New Insights into Glycosphingolipid Functions—Storage, Lipid Rafts, and Translocators

https://doi.org/10.1016/S0074-7696(07)62003-8Get rights and content

Glycosphingolipids are key components of eukaryotic cellular membranes. Through their propensity to form lipid rafts, they are important in membrane transport and signaling. At the cell surface, they are required for caveolar‐mediated endocytosis, a process required for the action of many glycosphingolipid‐binding toxins. Glycosphingolipids also exist intracellularly, on both leaflets of organelle membranes. It is expected that dissecting the mechanisms of cell pathology seen in the glycosphingolipid storage diseases, where lysosomal glycosphingolipid degradation is defective, will reveal their functions. Disrupted cation gradients in Mucolipidosis type IV disease are interlinked with glycosphingolipid storage, defective rab 7 function, and the activation of autophagy. Relationships between drug translocators and glycosphingolipid synthesis are also discussed. Mass spectrometry of cell lines defective in drug transporters reveal clear differences in glycosphingolipid mass and fatty acid composition. The potential roles of glycosphingolipids in lipid raft formation, endocytosis, and cationic gradients are discussed.

Introduction

Eukaryotes and some bacteria contain a heterogeneous class of lipids called glycosphingolipids (GSLs). These lipids are composed of a ceramide backbone and a sugar headgroup (Fig. 1). The hydrophobic ceramide part, consisting of a sphingoid base and N‐linked fatty acid, inserts in a cellular membrane, whereas the hydrophilic sugar headgroup mostly faces the extracellular space. At the cellular level, GSLs organize protein function by forming glycosignaling domains, where clustering of specific proteins within bilayer‐spanning lipid rafts regulates their signal transduction (Hakomori and Igarashi, 1995). In addition, cells use the capacity of GSLs to form ordered domains to create selectivity in membrane sorting (Simons and van Meer, 1988).

GSLs exhibit a huge heterogeneity of structure. The sphingoid bases can vary in length, saturation, hydroxylation, and branching. The main sphingoid base in mammals is sphingosine (Karlsson et al., 1973). Sphinganine, which corresponds to sphingosine without the trans double bond, and phytosphingosine (hydroxyl‐sphinganine) are also common sphingoid bases. The fatty acid is amide‐linked to the amino group of the sphingoid base. The fatty acid species are cell‐type–dependent, can vary in length, saturation, and hydroxylation, but are mostly long (≥C16) and saturated. GSLs can be divided into two main classes based on the first sugar linked to the ceramide backbone, glucose or galactose.

The majority of glycosphingolipid synthesis in resting cells occurs from recycled breakdown products (e.g., sphingosine) derived from the lysosome (Gillard et al., 1998). It is unclear whether the lysosomal breakdown products, such as sphingosine, escape from lysosomes via a transporter or are able to passively slip across the lysosomal membrane. Transport to the endoplasmic reticulum (ER) may occur via discrete protein complexes in areas of close membrane contact between the late endosome and the ER (Holthuis 2005, Ko 2001). The accumulation of sphingosine in the lysosome of many GSL storage diseases (Rodriguez‐Lafrasse et al., 1994) suggests transport from the lysosome can be frustrated, but the consequences of this, if any, have yet to be determined. Once transported, sphingosine is phosphorylated at the ER, forming sphingosine‐1‐phosphate, an important intracellular messenger (Funato 2003, Ghosh 1994). Once dephosphorylated, sphingosine is then acylated to ceramide. In specialized cell types, luminal ceramide is then converted to GalCer. From the ER, ceramide can also follow the vesicular pathway to early Golgi compartments where it is converted to glucosylceramide (GlcCer). Alternatively, ceramide reaches the trans Golgi via CERT, a ceramide transfer protein that is able to extract ceramide from the ER and deliver it to the Golgi after docking to phosphatidylinositol‐4‐phosphate (Hanada et al., 2003). This pathway may also involve the membrane contacts between the ER and the trans Golgi (Ladinsky et al., 1994). Synthesis of the phosphosphingolipid sphingomyelin (SM) in the trans Golgi (Huitema et al., 2004) depends on this pathway for ceramide supply (Hanada et al., 2003). It has been reported that ER to Golgi transport of ceramide may be regulated via an ER sphingosine‐1‐phosphate phosphohydrolase. Anterograde transport of ceramide, probably via CERT, and at least some proteins are regulated by the degradation of ER sphingosine‐1‐phosphate (Giussani et al., 2006). Feedback regulation via the formation of sphingosine in the lysosome and lipid transport and synthesis is an interesting topic for future research.

In actively dividing cells “de novo” synthesis is more active and starts with condensation of serine and palmitoyl‐CoA to 3‐ketosphinganine (Fig. 2). Formation of sphinganine and subsequent acylation and desaturation on the cytosolic side of the ER produces ceramide without production of sphingosine as an intermediate. A variety of ceramide synthases have been identified with both sphingoid base (Venkataraman et al., 2002) and fatty acid preferences (LASS1–LASS6). LASS1 and LASS4 expression preferentially increase C18 ceramide, LASS2 mainly increases C16 and C24:1 ceramide, and LASS5 and LASS6 produce C16 ceramide (Mizutani 2005, Mizutani 2006, Riebeling 2003, Venkataraman 2002).

Ceramide is also the direct precursor for GlcCer. GlcCer is present in most eukaryotic cells and a few bacteria and serves as the major precursor for complex GSLs. It is synthesized by the UDP‐Glc:ceramide glucosyltransferase or GlcCer synthase CGlcT that occurs on the cytosolic side of Golgi membranes (Coste 1986, Futerman 1991, Jeckel 1992). Although the functions of GlcCer are unknown, in some yeast and bacteria, GlcCer may play an important role in the modulation of pH gradients; knocking out GlcCer synthase reduces growth in alkaline conditions (Rittershaus et al., 2006). Complex GSLs are made by the stepwise addition of individual sugars from their activated nucleotide precursors onto GlcCer. In mammals, the first reaction is the conversion of GlcCer to lactosylceramide by the LacCer synthase. This enzyme, like all further glycosyltransferases and sulfotransferases involved in the glycosylation of GSLs, acts in the Golgi lumen (Lannert et al., 1994). Sialoglycosphingolipids are also called “gangliosides,” and abbreviations have been assigned to them according to the number of sialic acids present and to their migration order in chromatography. Except for ganglioside GM4, which corresponds to NeuAcα2‐3GalCer, all gangliosides have LacCer as a precursor.

In contrast to sphinganine1‐phosphate, the synthesis of sphingosine‐ 1‐phosphate is dependent on the degradation of GSLs and sphingomyelin in late endosomes and lysosomes (Berdyshev et al., 2006). Once formed, sphingosine‐1‐phosphate is involved in signaling events such as the regulation of calcium gradients (Bagshaw 2005, Masgrau 2003). Termination of the signal proceeds via a sphingosine‐1‐phosphate phosphohydrolase (Le Stunff et al., 2002). The observation that exogenous addition of sphingosine‐1‐phosphate regulates cell proliferation (Zhang et al., 1991) is also consistent with the role of this second messenger in the control of proliferation. Overexpression of sphingosine kinase has been shown to activate autophagy, a process that occurs constitutively but is also activated in neurodegenerative diseases such as the secondary glycosphingolipid storage diseases (see Section IV.C). Whether there is any relationship between GSL storage and sphingosine‐1‐phosphate signaling has yet to be determined. Both sphingosine kinase 1 and sphingosine‐1‐phosphate phosphohydrolase modulate de novo sphingolipid synthesis (Berdyshev 2006, Le Stunff 2002). The interrelationships between sphingosine‐1‐phosphate signaling and the regulation of new membrane sphingolipid has yet to be determined.

Structural differences between GSLs and glycerolipids (Fig. 1) underlie the special behavior of GSLs in membranes. In GSLs, the region between the polar headgroup and the hydrophobic backbone contains chemical groups that can function both as hydrogen bond donor and hydrogen bond acceptor, in contrast to the glycerolipids, which only have hydrogen bond accepting properties in that part of the molecule (Pascher, 1976). Additional hydrogen bonding can occur between the sugar headgroups of GSLs. Another striking difference between GSLs and most glycerolipids is that the lipid chains are saturated over at least the first 15 carbons of both chains. In combination with the higher hydrogen bonding, the saturated nature results in denser packing. This is measured as an increased melting temperature, or Tm, which corresponds to the temperature above which a bilayer of a single lipid switches from a frozen state, the gel or solid‐ordered (so) phase, to a fluid state, the liquid‐crystalline or liquid‐disordered (ld) phase. In addition, cholesterol, a rigid and flat cylindrical lipid, also interacts preferentially with sphingolipids via van der Waals interactions (Boggs, 1987). In model membranes containing mixtures of a high Tm lipid and cholesterol, a fluid‐fluid phase separation has been observed between the ld and a liquid‐ordered lo phase (Recktenwald and McConnell, 1981).

Unfortunately, most evidence to date that domains enriched in GSLs exist in biological membranes is indirect (Harder, 2003). However, GSLs are clustered on erythrocytes (Thompson and Tillack, 1985), gangliosides GM1 and GM3 are concentrated in domains (Parton, 1994), and GM1 can be enriched in different domains than GM3 on the same cell (Gomez‐Mouton et al., 2001). Most indications for the lipid domain association have been obtained by using the detergent‐resistance criterion. Although this technique has had great prospective value for how membrane signaling may work (Harder, 2003), physical studies have suggested there is no straightforward physical basis for why extraction at 4 °C would provide information concerning the situation at 37 °C (Heerklotz, 2002).

To avoid detergent extraction, pulse electron paramagnetic resonance (EPR) spin‐labeling methods and single‐molecule optical techniques have been applied (Dietrich 2002, Kenworthy 2000, Subczynski 2003) to monitor the entry and exit of probe molecules in domains in model membranes or in living cells. Results obtained using single molecule microscopy of saturated fluorescent analogues of phosphatidylethanolamine (Schutz et al., 2000) suggest that lipid analogues can be used to detect small (50 nm), stable (0.5–2 min) cell surface microdomains that may be related to the well‐established annular lipids that surround membrane proteins. Overall results have provided evidence for small/unstable rafts in unstimulated cells and for larger stabilized rafts induced by oligomerization of GPI‐anchored proteins or ligand binding (Harder, 2003), in which some proteins preferentially partition. The size of the confining domain for a GPI‐anchored protein is reduced when cells are treated with inhibitors of GSL synthesis; other studies have shown increases in phospholipase C susceptibility as well as changes in expression, suggesting GSLs contribute to the physical properties of microdomains. Using a fluorescent analogue of lactosylceramide, which is taken up by caveolar‐dependent mechanism, has shown that this specific form of endocytosis is dependent on the level of endogenous GSL, suggesting that some level of cell surface GSL is necessary for this mechanism of endocytosis, which may be lipid raft‐mediated (Cheng et al., 2006c). These processes are discussed in Section II.B and C. In conclusion, many studies suggest GSL can be heterogeneously organized on living cells, and this has functional consequences, although there is no strong consensus yet on the size, shape, and dynamics of lipid rafts.

One of the major setbacks in glycosphingolipid research has been a lack of tools with which to analyze them. Glycosphingolipids show increasing hydrophilicity with increasing size of the sugar headgroup, this leads to an increasing ability to form monomers in aqueous solution, decreasing solubility in organic solvents. Hence, difficulty arises with complete extraction from tissues and cells. Once extracted, utility of a phase split to purify the lipid components; higher GSLs escape into the aqueous phase which then leads to unacceptable losses. Higher GSLs are also more prone to adhere to the side of plastic tubes. The following techniques overcome many of these problems.

For tissue homogenates, GSLs can be extracted by the addition of 3.2 vol of CHCl3/MeOH (1:2.2) for 10 min at room temperature, and the phases are split by the addition of 1 vol of CHCl3 and 1 vol of H2O (Bligh and Dyer, 1959). This quick extraction procedure gives similar recoveries of GSLs to other commonly used methods of lipid extraction (Miller Podraza 1992, van Echten 1990). Upper phase GSLs can then be recovered by SePak C18 columns (Sillence et al., 2000). Briefly, C18 SepPak columns (Waters, Milford, MA) are washed with 1 mL MeOH and 1 mL water. The CHCl3 phase was dried down and loaded in 50‐μL of CHCl3/MeOH/H2O (1:2.2:1). The corresponding aqueous phase was loaded onto the columns and washed with 5 × 1 mL water. GSLs are then eluted with 5 × 1 mL CHCl3:MeOH (1:3) and 1 mL of MeOH, and the samples are dried under nitrogen. At this point, the cholesterol and GSLs can be quantified as the following describes. Sometimes further purification is necessary (e.g., for mass spectrometry), in which case glycerophospholipids are removed from samples by saponification by the addition of 1 mL of chloroform and 1 mL of 0.2 M NaOH in methanol and incubated overnight at 37 °C and further purified by silicic acid chromatography (Vance and Sweeley, 1967). Columns are washed with 5 mL of CHCl3 to remove fatty acids, nonalkali labile sterols, and alkylglycerols. Neutral GSLs are eluted with 6 mL acetone/MeOH (9:1). Gangliosides are eluted with 6 mL CHCl3:MeOH (1:3) and the eluates dried under nitrogen. The samples are resuspended with 1 × 50 μL and then 1 × 100 μL of CHCl3:MeOH (1:2.2) and transferred to a 1.5‐mL tube. For mass spectrometry, the samples are dried down in a SpeedVac and resuspended in 20 μL MeOH.

Matrix‐assisted laser desorption/ionization (MALDI) mass spectrometry can be performed as described previously (Hunnam et al., 2001). Briefly, a mixture of the sample (1 μL, 100 pmol) and matrix (1 μL of a saturated solution of 2,5‐dihydroxybenzoic acid in acetonitrile) are crystallized on the MALDI target. Positive ion reflectron MALDI spectra can then be obtained using a time‐of‐flight (TOF) mass spectrometer calibrated externally with hydrolyzed dextran sugars.

To dried lipid extracts, 10 μL incubation buffer (1 mg/mL sodium cholate in 50 mmol sodium acetate pH 5.0) was added. After vigorous vortexing and spinning in a benchtop picofuge, 10 μL of 50 mU/10 μL of ceramide glycanase (E.C. 3.2.1.123, Calbiochem, La Jolla, CA) in incubation buffer is added to cleave the glycans. Glucosylceramide is only partially digested due to the specificity of the glycanase (Wing et al., 2001). After this period, 10 μL of water is added to each enzyme digest followed by 80 μL of anthranilic acid and sodium cyanoborohydride to each digest and incubated in the 80 °C oven for 1 h. Derivatized oligosaccharides can be purified on DPA‐6S (Supelco, Ballafonte, PA) columns preequilibrated with 2 × 1 mL CH3CN. One milliliter 97:3 CH3CN:H2O is added to each sample and vortexed prior to loading the samples onto the columns. Columns are washed with 4 × 1 mL 99:1 CH3CN:H2O and then with 0.5‐mL 97:3 CH3CN:H2O and the derivatized oligosaccharides eluted with 2 × 0.6 mL water into screw‐cap Eppendorfs and stored at 4 °C in the dark until ready for NP‐HPLC.

Because of the specificity of the glycanase, GlcCer cannot be quantitated using the procedure applied for higher GSL assay. Fortunately GlcCer can be quantitated using glucocerebrosidase instead of the ceramide glycanase. Due to the preponderance of free glucose in most samples, GlcCer is first purified on silicic acid columns. After GSL extraction and sample loading (see earlier), columns are washed with 2 mL of CHCl3 and 2 mL of CHCl3:MeOH 98:2. GlcCer is then eluted with 2 mL of CHCl3:MeOH 97:3, 2 mL of CHCl3:MeOH 96:4, 2 mL of CHCl3:MeOH 95:5, and 2 mL of CHCl3:MeOH 94:6. Samples are dried down and resuspended in 15 μL of 50 mmol sodium acetate buffer (pH 5) with 0.1% triton and 0.25% taurocholate; 0.4 U of purified glucocerebrosidase (Cerezyme or Ceredase) is added and incubated at 37 °C for 18 h. Samples of glucocerebrosidase can have significant oligosaccharide contamination; therefore, the enzyme has to be purified using a spin column (MW cut off 10 kDa). Samples are then derivatized using the same method as for the higher GSLs.

Section snippets

Subcellular Distribution of Glycosphingolipids

Consistent with the topology of their synthesis, many GSLs have been found at the cell surface where they constitute a small proportion of the total plasma membrane lipid. One exception is the apical plasma membrane of apical cells, where GSLs are found in relatively high concentrations (Brasitus 1980, Forstner 1968).

At the cellular level, the simplest and evolutionarily most ancient GSL, glucosylceramide GlcCer, has been found to be enriched within intracellular membranes, in contrast to more

Glycosphingolipid Storage

Sphingolipids are degraded in late endosomes and lysosomes by the action of acid hydrolases and their cofactors. If a defect is inherited in a gene encoding a protein essential for lysosomal catabolism, storage of the substrate ensues. These diseases are termed the GSL storage diseases. To date, approximately 40 genetically distinct forms of GSL storage diseases have been described. They are inherited either as autosomal recessive or X‐linked traits and have a collective frequency of 1:8000

Mucolipidosis Type IV

In contrast to many lysosomal storage diseases, some do not show deficiencies in any enzyme activity or cofactor (Table I). In many of these diseases, the underlying reason for GSL storage is unclear. One example of secondary storage is α‐mannosidosis, where the lysosomal accumulation of glycoconjugates is thought to inhibit the activity of acid hydrolases and induce the accumulation of GSL. In other diseases with secondary GSL accumulation (e.g., Niemann‐Pick type C [NPC] and mucolipidosis

Effect of MDR1a Deficiency on Glycosphingolipids in Both Normal and Niemann‐Pick Type C Mice

It has been speculated that there may be a close relationship between multiple drug resistance (MDR) and NPC1 as lipid translocators/cation permeases (Ioannou, 2001). Evidence that supports such a relationship includes the observation that NPC1 can mediate drug efflux (Gong et al., 2006) and that a deficiency in mdr1a has been shown to be able to correct NPC1‐induced sterility in NPC1−/− female mice (Erickson et al., 2002). One further relationship between MDR and NPC1 is that they have both

Summary

In this review, the significance of some of the subtleties of GSL transport and synthesis have been discussed. Multiple pathways for synthesis via de novo and recycling routes may lead to changes in the levels of signaling sphingosine‐1‐phosphate. This may in turn affect both anterograde (Giussani et al, 2006) and retrograde endocytic trafficking. GSLs can be internalized via a specific retrograde endocytic pathway utilized by toxins to gain access to ER subdomains. Moreover, this endocytic

References (176)

  • M.F. De Rosa et al.

    Role of multiple drug resistance protein 1 in neutral but not acidic glycosphingolipid biosynthesis

    J. Biol. Chem.

    (2004)
  • C. Dietrich et al.

    Relationship of lipid rafts to transient confinement zones detected by single particle tracking

    Biophys. J.

    (2002)
  • R.G. Farrer et al.

    Acylation of exogenous glycosylsphingosines by intact neuroblastoma (NCB‐20) cells

    J. Biol. Chem.

    (1990)
  • K. Funato et al.

    Lcb4p is a key regulator of ceramide synthesis from exogenous long chain sphingoid base in Saccharomyces cerevisiae

    J. Biol. Chem.

    (2003)
  • T.K. Ghosh et al.

    Sphingosine 1–phosphate generated in the endoplasmic reticulum membrane activates release of stored calcium

    J. Biol. Chem.

    (1994)
  • E. Goldin et al.

    Type C Niemann‐Pick disease: A murine model of the lysosomal cholesterol lipidosis accumulates sphingosine and sphinganine in liver

    Biochim. Biophys. Acta

    (1992)
  • H. Heerklotz

    Triton promotes domain formation in lipid raft mixtures

    Biophys. J.

    (2002)
  • M.E. Higgins et al.

    Niemann‐Pick C1 is a late endosome‐resident protein that transiently associates with lysosomes and the trans‐Golgi network

    Mol. Genet. Metab.

    (1999)
  • J.W. Hinrichs et al.

    Drug resistance‐associated changes in sphingolipids and ABC transporters occur in different regions of membrane domains

    J. Lipid Res.

    (2005)
  • D. Hoekstra et al.

    Trafficking of glycosphingolipids in eukaryotic cells; sorting and recycling of lipids

    Biochim. Biophys. Acta

    (1992)
  • M. Holtta‐Vuori et al.

    Mobilization of late‐endosomal cholesterol is inhibited by Rab guanine nucleotide dissociation inhibitor

    Curr. Biol.

    (2000)
  • V. Hunnam et al.

    Ionization and fragmentation of neutral and acidic glycosphingolipids with a Q‐TOF mass spectrometer fitted with a MALDI ion source

    J. Am. Soc. Mass Spectrom.

    (2001)
  • K. Itagaki et al.

    Sphingosine 1‐phosphate, a diffusible calcium influx factor mediating store‐operated calcium entry

    J. Biol. Chem.

    (2003)
  • J.J. Jennings et al.

    Mitochondrial aberrations in mucolipidosis type IV

    J. Biol. Chem.

    (2006)
  • K.A. Karlsson et al.

    The sphingolipid composition of bovine kidney cortex, medulla and papilla

    Biochim. Biophys. Acta

    (1973)
  • T.W. Keenan et al.

    Distribution of gangliosides among subcellular fractions from rat liver and bovine mammary gland

    FEBS Lett.

    (1972)
  • E. Korkotian et al.

    Elevation of intracellular glucosylceramide levels results in an increase in endoplasmic reticulum density and in functional calcium stores in cultured neurons

    J. Biol. Chem.

    (1999)
  • P. Lala et al.

    Retroviral transfection of Madin‐Darby canine kidney cells with human MDR1 results in a major increase in globotriaosylceramide and 10(5)‐ to 10(6)‐fold increased cell sensitivity to verocytotoxin. Role of p‐ glycoprotein in glycolipid synthesis

    J. Biol. Chem.

    (2000)
  • H. Lannert et al.

    Lactosylceramide is synthesized in the lumen of the Golgi apparatus

    FEBS Lett.

    (1994)
  • J.M. LaPlante et al.

    Identification and characterization of the single channel function of human mucolipin‐1 implicated in mucolipidosis type IV, a disorder affecting the lysosomal pathway

    FEBS Lett.

    (2002)
  • J.M. LaPlante et al.

    Functional links between mucolipin‐1 and Ca2+‐dependent membrane trafficking in mucolipidosis IV

    Biochem. Biophys. Res. Commun.

    (2004)
  • J.M. LaPlante et al.

    Lysosomal exocytosis is impaired in mucolipidosis type IV

    Mol. Genet. Metab.

    (2006)
  • Y. Lavie et al.

    Accumulation of glucosylceramides in multidrug‐resistant cancer cells

    J. Biol. Chem.

    (1996)
  • E.H. Lee et al.

    Dimethylsphingosine regulates intracellular pH and Ca(2+) in human monocytes

    J. Pharmacol. Sci.

    (2006)
  • W.I. Lencer et al.

    Membrane traffic and the cellular uptake of cholera toxin

    Biochim. Biophys. Acta

    (1999)
  • R. Masgrau et al.

    NAADP: A new second messenger for glucose‐induced Ca2+ responses in clonal pancreatic beta cells

    Curr. Biol.

    (2003)
  • G.R. Matyas et al.

    Subcellular distribution and biosynthesis of rat liver gangliosides

    Biochim. Biophys. Acta

    (1987)
  • H. Miller Podraza et al.

    Isolation of complex gangliosides from human brain

    Biochim. Biophys. Acta

    (1992)
  • J.D. Allen et al.

    Extensive contribution of the multidrug transporters P‐Glycoprotein and Mrp1 to basal drug resistance

    Cancer Res.

    (2000)
  • S. Arab et al.

    Intracellular targeting of the endoplasmic reticulum/nuclear envelope by retrograde transport may determine cell hypersensitivity to verotoxin via globotriaosyl ceramide fatty acid isoform traffic

    J. Cell Physiol.

    (1998)
  • D. Ardail et al.

    The mitochondria‐associated endoplasmic‐reticulum subcompartment (MAM fraction) of rat liver contains highly active sphingolipid‐specific glycosyltransferases

    Biochem. J.

    (2003)
  • K. Badizadegan et al.

    Heterogeneity of detergent‐insoluble membranes from human intestine containing caveolin‐1 and ganglioside G(M1)

    Am. J. Physiol. Gastrointest. Liver Physiol.

    (2000)
  • E. Beutler et al.

    Gaucher disease

  • E.G. Bligh et al.

    A rapid method of total lipid extraction and purification

    Can. J. Biochem. Physiol.

    (1959)
  • J. Bodennec et al.

    Phosphatidylcholine synthesis is elevated in neuronal models of Gaucher disease due to direct activation of CTP: Phosphocholine cytidylyltransferase by glucosylceramide

    FASEB J.

    (2002)
  • R.G. Boot et al.

    Identification of the non‐lysosomal glucosylceramidase as beta‐glucosidase 2

    J. Biol. Chem.

    (2006)
  • T.A. Brasitus et al.

    Lipid dynamics and lipid‐protein interactions in rat enterocyte basolateral and microvillus membranes

    Biochemistry

    (1980)
  • K.N.J. Burger et al.

    Topology of sphingolipid galactosyltransferases in ER and Golgi: Transbilayer movement of monohexosyl sphingolipids is required for higher glycosphingolipid biosynthesis

    J. Cell Biol.

    (1996)
  • X. Buton et al.

    Transbilayer movement of monohexosylsphingolipids in endoplasmic reticulum and Golgi membranes

    Biochemistry

    (2002)
  • S. Chatterjee et al.

    Localization of urinary lactosylceramide in cytoplasmic vesicles of renal tubular cells in homozygous familial hypercholesterolemia

    Proc. Natl. Acad. Sci. USA

    (1983)
  • Cited by (0)

    View full text