Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Feb 1.
Published in final edited form as: Prostaglandins Other Lipid Mediat. 2007 Oct 13;85(1-2):1–16. doi: 10.1016/j.prostaglandins.2007.10.002

Sphingolipids and Membrane Biology as Determined from Genetic Models

Raghavendra Pralhada Rao 1, Jairaj K Acharya 1,*
PMCID: PMC2242731  NIHMSID: NIHMS39520  PMID: 18035569

Abstract

The importance of sphingolipids in membrane biology was appreciated early in the twentieth century when several human inborn errors of metabolism were linked to defects in sphingolipid degradation. The past two decades have seen an explosion of information linking sphingolipids with cellular processes. Studies have unraveled mechanistic details of the sphingolipid metabolic pathways, and these findings are being exploited in the development of novel therapies, some now in clinical trials. Pioneering work in yeast has laid the foundation for identifying genes encoding the enzymes of the pathways. The advent of the era of genomics and bioinformatics has led to the identification of homologous genes in other species and the subsequent creation of animal knock-out lines for these genes. Discoveries from these efforts have re-kindled interest in the role of sphingolipids in membrane biology. This review highlights some of the recent advances in understanding sphingolipids’ roles in membrane biology as determined from genetic models.

Introduction

Sphingolipids are essential components of all eukaryotic cell membranes, and are required for the survival of Drosophila, yeast, and mammals [1-4]. The biological and clinical importance of sphingolipid metabolism was revealed by the discovery that diseases such as Gaucher’s disease and Niemann-Pick were caused by defective sphingolipid degradation. However, the lack of genetic model systems that would permit molecular analyses impaired rapid progress in understanding the role of sphingolipids in the etiopathogenesis of these syndromes. In the mean-time, however, in vitro studies began to provide convincing evidence that sphingolipid metabolites function as second messengers [5-9] and that they influence signal-transduction events through their ability to segregate the plasma membrane into microdomains [10-12]. In addition, complex sphingolipids were shown to act as receptors to permit the entry of microorganisms into the cell and to alter the sensitivity of cells to cell-surface ligands [13-15]. Sphingolipid metabolites also facilitate cellular processes such as cell division, differentiation, and cell death [2, 16-18]. The relative concentrations of sphingolipid metabolites are thought to be regulated tightly, because each of them could have a different effect on cells. For example, ceramide and sphingosine have ‘pro-apoptotic’ effects, while sphingosine 1 phosphate (S1P) has a ‘pro-growth’ function [19-21].

These findings have led to a model of sphingolipid function called the ‘sphingolipid rheostat.’ According to this model, enzymes that interconvert active sphingolipids act as switches that control the relative levels of each of metabolite, thereby regulating vital cellular processes [21-23]. The subsequent identification of genes that encode proteins that catalyze these reactions in yeast, data from genome annotation, and the identification of homologous genes in higher eukaryotes have led to the generation of knock-out transgenic animals. Phenotypic analyses of these model systems and the integration of information obtained both in vitro and in vivo are beginning to elucidate the molecular nature of sphingolipid function in membrane biology. Since membrane biology is a prominent contributor to the genesis and/or progress of numerous diseases, including those caused by metabolic disorders, such as diabetes and some cardiac diseases, and by cancer, infections, and genetic mutations, a detailed understanding of how sphingolipids influence membrane biology may prove important for developing both diagnostic and therapeutic approaches.

Most sphingolipids exist as components of membranous structures, and the composition and concentrations of sphingolipids vary among organs and within sub-cellular organelles. Therefore, it seems intuitively likely that the biological functions of sphingolipids would most immediately affect the compartments in which they reside (assuming they are not merely cargo in transit). Membranes and membrane-derived structures participate in a number of cell functions. The cell membrane defines the first physical boundary of the cell and acts as a protective barrier. Because it is selectively permeable, it regulates the movement of small molecules in and out of the cell, thereby regulating the chemical composition of its internal contents. The membrane participates in endo- and exocytosis, signaling between the intracellular and extracellular spaces, and cell-cell interactions, and it functions as a reservoir of signaling lipids. Biochemical experiments suggest that sphingolipids influence these functions.

The roles of sphingolipid metabolites have been studied in vitro and in vivo. In vitro studies have used enzyme inhibitors, structural analogues of substrates, activator molecules, and gene-knock down (RNAi) approaches to this issue. These studies have provided a great deal of information on the kinetic and the mechanistic details of the enzyme reactions. Studies making use of mutant cell lines have provided the primary information on these pathways and have led to the discovery of novel proteins. In vivo approaches have also provided a great deal of insight into the biochemical roles as well as the physiological significance of these proteins and pathways.

Eukaryotic model systems that are commonly used for in vivo studies of sphingolipid function and biochemistry include yeast, Leishmania, Drosophila, Arabidopsis, and mouse. Pioneering work in yeast provided a wealth of information on the genes and enzymes that regulate the sphingolipid metabolic pathway. Phenotype-based genetic screening has elucidated the link between this metabolic pathway and several cellular processes and facilitated a better understanding of how this pathway interacts with signal-transduction cascades [24]. Drosophila is emerging as a model system for studying sphingolipid metabolism, and although a detailed chemical profile of sphingolipids is not available currently, recent studies have attempted to bridge this gap [25]. Finally, mouse models have been used to great advantage to understand the workings of sphingolipids in the mammalian system.

Here, we will first briefly describe the best understood sphingolipids and their metabolism. We will then highlight genetic studies that have addressed the role played by sphingolipids in membrane biology, including in endocytosis, exocytosis, the formation of micro-domains, membrane integrity, and membrane signaling.

Structural Overview and Biosynthesis of Sphingolipids

All sphingolipids contain a Long Chain Base (LCB), commonly referred to as the sphingoid base, as their primary structural element. The ‘sphingoid base’ itself comprises a broad range of 2-amino-1, 3-dihydroxyalkanes or -enes [26]. In yeast there are two types of LCBs, dihydrosphingosine (DHS) and phytosphingosine (PHS), which have a carbon chain length of 18 to 20 (C-18-20). Mammals have sphingosine (predominantly C-18 [17]) as their major sphingoid base, although they also make small amounts of DHS and PHS [27]. In Drosophila and other dipterans, the sphingoid bases are not C-18, but rather C-14 and C-16 [28, 29]. The existence of shorter LCBs correlates well with Drosophila’s requirement for lower ambient temperatures to survive [22]. A long fatty-acid chain attached to an LCB through an amide linkage forms ceramide, which is esterified to form different head groups. The polar head groups of different sphingolipid species show wide biochemical variations.

The de novo biosynthesis of sphingolipids begins in the endoplasmic reticulum (ER), with the condensation of serine with acyl CoA, which forms 3-keto dihydro sphingosine. The subsequent reaction steps—reduction, acylation, and desaturation—lead to the formation of ceramide as depicted in Fig. 1. This general scheme of reactions is essentially the same in yeast, mammals, and Drosophila, although the intermediates in the biosynthetic steps vary in chain length and degree of saturation and hydroxylation, between eukaryotic species [22, 30, 31].

Figure 1.

Figure 1

de novo sphingolipid biosynthetic pathways in yeast, mammals and Drosophila The figure outlines the steps that lead to biosynthesis of major sphingolipids in the three different species.

Ceramide must localize to the lumen of the Golgi apparatus to serve as the substrate for the synthesis of sphingomyelin and complex sphingolipids [32-34]. To do so, ceramide must first traverse the cytosolic compartment. The spontaneous translocation of ceramide across the interbilayer is too slow to be biologically significant, since the t1/2 for such movement is on the order of days [35]. Thus, ceramide requires active transport from the ER to the Golgi complex. Both vesicular and nonvesicular transport mechanisms can mediate this process [36]. The nonvesicular transport is mediated by the CERT protein in mammals, in an ATP-dependent manner [37, 38]. Recent studies from our laboratory have confirmed that a CERT homolog in Drosophila has the same role, but yeast lacks a CERT homolog [37, 39]. Once transported to the Golgi complex, several different head groups can be added to ceramide to form different classes of complex sphingolipids (Fig. 2). The dominant head groups vary among animal species. In yeast, the major sphingolipids are inositol phosphoceramide (IPC), mannose inositol phosphoceramide, and mannose inositol (P2) ceramide (M(IP)2C) and they have inositol phosphate and mannose inositol phosphate(s) as head groups [40]. Mammals have sphingomyelin (SM), with a phosphocholine moiety as the head group, glucosyl ceramide (with glucose as the head group), and gangliosides, which have a head group of several sugar residues linked by various types of glycosidic linkages [41]. In Drosophila, the major sphingolipids are ceramide phosphoethanolamine (CPE), a structural analogue of SM, with phosphoethanolamine as the head group, and glycosphingolipids (GSLs), which have sugar residues as the head group [22].

Figure 2. Core Structure of sphingolipids.

Figure 2

Long chain bases (LCB) form the core structure of sphingolipids. The structure base varies in it chain length (represented by // in the figure), degree of saturation and hydroxylation. N-acylation of LCB with very long chain fatty acid (VLFA) leads to the formation of respective ceramides. Further esterificationof the R-group of the LCB leads to generation of complex sphingolipids. The nature and complexity of the R groups defines the class of complex sphingolipids.

Sphingolipids in endocytosis and exocytosis

Endocytosis is a membrane-mediated phenomenon that permits the entry of substances into cells without passing through the plasma membrane. It requires both membrane-invagination and vesicle-formation steps. Endocytosis is a key process for the cellular uptake of extracellular nutrients. It also regulates membrane dynamics and cellular responses to external stimuli by modulating cell surface-receptor activity [42]. Endocytosed vesicles are dynamic. They are continuously modified as they traverse the cell to their final destination, showing changes in composition, size, and shape, and switching from actin-based transport to microtubule-mobilized structures [43]. In addition, sphingolipids can affect endocytosis by physically altering the local environment or by initiating signaling events.

In yeast, phenotype-based classical genetic screens have led to the isolation of mutants that are defective in different endocytotic steps. Several of these mutants are associated with genes encoding proteins involved in lipid metabolism [44]. One of them, end8, has a mutation in the LCB1 gene, which encodes Lcb 1p, a component of serine palmitoyl transferase, which catalyzes the formation of 3-ketosphinganine [45, 46]. lcb1-100 is a temperature-sensitive yeast line that carries a mutation in the LCB1 gene. At the permissive temperature (24°C), lcb1-100 grows like wild-type yeast, but upon being shifted to the restrictive temperature (37°C), it fails to synthesize sphingolipids de novo and, as a consequence, fail to grow. A further study showed that lcb1-100 exhibits endocytosis defects [47]. This phenotype is rescued by providing medium containing LCB [47], indicating that the defective sphingoid base synthesis is correlated with the faulty endocytosis. Other studies extended this finding by showing that at restrictive temperatures the mutant cells have an abnormal non-polarized cortical distribution of actin filaments associated with the endocytic defects. This is corrected by providing dihydrosphingosine (DHS) or phytosphingosine (PHS), indicating that sphingoid base synthesis is required for proper actin localization. Moreover, the deregulation of actin dynamics has been shown to affect the internalization step in endocytosis [48, 49] (Fig. 3). Subsequent studies indicated that proper sphingoid base synthesis is required to control protein phosphorylation events that are crucial for endocytosis and overexpression of protein kinase C1 (PKC1) and PKH1 or PKH2 (ortholgues of mammalian protein-phosphoinositide dependent protein kinase 1 or PDK1) restores the actin localization defects in the lcb1-100 mutant [50, 51]. These studies suggest that LCB can mediate the activation of Pkh1/2.

Figure 3.

Figure 3

A schematic representation of modulation of endocytosis by sphingolipids, as determined from studies in yeast.

It is also worth mentioning here that a known target of Pkh1 and Pkh2 is the Pkc1-MAP kinase pathway [52, 53], which is involved in maintaining cell-wall integrity in yeast, supporting the idea of a connection between Pkh1/2 and the cell membrane [54, 55]. Furthermore, another study showed that the mutant rvs161Δ (reduced viability upon starvation 161, an amphiphysin homolog) is unable to repolarize actin cytoskeleton following salt stress, and this phenotype was attributed to the failure of the Rvs161p protein to associate with lipid rafts. Mutations that suppress the salt sensitivity of rvs161Δ map to genes required for sphingolipid biosynthesis, thus providing a genetic link between sphingolipids and cytoskeletal organization [56] .

Earlier studies from our laboratory demonstrated that the modulation of sphingolipid biosynthesis in vivo rescues photoreceptor degeneration in a mutant Drosophila defective in endocytosis [57]. Phototransduction in Drosophila is a prototypic guanine nucleotide-binding protein-coupled receptor signaling system and is initiated by the activation of the visual pigment Rhodopsin by incident light. Arrestins bind to light-activated Rhodopsin and assist in its inactivation. They also facilitate the endocytic turnover of light-activated Rhodopsin. In Arrestin mutants, the photoreceptors degenerate because of endocytic defects [58-60]. In these mutants, chronically active rhodopsin leads to increased calcium entry, calcium necrotoxicity, and the defective endocytosis of rhodopsin. The targeted overexpression of neutral ceramidase in the mutant background rescues the photoreceptor degeneration. In addition, ceramidase overexpression rescues retinal degeneration caused by targeted expression of a temperature-sensitive allele of dynamin, UAS-shits1, in photoreceptors suggesting that the modulation of ceramide levels could affect endocytosis [61]. Increased ceramidase expression also facilitates Rhodopsin and membrane turnover. Rhodopsin (Rh1) is required not only for transducing the light signal in the adult retina but also for stabilizing the actin-based cytoskeletal scaffold of developing photoreceptors, which is called the rhabdomere terminal web (RTW). In ninaE117 mutants, which lack Rh1, the RTW is not organized and rhabdomere biogenesis is defective. Newly eclosed flies show inappropriately formed rhabdomeric components that involute into the cytosol [62]. Ceramidase overexpression facilitates the clearing of these inappropriately formed rhabdomeric components by internalization [63]. These results suggest a potential role for sphingolipid metabolites in endocytosis and membrane turnover.

Although these studies show that ceramidase affects endocytosis, the molecular details still need to be worked out. Given that sphingoid base assembly affects actin dynamics in yeast, it would be interesting to see if a similar mechanism operates in Drosophila. The conversion of ceramide to sphingosine should affect the physical attributes of the membrane locally and could affect endocytosis. Theoretically, the asymmetric generation or removal of ceramides in the leaflets of membranes could cause vesicle aggregation and facilitate fusion and fission reactions [64]. In addition, alterations in sphingolipid composition affect the phase behavior of cell membranes. Such changes could influence lipid mobility and sphingolipids’ access to other lipids, such as phosphoinositides, thereby affecting the interactions of phosphoinositide-binding proteins with the membranes. Future genetic and biochemical studies will be required to address these and other possibilities.

In recent years, studies have also linked sphingolipids to exocytosis. During exocytosis, the secretory vesicles are directed from donor membranes to target membranes. In yeast, this process is mediated in part by membrane-associated proteins called SNAREs, which are encoded by the SNC genes [65-67]. The disruption of genes encoding the SNARE proteins can be detrimental to the organism or even lethal [68, 69].

A temperature-sensitive yeast mutant, snc, which is null for the SNC genes, grows slowly at 30°C, and its growth is blocked at restrictive temperatures due to the failure of secretory vesicles to fuse with the plasma membrane and to other defective secretory events [70]. Interestingly, the suppressor mutations of this phenotype map to the VBM1 and VBM2 genes, which are identical to ELO 2 and 3, whose gene products are involved in the elongation of fatty acids to generate the long-chain fatty acids required for ceramide synthesis [71]. While attempting to understand the role of sphingolipids in the growth and secretion phenotypes of the snc mutant cells (see below) cells, Marash and Gerst found that treating the cells with ceramide precursors or analogs restores normal growth at restrictive temperatures [72]. These results suggest that sphingolipids (ceramide and PHS) regulate the dephosphorylation event required for membrane fusion during exocytosis.

Another study showed that the mutant strain M42A snc2Δ [73], in which methionine 42 in snc1p is changed to alanine and the SNC2 gene is deleted, also has defective exocytosis. The SNC1 and SNC2 genes encode snc1p and snc2p respectively, which are archetypal v-SNARES involved in the exocytosis process [68, 69, 71, 74]. The mutants cease growth at 38°C, and at 37°C they accumulate secretory vesicles and show defective protein secretion. High copy suppressor screening showed that overexpression of the DPL1 gene (encoding dihydrosphingosine phosphate lyase) restores the growth of this mutant at 38°C. Therefore, the enhanced degradation of DHS-1P mediated by the overexpressed DPL1 gene product probably downregulates DHS-1P-mediated signaling, which leads to growth restoration. Alternatively, DHS-1P degradation could induce compositional changes of the plasma membrane that contribute to growth restoration.

In this context, it is worth noting that, in Drosophila sphingosine 1 phosphate lyase (S1P lyase) mutants, pathology of the membrane structure has been invoked as a probable cause of the reduced viability and fertility and of abnormality in the dorsal longitudinal muscle [75]. In addition, Dictyostelium discoideum S1P lyase mutants are resistant to cis-platin treatment, have defects in their ability to form migrating slugs, and show defects in F-actin distribution, again linking sphingolipid metabolism with membrane dynamics [76, 77].

S1P lyase (called SGPL in mouse) has been identified as an immediate early gene (IEG) product of PDGF activation [78]. In a recent report on mouse knock-out lines for twelve of these IEGs, including SGPL, all the lines showed similar phenotypes, in spite of the disparate functions ascribed to the individual proteins. Although SGPL mice are viable at birth, they all die by about 8 wks of age. They show reduced gains in weight and size after birth, compared with normal mice. The SGPL mutants also show hemorrhages and microaneurysms at birth and are anemic by six weeks of age. Their blood contains circulating immature blood cells, probably due to the chronic blood loss. SGPL mutants also have a high blood urea concentration and glomerular and other renal defects, along with an apparent failure in the terminal differentiation of the kidney. The skeletal system shows defects in bone development and in ossification. Finally, fibroblasts from these mice also show reduced cell migration after scratch induction [79]. Additional information about the secretory functions of cell lines from SGPL mutant mice could help shed light on the physiological role of sphingolipids.

Defective protein sorting can also lead to defects in exocytosis. In a study designed to isolate mutants of non-essential genes that are required for the sorting of proteins to the plasma membrane, mutations in four of the ten genes involved in ceramide biosynthesis, elo3, SurP, YPC1, and Ayr1 were identified (the other five genes of the pathway are essential genes and were not evaluated in this screen) [80]. Mutants of elo3, which generates a C26:0 fatty acid, cannot synthesize long-chain fatty acids for ceramide biosynthesis and mis-sort cargo to the vacuole instead of the plasma membrane. A mutation in the SurP protein affects the hydroxylation of the sphingosine backbone, and these mutants accumulate cargo in the trans-Golgi network. Mutants of YPC1 (encoding a ceramidase) show defects in the exit of the cargo protein from the Golgi. Mutation in Ayr1p (which has 3-ketoreductase activity and could therefore be involved in fatty acid elongation and thus contribute to ceramide synthesis), also affects protein targeting to the plasma membrane. It is not clear if the mistargeting of the sphingolipid metabolites that accumulate in these mutants is caused by general deleterious effects on protein sorting or by changes in the membrane composition of vesicles owing to faulty sphingolipid biosynthesis.

Data from Drosophila also provide evidence for sphingolipid involvement in exocytosis. Phenotype-based genetic screening for mutants with defective synaptic function led to the identification of slab [81]. The slab mutants arrest their growth as fully developed embryos and have severely reduced movement. They display defects in synaptic vesicle fusion and trafficking. Ultrastructural studies indicated an increased tethering of the vesicles to the plasma membrane, and impaired neurotransmission at the neuromuscular junctions was also observed. slab turned out to encode neutral ceramidase, which was described previously, supporting a role for sphingolipids in exocytosis [57, 82]. It has also been hypothesized that ceramide might modulate membrane trafficking, depending on its topological location in the membrane [64]. Conceivably, in the absence of SLAB ceramidase, an imbalance in sphingolipid composition could lead to an altered lipid raft environment and the subsequent impairment of exocytosis, which is required for proper NMJ functioning.

Microdomains and membrane integrity

Studies done in the past decade have recognized the important role of sphingolipids in the organization of localized regions called membrane microdomains (“lipid rafts”). These microdomains have distinguishable phase-separation properties and are enriched in sphingolipids and cholesterol. The preferential distribution of sphingolipids into localized domains occurs because of their unique structural properties: sphingolipids contain OH and amido groups in the ceramide moiety, and these functional groups can act simultaneously both as hydrogen-bond acceptors and as hydrogen-bond donors [83-87]. Hence, sphingolipids can form a compact hydrogen-bonded network. In the case of glycosphingolipids, the sugar residues offer substantial inter-residue binding interactions [88, 89], and polar head group interactions are reported for sphingomyelin as well. In addition, due to favorable Van der Waals interactions, sphingolipids strongly bind cholesterol, which further promotes phase separation [85]. Collectively, these interactions act as the driving force for the lateral segregation of sphingolipids into microdomains. These microdomains are detergent resistant at cold temperatures, and this characteristic is used to isolate microdomains as the detergent-resistant membrane fraction (DRM) [90, 91]. These membrane microdomains can affect cellular functions like calcium homeostasis, endocytosis, and protein sorting and they can influence signaling pathways [92-94].

Some functions of lipid rafts in vivo have been demonstrated in yeast genetic studies. These studies indicate a possible link between lipid raft integrity and the appropriate membrane targeting of proteins. Bagnat et al. studied the sorting of membrane H+-ATPase (Pma 1p) to the cell surface as a function of its association with lipid rafts [95]. Pma 1p is an essential plasma membrane protein in yeast that requires an association with lipid rafts to relocate from the Golgi complex to the membrane. In the temperature-sensitive mutant lcb1-100, at restrictive temperatures that impair sphingolipid biosynthesis, the Pma 1p protein fails to reach the cell surface and is degraded. However, this phenotype is rescued at permissive temperatures that coincide with restored sphingolipid biosynthesis [95]. Another study showed that defects in the oligomerization of Pma 1p are attributed to defective sphingolipid biosynthesis, and externally added PHS rescues the phenotype [96]. In addition, in the yeast mutant pma 1-7, in which lipids rafts are perturbed, newly synthesized Pma 1p is mistargeted to the endosomal system for degradation instead of being delivered to the plasma membrane. These studies suggest a model in which the targeting of Pma 1p to the plasma membrane requires its association with lipid rafts, although additional parameters appear to influence it as well [97, 98]

In yeast, sphingolipids are also required for GPI-anchored proteins to associate stably with the membrane. In the lcb1-100 mutant line, GPI-anchored proteins fail to associate with the membrane, behaving instead like peripheral membrane proteins. Furthermore, sphingolipid biosynthesis is required for the transport of GPI-anchored proteins from the ER to the Golgi complex. Exogenously added D-erythro-dihydroceramide partially rescues this phenotype [99].

Studies in Candida albicans (C. albicans) also emphasize a role for lipid microdomains in membrane-mediated events. erg mutants of C. albicans are deficient in ergosterol synthesis and are hypersensitive to antifungal drugs such as fluconazole, ketoconazole, and terbinafine. It was initially believed that the increased plasma membrane fluidity caused by the ergosterol deficiency was responsible for the increased drug sensitivity. However, artificially fluidizing the membrane did not increase the mutant cells’ susceptibility to the drugs, indicating that increased membrane fluidity could not be solely responsible. However, the drug susceptibilities of the erg mutants to squalene synthase were further sensitized by perturbations in the sphingolipid metabolism caused by fumonisin 1B, an inhibitor of ceramide synthase. The authors suggest that a stable interaction between ergosterol and sphingolipid is essential to maintain the stability of sphingolipid-rich microdomains [100].

Later studies with C. albicans mutants in which the ERG1 gene was conditionally disrupted or the gene for IPT1 (the synthase responsible for generating mannosyl di-inositol diphosphoceramide from mannosyl inositol phosphoceramide) was homozygously disrupted [101] showed that the lack of either gene renders the organism more susceptible to the antifungal drugs compared with wild type. In these mutant backgrounds, the authors also analyzed the surface localization of Cdr1p, a major ABC drug efflux protein that functions in the drug resistance mechanism [102, 103]. In the mutant C. albicans, Cdr1p shows impaired surface localization and defective efflux function. While it is not yet certain that Cdr1p associates with lipid rafts, these results support a model in which the drug efflux protein Cdr1p requires a sphingolipid and ergosterol for its proper localization at the membrane, which is required for it to function.

A Leishmania mutant missing one of the two subunits (SPT2) of serine palmitoyltransferase (SPT) is completely deficient in de novo sphingolipid biosynthesis. This mutant is viable, but its growth rate and doubling time in log phase are compromised. Once it reaches stationary phase, it displays a progressively higher incidence of cell-shape abnormalities and loss of viability. In stationary phase, these parasites display defects in membrane structure such as increased vacuolation, alterations in flagellar surface, and the accumulation of small vesicles reminiscent of multivesicular bodies [104]. The stationary-phase parasites also fail to differentiate into the infective metacyclic stage. In a different study, SPT2-mutant Leishmania amastigotes showed a delay in the onset of infection compared with wild-type amastigotes in vivo. In addition, these mutants show defects in membrane trafficking, and a failure of the virulence factor GP63 to associate with lipid rafts, which subsequently leads to defects in the formation of infective extracellular parasites [105].

Although a large body of work and the genetic evidence point to their significance, the definition of the lipid raft as the detergent-resistant membrane fraction is controversial, given that rafts isolated using different detergents show compositional differences [106]. The use of detergents for isolating and defining lipid rafts presents a contradiction since the detergents themselves induce phase separation in vitro [107]. Nevertheless, the studies conducted in genetic model systems provide us with good indirect evidence on the composition of lipid rafts. The study of lipid rafts encompasses multiple disciplines and requires biochemical, biophysical, and cell biological analyses. Genetic studies can provide additional information that would bridge many of the lacunae in our understanding of these intriguing structures and their role in cell function.

Sphingolipids such as sphingomyelin are considered to be structural components of the membrane. In this role they are thought to assist in the physiological interactions of membranes with their environment. Several lines of evidence indicate a close link between sphingolipid signaling and aging. Studies indicate that sphingolipids accumulate in several tissues in an age-dependent manner [108, 109], and caloric restriction decreases sphingolipid accumulation by decreasing serine palmitoyltransferase (SPT) activity [110]. Recently, we established a link between sphingolipid biosynthesis, membrane integrity, oxidative stress response, and aging in Drosophila [111]

Sphingolipids are synthesized vectorially. The early biosynthetic reactions that lead to the formation of ceramide take place in the ER, and most of the complex sphingolipids (barring some glucosylceramide) are synthesized in the Golgi complex. This necessitates the transport of ceramide from the ER to the Golgi complex. The transport of ceramide is achieved mostly through an active ATP-dependent non-vesicular mode and by a minor ATP-independent pathway that may include vesicular transport [36, 112-114]. Hanada and his colleagues recently discovered a novel protein that is responsible for the former transport mode. It was identified in studies using a mutant chinese hamster ovary (CHO) cell line (LY-A) that is defective in sphingomyelin synthesis. The LY-A cell line, when grown in sphingolipid-deficient media, synthesizes only 40% of the sphingomyelin synthesized by control CHO cells, and this defect is rescued by transfecting the cells with a cDNA clone from a human library. The responsible protein was named CERT [37].

CERT is a 68-Kda protein that possesses ceramide-transfer activity in vitro [37]. The Drosophila genome encodes a CERT homolog that maps to the left arm of the third chromosome (CG7207) at 63C3 and is named Dcert. Like the mammalian protein, it has an N-terminal PH domain followed by a coiled-coil domain with an FFAT tract and a C-terminal START domain. Drosophila lacking a functional Dcert were isolated in a western blot-based genetic screen and named dcert1 [111, 115]. Mass spectrometric analyses revealed that these flies had a defective sphingolipid metabolism with total CPE and ceramide levels less than 30% of those in wild-type control flies. These flies have a short life span and live only up to 30 days compared with the 90-day life span of wild-type flies.

In an attempt to understand the possible link between decreased CPE synthesis and reduced life span, we investigated the effect of decreased CPE biosynthesis on the plasma membrane. Drosophila CPE, like mammalian SM, localizes to the membrane and has similar phase-distribution properties. We found that the plasma membrane of the mutant flies has compromised physical properties and is more fluid than normal. Plasma membranes harbor unsaturated fatty acids (UFA) and are vulnerable to attack by reactive oxygen species (ROS), which are constantly generated in cells. The sensitivity of the plasma membrane to ROS is regulated by several factors, including the UFA composition. We hypothesized that a less compact (more fluid) plasma membrane would be more accessible to ROS and hence would suffer increased lipid peroxidation. We found that plasma membrane preparations from the mutant flies were susceptible to ferrous-induced lipid peroxidation in vitro, which translated into increased oxidative damage to the protein in vivo. The mutant flies accumulate larger amounts of oxidatively modified proteins with age than do wild-type flies. They also display an age-dependent reduction in thermal tolerance, increased mitochondrial dysfunction, decreased ATP production, and ineffective glucose utilization. These are some of the common features of premature aging [116]. These studies demonstrate that sphingolipids, acting as structural components of the membrane, can influence the sensitivity of UFAs in the plasma membrane to oxidative stress. Thus, perturbations in the integrity of the plasma membrane can have detrimental effects on the stress response and life span.

It is of interest to note here that a mouse model of Niemann-Pick disease C (Table 1) also shows abnormalities of the plasma membrane. Tissues derived from these mice show increased sphingomyelin content and increased ratios of saturated fatty acid to unsaturated fatty acids. These effects are accompanied by decreased plasma membrane fluidity [117].

Table 1.

Mouse Models of Sphingolipidoses

Mouse Models for the Disease Enzyme Mutated Phenotype and Features
Gaucher’s Glucosylceramide-β-glucosidase Animals die perinatally with accumulation of glucosylceramide in the lysosomes of the reticuloendothelial system [191].
Tay-Sach’s α- chain of Hexosaminidase Late onset behavioral changes Accumulation of GM2 in cerebral cortex, hippocampus, amygdala, hypothalamus, mamillary bodies etc [192].
Sandhoff β- chain of Hexosaminidase Show progressive motor incoordination and by 5 months almost gravely ill. Accumulation of PAS positive laden cells throughout the CNS [193].
Fabry Disease Lysosomal α-galactosidase A Accumulation of Gb3 in liver and Kidneys. Life span not affected till 80 wks [194, 195].
Krabbe Disease Galactosylceramide-β-galactosidase Twitcher mouse, die around 42 days. Accumulation of galactosylceramide and psychosine in the oligodendrocytes and Schwann cells with abundant cell death and dysmyelination of nerves. Increase in latency of neurotransmission [196].
Farber’s Disease Acid-ceramidase Embryonic lethal around 2 cell stage [197, 198]
Metachromatic leukodystrophy Arylsulfatase A Mice turn deaf because of loss of spiral ganglion cells. Reduced galactosylceramide and cholesterol [199, 200].
Niemann-Pick A Acid Sphingomyelinase Mice die between 4 and 6 months of age. Accumulation of sphingomyelin in vesicular structures of macrophages and reticuloendothelial cells. Degeneration of Purkinje cell layer of cerebellum that leads to impaired motor coordination [201].
Niemann-Pick C NPC protein Mice die around 10 weeks of age. (Accumulates unesterified cholesterol in many tissues. Age related loss of Purkinje cells in the Cerebellum [202]}.
Neutral ceramidase Inability of intestinal cells to metabolize dietary ceramide [203].
Sphingolipid activator proteins Prosaposin (deletes A, B, C and D) Symptoms early or late onset. Mice die neonatally or around 6 weeks of age. Hypomyelinated nerve fibers and complex sphingolipid accumulation in brain, liver and kidney
Saposin A Milder form compared knockout of all saposins
Saposin C and D Die by about 56 days but no myelination defect. However, these mice loose all their Purkinje cells by week 6 and accumulate glucosylceramide and α-hydroxy ceramides in brain and kidney reference [204-206]
GalNAcT β1,4-N-acetylgalactosaminyltransferase Wallerian degeneration of the myelinated neurons.
Accumulation of GM3/GD3 lack of GM2/GD2
Glucosylceramide synthase Die during gastrulation due to enhanced apoptosis in the ectoderm.
Conditional KO show loss of axonal branching in Purkinje cells [207-209]

Models of Sphingolipidoses and Membrane Biology

Several human inborn errors of metabolism affect sphingolipid metabolism, chiefly that of complex sphingolipids, which leads to the accumulation of sphingolipids in intracellular compartments. These effects can cause degenerative changes in tissues, with devastating clinical implications.

Glycosphingolipids (GSL) are complex sphingolipids that are synthesized from ceramides. They are extended by the sequential addition of sugar moieties and, in mammals, by the addition of sialic acids and in some instances of sulfated galactosylceramides. They have distinct sub-cellular localizations, for example, an apical distribution in the epithelial cells of the intestinal and urinary tracts, suggesting a role in cell polarity [118, 119]. They have also been implicated in vesicular transport along the exocytic and endocytic recycling pathways [120-122]. Mouse models for several human disorders of sphingolipid metabolism have now been generated and used for the molecular dissection of the factors leading to disease states and to evaluate therapeutic strategies.

Gangliosides are complex glycosphingolipids that contain sialic acid. They are most abundant in the central nervous system and are thought to play crucial roles in cellular interactions and the control of cell proliferation [123, 124]. The biosynthesis of gangliosides begins with lactosyl ceramide (Fig. 4). The transfer of sialic acid to lactosyl ceramide forms GM3, which serves as substrate for two different pathways. The product of GD3 synthase action on GM3 is GD3 and that of β1,4-N-acetylgalactosaminyltransferase (also known as GalNAcT EC 2.4.1.92) is GM2 and other complex gangliosides [125].

Figure 4.

Figure 4

Biosynthesis of gangliosides. In this figure we outline the committed steps that lead to the generation of GM3, GM2 and GD2 in mammals.

Table 1 lists some of the knock-out mouse lines for sphingolipid-related genes that have been generated over the last two decades. Many of these mutants manifest at least some of the symptoms seen in human patients, and almost all of them show biochemical defects that are associated with human disorders of sphingolipid metabolism. Broadly speaking, the disease states of the mutant mice result from the accumulation of the sphingolipids upstream of the enzymatic deficiency and result in their death or in tissue dysfunction, particularly of the tissues that show abnormal accumulation of the sphingolipid product(s). Hence, the symptoms manifest as loss of function of the relevant tissue. Although in some cases there is a one-to-one correlation with human symptoms, in others there are not. For example, in the mouse model of Tay-Sach’s disease, the null-mutant mice do not display symptoms like those observed in humans, despite the absence of hexosaminidase A activity. This is because, in mouse, other sialidases can convert GM2 to GA2, which is further degraded by other glycosidases, thus alleviating the symptoms of GM2 accumulation. One note of interest is that many sphingolipidoses result in myelination defects. This is not surprising, given that myelin, the fatty substance that ensheathes vertebrate axons, is rich in sphingolipids, and any impairment in the sphingolipid metabolism can disrupt the synthesis and composition of myelin.

Myelin defects can have a wide variety of consequences. As seen in Table 1, many of the mouse mutants also show defects in myelinated neurons. Galactosyl ceramide is synthesized by the enzyme ceramide galactosyl transferase. Knock-out mice for this gene do not synthesize galactosyl ceramide but can make myelin, possibly by substituting glucosyl ceramide for galactosyl ceramide. However, although the ultrastructural features of the myelin sheath from mutant mice appear normal, myelin function was compromised. The mutant mice exhibit a noticeable tremor at rest that worsens with movement. Electrophysiological data indicated the impaired conduction of nerve impulses, and the action potential parameters suggested that the myelin is defective in its ability to insulate the nerve fiber [126]. Defective sphingolipid metabolism can also affect neuronal function directly. It is interesting to note that while sialyltransferase (GD3 synthase) knockout mice are homozygous viable without clinical defects, a double knockout of β1,4-N-acetylgalactosaminyltransferase and GD3 synthase results in sudden death and susceptibility to seizures, indicating altered membrane properties [125]. GM3 synthase knock-out mice have a hypersensitive insulin response, and progress to hypoglycemia faster than wild-type mice when challenged with parenteral insulin. This result highlights the importance of gangliosides in regulating insulin receptor activity, probably by affecting glycosphingolipid rich micro-domains [15].

Spermatogenesis and fertilization require dynamic membrane remodeling, and sperm membranes have specific lipid requirements for normal function [127, 128]. Therefore, several studies have focused on sperm and fertilization in sphingolipid mutants. Acid sphingomyelinase knock-out mice die by about 6 months of age. They also have aberrations in their sperm membrane. Mature spermatozoa in these mutants show accumulations of sphingomyelin and cholesterol. The mitochondria are depolarized, and the plasma and acrosomal membranes are disrupted. The spermatozoa have an abnormal morphology with kinks and bends and faulty capacitation [129]. Morphologically abnormal sperm has also been detected in heterozygote males with this knock-out mutation [130].

Krabbe disease is caused by the absence of galactocerebrosidase in humans and results in central and peripheral neurodegeneration [131, 132]. The mouse model for Krabbe disease, caused by a defect in the homologous gene, die before the age of mating. Although these mice complete spermiogenesis, they have abnormal sperm. The acrosomes are swollen and the flagellum is angulated. In addition, the acrosomal membranes are redundant, detached from the nucleus, and folded [133].

Homozygous acid ceramidase knockout mice die as embryos and show extensive apoptosis owing to the accumulated ceramide (Table 1). The heterozygotes appear normal, although they too accumulate ceramide over time and show abnormal histopathology. One review points out that the young mice produce a 2:1 ratio of heterozygotes to wild-type progeny from heterozygous intercross matings, but aged mice produced litters with a 10:1 ratio of heterozygotes to wild-type progeny, indicating a reproductive advantage for heterozygotes that increases with age. It is possible that increased ceramide in the heterozygotes leads to secondary changes in the membrane composition of the sperm that favors fertilization [134]. In addition, acrosomal exocytosis is a crucial event in fertilization and is facilitated by ceramide [135].

In ceramide galactosyl transferase mutants, spermatogenesis in the homozygous males is arrested at the late pachytene stage, and the spermatogenic cells degenerate through apoptosis [136].

Sphingolipidoses models in Drosophila

Drosophila glycosphingolipids are synthesized from ceramide but differ from mammalian glycosphingolipids in their complexity, chain length, and the sequence in which sugar residues are added. The core structure of Drosophila glycosphingolipids consists of mannosylglucosylceramide (Manb1-4Glcb1-cer). The first step in the synthesis is catalyzed by glucosylceramide synthase. RNAi knockdown of this enzyme leads to increased apoptosis in Drosophila [137]. The Drosophila glycosyltransferases encoded by egghead (egh) and brainiac (brn) are responsible for the 2nd and 3rd steps of GSL sugar chain elongation, respectively. Egghead is a β-4-mannosyltransferase that adds mannose to GlcCer, and Brn is a β3-N-acetylglucosaminyltransferase that elongates the chain by adding GlcNAc to synthesize the glycosphingolipid core structure, GlcNAcβ1-3Manβ1-4Glcβ1-Cer (N3, triglycosylceramide), which is found widely in invertebrates [138, 139]. Both mutants show defects in epithelial morphogenesis during oogenesis and embryogenesis. They demonstrate abnormal neurogenesis, compound egg chambers, and a dorsal appendage fusion phenotype [140, 141]. Their absence leads to the loss of elongated glycosphingolipids and the accumulation of truncated precursor glycosphingolipids [138]. These mutants die as pupae, and those lacking the maternal gene product die as embryos with incorrect specification of epidermal and neural cell fates. The absence of these genes also affects ovarian follicles, which lose their apical-basal polarity and become surrounded by multiple layers of accumulated follicular cells. It has been proposed that the products of these genes are required for mediating the morphogenesis of the follicle cells by regulating germline-follicle cell adhesion [140]. Glycosyltransferase β1,4-N-acetylgalactosaminyltransferase-A (β4GalNAcTA)- and glycosyltransferase β1,4-N-acetylgalactosaminyltransferase-B (β4GalNAcTB)-null mutants have been generated in Drosophila [142]. These two proteins catalyze the fourth step in Drosophila GSL biosynthesis, by adding N-acetylgalactosamine to N3. The null mutants of β4GalNAcTA have an abnormal locomotion phenotype, i.e., larval crawling defects, a reduction in the nerve terminal bouton number in larval neuromuscular junctions, and defects in spontaneous neurotransmitter release, indicating an inefficient secretory function [143]. The β4GalNAcTB mutant, on the other hand, has no apparent phenotypes, but shows abnormal oogenesis in a third of the ovarioles, with degenerative changes from stages 6 to 9 [144]. Fifty percent of the eggs from the remaining ovarioles show a dorsal appendage fusion phenotype, milder than in egg and brn, which is a sign of defective EGFR signaling between the germ-line and the follicle cells with the defect in the oocyte. Studies suggest that N3 is absolutely critical for survival, but its subsequent elongation, though not critical, is still important for the reproductive capability of the flies.

Sphingolipids as ligands for membrane receptors

In addition to their functions as membrane components and as signaling molecules inside cells, sphingolipids, specifically S1P, can act as extracellular ligand. S1P regulates cell migration and trafficking, also acts as a chemo attractant for migrating cells, and modulates permeability barriers (probably by engaging specific receptors on the target cell surface). S1P is a second messenger molecule with both intra- and extra-cellular functions. Its signal is important for cell growth, differentiation, and organismal development. In mammals, the serum concentrations of S1P are fairly high and believed to be derived from red blood cells and platelets, among others [145].

S1P precursors are found in eukaryotes, and sphingosine kinase, S1P, and sphingosine phosphate lyase are found in yeast, humans, and plants. However, a recognizable S1P receptor has been identified only in chordates and vertebrates [146]. Thus, the evolution of S1P into a signaling ligand at the cell surface is of recent origin and might contribute to the sophistication and complexity of these phyla. S1P is generated by the phosphorylation of sphingosine by sphingosine kinase (SphK) [147, 148]. Because recent reviews have dealt extensively with the S1P family receptors we will make only brief mention of them here, focusing on the findings of in vivo studies with an emphasis on membrane biology [149, 150].

Overview of S1P Family of Receptors

Hla and Maciag found that, following PMA stimulation, human endothelial cells produced abundant mRNA for EDG-1 (now recognized as S1P1), and they suggested a role for this receptor in the differentiation of endothelial cells [151]. Subsequent studies established that at least in chordates and mammals, S1P act as extracellular ligands for a family of G-protein coupled receptors called S1P receptors [152-154]. In mammals, S1P executes most of its extracellular functions by engaging these receptors. There are five members of this family, designated S1P1 - S1P5, and knock-out lines are available for four of them [146, 155, 156]. S1P1, S1P2, and S1P3 show a widespread distribution in various cell and tissue types, whereas S1P4 and S1P5 apparently are more restricted in their distribution pattern and are found mainly in the hematopoeitic system [157]. Although all these receptors are activated by S1P binding, studies with receptor inhibitors suggest variations in the S1P binding site among these receptors [152, 158, 159]. There are also differences in the coupling of receptor activation to downstream G proteins [160-163]. S1P 1 couples to Gi, S1P2 and S1P3 couple to Gi, Gq, and G13. S1P4 couples to Gi and G13 and S1P5 couples to Gi and G12. G0/Gi mediate the activation of Ras, Rac, ERK, PI3K and AKT. They can facilitate endothelial barrier function, induce vasodilation and cause migration. These signals also activate PKC and PLC to increase intracellular calcium while concomitantly inhibiting cAMP. When coupled to Gq, the S1P receptors activate PLC. G12/13 activate Rho and ROCK that can inhibit migration, decrease the effectiveness of the endothelial barrier and cause vasoconstriction. Many effects of the S1P receptor signaling pathways could occur via the regulation of actin assembly and disassembly. The coupling of S1P1 or S1P3 with G0 or Gi activates the Rac pathway, which can lead to cell polarization, lamellipodium formation and expansion, organization of focal complexes at the frontal edge of the cell, and adhesion. S1P2 or S1P3 coupled with G12/13 stimulates Rho/ROCK, which cause the cells to produce stress fibers and to detach from the substrate. Therefore Rac and Rho are utilized in a concerted fashion to affect migration, the formation and dissociation of focal complexes, and cell-cell adhesion. Endothelial cells express S1P1 and that S1P1 is critical for the interactions of endothelial cells with mural cells in the maturation of blood vessels, the failure of which leads to hemorrhage and death in S1P1-null mice [164]. The mechanism is through S1P1 activation by S1P, which activates the Rac GTPase (via Gi), which utilizes microtubules to transport N-cadherin to the cell surface where it interacts with the mural cells. The S1P1 expressed on the mural cells is irrelevant for this function. S1P1 also stabilizes the cadherin-cadherin interaction by promoting the association of α-, β-, and p120-catenin with N-cadherin and the underlying actin filaments. S1P1 has also been implicated in cardiovascular and neuronal development, angiogenesis, tonicity of blood vessels, vascular barrier functions, and gap-junction formation and maintenance [165-167]. Through its association with the Rho and Rac family of G-proteins, the S1P1 receptor induces the formation of cortical actin rings, enforces the formation of cell junction assembly, induces stress fibers, and perturbs or strengthens cell-cell junctions. Thus, fundamental aspects of membrane biology are affected by the S1P-S1P1 receptor interaction, which is mostly mediated through the actin cytoskeleton.

In an important study, Lee et al. showed that the engagement of S1P1 (and probably S1P3) leads to the recruitment of VE-Cadherin, and α-, β-, and γ-catenin to cell-cell junctions as well as a dramatic increase in actin stress fibers and cortical actin structures in endothelial cells [168]. They showed it also promotes the homophilic binding of intercellular cadherins and the attachment of these structures to the cytoskeleton. Rac localization at the cell-cell contact area also increases. Several lines of evidence now indicate that S1P activates the Rac pathway through G-protein coupled receptor activation, resulting in the cortical assembly of actin, and that this process of cytoskeletal reorganization is an important event in angiogenesis [168-170] [171]. Lee et al., also found that the S1P-induced assembly of focal contact sites was Rho dependent on three different extracellular matrices. Interestingly, Rac was not required for the initial focal contact assembly and cell spreading. (Rho can be activated in non-adherent human umbilical vein endothelial cells (HUVECs), whereas Rac requires cell adhesion.) Direct measurement showed that both S1P1 and S1P3 can activate both Rho and Rac. In addition, integrin alpha-5 beta 3 and 1 were shown to be required for HUVEC migration and morphogenesis in coordination with S1P1 and S1P3 activation. This pioneering study laid the groundwork for future studies on S1P receptor knock-out mouse lines.

Thus S1P receptors, activated by binding S1P, control several aspects of endothelial cell function that are important for angiogenesis [172, 173].

Mouse models

S1P1 knock-out mouse embryos are normal up to E11.5. On E12.5, the embryos exhibit abnormal yolk sacs, are edematous, and lack blood. They die between E12.5 and E14. These findings suggest that the S1P1 expressed on these cells may play a role in regulating the adhesion of endothelial cells and pericytes [174]. S1P is also believed to be an important signaling molecule in the regulation of cell-cell junction formation, by organizing actin into a strong cortical ring and reinforcing cell-matrix adherence [168, 175]. Thus, S1P is likely to regulate vascular permeability by maintaining the integrity of the vascular barrier.

Lymphocyte egress (a form of chemotaxis) is thought to be mediated by S1P1 and probably by S1P4 [145, 176]. Convincing evidence from studies with S1P1-deficient mice have demonstrated that ligation of the S1P1 receptor by S1P is critical for the re-entry of T lymphocytes from secondary lymphoid organs to peripheral tissues, where inflammation takes place [176, 177]. S1P increases intracellular calcium by activating the S1P2 receptor and induces smooth muscle cell contraction [178, 179].

S1P2 receptor-null mice are apparently normal at birth, with no anatomical defects, but between weeks 3 and 7, 59% of them are affected by sporadic seizures that are occasionally lethal (14% die) [180]. Interestingly, a mutation in a similar gene in zebrafish leads to embryonic lethality [181]. Therefore S1P2 receptors could affect either microscopic development or membrane excitability by affecting intracellular calcium. The S1P3 receptor is highly expressed in the heart, lung, spleen, kidney, intestine, diaphragm, and certain cartilaginous tissues. Mice null for this gene do not show any overt phenotype. S1P4 is expressed mostly in immune compartments, and no knock-out has been reported. S1P5 is expressed in the CNS, and knock-outs show no obvious phenotype.

The concerted action of S1P1, S1P2, and S1P3 is required for endothelial barrier function [149, 171, 182, 183]. S1P3 receptors play a prominent role in the regulation (slowing) of the heart rate [184]. The S1P-mediated signaling pathway also assumes importance in driving the migration of cells. Likewise, the S1P2 and S1P3 receptors, functioning in concert with the S1P1 receptor, participate in the formation of an intact embryonic vasculature, as demonstrated by the hemorrhage caused by their simultaneous genetic deletion [185].

Sphingosine Kinases and S1P

Serum S1P must be mostly derived from the action of Sphingosine Kinase 1 (SphK1), since SphK1 null mice have substantially lower levels of circulating S1P [186]. Also, unlike what was previously thought, SPHK1 does not have a determining role in inflammatory responses [187]. Furthermore, the loss of either SPHK1 or SPHK2 alone does not lead to developmental defects in mice. SphK1 knock-out mice are viable, fertile, and have no obvious developmental abnormalities [165, 186]. SphK2-/- mice are viable and sub-fertile [188]. Two recent reports have shown that SPHK2 is involved in the phosphorylation of the immune modulator FTY720 (a sphingosine analog) and that SPHK2 null mice do not respond to this drug by developing lymphopenia, as normal mice do [188, 189]. In addition, circulating S1P (generated mostly by SphK1) and intrinsic mast cell SPHK2 together determine the anaphylactic response, a phenomenon that involves degranulation, which is a specialized form of exocytosis. SphK2-null mice show defective PKC translocation, NF-κB activation, and calcium influx, which are all part of the regular anaphylactic response [190]. SphK1-/- SphK2-/- double mutants die as embryos, although the embryos appear normal until E9.5. By E11.5 and E12.5, however, all the embryos exhibit cranial hemorrhage and none survive beyond E13.5. S1P is undetectable in embryonic homogenates and no accumulation of sphingosine is observed. The vascular defects in these double-knock out embryos are prominent: they exhibit a poorly developed dorsal aortal wall, and the endothelial cells in all the blood vessels are severely vacuolated. The embryos also show mesenchymal defects, cranial neural tube defects, and disrupted cell-cell junctions [165]. These phenotypes are similar to those seen in S1P1 receptor knock-out mice.

Conclusions

Biochemical studies have implicated sphingolipids in membrane-mediated processes, and studies in yeast have further strengthened this concept. In the last fifteen years, animal knock-out lines for sphingolipid genes and receptors have become available. Most studies have concentrated on gross phenotypic analyses and the functional consequences of these mutations. In some of these mutant lines, however, the focus has now shifted from phenotypic observations to cell biological studies. In the near future, the molecular changes for all the available sphingolipid-related knock-out lines will be dissected. These studies will greatly elucidate how sphingolipids contribute to membrane biology and intracellular signaling.

Acknowledgements

The authors acknowledge the support from the intramural division of the National Cancer Institute, National Institutes of Health and Department of Health and Human Services.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Adachi-Yamada T, Gotoh T, Sugimura I, Tateno M, Nishida Y, Onuki T, Date H. De novo synthesis of sphingolipids is required for cell survival by down-regulating c-Jun N-terminal kinase in Drosophila imaginal discs. Mol Cell Biol. 1999;19:7276–7286. doi: 10.1128/mcb.19.10.7276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Hannun YA, Luberto C, Argraves KM. Enzymes of sphingolipid metabolism: from modular to integrative signaling. Biochemistry. 2001;40:4893–4903. doi: 10.1021/bi002836k. [DOI] [PubMed] [Google Scholar]
  • 3.Huwiler A, Kolter T, Pfeilschifter J, Sandhoff K. Physiology and pathophysiology of sphingolipid metabolism and signaling. Biochim Biophys Acta. 2000;1485:63–99. doi: 10.1016/s1388-1981(00)00042-1. [DOI] [PubMed] [Google Scholar]
  • 4.Wells GB, 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. 1983;258:10200–10203. [PubMed] [Google Scholar]
  • 5.Kim MY, Linardic C, Obeid L, Hannun Y. Identification of sphingomyelin turnover as an effector mechanism for the action of tumor necrosis factor alpha and gamma-interferon. Specific role in cell differentiation. J Biol Chem. 1991;266:484–489. [PubMed] [Google Scholar]
  • 6.Morales A, Lee H, Goni FM, Kolesnick R, Fernandez-Checa JC. Sphingolipids and cell death. Apoptosis. 2007;12:923–939. doi: 10.1007/s10495-007-0721-0. [DOI] [PubMed] [Google Scholar]
  • 7.Spiegel S, Milstien S. Sphingosine 1-phosphate, a key cell signaling molecule. J Biol Chem. 2002;277:25851–25854. doi: 10.1074/jbc.R200007200. [DOI] [PubMed] [Google Scholar]
  • 8.Kashiwagi K, Shirai Y, Kuriyama M, Sakai N, Saito N. Importance of C1B domain for lipid messenger-induced targeting of protein kinase C. J Biol Chem. 2002;277:18037–18045. doi: 10.1074/jbc.M111761200. [DOI] [PubMed] [Google Scholar]
  • 9.Hannun YA, Loomis CR, Merrill AH, Jr., Bell RM. Sphingosine inhibition of protein kinase C activity and of phorbol dibutyrate binding in vitro and in human platelets. J Biol Chem. 1986;261:12604–12609. [PubMed] [Google Scholar]
  • 10.Allen JA, Halverson-Tamboli RA, Rasenick MM. Lipid raft microdomains and neurotransmitter signalling. Nat Rev Neurosci. 2007;8:128–140. doi: 10.1038/nrn2059. [DOI] [PubMed] [Google Scholar]
  • 11.Brown DA, London E. Functions of lipid rafts in biological membranes. Annu Rev Cell Dev Biol. 1998;14:111–136. doi: 10.1146/annurev.cellbio.14.1.111. [DOI] [PubMed] [Google Scholar]
  • 12.Simons K, Toomre D. Lipid rafts and signal transduction. Nat Rev Mol Cell Biol. 2000;1:31–39. doi: 10.1038/35036052. [DOI] [PubMed] [Google Scholar]
  • 13.Tsai B, Gilbert JM, Stehle T, Lencer W, Benjamin TL, Rapoport TA. Gangliosides are receptors for murine polyoma virus and SV40. Embo J. 2003;22:4346–4355. doi: 10.1093/emboj/cdg439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Low JA, Magnuson B, Tsai B, Imperiale MJ. Identification of gangliosides GD1b and GT1b as receptors for BK virus. J Virol. 2006;80:1361–1366. doi: 10.1128/JVI.80.3.1361-1366.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Yamashita T, Hashiramoto A, Haluzik M, Mizukami H, Beck S, Norton A, Kono M, Tsuji S, Daniotti JL, Werth N, Sandhoff R, Sandhoff K, Proia RL. Enhanced insulin sensitivity in mice lacking ganglioside GM3. Proc Natl Acad Sci U S A. 2003;100:3445–3449. doi: 10.1073/pnas.0635898100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Holthuis JC, Pomorski T, Raggers RJ, Sprong H, Van Meer G. The organizing potential of sphingolipids in intracellular membrane transport. Physiol Rev. 2001;81:1689–1723. doi: 10.1152/physrev.2001.81.4.1689. [DOI] [PubMed] [Google Scholar]
  • 17.Shayman JA. Sphingolipids. Kidney Int. 2000;58:11–26. doi: 10.1046/j.1523-1755.2000.00136.x. [DOI] [PubMed] [Google Scholar]
  • 18.Posse de Chaves EI. Sphingolipids in apoptosis, survival and regeneration in the nervous system. Biochim Biophys Acta. 2006;1758:1995–2015. doi: 10.1016/j.bbamem.2006.09.018. [DOI] [PubMed] [Google Scholar]
  • 19.Chalfant CE, Spiegel S. Sphingosine 1-phosphate and ceramide 1-phosphate: expanding roles in cell signaling. J Cell Sci. 2005;118:4605–4612. doi: 10.1242/jcs.02637. [DOI] [PubMed] [Google Scholar]
  • 20.Cuvillier O. Sphingosine in apoptosis signaling. Biochim Biophys Acta. 2002;1585:153–162. doi: 10.1016/s1388-1981(02)00336-0. [DOI] [PubMed] [Google Scholar]
  • 21.Spiegel S, Milstien S. Sphingosine-1-phosphate: an enigmatic signalling lipid. Nat Rev Mol Cell Biol. 2003;4:397–407. doi: 10.1038/nrm1103. [DOI] [PubMed] [Google Scholar]
  • 22.Acharya U, Acharya JK. Enzymes of sphingolipid metabolism in Drosophila melanogaster. Cell Mol Life Sci. 2005;62:128–142. doi: 10.1007/s00018-004-4254-1. [DOI] [PubMed] [Google Scholar]
  • 23.Taha TA, Mullen TD, Obeid LM. A house divided: ceramide, sphingosine, and sphingosine-1-phosphate in programmed cell death. Biochim Biophys Acta. 2006;1758:2027–2036. doi: 10.1016/j.bbamem.2006.10.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Dickson RC, Sumanasekera C, Lester RL. Functions and metabolism of sphingolipids in Saccharomyces cerevisiae. Prog Lipid Res. 2006;45:447–465. doi: 10.1016/j.plipres.2006.03.004. [DOI] [PubMed] [Google Scholar]
  • 25.Fyrst H, Herr DR, Harris GL, Saba JD. Characterization of free endogenous C14 and C16 sphingoid bases from Drosophila melanogaster. J Lipid Res. 2004;45:54–62. doi: 10.1194/jlr.M300005-JLR200. [DOI] [PubMed] [Google Scholar]
  • 26.Merrill AH., Jr. De novo sphingolipid biosynthesis: a necessary, but dangerous, pathway. J Biol Chem. 2002;277:25843–25846. doi: 10.1074/jbc.R200009200. [DOI] [PubMed] [Google Scholar]
  • 27.Dickson RC, Lester RL. Sphingolipid functions in Saccharomyces cerevisiae. Biochim Biophys Acta. 2002;1583:13–25. doi: 10.1016/s1388-1981(02)00210-x. [DOI] [PubMed] [Google Scholar]
  • 28.Hildenbrandt GR, Abraham T, Bieber LL. Metabolism of ceramide phosphorylethanolamine, phosphatidylinositol, phosphatidylserine and phosphatidylglycerol by housefly larvae. Lipids. 1971;6:508–516. doi: 10.1007/BF02531237. [DOI] [PubMed] [Google Scholar]
  • 29.Bieber LL, O’Connor JD, Sweeley CC. The occurrence of tetradecasphing-4-enine and hexadecasphing-4-enine as the principal sphingosines of Musca domestica larvae and adults. Biochim Biophys Acta. 1969;187:157–159. doi: 10.1016/0005-2760(69)90145-3. [DOI] [PubMed] [Google Scholar]
  • 30.Hanada K. Discovery of the molecular machinery CERT for endoplasmic reticulum-to-Golgi trafficking of ceramide. Mol Cell Biochem. 2006;286:23–31. doi: 10.1007/s11010-005-9044-z. [DOI] [PubMed] [Google Scholar]
  • 31.Sims KJ, Spassieva SD, Voit EO, Obeid LM. Yeast sphingolipid metabolism: clues and connections. Biochem Cell Biol. 2004;82:45–61. doi: 10.1139/o03-086. [DOI] [PubMed] [Google Scholar]
  • 32.Futerman AH, Stieger B, Hubbard AL, Pagano RE. Sphingomyelin synthesis in rat liver occurs predominantly at the cis and medial cisternae of the Golgi apparatus. J Biol Chem. 1990;265:8650–8657. [PubMed] [Google Scholar]
  • 33.Futerman AH, Pagano RE. Determination of the intracellular sites and topology of glucosylceramide synthesis in rat liver. Biochem J. 1991;280(Pt 2):295–302. doi: 10.1042/bj2800295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Jeckel D, Karrenbauer A, Burger KN, van Meer G, Wieland F. Glucosylceramide is synthesized at the cytosolic surface of various Golgi subfractions. J Cell Biol. 1992;117:259–267. doi: 10.1083/jcb.117.2.259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Simon CG, Jr., Holloway PW, Gear AR. Exchange of C(16)-ceramide between phospholipid vesicles. Biochemistry. 1999;38:14676–14682. doi: 10.1021/bi991537w. [DOI] [PubMed] [Google Scholar]
  • 36.Funato K, Riezman H. Vesicular and nonvesicular transport of ceramide from ER to the Golgi apparatus in yeast. J Cell Biol. 2001;155:949–959. doi: 10.1083/jcb.200105033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Hanada K, Kumagai K, Yasuda S, Miura Y, Kawano M, Fukasawa M, Nishijima M. Molecular machinery for non-vesicular trafficking of ceramide. Nature. 2003;426:803–809. doi: 10.1038/nature02188. [DOI] [PubMed] [Google Scholar]
  • 38.Kumagai K, Yasuda S, Okemoto K, Nishijima M, Kobayashi S, Hanada K. CERT mediates intermembrane transfer of various molecular species of ceramides. J Biol Chem. 2005;280:6488–6495. doi: 10.1074/jbc.M409290200. [DOI] [PubMed] [Google Scholar]
  • 39.Perry RJ, Ridgway ND. Molecular mechanisms and regulation of ceramide transport. Biochim Biophys Acta. 2005;1734:220–234. doi: 10.1016/j.bbalip.2005.04.001. [DOI] [PubMed] [Google Scholar]
  • 40.Dickson RC, Lester RL. Metabolism and selected functions of sphingolipids in the yeast Saccharomyces cerevisiae. Biochim Biophys Acta. 1999;1438:305–321. doi: 10.1016/s1388-1981(99)00068-2. [DOI] [PubMed] [Google Scholar]
  • 41.Holthuis JC, van Meer G, Huitema K. Lipid microdomains, lipid translocation and the organization of intracellular membrane transport (Review) Mol Membr Biol. 2003;20:231–241. [PubMed] [Google Scholar]
  • 42.D’Hondt K, Heese-Peck A, Riezman H. Protein and lipid requirements for endocytosis. Annu Rev Genet. 2000;34:255–295. doi: 10.1146/annurev.genet.34.1.255. [DOI] [PubMed] [Google Scholar]
  • 43.Mellman I. Endocytosis and molecular sorting. Annu Rev Cell Dev Biol. 1996;12:575–625. doi: 10.1146/annurev.cellbio.12.1.575. [DOI] [PubMed] [Google Scholar]
  • 44.Souza CM, Pichler H. Lipid requirements for endocytosis in yeast. Biochim Biophys Acta. 2007;1771:442–454. doi: 10.1016/j.bbalip.2006.08.006. [DOI] [PubMed] [Google Scholar]
  • 45.Sutterlin C, Doering TL, Schimmoller F, Schroder S, Riezman H. Specific requirements for the ER to Golgi transport of GPI-anchored proteins in yeast. J Cell Sci. 1997;110(Pt 21):2703–2714. doi: 10.1242/jcs.110.21.2703. [DOI] [PubMed] [Google Scholar]
  • 46.Buede R, Rinker-Schaffer C, Pinto WJ, Lester RL, Dickson RC. Cloning and characterization of LCB1, a Saccharomyces gene required for biosynthesis of the long-chain base component of sphingolipids. J Bacteriol. 1991;173:4325–4332. doi: 10.1128/jb.173.14.4325-4332.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Zanolari B, Friant S, Funato K, Sutterlin C, Stevenson BJ, Riezman H. Sphingoid base synthesis requirement for endocytosis in Saccharomyces cerevisiae. Embo J. 2000;19:2824–2833. doi: 10.1093/emboj/19.12.2824. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Kubler E, Riezman H. Actin and fimbrin are required for the internalization step of endocytosis in yeast. Embo J. 1993;12:2855–2862. doi: 10.1002/j.1460-2075.1993.tb05947.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Lamaze C, Fujimoto LM, Yin HL, Schmid SL. The actin cytoskeleton is required for receptor-mediated endocytosis in mammalian cells. J Biol Chem. 1997;272:20332–20335. doi: 10.1074/jbc.272.33.20332. [DOI] [PubMed] [Google Scholar]
  • 50.Friant S, Zanolari B, Riezman H. Increased protein kinase or decreased PP2A activity bypasses sphingoid base requirement in endocytosis. Embo J. 2000;19:2834–2844. doi: 10.1093/emboj/19.12.2834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Friant S, Lombardi R, Schmelzle T, Hall MN, Riezman H. Sphingoid base signaling via Pkh kinases is required for endocytosis in yeast. Embo J. 2001;20:6783–6792. doi: 10.1093/emboj/20.23.6783. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Inagaki M, Schmelzle T, Yamaguchi K, Irie K, Hall MN, Matsumoto K. PDK1 homologs activate the Pkc1-mitogen-activated protein kinase pathway in yeast. Mol Cell Biol. 1999;19:8344–8352. doi: 10.1128/mcb.19.12.8344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Zhang X, Lester RL, Dickson RC. Pil1p and Lsp1p negatively regulate the 3-phosphoinositide-dependent protein kinase-like kinase Pkh1p and downstream signaling pathways Pkc1p and Ypk1p. J Biol Chem. 2004;279:22030–22038. doi: 10.1074/jbc.M400299200. [DOI] [PubMed] [Google Scholar]
  • 54.Errede B, Levin DE. A conserved kinase cascade for MAP kinase activation in yeast. Curr Opin Cell Biol. 1993;5:254–260. doi: 10.1016/0955-0674(93)90112-4. [DOI] [PubMed] [Google Scholar]
  • 55.Levin DE. Cell wall integrity signaling in Saccharomyces cerevisiae. Microbiol Mol Biol Rev. 2005;69:262–291. doi: 10.1128/MMBR.69.2.262-291.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Balguerie A, Bagnat M, Bonneu M, Aigle M, Breton AM. Rvs161p and sphingolipids are required for actin repolarization following salt stress. Eukaryot Cell. 2002;1:1021–1031. doi: 10.1128/EC.1.6.1021-1031.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Acharya U, Patel S, Koundakjian E, Nagashima K, Han X, Acharya JK. Modulating sphingolipid biosynthetic pathway rescues photoreceptor degeneration. Science. 2003;299:1740–1743. doi: 10.1126/science.1080549. [DOI] [PubMed] [Google Scholar]
  • 58.Alloway PG, Howard L, Dolph PJ. The formation of stable rhodopsin-arrestin complexes induces apoptosis and photoreceptor cell degeneration. Neuron. 2000;28:129–138. doi: 10.1016/s0896-6273(00)00091-x. [DOI] [PubMed] [Google Scholar]
  • 59.Kiselev A, Socolich M, Vinos J, Hardy RW, Zuker CS, Ranganathan R. A molecular pathway for light-dependent photoreceptor apoptosis in Drosophila. Neuron. 2000;28:139–152. doi: 10.1016/s0896-6273(00)00092-1. [DOI] [PubMed] [Google Scholar]
  • 60.Orem NR, Dolph PJ. Loss of the phospholipase C gene product induces massive endocytosis of rhodopsin and arrestin in Drosophila photoreceptors. Vision Res. 2002;42:497–505. doi: 10.1016/s0042-6989(01)00229-2. [DOI] [PubMed] [Google Scholar]
  • 61.Kitamoto T. Conditional modification of behavior in Drosophila by targeted expression of a temperature-sensitive shibire allele in defined neurons. J Neurobiol. 2001;47:81–92. doi: 10.1002/neu.1018. [DOI] [PubMed] [Google Scholar]
  • 62.Chang HY, Ready DF. Rescue of photoreceptor degeneration in rhodopsin-null Drosophila mutants by activated Rac1. Science. 2000;290:1978–1980. doi: 10.1126/science.290.5498.1978. [DOI] [PubMed] [Google Scholar]
  • 63.Acharya U, Mowen MB, Nagashima K, Acharya JK. Ceramidase expression facilitates membrane turnover and endocytosis of rhodopsin in photoreceptors. Proc Natl Acad Sci U S A. 2004;101:1922–1926. doi: 10.1073/pnas.0308693100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.van Blitterswijk WJ, van der Luit AH, Veldman RJ, Verheij M, Borst J. Ceramide: second messenger or modulator of membrane structure and dynamics? Biochem J. 2003;369:199–211. doi: 10.1042/BJ20021528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Nichols BJ, Ungermann C, Pelham HR, Wickner WT, Haas A. Homotypic vacuolar fusion mediated by t- and v-SNAREs. Nature. 1997;387:199–202. doi: 10.1038/387199a0. [DOI] [PubMed] [Google Scholar]
  • 66.Ferro-Novick S, Jahn R. Vesicle fusion from yeast to man. Nature. 1994;370:191–193. doi: 10.1038/370191a0. [DOI] [PubMed] [Google Scholar]
  • 67.Burri L, Lithgow T. A complete set of SNAREs in yeast. Traffic. 2004;5:45–52. doi: 10.1046/j.1600-0854.2003.00151.x. [DOI] [PubMed] [Google Scholar]
  • 68.Gurunathan S, Chapman-Shimshoni D, Trajkovic S, Gerst JE. Yeast exocytic v-SNAREs confer endocytosis. Mol Biol Cell. 2000;11:3629–3643. doi: 10.1091/mbc.11.10.3629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Gerst JE, Rodgers L, Riggs M, Wigler M. SNC1, a yeast homolog of the synaptic vesicle-associated membrane protein/synaptobrevin gene family: genetic interactions with the RAS and CAP genes. Proc Natl Acad Sci U S A. 1992;89:4338–4342. doi: 10.1073/pnas.89.10.4338. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Protopopov V, Govindan B, Novick P, Gerst JE. Homologs of the synaptobrevin/VAMP family of synaptic vesicle proteins function on the late secretory pathway in S. cerevisiae. Cell. 1993;74:855–861. doi: 10.1016/0092-8674(93)90465-3. [DOI] [PubMed] [Google Scholar]
  • 71.David D, Sundarababu S, Gerst JE. Involvement of long chain fatty acid elongation in the trafficking of secretory vesicles in yeast. J Cell Biol. 1998;143:1167–1182. doi: 10.1083/jcb.143.5.1167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Marash M, Gerst JE. t-SNARE dephosphorylation promotes SNARE assembly and exocytosis in yeast. Embo J. 2001;20:411–421. doi: 10.1093/emboj/20.3.411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Grote E, Vlacich G, Pypaert M, Novick PJ. A snc1 endocytosis mutant: phenotypic analysis and suppression by overproduction of dihydrosphingosine phosphate lyase. Mol Biol Cell. 2000;11:4051–4065. doi: 10.1091/mbc.11.12.4051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Gerst JE. Conserved alpha-helical segments on yeast homologs of the synaptobrevin/VAMP family of v-SNAREs mediate exocytic function. J Biol Chem. 1997;272:16591–16598. doi: 10.1074/jbc.272.26.16591. [DOI] [PubMed] [Google Scholar]
  • 75.Herr DR, Fyrst H, Phan V, Heinecke K, Georges R, Harris GL, Saba JD. Sply regulation of sphingolipid signaling molecules is essential for Drosophila development. Development. 2003;130:2443–2453. doi: 10.1242/dev.00456. [DOI] [PubMed] [Google Scholar]
  • 76.Li G, Foote C, Alexander S, Alexander H. Sphingosine-1-phosphate lyase has a central role in the development of Dictyostelium discoideum. Development. 2001;128:3473–3483. doi: 10.1242/dev.128.18.3473. [DOI] [PubMed] [Google Scholar]
  • 77.Kumar A, Wessels D, Daniels KJ, Alexander H, Alexander S, Soll DR. Sphingosine-1-phosphate plays a role in the suppression of lateral pseudopod formation during Dictyostelium discoideum cell migration and chemotaxis. Cell Motil Cytoskeleton. 2004;59:227–241. doi: 10.1002/cm.20035. [DOI] [PubMed] [Google Scholar]
  • 78.Chen WV, Delrow J, Corrin PD, Frazier JP, Soriano P. Identification and validation of PDGF transcriptional targets by microarray-coupled gene-trap mutagenesis. Nat Genet. 2004;36:304–312. doi: 10.1038/ng1306. [DOI] [PubMed] [Google Scholar]
  • 79.Schmahl J, Raymond CS, Soriano P. PDGF signaling specificity is mediated through multiple immediate early genes. Nat Genet. 2007;39:52–60. doi: 10.1038/ng1922. [DOI] [PubMed] [Google Scholar]
  • 80.Proszynski TJ, Klemm RW, Gravert M, Hsu PP, Gloor Y, Wagner J, Kozak K, Grabner H, Walzer K, Bagnat M, Simons K, Walch-Solimena C. A genome-wide visual screen reveals a role for sphingolipids and ergosterol in cell surface delivery in yeast. Proc Natl Acad Sci U S A. 2005;102:17981–17986. doi: 10.1073/pnas.0509107102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Rohrbough J, Rushton E, Palanker L, Woodruff E, Matthies HJ, Acharya U, Acharya JK, Broadie K. Ceramidase regulates synaptic vesicle exocytosis and trafficking. J Neurosci. 2004;24:7789–7803. doi: 10.1523/JNEUROSCI.1146-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Yoshimura Y, Okino N, Tani M, Ito M. Molecular cloning and characterization of a secretory neutral ceramidase of Drosophila melanogaster. J Biochem (Tokyo) 2002;132:229–236. doi: 10.1093/oxfordjournals.jbchem.a003215. [DOI] [PubMed] [Google Scholar]
  • 83.Ferraretto A, Pitto M, Palestini P, Masserini M. Lipid domains in the membrane: thermotropic properties of sphingomyelin vesicles containing GM1 ganglioside and cholesterol. Biochemistry. 1997;36:9232–9236. doi: 10.1021/bi970428j. [DOI] [PubMed] [Google Scholar]
  • 84.Masserini M, Ravasi D. Role of sphingolipids in the biogenesis of membrane domains. Biochim Biophys Acta. 2001;1532:149–161. doi: 10.1016/s1388-1981(01)00128-7. [DOI] [PubMed] [Google Scholar]
  • 85.McIntosh TJ, Simon SA, Needham D, Huang CH. Structure and cohesive properties of sphingomyelin/cholesterol bilayers. Biochemistry. 1992;31:2012–2020. doi: 10.1021/bi00122a017. [DOI] [PubMed] [Google Scholar]
  • 86.Pascher I. Molecular arrangements in sphingolipids. Conformation and hydrogen bonding of ceramide and their implication on membrane stability and permeability. Biochim Biophys Acta. 1976;455:433–451. doi: 10.1016/0005-2736(76)90316-3. [DOI] [PubMed] [Google Scholar]
  • 87.Takagi A, Ikeda Y, Takeda E, Yamamoto A. A newly identified lipoprotein lipase (LPL) gene mutation (F270L) in a Japanese patient with familial LPL deficiency. Biochim Biophys Acta. 2000;1502:433–446. doi: 10.1016/s0925-4439(00)00067-3. [DOI] [PubMed] [Google Scholar]
  • 88.Kojima N, Sakata S, Nakamura S, Nagai K, Takuno H, Ogawa T, Matsui I, Sarui H, Miura K. Serum concentrations of osteocalcin in patients with hyperthyroidism, hypothyroidism and subacute thyroiditis. J Endocrinol Invest. 1992;15:491–496. doi: 10.1007/BF03348786. [DOI] [PubMed] [Google Scholar]
  • 89.Spillmann D, Burger MM. Carbohydrate-carbohydrate interactions in adhesion. J Cell Biochem. 1996;61:562–568. doi: 10.1002/(sici)1097-4644(19960616)61:4<562::aid-jcb9>3.0.co;2-m. [DOI] [PubMed] [Google Scholar]
  • 90.Brown D. GPI-anchored proteins and detergent-resistant membrane domains. Braz J Med Biol Res. 1994;27:309–315. [PubMed] [Google Scholar]
  • 91.Jacobson K, Mouritsen OG, Anderson RG. Lipid rafts: at a crossroad between cell biology and physics. Nat Cell Biol. 2007;9:7–14. doi: 10.1038/ncb0107-7. [DOI] [PubMed] [Google Scholar]
  • 92.Isshiki M, Anderson RG. Calcium signal transduction from caveolae. Cell Calcium. 1999;26:201–208. doi: 10.1054/ceca.1999.0073. [DOI] [PubMed] [Google Scholar]
  • 93.Simons K, Ikonen E. Functional rafts in cell membranes. Nature. 1997;387:569–572. doi: 10.1038/42408. [DOI] [PubMed] [Google Scholar]
  • 94.Smart EJ, Graf GA, McNiven MA, Sessa WC, Engelman JA, Scherer PE, Okamoto T, Lisanti MP. Caveolins, liquid-ordered domains, and signal transduction. Mol Cell Biol. 1999;19:7289–7304. doi: 10.1128/mcb.19.11.7289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Bagnat M, Chang A, Simons K. Plasma membrane proton ATPase Pma1p requires raft association for surface delivery in yeast. Mol Biol Cell. 2001;12:4129–4138. doi: 10.1091/mbc.12.12.4129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Lee MC, Hamamoto S, Schekman R. Ceramide biosynthesis is required for the formation of the oligomeric H+-ATPase Pma1p in the yeast endoplasmic reticulum. J Biol Chem. 2002;277:22395–22401. doi: 10.1074/jbc.M200450200. [DOI] [PubMed] [Google Scholar]
  • 97.Pizzirusso M, Chang A. Ubiquitin-mediated targeting of a mutant plasma membrane ATPase, Pma1-7, to the endosomal/vacuolar system in yeast. Mol Biol Cell. 2004;15:2401–2409. doi: 10.1091/mbc.E03-10-0727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Wang Q, Chang A. Sphingoid base synthesis is required for oligomerization and cell surface stability of the yeast plasma membrane ATPase, Pma1. Proc Natl Acad Sci U S A. 2002;99:12853–12858. doi: 10.1073/pnas.202115499. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Watanabe R, Funato K, Venkataraman K, Futerman AH, Riezman H. Sphingolipids are required for the stable membrane association of glycosylphosphatidylinositol-anchored proteins in yeast. J Biol Chem. 2002;277:49538–49544. doi: 10.1074/jbc.M206209200. [DOI] [PubMed] [Google Scholar]
  • 100.Mukhopadhyay K, Prasad T, Saini P, Pucadyil TJ, Chattopadhyay A, Prasad R. Membrane sphingolipid-ergosterol interactions are important determinants of multidrug resistance in Candida albicans. Antimicrob Agents Chemother. 2004;48:1778–1787. doi: 10.1128/AAC.48.5.1778-1787.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Pasrija R, Prasad T, Prasad R. Membrane raft lipid constituents affect drug susceptibilities of Candida albicans. Biochem Soc Trans. 2005;33:1219–1223. doi: 10.1042/BST20051219. [DOI] [PubMed] [Google Scholar]
  • 102.Sanglard D, Kuchler K, Ischer F, Pagani JL, Monod M, Bille J. Mechanisms of resistance to azole antifungal agents in Candida albicans isolates from AIDS patients involve specific multidrug transporters. Antimicrob Agents Chemother. 1995;39:2378–2386. doi: 10.1128/aac.39.11.2378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Prasad R, De Wergifosse P, Goffeau A, Balzi E. Molecular cloning and characterization of a novel gene of Candida albicans, CDR1, conferring multiple resistance to drugs and antifungals. Curr Genet. 1995;27:320–329. doi: 10.1007/BF00352101. [DOI] [PubMed] [Google Scholar]
  • 104.Zhang K, Showalter M, Revollo J, Hsu FF, Turk J, Beverley SM. Sphingolipids are essential for differentiation but not growth in Leishmania. Embo J. 2003;22:6016–6026. doi: 10.1093/emboj/cdg584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Denny PW, Goulding D, Ferguson MA, Smith DF. Sphingolipid-free Leishmania are defective in membrane trafficking, differentiation and infectivity. Mol Microbiol. 2004;52:313–327. doi: 10.1111/j.1365-2958.2003.03975.x. [DOI] [PubMed] [Google Scholar]
  • 106.Schuck S, Honsho M, Ekroos K, Shevchenko A, Simons K. Resistance of cell membranes to different detergents. Proc Natl Acad Sci U S A. 2003;100:5795–5800. doi: 10.1073/pnas.0631579100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Munro S. Lipid rafts: elusive or illusive? Cell. 2003;115:377–388. doi: 10.1016/s0092-8674(03)00882-1. [DOI] [PubMed] [Google Scholar]
  • 108.Giusto NM, Roque ME, Ilincheta de Boschero MG. Effects of aging on the content, composition and synthesis of sphingomyelin in the central nervous system. Lipids. 1992;27:835–839. doi: 10.1007/BF02535859. [DOI] [PubMed] [Google Scholar]
  • 109.Lightle SA, Oakley JI, Nikolova-Karakashian MN. Activation of sphingolipid turnover and chronic generation of ceramide and sphingosine in liver during aging. Mech Ageing Dev. 2000;120:111–125. doi: 10.1016/s0047-6374(00)00191-3. [DOI] [PubMed] [Google Scholar]
  • 110.Tacconi MT, Lligona L, Salmona M, Pitsikas N, Algeri S. Aging and food restriction: effect on lipids of cerebral cortex. Neurobiol Aging. 1991;12:55–59. doi: 10.1016/0197-4580(91)90039-m. [DOI] [PubMed] [Google Scholar]
  • 111.Rao RP, Yuan C, Allegood JC, Rawat SS, Edwards MB, Wang X, Merrill AH, Jr., Acharya U, Acharya JK. Ceramide transfer protein function is essential for normal oxidative stress response and lifespan. Proc Natl Acad Sci U S A. 2007;104:11364–11369. doi: 10.1073/pnas.0705049104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Fukasawa M, Nishijima M, 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. 1999;144:673–685. doi: 10.1083/jcb.144.4.673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Fukasawa M, Nishijima M, Itabe H, Takano T, Hanada K. Reduction of sphingomyelin level without accumulation of ceramide in Chinese hamster ovary cells affects detergent-resistant membrane domains and enhances cellular cholesterol efflux to methyl-beta -cyclodextrin. J Biol Chem. 2000;275:34028–34034. doi: 10.1074/jbc.M005151200. [DOI] [PubMed] [Google Scholar]
  • 114.Funakoshi T, Yasuda S, Fukasawa M, Nishijima M, 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. 2000;275:29938–29945. doi: 10.1074/jbc.M004470200. [DOI] [PubMed] [Google Scholar]
  • 115.Acharya U, Edwards MB, Jorquera RA, Silva H, Nagashima K, Labarca P, Acharya JK. Drosophila melanogaster Scramblases modulate synaptic transmission. J Cell Biol. 2006;173:69–82. doi: 10.1083/jcb.200506159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Halliwell B, Gutteridge JMC. Free radicals in biology and medicine. 3rd. Oxford University Press; Oxford; New York: 1999. [Google Scholar]
  • 117.Koike T, Ishida G, Taniguchi M, Higaki K, Ayaki Y, Saito M, Sakakihara Y, Iwamori M, Ohno K. Decreased membrane fluidity and unsaturated fatty acids in Niemann-Pick disease type C fibroblasts. Biochim Biophys Acta. 1998;1406:327–335. doi: 10.1016/s0925-4439(98)00019-2. [DOI] [PubMed] [Google Scholar]
  • 118.Brasitus TA, Schachter D. Lipid dynamics and lipid-protein interactions in rat enterocyte basolateral and microvillus membranes. Biochemistry. 1980;19:2763–2769. doi: 10.1021/bi00553a035. [DOI] [PubMed] [Google Scholar]
  • 119.Stubbs CD, Ketterer B, Hicks RM. The isolation and analysis of the luminal plasma membrane of calf urinary bladder epithelium. Biochim Biophys Acta. 1979;558:58–72. doi: 10.1016/0005-2736(79)90315-8. [DOI] [PubMed] [Google Scholar]
  • 120.Sillence DJ, Platt FM. Glycosphingolipids in endocytic membrane transport. Semin Cell Dev Biol. 2004;15:409–416. doi: 10.1016/j.semcdb.2004.03.004. [DOI] [PubMed] [Google Scholar]
  • 121.Sillence DJ, Platt FM. Introduction: Glycosphingolipids in cell biology and disease. Semin Cell Dev Biol. 2004;15:371–373. doi: 10.1016/j.semcdb.2004.03.008. [DOI] [PubMed] [Google Scholar]
  • 122.van Meer G, Sprong H. Membrane lipids and vesicular traffic. Curr Opin Cell Biol. 2004;16:373–378. doi: 10.1016/j.ceb.2004.06.004. [DOI] [PubMed] [Google Scholar]
  • 123.Kasahara K, Watanabe K, Takeuchi K, Kaneko H, Oohira A, Yamamoto T, Sanai Y. Involvement of gangliosides in glycosylphosphatidylinositol-anchored neuronal cell adhesion molecule TAG-1 signaling in lipid rafts. J Biol Chem. 2000;275:34701–34709. doi: 10.1074/jbc.M003163200. [DOI] [PubMed] [Google Scholar]
  • 124.Hakomori S. Bifunctional role of glycosphingolipids. Modulators for transmembrane signaling and mediators for cellular interactions. J Biol Chem. 1990;265:18713–18716. [PubMed] [Google Scholar]
  • 125.Kawai H, Allende ML, Wada R, Kono M, Sango K, Deng C, Miyakawa T, Crawley JN, Werth N, Bierfreund U, Sandhoff K, Proia RL. Mice expressing only monosialoganglioside GM3 exhibit lethal audiogenic seizures. J Biol Chem. 2001;276:6885–6888. doi: 10.1074/jbc.C000847200. [DOI] [PubMed] [Google Scholar]
  • 126.Coetzee T, Fujita N, Dupree J, Shi R, Blight A, Suzuki K, Suzuki K, Popko B. Myelination in the absence of galactocerebroside and sulfatide: normal structure with abnormal function and regional instability. Cell. 1996;86:209–219. doi: 10.1016/s0092-8674(00)80093-8. [DOI] [PubMed] [Google Scholar]
  • 127.Flesch FM, Gadella BM. Dynamics of the mammalian sperm plasma membrane in the process of fertilization. Biochim Biophys Acta. 2000;1469:197–235. doi: 10.1016/s0304-4157(00)00018-6. [DOI] [PubMed] [Google Scholar]
  • 128.Alonso TS, Bonini de Romanelli IC, Pechen de D’Angelo AM, Bazan NG. Dynamics of cellular membranes during amphibian fertilization and early development. Prog Clin Biol Res. 1982;112(Pt B):55–62. [PubMed] [Google Scholar]
  • 129.Butler A, He X, Gordon RE, Wu HS, Gatt S, Schuchman EH. Reproductive pathology and sperm physiology in acid sphingomyelinase-deficient mice. Am J Pathol. 2002;161:1061–1075. doi: 10.1016/S0002-9440(10)64267-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Butler A, Gordon RE, Gatt S, Schuchman EH. Sperm Abnormalities in Heterozygous Acid Sphingomyelinase Knockout Mice Reveal a Novel Approach for the Prevention of Genetic Diseases. Am J Pathol. 2007 doi: 10.2353/ajpath.2007.061002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Suzuki K, Suzuki Y. Globoid cell leucodystrophy (Krabbe’s disease): deficiency of galactocerebroside beta-galactosidase. Proc Natl Acad Sci U S A. 1970;66:302–309. doi: 10.1073/pnas.66.2.302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Kolter T, Sandhoff K. Sphingolipid metabolism diseases. Biochim Biophys Acta. 2006;1758:2057–2079. doi: 10.1016/j.bbamem.2006.05.027. [DOI] [PubMed] [Google Scholar]
  • 133.Luddi A, Strazza M, Carbone M, Moretti E, Costantino-Ceccarini E. Galactosylceramidase deficiency causes sperm abnormalities in the mouse model of globoid cell leukodystrophy. Exp Cell Res. 2005;304:59–68. doi: 10.1016/j.yexcr.2004.10.034. [DOI] [PubMed] [Google Scholar]
  • 134.Park JH, Schuchman EH. Acid ceramidase and human disease. Biochim Biophys Acta. 2006;1758:2133–2138. doi: 10.1016/j.bbamem.2006.08.019. [DOI] [PubMed] [Google Scholar]
  • 135.Murase T, Imaeda N, Kondoh N, Tsubota T. Ceramide enhances acrosomal exocytosis triggered by calcium and the calcium ionophore A23187 in boar spermatozoa. J Reprod Dev. 2004;50:667–674. doi: 10.1262/jrd.50.667. [DOI] [PubMed] [Google Scholar]
  • 136.Fujimoto H, Tadano-Aritomi K, Tokumasu A, Ito K, Hikita T, Suzuki K, Ishizuka I. Requirement of seminolipid in spermatogenesis revealed by UDP-galactose: Ceramide galactosyltransferase-deficient mice. J Biol Chem. 2000;275:22623–22626. doi: 10.1074/jbc.C000200200. [DOI] [PubMed] [Google Scholar]
  • 137.Kohyama-Koganeya A, Sasamura T, Oshima E, Suzuki E, Nishihara S, Ueda R, Hirabayashi Y. Drosophila glucosylceramide synthase: a negative regulator of cell death mediated by proapoptotic factors. J Biol Chem. 2004;279:35995–36002. doi: 10.1074/jbc.M400444200. [DOI] [PubMed] [Google Scholar]
  • 138.Wandall HH, Pizette S, Pedersen JW, Eichert H, Levery SB, Mandel U, Cohen SM, Clausen H. Egghead and brainiac are essential for glycosphingolipid biosynthesis in vivo. J Biol Chem. 2005;280:4858–4863. doi: 10.1074/jbc.C400571200. [DOI] [PubMed] [Google Scholar]
  • 139.Muller R, Altmann F, Zhou D, Hennet T. The Drosophila melanogaster brainiac protein is a glycolipid-specific beta 1,3N-acetylglucosaminyltransferase. J Biol Chem. 2002;277:32417–32420. doi: 10.1074/jbc.C200381200. [DOI] [PubMed] [Google Scholar]
  • 140.Goode S, Melnick M, Chou TB, Perrimon N. The neurogenic genes egghead and brainiac define a novel signaling pathway essential for epithelial morphogenesis during Drosophila oogenesis. Development. 1996;122:3863–3879. doi: 10.1242/dev.122.12.3863. [DOI] [PubMed] [Google Scholar]
  • 141.Goode S, Wright D, Mahowald AP. The neurogenic locus brainiac cooperates with the Drosophila EGF receptor to establish the ovarian follicle and to determine its dorsal-ventral polarity. Development. 1992;116:177–192. doi: 10.1242/dev.116.1.177. [DOI] [PubMed] [Google Scholar]
  • 142.Haines N, Irvine KD. Functional analysis of Drosophila beta1,4-N-acetlygalactosaminyltransferases. Glycobiology. 2005;15:335–346. doi: 10.1093/glycob/cwi017. [DOI] [PubMed] [Google Scholar]
  • 143.Haines N, Stewart BA. Functional roles for beta1,4-N-acetlygalactosaminyltransferase-A in Drosophila larval neurons and muscles. Genetics. 2007;175:671–679. doi: 10.1534/genetics.106.065565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Chen YW, Pedersen JW, Wandall HH, Levery SB, Pizette S, Clausen H, Cohen SM. Glycosphingolipids with extended sugar chain have specialized functions in development and behavior of Drosophila. Dev Biol. 2007;306:736–749. doi: 10.1016/j.ydbio.2007.04.013. [DOI] [PubMed] [Google Scholar]
  • 145.Pappu R, Schwab SR, Cornelissen I, Pereira JP, Regard JB, Xu Y, Camerer E, Zheng YW, Huang Y, Cyster JG, Coughlin SR. Promotion of lymphocyte egress into blood and lymph by distinct sources of sphingosine-1-phosphate. Science. 2007;316:295–298. doi: 10.1126/science.1139221. [DOI] [PubMed] [Google Scholar]
  • 146.Hla T. Sphingosine 1-phosphate receptors. Prostaglandins Other Lipid Mediat. 2001;64:135–142. doi: 10.1016/s0090-6980(01)00109-5. [DOI] [PubMed] [Google Scholar]
  • 147.Olivera A, Kohama T, Tu Z, Milstien S, Spiegel S. Purification and characterization of rat kidney sphingosine kinase. J Biol Chem. 1998;273:12576–12583. doi: 10.1074/jbc.273.20.12576. [DOI] [PubMed] [Google Scholar]
  • 148.Kohama T, Olivera A, Edsall L, Nagiec MM, Dickson R, Spiegel S. Molecular cloning and functional characterization of murine sphingosine kinase. J Biol Chem. 1998;273:23722–23728. doi: 10.1074/jbc.273.37.23722. [DOI] [PubMed] [Google Scholar]
  • 149.Brinkmann V. Sphingosine 1-phosphate receptors in health and disease: Mechanistic insights from gene deletion studies and reverse pharmacology. Pharmacol Ther. 2007;115:84–105. doi: 10.1016/j.pharmthera.2007.04.006. [DOI] [PubMed] [Google Scholar]
  • 150.Hla T. Physiological and pathological actions of sphingosine 1-phosphate. Semin Cell Dev Biol. 2004;15:513–520. doi: 10.1016/j.semcdb.2004.05.002. [DOI] [PubMed] [Google Scholar]
  • 151.Hla T, Maciag T. An abundant transcript induced in differentiating human endothelial cells encodes a polypeptide with structural similarities to G-protein-coupled receptors. J Biol Chem. 1990;265:9308–9313. [PubMed] [Google Scholar]
  • 152.Ancellin N, Hla T. Differential pharmacological properties and signal transduction of the sphingosine 1-phosphate receptors EDG-1, EDG-3, and EDG-5. J Biol Chem. 1999;274:18997–19002. doi: 10.1074/jbc.274.27.18997. [DOI] [PubMed] [Google Scholar]
  • 153.Lee MJ, Van Brocklyn JR, Thangada S, Liu CH, Hand AR, Menzeleev R, Spiegel S, Hla T. Sphingosine-1-phosphate as a ligand for the G protein-coupled receptor EDG-1. Science. 1998;279:1552–1555. doi: 10.1126/science.279.5356.1552. [DOI] [PubMed] [Google Scholar]
  • 154.Van Brocklyn JR, Lee MJ, Menzeleev R, Olivera A, Edsall L, Cuvillier O, Thomas DM, Coopman PJ, Thangada S, Liu CH, Hla T, Spiegel S. Dual actions of sphingosine-1-phosphate: extracellular through the Gi-coupled receptor Edg-1 and intracellular to regulate proliferation and survival. J Cell Biol. 1998;142:229–240. doi: 10.1083/jcb.142.1.229. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Rapizzi E, Donati C, Cencetti F, Pinton P, Rizzuto R, Bruni P. Sphingosine 1-phosphate receptors modulate intracellular Ca2+ homeostasis. Biochem Biophys Res Commun. 2007;353:268–274. doi: 10.1016/j.bbrc.2006.12.010. [DOI] [PubMed] [Google Scholar]
  • 156.Sanchez T, Hla T. Structural and functional characteristics of S1P receptors. J Cell Biochem. 2004;92:913–922. doi: 10.1002/jcb.20127. [DOI] [PubMed] [Google Scholar]
  • 157.Kluk MJ, Hla T. Signaling of sphingosine-1-phosphate via the S1P/EDG-family of G-protein-coupled receptors. Biochim Biophys Acta. 2002;1582:72–80. doi: 10.1016/s1388-1981(02)00139-7. [DOI] [PubMed] [Google Scholar]
  • 158.Graler MH, Bernhardt G, Lipp M. EDG6, a novel G-protein-coupled receptor related to receptors for bioactive lysophospholipids, is specifically expressed in lymphoid tissue. Genomics. 1998;53:164–169. doi: 10.1006/geno.1998.5491. [DOI] [PubMed] [Google Scholar]
  • 159.Van Brocklyn JR, Tu Z, Edsall LC, Schmidt RR, Spiegel S. Sphingosine 1-phosphate-induced cell rounding and neurite retraction are mediated by the G protein-coupled receptor H218. J Biol Chem. 1999;274:4626–4632. doi: 10.1074/jbc.274.8.4626. [DOI] [PubMed] [Google Scholar]
  • 160.Graler MH, Grosse R, Kusch A, Kremmer E, Gudermann T, Lipp M. The sphingosine 1-phosphate receptor S1P4 regulates cell shape and motility via coupling to Gi and G12/13. J Cell Biochem. 2003;89:507–519. doi: 10.1002/jcb.10537. [DOI] [PubMed] [Google Scholar]
  • 161.Im DS, Heise CE, Ancellin N, O’Dowd BF, Shei GJ, Heavens RP, Rigby MR, Hla T, Mandala S, McAllister G, George SR, Lynch KR. Characterization of a novel sphingosine 1-phosphate receptor, Edg-8. J Biol Chem. 2000;275:14281–14286. doi: 10.1074/jbc.275.19.14281. [DOI] [PubMed] [Google Scholar]
  • 162.Van Brocklyn JR, Graler MH, Bernhardt G, Hobson JP, Lipp M, Spiegel S. Sphingosine-1-phosphate is a ligand for the G protein-coupled receptor EDG-6. Blood. 2000;95:2624–2629. [PubMed] [Google Scholar]
  • 163.Windh RT, Lee MJ, Hla T, An S, Barr AJ, Manning DR. Differential coupling of the sphingosine 1-phosphate receptors Edg-1, Edg-3, and H218/Edg-5 to the G(i), G(q), and G(12) families of heterotrimeric G proteins. J Biol Chem. 1999;274:27351–27358. doi: 10.1074/jbc.274.39.27351. [DOI] [PubMed] [Google Scholar]
  • 164.Paik JH, Skoura A, Chae SS, Cowan AE, Han DK, Proia RL, Hla T. Sphingosine 1-phosphate receptor regulation of N-cadherin mediates vascular stabilization. Genes Dev. 2004;18:2392–2403. doi: 10.1101/gad.1227804. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Mizugishi K, Yamashita T, Olivera A, Miller GF, Spiegel S, Proia RL. Essential role for sphingosine kinases in neural and vascular development. Mol Cell Biol. 2005;25:11113–11121. doi: 10.1128/MCB.25.24.11113-11121.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Sato K, Tomura H, Igarashi Y, Ui M, Okajima F. Exogenous sphingosine 1-phosphate induces neurite retraction possibly through a cell surface receptor in PC12 cells. Biochem Biophys Res Commun. 1997;240:329–334. doi: 10.1006/bbrc.1997.7666. [DOI] [PubMed] [Google Scholar]
  • 167.Rouach N, Pebay A, Meme W, Cordier J, Ezan P, Etienne E, Giaume C, Tence M. S1P inhibits gap junctions in astrocytes: involvement of G and Rho GTPase/ROCK. Eur J Neurosci. 2006;23:1453–1464. doi: 10.1111/j.1460-9568.2006.04671.x. [DOI] [PubMed] [Google Scholar]
  • 168.Lee MJ, Thangada S, Claffey KP, Ancellin N, Liu CH, Kluk M, Volpi M, Sha’afi RI, Hla T. Vascular endothelial cell adherens junction assembly and morphogenesis induced by sphingosine-1-phosphate. Cell. 1999;99:301–312. doi: 10.1016/s0092-8674(00)81661-x. [DOI] [PubMed] [Google Scholar]
  • 169.Okamoto H, Takuwa N, Yokomizo T, Sugimoto N, Sakurada S, Shigematsu H, Takuwa Y. Inhibitory regulation of Rac activation, membrane ruffling, and cell migration by the G protein-coupled sphingosine-1-phosphate receptor EDG5 but not EDG1 or EDG3. Mol Cell Biol. 2000;20:9247–9261. doi: 10.1128/mcb.20.24.9247-9261.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Paik JH, Chae S, Lee MJ, Thangada S, Hla T. Sphingosine 1-phosphate-induced endothelial cell migration requires the expression of EDG-1 and EDG-3 receptors and Rho-dependent activation of alpha vbeta3- and beta1-containing integrins. J Biol Chem. 2001;276:11830–11837. doi: 10.1074/jbc.M009422200. [DOI] [PubMed] [Google Scholar]
  • 171.Singleton PA, Dudek SM, Chiang ET, Garcia JG. Regulation of sphingosine 1-phosphate-induced endothelial cytoskeletal rearrangement and barrier enhancement by S1P1 receptor, PI3 kinase, Tiam1/Rac1, and alpha-actinin. Faseb J. 2005;19:1646–1656. doi: 10.1096/fj.05-3928com. [DOI] [PubMed] [Google Scholar]
  • 172.Allende ML, Proia RL. Sphingosine-1-phosphate receptors and the development of the vascular system. Biochim Biophys Acta. 2002;1582:222–227. doi: 10.1016/s1388-1981(02)00175-0. [DOI] [PubMed] [Google Scholar]
  • 173.Saba JD, Hla T. Point-counterpoint of sphingosine 1-phosphate metabolism. Circ Res. 2004;94:724–734. doi: 10.1161/01.RES.0000122383.60368.24. [DOI] [PubMed] [Google Scholar]
  • 174.Liu Y, Wada R, Yamashita T, Mi Y, Deng CX, Hobson JP, Rosenfeldt HM, Nava VE, Chae SS, Lee MJ, Liu CH, Hla T, Spiegel S, Proia RL. Edg-1, the G protein-coupled receptor for sphingosine-1-phosphate, is essential for vascular maturation. J Clin Invest. 2000;106:951–961. doi: 10.1172/JCI10905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.McVerry BJ, Garcia JG. Endothelial cell barrier regulation by sphingosine 1-phosphate. J Cell Biochem. 2004;92:1075–1085. doi: 10.1002/jcb.20088. [DOI] [PubMed] [Google Scholar]
  • 176.Matloubian M, Lo CG, Cinamon G, Lesneski MJ, Xu Y, Brinkmann V, Allende ML, Proia RL, Cyster JG. Lymphocyte egress from thymus and peripheral lymphoid organs is dependent on S1P receptor 1. Nature. 2004;427:355–360. doi: 10.1038/nature02284. [DOI] [PubMed] [Google Scholar]
  • 177.Allende ML, Dreier JL, Mandala S, Proia RL. Expression of the sphingosine 1-phosphate receptor, S1P1, on T-cells controls thymic emigration. J Biol Chem. 2004;279:15396–15401. doi: 10.1074/jbc.M314291200. [DOI] [PubMed] [Google Scholar]
  • 178.Watterson KR, Ratz PH, Spiegel S. The role of sphingosine-1-phosphate in smooth muscle contraction. Cell Signal. 2005;17:289–298. doi: 10.1016/j.cellsig.2004.09.013. [DOI] [PubMed] [Google Scholar]
  • 179.Oskeritzian CA, Milstien S, Spiegel S. Sphingosine-1-phosphate in allergic responses, asthma and anaphylaxis. Pharmacol Ther. 2007 doi: 10.1016/j.pharmthera.2007.05.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.MacLennan AJ, Carney PR, Zhu WJ, Chaves AH, Garcia J, Grimes JR, Anderson KJ, Roper SN, Lee N. An essential role for the H218/AGR16/Edg-5/LP(B2) sphingosie 1-phosphate receptor in neuronal excitability. Eur J Neurosci. 2001;14:203–209. doi: 10.1046/j.0953-816x.2001.01634.x. [DOI] [PubMed] [Google Scholar]
  • 181.Kupperman E, An S, Osborne N, Waldron S, Stainier DY. A sphingosine-1-phosphate receptor regulates cell migration during vertebrate heart development. Nature. 2000;406:192–195. doi: 10.1038/35018092. [DOI] [PubMed] [Google Scholar]
  • 182.Sanna MG, Wang SK, Gonzalez-Cabrera PJ, Don A, Marsolais D, Matheu MP, Wei SH, Parker I, Jo E, Cheng WC, Cahalan MD, Wong CH, Rosen H. Enhancement of capillary leakage and restoration of lymphocyte egress by a chiral S1P1 antagonist in vivo. Nat Chem Biol. 2006;2:434–441. doi: 10.1038/nchembio804. [DOI] [PubMed] [Google Scholar]
  • 183.Gon Y, Wood MR, Kiosses WB, Jo E, Sanna MG, Chun J, Rosen H. S1P3 receptor-induced reorganization of epithelial tight junctions compromises lung barrier integrity and is potentiated by TNF. Proc Natl Acad Sci U S A. 2005;102:9270–9275. doi: 10.1073/pnas.0501997102. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 184.Forrest M, Sun SY, Hajdu R, Bergstrom J, Card D, Doherty G, Hale J, Keohane C, Meyers C, Milligan J, Mills S, Nomura N, Rosen H, Rosenbach M, Shei GJ, Singer II, Tian M, West S, White V, Xie J, Proia RL, Mandala S. Immune cell regulation and cardiovascular effects of sphingosine 1-phosphate receptor agonists in rodents are mediated via distinct receptor subtypes. J Pharmacol Exp Ther. 2004;309:758–768. doi: 10.1124/jpet.103.062828. [DOI] [PubMed] [Google Scholar]
  • 185.Kono M, Mi Y, Liu Y, Sasaki T, Allende ML, Wu YP, Yamashita T, Proia RL. The sphingosine-1-phosphate receptors S1P1, 1P2, and S1P3 function coordinately during embryonic angiogenesis. J Biol Chem. 2004;279:29367–29373. doi: 10.1074/jbc.M403937200. [DOI] [PubMed] [Google Scholar]
  • 186.Allende ML, Sasaki T, Kawai H, Olivera A, Mi Y, van Echten-Deckert G, Hajdu R, Rosenbach M, Keohane CA, Mandala S, Spiegel S, Proia RL. Mice deficient in sphingosine kinase 1 are rendered lymphopenic by FTY720. J Biol Chem. 2004;279:52487–52492. doi: 10.1074/jbc.M406512200. [DOI] [PubMed] [Google Scholar]
  • 187.Michaud J, Kohno M, Proia RL, Hla T. Normal acute and chronic inflammatory responses in sphingosine kinase 1 knockout mice. FEBS Lett. 2006;580:4607–4612. doi: 10.1016/j.febslet.2006.07.035. [DOI] [PubMed] [Google Scholar]
  • 188.Kharel Y, Lee S, Snyder AH, Sheasley-O’neill SL, Morris MA, Setiady Y, Zhu R, Zigler MA, Burcin TL, Ley K, Tung KS, Engelhard VH, Macdonald TL, Pearson-White S, Lynch KR. Sphingosine kinase 2 is required for modulation of lymphocyte traffic by FTY720. J Biol Chem. 2005;280:36865–36872. doi: 10.1074/jbc.M506293200. [DOI] [PubMed] [Google Scholar]
  • 189.Zemann B, Kinzel B, Muller M, Reuschel R, Mechtcheriakova D, Urtz N, Bornancin F, Baumruker T, Billich A. Sphingosine kinase type 2 is essential for lymphopenia induced by the immunomodulatory drug FTY720. Blood. 2006;107:1454–1458. doi: 10.1182/blood-2005-07-2628. [DOI] [PubMed] [Google Scholar]
  • 190.Olivera A, Mizugishi K, Tikhonova A, Ciaccia L, Odom S, Proia RL, Rivera J. The sphingosine kinase-sphingosine-1-phosphate axis is a determinant of mast cell function and anaphylaxis. Immunity. 2007;26:287–297. doi: 10.1016/j.immuni.2007.02.008. [DOI] [PubMed] [Google Scholar]
  • 191.Tybulewicz VL, Tremblay ML, LaMarca ME, Willemsen R, Stubblefield BK, Winfield S, Zablocka B, Sidransky E, Martin BM, Huang SP, et al. Animal model of Gaucher’s disease from targeted disruption of the mouse glucocerebrosidase gene. Nature. 1992;357:407–410. doi: 10.1038/357407a0. [DOI] [PubMed] [Google Scholar]
  • 192.Miklyaeva EI, Dong W, Bureau A, Fattahie R, Xu Y, Su M, Fick GH, Huang JQ, Igdoura S, Hanai N, Gravel RA. Late onset Tay-Sachs disease in mice with targeted disruption of the Hexa gene: behavioral changes and pathology of the central nervous system. Brain Res. 2004;1001:37–50. doi: 10.1016/j.brainres.2003.11.067. [DOI] [PubMed] [Google Scholar]
  • 193.Sango K, Yamanaka S, Hoffmann A, Okuda Y, Grinberg A, Westphal H, McDonald MP, Crawley JN, Sandhoff K, Suzuki K, Proia RL. Mouse models of Tay-Sachs and Sandhoff diseases differ in neurologic phenotype and ganglioside metabolism. Nat Genet. 1995;11:170–176. doi: 10.1038/ng1095-170. [DOI] [PubMed] [Google Scholar]
  • 194.Ohshima T, Murray GJ, Swaim WD, Longenecker G, Quirk JM, Cardarelli CO, Sugimoto Y, Pastan I, Gottesman MM, Brady RO, Kulkarni AB. alpha-Galactosidase A deficient mice: a model of Fabry disease. Proc Natl Acad Sci U S A. 1997;94:2540–2544. doi: 10.1073/pnas.94.6.2540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Ohshima T, Schiffmann R, Murray GJ, Kopp J, Quirk JM, Stahl S, Chan CC, Zerfas P, Tao-Cheng JH, Ward JM, Brady RO, Kulkarni AB. Aging accentuates and bone marrow transplantation ameliorates metabolic defects in Fabry disease mice. Proc Natl Acad Sci U S A. 1999;96:6423–6427. doi: 10.1073/pnas.96.11.6423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Ezoe T, Vanier MT, Oya Y, Popko B, Tohyama J, Matsuda J, Suzuki K, Suzuki K. Biochemistry and neuropathology of mice doubly deficient in synthesis and degradation of galactosylceramide. J Neurosci Res. 2000;59:170–178. doi: 10.1002/(sici)1097-4547(20000115)59:2<170::aid-jnr3>3.0.co;2-g. [DOI] [PubMed] [Google Scholar]
  • 197.Eliyahu E, Park JH, Shtraizent N, He X, Schuchman EH. Acid ceramidase is a novel factor required for early embryo survival. Faseb J. 2007;21:1403–1409. doi: 10.1096/fj.06-7016com. [DOI] [PubMed] [Google Scholar]
  • 198.Li CM, Park JH, Simonaro CM, He X, Gordon RE, Friedman AH, Ehleiter D, Paris F, Manova K, Hepbildikler S, Fuks Z, Sandhoff K, Kolesnick R, Schuchman EH. Insertional mutagenesis of the mouse acid ceramidase gene leads to early embryonic lethality in homozygotes and progressive lipid storage disease in heterozygotes. Genomics. 2002;79:218–224. doi: 10.1006/geno.2002.6686. [DOI] [PubMed] [Google Scholar]
  • 199.Hess B, Saftig P, Hartmann D, Coenen R, Lullmann-Rauch R, Goebel HH, Evers M, von Figura K, D’Hooge R, Nagels G, De Deyn P, Peters C, Gieselmann V. Phenotype of arylsulfatase A-deficient mice: relationship to human metachromatic leukodystrophy. Proc Natl Acad Sci U S A. 1996;93:14821–14826. doi: 10.1073/pnas.93.25.14821. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Lutjohann D, Harzer K, Gieselmann V, Eckhardt M. Reduced brain cholesterol content in arylsulfatase A-deficient mice. Biochem Biophys Res Commun. 2006;344:647–650. doi: 10.1016/j.bbrc.2006.03.186. [DOI] [PubMed] [Google Scholar]
  • 201.Otterbach B, Stoffel W. Acid sphingomyelinase-deficient mice mimic the neurovisceral form of human lysosomal storage disease (Niemann-Pick disease) Cell. 1995;81:1053–1061. doi: 10.1016/s0092-8674(05)80010-8. [DOI] [PubMed] [Google Scholar]
  • 202.Pentchev PG, Gal AE, Booth AD, Omodeo-Sale F, Fouks J, Neumeyer BA, Quirk JM, Dawson G, Brady RO. A lysosomal storage disorder in mice characterized by a dual deficiency of sphingomyelinase and glucocerebrosidase. Biochim Biophys Acta. 1980;619:669–679. doi: 10.1016/0005-2760(80)90116-2. [DOI] [PubMed] [Google Scholar]
  • 203.Kono M, Dreier JL, Ellis JM, Allende ML, Kalkofen DN, Sanders KM, Bielawski J, Bielawska A, Hannun YA, Proia RL. Neutral ceramidase encoded by the Asah2 gene is essential for the intestinal degradation of sphingolipids. J Biol Chem. 2006;281:7324–7331. doi: 10.1074/jbc.M508382200. [DOI] [PubMed] [Google Scholar]
  • 204.Fujita N, Suzuki K, Vanier MT, Popko B, Maeda N, Klein A, Henseler M, Sandhoff K, Nakayasu H, Suzuki K. Targeted disruption of the mouse sphingolipid activator protein gene: a complex phenotype, including severe leukodystrophy and wide-spread storage of multiple sphingolipids. Hum Mol Genet. 1996;5:711–725. doi: 10.1093/hmg/5.6.711. [DOI] [PubMed] [Google Scholar]
  • 205.Matsuda J, Vanier MT, Saito Y, Tohyama J, Suzuki K, Suzuki K. A mutation in the saposin A domain of the sphingolipid activator protein (prosaposin) gene results in a late-onset, chronic form of globoid cell leukodystrophy in the mouse. Hum Mol Genet. 2001;10:1191–1199. doi: 10.1093/hmg/10.11.1191. [DOI] [PubMed] [Google Scholar]
  • 206.Sun Y, Witte DP, Zamzow M, Ran H, Quinn B, Matsuda J, Grabowski GA. Combined saposin C and D deficiencies in mice lead to a neuronopathic phenotype, glucosylceramide and alpha-hydroxy ceramide accumulation, and altered prosaposin trafficking. Hum Mol Genet. 2007;16:957–971. doi: 10.1093/hmg/ddm040. [DOI] [PubMed] [Google Scholar]
  • 207.Jennemann R, Sandhoff R, Wang S, Kiss E, Gretz N, Zuliani C, Martin-Villalba A, Jager R, Schorle H, Kenzelmann M, Bonrouhi M, Wiegandt H, Grone HJ. Cell-specific deletion of glucosylceramide synthase in brain leads to severe neural defects after birth. Proc Natl Acad Sci U S A. 2005;102:12459–12464. doi: 10.1073/pnas.0500893102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Sheikh KA, Sun J, Liu Y, Kawai H, Crawford TO, Proia RL, Griffin JW, Schnaar RL. Mice lacking complex gangliosides develop Wallerian degeneration and myelination defects. Proc Natl Acad Sci U S A. 1999;96:7532–7537. doi: 10.1073/pnas.96.13.7532. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Yamashita T, Wada R, Sasaki T, Deng C, Bierfreund U, Sandhoff K, Proia RL. A vital role for glycosphingolipid synthesis during development and differentiation. Proc Natl Acad Sci U S A. 1999;96:9142–9147. doi: 10.1073/pnas.96.16.9142. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES