramide droplets in between the two leaflets of
the bilayer (404). At intermediate concentrations (⬃10
mol%; Ref. 141), the ceramides can aggregate to form a
ceramide domain. The higher the affinity between ceramides, e.g., long acyl chain containing hydroxylated ceramides of yeast versus short acyl chain and sphingosine
containing ceramides of mammals (204), the lower the
ceramide concentration needed for segregation.
In a membrane with coexisting ld and lo phases, the
distribution of each lipid species over the two phases is
not governed solely by its relative affinity for the lipids in
either phase. A different energy parameter is the longrange order of the lipids. According to the superlattice
view (350, 372), lipids tend to be regularly distributed.
Cholesterol with its different shape and smaller surface
area imposes a steric strain on the alkyl chain matrix. To
minimize the strain, cholesterol will adopt a regular lateral distribution, which yields a so-called superlattice. In
a particular phospholipid mixture, only a defined set of
cholesterol concentrations fits a superlattice, and this has
been experimentally confirmed (349). At any given cholesterol concentration, a membrane would thus consist of
domains displaying one of the allowed, minimum-energy
superlattice organizations. A particular lipid will be arranged according to some superlattice pattern whether it
is present in an lo or in an ld phase. Still, individual
molecules will rapidly exchange positions, whereby the
rate of exchange, and thus lateral diffusion, will be higher
in the ld phase. The degree to which each lipid type fits the
superlattice of one phase will contribute to its partitioning
between the two phases.
B. Evidence for the Existence of Ordered
Sphingolipid Domains in Cellular Membranes
An overwhelming number of studies in artificial bilayers have demonstrated that lipids have a strong selforganizing capacity; lipid immiscibility can drive phase
separation and give rise to domains (used in the meaning
of “environments”) with unique lipid compositions and
biophysical properties. The current evidence does support the existence of such phase-separated lipid domains
6
Phase separation will only occur above a critical concentration of
a lipid in the membrane. In analogy to the aggregation of detergent into
micelles above the critical micellar concentration (CMC), the concentration above which a lipid aggregates in an environment of other lipids
may be called the critical domain concentration (CDC). Just as the CMC
depends on the molecular characteristics of the detergent, the CDC
depends on the molecular characteristics of the aggregating lipid. The
CDC also depends on the lipid matrix within which the aggregate forms.
Furthermore, it depends on the presence of similarly aggregating lipids,
just like detergents in water coaggregate with other detergents.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
termed the liquid-crystalline, liquid disordered (ld), or l␣
phase (173, 319). Sphingolipids (especially glycosphingolipids) display a much higher Tm than glycerolipids, due to
their denser packing, which in turn is due to the saturation of their fatty chains and to the fact that they are more
prone to intermolecular hydrogen bonding. While the
phase transition for the typical mono-unsaturated PCs
C16:0-C18:1 PC and C18:0-C18:1 PC occurs at ⫺3 and 7°C,
this temperature is 42°C for diC16:0 PC, 45°C for C18:0
SM, but 85°C for C18:0 GlcCer and GalCer (see Refs. 172,
173). Studies with model membranes consisting of binary
mixtures of lipids with different Tm revealed that at temperatures between the Tm of the two lipids, a cooperative
phase separation can occur, resulting in the coexistence
of gel and ld phase. The physiological significance of this
type of phase separation in eukaryotes is questionable,
mainly because membranes that contain high amounts of
lipids with Tm values at or above room temperature, like the
plasma membrane (160), also contain high levels of cholesterol. Cholesterol at concentrations of 30 –50 mol% abolishes the gel-liquid crystalline phase transition (179, 319).
Remarkably, a phase separation between two fluid
phases has been described in model membranes containing binary mixtures of a high-Tm lipid (saturated-chain
lipid) and cholesterol (299). At temperatures above the Tm
of the lipid, with increasing cholesterol concentrations a
liquid-ordered (lo) phase can separate from the liquiddisordered (ld) phase. Above a threshold concentration of
cholesterol, only the lo phase exists, independent of the
temperature. For diC16:0 PC, this occurs at 30 mol%
cholesterol (319). Acyl chains of lipids in the lo phase have
properties that are intermediate between those of the gel
and ld phases; they are extended and ordered as in the gel
phase but are laterally mobile in the bilayer as in the ld
phase. Also in bilayers of more complex composition the
coexistence of lo and ld phases depends on the cholesterol
concentration (reviewed in detail by Refs. 37, 40), and
three fluid phases can coexist, of which two are lo phases
(294). The phase separation is especially pronounced
when the Tm of the lipids is greatly different, like in the
case of PCs of different chain length (338). Disaturated
phospholipids and sphingolipids would preferentially partition into the lo phase, which as a consequence would
contain more cholesterol, whereas (poly)unsaturated lipids would prefer the ld phase. Using a fluorescence
quenching assay, Ahmed et al. (1) found that cholesterol
induces the formation of a SM-enriched lo phase at 37°C in
a mixture of SM and monounsaturated PC. Mixtures of
GlcCer, GalCer, and monounsaturated PC were found to
be significantly inhomogeneous at 37°C, regardless of
whether cholesterol was absent or present at concentrations at which the lateral separation in mixtures of SM
and PC is abolished. Hence, glycosphingolipids appear to
have a stronger tendency than SM to segregate from
1697
1698
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
in cellular membranes. However, many uncertainties remain concerning their composition, size, and dynamics
and concerning their location. In which of the subcellular
membranes do they occur? Do they exist on both sides of
the membrane, and if so, do domains on the one surface
colocalize with domains on the opposite surface? Are domains the cause of variations in biomembrane thickness?
Finally, a most relevant question: by what molecular interactions can membrane proteins discriminate between
different lipid environments (69), and what is the contribution of membrane proteins to the properties of the domains?
1. Early evidence for lipid lateral heterogeneity
2. From macrodomain in epithelial plasma membrane
to microdomain in the trans-Golgi network
A unique type of sphingolipid (macro)domain organization exists in the plasma membrane of epithelial cells,
where originally an accumulation of glycosphingolipids
was reported for the apical plasma membrane domain of
intestinal epithelial cells (92). Later work (30, 81, 159)
demonstrated a two- to fourfold enrichment of glycosphingolipids in the apical versus the basolateral domain
of the continuous plasma membrane of these cells. Similar differences have been reported for urinary bladder
epithelium (369) and, indirectly, for Madin-Darby canine
kidney (MDCK) cells derived from the distal nephron
(341). Apical membranes from kidney cortex, which contain low levels of glycosphingolipids (358), contained an
exceptionally high level of SM (35% compared with 13% in
basolateral membranes; Refs. 48, 137, 247, 330, 402).
These differences reside in the outer leaflet of the plasma
membrane where they are maintained by the presence of
a barrier to lipid diffusion, the tight junctions (reviewed in
Ref. 341). Similar surface domains have been proposed to
exist in sperm (reviewed in Refs. 14, 102), and a lipidphase separation has been held responsible for their
Physiol Rev • VOL
3. Resistance to detergent extraction
The first candidate proteins of sphingolipid membrane domains were the proteins that are anchored in the
noncytoplasmic bilayer leaflet of apical membranes by a
GPI anchor (201). These proteins were found to be targeted to the apical cell surface, obviously via some interaction in the luminal leaflet of the Golgi membrane (34,
200). At the same time it was observed that influenza virus
hemagglutinin, which is targeted to the apical membrane
of infected epithelial cells, became partially resistant to
extraction by Triton X-100 (TX-100) in the cold during
transport to the surface (344), and it was proposed that
the protein may complex with glycosphingolipids, as glycosphingolipids had been shown to be TX-100 insoluble
as well. Progress in the field was tremendously accelerated when it was realized that not only hemagglutinin and
sphingolipids but also the GPI proteins show resistance to
extraction by detergent at low temperature, that this behavior might reflect the presence of these components in
GPI protein/sphingolipid microdomains in the membrane,
and that detergent insolubility might be an experimental
tool to study the domains (38). The method consisted of
the addition of 1% TX-100 to cells on ice for 20 min,
homogenization of the cells, addition of sucrose to 40%,
and flotation in a 30%-5% linear sucrose gradient without
detergent. A GPI protein was found to float to a density of
1.081 g/ml,9 and closer inspection showed that the GPI
protein was present in closed membrane vesicles that
were now resistant to TX-100 (detergent-resistant mem-
7
Although GM1 was found to be exclusively localized in axons of
hippocampal neurons in culture (188), this was not confirmed in independent studies in the same system (335). Also, the presence of a lipid
diffusion barrier between axon and dendrites (165) has been contested
(99).
8
With an SM content of 8 –12% in the total Golgi (90, 160) and with
all SM in the luminal leaflet, 16 –24% of that surface might be covered by
a sphingolipid microdomain. The expected enrichment of SM in the TGN
to plasma membrane levels (16 –19%), the presence of cholesterol, and
participation of disaturated PC (9% of total Golgi lipids; Ref. 160) increases the relative area of the TGN luminal surface covered by the
microdomain to numbers above 40%. In this sense, the term microdomain is misleading.
9
A density of 1.081 g/ml equals 20% sucrose (wt/wt), 21.6% sucrose
(wt/vol), or 0.63 M sucrose.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
The fluid-mosaic models of biomembranes in which
proteins freely float around in an oily lipid bilayer (148,
343) marked a breakthrough in the thinking about membrane structure and function. However, already at that
time, evidence became available suggesting that lipids
and proteins were not randomly distributed in the membrane. For the lipids, phase separations had been demonstrated in the plasma membrane of Acholeplasma by calorimetry (363) and by freeze-fracture electron microscopy
(403). Electron spin resonance studies suggested lipid
phase heterogeneity in the ER (234, 364), and subsequently, a variety of techniques suggested the same for
the plasma membrane of mammalian cells (157). Ideas on
long-range order in the organization of lipids and proteins
and the occurrence of microdomains were then formulated (148).
maintenance.7 Based on sphingolipid transport experiments (398), it was subsequently postulated that the
sphingolipid-rich apical macrodomains originate from
sphingolipid- and cholesterol-enriched microdomains in
trans-Golgi membranes (339, 341, 395).8 These would
then act as a sorting device for the delivery of apically
directed cell surface components in polarized epithelial
cells (see sect. VI).
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
10
The term DRM describes the remnant left after detergent extraction. Formally, this term cannot be used to describe a domain that
existed in the original membrane before detergent addition, since this
would assume that detergent has no effect on the properties of the
domain. The latter must be proven for each property studied.
11
Although also hemagglutinin became detergent insoluble after
20 min (344), this may have a different reason. The insolubility of
hemagglutinin requires its triple palmitoylation (sites), and the organelle
where palmitoylation takes place has not been established (237). So
insolubility 20 min after synthesis might be due either to contact with
sphingolipids in the Golgi or to palmitoylation in the Golgi.
12
However, 60 – 80% of the GalCer and cholesterol in myelin were
resistant to each detergent, showing that at least part of the GalCer was
resistant to both detergents. One possible explanation is that there were
three different domains: one domain containing TX-100-resistant proteins, one containing CHAPS-resistant proteins, and a third domain
resistant to both detergents and containing some 40% of the GalCer.
Alternatively, there was only one type of domain, but TX-100 and CHAPS
selectively extracted certain proteins.
Physiol Rev • VOL
by the branched isoprenyl groups are not recovered in
DRMs (237). This supports the idea that an ordered lipid
environment is present on the cytoplasmic surface of the
sphingolipid rafts. Some of the DRMs from the plasma
membrane were derived from caveolae, morphologically
well-established flasklike invaginations of the plasma
membrane (7, 37). Caveolae are defined by the presence
on the cytosolic surface of the palmitoylated, cholesterolbinding protein caveolin (7, 256). Multiple palmitoylation
of caveolin was found to be required for its interaction
with dually acylated G proteins (103), but not for binding
to caveolae (78). Caveolin (VIP21) was also localized to
the Golgi (177, 207), where it is supposed to be involved
in membrane transport. DRMs have been isolated from
many different mammalian cell types, but also from Drosophila (304), Dictyostelium (419), yeast (11, 174), and
protozoans (430).
Obviously, the detergent extraction method has its
limitations. It cannot be concluded that the organization
of the lipids (and proteins) in the resulting DRMs reflects
their organization in the cellular membrane of origin in
full detail, because some rearrangement may have occurred as a consequence of low temperature, detergent
addition, and detergent dilution. Low temperature favors
formation of ordered domains, while detergent will insert
into the membrane and influence the partitioning of the
various lipids. Notably, the size of the original sphingolipid domains cannot be assessed by detergent extraction,
because separate domains within the same membrane
may coalesce due to removal of the surrounding glycerolipids. Because the bulk of the sphingolipids has been
localized to the noncytoplasmic surface of cellular membranes (see sect. VB), one other uncertainty is the composition of the lipids on the cytosolic surface facing the
sphingolipid domain and their behavior during detergent
extraction (discussed in detail in Ref. 36). Various groups
have published methods to isolate caveolae without the
use of detergent (326, 348, 351, 361). Whereas three
groups set out by isolating plasma membranes, the fourth
group (351) followed the detergent extraction method but
without detergent. Cells were homogenized and sonicated, and the membranes were floated up in a sucrose
gradient. Membranes at the 35/5% sucrose interface were
separated from membranes at the 45/35% sucrose interface and were designated “purified caveolae membranes.”13 It will be interesting to see how these protocols
will complement the DRM method.
13
According to common cell fractionation experience, the 35%/5%
sucrose interface should contain most of the Golgi, early endosomes,
and plasma membrane. The fraction at the 45%/35% sucrose interface
typically contains ER and lysosomes, while the alkaline carbonate used
for homogenization should cause a large part of the ER to float up to the
35%/5% interface as well. Without any determination of the flotation of
these membranes, the use of the term purified caveolae membranes is
optimistic, if not naive.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
branes; DRMs).10 The low-density vesicles contained
some 10% of the total cellular lipids: all of the SM, 50% of
the glycosphingolipids, 25% of the cellular cholesterol, but
only 5% of the glycerolipids (38). Subsequent work on
model membranes has suggested that SM is detergent
insoluble only when it was present in an l0 liquid-ordered
phase before the addition of the detergent (1), supporting
the idea that DRMs are derived from preexisting l0 phase
domains. The power of the DRM isolation method to study
the structure and function of domains was demonstrated by
the finding in the original paper (38) that the GPI proteins
became detergent insoluble at 20 min after synthesis only,
which suggested that the proteins became insoluble upon
entering the Golgi where they are thought to come into
contact with the sphingolipids (see sect. VI).11 The method
has been applied to identify proteins involved in the transport pathway from the trans-Golgi network (TGN) to the
apical domain (180, 304). Evidence has been reported for
the coexistence in the same membrane of domains with
different (GPI) protein compositions and different physical properties (211, 307). The same was reported for
myelin where TX-100 and CHAPS yielded DRMs of different protein compositions (342).12 SM-enriched DRMs
have been isolated from nuclei, where they were assumed
to represent the inner nuclear membrane (18).
DRMs have been isolated from the plasma membrane, suggesting the existence of phase-separated sphingolipid/cholesterol domains on the cell surface where
they would compartmentalize, modulate, and integrate
signaling events by providing sites for the assembly of
components involved in signal transduction (202, 340).
Exciting new developments in the field are the regulation,
by ligand binding, of the association with the domain of
signaling receptors and of the Src family kinases that
transduce the signals into the cell. The latter kinases are
anchored to the cytoplasmic bilayer leaflet by two saturated acyl chains, whereas in contrast proteins anchored
1699
1700
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
5. Optical studies on living cells
Independent evidence for clustering of glycosphingolipids on the cell surface has been obtained by microscopy. A first approach has utilized glycosphingolipid-binding proteins, which were visualized by a fluorescent or
electron-dense tag or by a secondary labeled protein. The
major problem in such studies is that lipids cannot be
fixed at their original location. Artificial clustering was
induced when the binding of a primary antibody to globoside (Gb4Cer) or Forssman glycosphingolipid (IV3-␣GalNAc-Gb4Cer) was followed by labeling with a dimeric
secondary antibody, tetrameric protein A, or when multimeric complexes were used of the primary or secondary
ligand to ferritin or colloidal gold (47, 96, 377). Clustering
of the ganglioside GM1 was observed when it was labeled
with pentameric cholera toxin B subunit by itself (9) or
conjugated with gold into a multimeric complex (276), or
when a biotinylated GM1 was labeled with anti-biotingold (246). The latter studies were performed on freezesubstituted samples in which redistribution seems rather
unlikely. In an independent approach, labeling of glycosphingolipids with a primary antibody was followed by
fixation before the addition of the secondary antibodygold complex, a condition that had been shown to prevent
redistribution of Forssman glycosphingolipid (47, 96).
Clusters of GM3 were still observed in one study (354),
while clustering of a number of sphingolipids in caveolae
was no longer observed under these stringent conditions
(96). Still, a local enrichment of the ganglioside GM1 in
caveolae was found by a postembedding labeling protocol
using cholera toxin where redistribution could be excluded (276).14 In cells transfected with influenza virus
hemagglutinin or GPI proteins, the distribution of these
proteins overlapped with that of GM1 (128). Clustering of
GPI proteins has been observed by microscopy under a
variety of conditions (7). However, in many cases clustering was induced by the protocol used. A major point of
concern in most microscopic studies on lipid and lipidlinked molecules is the use of multimeric reporter ligands
without proper controls.
Already in the early 1980s, measurements on the
behavior of fluorescent probes in biomembranes supported the concept of lipid domains in membranes (157),
and microscopy was performed on living cells using fluorescent lipids (357). In the latter study, the redistribution
of GM1 by cholera toxin caused cocapping of the unrelated ganglioside GM3, which suggested that the headgroup-labeled GM1 and GM3 were associated by lipidlipid interactions. Much more recently, a major
breakthrough in the field has been the application to living
cells of novel high-resolution optical techniques, like resonance energy transfer between fluorescent membrane
molecules, single particle tracking, two-dimension scanning resistance, and single dye tracing. A number of these
studies support the existence of locations on the cell
surface that are enriched in GPI proteins and glycosphingolipids and, in addition, of areas with enhanced resistance to lateral diffusion that are preferred by some but
not by other probes. One controversial issue is the size of
the domains. What is their diameter? Whereas the detergent-extraction studies yielded DRM vesicles with a diameter of 0.1–1 ␮m (39), suggesting a diameter size of 200 –
2,000 nm, this was a few hundred nanometers for the
small regions to which a GPI protein and GM1 were found
to be confined in particle tracking studies (290, 331).15
The newest single particle tracking measurements on
these domains suggest that they are even smaller with a
diameter of 50 nm (roughly 3,500 lipids), exist for more
than 1 min, and comprise ⬍50 proteins (289). Chemical
cross-linking and fluorescence resonance energy transfer
to measure GPI-protein interactions led to estimates of 70
nm (94, 401). In contrast to these data, no clustering of a
GPI protein was observed on the apical surface of MDCK
cells (162). In addition, a comparison between various
GPI proteins on various cell surfaces did not provide
evidence for the occurrence of a sizeable fraction of the
GPI proteins as stable clusters (163). The authors explained the discrepancies with the earlier work, by concluding that lipid rafts either exist only as transiently
stabilized structures or, if stable, comprise at most a
minor fraction of the cell surface.
Here, it becomes relevant to discuss the area of the
membrane covered by rafts. SM constitutes ⬃20% of the
plasma membrane phospholipids. If, as generally believed, SM is located in the outer leaflet, it covers 40% of
the surface. When saturated phospholipids and choles-
14
When sections of freeze-substituted A431 cells were labeled
with cholera toxin B-gold, clusters were observed in uncoated invaginations of the plasma membrane. It seems unlikely that clustering was
induced by the multimeric ligand as lipids most likely do not diffuse on
the surface of the section which in this case consists of Lowicryl
polymer. This is different in frozen sections, where during the labeling of
the thawed sections lipids are free to diffuse in the membranes. In the
freeze-substituted cells, GM1 was also found in small vesicles in the
cytosol in close approximation with the patches on the plasma membrane. Because the vesicle profiles were not in contact with the plasma
membrane during the labeling of the section, a higher concentration of
GM1 in these vesicles must have been present already in the living cell.
In the cell, the vesicle membranes are thought to be continuous with the
plasma membrane or to be derived from the plasma membrane by
endocytosis. In both cases, the results demonstrate that GM1 is enriched
in subdomains of the plasma membrane of these A431 cells.
Physiol Rev • VOL
15
It should be noted that each gold particle possessed many
binding sites against the GPI protein (antibodies conjugated to gold) and
to GM1 (cholera toxin B subunits conjugated to gold). In the case of the
GPI protein, the number of binding sites did not affect the results.
Concerning GM1, cholera toxin has been shown to reduce its solubility
in detergent, which implies that binding of the pentavalent toxin changes
the phase behavior of GM1 (117).
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
4. Microscopy on fixed cells
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
V. SPHINGOLIPID ASSEMBLY AND TRANSPORT
A. Enzymes of Sphingolipid Metabolism
and Their Topology
1. Ceramide synthesis
Sphingolipid synthesis in all eukaryotes starts in the
ER with the condensation of L-serine with palmitoyl coenzyme A, a reaction catalyzed by serine palmitoyl-transferase (SPT) and yielding 3-ketosphinganine (Fig. 2). SPT
forms the target of several potent natural inhibitors that
include sphingofungin (436), lipoxamycin (214), myriocin
(245), and viridiofungins (217). Genetic studies in yeast
Physiol Rev • VOL
revealed that at least two genes, LCB1 and LCB2, are
required for SPT activity, suggesting that the enzyme is
composed of at least two distinct proteins (43, 258). This
was confirmed when the proteins encoded by mammalian
homologs of these genes were characterized and found to
be part of a functional enzyme complex (121, 122). The
product of SPT, 3-ketosphinganine, is reduced to sphinganine (also called dihydrosphingosine), which in yeast
requires the TSC10 gene product (15). Reactions up to the
formation of sphinganine appear to be the same from
mammals down to yeast. Sphinganine in mammals is
N-acylated by ceramide synthase to form dihydroceramide, which is rapidly desaturated to give ceramide; a 4,5trans-double bond is introduced in the sphinganine moiety to form ceramide containing sphingosine as its longchain base (241). In yeast, but probably also in mammals,
sphinganine is hydroxylated to form 4-hydroxy-sphinganine, commonly named phytosphingosine. This reaction
requires the SYR2 gene product and, unlike earlier enzymatic steps in the biosynthetic pathway, is not essential
for cell growth in yeast (116). Amide linkage of a fatty
acid to phytosphingosine yields phytoceramide, the ceramide found in most fungal and plant sphingolipids and
abundant in some mammalian tissues (see sect. IIIA).
Ceramide synthesis is believed to take place on the cytosolic face of the ER (220). In mammals, this reaction can
be inhibited by fumonisins (405), while australifungin effectively blocks ceramide synthesis in yeast (216). So far,
no single acyl CoA-dependent ceramide synthase has
been purified or cloned. However, the alkaline ceramidases encoded by the yeast genes YPC1 and YDC1 possess a reverse, acyl CoA-independent ceramide synthase
activity that is insensitive to classical ceramide synthesis
inhibitors (222, 223). The very-long-chain (20⫹ carbon)
fatty acids found in (phyto)ceramides are formed by ER
membrane-bound fatty acid elongation systems whose
components belong to a evolutionary conserved family of
transmembrane proteins (269, 382, 383). Disruption of the
corresponding genes in yeast reduces the cellular sphingolipid levels (269) and induces pleiotropic phenotypes
such as bud site localization defects, changes in plasma
membrane H⫹-ATPase levels, and resistance to sterol
synthesis inhibitors (104, 303). The fatty acid moiety in
ceramides can be mono- or dihydroxylated. In yeast, the
first hydroxylation step occurs in the ER and requires
SCS7, a gene structurally related to the SUR2 gene (116).
The second hydroxylation takes place in the Golgi and
requires the Golgi copper transporter encoded by the
CCC2 gene (16).
2. Synthesis of sphingomyelin and glycosphingolipids
In animal cells, ceramide is used to produce SM and
GlcCer, the precursor for the higher glycosphingolipids
(Fig. 3). SM production occurs in the lumen of the cis and
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
terol are added to this figure, one would expect most of
the plasma membrane to be covered by a liquid-ordered
domain. In a special case, the apical surface of epithelial
cells of kidney and intestine, the surface is completely
covered by sphingolipids (341). This supports the idea
that the apical membrane of kidney cells, like MDCK, is
one big raft (162). Clustering on apical surfaces may
therefore not occur (162). When it occurs, it may imply
that there are different types of rafts (211, 307, 342).
Interestingly, when beads attached to a GPI protein (major histocompatibility complex class I) were scanned over
the surface of hepatoma cells by laser tweezers, the beads
experienced areas of increased resistance of 100 nm diameter, that were independent of the actin skeleton and
covered an area of ⬍10% of the surface, which is similar
to caveolae (370). The physicochemical basis of the existence of multiple types of rafts remains to be resolved but
may involve cholesterol-induced domains (294) and
caveolae (7), both of which are very stable. In contrast to
measurements on beads attached to diC16:0 phosphatidylethanolamine (PE) where no indications for areas of
higher resistance were observed (370), in tracings of single fluorescent molecules on muscle cells, the lipid analog
Cy5-diC14:0-PE, but not Cy5-diC18:1-PE (which were assumed to represent lipids preferring areas of lower and
higher fluidity, respectively), was 100-fold enriched in
domains of ⬃0.7 ␮m, that were stable for several minutes
(329). No ultrastructural correlate for such domains has
been found.
In line with the model of GPI protein sorting in
epithelial cells via sphingolipid domains (201, 341), a GPI
protein was found by various techniques to be clustered
during transport to the apical MDCK cell surface, after
which it diffused over the apical surface (125). However,
in contrast to inhibition of sphingolipid synthesis (233),
cholesterol depletion did not change polarity of cell surface delivery of the GPI protein (124). Cholesterol depletion did inhibit transport of hemagglutinin to the apical
membrane (161).
1701
1702
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
Physiol Rev • VOL
3. Synthesis of inositol sphingolipids
In contrast to animals, all fungi and plants studied
sofar, as well as several protozoa, add inositol phosphate
to phytoceramide to form IPC (192). The biosynthetic
route of inositol sphingolipids in yeast has been studied in
great detail. Yeast produces only three types of inositol
sphingolipids: IPC, mannose ␣1–2IPC: MIPC, and inositol1-P-6 mannose ␣1–2IPC or mannose-(inositol-P)2-ceramide: M(IP)2C. The IPC synthase, or an essential subunit
thereof, is encoded by the AUR1 gene (259). AUR1 homologs have been identified in a wide variety of fungi
(132, 176). IPC activity is effectively blocked by the antifungal agents aureobasidin A (259), khafrefungin (219),
and rustmycin (218). Because IPC synthase activity is
essential and not found in mammals, it provides an ideal
target for therapeutic drugs to fight pathogenic fungi in
immunocompromised individuals. IPC synthesis in yeast
was previously thought to occur in the ER. This idea was
based on the fact that the formation of IPC from ERderived ceramide and PI continues under conditions
when ER-to-Golgi vesicular transport is blocked in temperature-sensitive secretion (sec) mutants (292). However, fluorescence microscopy and membrane fractionation experiments have recently shown that both the
AUR1 gene product and IPC synthase activity are located
in the Golgi (195). This discrepancy in results remains to
be clarified but could be explained if the IPC synthase
would constitutively cycle between the ER and the Golgi,
a feature inherent of several Golgi-based proteins (60, 146,
210). An alternative explanation could be that ceramide
made in the ER can reach the Golgi by a vesicle-independent transport mechanism (see sect. VC).
Whereas the sidedness of IPC production is unknown, mannosylation to form MIPC most likely occurs
in the lumen of the Golgi. The latter reaction requires at
least three genes: SUR1, VRG4, and CSG2. SUR1 most
likely encodes a mannosyltransferase (16), whereas VRG4
is required for GDP-mannose transport into the Golgi
lumen (68). CSG2 encodes a member of the major facilitator superfamily, and its role in MIPC production is
unclear (432). The final and most abundant sphingolipid
in yeast, M(IP)2C, is formed by transfer of inositol phosphate from PI onto MIPC and requires a protein encoded
by the IPT1 gene (77). This reaction resembles the one
that yields IPC. Accordingly, the IPT1 and AUR1 gene
products exhibit a striking structural similarity.
4. Sphingolipid hydrolysis and signaling
sphingolipids
The major pathway along which sphingolipids are
degraded is removal of the head group and subsequent
hydrolysis of ceramide to sphingoid base and free fatty
acid. The breakdown products are then further metabolized or reutilized. Glycosphingolipids are degraded via a
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
medial Golgi and involves the transfer of phosphocholine
from PC to ceramide, yielding diacylglycerol as a (potentially important) side product (101, 150). A second SM
synthase activity has been located on the cell surface
(392). Two SM synthases with different properties have
also been found in the intra-erythrocyte stage of Plasmodium (120). The SM synthases remain to be identified.
The ceramide:glucosyltransferase involved in GlcCer production is located on the cytosolic surface of the early
Golgi (45, 65, 100, 151, 224). Whereas the GlcCer synthase
in mammals is encoded by a single gene, three GlcCer
synthase analogs have been identified in the C. elegans
genome (144). The physiological significance of this multiplicity remains to be solved. By an unknown mechanism,
GlcCer is translocated to the Golgi lumen where it can be
trapped by galactosylation to Galß1– 4GlcCer (LacCer;
Refs. 45, 183, 267). Various series of complex glycosphingolipids can then be generated by stepwise addition of
sugars to LacCer, whereby the first sugar and its glycosidic linkage determine the name of the series (53), e.g.,
Gal␣1– 4: globo- or Gb, Gal␣1–3: isoglobo or iGb, GlcNAcß1–3: lacto or Lc (Fig. 3). The ganglio series (or Gg)
is based on GalNAcß1– 4LacCer. The ganglio series also
comprises the simple gangliosides NeuAc1–3-LacCer, wellknown under the Svennerholm nomenclature as GM3,
GD3, and GT3, respectively (53). Glycosphingolipids in
the various series may be fucosylated or sulfated. LacCer
synthesis and all further conversions occur in the lumen
of the Golgi and require import of the necessary sugar
nucleotides (45, 183; reviewed in Ref. 209). With the use of
subcellular fractionation, a sequential distribution over
the Golgi was observed for glycosphingolipid glycosyltransferases (143, 380, 381), with a significant overlap in
distributions. A later paper (184) assigned the synthesis of
LacCer, GM3, and GM2 to the trans-Golgi and the TGN,
whereby a considerable fraction of the LacCer and GM3
synthases localized to the cis-Golgi. Pharmacological
studies, using the drugs brefeldin A (308, 333, 384, 427)
and monensin (244, 316, 317, 385) to discriminate enzymes of the Golgi stack from those in the TGN, localized
GM2 synthase (GalNAc-transferase) and two galactosyltransferases of complex glycosphingolipid synthesis to
the TGN (see, however, Ref. 426). The same conclusion
was reached in a study on mitotic cells (62).
Arthropods and mollusks transfer a mannose onto
GlcCer and further extend this chain in the arthro (At) or
the mollu (Mu) series. In some mammalian tissues, notably myelin, but in humans also the epithelia of the gastrointestinal and urogenital tracts, ceramide is mainly glycosylated to GalCer by the ceramide:galactosyltransferase,
an enzyme situated on the luminal face of the ER (328,
359, 360). The enzyme displays a preference for ceramides
containing a 2-hydroxylated fatty acid which are abundant
in these tissues. GalCer can be further galactosylated and/or
sulfated (or sialylated) in the Golgi lumen (45, 183, 373).
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
B. Subcellular Distribution and Topology
of Sphingolipids
A sphingolipid gradient exists along the organelles of
the secretory pathway. Although primarily assembled in
Physiol Rev • VOL
the Golgi, sphingolipids are enriched in the plasma membrane and endocytic membranes of cells, whereas only
low amounts are found in the ER. This has been demonstrated both in yeast (131, 280) and in animal cells (59, 90,
160, 302). Although over 90% of the cellular SM was
assigned to the plasma membrane by a cell fractionation
approach (182), some 60% of the cellular SM has been
routinely found in the plasma membrane by an assay
based on exogenous sphingomyelinase (5, 297, 334, 347,
389). Very high concentrations of glycosphingolipids (30 –
40% of total membrane lipids) have been found in the
apical plasma membrane domain of epithelial cells (see
sect. VB) and in myelin, a specialized plasma membrane
domain of Schwann cells and oligodendrocytes (366). By
electron microscopy, the complex glycosphingolipid
Forssman antigen was found absent from mitochondria
and peroxisomes, low in ER and Golgi and enriched in
plasma membrane and endocytic structures (387). Although generally enriched in the plasma membrane and
related membranes, there are differences with respect to
the distribution of individual sphingolipid classes between organelles (231) or between apical and basolateral
plasma membrane domains (159, 358). The intracellular
distribution of ceramides, sphingoid bases, and their
1-phosphates is less clear.
The accessibility of glycosphingolipids and SM to
reagents, antibodies, and enzymes on the cell surface has
led to the general belief that sphingolipids are primarily
situated in the noncytoplasmic leaflet of cellular membranes. This is in line with their site of synthesis that is on
the luminal aspect of the Golgi (for GalCer on the luminal
aspect of the ER). Only GlcCer is synthesized on the
cytosolic surface of the Golgi (see sect. VA). The best
evidence seems available for SM. In the original studies
on erythrocytes, bacterial sphingomyelinase hydrolyzed
80 – 85% of the SM (404), suggesting that the bulk of the
SM is situated in the outer, noncytoplasmic leaflet. A
similar conclusion can be drawn from the accessibility of
60% of the SM to sphingomyelinase in intact nucleated
cells (5, 297, 334, 347, 389), where a significant fraction of
the remaining 40% can be expected to be present in
intracellular membranes like endosomes and Golgi.16 Indeed, no SM was found accessible to sphingomyelinase in
microsomal membranes (263, 264), and ⱕ20% was hydrolyzed in isolated chromaffin granules (42), in which cases
the cytosolic surface is exposed. Only little quantitative
data are available on the transbilayer distribution of gly-
16
A similar SM hydrolysis of 50 –70% was observed when sphingomyelinase was applied after glutaraldehyde fixation (5) or at 15°C
(334, 389), conditions under which there is no vesicular traffic between
the plasma membrane and the endosomes/Golgi. Still, it should be
realized that the sphingomyelinase method is an invasive technique; it
modifies the membrane under study and may not be free from artifacts
(336).
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
complex cascade of glycosidases and activator proteins
or saposins in the lysosomes, and defects in any of the
steps result in lysosomal storage diseases (171). A nonlysosomal glucocerebrosidase activity has been found
(400). It appears to be active on the cytosolic surface of a
cellular membrane, possibly the plasma membrane. GlcCer is the only glycosphingolipid that is synthesized on
the cytosolic surface of the Golgi. The enzyme may be
involved in regulating GlcCer availability on cytosolic
surfaces (R. Raggers and G. van Meer, personal communication). SM is degraded by a lysosomal acid sphingomyelinase (327) in the lysosomal lumen. Alternatively, the
acid sphingomyelinase and neutral sphingomyelinases
(51, 85, 140, 378) have been invoked as the enzymes that
generate ceramide in signal transduction on the cytosolic
surface of cellular membranes (see sect. IIC). Sphingomyelinase activity has also been reported for yeast (84) and
requires ISC1, a gene whose product is structurally related to neutral sphingomyelinases (320). The ISC1-encoded protein displays phospholipase C activity toward
SM, IPC, MIPC, and M(IP)2C, but not to PI, PC, or lyso-PC,
indicating that it functions as a inositol phosphosphingolipid-specific phospholipase C in yeast (320). Where in the
cell and on which side of the membrane Isc1p-mediated
hydrolysis of sphingolipids occurs is unclear. Whether
yeast contains additional enzymes for breaking down its
sphingolipids remains to be established. Ceramides are
broken down by a number of ceramidase activities that
are classified as acidic, neutral, or alkaline, based on their
pH optimum. Acidic ceramidase is localized in lysosomes
and its gene has been cloned from human (166). Two
alkaline ceramidases associated with the ER have been
identified in yeast (222, 223).
Although the bulk of sphingoid bases and ceramides
is incorporated into sphingolipids, cells do contain small
quantities of free sphingoid bases and ceramides, including some phosphorylated derivatives; these are newly
synthesized or derived from sphingolipid breakdown and
are generally (re)utilized for sphingolipid synthesis in ER
and Golgi. Sphingosine-1-phosphate and sphinganine-1phosphate are special in the sense that they are the final
substrates in sphingolipid hydrolysis, being converted by
a lyase to ethanolamine phosphate and a C16 aldehyde
(315, 399, 433). At the same time, sphingoid long-chain
base-1-phosphates are important intra- and intercellular
second messengers (see sect. IIC), and their concentration
is tightly regulated by the combination of kinases (167,
186, 260), the lyase, and a phosphatase (149, 215, 221,
399).
1703
1704
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
C. Sphingolipid Transport and Sorting
1. Concepts
After synthesis, sphingolipids can move around the
cell in various ways. Intracellular transport processes are
fast (minutes) compared with sphingolipid turnover
(many hours). So, to maintain the differences in sphingolipid concentration between cellular membranes, there
must be specificity in sphingolipid transport. We recently
discussed sphingolipid transport in a separate review
(397). The present paper focuses on the specificity in
sphingolipid transport and the involvement of sphingolipids in sorting other membrane components.
When situated in a membrane, sphingolipids first of
all can diffuse as monomers in four directions. If we do
not take into account the motions of the entire molecule
that do not result in transport, like the rotation around
their longitudinal axis and the wobble (279), sphingolipids
can diffuse laterally in the two-dimensional plane of the
membrane; they can diffuse out of the membrane into the
aqueous phase, and they can flip across the membrane
into the opposite lipid monolayer.18 Of these movements,
17
Because the cells had been treated with acetone and/or TX-100
before antibody addition, it is possible that these lipids were originally
present on the luminal side of cytosolic vesicles or vacuoles.
18
Diffusion into the opposite leaflet of the bilayer without a
change in the longitudinal orientation of the molecule, by which the
molecule merely dips deeper into the membrane, is followed by movement back to its original position, without consequences for transport.
Only transversal diffusion plus rotation around one of its short axes, by
which the head group changes orientation, flips the molecule into the
Physiol Rev • VOL
only diffusion out of the membrane may result in transport between cellular organelles. The second mechanism
of lipid transfer in cells is by the vesicular transport
pathways that connect most cellular organelles. Finally,
lipids may be transported between organelles via transient contacts between the membranes of the two organelles.19
The word sorting is used to indicate the process by
which the cell generates the differences in protein and
lipid composition between two membranes, starting from
a membrane where these components were mixed. Because the intracellular traffic is practically a closed system for membrane components, the term generates is
equivalent to the term maintains. Where transport between membranes occurs by aqueous diffusion of a certain component as monomers, sorting requires a different
affinity of this component for the two membranes. This
may concern the affinity of the component for other membrane components, or, theoretically, the aqueous diffusion could be made unidirectional by transfer proteins.20
In vesicular transport pathways, bidirectional transport of
vesicles of random composition would result in mixing of
all components and dissipation of differences between
the two compartments. In this case sorting requires preferential inclusion of a specific component in the budding
vesicle in at least one of the two compartments. This
means lateral concentration of this component and locating the site of higher concentration to the site of vesicle
budding.
2. Monomeric transport through the cytosol
The rate of monomeric diffusion of a sphingolipid
between two membranes strongly depends on its physical
structure. The smaller the hydrophobic part, and the
larger or more polar the hydrophilic part, the higher the
rate of exchange. Sphingoid bases and their phosphory-
opposite membrane leaflet. The degree of wobble experienced by a lipid
in a certain membrane is clearly one parameter that determines the flip
rate. Other parameters are the size and polarity of the polar head group
and the dielectric constant of the membrane interior.
19
Membranes may come into close proximity, which stimulates
monomeric exchange though the aqueous phase by enhancing the desorption from the membrane (treated in detail in Ref. 39). The cytosolic
monolayers of the two organelles may transiently become continuous
(hemi-fusion), or transient fusion between the apposed membranes may
occur. The last possibility can be considered a special case of vesicular
traffic.
20
In an in vitro study in a mixture of ER and plasma membrane,
cholesterol preferentially partitioned into the plasma membrane, which
is thought to be due to its high affinity for the sphingolipids abundant in
plasma membranes (408; for a review see Ref. 203). On the other hand,
transfer proteins might pick up a lipid from one membrane and deliver
it to another membrane after being modified by, e.g., phosphorylation at
that target membrane. The activated protein would then leave the membrane empty, be inactivated at the donor membrane, and start a new
round of delivery. No experimental support for this mechanism exists.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
cosphingolipids. From studies on GalCer in myelin (199),
GM3 in membrane viruses (365, 367), and GM3 and GD1a
in a number of cells (242) it has been concluded that the
bulk of these lipids is situated in the noncytoplasmic
surface of the plasma membrane (for discussion of the
methodology, see Ref. 336).
Still, pools of sphingolipids may exist on the cytosolic surface of membranes. This is especially true for
GlcCer, which after synthesis is initially present in the
cytosolic leaflet of the Golgi membrane (65). A cytosolic
protein that can interact with GlcCer has been isolated
from cytosol and its gene cloned (198). Other cytosolic
proteins have been found to be capable of interacting with
complex glycosphingolipids (49, 54, 134, 135, 243, 352,
353), whereas also glycosphingolipids have been colocalized with cytoskeletal elements (106, 107, 318).17 This may
suggest that also complex glycosphingolipids are present
in cytosolic surfaces. This could be a consequence of lipid
mixing caused by fission and fusion events during membrane traffic, or of transbilayer equilibration of a small
sphingolipid fraction reaching the ER (see below).
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
Physiol Rev • VOL
moved from the cytosolic surface of the Golgi to the Golgi
lumen where they can be utilized for higher glycosphingolipid synthesis. This process is apparently mediated by
a non-energy-consuming translocator (45, 183). In addition, GlcCer is actively translocated from the cytosolic
surface of the plasma membrane to the outer bilayer
leaflet by the multidrug transporter MDR1 P-glycoprotein
(390; Raggers and van Meer, personal communication).
Translocation across the ER membrane may not be limited to GalCer, since this event is probably mediated by a
nonspecific translocator (46, 133). Although only a small
fraction of the sphingolipids is in the ER, these molecules
may have access to the cytosolic surface and transfer
through the aqueous phase. It is presently unclear
whether the cell contains translocators to flip them back
to the noncytoplasmic surface (295). Finally, also a small
fraction of the SM may reside in the cytosolic leaflet (see
above). Under some conditions SM may flow into the
cytosolic leaflet of the plasma membrane due to the activation of a scramblase (295, 434). It is unclear what is its
fate: transfer to other organelles via a transfer protein in
the cytosol (73), removal by a translocator like MDR1
P-glycoprotein (390), or hydrolysis by a cytosolic neutral
sphingomyelinase. In cancer cells, a remarkable increase
in SM content has been observed in microsomes (ER) and
mitochondria (20, 21, 142, 301). Because mitochondria are
thought to receive lipids from the ER by a monomeric
transfer process only, SM may have reached the mitochondria via translocation across the ER membrane followed by monomeric transport, maybe via sites of ERmitochondrial contact (see Ref. 66).
3. Vesicular traffic
The organelles along the exocytic and endocytic
routes are connected through a series of fusion and fission events. Essentially, membrane vesicles bud from one
compartment and fuse to the next, while at the same time
there is a vesicular pathway in the opposite direction. As
a variation on this theme, both at the level of the Golgi
and that of the endosomes, the complete organelle may
move forward in a maturation process, while generating
vesicles that follow the opposite direction (discussed in
detail in sect. VI). The rate of the process is sufficiently
high that, if each budding vesicle would randomly reflect
the composition of the membrane of origin, complete
mixing of all membrane components could be expected
on a time scale of hours.21 Such mixing is not observed,
and the different cellular membranes are capable of main21
Calculations on fibroblasts suggested that per hour a surface
area is endocytosed equivalent to the plasma membrane (112). An area
equivalent to five times their surface area would pass through the
endosomes per hour. Endocytic uptake of 50% of the plasma membrane
surface area per hour was measured for the apical and the basolateral
plasma membrane domain of kidney epithelial MDCKII cells (386).
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
lated derivatives can therefore be expected to readily
equilibrate between membranes with half times of seconds, whereas ceramide is on the other end of the spectrum. For SM and glycosphingolipids, the rate of spontaneous exchange is very low with half times of days, even
when they are present as monomers in a liquid-disordered
membrane (39, 375). Over the last decade, many studies
have utilized analogs of SM or GlcCer carrying a shortened radiolabeled or fluorescent fatty acyl chain. In contrast to the natural lipids, these lipids display greatly
enhanced rates of intermembrane exchange, and interpretations of results obtained with such lipids must take this
difference in properties into account (reviewed in Ref.
397). In cells, transfer of sphingolipids may be stimulated
by protein factors. One protein has been identified that in
vitro stimulates transfer of glycosphingolipids between
membranes (198), whereas proteins that can transfer SM
have been identified as well (73, 83). Whether these proteins actually function as transfer proteins in the cell
remains to be demonstrated, as is the case for the phospholipid transfer proteins first identified over 30 years ago
(388, 417).
For monomeric transport of a sphingolipid to another cellular membrane to occur, the sphingolipid must
be present on the cytosolic surface of the donor compartment. As discussed above, this is the case for sphingolipids synthesized on the cytosolic surface like the sphingoid
bases and their phosphates. Little is known concerning
their intracellular distribution and transport and concerning the site of action of the sphingoid long-chain base
phosphates in intracellular signaling. Also, ceramide is
synthesized on the cytosolic surface. Evidence has been
presented to suggest that ceramide, newly synthesized in
the ER and destined for GlcCer synthesis in the Golgi, can
be transported from the ER to the Golgi by nonvesicular
traffic (62, 97, 98, 169, 249). However, new evidence raises
the alternative interpretation that some ceramide glucosyltransferase is located in the ER, while the pathway
of ceramide transport from the ER to the Golgi site of SM
synthesis is ATP and cytosol dependent, and therefore
likely mediated by vesicles (K. Hanada, personal communication). Finally, also GlcCer is synthesized on a cytosolic surface, that of the Golgi. Indeed, GlcCer transport
from the Golgi to the plasma membrane proceeded in the
presence of brefeldin A when vesicular traffic was inhibited (406), suggesting that it may have equilibrated via the
aqueous phase. In view of its very low rate of spontaneous
exchange, such transport would most likely be mediated
by a transfer protein.
Sphingolipids could also become available on the
cytosolic surface by a transbilayer translocation event.
This has been suggested to occur for GalCer after synthesis on the luminal aspect of the ER (45). It is unclear
whether GalCer exchanges between membranes through
the aqueous phase. Both GlcCer and GalCer can be re-
1705
1706
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
22
In principle, domains of GlcCer may also exist on the cytosolic
surface of cellular membranes. In cells that synthesize GalCer, sphingolipid domains may already form in the luminal leaflet of the ER, where
GalCer is synthesized (328, 359, 360). As discussed in section IIIB, the
possibility exists that ceramide in the ER forms a domain, especially in
yeast with its long-chain hydroxylated ceramides (252). Because shortchain analogs of Cer and GalCer readily translocate across the ER
membrane (45), domains of these lipids may be situated in either membrane leaflet.
23
While of the PS in the plasma membrane, which is exclusively
present in the cytosolic leaflet, at least 75% was found to be disaturated,
Physiol Rev • VOL
additional indication that the lipid environment on the
cytosolic leaflet opposed to a sphingolipid domain in the
noncytoplasmic leaflet may be special comes from studies
on sphingolipid domains involved in signaling in the
plasma membrane (see sect. IVB). By various techniques,
the coat protein caveolin and signaling proteins that are
anchored to the cytosolic surface via saturated fatty acids
were found to colocalize with the domains in the opposite
noncytoplasmic leaflet (202, 237).
VI. SPHINGOLIPIDS AND THE CREATION
OF SELECTIVITY IN INTRACELLULAR
MEMBRANE TRANSPORT
A. The Golgi: a Central Sorting Device on the
Exocytic and Endocytic Pathways
Remarkably, sphingolipid synthesis in eukaryotic
cells primarily takes place in the Golgi, spatially separated
from glycerolipid and sterol production in the ER (see
sect. VA). To better understand the biological significance
of this arrangement, the first two parts of this section
serve to describe the elementary features of the Golgi and
to discuss how the organelle is organized to perform its
vital function in the cell.
The Golgi occupies a key position in the secretory
membrane system of cells. Arguably its most fundamental
task is to act as a filter between the glycerolipid-rich ER
and sterol/sphingolipid-rich plasma membrane. A filter
function is reflected by the organelle’s unique morphological appearance: a stack of flattened cisternae flanked by
two tubular networks that serve as the polarized entry
and exit faces (Fig. 4). The number of cisternae within the
stack can vary from 3 to ⬎20, depending on cell type
(248). The multicisternal nature of the Golgi is thought to
allow repeated opportunities for sorting out ER-based
and Golgi-resident components from constituents to be
delivered to the plasma membrane or elsewhere in the cell.
The cis-Golgi network, also called ER-Golgi intermediate compartment, forms the site where ER-derived
cargo is received in a process that involves the formation
and fusion of vesicular tubular clusters. It also provides a
major site for the budding of vesicles mediating retrograde transport of selected proteins and lipids from the
Golgi back to the ER. This retrograde flow serves to
maintain the surface area of the ER in the face of extensive membrane outflow into the secretory pathway (415),
to return escaped ER resident proteins (284), and to
recycle membrane machinery involved in ER to Golgi
in the ER at least 50% of the PS contained two unsaturated fatty acids.
For comparison, the major species of cellular PC contains a saturated
C16 or C18 at the sn-1 position and an unsaturated fatty acid at the sn-2
position of the glycerol.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
taining their unique protein and lipid compositions. For
proteins, the underlying process is clearly not ongoing
local synthesis and hydrolysis, and this is also true for
lipids. The enrichment of sphingolipids in the plasma
membrane compared with the ER is not due to synthesis
of these lipids in the plasma membrane nor to hydrolysis
of sphingolipids in the ER. As a consequence, there must
be specificity in vesicular transport; membrane components must be sorted. As defined above, this implies that
at each budding event, specific membrane components
must be included or excluded from the budding vesicle. In
the exocytic pathway, the sphingolipids are preferentially
included in membranes that travel in the forward direction
and/or excluded from transport in the retrograde direction.
Because SM and all glycosphingolipids with the exception of GlcCer are synthesized on the luminal surface
of the Golgi, sphingolipid domains would form in the
luminal leaflet of the Golgi membrane (398).22 From the
preferential interaction between cholesterol and sphingolipids (see sect. IIIB), it was then suggested that sorting of
the sphingolipids could be the driving force for sorting of
cholesterol along the exocytic pathway (395). Obviously,
the presence of cholesterol may be a prerequisite for
segregation to occur. To convert the lateral segregation
into a sorting event whereby sphingolipids and cholesterol are sorted to the plasma membrane, minimally one
of the two domains must be selectively included into a
budding vesicle. Vesicle budding along the exocytic pathway is mediated by protein coats on the cytosolic surface.
Therefore, sorting via lipid domains requires the recruitment of a specific set of coat components to one domain.
Because the coat proteins are recruited on the cytosolic
surface while the domains reside on the opposite, luminal
surface, there must be components that recognize both.
These could be proteins that span the membrane like
SNAREs (discussed in sect. VI). Alternatively, or in addition, lipid segregation may also occur in the cytosolic
surface, and a liquid-ordered domain on the luminal surface may colocalize with a similar domain on the cytosolic
surface. Evidence for such a mechanism may be found in
the fact that at least one type of lipid on the cytosolic
surface of the plasma membrane, phosphatidylserine
(PS), is highly saturated compared with the same lipid
class in the ER of rat liver (160)23 and yeast (325). One
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
1707
transport (196, 310). As cargo moves through the stack, it
is modified by Golgi-associated processing enzymes.
These enzymes, which include numerous glycosidases
and glycosyltransferases, are generally not distributed
evenly between the cisternae, but often found in the order
in which they act on their substrates (see sect. VA).24 An
advantage of this compartmental organization is that
cargo can be exposed to an ordered array of processing
steps, allowing the cell to generate highly complex glycoproteins and glycosphingolipids. Positioned at the transside of the Golgi stack is the TGN. Here, processed cargo
24
It should be noted that this intra-Golgi separation is not precise
because the enzymes are generally spread over several cisternae, displaying considerable overlap in their distributions (reviewed in Ref.
109). Therefore, different Golgi cisternae have unique mixtures of enzymes rather than different sets of enzymes. Moreover, the distribution
of a given enzyme can vary between different cell types. The basis for
this plasticity is unknown but may reflect an underlying dynamic behavior of Golgi membranes, as discussed further below.
Physiol Rev • VOL
is sorted, packaged into distinct vesicles (or larger membrane carriers), and shipped to various destinations.
These post-Golgi destinations include the cell surface
(either the apical or basolateral surface of epithelial
cells), secretory storage granules, and the various compartments of the endosomal/lysosomal system. The TGN
not only serves as a major branching point in the secretory pathway but also forms the major site where the
secretory and endocytic pathways become interconnected.
This interconnection enables cells to balance membrane
flow between the pathways and to maintain the proper
composition of their surfaces and intracellular organelles.
B. A Maturing View on Golgi Organization
A fundamental problem currently being addressed in
cell biology is how the Golgi is organized to perform its
vital tasks and how its functional integrity is maintained
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
FIG. 4. Model depicting how sphingolipid synthesis may contribute to the sorting power of the Golgi. As Golgi
cisternae mature (large gray arrow), sphingolipids are synthesized and gradually accumulate by being specifically
excluded from the tightly curved membrane in percolating COPI vesicles or tubular connections. By attracting endoplasmic reticulum (ER)-synthesized cholesterol, the flat sphingolipid-rich regions in the cisternal bilayer grow thicker
(see footnote 4). Due to their short transmembrane segments, Golgi enzymes and escaped ER-resident proteins are
excluded from the thick, sphingolipid/sterol-rich membrane regions where transport intermediates depart for the cell
surface. Instead, they will partition preferentially into earlier cisternae. Upon thickening of the bilayer, basolateral (or
lysosomal) membrane proteins adopt a conformation recognized by adaptor protein complexes (AP1/AP3; see footnote
30). AP-dependent removal of basolateral (lysosomal) material represents the terminal step in the cisternal maturation
process, leaving behind an apical membrane carrier whose content is determined by a “sorting by retention” principle
based on coaggregative properties of apical sphingolipids and proteins. Note that although the model implies a gradual
increase in membrane thickness across the Golgi stack, the shape of this thickness gradient is not known (see footnote
28). See text for further details. VTC, vesicular-tubular clusters; PGC, post-Golgi containers; PM, plasma membrane.
1708
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
25
Coat proteins are believed to have a dual function in membrane
trafficking, serving both to shape a transport vesicle and to select, by
direct or indirect interactions, the desired set of cargo molecules.
Among the cargo molecules recruited by these vesicle coats are components that specify the docking and fusion of transport vesicles with
their appropriate target organelles. These include members of the
SNARE family of membrane proteins (311). Coat recruitment, cargo
selection, and the subsequent fission and fusion of transport vesicles is
coordinated by distinct families of GTPases. Hence, coat proteins function together with targeting and fusion components to form so-called
“vesicle sorting machines” by which specific types of cargo can be
moved from one destination to the next. Cells have a variety of coat
proteins, allowing different species of vesicles to depart from various
subcellular locations.
Physiol Rev • VOL
movement of green fluorescent protein (GFP)-tagged
Golgi enzymes was followed with fluorescence video microscopy. In contrast to previous assumptions (265), it
was found that Golgi enzymes are highly mobile (61),
undergo substantial recycling within Golgi stacks (418),
and redistribute into the ER when export from this compartment is blocked (60, 368, 428).
The existence of recycling pathways both within and
from the Golgi stack together with the unsolved problem
of how COPI vesicles, the only type of vesicles known to
mediate intra-Golgi transport, would contribute to vectorial transport of secretory cargo led to a renewed interest
into an old idea, namely, that the Golgi is a collection of
maturing compartments (111).26 This “cisternal progression” or “maturation” model arose from previous observations that large cargo complexes produced by certain
cell types (e.g., algal scales, procollagen) move through
the Golgi stack without entering transport vesicles or
leaving the cisternae (25, 238). According to the model,
cisternae are continuously formed de novo by fusion of
ER-derived vesicles (314), pass through the stack as they
mature, and eventually disintegrate at the level of the
TGN. While secretory cargo stays put in cisternae, Golgi
enzymes are delivered at the appropriate time and in the
appropriate order by recycling vesicles so that each cisterna matures into the next (reviewed in Refs. 4, 110, 285).
On the basis of the currently available data, evidence
in support of cisternal maturation cannot preclude the
possibility of coexisting anterograde vesicular transport,
and vice versa. In fact, there is reason to believe that both
mechanisms operate simultaneously (286). First, procollagen aggregates appear to traverse the Golgi stack at a
much slower pace (hours) than other cargo proteins such
as VSV-G (10 –20 min; Refs. 24, 25). This finding suggests
that cisternal maturation is too slow to account for all
anterograde cargo transport. Second, Golgi proteins involved in vesicle targeting, like SNAREs, are typically
distributed over multiple cisternae across the stack (130,
272). Because there are no known components of the
targeting machinery that can specify a single cisterna, it is
hard to envision how an exclusive unidirectional movement of COPI vesicles, whether forward or backward, can
be achieved. On the basis of these notions, it has been
postulated that COPI vesicles may “percolate” up and
down the stack in a bidirectional fashion, allowing a rapid
26
A role for COPI in bidirectional vesicular transport has been
postulated based on the observation that pancreatic ␤-cells contain two
distinct populations of Golgi-associated COPI vesicles: one containing
recycling components and the other one secretory cargo (274). However, because the concentration of secretory cargo found in the second
vesicle population was comparable to that in the cisternae, it is unclear
how these vesicles would contribute to vectorial transport through the
stack. Moreover, the model leaves unresolved by what mechanism proteins bearing dilysine retrieval signals can be excluded from anterograde
COPI vesicles while being included in retrograde ones.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
despite the constant flow of membrane that enters and
leaves the organelle on both sides. Until recently, it was
widely held that the structural elements of the Golgi, the
cisternae, have a stable existence and that anterograde
movement of cargo is achieved exclusively by transport
vesicles that pinch off from one cisternae and fuse with
the next, while Golgi resident enzymes stay put in the
appropriate cisternae. This idea was largely based on
results obtained with a cell-free system in which isolated
Golgi membranes were used to measure the sequential
processing of cargo by Golgi-associated enzymes (309,
312). These studies offered key insights into the intraGolgi transport machinery. They allowed the identification of a cytoplasmic coat protein complex, termed COPI,
whose regulated assembly on Golgi membranes was
found to promote the formation of fusion-competent vesicles containing secretory cargo (270, 275).25 In the light
of these data, it was assumed that COPI-coated vesicles
mediate transport of secretory cargo through sequential
Golgi compartments in a cis-to-trans (or anterograde)
direction.
However, this view was challenged by the discovery
that COPI plays a crucial role in retrograde vesicular
transport. This notion emerged when COPI subunits were
found to bind the dilysine retrieval signal on the cytoplasmic tails of ER membrane proteins (64). Moreover, mutations in COPI blocked retrieval of these proteins from
the Golgi complex (194). Consistent with a function of
COPI in retrograde transport, quantitative immunoelectron microscopy revealed that COPI-coated tips at the
CGN are enriched for recycling membrane proteins
whereas secretory cargo prevailed at COPI-negative regions (228). The finding that COPI vesicles contain Golgi
enzymes and serve as their transport intermediates in
vitro (185, 197, 206) offered an alternative explanation for
the results previously obtained with the Golgi cell-free
transport assay; instead of moving secretory cargo forward, COPI may serve to retrieve transport machinery
and Golgi enzymes to cisternae containing cargo to be
modified. These novel data predicted a prominent role for
COPI-mediated retrograde traffic in Golgi function and
maintenance. Support for this prediction came when
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
C. Sphingolipids: Integral Parts of the Golgi-Based
Sorting Machinery?
1. Segregating forward-moving cargo from
recycling components
After assembly in the Golgi, sphingolipids accumulate in the plasma membrane (see sect. VB). Their long
and saturated fatty acyl chains, capacity for hydrogen
bonding, and association with sterols is thought to
promote the compactness and thickness, and hence the
impermeability of the cell’s outer bilayer (see sect. IIIB).
Compared with the plasma membrane, the ER contains
much lower concentrations of sphingolipids (396), despite extensive bidirectional membrane trafficking at
the ER-Golgi interface (415). Hence, a mechanism must
exist by which the Golgi prevents backflow of sphingolipids to the ER and promotes their concentration toward the trans-site where membrane carriers depart
for the plasma membrane. This could in principle be
achieved if sphingolipids were prevented from leaving
the cisternae in which they are made. The result would
be a gradual rise in their concentration as cisternae
mature. Such a scenario is in line with the recent
finding that COPI vesicles contain significantly less
sphingolipids and cholesterol than their parental Golgi
membranes (41). Indeed, a segregation of sphingolipids
and cholesterol from COPI vesicles could explain why
the interleaflet clear space of these vesicles appears
thinner than that of the cisternal membranes from
which they bud (273).
Physiol Rev • VOL
How are sphingolipids and sterols segregated from
COPI vesicles? Two distinct mechanisms can be envisaged: 1) segregation may occur because the coat machinery influences the lipid composition at the bud site.
According to this scheme, sorting could result from
preferential interactions of certain lipids with the membrane-proximal surfaces of the coat proteins and/or the
transmembrane segments of selected proteins (e.g.,
COPI coat receptors; Ref. 31). Alternatively, the COPI
coat machinery might influence the lipid composition at
the bud site by imposing a high curvature onto the
vesicle membrane. This would result in a preferential
exclusion of those lipid species that contribute to the
rigidity of flat bilayers (e.g., sphingolipids, sterols, and
saturated glycerolipids). In this respect, it is of interest
to note that COPI assembly on liposomes is negatively influenced by acyl chain saturation.27 2) Segregation may occur because the coat machinery is specifically recruited to lipid domains in the Golgi bilayer.
Sphingolipids have a strong intrinsic tendency to segregate from unsaturated glycerolipids and display a
high affinity for cholesterol (see sect. IV). Hence, sphingolipids synthesized in Golgi cisternae may trigger a
phase separation, thus creating domains enriched in
unsaturated glycerolipids that function as donor sites
for COPI vesicle biogenesis versus sphingolipid- and
sterol-enriched domains that do not. There is evidence
that such lipid subdomains exist in early Golgi cisternae (108).
Of the two lipid-sorting mechanisms described
above, the latter has the particular appeal that it is inherently self-organizing, thus escaping the requirement for
proteins that to trigger a lateral segregation of lipids need
themselves to be retained in place. It is tempting to speculate that ongoing sphingolipid synthesis in the Golgi acts
as a sink for ER-synthesized cholesterol and that a phase
separation is nucleated once these lipids have reached a
critical concentration in the cisternal bilayer (see sect. IV).
Regardless of the actual mechanism of lipid segregation, the exclusion of sphingolipids and sterols from COPI
vesicles would cause their progressive accumulation in
the maturing cisternae. As a result, anterograde transport
of sphingolipids and cholesterol would be coupled to a
gradual depletion of unsaturated glycerolipids along the
secretory pathway. A sphingolipid/sterol concentration
gradient would form across the Golgi stack (26, 56, 57,
271), causing an increase in bilayer thickness toward the
27
In a chemically defined in vitro assay, the assembly of COPI
coats on synthetic liposomes was found to be strongly affected by acyl
chain saturation; diC18:1 PC/diC18:1 PE vesicles, which contain unsaturated fatty acyl chains, allowed coat recruitment, whereas C16:0-C18:1
PC/C16:0-C18:1 PE vesicles, carrying one saturated and one unsaturated
fatty acid, did not (355).
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
equilibration of plasma membrane and soluble secretory
cargo between adjacent cisternae (272, 286). Percolating
vesicles (or transient tubular connections; Ref. 410)
would give rise to a “fast track” by which cargo can flow
across the entire Golgi stack. Larger cargo structures and
other molecules that cannot enter COPI vesicles (or tubules) would then transit the stack at the slower pace of
cisternal maturation.
In view of these coexisting modes of transport, how
does the Golgi ensure efficient vectorial movement of cell
surface-destined proteins and lipids, while maintaining at
the same time an asymmetric arrangement of its processing enzymes? Without a mechanism that regulates the
partitioning of cargo molecules and Golgi residents into
COPI vesicles, a rapid interchange would result in complete uniformity and undermine directionality of transport. In the next section, we discuss how the assembly of
sphingolipids in Golgi cisternae and their impact on the
lateral organization of other membrane molecules may
provide a physical basis exploited by the Golgi to segregate forward-moving cargo from recycling components,
and once this is accomplished, to organize specialized
TGN export sites used for polarized secretion.
1709
1710
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
28
The shape of the thickness gradient is not known. In one extreme model, the phase separation between sphingolipids/cholesterol
and unsaturated glycerophospholipids gives rise to a thick domain that
is transported to the plasma membrane and a thin domain that is
transported back to the ER. However, it is clear from the literature that
glycerolipid membranes become thicker upon addition of cholesterol
(see footnote 4). Also, it is clear that sphingolipid-rich membranes of
different thickness can leave the TGN (like the apical and basolateral
transport vesicles in epithelial cells). We therefore hypothesize that the
membrane thickness increases gradually across the Golgi stack. Reasons for this may be the presence in cis-Golgi cisternae of sphingolipids
at a concentration that is insufficient to cause phase separation, a
gradient in sphingolipid synthesis, and a cholesterol gradient. However,
a proper understanding of this phenomenon will require a quantitative
analysis of these parameters.
29
Previous studies had revealed that the steady-state localization
of intra-Golgi processing enzymes is largely determined not by coatbinding motifs but by their TMDs. In all cases examined sofar, the TMDs
of glycosyltransferases are sufficient to direct localization of hybrid
proteins to the Golgi (reviewed in Ref. 109). These TMDs lack obvious
sequence motifs that could serve as sorting determinants, and attempts
to identify such motifs by mutagenesis have failed.
30
It is important to note that the extent to which lipids in a bilayer
can distort to match the length of a transmembrane ␣-helix in a protein
is limited. Although a short transmembrane ␣-helix can incorporate into
a thin bilayer, it cannot incorporate into a bilayer that is too thick, and
instead will form aggregates (164, 190). A long transmembrane ␣-helix,
on the other hand, can incorporate into either a thick or a thin bilayer,
in the latter case by tilting. Tilting of a transmembrane protein may alter
its conformation and function. The effect of bilayer thickness on protein
function will probably be stronger in the case of proteins containing a
bundle of multiple transmembrane helices (e.g., an ion channel) as a
sliding of helices within a bundle may be neccessary to achieve the most
optimal hydrophobic match. Interestingly, a coupling between hydrophobic membrane thickness and protein activity has been observed for
several transporter proteins, including ion pumps (63, 362).
Physiol Rev • VOL
bypassed by genetic mutations in enzymes responsible for
the elongation of the long-chain fatty acids found in ceramide and sphingolipids (67).31
Apart from being short, some Golgi TMDs display an
unusual high content of the bulky residue phenylalanine, a
feature that may well promote a segregation from the ordered sphingolipid/sterol-rich membrane regions (32). Oligomerization into multi-enzyme complexes (265) provides
another factor that could influence phase behavior of Golgi
enzymes. Collectively, the variation in such properties would
offer considerable scope for establishing different steadystate distributions of enzymes within the Golgi stack. Bidirectional transport mediated by percolating COPI vesicles or
transient tubular connections would allow an enzyme to
transiently explore the entire stack until it finds a cisterna
whose lipid composition suits its TMD (283). Ongoing sphingolipid synthesis and depletion of unsaturated glycerolipids
will gradually remodel the lipid composition of the cisternal
bilayer to that of the plasma membrane, and eventually drive
all Golgi-resident proteins, including the sphingolipid-synthesizing enzymes themselves, into recycling COPI vesicles.
Likewise, plasma membrane proteins whose rapid
diffusion through the stack is mediated by percolating
vesicles or tubules could achieve the desired directionality in transport if their membrane anchors would favor
cisternal bilayers that have become thicker and more
organized due to a progressive accumulation of sphingolipids and sterols. Indeed, GPI-anchored proteins have a
high affinity for glycosphingolipid/cholesterol-rich domains (see sect. IVB), a feature thought to be required for
their efficient transport from the Golgi to the cell surface
(233, 345). The same has been reported for influenza virus
hemagglutinin (see sect. IVB; Ref. 161).
The notions that sorting in the Golgi depends on transitions in bilayer thickness and that these transitions are the
result of sphingolipid-induced phase separations provide a
compelling explanation for why structural features render
sphingolipids essential for the viabilty of cells; they form an
integral part of the mechanisms by which cells generate the
compositional and functional differences between their
plasma membrane and internal organelles.
2. Organizing specialized export sites
for polarized secretion
The TGN is a major branching point in secretory
traffic where apical and basolateral cell surface compo-
31
Membrane fusion reconstitution experiments have revealed that
v-SNAREs active in the lysosomal system of yeast can replace the
exocytic v-SNAREs in a reaction that mimics fusion of secretory vesicles
with the plasma membrane (236). With respect to the observed redundancy of exocytic v-SNAREs in a fatty acid elongation mutant (67), it is
tempting to speculate that a shortening of sphingolipid-associated fatty
acids would cause lysosomal v-SNAREs to leak into secretory vesicles
and take over the function of the exocytic v-SNAREs.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
TGN (114; Fig. 4).28 Preferential interactions between lipids
and transmembrane proteins with the best matching hydrophobic length has been recognized as a potential mechanism
for protein sorting in the Golgi (32, 255). This idea emerged
with the observation that the transmembrane domains
(TMDs) of Golgi-resident proteins are on average five residues shorter than those of plasma membrane proteins
(32).29 This phenomenon is well conserved among eukaryotic cells, from mammals (32) to yeast (195). It has been
postulated that due to their shorter TMDs, Golgi-resident
proteins would be excluded from the thicker sphingolipid/
sterol-enriched membrane regions destined for the cell surface and hence be retained in the Golgi (32, 255).30 In support of this idea, lengthening the TMDs of several Golgi
enzymes results in their movement to the plasma membrane
(229, 253, 254), while shortening the TMD of a plasma membrane protein causes its accumulation in the Golgi (60). In
addition, the ER residency of several membrane proteins
can be disrupted by elongating their TMDs (282, 422). TMD
length has also been recognized as a critical parameter in the
sorting of SNAREs (298), membrane proteins whose specific
interactions contribute significantly to the specificity of vesicle docking and fusion (236, 311). Particularly intruiging in
this respect is the finding that the requirement of exocytic
v-SNAREs (vesicular SNAREs) for secretion in yeast can be
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
32
It should be noted that a fraction of the apical proteins are
transported to the apical surface via the basolateral surface in some
cells (187, 230), but not in other cells (413). Interestingly, the polarity of
apical and basolateral proteins may alter with the physiological state of
a cell (2). In retinal pigmented epithelial cells, developmental switches
control apical and basolateral trafficking of a number of proteins (225).
Physiol Rev • VOL
culminating in an apical domain and a basolateral domain
at the TGN.
It is believed that apical protein sorting relies on the
formation of glycosphingolipid and cholesterol-enriched
domains in the luminal leaflet of the TGN (341, 398). This
idea is supported by the following observations: 1) GPI
proteins use their lipid anchors as apical sorting determinants (34, 200), while sorting information in at least two
apically expressed transmembrane proteins (influenza
hemagglutinin and neuraminidase) resides in their transmembrane segments (175, 322); 2) influenza hemagglutinin and a GPI protein acquire detergent resistance in the
Golgi during biosynthetic transport (38, 86); and 3) inhibition of sphingolipid synthesis or deprivation of cholesterol causes a reduction in detergent insolubility and randomizes the cell surface distribution of influenza
hemagglutinin and a GPI protein without affecting sorting
of basolateral proteins (161, 233). A lipid-based sorting
mechanism similar to that reported for the apical delivery
of proteins in epithelial cells has been suggested to mediate axonal delivery of proteins in neurons (80, 188, 189).
In addition, it appears that carbohydrates in glycoproteins also provide apical sorting information (88). Several secretory and membrane proteins have been reported
to rely on their N- or O-glycans for apical delivery (3, 115,
321), including GPI-anchored proteins (19). The mechanism of carbohydrate-mediated apical sorting is unclear
(see Ref. 306). One potential mechanism involves recognition of N-glycans by sorting lectins such as VIP36 (87,
420), which is thought to partition into glycosphingolipid/
cholesterol-rich domains in the TGN along with other
apical membrane proteins. In summary, protein sorting to
the apical membrane is variable and largely depends on
targeting signals present in the luminal domain or membrane anchor.
In contrast, basolateral sorting is typically mediated
by discrete targeting signals that are confined to the cytoplasmic tails of transmembrane proteins. These signals
often consist of tyrosine- or dileucine-based amino acid
motifs and resemble the signals that mediate endocytosis
and protein sorting to endosomes and lysosomes. It has
been well established that interactions between tyrosineand leucine-based sorting signals and AP coats are responsible for selective cargo uptake into TGN- and
plasma membrane-derived clathrin-coated vesicles (239).
Basolateral sorting is most likely based on a similar coatmediated mechanism. An epithelial cell type-specific isoform of the AP-1 coat complex, designated AP-1B, has
recently been identified and found to play a critical role in
the basolateral targeting of a number of membrane proteins (91). Still, AP-1B-independent mechanisms for basolateral sorting must exist as epithelial cells are capable of
correctly delivering some basolateral membrane proteins
in the absence of AP-1B (313). Moreover, AP-1B is not
expressed in hepatocytes and neurons, yet these cell
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
nents are segregated and delivered via separate transport
routes. Although previously viewed as a special feature of
epithelial cells, it has become clear that apical and basolateral delivery pathways operate in both polarized and
nonpolarized cell types (257, 425). However, only in epithelial cells, they would result in macroscopic domains
laterally separated by a diffusion barrier at the tight junction. Recent studies have indicated that the establishment
of cell polarity is initiated by signaling pathways that are
responsive to cell-cell and cell-matrix contacts (423). This
would allow a molecular definition of contacting and
noncontacting cell surfaces (the precursors of the basolateral and apical membrane domains, respectively). The
next step would then involve a localized assembly of
cytoskeletal elements and a positioning of the TGN relative to the spatial cues. Finally, the cell surface polarity
that was initially defined by the spatial cues would be
reinforced by targeted delivery of apical and basolateral
cell surface components that are constitutively sorted in
the TGN.32
As discussed in the previous section, the enrichment
of sphingolipids and saturated glycerolipids along the
secretory pathway could be explained if COPI vesicles
budding from the Golgi cisternae would select thinner,
unsaturated lipid-enriched membrane regions. Still an additional level of lipid sorting must be invoked to explain
the difference in lipid composition between the apical and
basolateral surface of epithelial cells. The basolateral cell
surface has a composition very similar to that of the
plasma membrane of nonpolarized cells, but the apical
surface is highly enriched in sphingolipids, either glycosphingolipids or SM (see sect. IVB; Ref. 341). It therefore
appears that superimposed on the lateral segregation of
sphingolipids from unsaturated glycerolipids throughout
the Golgi stack, a second lateral segregation in the TGN
membrane would occur. This would give rise to domains
highly enriched in glycosphingolipids or SM and with
targeting information for the apical surface, and domains
with a more moderate enrichment in sphingolipids that
would contain basolateral information (395). The segregation of apical from basolateral lipids could imply the
existence of three phases in the TGN membrane. However, a simpler mechanism would be extension to the
TGN of the segregation of the two fluid lipid phases
occurring in the Golgi stack. Depletion of unsaturated
lipids and ongoing sphingolipid synthesis would result in
a gradual change in the composition of both subdomains,
1711
1712
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
3. Apical sorting by retention?
In contrast to the high-affinity protein-protein interactions involved in the formation of basolateral vesicles,
apical sorting seems largely based on a cooperativity of
weak interactions between lipids, proteins, and sugars
that occur predominantly within the bilayer domain or
lumen of the TGN. Therefore, a role for vesicular coats in
the formation of apical transport carriers has remained
rather elusive. The production of apical transport vesicles
has been proposed to involve VIP21/caveolin, a membrane protein with high affinity for cholesterol (256). It
was suggested that oligomerization and intercalation of
VIP21/caveolin into the glycosphingolipid/cholesterol-enriched domains would cause the bilayer to bend into
apical transport vesicles (277). However, evidence in support of this hypothesis remains indirect. An alternative
possibility is that the inclusion of basolateral molecules
into coated vesicles budding from the TGN provides the
primary means by which apical and basolateral components are physically segregated into different membrane
carriers. Interestingly, correlative light and electron microscopic studies have indicated that the membrane carriers involved in transport from the Golgi to the plasma
membrane are not only small vesicles, but also larger
tubulovesicular structures (136, 379), sometimes even as
large as half a Golgi cisternea (288). It appears that these
structures arise by a controlled drawing out and progressive breakdown of TGN material from the Golgi complex
according to a mechanism that bears similarities with the
formation and maturation of secretory storage granules in
neuroendocrine cells (10, 288). An intruiging possibility
would be that apical transport intermediates are formed
by means of a progressive maturation of TGN membranes
Physiol Rev • VOL
via AP-dependent removal of endosomal/lysosomal and
basolateral material (Fig. 4). Such a mechanism would
escape the requirement of a specific apical coat complex
as it would represent a terminal step in the cisternal
maturation process. Analogous to the mechanism of
cargo sorting in secretory storage granules (374), sorting
of apical cargo may rely on a “sorting by retention” principle, based on the coaggregative properties of the apical
proteins and membrane lipids. To become a functional
transport intermediate, a maturing apical container must
recruit components of the docking and fusion machinery.
Specific members of the SNARE family (TI-VAMP and
syntaxin 3) have been identified on apical membrane
carriers and were shown to play a critical role in membrane trafficking from the TGN to the apical cell surface
(180). The MAL/VIP17 proteolipid may represent another
part of the apical transport machinery (227, 291), but its
mode of action remains to be clarified. All these components occurred in detergent-resistent membrane fractions
along with protein cargo destined for the apical cell surface, suggesting that both apical cargo and transport machinery are subject to the same lipid-based sorting mechanism.
VII. CONCLUDING REMARKS
A. Sphingolipids and Golgi Maturation
Above, we have worked out the concept that a separation between two fluid phases of lipids lies at the heart
of the sorting potential of the Golgi complex. This phase
separation would be based on the differential miscibility
of sphingolipids, glycerolipids, and sterols. An essential
feature of the model is that the Golgi membrane matures
along the cis-trans axis, based on cisternal progression, a
gradual change in sphingolipid composition by biosynthetic reactions, and continuous retrieval of the most fluid
lipids. As a consequence, the Golgi membrane as a whole
becomes less fluid and thickens, resulting in a sequential
loss of Golgi-resident enzymes and transport machinery
to recycling pathways that run from the cis-Golgi to the
ER and from within the Golgi stack to more proximal
cisternae. After distillation, a membrane persists that is
highly enriched in sphingolipids, cholesterol, and selected
proteins and forms the prototype “apical” membrane carrier. The ordered array of sphingolipid synthetic enzymes
that extends from the ER to the trans-Golgi seems an
ideal device to gradually convert the highly fluid and
flexible ER bilayer into a rigid and robust plasma membrane. In fact, it may form an integral part of the mechanistic framework by which the compartmental organization of the secretory pathway is established.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
types rely on basolateral-type targeting signals for transport to their sinusoidal and somatodendritic surfaces,
respectively (416, 424). A novel, ubiquitously expressed
adaptor protein complex, AP-4, localizes to the TGN (70)
and may be involved in basolateral sorting as well, but
this remains to be shown.
Interestingly, there appears to be a hierarchical readout of basolateral and apical sorting signals. Basolateral
sorting signals are usually dominant; influenza hemagglutinin equipped with a tyrosine-based sorting signal is converted from an apical to a basolateral protein (33). On the
other hand, inactivation of basolateral signals does not
always lead to apical delivery, even if the protein is glycosylated. It has been suggested that some basolateral
membrane proteins are sorted by exclusion from glycosphingolipid rafts; such a scenario may apply to a number
of proteins whose efficient basolateral distribution is
achieved in the absence of an active basolateral sorting
signal (233, 423).
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
B. Domains of Saturated Lipids on the
Cytosolic Surface
C. Sphingolipid Domains as a General Theme
in Cellular Membrane Traffic
Lipid segregation does not necessarily start in the
cis-Golgi. As discussed in section IIIB, the possibility exists that ceramides in the ER, especially in yeast with its
long-chain hydroxylated ceramides, occur in domains
(252). In cells that synthesize GalCer, this sphingolipid
may organize in domains already in the luminal leaflet of
the ER, where it is synthesized (328, 359, 360). Short-chain
analogs of Cer and GalCer (45) and other (sphingo)lipids
(133) readily translocate across the ER membrane. If true
for natural lipids, domains of these lipids may be situated
in either membrane leaflet.
Physiol Rev • VOL
Moreover, the ultimate formation of an “apical”
sphingolipid/cholesterol remnant in the TGN is a final
stage in Golgi maturation but clearly is not a final stage in
cellular membrane transport. Membrane recycles from
the plasma membrane via endosomes to the TGN, and
evidence for lipid-based sorting has been found in the
endocytic recycling pathway (reviewed in Ref. 397). Although, unfortunately, essentially all evidence has been
obtained using fluorescent analogs of membrane lipids,
still it leads to the conclusion that certain lipidic molecules can be sorted in this pathway. In short, fluorescent
sphingolipids are concentrated during the first step(s) of
endocytosis (52, 407). In some cells, fluorescent GlcCer
segregates from fluorescent SM during recycling from the
plasma membrane (168, 393, 394). However, such sorting
was not observed in different cell types (386), and it
cannot be assumed that these lipid analogs strictly monitor vesicular transport processes (226). Independent evidence for lipid sorting via phase separation was obtained
by studying transport of lipid probes of different chain
length and saturation (251). The data in the latter study
suggest that lipids with longer or more saturated chains
are preferentially transported to late endosomes versus
the recycling endosome. Furthermore, cholesterol had a
dramatic effect on endocytic lipid sorting (293). Because
endosome maturation has many parallels with the maturation of Golgi cisternae, a sphingolipid-driven sorting
device may also be central to the endocytic pathway.
Because sphingolipid synthesis is limited to the Golgi (see
sect. VA; Ref. 391), and all components are present at the
start of the endosomal maturation process, in the plasma
membrane, the endosomal lipid maturation can be considered to be an extension of the Golgi system.
We thank the other members of our lipid cell biology group
for many useful discussions. In addition, we are grateful to Rick
Brown, Ken Hanada, Richard Pagano, Rob Parton, Rick Proia,
Peter Slotte, Pentti Somerharju, Paul van Veldhoven, and Rob
Zwaal for discussing distinct issues in the review and for sending preprints of unpublished work.
We thank the following organizations for financial support:
the Royal Netherlands Academy of Arts and Sciences (to J. C. M.
Holthuis), the Netherlands Foundations for Chemical Research
and for Life Sciences with financial aid from the Netherlands
Organization for Scientific Research (to R. J. Raggers and H.
Sprong), and the European Molecular Biology Organization (to
T. Pomorski).
Address for reprint requests and other correspondence: G.
van Meer, Dept. of Cell Biology and Histology, Academic Medical Center L3, Meibergdreef 15, 1105 AZ Amsterdam, The Netherlands (E-mail: [email protected]).
REFERENCES
1. AHMED SN, BROWN DA, AND LONDON E. On the origin of sphingolipid/
cholesterol-rich detergent-insoluble cell membranes: physiological
concentrations of cholesterol and sphingolipid induce formation of
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
Because SM and all glycosphingolipids with the exception of GlcCer are synthesized on the luminal surface
of the Golgi, sphingolipid domains would form in the
luminal leaflet of the Golgi membrane (398). We have
recently obtained evidence that the pool of GlcCer in the
cytosolic leaflet is regulated by the combination of MDR1
P-glycoprotein, a GlcCer translocator, and a nonlysosomal glucoceramidase (390). In addition, our studies show
that glycosphingolipid synthesis is required for an intact
AP-3 pathway to melanosomes (H. Sprong, P. van der
Sluijs, and G. van Meer, personal communication). The
data suggest that GlcCer on the cytosolic surface is involved in the recruitment of specific coat proteins at the
level of the TGN. This could be by a direct interaction of
GlcCer with a cytosolic protein (complex). Alternatively,
GlcCer could form (or be part of) a domain on the cytosolic surface.
Evidence for lipid phase separation in the cytosolic
bilayer leaflet of Golgi membranes may be found in the
fact that at least one type of lipid on the cytosolic surface
of the plasma membrane is highly saturated compared
with the same lipid class in the ER of rat liver (160) and
yeast (325). Such a domain would mature along with the
luminal sphingolipid/cholesterol domain. One additional
indication that the lipid environment on the cytosolic
leaflet opposed to a sphingolipid domain in the noncytoplasmic leaflet may be special comes from studies on
sphingolipid domains involved in signaling in the plasma
membrane (see sect. IVB). By various techniques, the coat
protein caveolin and signaling proteins that are anchored
to the cytosolic surface via saturated fatty acids colocalize with the domains in the opposite noncytoplasmic leaflet (202, 237). Whether the opposed domains would be
connected via transmembrane lipid-lipid interactions (see
sect. IIIB) or via membrane-spanning proteins is an open
issue.
1713
1714
2.
3.
4.
5.
6.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
a detergent-insoluble, liquid-ordered lipid phase in model membranes. Biochemistry 36: 10944 –10953, 1997.
AL-AWQATI Q. Plasticity in epithelial polarity of renal intercalated
cells: targeting of the H⫹-ATPase and band 3. Am J Physiol Cell
Physiol 270: C1571–C1580, 1996.
ALFALAH M, JACOB R, PREUSS U, ZIMMER KP, NAIM H, AND NAIM HY.
O-linked glycans mediate apical sorting of human intestinal sucrase-isomaltase through association with lipid rafts. Curr Biol 9:
593–596, 1999.
ALLAN BB AND BALCH WE. Protein sorting by directed maturation of
Golgi compartments. Science 285: 63– 66, 1999.
ALLAN D AND QUINN P. Resynthesis of sphingomyelin from plasmamembrane phosphatidylcholine in BHK cells treated with Staphylococcus aureus sphingomyelinase. Biochem J 254: 765–771, 1988.
AN S, ZHENG Y, AND BLEU T. Sphingosine 1-phosphate-induced cell
proliferation, survival, and related signaling events mediated by G
protein-coupled receptors edg3 and edg5. J Biol Chem 275: 288 –
296, 2000.
ANDERSON RGW. The caveolae membrane system. Annu Rev Biochem 67: 199 –225, 1998.
NGSTRÖM J, BREIMER ME, FALK KE, GRIPH I, HANSSON GC, KARLSSON
KA, AND LEFFLER H. Separation and characterization of hematosides
with different sialic acids and ceramides from rat small intestine.
Different composition of epithelial cells versus non-epithelial tissue and of duodenum versus jejunum-ileum. J Biochem 90: 909 –
921, 1981.
ANTES P, SCHWARZMANN G, AND SANDHOFF K. Detection of protein
mediated glycosphingolipid clustering by the use of resonance
energy transfer between fluorescent labelled lipids. A method established by applying the system ganglioside GM1 and cholera
toxin B subunit. Chem Phys Lipids 62: 269 –280, 1992.
ARVAN P AND CASTLE D. Sorting and storage during secretory granule
biogenesis: looking backward and looking forward. Biochem J 332:
593– 610, 1998.
BAGNAT M, KERANEN S, SHEVCHENKO A, AND SIMONS K. Lipid rafts
function in biosynthetic delivery of proteins to the cell surface in
yeast. Proc Natl Acad Sci USA 97: 3254 –3259, 2000.
BARENHOLZ Y AND GATT S. Sphingomyelin: metabolism, chemical
synthesis, chemical and physical properties. In: Phospholipids.
New Comprehensive Biochemistry, edited by Hawthorne JN and
Ansell GB. Amsterdam: Elsevier, 1982, vol. 4, p. 129 –177.
BASU S AND KOLESNICK R. Stress signals for apoptosis: ceramide and
c-Jun kinase. Oncogene 17: 3277–3285, 1998.
BEARER EL AND FRIEND DS. Morphology of mammalian sperm membranes during differentiation, maturation, and capacitation. J Electron Microsc Tech 16: 281–297, 1990.
BEELER T, BACIKOVA D, GABLE K, HOPKINS L, JOHNSON C, SLIFE H, AND
DUNN T. The Saccharomyces cerevisiae TSC10/YBR265w gene
encoding 3-ketosphinganine reductase is identified in a screen for
temperature-sensitive suppressors of the Ca2⫹-sensitive csg2⌬ mutant. J Biol Chem 273: 30688 –30694, 1998.
BEELER TJ, FU D, RIVERA J, MONAGHAN E, GABLE K, AND DUNN TM.
SUR1 (CSG1/BCL21), a gene necessary for growth of Saccharomyces cerevisiae in the presence of high Ca2⫹ concentrations at
37°C, is required for mannosylation of inositolphosphorylceramide.
Mol Gen Genet 255: 570 –579, 1997.
BEHNE M, UCHIDA Y, SEKI T, DE MONTELLANO PO, ELIAS PM, AND
HOLLERAN WM. Omega-hydroxyceramides are required for corneocyte lipid envelope (CLE) formation and normal epidermal permeability barrier function. J Invest Dermatol 114: 185–192, 2000.
BEN-PORAT T AND KAPLAN AS. Phospholipid metabolism of herpesvirus-infected and uninfected rabbit kidney cells. Virology 45: 252–
264, 1971.
BENTING JH, RIETVELD AG, AND SIMONS K. N-glycans mediate the
apical sorting of a GPI-anchored, raft-associated protein in MadinDarby canine kidney cells. J Cell Biol 146: 313–320, 1999.
BERGELSON LD, DYATLOVITSKAYA EV, SOROKINA IB, AND GORKOVA NP.
Phospholipid compositon of mitochondria and microsomes from
regenerating rat liver and hepatomas of different growth rate.
Biochim Biophys Acta 360: 361–365, 1974.
BERGELSON LD, DYATLOVITSKAYA EV, TORKHOVSKAYA TI, SOROKINA IB,
AND GORKOVA NP. Phospholipid composition of membranes in the
tumor cell. Biochim Biophys Acta 210: 287–298, 1970.
Physiol Rev • VOL
22. BEUTLER E. Enzyme replacement therapy for Gaucher’s disease.
Baillieres Clin Haematol 10: 751–763, 1997.
23. BOGGS JM. Lipid intermolecular hydrogen bonding: influence on
structural organization and membrane function. Biochim Biophys
Acta 906: 353– 404, 1987.
24. BONFANTI L, MARTELLA O, MIRONOV A, AND LUINI A. Comparative
analysis of procollagen I and VSV G protein transport in the same
cell (Abstract). Mol Biol Cell Suppl 10: 114a, 1999.
25. BONFANTI L, MIRONOV AA, MARTINEZ-MENARGUEZ JA, MARTELLA O,
FUSELLA A, BALDASSARRE M, BUCCIONE R, GEUZE HJ, AND LUINI A.
Procollagen traverses the Golgi stack without leaving the lumen of
cisternae: evidence for cisternal maturation. Cell 95: 993–1003,
1998.
26. BORGESE N AND MELDOLESI J. Localization and biosynthesis of
NADH-cytochrome b5 reductase, an integral membrane protein, in
rat liver cells. I. Distribution of the enzyme activity in microsomes,
mitochondria, and Golgi complex. J Cell Biol 85: 501–515, 1980.
27. BOSIO A, BINCZEK E, AND STOFFEL W. Functional breakdown of the
lipid bilayer of the myelin membrane in central and peripheral
nervous system by disrupted galactocerebroside synthesis. Proc
Natl Acad Sci USA 93: 13280 –13285, 1996.
28. BOUHOURS JF AND GLICKMAN RM. Rat intestinal glycolipids. III. Fatty
acids and long chain bases of glycolipids from villus and crypt cells.
Biochim Biophys Acta 487: 51– 60, 1977.
29. BRADY RO. Gaucher’s disease: past, present and future. Baillieres
Clin Haematol 10: 621– 634, 1997.
30. BRASITUS TA AND SCHACHTER D. Lipid dynamics and lipid-protein
interactions in the rat enterocyte basolateral and microvillus membranes. Biochemistry 19: 2763–2769, 1980.
31. BREMSER M, NICKEL W, SCHWEIKERT M, RAVAZZOLA M, AMHERDT M,
HUGHES CA, SÖLLNER TH, ROTHMAN JE, AND WIELAND FT. Coupling of
coat assembly and vesicle budding to packaging of putative cargo
receptors. Cell 96: 495–506, 1999.
32. BRETSCHER MS AND MUNRO S. Cholesterol and the Golgi apparatus.
Science 261: 1280 –1281, 1993.
33. BREWER CB AND ROTH MG. A single amino acid change in the
cytoplasmic domain alters the polarized delivery of influenza virus
hemagglutinin. J Cell Biol 114: 413– 421, 1991.
34. BROWN DA, CRISE B, AND ROSE JK. Mechanism of membrane anchoring affects polarized expression of two proteins in MDCK cells.
Science 245: 1499 –1501, 1989.
35. BROWN DA AND LONDON E. Functions of lipid rafts in biological
membranes. Annu Rev Cell Dev Biol 14: 111–136, 1998.
36. BROWN DA AND LONDON E. Structure and origin of ordered lipid
domains in biological membranes. J Membr Biol 164: 103–114,
1998.
37. BROWN DA AND LONDON E. Structure and function of sphingolipidand cholesterol-rich membrane rafts. J Biol Chem 275: 17221–
17224, 2000.
38. BROWN DA AND ROSE JK. Sorting of GPI-anchored proteins to glycolipid-enriched membrane subdomains during transport to the
apical cell surface. Cell 68: 533–544, 1992.
39. BROWN RE. Spontaneous lipid transfer between organized lipid
assemblies. Biochim Biophys Acta 1113: 375–389, 1992.
40. BROWN RE. Sphingolipid organization in biomembranes: what physical studies of model membranes reveal. J Cell Sci 111: 1–9, 1998.
41. BRÜGGER B, SANDHOFF R, WEGEHINGEL S, GORGAS K, MALSAM J, HELMS
JB, LEHMANN WD, NICKEL W, AND WIELAND FT. Evidence for segregation of sphingomyelin and cholesterol during formation of COPIcoated vesicles. J Cell Biol 151: 507–518, 2000.
42. BUCKLAND RM, RADDA GK, AND SHENNAN CD. Accessibility of phospholipids in the chromaffin granule membrane. Biochim Biophys
Acta 513: 321–337, 1978.
43. BUEDE R, RINKER-SCHAFFER C, PINTO WJ, LESTER RL, AND DICKSON RC.
Cloning and characterization of LCB1, a Saccharomyces gene required for biosynthesis of the long-chain base component of sphingolipids. J Bacteriol 173: 4325– 4332, 1991.
44. BUNOW MR AND LEVIN IW. Molecular conformations of cerebrosides
in bilayers determined by Raman spectroscopy. Biophys J 32:
1007–1021, 1980.
45. BURGER KNJ, VAN DER BIJL P, AND VAN MEER G. Topology of sphingolipid galactosyltransferases in ER and Golgi: transbilayer move-
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
7.
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
46.
47.
48.
49.
50.
51.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
Physiol Rev • VOL
69. DE KRUIJFF B (Editor). Lipid-protein interactions. Biochim Biophys
Acta 1376: 243– 482, 1999.
70. DELL’ANGELICA EC, MULLINS C, AND BONIFACINO JS. AP-4, a novel
protein complex related to clathrin adaptors. J Biol Chem 274:
7278 –7285, 1999.
71. DEMEL RA AND DE KRUYFF B. The function of sterols in membranes.
Biochim Biophys Acta 457: 109 –132, 1976.
72. DEMEL RA, PALTAUF F, AND HAUSER H. Monolayer characteristics and
thermal behavior of natural and synthetic phosphatidylserines.
Biochemistry 26: 8659 – 8665, 1987.
73. DE VRIES KJ, HEINRICHS AAJ, CUNNINGHAM E, BRUNINK F, WESTERMAN
J, SOMERHARJU PJ, COCKCROFT S, WIRTZ KWA, AND SNOEK GT. An
isoform of the phosphatidylinositol-transfer protein transfers
sphingomyelin and is associated with the Golgi system. Biochem J
310: 643– 649, 1995.
74. DICKSON RC. Sphingolipid functions in Saccharomyces cerevisiae:
comparison to mammals. Annu Rev Biochem 67: 27– 48, 1998.
75. DICKSON RC AND LESTER RL. Metabolism and selected functions of
sphingolipids in the yeast Saccharomyces cerevisiae. Biochim Biophys Acta 1438: 305–321, 1999.
76. DICKSON RC, NAGIEC EE, SKRZYPEK M, TILLMAN P, WELLS GB, AND
LESTER RL. Sphingolipids are potential heat stress signals in Saccharomyces. J Biol Chem 272: 30196 –30200, 1997.
77. DICKSON RC, NAGIEC EE, WELLS GB, NAGIEC MM, AND LESTER RL.
Synthesis of mannose-(inositol-P)2-ceramide, the major sphingolipid in Saccharomyces cerevisiae, requires the IPT1 (YDR072c)
gene. J Biol Chem 272: 29620 –29625, 1997.
78. DIETZEN DJ, HASTINGS WR, AND LUBLIN DM. Caveolin is palmitoylated
on multiple cysteine residues. Palmitoylation is not necessary for
localization of caveolin to caveolae. J Biol Chem 270: 6838 – 6842,
1995.
79. DOERING T, PROIA RL, AND SANDHOFF K. Accumulation of proteinbound epidermal glucosylceramides in beta-glucocerebrosidase deficient type 2 Gaucher mice. FEBS Lett 447: 167–170, 1999.
80. DOTTI CG, PARTON RG, AND SIMONS K. Polarized sorting of glypiated
proteins in hippocampal neurons. Nature 349: 158 –161, 1991.
81. DOUGLAS AP, KERLEY R, AND ISSELBACHER KJ. Preparation and characterization of the lateral and basal plasma membranes of the rat
intestinal epithelial cell. Biochem J 128: 1329 –1338, 1972.
82. DUNN TM, HAAK D, MONAGHAN E, AND BEELER TJ. Synthesis of
monohydroxylated inositolphosphorylceramide (IPC-C) in Saccharomyces cerevisiae requires Scs7p, a protein with both a cytochrome b(5)-like domain and a hydroxylase/desaturase domain.
Yeast 14: 311–321, 1998.
83. DYATLOVITSKAYA EV, TIMOFEEVA NG, YAKIMENKO EF, BARSUKOV LI,
MUZYA GI, AND BERGELSON LD. A sphingomyelin transfer protein in
rat tumors and fetal liver. Eur J Biochem 123: 311–315, 1982.
84. ELLA KM, QI C, DOLAN JW, THOMPSON RP, AND MEIER KE. Characterization of a sphingomyelinase activity in Saccharomyces cerevisiae. Arch Biochem Biophys 340: 101–110, 1997.
85. FENSOME AC, RODRIGUES-LIMA F, JOSEPHS M, PATERSON HF, AND KATAN
M. A neutral magnesium-dependent sphingomyelinase isoform associated with intracellular membranes and reversibly inhibited by
reactive oxygen species. J Biol Chem 275: 1128 –1136, 2000.
86. FIEDLER K, KOBAYASHI T, KURZCHALIA TV, AND SIMONS K. Glycosphingolipid-enriched, detergent-insoluble complexes in protein sorting
in epithelial cells. Biochemistry 32: 6365– 6373, 1993.
87. FIEDLER K, PARTON RG, KELLNER R, ETZOLD T, AND SIMONS K. Vip36, a
novel component of glycolipid rafts and exocytic carrier vesicles in
epithelial cells. EMBO J 13: 1729 –1740, 1994.
88. FIEDLER K AND SIMONS K. The role of N-glycans in the secretory
pathway. Cell 81: 309 –312, 1995.
89. FISHMAN MA, MADYASTHA P, AND PRENSKY AL. The effect of undernutrition on the development of myelin in the rat central nervous
system. Lipids 6: 458 – 465, 1971.
90. FLEISCHER B, ZAMBRANO F, AND FLEISCHER S. Biochemical characterization of the Golgi complex of mammalian cells. J Supramol
Struct 2: 737–750, 1974.
91. FOLSCH H, OHNO H, BONIFACINO JS, AND MELLMAN I. A novel clathrin
adaptor complex mediates basolateral targeting in polarized epithelial cells. Cell 99: 189 –198, 1999.
92. FORSTNER GG, TANAKA K, AND ISSELBACHER KJ. Lipid composition of
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
52.
ment of monohexosyl sphingolipids is required for higher glycosphingolipid biosynthesis. J Cell Biol 133: 15–28, 1996.
BUTON X, MORROT G, FELLMANN P, AND SEIGNEURET M. Ultrafast
glycerophospholipid-selective transbilayer motion mediated by a
protein in the endoplasmic reticulum membrane. J Biol Chem 271:
6651– 6657, 1996.
BUTOR C, STELZER EH, SONNENBERG A, AND DAVOUST J. Apical and
basal Forssman antigen in MDCK II cells: a morphological and
quantitative study. Eur J Cell Biol 56: 269 –285, 1991.
CARMEL G, RODRIGUE F, CARRIÈRE S, AND LE GRIMELLEC C. Composition and physical properties of lipids from plasma membranes of
dog kidney. Biochim Biophys Acta 818: 149 –157, 1985.
CHAN KF AND LIU Y. Ganglioside-binding proteins in skeletal and
cardiac muscle. Glycobiology 1: 193–203, 1991.
CHAPMAN D AND COLLIN DT. Differential thermal analysis of phospholipids. Nature 206: 189, 1965.
CHATTERJEE S, HAN H, ROLLINS S, AND CLEVELAND T. Molecular cloning, characterization, and expression of a novel human neutral
sphingomyelinase. J Biol Chem 274: 37407–37412, 1999.
CHEN CS, MARTIN OC, AND PAGANO RE. Changes in the spectral
properties of a plasma membrane lipid analog during the first
seconds of endocytosis in living cells. Biophys J 72: 37–50, 1997.
CHESTER MA. IUPAC-IUB Joint Commission on Biochemical Nomenclature (JCBN). Nomenclature of glycolipids: recommendations 1997. Eur J Biochem 257: 293–298, 1998.
CHIGORNO V, VALSECCHI M, ACQUOTTI D, SONNINO S, AND TETTAMANTI G.
Formation of a cytosolic ganglioside-protein complex following
administration of photoreactive ganglioside GM1 to human fibroblasts in culture. FEBS Lett 263: 329 –331, 1990.
CHITWOOD DJ, LUSBY WR, THOMPSON MJ, KOCHANSKY JP, AND HOWARTH
OW. The glycosylceramides of the nematode Caenorhabditis elegans contain an unusual, branched-chain sphingoid base. Lipids
30: 567–573, 1995.
CLUETT EB, KUISMANEN E, AND MACHAMER CE. Heterogeneous distribution of the unusual phospholipid semilysobisphosphatidic acid
through the Golgi complex. Mol Biol Cell 8: 2233–2240, 1997.
CLUETT EB AND MACHAMER CE. The envelope of vaccinia virus
reveals an unusual phospholipid in Golgi complex membranes.
J Cell Sci 109: 2121–2131, 1996.
COETZEE T, FUJITA N, DUPREE J, SHI R, BLIGHT A, SUZUKI K, SUZUKI K,
AND POPKO B. Myelination in the absence of galactocerebroside and
sulfatide: normal structure with abnormal function and regional
instability. Cell 86: 209 –219, 1996.
COLBEAU A, NACHBAUR J, AND VIGNAIS PM. Enzymic characterization
and lipid composition of rat liver subcellular membranes. Biochim
Biophys Acta 249: 462– 492, 1971.
COLE NB, ELLENBERG J, SONG J, DIEULIIS D, AND LIPPINCOTT-SCHWARTZ
J. Retrograde transport of Golgi-localized proteins to the ER. J Cell
Biol 140: 1–15, 1998.
COLE NB, SCIAKY N, MAROTTA A, SONG J, AND LIPPINCOTT-SCHWARTZ J.
Golgi dispersal during microtubule disruption: regeneration of
Golgi stacks at peripheral endoplasmic reticulum sites. Mol Cell
Biol 7: 631– 650, 1996.
COLLINS RN AND WARREN G. Sphingolipid transport in mitotic HeLa
cells. J Biol Chem 267: 24906 –24911, 1992.
CORNEA RL AND THOMAS DD. Effects of membrane thickness on the
molecular dynamics and enzymatic activity of reconstituted CaATPase. Biochemistry 33: 2912–2920, 1994.
COSSON P AND LETOURNEUR F. Coatomer interaction with di-lysine
endoplasmic reticulum retention motifs. Science 263: 1629 –1631,
1994.
COSTE H, MARTEL MB, AND GOT R. Topology of glucosylceramide
synthesis in Golgi membranes from porcine submaxillary glands.
Biochim Biophys Acta 858: 6 –12, 1986.
DAUM G AND VANCE JE. Import of lipids into mitochondria. Prog
Lipid Res 36: 103–130, 1997.
DAVID D, SUNDARABABU S, AND GERST JE. Involvement of long chain
fatty acid elongation in the trafficking of secretory vesicles in yeast.
J Cell Biol 143: 1167–1182, 1998.
DEAN N, ZHANG YB, AND POSTER JB. The VRG4 gene is required for
GDP-mannose transport into the lumen of the Golgi in the yeast,
Saccharomyces cerevisiae. J Biol Chem 272: 31908 –31914, 1997.
1715
1716
93.
94.
95.
96.
97.
99.
100.
101.
102.
103.
104.
105.
106.
107.
108.
109.
110.
111.
112.
113.
the isolated rat intestinal microvillus membrane. Biochem J 109:
51–59, 1968.
FRIANT S, ZANOLARI B, AND RIEZMAN H. Increased protein kinase or
decreased PP2A activity bypasses sphingoid base requirement in
endocytosis. EMBO J 19: 2834 –2844, 2000.
FRIEDRICHSON T AND KURZCHALIA TV. Microdomains of GPI-anchored
proteins in living cells revealed by crosslinking. Nature 394: 802–
805, 1998.
FUJIMOTO H, TADANO-ARITOMI K, TOKUMASU A, ITO K, HIKITA T, SUZUKI
K, AND ISHIZUKA I. Requirement of seminolipid in spermatogenesis
revealed by UDP-galactose: ceramide galactosyltransferase-deficient mice. J Biol Chem 275: 22623–22626, 2000.
FUJIMOTO T. GPI-anchored proteins, glycosphingolipids, and sphingomyelin are sequestered to caveolae only after crosslinking.
J Histochem Cytochem 44: 929 –941, 1996.
FUKASAWA M, NISHIJIMA M, AND HANADA K. Genetic evidence for
ATP-dependent endoplasmic reticulum-to-Golgi apparatus trafficking of ceramide for sphingomyelin synthesis in Chinese hamster
ovary cells. J Cell Biol 144: 673– 685, 1999.
FUNAKOSHI T, YASUDA S, FUKASAWA M, NISHIJIMA M, AND HANADA K.
Reconstitution of ATP- and cytosol-dependent transport of de novo
synthesized ceramide to the site of sphingomyelin synthesis in
semi-intact cells. J Biol Chem 275: 29938 –29945, 2000.
FUTERMAN AH, KHANIN R, AND SEGEL LA. Lipid diffusion in neurons.
Nature 362: 119, 1993.
FUTERMAN AH AND PAGANO RE. Determination of the intracellular
sites and topology of glucosylceramide synthesis in rat liver. Biochem J 280: 295–302, 1991.
FUTERMAN AH, STIEGER B, HUBBARD AL, AND PAGANO RE. Sphingomyelin synthesis in rat liver occurs predominantly at the cis and
medial cisternae of the Golgi apparatus. J Biol Chem 265: 8650 –
8657, 1990.
GADELLA BM, GADELLA TWJ JR, COLENBRANDER B, VAN GOLDE LMG,
AND LOPES-CARDOZO M. Visualization and quantification of glycolipid
polarity dynamics in the plasma membrane of the mammalian
spermatozoon. J Cell Sci 107: 2151–2163, 1994.
GALBIATI F, VOLONTE D, MEANI D, MILLIGAN G, LUBLIN DM, LISANTI MP,
AND PARENTI M. The dually acylated NH2-terminal domain of G(i1
alpha) is sufficient to target a green fluorescent protein reporter to
caveolin-enriched plasma membrane domains: palmitoylation of
caveolin-1 is required for the recognition of dually acylated Gprotein alpha subunits in vivo. J Biol Chem 274: 5843–5850, 1999.
GARCIA-ARRANZ M, MALDONADO AM, MAZON MJ, AND PORTILLO F. Transcriptional control of yeast plasma membrane H(⫹)-ATPase by
glucose. Cloning and characterization of a new gene involved in
this regulation. J Biol Chem 269: 18076 –18082, 1994.
GERDT S, LOCHNIT G, DENNIS RD, AND GEYER R. Isolation and structural analysis of three neutral glycosphingolipids from a mixed
population of Caenorhabditis elegans (Nematoda: Rhabditida).
Glycobiology 7: 265–275, 1997.
GILLARD BK, HEATH JP, THURMON LT, AND MARCUS DM. Association of
glycosphingolipids with intermediate filaments of human umbilical
vein endothelial cells. Exp Cell Res 192: 433– 444, 1991.
GILLARD BK, THURMON LT, AND MARCUS DM. Association of glycosphingolipids with intermediate filaments of mesenchymal, epithelial, glial, and muscle cells. Cell Motil Cytoskeleton 21: 255–271,
1992.
GKANTIRAGAS I, BRÜGGER B, STÜVEN E, KALOYANOVA D, LI XY, LÖHR K,
LOTTSPEICH F, WIELAND FT, AND HELMS JB. Sphingomyelin-enriched
microdomains at the Golgi complex. Mol Biol Cell. In press.
GLEESON PA. Targeting of proteins to the Golgi apparatus. Histochem Cell Biol 109: 517–532, 1998.
GLICK BS AND MALHOTRA V. The curious status of the Golgi apparatus. Cell 95: 883– 889, 1998.
GRASSÉ MPP. Ultrastructure, polarite et reproduction de l’appareil
de Golgi. CR Acad Sci Paris 245: 1278 –1281, 1957.
GRIFFITHS G, BACK R, AND MARSH M. A quantitative analysis of the
endocytic pathway in baby hamster kidney cells. J Cell Biol 109:
2703–2720, 1989.
GROSS SK, LYERLA TA, EVANS JE, AND MCCLUER RH. Expression of
glycosphingolipids in serum-free primary cultures of mouse kidney
cells: male-female differences and androgen sensitivity. Mol Cell
Biochem 137: 25–31, 1994.
Physiol Rev • VOL
114. GROVE SN, BRACKER CE, AND MORRÉ DJ. Cytomembrane differentiation in the endoplasmic reticulum-Golgi apparatus-vesicle complex. Science 161: 171–173, 1968.
115. GUT A, KAPPELER F, HYKA N, BALDA MS, HAURI HP, AND MATTER K.
Carbohydrate-mediated Golgi to cell surface transport and apical
targeting of membrane proteins. EMBO J 17: 1919 –1929, 1998.
116. HAAK D, GABLE K, BEELER T, AND DUNN T. Hydroxylation of Saccharomyces cerevisiae ceramides requires Sur2p and Scs7p. J Biol
Chem 272: 29704 –29710, 1997.
117. HAGMANN J AND FISHMAN PH. Detergent extraction of cholera toxin
and gangliosides from cultured cells and isolated membranes. Biochim Biophys Acta 720: 181–187, 1982.
118. HAKOMORI S, HANDA K, IWABUCHI K, YAMAMURA S, AND PRINETTI A. New
insights in glycosphingolipid function: “glycosignaling domain,” a
cell surface assembly of glycosphingolipids with signal transducer
molecules, involved in cell adhesion coupled with signaling. Glycobiology 8: xi-xix, 1998.
119. HAKOMORI SI AND IGARASHI Y. Functional role of glycosphingolipids
in cell recognition and signaling. J Biochem 118: 1091–1103, 1995.
120. HALDAR K. Sphingolipid synthesis and membrane formation by
Plasmodium. Trends Cell Biol 6: 398 – 405, 1996.
121. HANADA K, HARA T, FUKASAWA M, YAMAJI A, UMEDA M, AND NISHIJIMA
M. Mammalian cell mutants resistant to a sphingomyelin-directed
cytolysin. Genetic and biochemical evidence for complex formation of the LCB1 protein with the LCB2 protein for serine palmitoyltransferase. J Biol Chem 273: 33787–33794, 1998.
122. HANADA K, HARA T, AND NISHIJIMA M. Purification of the serine
palmitoyltransferase complex responsible for sphingoid base synthesis by using affinity peptide chromatography techniques. J Biol
Chem 275: 8409 – 8415, 2000.
123. HANADA K, NISHIJIMA M, KISO M, HASEGAWA A, FUJITA S, OGAWA T, AND
AKAMATSU Y. Sphingolipids are essential for the growth of Chinese
hamster ovary cells. Restoration of the growth of a mutant defective in sphingoid base biosynthesis by exogenous sphingolipids.
J Biol Chem 267: 23527–23533, 1992.
124. HANNAN LA AND EDIDIN M. Traffic, polarity, and detergent solubility
of a glycosylphosphatidylinositol-anchored protein after LDL-deprivation of MDCK cells. J Cell Biol 133: 1265–1276, 1996.
125. HANNAN LA, LISANTI MP, RODRIGUEZ-BOULAN E, AND EDIDIN M. Correctly sorted molecules of a GPI-anchored protein are clustered
and immobile when they arrive at the apical surface of MDCK cells.
J Cell Biol 120: 353–358, 1993.
126. HANNUN YA. Functions of ceramide in coordinating cellular responses to stress. Science 274: 1855–1859, 1996.
127. HANNUN YA AND OBEID LM. Ceramide and the eukaryotic stress
response. Biochem Soc Trans 25: 1171–1175, 1997.
128. HARDER T, SCHEIFFELE P, VERKADE P, AND SIMONS K. Lipid domain
structure of the plasma membrane revealed by patching of membrane components. J Cell Biol 141: 929 –942, 1998.
129. HARRIS JS, EPPS DE, DAVIO SR, AND KEZDY FJ. Evidence for transbilayer, tail-to-tail cholesterol dimers in dipalmitoylglycerophosphocholine liposomes. Biochemistry 34: 3851–3857, 1995.
130. HAY JC, KLUMPERMAN J, OORSCHOT V, STEEGMAIER M, KUO CS, AND
SCHELLER RH. Localization, dynamics, and protein interactions reveal distinct roles for ER and Golgi SNAREs. J Cell Biol 141:
1489 –1502, 1998.
131. HECHTBERGER P, ZINSER E, SAF R, HUMMEL K, PALTAUF F, AND DAUM G.
Characterization, quantification and subcellular localization of inositol-containing sphingolipids of the yeast, Saccharomyces cerevisiae. Eur J Biochem 225: 641– 649, 1994.
132. HEIDLER SA AND RADDING JA. Inositol phosphoryl transferases from
human pathogenic fungi. Biochim Biophys Acta 1500: 147–152,
2000.
133. HERRMANN A, ZACHOWSKI A, AND DEVAUX PF. Protein-mediated phospholipid translocation in the endoplasmic reticulum with a low
lipid specificity. Biochemistry 29: 2023–2027, 1990.
134. HERSCHKOWITZ N, MCKHANN GM, SAXENA S, AND SHOOTER EM. Characterization of sulphatide-containing lipoproteins in rat brain.
J Neurochem 15: 1181–1188, 1968.
135. HIGASHI H, OMORI A, AND YAMAGATA T. Calmodulin, a gangliosidebinding protein. Binding of gangliosides to calmodulin in the presence of calcium. J Biol Chem 267: 9831–9838, 1992.
136. HIRSCHBERG K, MILLER CM, ELLENBERG J, PRESLEY JF, SIGGIA ED,
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
98.
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
137.
138.
139.
140.
141.
143.
144.
145.
146.
147.
148.
149.
150.
151.
152.
153.
154.
155.
156.
157.
158.
159.
Physiol Rev • VOL
160.
161.
162.
163.
164.
165.
166.
167.
168.
169.
170.
171.
172.
173.
174.
175.
176.
177.
178.
179.
180.
181.
182.
regions of an intestinal epithelial cell membrane of mouse. Biochim Biophys Acta 369: 222–233, 1974.
KEENAN TW AND MORRÉ DJ. Phospholipid class and fatty acid composition of Golgi apparatus isolated from rat liver and comparison
with other cell fractions. Biochemistry 9: 19 –25, 1970.
KELLER P AND SIMONS K. Cholesterol is required for surface transport
of influenza virus hemagglutinin. J Cell Biol 140: 1357–1367, 1998.
KENWORTHY AK AND EDIDIN M. Distribution of a glycosylphosphatidylinositol-anchored protein at the apical surface of MDCK cells
examined at a resolution of ⬍100 A using imaging fluorescence
resonance energy transfer. J Cell Biol 142: 69 – 84, 1998.
KENWORTHY AK, PETRANOVA N, AND EDIDIN M. High-resolution FRET
microscopy of cholera toxin B-subunit and GPI-anchored proteins
in cell plasma membranes. Mol Biol Cell 11: 1645–1655, 2000.
KILLIAN JA. Hydrophobic mismatch between proteins and lipids in
membranes. Biochim Biophys Acta 1376: 401– 415, 1998.
KOBAYASHI T, STORRIE B, SIMONS K, AND DOTTI CG. A functional
barrier to movement of lipids in polarized neurons. Nature 359:
647– 650, 1992.
KOCH J, GARTNER S, LI CM, QUINTERN LE, BERNARDO K, LEVRAN O,
SCHNABEL D, DESNICK RJ, SCHUCHMAN EH, AND SANDHOFF K. Molecular cloning and characterization of a full-length complementary
DNA encoding human acid ceramidase. Identification of the first
molecular lesion causing Farber disease. J Biol Chem 271: 33110 –
33115, 1996.
KOHAMA T, OLIVERA A, EDSALL L, NAGIEC MM, DICKSON R, AND SPIEGEL
S. Molecular cloning and functional characterization of murine
sphingosine kinase. J Biol Chem 273: 23722–23728, 1998.
KOK JW, BABIA T, AND HOEKSTRA D. Sorting of sphingolipids in the
endocytic pathway of HT29 cells. J Cell Biol 114: 231–239, 1991.
KOK JW, BABIA T, KLAPPE K, EGEA G, AND HOEKSTRA D. Ceramide
transport from endoplasmic reticulum to Golgi apparatus is not
vesicle-mediated. Biochem J 333: 779 –786, 1998.
KOLESNICK R AND HANNUN YA. Ceramide and apoptosis. Trends
Biochem Sci 24: 224 –225, 1999.
KOLTER T AND SANDHOFF K. Sphingolipids: their metabolic pathways
and the pathobiochemistry of neurodegenerative diseases. Angew
Chem Int Ed 38: 1532–1568, 1999.
KOYNOVA R AND CAFFREY M. Phases and phase transitions of the
sphingolipids. Biochim Biophys Acta 1255: 213–236, 1995.
KOYNOVA R AND CAFFREY M. Phases and phase transitions of the
phosphatidylcholines. Biochim Biophys Acta 1376: 91–145, 1998.
KÜBLER E, DOHLMAN HG, AND LISANTI MP. Identification of Triton
X-100 insoluble membrane domains in the yeast Saccharomyces
cerevisiae. Lipid requirements for targeting of heterotrimeric Gprotein subunits. J Biol Chem 271: 32975–32980, 1996.
KUNDU A, AVALOS RT, SANDERSON CM, AND NAYAK DP. Transmembrane domain of influenza virus neuraminidase, a type II protein,
possesses an apical sorting signal in polarized MDCK cells. J Virol
70: 6508 – 6515, 1996.
KURODA M, HASHIDA-OKADO T, YASUMOTO R, GOMI K, KATO I, AND
TAKESAKO K. An aureobasidin A resistance gene isolated from Aspergillus is a homolog of yeast AUR1, a gene responsible for
inositol phosphorylceramide (IPC) synthase activity. Mol Gen
Genet 261: 290 –296, 1999.
KURZCHALIA TV, DUPREE P, PARTON RG, KELLNER R, VIRTA H, LEHNERT
M, AND SIMONS K. VIP21, a 21-kD membrane protein is an integral
component of trans-Golgi-network-derived transport vesicles.
J Cell Biol 118: 1003–1014, 1992.
KURZCHALIA TV AND PARTON RG. Membrane microdomains and
caveolae. Curr Opin Cell Biol 11: 424 – 431, 1999.
LADBROOKE BD, WILLIAMS RM, AND CHAPMAN D. Studies on lecithincholesterol-water interactions by differential scanning calorimetry
and X-ray diffraction. Biochim Biophys Acta 150: 333–340, 1968.
LAFONT F, VERKADE P, GALLI T, WIMMER C, LOUVARD D, AND SIMONS K.
Raft association of SNAP receptors acting in apical trafficking in
Madin-Darby canine kidney cells. Proc Natl Acad Sci USA 96:
3734 –3738, 1999.
LANGE Y, D’ALESSANDRO JS, AND SMALL DM. The affinity of cholesterol for phosphatidylcholine and sphingomyelin. Biochim Biophys Acta 556: 388 –398, 1979.
LANGE Y, SWAISGOOD MH, RAMOS BV, AND STECK TL. Plasma membranes contain half the phospholipid and 90% of the cholesterol
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
142.
PHAIR RD, AND LIPPINCOTT-SCHWARTZ J. Kinetic analysis of secretory
protein traffic and characterization of Golgi to plasma membrane
transport intermediates in living cells. J Cell Biol 143: 1485–1503,
1998.
HISE MK, MANTULIN WW, AND WEINMAN EJ. Fluidity and composition
of brush border and basolateral membranes from rat kidney. Am J
Physiol Renal Fluid Electrolyte Physiol 247: F434 –F439, 1984.
HOFMANN K AND DIXIT VM. Ceramide in apoptosis: does it really
matter? Trends Biochem Sci 23: 374 –377, 1998.
HOFMANN K AND DIXIT VM. Reply to Kolesnick and Hannun, and
Perry and Hannun. Trends Biochem Sci 24: 227, 1999.
HOFMANN K, TOMIUK S, WOLFF G, AND STOFFEL W. Cloning and characterization of the mammalian brain-specific, Mg2⫹-dependent neutral sphingomyelinase. Proc Natl Acad Sci USA 97: 5895–5900, 2000.
HOLOPAINEN JM, LEHTONEN JY, AND KINNUNEN PK. Lipid microdomains in dimyristoylphosphatidylcholine-ceramide liposomes.
Chem Phys Lipids 88: 1–13, 1997.
HOSTETLER KY, ZENNER BD, AND MORRIS HP. Phospholipid content of
mitochondrial and microsomal membranes from Morris hepatomas
of varying growth rates. Cancer Res 39: 2978 –2983, 1979.
IBER H, VAN ECHTEN G, AND SANDHOFF K. Fractionation of primary
cultured cerebellar neurons: distribution of sialyltransferases involved in ganglioside biosynthesis. J Neurochem 58: 1533–1537,
1992.
ICHIKAWA S AND HIRABAYASHI Y. Glucosylceramide synthase and glycosphingolipid synthesis. Trends Cell Biol 8: 198 –202, 1998.
ICHIKAWA S, NAKAJO N, SAKIYAMA H, AND HIRABAYASHI Y. A mouse B16
melanoma mutant deficient in glycolipids. Proc Natl Acad Sci USA
91: 2703–2707, 1994.
ITIN C, SCHINDLER R, AND HAURI HP. Targeting of protein ERGIC-53
to the ER/ERGIC/cis-Golgi recycling pathway. J Cell Biol 131:
57– 67, 1995.
IWABUCHI K, HANDA K, AND HAKOMORI S. Separation of “glycosphingolipid signaling domain” from caveolin-containing membrane fraction in mouse melanoma B16 cells and its role in cell adhesion
coupled with signaling. J Biol Chem 273: 33766 –33773, 1998.
JAIN MK AND WHITE HBD. Long-range order in biomembranes. Adv
Lipid Res 15: 1– 60, 1977.
JASINSKA R, ZHANG QX, PILQUIL C, SINGH I, XU J, DEWALD J, DILLON DA,
BERTHIAUME LG, CARMAN GM, WAGGONER DW, AND BRINDLEY DN.
Lipid phosphate phosphohydrolase-1 degrades exogenous glycerolipid and sphingolipid phosphate esters. Biochem J 340: 677– 686,
1999.
JECKEL D, KARRENBAUER A, BIRK R, SCHMIDT RR, AND WIELAND F.
Sphingomyelin is synthesized in the cis Golgi. FEBS Lett 261:
155–157, 1990.
JECKEL D, KARRENBAUER A, BURGER KNJ, VAN MEER G, AND WIELAND F.
Glucosylceramide is synthesized at the cytosolic surface of various
Golgi subfractions. J Cell Biol 117: 259 –267, 1992.
JENKINS GM, RICHARDS A, WAHL T, MAO C, OBEID L, AND HANNUN Y.
Involvement of yeast sphingolipids in the heat stress response of
Saccharomyces cerevisiae. J Biol Chem 272: 32566 –32572, 1997.
JEYAKUMAR M, BUTTERS TD, CORTINA-BORJA M, HUNNAM V, PROIA RL,
PERRY VH, DWEK RA, AND PLATT FM. Delayed symptom onset and
increased life expectancy in Sandhoff disease mice treated with
N-butyldeoxynojirimycin. Proc Natl Acad Sci USA 96: 6388 – 6393,
1999.
KARLSSON KA. On the chemistry and occurrence of sphingolipid
long-chain bases. Chem Phys Lipids 5: 6 – 43, 1970.
KARLSSON KA, SAMUELSSON BE, AND STEEN GO. The sphingolipid
composition of bovine kidney cortex, medulla and papilla. Biochim
Biophys Acta 316: 317–335, 1973.
KARLSSON KA, SAMUELSSON BE, AND STEEN GO. Detailed structure of
sphingomyelins and ceramides from different regions of bovine
kidney with special reference to long-chain bases. Biochim Biophys Acta 316: 336 –362, 1973.
KARNOVSKY MJ, KLEINFELD AM, HOOVER RL, AND KLAUSNER RD. The
concept of lipid domains in membranes. J Cell Biol 94: 1– 6, 1982.
KAWAI H, ALLENDE ML, WADA R, KONO M, SANGO K, DENG C, MIYAKAWA
T, CRAWLEY JN, WERTH N, BIERFREUND U, SANDHOFF K, AND PROIA RL.
Mice expressing only monosialoganglioside GM3 exhibit lethal audiogenic seizures. J Biol Chem 276: 6885– 6888, 2001.
KAWAI K, FUJITA M, AND NAKAO M. Lipid components of two different
1717
1718
183.
184.
185.
186.
188.
189.
190.
191.
192.
193.
194.
195.
196.
197.
198.
199.
200.
201.
202.
203.
204.
and sphingomyelin in cultured human fibroblasts. J Biol Chem 264:
3786 –3793, 1989.
LANNERT H, BÜNNING C, JECKEL D, AND WIELAND FT. Lactosylceramide is synthesized in the lumen of the Golgi apparatus. FEBS Lett
342: 91–96, 1994.
LANNERT H, GORGAS K, MEIßNER I, WIELAND FT, AND JECKEL D. Functional organization of the Golgi apparatus in glycosphingolipid
biosynthesis. Lactosylceramide and subsequent glycosphingolipids
are formed in the lumen of the late Golgi. J Biol Chem 273:
2939 –2946, 1998.
LANOIX J, OUWENDIJK J, LIN CC, STARK A, LOVE HD, OSTERMANN J, AND
NILSSON T. GTP hydrolysis by Arf-1 mediates sorting and concentration of Golgi resident enzymes into functional COP I vesicles.
EMBO J 18: 4935– 4948, 1999.
LANTERMAN MM AND SABA JD. Characterization of sphingosine kinase (SK) activity in Saccharomyces cerevisiae and isolation of
SK-deficient mutants. Biochem J 332: 525–531, 1998.
LE BIVIC A, QUARONI A, NICHOLS B, AND RODRIGUEZ-BOULAN E. Biogenetic pathways of plasma membrane proteins in Caco-2, a human
intestinal epithelial cell line. J Cell Biol 111: 1351–1361, 1990.
LEDESMA MD, BRÜGGER B, BÜNNING C, WIELAND FT, AND DOTTI CG.
Maturation of the axonal plasma membrane requires upregulation
of sphingomyelin synthesis and formation of protein-lipid complexes. EMBO J 18: 1761–1771, 1999.
LEDESMA MD, SIMONS K, AND DOTTI CG. Neuronal polarity: essential
role of protein-lipid complexes in axonal sorting. Proc Natl Acad
Sci USA 95: 3966 –3971, 1998.
LEE AG. How lipids interact with an intrinsic membrane protein:
the case of the calcium pump. Biochim Biophys Acta 1376: 381–
390, 1998.
LEE MJ, VAN BROCKLYN JR, THANGADA S, LIU CH, HAND AR, MENZELEEV R, SPIEGEL S, AND HLA T. Sphingosine-1-phosphate as a ligand
for the G protein coupled receptor EDG-1. Science 279: 1552–1555,
1998.
LESTER RL AND DICKSON RC. Sphingolipids with inositolphosphatecontaining head groups. Adv Lipid Res 26: 253–274, 1993.
LESTER RL, WELLS GB, OXFORD G, AND DICKSON RC. Mutant strains of
Saccharomyces cerevisiae lacking sphingolipids synthesize novel
inositol glycerophospholipids that mimic sphingolipid structures.
J Biol Chem 268: 845– 856, 1993.
LETOURNEUR F, GAYNOR EC, HENNECKE S, DEMOLLIERE C, DUDEN R,
EMR SD, RIEZMAN H, AND COSSON P. Coatomer is essential for retrieval of dilysine-tagged proteins to the endoplasmic reticulum.
Cell 79: 1199 –1207, 1994.
LEVINE TP, WIGGINS CA, AND MUNRO S. Inositol phosphorylceramide
synthase is located in the Golgi apparatus of Saccharomyces cerevisiae. Mol Biol Cell 11: 2267–2281, 2000.
LEWIS MJ AND PELHAM HR. SNARE-mediated retrograde traffic from
the Golgi complex to the endoplasmic reticulum. Cell 85: 205–215,
1996.
LIN CC, LOVE HD, GUSHUE JN, BERGERON JJ, AND OSTERMANN J.
ER/Golgi intermediates acquire Golgi enzymes by brefeldin A-sensitive retrograde transport in vitro. J Cell Biol 147: 1457–1472, 1999.
LIN X, MATTJUS P, PIKE HM, WINDEBANK AJ, AND BROWN RE. Cloning
and expression of glycolipid transfer protein from bovine and
porcine brain. J Biol Chem 275: 5104 –5110, 2000.
LININGTON C AND RUMSBY MG. Accessibility of galactosyl ceramides
to probe reagents in central nervous system myelin. J Neurochem
35: 983–992, 1980.
LISANTI MP, CARAS IW, DAVITZ MA, AND RODRIGUEZ-BOULAN E. A
glycophospholipid membrane anchor acts as an apical targeting
signal in polarized epithelial cells. J Cell Biol 109: 2145–2156, 1989.
LISANTI MP, SARGIACOMO M, GRAEVE L, SALTIEL AR, AND RODRIGUEZBOULAN E. Polarized apical distribution of glycosyl-phosphatidylinositol-anchored proteins in a renal epithelial cell line. Proc Natl
Acad Sci USA 85: 9557–9561, 1988.
LISANTI MP, SCHERER PE, TANG ZL, AND SARGIACOMO M. Caveolae,
caveolin and caveolin-rich membrane domains: a signalling hypothesis. Trends Cell Biol 4: 231–235, 1994.
LISCUM L AND MUNN NJ. Intracellular cholesterol transport. Biochim
Biophys Acta 1438: 19 –37, 1999.
LÖFGREN H AND PASCHER I. Molecular arrangements of sphingolipids.
Physiol Rev • VOL
205.
206.
207.
208.
209.
210.
211.
212.
213.
214.
215.
216.
217.
218.
219.
220.
221.
222.
223.
The monolayer behaviour of ceramides. Chem Phys Lipids 20:
273–284, 1977.
LOURA LM AND PRIETO M. Dehydroergosterol structural organization
in aqueous medium and in a model system of membranes. Biophys
J 72: 2226 –2236, 1997.
LOVE HD, LIN CC, SHORT CS, AND OSTERMANN J. Isolation of functional Golgi-derived vesicles with a possible role in retrograde
transport. J Cell Biol 140: 541–551, 1998.
LUETTERFORST R, STANG E, ZORZI N, CAROZZI A, WAY M, AND PARTON
RG. Molecular characterization of caveolin association with the
Golgi complex: identification of a cis-Golgi targeting domain in the
caveolin molecule. J Cell Biol 145: 1443–1459, 1999.
LUND-KATZ S, LABODA HM, MCLEAN LR, AND PHILLIPS MC. Influence of
molecular packing and phospholipid type on rates of cholesterol
exchange. Biochemistry 27: 3416 –3423, 1988.
MACCIONI HJ, DANIOTTI JL, AND MARTINA JA. Organization of ganglioside synthesis in the Golgi apparatus. Biochim Biophys Acta
1437: 101–118, 1999.
MACEYKA M AND MACHAMER CE. Ceramide accumulation uncovers a
cycling pathway for the cis-Golgi network marker, infectious bronchitis virus M protein. J Cell Biol 139: 1411–1418, 1997.
MADORE N, SMITH KL, GRAHAM CH, JEN A, BRADY K, HALL S, AND
MORRIS R. Functionally different GPI proteins are organized in
different domains on the neuronal surface. EMBO J 18: 6917– 6926,
1999.
MAGGIO B, CUMAR FA, AND CAPUTTO R. Interactions of gangliosides
with phospholipids and glycosphingolipids in mixed monolayers.
Biochem J 175: 1113–1118, 1978.
MALGAT M, MAURICE A, AND BARAUD J. Sphingomyelin and ceramidephosphoethanolamine synthesis by microsomes and plasma membranes from rat liver and brain. J Lipid Res 27: 251–260, 1986.
MANDALA SM, FROMMER BR, THORNTON RA, KURTZ MB, YOUNG NM,
CABELLO MA, GENILLOUD O, LIESCH JM, SMITH JL, AND HORN WS.
Inhibition of serine palmitoyl-transferase activity by lipoxamycin. J
Antibiotics 47: 376 –379, 1994.
MANDALA SM, THORNTON R, TU Z, KURTZ MB, NICKELS J, BROACH J,
MENZELEEV R, AND SPIEGEL S. Sphingoid base 1-phosphate phosphatase: a key regulator of sphingolipid metabolism and stress response. Proc Natl Acad Sci USA 95: 150 –155, 1998.
MANDALA SM, THORNTON RA, FROMMER BR, CUROTTO JE, ROZDILSKY W,
KURTZ MB, GIACOBBE RA, BILLS GF, CABELLO MA, AND MARTIN I. The
discovery of australifungin, a novel inhibitor of sphinganine Nacyltransferase from Sporormiella australis producing organism,
fermentation, isolation, and biological activity. J Antibiotics 48:
349 –356, 1995.
MANDALA SM, THORNTON RA, FROMMER BR, DREIKORN S, AND KURTZ
MB. Viridiofungins, novel inhibitors of sphingolipid synthesis. J
Antibiotics 50: 339 –343, 1997.
MANDALA SM, THORNTON RA, MILLIGAN J, ROSENBACH M, GARCIA-CALVO
M, BULL HG, HARRIS G, ABRUZZO GK, FLATTERY AM, GILL CJ, BARTIZAL
K, DREIKORN S, AND KURTZ MB. Rustmicin, a potent antifungal agent,
inhibits sphingolipid synthesis at inositol phosphoceramide synthase. J Biol Chem 273: 14942–14949, 1998.
MANDALA SM, THORNTON RA, ROSENBACH M, MILLIGAN J, GARCIA-CALVO
M, BULL HG, AND KURTZ MB. Khafrefungin, a novel inhibitor of
sphingolipid synthesis. J Biol Chem 272: 32709 –32714, 1997.
MANDON EC, EHSES I, ROTHER J, VAN ECHTEN G, AND SANDHOFF K.
Subcellular localization and membrane topology of serine palmitoyltransferase, 3-dehydrosphinganine reductase, and sphinganine
N-acyltransferase in mouse liver. J Biol Chem 267: 11144 –11148,
1992.
MAO C, WADLEIGH M, JENKINS GM, HANNUN YA, AND OBEID LM.
Identification and characterization of Saccharomyces cerevisiae
dihydrosphingosine-1-phosphate phosphatase. J Biol Chem 272:
28690 –28694, 1997.
MAO C, XU R, BIELAWSKA A, AND OBEID LM. Cloning of an alkaline
ceramidase from Saccharomyces cerevisiae. An enzyme with reverse (CoA-independent) ceramide synthase activity. J Biol Chem
275: 6876 – 6884, 2000.
MAO C, XU R, BIELAWSKA A, SZULC ZM, AND OBEID LM. Cloning and
characterization of a Saccharomyces cerevisiae alkaline ceramidase with specificity for dihydroceramide. J Biol Chem 275: 31369 –
31378, 2000.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
187.
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
Physiol Rev • VOL
246.
247.
248.
249.
250.
251.
252.
253.
254.
255.
256.
257.
258.
259.
260.
261.
262.
263.
264.
265.
266.
267.
like immunosuppressant, ISP-1/myriocin. Biochem Biophys Res
Commun 211: 396 – 403, 1995.
MÖBIUS W, HERZOG V, SANDHOFF K, AND SCHWARZMANN G. Intracellular
distribution of a biotin-labeled ganglioside, GM1, by immunoelectron microscopy after endocytosis in fibroblasts. J Histochem Cytochem 47: 1005–1014, 1999.
MOLITORIS BA AND SIMON FR. Renal cortical brush-border and basolateral membranes: cholesterol and phospholipid composition and
relative turnover. J Membr Biol 83: 207–215, 1985.
MOLLENHAUER HH AND MORRE DJ. Perspectives on Golgi apparatus
form and function. J Electron Microsc Tech 17: 2–14, 1991.
MOREAU P, CASSAGNE C, KEENAN TW, AND MORRÉ DJ. Ceramide
excluded from cell-free vesicular lipid transfer from endoplasmic
reticulum to Golgi apparatus: evidence for lipid sorting. Biochim
Biophys Acta 1146: 9 –16, 1993.
MUKHERJEE S AND CHATTOPADHYAY A. Membrane organization at low
cholesterol concentrations: a study using 7-nitrobenz-2-oxa-1,3-diazol-4-yl-labeled cholesterol. Biochemistry 35: 1311–1322, 1996.
MUKHERJEE S, SOE TT, AND MAXFIELD FR. Endocytic sorting of lipid
analogues differing solely in the chemistry of their hydrophobic
tails. J Cell Biol 144: 1271–1284, 1999.
MUNIZ M AND RIEZMAN H. Intracellular transport of GPI-anchored
proteins. EMBO J 19: 10 –15, 2000.
MUNRO S. Sequences within and adjacent to the transmembrane
segment of alpha-2,6-sialyltransferase specify Golgi retention.
EMBO J 10: 3577–3588, 1991.
MUNRO S. An investigation of the role of transmembrane domains in
Golgi protein retention. EMBO J 14: 4695– 4704, 1995.
MUNRO S. Localization of proteins to the Golgi apparatus. Trends
Cell Biol 8: 11–15, 1998.
MURATA M, PERÄNEN J, SCHREINER R, WIELAND F, KURZCHALIA TV, AND
SIMONS K. VIP21/caveolin is a cholesterol-binding protein. Proc Natl
Acad Sci USA 92: 10339 –10343, 1995.
MUSCH A, XU H, SHIELDS D, AND RODRIGUEZ-BOULAN E. Transport of
vesicular stomatitis virus G protein to the cell surface is signal
mediated in polarized and nonpolarized cells. J Cell Biol 133:
543–558, 1996.
NAGIEC MM, BALTISBERGER JA, WELLS GB, LESTER RL, AND DICKSON
RC. The LCB2 gene of Saccharomyces and the related LCB1 gene
encode subunits of serine palmitoyltransferase, the initial enzyme
in sphingolipid synthesis. Proc Natl Acad Sci USA 91: 7899 –7902,
1994.
NAGIEC MM, NAGIEC EE, BALTISBERGER JA, WELLS GB, LESTER RL, AND
DICKSON RC. Sphingolipid synthesis as a target for antifungal drugs.
Complementation of the inositol phosphorylceramide synthase defect in a mutant strain of Saccharomyces cerevisiae by the AUR1
gene. J Biol Chem 272: 9809 –9817, 1997.
NAGIEC MM, SKRZYPEK M, NAGIEC EE, LESTER RL, AND DICKSON RC.
The LCB4 (YOR171c) and LCB5 (YLR260w) genes of Saccharomyces encode sphingoid long chain base kinases. J Biol Chem 273:
19437–19442, 1998.
NAGIEC MM, WELLS GB, LESTER RL, AND DICKSON RC. A suppressor
gene that enables Saccharomyces cerevisiae to grow without making sphingolipids encodes a protein that resembles an Escherichia
coli fatty acyltransferase. J Biol Chem 268: 22156 –22163, 1993.
NEZIL FA AND BLOOM M. Combined influence of cholesterol and
synthetic amphiphillic peptides upon bilayer thickness in model
membranes. Biophys J 61: 1176 –1183, 1992.
NILSSON OS AND DALLNER G. Enzyme and phospholipid asymmetry in
liver microsomal membranes. J Cell Biol 72: 568 –583, 1977.
NILSSON OS AND DALLNER G. Transverse asymmetry of phospholipids
in subcellular membranes of rat liver. Biochim Biophys Acta 464:
453– 458, 1977.
NILSSON T AND WARREN GB. Retention and retrieval in the endoplasmic reticulum and the Golgi apparatus. Curr Opin Cell Biol 6:
517–521, 1994.
NISHIMURA K. Phytosphingosine is a characteristic component of the
glycolipids in the vertebrate intestine. Comp Biochem Physiol 86:
149 –154, 1987.
NOMURA T, TAKIZAWA M, AOKI J, ARAI H, INOUE K, WAKISAKA E,
YOSHIZUKA N, IMOKAWA G, DOHMAE N, TAKIO K, HATTORI M, AND
MATSUO N. Purification, cDNA cloning, and expression of UDP-Gal:
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
224. MARKS DL, WU KJ, PAUL P, KAMISAKA Y, WATANABE R, AND PAGANO RE.
Oligomerization and topology of the Golgi membrane protein glucosylceramide synthase. J Biol Chem 274: 451– 456, 1999.
225. MARRS JA, ANDERSSON-FISONE C, JEONG MC, COHEN-GOULD L, ZURZOLO
C, NABI IR, RODRIGUEZ-BOULAN E, AND NELSON WJ. Plasticity in
epithelial cell phenotype: modulation by expression of different
cadherin cell adhesion molecules. J Cell Biol 129: 507–519, 1995.
226. MARTIN OC AND PAGANO RE. Internalization and sorting of a fluorescent analogue of glucosylceramide to the Golgi apparatus of
human skin fibroblasts: utilization of endocytic and nonendocytic
transport mechanisms. J Cell Biol 125: 769 –781, 1994.
227. MARTIN-BELMONTE F, PUERTOLLANO R, MILLAN J, AND ALONSO MA. The
MAL proteolipid is necessary for the overall apical delivery of
membrane proteins in the polarized epithelial Madin-Darby canine
kidney and Fischer rat thyroid cell lines. Mol Biol Cell 11: 2033–
2045, 2000.
228. MARTINEZ-MENARGUEZ JA, GEUZE HJ, SLOT JW, AND KLUMPERMAN J.
Vesicular tubular clusters between the ER and Golgi mediate concentration of soluble secretory proteins by exclusion from COPIcoated vesicles. Cell 98: 81–90, 1999.
229. MASIBAY AS, BALAJI PV, BOEGGEMAN EE, AND QASBA PK. Mutational
analysis of the Golgi retention signal of bovine beta-1,4-galactosyltransferase. J Biol Chem 268: 9908 –9916, 1993.
230. MATTER K, BRAUCHBAR M, BUCHER K, AND HAURI HP. Sorting of
endogenous plasma membrane proteins occurs from two sites in
cultured human intestinal epithelial cells (Caco-2). Cell 60: 429 –
437, 1990.
231. MATYAS GR AND MORRÉ DJ. Subcellular distribution and biosynthesis of rat liver gangliosides. Biochim Biophys Acta 921: 599 – 614,
1987.
232. MAULIK PR AND SHIPLEY GG. Interactions of N-stearoyl sphingomyelin with cholesterol and dipalmitoylphosphatidylcholine in bilayer
membranes. Biophys J 70: 2256 –2265, 1996.
233. MAYS RW, SIEMERS KA, FRITZ BA, LOWE AW, VAN MEER G, AND NELSON
WJ. Hierarchy of mechanisms involved in generating Na/K-ATPase
polarity in MDCK epithelial cells. J Cell Biol 130: 1105–1115, 1995.
234. MCCONNELL HM, WRIGHT KL, AND MCFARLAND BG. The fraction of the
lipid in a biological membrane that is in a fluid state: a spin label
assay. Biochem Biophys Res Commun 47: 273–281, 1972.
235. MCINTOSH TJ, SIMON SA, NEEDHAM D, AND HUANG C. Structure and
cohesive properties of sphingomyelin/cholesterol bilayers. Biochemistry 31: 2012–2020, 1992.
236. MCNEW JA, PARLATI F, FUKUDA R, JOHNSTON RJ, PAZ K, PAUMET F,
SOLLNER TH, AND ROTHMAN JE. Compartmental specificity of cellular
membrane fusion encoded in SNARE proteins. Nature 407: 153–
159, 2000.
237. MELKONIAN KA, OSTERMEYER AG, CHEN JZ, ROTH MG, AND BROWN DA.
Role of lipid modifications in targeting proteins to detergent-resistant membrane rafts. Many raft proteins are acylated, while few are
prenylated. J Biol Chem 274: 3910 –3917, 1999.
238. MELKONIAN M, BECKER B, AND BECKER D. Scale formation in algae. J
Electron Microsc Tech 17: 165–178, 1991.
239. MELLMAN I. Endocytosis and molecular sorting. Annu Rev Cell Dev
Biol 12: 575– 625, 1996.
240. MERRILL AH JR AND SWEELEY CC. Sphingolipids: metabolism and cell
signalling. In: Biochemistry of Lipids, Lipoproteins and Membranes, edited by Vance D and Vance JE. Amsterdam: Elsevier,
1996, p. 309 –339, 1996.
241. MICHEL C, VAN ECHTEN-DECKERT G, ROTHER J, SANDHOFF K, WANG E,
AND MERRILL AH JR. Characterization of ceramide synthesis: a dihydroceramide desaturase introduces the 4,5-trans-double bond of
sphingosine at the level of dihydroceramide. J Biol Chem 272:
22432–22437, 1997.
242. MILLER-PODRAZA H, BRADLEY RM, AND FISHMAN PH. Biosynthesis and
localization of gangliosides in cultured cells. Biochemistry 21:
3260 –3265, 1982.
243. MILLER-PODRAZA H AND FISHMAN PH. Soluble gangliosides in cultured
neurotumor cells. J Neurochem 41: 860 – 867, 1983.
244. MILLER-PRODRAZA H AND FISHMAN PH. Effect of drugs and temperature on biosynthesis and transport of glycosphingolipids in cultured neurotumor cells. Biochim Biophys Acta 804: 44 –51, 1984.
245. MIYAKE Y, KOZUTSUMI Y, NAKAMURA S, FUJITA T, AND KAWASAKI T.
Serine palmitoyltransferase is the primary target of a sphingosine-
1719
1720
268.
269.
270.
271.
273.
274.
275.
276.
277.
278.
279.
280.
281.
282.
283.
284.
285.
286.
287.
288.
289.
290.
glucosylceramide beta-1,4-galactosyltransferase from rat brain.
J Biol Chem 273: 13570 –13577, 1998.
O’BRIEN JS, SAMPSON EL, AND STERN MB. Lipid composition of
myelin from the peripheral nervous system. Intradural spinal roots.
J Neurochem 14: 357–365, 1967.
OH CS, TOKE DA, MANDALA S, AND MARTIN CE. ELO2 and ELO3,
homologues of the Saccharomyces cerevisiae ELO1 gene, function
in fatty acid elongation and are required for sphingolipid formation.
J Biol Chem 272: 17376 –17384, 1997.
ORCI L, MALHOTRA V, AMHERDT M, SERAFINI T, AND ROTHMAN JE.
Dissection of a single round of vesicular transport: sequential
intermediates for intercisternal movement in the Golgi stack. Cell
56: 357–368, 1989.
ORCI L, MONTESANO R, MEDA P, MALAISSE-LAGAE F, BROWN D, PERRELET A, AND VASSALLI P. Heterogeneous distribution of filipin-cholesterol complexes across the cisternae of the Golgi apparatus. Proc
Natl Acad Sci USA 78: 293–297, 1981.
ORCI L, RAVAZZOLA M, VOLCHUK A, ENGEL T, GMACHL M, AMHERDT M,
PERRELET A, SÖLLNER TH, AND ROTHMAN JE. Anterograde flow of
cargo across the Golgi stack potentially mediated via bidirectional
“percolating” COPI vesicles. Proc Natl Acad Sci USA 97: 10400 –
10405, 2000.
ORCI L, SCHEKMAN R, AND PERRELET A. Interleaflet clear space is
reduced in the membrane of COP I and COP II-coated buds/vesicles. Proc Natl Acad Sci USA 93: 8968 – 8970, 1996.
ORCI L, STAMNES M, RAVAZZOLA M, AMHERDT M, PERRELET A, SOLLNER
TH, AND ROTHMAN JE. Bidirectional transport by distinct populations of COPI-coated vesicles. Cell 90: 335–349, 1997.
OSTERMANN J, ORCI L, TANI K, AMHERDT M, RAVAZZOLA M, ELAZAR Z,
AND ROTHMAN JE. Stepwise assembly of functionally active transport vesicles. Cell 75: 1015–1025, 1993.
PARTON RG. Ultrastructural localization of gangliosides; GM1 is
concentrated in caveolae. J Histochem Cytochem 42: 155–166,
1994.
PARTON RG. Caveolae and caveolins. Curr Opin Cell Biol 8: 542–
548, 1996.
PASCHER I. Molecular arrangements in sphingolipids. Conformation
and hydrogen bonding of ceramide and their implication on membrane stability and permeability. Biochim Biophys Acta 455: 433–
451, 1976.
PASTOR RW, VENABLE RM, AND KARPLUS M. Model for the structure of
the lipid bilayer. Proc Natl Acad Sci USA 88: 892– 896, 1991.
PATTON JL AND LESTER RL. The phosphoinositol sphingolipids of
Saccharomyces cerevisiae are highly localized in the plasma membrane. J Bacteriol 173: 3101–3108, 1991.
PATTON JL, SRINIVASAN B, DICKSON RC, AND LESTER RL. Phenotypes of
sphingolipid-dependent strains of Saccharomyces cerevisiae. J
Bacteriol 174: 7180 –7184, 1992.
PEDRAZZINI E, VILLA A, AND BORGESE N. A mutant cytochrome b5 with
a lengthened membrane anchor escapes from the endoplasmic
reticulum and reaches the plasma membrane. Proc Natl Acad Sci
USA 93: 4207– 4212, 1996.
PELHAM H. Getting stuck in the Golgi. Traffic 1: 191–192, 2000.
PELHAM HR. Sorting and retrieval between the endoplasmic reticulum and Golgi apparatus. Curr Opin Cell Biol 7: 530 –535, 1995.
PELHAM HR. Getting through the Golgi complex. Trends Cell Biol 8:
45– 49, 1998.
PELHAM HR AND ROTHMAN JE. The debate about transport in the
Golgi. Two sides of the same coin? Cell 102: 713–719, 2000.
PLATT FM, NEISES GR, REINKENSMEIER G, TOWNSEND MJ, PERRY VH,
PROIA RL, WINCHESTER B, DWEK RA, AND BUTTERS TD. Prevention of
lysosomal storage in Tay-Sachs mice treated with N-butyldeoxynojirimycin. Science 276: 428 – 431, 1997.
POLISHCHUK RS, POLISHCHUK EV, MARRA P, ALBERTI S, BUCCIONE R,
LUINI A, AND MIRONOV AA. Correlative light-electron microscopy
reveals the tubular-saccular ultrastructure of carriers operating
between Golgi apparatus and plasma membrane. J Cell Biol 148:
45–58, 2000.
PRALLE A, KELLER P, FLORIN EL, SIMONS K, AND HORBER JK. Sphingolipid-cholesterol rafts diffuse as small entities in the plasma membrane of mammalian cells. J Cell Biol 148: 997–1008, 2000.
PRALLE A, PRUMMER M, FLORIN EL, STELZER EHK, AND HORBER JKH.
Three-dimensional high-resolution particle tracking for optical
Physiol Rev • VOL
291.
292.
293.
294.
295.
296.
297.
298.
299.
300.
301.
302.
303.
304.
305.
306.
307.
308.
309.
310.
311.
312.
tweezers by forward scattered light. Microsc Res Tech 44: 378 –386,
1999.
PUERTOLLANO R, MARTIN-BELMONTE F, MILLAN J, DE MARCO MC, ALBAR
JP, KREMER L, AND ALONSO MA. The MAL proteolipid is necessary for
normal apical transport and accurate sorting of the influenza virus
hemagglutinin in Madin-Darby canine kidney cells. J Cell Biol 145:
141–151, 1999.
PUOTI A, DESPONDS C, AND CONZELMANN A. Biosynthesis of mannosylinositolphosphoceramide in Saccharomyces cerevisiae is dependent on genes controlling the flow of secretory vesicles from
the endoplasmic reticulum to the Golgi. J Cell Biol 113: 515–525,
1991.
PURI V, WATANABE R, DOMINGUEZ M, SUN X, WHEATLEY CL, MARKS DL,
AND PAGANO RE. Cholesterol modulates membrane traffic along the
endocytic pathway in sphingolipid-storage diseases. Nature Cell
Biol 1: 386 –388, 1999.
RADHAKRISHNAN A, ANDERSON TG, AND MCCONNELL HM. Condensed
complexes, rafts, and the chemical activity of cholesterol in membranes. Proc Natl Acad Sci USA 97: 12422–12427, 2000.
RAGGERS RJ, POMORSKI T, HOLTHUIS JCM, KAELIN N, AND VAN MEER G.
Lipid traffic: the ABC of transbilayer movement. Traffic 1: 226 –234,
2000.
RAMSTEDT B AND SLOTTE JP. Interaction of cholesterol with sphingomyelins and acyl-chain-matched phosphatidylcholines: a comparative study of the effect of the chain length. Biophys J 76:
908 –915, 1999.
RAWYLER AJ, ROELOFSEN B, OP DEN KAMP JAF, AND VAN DEENEN LLM.
Isolation and characterization of plasma membranes from Friend
erythroleukaemic cells. A study with sphingomyelinase C. Biochim
Biophys Acta 730: 130 –138, 1983.
RAYNER JC AND PELHAM HR. Transmembrane domain-dependent
sorting of proteins to the ER and plasma membrane in yeast. EMBO
J 16: 1832–1841, 1997.
RECKTENWALD DJ AND MCCONNELL HM. Phase equilibria in binary
mixtures of phosphatidylcholine and cholesterol. Biochemistry 20:
4505– 4510, 1981.
REISS-HUSSON F. Structure of liquid-crystalline phases of different
phospholipids, monoglycerides, sphingolipids in the absence or
presence of water. J Mol Biol 25: 363–382, 1967.
REITZ RC, THOMPSON JA, AND MORRIS HP. Mitochondrial and microsomal phospholipids of Morris hepatoma 7777. Cancer Res 37:
561–567, 1977.
RENKONEN O, GAHMBERG CG, SIMONS K, AND KÄÄRIÄINEN L. The lipids
of the plasma membranes and endoplasmic reticulum from cultured baby hamster kidney cells (BHK21). Biochim Biophys Acta
255: 66 –78, 1972.
REVARDEL E, BONNEAU M, DURRENS P, AND AIGLE M. Characterization
of a new gene family developing pleiotropic phenotypes upon
mutation in Saccharomyces cerevisiae. Biochim Biophys Acta
1263: 261–265, 1995.
RIETVELD A, NEUTZ S, SIMONS K, AND EATON S. Association of steroland glycosylphosphatidylinositol-linked proteins with Drosophila
raft lipid microdomains. J Biol Chem 274: 12049 –12054, 1999.
RIETVELD A AND SIMONS K. The differential miscibility of lipids as the
basis for the formation of functional membrane rafts. Biochim
Biophys Acta 1376: 467– 479, 1998.
RODRIGUEZ-BOULAN E AND GONZALEZ A. Glycans in post-Golgi apical
targeting: sorting signals or structural props? Trends Cell Biol 9:
291–294, 1999.
ROPER K, CORBEIL D, AND HUTTNER WB. Retention of prominin in
microvilli reveals distinct cholesterol-based lipid micro-domains in
the apical plasma membrane. Nature Cell Biol 2: 582–592, 2000.
ROSALES FRITZ VM AND MACCIONI HJ. Effects of brefeldin A on
synthesis and intracellular transport of ganglioside GT3 by chick
embryo retina cells. J Neurochem 65: 1859 –1864, 1995.
ROTHMAN JE. Mechanism of intracellular protein transport. Nature
372: 55– 63, 1994.
ROTHMAN JE. The protein machinery of vesicle budding and fusion.
Protein Sci 5: 185–194, 1996.
ROTHMAN JE AND SÖLLNER TH. Throttles and dampers: controlling
the engine of membrane fusion. Science 276: 1212–1213, 1997.
ROTHMAN JE AND WIELAND FT. Protein sorting by transport vesicles.
Science 272: 227–234, 1996.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
272.
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
Physiol Rev • VOL
333.
334.
335.
336.
337.
338.
339.
340.
341.
342.
343.
344.
345.
346.
347.
348.
349.
350.
351.
352.
353.
Wallerian degeneration and myelination defects. Proc Natl Acad
Sci USA 96: 7532–7537, 1999.
SHERWOOD AL AND HOLMES EH. Brefeldin A induced inhibition of de
novo globo- and neolacto-series glycolipid core chain biosynthesis
in human cells. Evidence for an effect on beta 134 galactosyltransferase activity. J Biol Chem 267: 25328 –25336, 1992.
SHIAO YJ AND VANCE JE. Sphingomyelin transport to the cell surface
occurs independently of protein secretion in rat hepatocytes. J Biol
Chem 268: 26085–26092, 1993.
SHOGOMORI H, BURACK MA, BANKER G, AND FUTERMAN AH. Gangliosides GM1 and GD1b are not polarized in mature hippocampal
neurons. FEBS Lett 458: 107–111, 1999.
SILLENCE DJ, RAGGERS RJ, AND VAN MEER G. Assays for transmembrane movement of sphingolipids. Methods Enzymol 312: 562–579,
2000.
SILVIUS JR. Cholesterol modulation of lipid intermixing in phospholipid and glycosphingolipid mixtures. Evaluation using fluorescent
lipid probes and brominated lipid quenchers. Biochemistry 31:
3398 –3408, 1992.
SILVIUS JR, DEL GIUDICE D, AND LAFLEUR M. Cholesterol at different
bilayer concentrations can promote or antagonize lateral segregation of phospholipids of differing acyl chain length. Biochemistry
35: 15198 –15208, 1996.
SIMONS K AND IKONEN E. Functional rafts in cell membranes. Nature
387: 569 –572, 1997.
SIMONS K AND TOOMRE D. Lipid rafts and signal transduction. Nature
Rev Mol Cell Biol 1: 31–39, 2000.
SIMONS K AND VAN MEER G. Lipid sorting in epithelial cells. Biochemistry 27: 6197– 6202, 1988.
SIMONS M, KRAMER EM, THIELE C, STOFFEL W, AND TROTTER J. Assembly of myelin by association of proteolipid protein with cholesteroland galactosylceramide-rich membrane domains. J Cell Biol 151:
143–154, 2000.
SINGER SJ AND NICOLSON GL. The fluid mosaic model of the structure
of cell membranes. Science 175: 720 –731, 1972.
SKIBBENS JE, ROTH MG, AND MATLIN KS. Differential extractability of
influenza virus hemagglutinin during intracellular transport in polarized epithelial cells and non-polar fibroblasts. J Cell Biol 108:
821– 832, 1989.
SKRZYPEK M, LESTER RL, AND DICKSON RC. Suppressor gene analysis
reveals an essential role for sphingolipids in transport of glycosylphosphatidylinositol-anchored proteins in Saccharomyces cerevisiae. J Bacteriol 179: 1513–1520, 1997.
SKRZYPEK MS, NAGIEC MM, LESTER RL, AND DICKSON RC. Analysis of
phosphorylated sphingolipid long-chain bases reveals potential
roles in heat stress and growth control in Saccharomyces. J Bacteriol 181: 1134 –1140, 1999.
SLOTTE JP, HÄRMÄLÄ AS, JANSSON C, AND PÖRN MI. Rapid turn-over of
plasma membrane sphingomyelin and cholesterol in baby hamster
kidney cells after exposure to sphingomyelinase. Biochim Biophys
Acta 1030: 251–257, 1990.
SMART EJ, YING YS, MINEO C, AND ANDERSON RGW. A detergent-free
method for purifying caveolae membrane from tissue culture cells.
Proc Natl Acad Sci USA 92: 10104 –10108, 1995.
SOMERHARJU P, VIRTANEN JA, AND CHENG KH. Lateral organisation of
membrane lipids. The superlattice view. Biochim Biophys Acta
1440: 32– 48, 1999.
SOMERHARJU PJ, VIRTANEN JA, EKLUND KK, VAINIO P, AND KINNUNEN
PK. 1-Palmitoyl-2-pyrenedecanoyl glycerophospholipids as membrane probes: evidence for regular distribution in liquid-crystalline
phosphatidylcholine bilayers. Biochemistry 24: 2773–2781, 1985.
SONG KS, LI S, OKAMOTO T, QUILLIAM LA, SARGIACOMO M, AND LISANTI
MP. Co-purification and direct interaction of Ras with caveolin, an
integral membrane protein of caveolae microdomains. Detergentfree purification of caveolae membranes. J Biol Chem 271: 9690 –
9697, 1996.
SONNINO S, GHIDONI R, FIORILLI A, VENERANDO B, AND TETTAMANTI G.
Cytosolic gangliosides of rat brain: their fractionation into proteinbound complexes of different ganglioside compositions. J Neurosci Res 12: 193–204, 1984.
SONNINO S, GHIDONI R, MARCHESINI S, AND TETTAMANTI G. Cytosolic
gangliosides: occurrence in calf brain as ganglioside-protein complexes. J Neurochem 33: 117–121, 1979.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
313. ROUSH DL, GOTTARDI CJ, NAIM HY, ROTH MG, AND CAPLAN MJ. Tyrosine-based membrane protein sorting signals are differentially
interpreted by polarized Madin-Darby canine kidney and LLC-PK1
epithelial cells. J Biol Chem 273: 26862–26869, 1998.
314. ROWE T, DASCHER C, BANNYKH S, PLUTNER H, AND BALCH WE. Role of
vesicle-associated syntaxin 5 in the assembly of pre-Golgi intermediates. Science 279: 696 –700, 1998.
315. SABA JD, NARA F, BIELAWSKA A, GARRETT S, AND HANNUN YA. The
BST1 gene of Saccharomyces cerevisiae is the sphingosine-1-phosphate lyase. J Biol Chem 272: 26087–26090, 1997.
316. SAITO M, SAITO M, AND ROSENBERG A. Action of monensin, a monovalent cationophore, on cultured human fibroblasts: evidence that
it induces high cellular accumulation of glucosyl- and lactosylceramide (gluco- and lactocerebroside). Biochemistry 23: 1043–1046,
1984.
317. SAITO M, SAITO M, AND ROSENBERG A. Influence of monovalent cation
transport on anabolism of glycosphingolipids in cultured human
fibroblasts. Biochemistry 24: 3054 –3059, 1985.
318. SAKAKIBARA K, MOMOI T, UCHIDA T, AND NAGAI Y. Evidence for association of glycosphingolipid with a colchicine-sensitive microtubule-like cytoskeletal structure of cultured cells. Nature 293: 76 –
79, 1981.
319. SANKARAM MB AND THOMPSON TE. Cholesterol-induced fluid-phase
immiscibility in membranes. Proc Natl Acad Sci USA 88: 8686 –
8690, 1991.
320. SAWAI H, OKAMOTO Y, LUBERTO C, MAO C, BIELAWSKA A, DOMAE N, AND
HANNUN YA. Identification of ISC1 (YER019w) as inositol phosphosphingolipid phospholipase C in Saccharomyces cerevisiae.
J Biol Chem 275: 39793–39798, 2000.
321. SCHEIFFELE P, PERÄNEN J, AND SIMONS K. N-glycans as apical sorting
signals in epithelial cells. Nature 378: 96 –98, 1995.
322. SCHEIFFELE P, ROTH MG, AND SIMONS K. Interaction of influenza virus
haemagglutinin with sphingolipid-cholesterol membrane domains
via its transmembrane domain. EMBO J 16: 5501–5508, 1997.
323. SCHIFFMANN R, MURRAY GJ, TRECO D, DANIEL P, SELLOS-MOURA M,
MYERS M, QUIRK JM, ZIRZOW GC, BOROWSKI M, LOVEDAY K, ANDERSON
T, GILLESPIE F, OLIVER KL, JEFFRIES NO, DOO E, LIANG TJ, KREPS C,
GUNTER K, FREI K, CRUTCHFIELD K, SELDEN RF, AND BRADY RO.
Infusion of alpha-galactosidase A reduces tissue globotriaosylceramide storage in patients with fabry disease. Proc Natl Acad Sci
USA 97: 365–370, 2000.
324. SCHMIDT CF, BARENHOLZ Y, HUANG C, AND THOMPSON TE. Monolayer
coupling in sphingomyelin bilayer systems. Nature 271: 775–777,
1978.
325. SCHNEITER R, BRUGGER B, SANDHOFF R, ZELLNIG G, LEBER A, LAMPL M,
ATHENSTAEDT K, HRASTNIK C, EDER S, DAUM G, PALTAUF F, WIELAND
FT, AND KOHLWEIN SD. Electrospray ionization tandem mass spectrometry (ESI-MS/MS) analysis of the lipid molecular species composition of yeast subcellular membranes reveals acyl chain-based
sorting/remodeling of distinct molecular species en route to the
plasma membrane. J Cell Biol 146: 741–754, 1999.
326. SCHNITZER JE, MCINTOSH DP, DVORAK AM, LIU J, AND OH P. Separation
of caveolae from associated microdomains of GPI-anchored proteins. Science 269: 1435–1439, 1995.
327. SCHUCHMAN EH, SUCHI M, TAKAHASHI T, SANDHOFF K, AND DESNICK RJ.
Human acid sphingomyelinase. Isolation, nucleotide sequence and
expression of the full-length and alternatively spliced cDNAs.
J Biol Chem 266: 8531– 8539, 1991.
328. SCHULTE S AND STOFFEL W. Ceramide UDPgalactosyltransferase
from myelinating rat brain: purification, cloning, and expression.
Proc Natl Acad Sci USA 90: 10265–10269, 1993.
329. SCHÜTZ GJ, KADA G, PASTUSHENKO VP, AND SCHINDLER H. Properties of
lipid microdomains in a muscle cell membrane visualized by single
molecule microscopy. EMBO J 19: 892–901, 2000.
330. SCHWERTZ DW, KREISBERG JI, AND VENKATACHALAM MA. Characterization of rat kidney proximal tubule brush border membrane-associated phosphatidylinositol phosphodiesterase. Arch Biochem Biophys 224: 555–567, 1983.
331. SHEETS ED, LEE GM, SIMSON R, AND JACOBSON K. Transient confinement of a glycosylphosphatidylinositol-anchored protein in the
plasma membrane. Biochemistry 36: 12449 –12458, 1997.
332. SHEIKH KA, SUN J, LIU Y, KAWAI H, CRAWFORD TO, PROIA RL, GRIFFIN
JW, AND SCHNAAR RL. Mice lacking complex gangliosides develop
1721
1722
HOLTHUIS, POMORSKI, RAGGERS, SPRONG, AND VAN MEER
Physiol Rev • VOL
376.
377.
378.
379.
380.
381.
382.
383.
384.
385.
386.
387.
388.
389.
390.
391.
392.
393.
394.
lipids in phosphatidylcholine bilayers and biological membranes.
In: Enzymes of Lipid Metabolism, edited by Freysz L. New York:
Plenum, 1986, p. 387–396.
THUDICHUM JLW. A Treatise on the Chemical Constitution of
Brain. London: Balliere, Tindall, and Cox, 1884.
TILLACK TW, ALLIETTA M, MORAN RE, AND YOUNG WW JR. Localization
of globoside and Forssman glycolipids on erythrocyte membranes.
Biochim Biophys Acta 733: 15–24, 1983.
TOMIUK S, HOFMANN K, NIX M, ZUMBANSEN M, AND STOFFEL W. Cloned
mammalian neutral sphingomyelinase. Functions in sphingolipid
signaling. Proc Natl Acad Sci USA 95: 3638 –3643, 1998.
TOOMRE D, KELLER P, WHITE J, OLIVO JC, AND SIMONS K. Dual-color
visualization of trans-Golgi network to plasma membrane traffic
along microtubules in living cells. J Cell Sci 112: 21–33, 1999.
TRINCHERA M AND GHIDONI R. Two glycosphingolipid sialyltransferases are localized in different sub-Golgi compartments in rat
liver. J Biol Chem 264: 15766 –15769, 1989.
TRINCHERA M, PIROVANO B, AND GHIDONI R. Sub-Golgi distribution in
rat liver of CMP-NeuAc GM3- and CMP-NeuAc:GT1b␣238 sialyltransferases and comparison with the distribution of the other
glycosyltransferase activities involved in ganglioside biosynthesis.
J Biol Chem 265: 18242–18247, 1990.
TVRDIK P, ASADI A, KOZAK LP, NEDERGAARD J, CANNON B, AND JACOBSSON A. Cig30, a mouse member of a novel membrane protein gene
family, is involved in the recruitment of brown adipose tissue.
J Biol Chem 272: 31738 –31746, 1997.
TVRDIK P, WESTERBERG R, SILVE S, ASADI A, JAKOBSSON A, CANNON B,
LOISON G, AND JACOBSSON A. Role of a new mammalian gene family
in the biosynthesis of very long chain fatty acids and sphingolipids.
J Cell Biol 149: 707–718, 2000.
VAN ECHTEN G, IBER H, STOTZ H, TAKATSUKI A, AND SANDHOFF K.
Uncoupling of ganglioside biosynthesis by Brefeldin A. Eur J Cell
Biol 51: 135–139, 1990.
VAN ECHTEN G AND SANDHOFF K. Modulation of ganglioside biosynthesis in primary cultured neurons. J Neurochem 52: 207–214, 1989.
VAN GENDEREN I AND VAN MEER G. Differential targeting of glucosylceramide and galactosylceramide analogues after synthesis but not
during transcytosis in Madin-Darby canine kidney cells. J Cell Biol
131: 645– 654, 1995.
VAN GENDEREN IL, VAN MEER G, SLOT JW, GEUZE HJ, AND VOORHOUT
WF. Subcellular localization of Forssman glycolipid in epithelial
MDCK cells by immuno-electronmicroscopy after freeze-substitution. J Cell Biol 115: 1009 –1019, 1991.
VAN HELVOORT A, DE BROUWER A, OTTENHOFF R, BROUWERS JF, WIJNHOLDS J, BEIJNEN JH, RIJNEVELD A, VAN DER POLL T, VAN DER VALK MA,
MAJOOR D, VOORHOUT W, WIRTZ KW, ELFERINK RP, AND BORST P. Mice
without phosphatidylcholine transfer protein have no defects in the
secretion of phosphatidylcholine into bile or into lung airspaces.
Proc Natl Acad Sci USA 96: 11501–11506, 1999.
VAN HELVOORT A, GIUDICI ML, THIELEMANS M, AND VAN MEER G.
Transport of sphingomyelin to the cell surface is inhibited by
brefeldin A and in mitosis, where C6-NBD-sphingomyelin is translocated across the plasma membrane by a multidrug transporter
activity. J Cell Sci 110: 75– 83, 1997.
VAN HELVOORT A, SMITH AJ, SPRONG H, FRITZSCHE I, SCHINKEL AH,
BORST P, AND VAN MEER G. MDR1 P-glycoprotein is a lipid translocase of broad specificity, while MDR3 P-glycoprotein specifically
translocates phosphatidylcholine. Cell 87: 507–517, 1996.
VAN HELVOORT A, STOORVOGEL W, VAN MEER G, AND BURGER KNJ.
Sphingomyelin synthase is absent from endosomes. J Cell Sci 110:
781–788, 1997.
VAN HELVOORT A, VAN’T HOF W, RITSEMA T, SANDRA A, AND VAN MEER
G. Conversion of diacylglycerol to phosphatidylcholine on the basolateral surface of epithelial (Madin-Darby canine kidney) cells.
Evidence for the reverse action of a sphingomyelin synthase. J Biol
Chem 269: 1763–1769, 1994.
VAN IJZENDOORN SCD AND HOEKSTRA D. (Glyco)sphingolipids are
sorted in sub-apical compartments in HepG2 cells: a role for nonGolgi-related intracellular sites in the polarized distribution of (glyco)sphingolipids. J Cell Biol 142: 683– 696, 1998.
VAN IJZENDOORN SCD, ZEGERS MMP, KOK JW, AND HOEKSTRA D. Segregation of glucosylceramide and sphingomyelin occurs in the
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
354. SORICE M, PAROLINI I, SANSOLINI T, GAROFALO T, DOLO V, SARGIACOMO
M, TAI T, PESCHLE C, TORRISI MR, AND PAVAN A. Evidence for the
existence of ganglioside-enriched plasma membrane domains in
human peripheral lymphocytes. J Lipid Res 38: 969 –980, 1997.
355. SPANG A, MATSUOKA K, HAMAMOTO S, SCHEKMAN R, AND ORCI L.
Coatomer, Arf1p, and nucleotide are required to bud coat protein
complex I-coated vesicles from large synthetic liposomes. Proc
Natl Acad Sci USA 95: 11199 –11204, 1998.
356. SPIEGEL S. Sphingosine 1-phosphate: a prototype of a new class of
second messengers. J Leukocyte Biol 65: 341–344, 1999.
357. SPIEGEL S, KASSIS S, WILCHEK M, AND FISHMAN PH. Direct visualization
of redistribution and capping of fluorescent gangliosides on lymphocytes. J Cell Biol 99: 1575–1581, 1984.
358. SPIEGEL S, MATYAS GR, CHENG L, AND SACKTOR B. Asymmetric distribution of gangliosides in rat renal brush-border and basolateral
membranes. Biochim Biophys Acta 938: 270 –278, 1988.
359. SPRONG H, KRUITHOF B, LEIJENDEKKER R, SLOT JW, VAN MEER G, AND
VAN DER SLUIJS P. UDP-galactose:ceramide galactosyltransferase is a
class I integral membrane protein of the endoplasmic reticulum.
J Biol Chem 273: 25880 –25888, 1998.
360. STAHL N, JUREVICS H, MORELL P, SUZUKI K, AND POPKO B. Isolation,
characterization, and expression of cDNA clones that encode rat
UDP-galactose: ceramide galactosyltransferase. J Neurosci Res 38:
234 –242, 1994.
361. STAN RV, ROBERTS WG, PREDESCU D, IHIDA K, SAUCAN L, GHITESCU L,
AND PALADE GE. Immunoisolation and partial characterization of
endothelial plasmalemmal vesicles (caveolae). Mol Biol Cell 8:
595– 605, 1997.
362. STARLING AP, EAST JM, AND LEE AG. Effects of phosphatidylcholine
fatty acyl chain length on calcium binding and other functions of
the (Ca2⫹-Mg2⫹)-ATPase. Biochemistry 32: 1593–1600, 1993.
363. STEIM JM, TOURTELLOTTE ME, REINERT JC, MCELHANEY RN, AND RADER
RL. Calorimetric evidence for the liquid-crystalline state of lipids in
a biomembrane. Proc Natl Acad Sci USA 63: 104 –109, 1969.
364. STIER A AND SACKMANN E. Spin labels as enzyme substrates. Heterogeneous lipid distribution in liver microsomal membranes. Biochim Biophys Acta 311: 400 – 408, 1973.
365. STOFFEL W, ANDERSON R, AND STAHL J. Studies on the asymmetric
arrangement of membrane-lipid-enveloped virions as a model system. Hoppe-Seylers Z Physiol Chem 356: 1123–1129, 1975.
366. STOFFEL W AND BOSIO A. Myelin glycolipids and their functions. Curr
Opin Neurobiol 7: 654 – 661, 1997.
367. STOFFEL W AND SORGO W. Asymmetry of the lipid-bilayer of Sindbis
virus. Chem Phys Lipids 17: 324 –335, 1976.
368. STORRIE B, WHITE J, ROTTGER S, STELZER EHK, SUGANUMA T, AND
NILSSON T. Recycling of Golgi-resident glycosyltransferases through
the ER reveals a novel pathway and provides an explanation for
nocodazole-induced Golgi scattering. J Cell Biol 143: 1505–1521,
1998.
369. STUBBS CD, KETTERER B, AND HICKS RM. The isolation and analysis of
the luminal plasma membrane of calf urinary bladder epithelium.
Biochim Biophys Acta 558: 58 –72, 1979.
370. SUZUKI K, STERBA RE, AND SHEETZ MP. Outer membrane monolayer
domains from two-dimensional surface scanning resistance measurements. Biophys J 79: 448 – 459, 2000.
371. TAKAMIYA K, YAMAMOTO A, FURUKAWA K, YAMASHIRO S, SHIN M, OKADA
M, FUKUMOTO S, HARAGUCHI M, TAKEDA N, FUJIMURA K, SAKAE M,
KISHIKAWA M, SHIKU H, FURUKAWA K, AND AIZAWA S. Mice with disrupted GM2/GD2 synthase gene lack complex gangliosides but
exhibit only subtle defects in their nervous system. Proc Natl Acad
Sci USA 93: 10662–10667, 1996.
372. TANG D, AND CHONG PL-G. E/M dips: evidence for lipids regularly
distributed into hexagonal super-lattices in pyrene-PC/DMPC binary mixtures at specific concentrations. Biophys J 63: 903–910,
1992.
373. TENNEKOON G, ZARUBA M, AND WOLINSKY J. Topography of cerebroside sulfotransferase in Golgi-enriched vesicles from rat brain.
J Cell Biol 97: 1107–1112, 1983.
374. THIELE C AND HUTTNER WB. Protein and lipid sorting from the
trans-Golgi network to secretory granules. Recent developments.
Semin Cell Dev Biol 9: 511–516, 1998.
375. THOMPSON TE, BARENHOLZ Y, BROWN RE, CORREA-FREIRE M, YOUNG
WW JR, AND TILLACK TW. Molecular organization of glycosphingo-
ORGANIZING POTENTIAL OF SPHINGOLIPIDS
395.
396.
397.
398.
399.
400.
402.
403.
404.
405.
406.
407.
408.
409.
410.
411.
412.
413.
414.
415.
416.
Physiol Rev • VOL
417. WIRTZ KW AND ZILVERSMIT DB. Participation of soluble liver proteins
in the exchange of membrane phospholipids. Biochim Biophys
Acta 193: 105–116, 1969.
418. WOODING S AND PELHAM HR. The dynamics of Golgi protein traffic
visualized in living yeast cells. Mol Biol Cell 9: 2667–2680, 1998.
419. XIAO Z AND DEVREOTES PN. Identification of detergent-resistant
plasma membrane microdomains in Dictyostelium: enrichment of
signal transduction proteins. Mol Biol Cell 8: 855– 869, 1997.
420. YAMASHITA K, HARA-KUGE S, AND OHKURA T. Intracellular lectins
associated with N-linked glycoprotein traffic. Biochim Biophys
Acta 1473: 147–160, 1999.
421. YAMASHITA T, WADA R, SASAKI T, DENG C, BIERFREUND U, SANDHOFF K,
AND PROIA RL. A vital role for glycosphingolipid synthesis during
development and differentiation. Proc Natl Acad Sci USA 96: 9142–
9147, 1999.
422. YANG M, ELLENBERG J, BONIFACINO JS, AND WEISSMAN AM. The transmembrane domain of a carboxyl-terminal anchored protein determines localization to the endoplasmic reticulum. J Biol Chem 272:
1970 –1975, 1997.
423. YEAMAN C, GRINDSTAFF KK, AND NELSON WJ. New perspectives on
mechanisms involved in generating epithelial cell polarity. Physiol
Rev 79: 73–98, 1999.
424. YOKODE M, PATHAK RK, HAMMER RE, BROWN MS, GOLDSTEIN JL, AND
ANDERSON RG. Cytoplasmic sequence required for basolateral targeting of LDL receptor in livers of transgenic mice. J Cell Biol 117:
39 – 46, 1992.
425. YOSHIMORI T, KELLER P, ROTH MG, AND SIMONS K. Different biosynthetic transport routes to the plasma membrane in BHK and CHO
cells. J Cell Biol 133: 247–256, 1996.
426. YOUNG WW JR, ALLENDE ML, AND JASKIEWICZ E. Reevaluating the
effect of brefeldin A (BFA) on ganglioside synthesis: the location of
GM2 synthase cannot be deduced from the inhibition of GM2
synthesis by BFA. Glycobiology 9: 689 – 695, 1999.
427. YOUNG WW JR, LUTZ MS, MILLS SE, AND LECHLER-OSBORN S. Use of
brefeldin A to define sites of glycosphingolipid synthesis: GA2/
GM2/GD2 synthase is trans to the brefeldin A block. Proc Natl Acad
Sci USA 87: 6838 – 6842, 1990.
428. ZAAL KJ, SMITH CL, POLISHCHUK RS, ALTAN N, COLE NB, ELLENBERG J,
HIRSCHBERG K, PRESLEY JF, ROBERTS TH, SIGGIA E, PHAIR RD, AND
LIPPINCOTT-SCHWARTZ J. Golgi membranes are absorbed into and
reemerge from the ER during mitosis. Cell 99: 589 – 601, 1999.
429. ZANOLARI B, FRIANT S, FUNATO K, SUTTERLIN C, STEVENSON BJ, AND
RIEZMAN H. Sphingoid base synthesis requirement for endocytosis
in Saccharomyces cerevisiae. EMBO J 19: 2824 –2833, 2000.
430. ZHANG X AND THOMPSON GA JR. An apparent association between
glycosylphosphatidylinositol-anchored proteins and a sphingolipid
in Tetrahymena mimbres. Biochem J 323: 197–206, 1997.
431. ZHANG Y, YAO B, DELIKAT S, BAYOUMY S, LIN XH, BASU S, MCGINLEY M,
CHAN-HUI PY, LICHENSTEIN H, AND KOLESNICK R. Kinase suppressor of
Ras is ceramide-activated protein kinase. Cell 89: 63–72, 1997.
432. ZHAO C, BEELER T, AND DUNN T. Suppressors of the Ca2⫹-sensitive
yeast mutant (csg2) identify genes involved in sphingolipid biosynthesis. Cloning and characterization of SCS1, a gene required for
serine palmitoyltransferase activity. J Biol Chem 269: 21480 –21488,
1994.
433. ZHOU J AND SABA JD. Identification of the first mammalian sphingosine phosphate lyase gene and its functional expression in yeast.
Biochem Biophys Res Commun 242: 502–507, 1998.
434. ZHOU Q, ZHAO J, STOUT JG, LUHM RA, WIEDMER T, AND SIMS PJ.
Molecular cloning of human plasma membrane phospholipid
scramblase. A protein mediating transbilayer movement of plasma
membrane phospholipids. J Biol Chem 272: 18240 –18244, 1997.
435. ZIEGLER RJ, YEW NS, LI C, CHERRY M, BERTHELETTE P, ROMANCZUK H,
IOANNOU YA, ZEIDNER KM, DESNICK RJ, AND CHENG SH. Correction of
enzymatic and lysosomal storage defects in Fabry mice by adenovirus-mediated gene transfer. Hum Gene Ther 10: 1667–1682, 1999.
436. ZWEERINK MM, EDISON AM, WELLS GB, PINTO W, AND LESTER RL.
Characterization of a novel, potent, and specific inhibitor of serine
palmitoyltransferase. J Biol Chem 267: 25032–25038, 1992.
81 • OCTOBER 2001 •
www.prv.org
Downloaded from http://physrev.physiology.org/ by 10.220.32.246 on July 4, 2017
401.
apical to basolateral transcytotic route in HepG2 cells. J Cell Biol
137: 347–357, 1997.
VAN MEER G. Lipid traffic in animal cells. Annu Rev Cell Biol 5:
247–275, 1989.
VAN MEER G. Lipids of the Golgi membrane. Trends Cell Biol 8:
29 –33, 1998.
VAN MEER G AND HOLTHUIS JCM. Sphingolipid transport in eukaryotic cells. Biochim Biophys Acta 1486: 145–170, 2000.
VAN MEER G, STELZER EHK, WIJNAENDTS-VAN-RESANDT RW, AND SIMONS K. Sorting of sphingolipids in epithelial (Madin-Darby canine
kidney) cells. J Cell Biol 105: 1623–1635, 1987.
VAN VELDHOVEN PP AND MANNAERTS GP. Sphinganine 1-phosphate
metabolism in cultured skin fibroblasts: evidence for the existence
of a sphingosine phosphatase. Biochem J 299: 597– 601, 1994.
VAN WEELY S, BRANDSMA M, STRIJLAND A, TAGER JM, AND AERTS JM.
Demonstration of the existence of a second, non-lysosomal glucocerebrosidase that is not deficient in Gaucher disease. Biochim
Biophys Acta 1181: 55– 62, 1993.
VARMA R AND MAYOR S. GPI-anchored proteins are organized in
submicron domains at the cell surface. Nature 394: 798 – 801, 1998.
VENIEN C AND LE GRIMELLEC C. Phospholipid asymmetry in renal
brush-border membranes. Biochim Biophys Acta 942: 159 –168,
1988.
VERKLEIJ AJ, VERVERGAERT PHJT, VAN DEENEN LLM, AND ELBERS PF.
Phase transitions of phospholipid bilayers and membranes of Acholeplasma laidlawii B visualized by freeze fracturing electron microscopy. Biochim Biophys Acta 288: 326 –332, 1972.
VERKLEIJ AJ, ZWAAL RFA, ROELOFSEN B, COMFURIUS P, KASTELIJN D,
AND VAN DEENEN LLM. The asymmetric distribution of phospholipids in the human red cell membrane. A combined study using
phospholipases and freeze-etch electron microscopy. Biochim
Biophys Acta 323: 178 –193, 1973.
WANG E, NORRED WP, BACON CW, RILEY RT, AND MERRILL AH JR.
Inhibition of sphingolipid biosynthesis by fumonisins. Implications
for diseases associated with Fusarium moniliforme. J Biol Chem
266: 14486 –14490, 1991.
WARNOCK DE, LUTZ MS, BLACKBURN WA, YOUNG WW JR, AND BAENZIGER JU. Transport of newly synthesized glucosylceramide to the
plasma membrane by a non-Golgi pathway. Proc Natl Acad Sci USA
91: 2708 –2712, 1994.
WATANABE R, ASAKURA K, RODRIGUEZ M, AND PAGANO RE. Internalization and sorting of plasma membrane sphingolipid analogues in
differentiating oligodendrocytes. J Neurochem 73: 1375–1383, 1999.
WATTENBERG BW AND SILBERT DF. Sterol partitioning among intracellular membranes. Testing a model for cellular sterol distribution. J Biol Chem 258: 2284 –2289, 1983.
WATTS JD, AEBERSOLD R, POLVERINO AJ, PATTERSON SD, AND GU M.
Ceramide second messengers and ceramide assays. Trends Biochem Sci 24: 228, 1999.
WEIDMAN PJ. Anterograde transport through the Golgi complex: do
Golgi tubules hold the key? Trends Cell Biol 5: 302–305, 1995.
WELLS GB AND LESTER RL. The isolation and characterization of a
mutant strain of Saccharomyces cerevisiae that requires a long
chain base for growth and for synthesis of phosphosphingolipids.
J Biol Chem 258: 10200 –10203, 1983.
WERTZ PW AND DOWNING DT. Covalently bound omega-hydroxyacylsphingosine in the stratum corneum. Biochim Biophys Acta 917:
108 –111, 1987.
WESSELS HP, HANSEN GH, FUHRER C, LOOK AT, SJÖSTRÖM H, NORÉN O,
AND SPIESS M. Aminopeptidase N is directly sorted to the apical
domain in MDCK cells. J Cell Biol 111: 2923–2930, 1990.
WIEGANDT H. Insect glycolipids. Biochim Biophys Acta 1123: 117–
126, 1992.
WIELAND FT, GLEASON ML, SERAFINI TA, AND ROTHMAN JE. The rate of
bulk flow from the endoplasmic reticulum to the cell surface. Cell
50: 289 –300, 1987.
WINCKLER B AND MELLMAN I. Neuronal polarity: controlling the sorting and diffusion of membrane components. Neuron 23: 637– 640,
1999.
1723