Skip to main content
Glycobiology logoLink to Glycobiology
. 2014 Mar 18;24(5):407–417. doi: 10.1093/glycob/cwu015

The role of protein N-glycosylation in neural transmission

Hilary Scott 2, Vladislav M Panin 1
PMCID: PMC3976283  PMID: 24643084

Abstract

Recent studies have explored the function of N-linked glycosylation in the nervous system, demonstrating essential roles of carbohydrate structures in neural development. The function of N-glycans in neural physiology remains less understood; however, increasing evidence indicates that N-glycans can play specific modulatory roles controlling neural transmission and excitability of neural circuits. These roles are mediated via effects on synaptic proteins involved in neurotransmitter release, transporters that regulate nerotransmitter concentrations, neurotransmitter receptors, as well as via regulation of proteins that control excitability and response to milieu stimuli, such as voltage-gated ion channels and transient receptor potential channels, respectively. Sialylated N-glycan structures are among the most potent modulators of cell excitability, exerting prominent effects on voltage gated Na+ and K+ channels. This modulation appears to be underlain by complex molecular mechanisms involving electrostatic effects, as well as interaction modes based on more specific steric effects and interactions with lectins and other molecules. Data also indicate that particular features of N-glycans, such as their location on a protein and structural characteristics, can be specifically associated with the effect of glycosylation. These features and their functional implications can vary between different cell types, which highlight the importance of in vivo analyses of glycan functions. Experimental challenges are associated with the overwhelming complexity of the nervous system and glycosylation pathways in vertebrates, and thus model organisms like Drosophila should help elucidate evolutionarily conserved mechanisms underlying glycan functions. Recent studies supported this notion and shed light on functions of several glycosylation genes involved in the regulation of the nervous system.

Keywords: Drosophila, glycosylation, ion channels, neural transmission, sialylation

Introduction

Protein N-glycosylation, one of the most abundant and well-studied types of protein glycosylation, has been found in all three domains of life: Eukarya, Bacteria and Archaea (Abu-Qarn et al. 2008; Stanley et al. 2009; Moremen et al. 2012). The majority of all secretory pathway glycoproteins in the human organism appear to be N-glycosylated (Apweiler et al. 1999). The genetic or pharmacological inhibition of N-glycosylation does not preclude cell viability in mammalian tissue cultures (Gottlieb et al. 1975; Stanley et al. 1975). At the same time, the abrogation of biosynthesis of hybrid and complex N-glycans causes embryonic lethality in mice (Ioffe and Stanley 1994; Metzler et al. 1994), which indicates the essential requirement of N-glycosylation for cell communications within mammalian organisms. General functions of N-linked glycans inside the secretory pathway have been extensively studied, including prominent roles in protein folding, quality control and trafficking (Helenius and Aebi 2001). At the same time, N-linked glycosylation also plays a variety of important roles outside of the cell. These roles are significantly less understood, because they are commonly pleiotropic, complex and not directly amenable to ex vivo analyses. Moreover, N-glycans of a protein can show significant heterogeneity at the cellular level, and their biosynthesis is intimately linked to metabolism, representing a dynamic physiological read-out of a metabolic state of the cell (Dennis et al. 2009). Many of the extracellular functions of N-glycans are underlain by mechanisms of interactions with lectins, proteins that bind specific glycan structures (Varki et al. 2009). Lectin interactions play prominent roles in cell adhesion and signaling, and they can sculpt the molecular landscape of cell surfaces (Sharon 2007; Dennis et al. 2009). Molecular functions of N-glycans can also include the stabilization of functional protein conformations and steric interactions protecting from proteolysis (Wittwer and Howard 1990, reviewed in Wormald and Dwek 1999). Several novel paradigms have started to emerge for functions of N-glycans in the nervous system. With the goal of focusing on some interesting findings that shed light on these paradigms, while not intending to provide a comprehensive account of data accumulated in the field, we will discuss the neural functions of N-glycosylation in the current review.

N-Glycans in the nervous system development

Studies of human congenital disorders of glycosylation unveiled that inborn defects of N-glycosylation almost inevitably result in prominent neurological abnormalities (reviewed in Freeze et al. 2012), which highlights the crucial role of this posttranslational protein modification in the nervous system. The genetic inactivation of glycosyltransferase genes in mice has been an invaluable tool for elucidating in vivo functions of glycan structures (Lowe and Marth 2003). Thus, brain-targeted inactivation of β-1,6-N-acetylglucosaminyltransferase (GlcNAcT)-I, the enzyme required for the biosynthesis of hybrid and complex N-linked carbohydrates, was found to result in severe neurological phenotypes, including locomotor abnormalities, tremors and paralysis (Ye and Marth 2004). While the complexity of mammalian glycosylation and its pleiotropic effects on development and physiology provide serious obstacles for studying mutants with defects in core structures, the knockouts affecting more specialized and certain terminal structures proved to be more amenable for analyses. They demonstrated the involvement of some N-linked modifications in specific regulatory events. For instance, genetic deletions of ST8Sia II and ST8Sia IV polysialyltransferases modifying N-glycans of neural cell adhesion molecule (NCAM) with polysialic acid (PSA) unveiled the remarkable role of PSA in the nervous system (Weinhold et al. 2005; Angata et al. 2007; Hildebrandt et al. 2009). PSA, a unique long homopolymer of α2,8-linked sialic acid (Sia) residues, is present on N-glycans of NCAM and regulates brain development, neurite outgrowth and targeting, while also affecting synaptic plasticity, learning and memory (the structure and function of NCAM-PSA has been discussed extensively in several reviews, e.g. Muhlenhoff et al. 1998; Rutishauser 2008; Muhlenhoff et al. 2009; Colley 2010). Interestingly, the severe malformation of major axonal tracts in the brain, progressive hydrocephalus and early postnatal lethality associated with PSA deficiency are caused by the gain-of-function effect of NCAM lacking PSA, and the genetic removal of NCAM rescues these gross defects in PSA-deficient mice (Weinhold et al. 2005). Another striking example of a specialized N-glycan structure affecting neural development is represented by poly-N-acetyllactosamine (poly-LacNAc) oligosaccharides synthesized by β3GnT2. Genetic deletion of the β3GnT2 glycosyltransferase in mice leads to multiple abnormalities of the olfactory system, including prominent axonal guidance defects. This phenotype appears to result primarily from hypoglycosylation of adenylyl cyclase 3, a multi-pass membrane glycoprotein highly expressed in olfactory sensory neurons, which leads to the loss of cyclase's enzymatic activity that generates cAMP, a key signaling molecule in olfactory axon targeting (Henion et al. 2011). These examples probably represent just a tip of the iceberg of many other crucial roles of N-glycans in the nervous system that are waiting to be discovered. The panoply of glycoproteins, including cell surface and extracellular matrix components involved in cell signaling and adhesion, is known to regulate neural development and physiology (Kleene and Schachner 2004; Dityatev et al. 2010). In some cases, functions of these glycoproteins were found to be affected by N-glycosylation in biological contexts outside of the nervous system. For example, laminin 332 and α3β1 integrins were shown to be regulated by bisecting and β1,6-branching N-acetylglucosamine (GlcNAc) modifications that modulated cell adhesion and contributed to increased cell motility (Zhao et al. 2006; Kariya et al. 2008). Another striking example of a carbohydrate modification that can affect different molecular interactions is α2,6-sialylation that was found to control the receptor protein tyrosine phosphatase CD45 and β1 integrins via modifying their conformation and interactions with functionally important partners (Amano et al. 2003; Woodard-Grice et al. 2008). Specific role of α2,6-linked Sia's in negative regulation of galectin binding has emerged as a paradigm that is likely pertinent to many cellular and molecular contexts (reviewed in Zhuo and Bellis 2011). These and other examples suggest that similar, yet unknown, glycan-dependent regulatory processes may operate in the nervous system. Recent studies have also shed light on several novel mechanisms implicating N-glycosylation in the modulation of neural transmission, which we will be discussed in more detail below.

Glycans in synaptic transmission

Recent genetic analysis of congenital myasthenic syndromes, a group of hereditary abnormalities of synaptic transmission at neuromuscular junctions (NMJs; e.g., OMIM 608931, Engel 2012), unveiled a novel genetic lesion associated with mutations in GFPT1, the glutamine-fructose-6-phosphate transaminase 1. GFPT1 mediates the first, rate-limiting step of the hexosamine pathway required for the biosynthesis of uridine diphospho-N-acetylglucosamine (UDP-GlcNAc), a key substrate for glycosylation pathways (Senderek et al. 2011); hence, these data highlight that glycosylation in general is an essential prerequisite for synaptic functions. The pertinence of synaptic glycosylation has been discussed in many studies (see reviews Kleene and Schachner 2004; Dityatev et al. 2010; Dani and Broadie 2012). However, the molecular and cellular mechanisms underlying the effects of glycans on synaptic transmission is complex and still poorly understood. The role of protein N-glycosylation at the synapse is multifaceted since it controls the function of a number of key players of synaptic processes. For instance, N-linked glycosylation was found to be crucial for the function of the synaptic vesicle protein 2 (SV2) that is ubiquitously present at vertebrate synapses. The importance of SV2 has been revealed by gene-targeting experiments that demonstrated that the deletion of two out of three existing SV2 isoforms resulted in postnatal lethality in mice due to severe seizures. No brain developmental abnormalities were found in these SV2 mutants, which indicated that this protein has a dedicated role in synaptic physiology (Janz et al. 1999). Experiments suggested that SV2 determines a novel maturation step of primed synaptic vesicles and enhances their responsiveness to Ca2+. Moreover, all SV2 isoforms are N-glycosylated at multiple sites within their intravesicular loop. The removal of all three glycosylation sites present on the most ubiquitous SV2a isoform abolishes its synaptic targeting along with its function (Chang and Sudhof 2009), suggesting that N-glycans are required for proper folding and trafficking of SV2 within the neuron. A more recent study of single N-glycosylation site mutants indicated that individual glycans are partially dispensable and their function is redundant for SV2a sorting to synaptic vesicles (Kwon and Chapman 2012). Similar analyses of two other major synaptic vesicle glycoproteins revealed that the role of glycosylation in their sorting to synaptic vesicles ranges from being dispensable (synaptotagmin 1) to essential (synaptophysin) (Kwon and Chapman 2012). This highlights a common theme in glycobiology: glycans can play highly individualized roles tailored for a particular glycoprotein and its specific function.

Ligand-gated channel proteins can function as synaptic neurotransmitter receptors that mediate communications among neurons or between neurons and muscles. These receptors usually carry several N-linked carbohydrate chains. Substantial evidence uncovering the biological importance of these carbohydrate modifications has started to accumulate. Thus, the effect of N-glycans on the function of nicotinic acetylcholine receptors (nAChRs) has been revealed by several studies. These receptors represent founding members of the pentameric ligand-gated super family of ion channels that also includes serotonin, γ-aminobutyric acid (GABA) and glycine receptors (Chen 2010). nAChRs regulate postsynaptic responses of NMJs, while also affecting diverse brain functions, such as learning, memory and processing of sensory information (Miwa et al. 2011). A number of studies suggested that glycosylation plays a role in channel gating and desensitization (Gehle et al. 1997; Chen et al. 1998). N-glycans of Torpedo nAChRs receptors have been implicated in receptor modulation, as receptors with mutated glycosylation sites have abnormal desensitization (the rate of current decay) and conductances (Nishizaki 2003). Interestingly, concanavalin A (ConA) can mimic this effect of mutations in vitro when pharmacologically included in the assays with wild-type receptors. ConA is a lectin that binds N-linked glycans, while showing a preference for oligomannosidic structures, and thus its effect on wild-type nAChRs suggested that these glycans may function as a modulating “lid” at the channel pore (Nishizaki 2003). More recent research on structure–function relationship of nAChR glycosylation indicated that carbohydrate modifications can be important for the surface expression and cholinergic agonist-dependent gating of nAChRs, but not for the binding affinity for the agonists or the stability of folded receptors (da Costa et al. 2005; Dellisanti et al. 2007).

The modulatory effect of ConA was also demonstrated for ionotropic glutamate receptors (iGluRs). iGluRs mediate fast transmission at the majority of excitatory synaptic connections within the mammalian nervous system, while also playing important roles in the regulation of synaptic plasticity and extrasynaptic modulation of neurons (reviewed in Traynelis et al. 2010). These receptors represent tetrameric complexes that function as ligand-gated ion channels and encompass large subfamilies of α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA), kainate and N-methyl-d-aspartate (NMDA) receptors (Traynelis et al. 2010). ConA exerts a pronounced effect on kainate receptors by inhibiting their desensitization (Partin et al. 1993; Everts, Villmann et al. 1997; Everts, Petroski et al. 1999). Most iGluRs appear to be modified by N-glycosylation, while having consensus glycosylation sites within the amino-terminal domain (ATD) involved in the receptor assembly and modulation, as well as in the ligand-binding domain (LBD) (Partin et al. 1993; Everts, Villmann et al. 1997, Everts, Petroski et al. 1999; Mah et al. 2005). The in vivo utilization of these sites has not been well characterized, but it was suggested that ConA interacts with ATD-attached N-glycans and affects conformational changes of the receptors (Everts, Villmann et al. 1997; Everts, Petroski et al. 1999; Fay and Bowie 2006). Other lectins, such as wheat germ agglutinin, soybean agglutinin and succinyl-ConA, can also potentiate kainate receptors (Thio et al. 1993; Yue et al. 1995). The effect of heterologous lectins in vitro suggests that glycans may play a specialized role in iGluR modulation (Everts et al. 1997; Nanao et al. 2005) mediated by interactions with some endogenous, yet unidentified lectins. However, currently, there is no direct evidence to support this scenario. Nevertheless, the appealing hypothesis that lectins recognizing N-glycans can be involved in the regulation of neurophysiology is consistent with several studies that shed light on lectin functions in the nervous system. Thus, experiments demonstrated the involvement of Galectin 1 in the degeneration of neuronal processes and the inhibition of neurotoxicity by affecting the expression of the NR1 NMDA receptor (Lekishvili et al. 2006; Plachta et al. 2007). Although the exact mechanisms of these functions remain to be elucidated, they appear to depend on galectin carbohydrate-binding activity and require galectin dimerization. More recently, Galectin 1 was implicated in the regulation of circadian rhythms in mice, which also points to its role in neural physiology (Casiraghi et al. 2010). Narp, a member of pentraxin family of putative endogenous lectins with structural similarity to wheat germ agglutinin, localizes to excitatory synapses and specifically clusters AMPA-type glutamate receptors (O'Brien et al. 1999). Finally, studies in Drosophila indicated that Mind-the-Gap (MTG), a putative fly lectin, is required for synaptic localization of iGluRs (Rushton et al. 2012).

While glycans can influence maximal currents and desensitization of AMPA and kainite receptors (Hollmann et al. 1994; Everts et al. 1997), N-glycosylation appears to be not generally required for the function of these iGluRs, and glycans do not significantly affect the synthesis, transport or subunit assembly of AMPA or kainate receptors (Sumikawa et al. 1988; Everts et al. 1997; Gill et al. 2009). Crystallization experiments indicated that glycans are not directly involved in ligand binding and subunit association of kainate receptors (Armstrong et al. 1998; Nanao et al. 2005). Unlike AMPA and kainate iGluRs, the NMDA-type receptors can be dramatically downregulated by inhibition of N-glycosylation. This effect is primarily due to the specific reduction in NR1 subunits (Everts et al. 1997). Although the exact mechanism of this defect remains largely unknown, data suggested a possible requirement for glycans in the folding or trafficking of these subunits.

Interestingly, a specialized structure that can be present on N-glycans, the HNK-1 glycoepitope, was shown to regulate the AMPA-type receptor subunit GluR2. HNK-1 is represented by a sulfated glucuronic acid linked to GlcNAc on the nonreducing termini of oligosaccharides (HSO3–3GlcAβ1–3Galβ1–4GlcNAc) and its expression is enriched in the nervous system. Mouse knockouts of enzymes involved in the biosynthesis of this structure (glucuronyltransferase GlcAT-P, sulfotransferase HNK-1ST and β4-galactosyltransferase-2) showed impairment of neural plasticity, learning and memory, suggesting the role of HNK-1 in synaptic functions (Senn et al. 2002; Yamamoto et al. 2002; Yoshihara et al. 2009). Experiments with GluR2 expressed in cultured hippocampal neurons and heterologous cells demonstrated that HNK-1 epitope downregulates receptor endocytosis and promotes GluR2 stability on neuronal membranes, probably via enhancing GluR2 interactions with N-cadherin (Morita et al. 2009).

Neurotransmitter transporters of the solute carrier 6 (SLC6) family represent another example of synaptic glycoproteins that play essential roles in synaptic transmission (Kristensen et al. 2011). They include ion-coupled plasma membrane cotransporters that control the synaptic cleft concentration of key neurotransmitters such as dopamine, GABA, glycine, norepinephrine and serotonin. These transporters share a similar structure including 12 transmembrane domains and a large N-glycosylated extracellular loop between transmembrane domains III and IV. The glycosylation of this loop is important for transporter functions, since removal of glycan chains by mutations, pharmacological inhibition of glycosylation or glycosidase treatment can dramatically reduce transporter activity at the cell surface (Melikian et al. 1996; Martinez-Maza et al. 2001; Li et al. 2004; Hahn et al. 2005). These and other experiments demonstrated that the most prominent effect of glycans on these transporters is due to the requirement of proper N-glycosylation for trafficking and cell surface retention, while glycans usually do not significantly influence ligand binding and catalytic activity of the transporters (Tate and Blakely 1994; Olivares et al. 1995; Nguyen and Amara 1996). Importantly, the structure of glycan chains can also affect the function of these transporters. Complex sialylated glycans were found to play a role in the functional oligomerization of serotonin transporter (SERT). Sia modifications can potentiate hetero-oligomeric interactions of SERT with the cytoskeletal protein myosin IIA and stimulate serotonin uptake, presumably due to their effect on transporter homo-oligomerization and the accessibility of domains involved in hetero-oligomeric interactions (Ozaslan et al. 2003). Sia was also found to influence the function of GAT1, the major GABA transporter in the brain. Deficient sialylation of GAT1 was found to slow down the kinetics of GABA transport cycle and decrease the apparent affinity for extracellular Na+ (Cai et al. 2005; Hu et al. 2011). Notably, several single-nucleotide polymorphisms (SNPs) affecting N-glycosylation and potentially associated with a disease condition have been identified in transporter genes in human populations. Thus, several SNPs that significantly decrease the glycosylation and activity of human norepinephrine transporter were found to be potentially associated with defects of synaptic transmission at postganglionic sympathetic nerve terminals implicated in cardiovascular abnormalities (Hahn et al. 2005). A SNP in the SLC6A4 gene encoding serotonin transporter was found to create an ectopic glycosylation site (K201N) that increases the glycosylation of hSERT along with its expression and activity. By analogy to another polymorphism that similarly affects the expression of serotonin transporter and was phenotypically characterized, it was suggested that the K201N mutation can potentially influence personality traits and psychiatric disease susceptibility (Rasmussen et al. 2009).

It is important to keep in mind that many experiments with neurotransmitter receptors and transporters have been performed in various types of heterologous cells (such as insect cells (Tate and Blakely 1994), frog oocytes (Everts et al. 1997; Gehle et al. 1997; Nishizaki 2003), mammalian tissue cultures (Dellisanti et al. 2007; Gurba et al. 2012)) that could execute glycosylation and glycan-dependent regulation differently from each other and from neural cells in vivo. Moreover, glycosylation of multisubunit receptor complexes could also depend on a particular combination of expressed receptor subunits (e.g., as in the case of GABAA receptors that can have differently glycosylated β2 subunits, depending on co-expression of other subunits (Gurba et al. 2012)). Therefore, a caution should be exercised when interpreting data from cell culture experiments and relating them to in vivo mechanisms. Nevertheless, taken together, experimental data clearly indicate that glycosylation can significantly influence the function of neurotransmitter receptors and transporters. This influence can be dissimilar for distinct neurotransmitter systems, and even within one family of closely related glycoproteins, such as iGluRs, glycosylation can engage various modulatory mechanisms resulting in different functional outcomes and depending on particular structures of carbohydrate chains. These specific effects of N-glycans add an additional dimension to the regulatory space of processes that control synaptic transmission and plasticity.

N-glycosylation regulates ion channels in vertebrate neurons

N-glycosylation was found to be an important modulator of ion channels in neural cells. Usually, glycans modulate channels via two mechanisms: by affecting their cell surface expression and by influencing biophysical properties of channel proteins. The first mechanism depends on the effect on protein folding, stability and subcellular trafficking, in particular exocytosis and internalization. The effect on cell surface expression has been demonstrated for several types of neuronal channels, including acid-sensing channels (e.g., acid-sensing ion channel (ASIC)1a and 1b (Kadurin et al. 2008; Jing et al. 2012)) and voltage-gated ion channels (e.g., potassium channels Kv1.3, Kv1.4 and HERG (Gong et al. 2002; Watanabe et al. 2004; Zhu et al. 2012) and calcium channels Cav3.2 (Weiss et al. 2013)). Many channel proteins have glycans attached to their pore loops, a protein region frequently influencing channel gating, which promotes the effect of glycans on channel's biophysical characteristics. Thus, the function of TRPM8, a member of a large family of transient receptor potential (TRP) ion channels playing essential roles in sensory physiology, is modulated by pore loop glycans that cause a shift in the voltage dependence of channel activation (Pertusa et al. 2012). TRPM8 is expressed in sensory neurons that respond to cold (McKemy et al. 2002; Peier et al. 2002). Since its glycosylation affects the temperature threshold of channel activation, it was proposed that TRPM8 glycans represent critical molecular determinants that establish cold sensitivity in primary sensory neurons (Pertusa et al. 2012). Interestingly, the membrane localization of TRPM8 appears to be largely unaffected by glycosylation (Pertusa et al. 2012). While glycans change the biophysical properties of several other TRP channels (TRPC3, TRPC6 and TRPV1 (Dietrich et al. 2003; Wirkner et al. 2005)), glycosylation can also control TRP channels by influencing their expression and subcellular localization (TRPV4 and TRPV5 (Chang, et al. 2005; Xu et al. 2006)). These different mechanisms appear to be not mutually exclusive and could operate at the same time, while one of them could dominate, depending on particular molecular and cellular contexts.

A fascinating glycan-dependent regulation has recently been described for the renal epithelial Ca2+ channel TRPV5. This channel can be efficiently internalized, and its retention on the cell surface is an essential process that regulates channel function. Although this regulation is not yet fully understood, it is most likely mediated by converging effects of several mechanisms. One of them was found to rely on N-glycan-dependent stabilization of the channel by the hormone Klotho. Klotho is a glycosidase that appears to modify channel glycans, which in turn facilitates channel interactions with galectin and potentiates cell surface retention (Cha et al. 2008; Chang et al. 2005; Leunissen et al. 2013). Yet another regulation engages sialylation and appears to work independently from N-glycosylation by promoting lipid-raft–mediated internalization of the channel (Leunissen et al. 2013). It would be interesting to investigate if similar mechanisms operate in the nervous system to control channels involved in neural transmission.

In addition to direct effects mediated by channel glycans, N-glycosylation can exert its effect indirectly, in molecular nonautonomous manner, via regulating glycoproteins that control channel functions, e.g. auxiliary subunits promoting membrane localization of a channel (Cotella et al. 2010). This extra layer of regulation, however, remains largely unexplored.

Sialylated N-glycans regulate voltage-gated ion channels and membrane excitability

Glycans can directly influence cell excitability in vertebrates by affecting the function of different members of a large superfamily of voltage-gated ion channels, including Na+, K+ and Ca2+ channel (e.g., Recio-Pinto et al. 1990; Zhang et al. 1999; Bennett 2002; Gong et al. 2002; Watanabe, Wang et al. 2003; Watanabe, Zhu et al. 2007; Johnson et al. 2004; Schwetz et al. 2010; Weiss et al. 2013). The glycosylation of mammalian voltage-gated Na+ and K+ channels has been shown to be developmentally and spatially regulated in the heart and the nervous system, suggesting that glycans are involved in the control of cellular excitability to fulfill different demands of particular cells and developmental stages (Castillo et al. 1997; Tyrrell et al. 2001; Schwalbe et al. 2008; Montpetit et al. 2009). Many studies have been specifically focused on sialylated glycans because of their negative charge and potential electrostatic effect on channel gating (reviewed in Ednie and Bennett 2012). Vertebrate voltage-gated Na+ channels are especially heavily glycosylated and sialylated: glycans can comprise up to 30% of their molecular mass, with Sia representing almost 50% of these glycan chains (Miller et al. 1983; Elmer et al. 1985; Messner and Catterall 1985; James and Agnew 1987; Roberts and Barchi 1987).

Electrophysiological analyses of cultured cells and in vitro experiments revealed that sialylation can significantly influence the conductance properties of Na+ channels (Recio-Pinto et al. 1990; Bennett et al. 1997; Zhang et al. 1999; Cronin et al. 2005). The effect of Sia is isoform- and subunit-specific, and it significantly varies for different channel proteins (Bennett 2002; Johnson et al. 2004; Johnson and Bennett 2006). This effect has been usually attributed to a large negative charge due to numerous Sia residues attached to vertebrate channel glycans. Interestingly, vertebrate voltage-gated Na+ channels represent one of only few types of glycoproteins shown to be modified with PSA (James and Agnew 1987; Zuber et al. 1992). It is estimated that these channels can carry more than 100 Sia residues, most of them being in a form of PSA (Miller et al. 1983; James and Agnew 1987; Zuber et al. 1992). The role of PSA in channel modulation was confirmed by experiments with ST8Sia II polysialyltransferase mutant mouse cardiomyocites that showed defects in excitability and gating of voltage-gated Na+ channels (Montpetit et al. 2009). However, several lines of evidence indicate that the effect of sialylation is not always associated with PSA and a significant charge accumulated in the vicinity of channel pore, suggesting that Sia can have a more specific effect on channel functions. For instance, endo-N treatment that specifically removes PSA did not significantly influence evoked synaptic transmission and action potentials in the rat brain (Muller et al. 1996). Similarly, cell culture experiments indicated that cardiac sodium channel could be affected by some “functional” Sia residues rather than by the total charge of sialylation (Stocker and Bennett 2006). Moreover, experiments in mutant isogenic Chinese hamster ovary (CHO) cell lines with different defects in sialylation found that the effects of PSA and non-PSA sialylation on the function of alpha-NaV1.4 channel were distinct. The complete loss of sialylation in the 6B2 clone due to a mutation in the Golgi CMP-Sia transporter resulted in opposite shifts of voltage-dependent activation and steady-state inactivation of the channel, when compared with the consequence of a PSA defect due to genetic inactivation of the polysialyltransferase ST8SiaIV in the 2A10 clone, while only the loss of Sia had a significant effect on recovery from fast inactivation (Ahrens et al. 2011). Finally, metabolic introduction of unnatural Sia's with N-acetyl groups changed to N-pentanoyl and N-propanoyl structures was found to have an effect on conductance properties of the Kv3.1 voltage-gated K+ channel, suggesting that Sia residues can modulate channels through specific steric interactions (Hall et al. 2011).

While many studies using transgenic approaches in various heterologous cultured cells have demonstrated the involvement of sialylation in the regulation of voltage-gated channels, the structure of glycosylation and its functional implication can significantly vary between different cell types, and between cultured cells and in vivo. At the same time, experiments that address the function of Sia's in vivo remain scarce. Biological importance of sialylation of voltage-gated Na+ channels has been most clearly shown in the context of cardiac functions using mouse genetic models and cardiomyocyte cultures. These experiments indicated that deficient sialylation of channels can be implicated in cardiac abnormalities and heart failure (Ufret-Vincenty, Baro, Lederer, et al. 2001; Ufret-Vincenty, Baro, Santana, et al. 2001; Stocker and Bennett 2006; Montpetit et al. 2009). Treatments with glycosidases that remove sialylated glycans and the inhibition of endogenous neuraminidase, an enzyme that trims Sia's from carbohydrate chains, demonstrated that sialylation influences network excitability and seizure threshold in kindling epilepsy models and suggested that Sia's modulate voltage-gated Na+ channels in brain neurons (Tyrrell et al. 2001; Isaev, Isaeva et al. 2007; Isaeva, Lushnikova et al. 2010; Isaev, Zhao et al. 2011). While these experiments revealed that sialylation can significantly potentiate the excitability of neural networks, whether this effect is due to the sialylation of voltage-gated channels or some other molecules remains unknown. It is challenging to address this question in vertebrates because of the complexity of the nervous system, intricacies of glycosylation pathways, as well as the ubiquity of sialylation affecting panoply of glycoconjugates. With its power of genetic approaches, a spectrum of well-established neurobiological approaches, and simplified glycosylation pathways, Drosophila has recently emerged as promising model for elucidating conserved genetic and molecular mechanisms of neural glycosylation.

N-glycosylation regulates the nervous system of Drosophila

The repertoire of N-glycans in insects is different from that in vertebrate species. Detailed analyses of Drosophila glycome revealed that paucimannose and high mannose structures dominate the spectrum of N-glycosylation, while the complex and hybrid-type oligosaccharides, the glycans corresponding to more processed mature structures prevalent in vertebrates cells, represent only 12% of the total N-glycan profile (Aoki et al. 2007). These minor glycan species, however, appear to play prominent roles in the nervous system, suggesting conservation of their functions in evolution (Schachter 2010). The importance of these glycans has become evident from a number of studies that analyzed neurological phenotypes of mutants affecting N-glycosylation pathway. Thus, genetic inactivation of N-acetylglucosaminyltransferase I, a key enzyme in the production of mature N-glycan structures, was found to result in neurological abnormalities in fruit flies, including locomotion phenotypes, shortened life span and developmental defect of the mushroom bodies, a specialized brain structure involved in memory formation (Sarkar et al. 2006; Sarkar et al. 2010). Mutations of fused lobes, a gene encoding Golgi β-N-acetylglucosaminidase that promotes the biosynthesis of paucimannose structures at the expense of hybrid and complex N-glycans, result in similar mushroom body defects (the mushroom body lobes become fused) and cause cell-lethal phenotype in mosaic clones of olfactory projection neurons (Boquet et al. 2000; Leonard et al. 2006; Sekine et al. 2013). Downregulation of sugar-free frosting, a gene encoding a Drosophila homolog of SAD kinase regulating secretory flux through the Golgi, increases N-glycan complexity and leads to NMJ defects in larvae and locomotor abnormalities in adult flies (Baas et al. 2011). Meigo, a putative nucleotide sugar transporter, affects N-glycosylation of ephrin and regulates dendrite and axonal targeting in the olfactory system (Sekine et al. 2013). Mutations in β1,4-N-acetylgalactosaminyltransferases A (β4GalNAcTA), a glycosyltransferase potentially involved in the biosynthesis of complex and hybrid N-glycans, including sialylated structures, result in prominent neurological phenotypes (Haines and Irvine 2005; Haines and Stewart 2007; Nakamura et al. 2012). All these examples illustrate that there are important requirements for a proper amount and structural diversity of N-glycans in the nervous system. These studies also highlight an intriguing possibility that many genes involved the N-glycosylation pathway may be associated with conserved mechanisms regulating the nervous system.

Sialylated N-glycans control neural excitability in Drosophila

Although sialylated glycans represent a minuscule fraction of Drosophila N-glycome (<0.1% of total N-glycan profile), they are unambiguously detected by multidimensional mass spectrometry in embryos and adult heads (Aoki et al. 2007; Koles et al. 2007) and play important roles in neural regulation (see below). Unlike vertebrate species that have multiple sialyltransferase enzymes, Drosophila possesses only one sialyltransferase, DSiaT, which facilitates the analysis of sialylation function in vivo (Koles et al. 2009). DSiaT modifies N-glycans with α2,6-linked Sia's and shows close evolutionary relationship to ST6Gal family of mammalian sialyltransferases (Koles et al. 2004; Repnikova et al. 2010). The expression of sialylation genes is dynamic and largely restricted to a subset of CNS neurons throughout development, indicating that sialylation undergoes a tight cell-specific and temporal regulation (Koles et al. 2009; Repnikova et al. 2010; Islam et al. 2013), which partially explains the low amount of sialylated glycans in the Drosophila CNS (Koles et al. 2007).

Targeted mutagenesis that inactivated the sialylation pathway has revealed that sialylated N-glycans have a specific role in the regulation of the nervous system. DSiaT mutations result in a significantly reduced life span, locomotion abnormalities and temperature-sensitive (TS) paralysis. DSiaT mutants have defects of larval NMJs, and electrophysiological analyses indicated that DSiaT regulates neuronal excitability and affects the function of Paralytic (Para), a major Drosophila voltage-gated Na+ channel (Repnikova et al. 2010). Similar phenotypes result from mutations in the CMP-sialic acid synthetase gene (CSAS) encoding an enzyme that produces activated sugar donor, CMP-Sia, for DSiaT-mediated sialylation (Islam et al. 2013). Notably, the paralysis phenotype of CSAS mutants can be ameliorated by an extra copy of para, which suggests that sialylation may control the amount of voltage-gated channels available for neural transmission (Islam et al. 2013). Moreover, genetic interactions between DSiaT and β4GalNAcTA indicated that Sia's may function by masking the interaction of LacNAc terminal structures of N-glycans with some endogenous lectins (Nakamura et al. 2012). These interesting hypotheses await further detailed investigation. Taken together, the data from Drosophila experiments shed light on a novel, nervous system-specific function of α2,6-sialylated N-glycans as essential modulators of neural transmission. This function probably represents one of the most ancient roles of Sia's in metazoan organisms. It is tempting to speculate that this role is evolutionarily conserved between flies and mammals. This intriguing possibility remains to be further investigated.

Concluding remarks

Functional modalities of N-glycosylation include effects on glycoproteins' folding, stability, trafficking, interactions with other molecules, such as recognition of specific sugar structures by lectins, as well as electrostatic and steric effects on protein dynamics and conformation. In the nervous system, N-linked glycans influence a spectrum of biological processes by mechanisms that largely depend on particular molecular and cellular contexts. Many major players of neural transmission are modified with N-glycans, and their effect on these proteins' functions can range from a modulation to the obligatory requirement (Table I). This wide range of possible effects appears to be well suited to the regulation of neural processes, providing a basis for achieving a whole additional continuum of states of neural transmission (Figure 1). Moreover, these effects can potentially provide an important link between neural transmission and metabolism. The modulation of neural excitability and synaptic transmission by N-glycans is predicted to influence essential neural functions, including plasticity and memory formation. This poses several central questions, such as what are the molecular, cellular and genetic mechanisms that underlie these roles of glycans in vivo? What are the pathobiological functions of these glycans in neurological diseases? What are the roles of lectins in mediating these glycan functions? The progress in this fundamentally important but challenging and underexplored area should be facilitated by combining in vitro and in vivo approaches, including experiments using genetically tractable model organisms.

Table I.

Examples of effects of N-glycosylation on glycoproteins involved in neural physiology

Glycoprotein Function in the nervous system Role of N-glycosylation References
SV2 (synaptic vesicle protein 2) Major synaptic vesicle protein, controls maturation step of primed synaptic vesicles Required for proper folding and trafficking to synapses (Chang and Sudhof 2009; Kwon and Chapman 2012)
Synaptophysin Major synaptic vesicle protein, regulates the kinetics of synaptic vesicle endocytosis Required for synaptic localization (Kwon and Chapman 2012)
Nicotinic acetylcholine receptors (nAChRs) Ligand-gated cation channels, regulate postsynaptic responses to neurotransmitter acetylcholine, regulates diverse brain functions Regulates desensitization and channel gating (Gehle et al. 1997; Chen et al. 1998; Nishizaki 2003)
Ionotropic glutamate receptors (iGluRs) Ligand-gated ion channels, regulate fast transmission at the majority of excitatory synapses Affects maximal currents and desensitization of AMPA and kainite receptors.
Required for folding or trafficking of NMDA receptors.
HNK-1 structure downregulates endocytosis of AMPA GluR2 subunit and promotes receptors' stability on neuronal membranes
(Partin et al. 1993; Thio et al. 1993; Yue et al. 1995; Everts et al. 1997, 1999)


(Senn et al. 2002; Yamamoto et al. 2002; Morita et al. 2009; Yoshihara et al. 2009)
Neurotransmitter transporters Major determinants of synaptic signaling, mediate uptake of neurotransmitters and regulate synaptic concentration of neurotransmitters Promotes protein stability and trafficking, increases the number of transporters at the cell surface. Sialylated glycans can affect the kinetics of GABA transporter activity and its affinity for Na+. (Tate and Blakely 1994; Olivares et al. 1995; Melikian et al. 1996; Nguyen and Amara 1996; Martinez-Maza et al. 2001; Ozaslan et al. 2003; Li et al. 2004; Cai et al. 2005; Hahn et al. 2005; Rasmussen et al. 2009; Hu et al. 2011; Kristensen et al. 2011)
Acid-sensing channels (ASICs) Acidosis-activated cation channels. Play roles in pain, neurological and psychiatric diseases, potential mechanosensory function in sensory neurons Effect on cell surface expression of ASIC1a and ASIC1b (Kadurin et al. 2008; Jing et al. 2012)
TRP channels (transient receptor potential ion channels) Playing essential roles in sensory physiology Affects the temperature threshold of TRPM8 activation in response to cold.
Affect biophysical properties of TRPC3, TRPC6 and TRPV1.
Affect expression and subcellular localization of TRPV4 and TRPV5.
(Dietrich et al. 2003; Chang et al. 2005; Wirkner et al. 2005; Xu et al. 2006; Pertusa et al. 2012)
Voltage-gated ion channels:
Potassium channels










Calcium channels


Sodium channels
Control cell excitability, generate action potentials, affect neural transmission

Affects cell surface expression and stability of Kv11.1, Kv1.3, Kv12.2 and Kv1.4.
Affects gating of Kv1.1, Kv1.5, Kv12.2, IsK
Affects trafficking and gating of Kv1.2
Affects simulated action potentials for Kv1.1 and Kv1.2 .
Sialylation: Affects gating of Kv1.1, Kv1.5, Kv3.1
Gating of Drosophila Shaker channel expressed heterologously in mammalian cells is affected by N-glycans and sialylation
Controls activity and cell surface expression of Cav3.2, affects glucose-dependent potentiation
Affects gating of Nav1.4 and Nav1.5,
Alters steady-state inactivation of Nav1.9
Sialylation: Affects gating of Nav1.4, and Nav1.5, electroplax channel. Sialylation of Nav beta(2) subunit affects gating of Nav1.5.Affects functional properties of Drosophila Nav Para (unknown if the effect is direct)


(Thornhill et al. 1996; Freeman et al. 2000; Gong et al. 2002; Watanabe et al. 2003, 2004; Sutachan et al. 2005; Watanabe et al. 2007; Johnson and Bennett 2008; Noma et al. 2009; Zhu et al. 2009; Schwetz et al. 2010; Hall et al. 2011; Zhu et al. 2012)




(Weiss et al. 2013)


(Recio-Pinto et al. 1990; Bennett et al. 1997; Zhang et al. 1999; Johnson et al. 2004; Cronin et al. 2005; Stocker and Bennett 2006; Repnikova et al. 2010)

Fig. 1.

Fig. 1.

N-glycosylation can regulate neural transmission by modulating voltage-gated ion channels that generate action potential on neuronal membranes, and by affecting synaptic transmission via impact on the function of synaptic vesicle proteins and postsynaptic neurotransmitter receptors. N-glycosylation is sketched as a single generic N-glycan (not to scale), while the number of modifications sites can be different for distinct proteins, and particular glycan structures can significantly vary and include some specific modifications, such as polysialylation and the HNK-1 epitopes (not shown).

Abbreviations

β4GalNAcTA, β1,4-N-acetylgalactosaminyltransferase A; AMPA, α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid; ASIC, acid-sensing ion channel; ConA, concanavalin A; CSAS, CMP-sialic acid synthetase; DSiaT, Drosophila sialyltransferase; GABA, γ-aminobutyric acid; GlcNAcT, β-1,6-N-acetylglucosaminyltransferase; NMJ, neuromuscular junction; iGluR, ionotropic glutamate receptor; LacNAc, N-acetyllactosamine; nAChR, nicotinic acetylcholine receptor; NCAM, neural cell adhesion molecule; NMDA, N-methyl-D-aspartate; Para, paralytic; PSA, polysialic acid; Sia, sialic acid(s); SV2, synaptic vesicle protein 2; TRP, transient receptor potential; UDP-GlcNAc, uridine diphospho-N-acetylglucosamine; GalNAc, N-acetylgalactosamine.

Acknowledgements

We are grateful to Dr. Mark Zoran for stimulating discussions, Dr. Linda Baum and Dr. Mark Lehrman for their inspiring encouragement to work on the manuscript; Dr. Daria Panina for valuable comments. We thank all members of the Panin Laboratory for helpful discussions. This work was supported in part by NIH grant NS075534 to V.M.P.

References

  1. Abu-Qarn M, Eichler J, Sharon N. Not just for Eukarya anymore: protein glycosylation in Bacteria and Archaea. Curr Opin Struct Biol. 2008;18:544–550. doi: 10.1016/j.sbi.2008.06.010. [DOI] [PubMed] [Google Scholar]
  2. Ahrens J, Foadi N, Eberhardt A, Haeseler G, Dengler R, Leffler A, Muhlenhoff M, Gerardy-Schahn R, Leuwer M. Defective polysialylation and sialylation induce opposite effects on gating of the skeletal Na+ channel NaV1.4 in Chinese hamster ovary cells. Pharmacology. 2011;87:311–317. doi: 10.1159/000327389. [DOI] [PubMed] [Google Scholar]
  3. Amano M, Galvan M, He J, Baum LG. The ST6Gal I sialyltransferase selectively modifies N-glycans on CD45 to negatively regulate galectin-1-induced CD45 clustering, phosphatase modulation, and T cell death. J Biol Chem. 2003;278:7469–7475. doi: 10.1074/jbc.M209595200. [DOI] [PubMed] [Google Scholar]
  4. Angata K, Huckaby V, Ranscht B, Terskikh A, Marth JD, Fukuda M. Polysialic acid-directed migration and differentiation of neural precursors are essential for mouse brain development. Mol Cell Biol. 2007;27:6659–6668. doi: 10.1128/MCB.00205-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Aoki K, Perlman M, Lim JM, Cantu R, Wells L, Tiemeyer M. Dynamic developmental elaboration of N-linked glycan complexity in the Drosophila melanogaster embryo. J Biol Chem. 2007;282:9127–9142. doi: 10.1074/jbc.M606711200. [DOI] [PubMed] [Google Scholar]
  6. Apweiler R, Hermjakob H, Sharon N. On the frequency of protein glycosylation, as deduced from analysis of the SWISS-PROT database. Biochim Biophys Acta. 1999;1473:4–8. doi: 10.1016/s0304-4165(99)00165-8. [DOI] [PubMed] [Google Scholar]
  7. Armstrong N, Sun Y, Chen GQ, Gouaux E. Structure of a glutamate-receptor ligand-binding core in complex with kainate. Nature. 1998;395:913–917. doi: 10.1038/27692. [DOI] [PubMed] [Google Scholar]
  8. Baas S, Sharrow M, Kotu V, Middleton M, Nguyen K, Flanagan-Steet H, Aoki K, Tiemeyer M. Sugar-free frosting, a homolog of SAD kinase, drives neural-specific glycan expression in the Drosophila embryo. Development. 2011;138:553–563. doi: 10.1242/dev.055376. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bennett ES. Isoform-specific effects of sialic acid on voltage-dependent Na+ channel gating: functional sialic acids are localized to the S5-S6 loop of domain I. J Physiol. 2002;538:675–690. doi: 10.1113/jphysiol.2001.013285. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bennett E, Urcan MS, Tinkle SS, Koszowski AG, Levinson SR. Contribution of sialic acid to the voltage dependence of sodium channel gating: a possible electrostatic mechanism. J General Physiol. 1997;109:327–343. doi: 10.1085/jgp.109.3.327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Boquet I, Hitier R, Dumas M, Chaminade M, Preat T. Central brain postembryonic development in Drosophila: implication of genes expressed at the interhemispheric junction. J Neurobiol. 2000;42:33–48. [PubMed] [Google Scholar]
  12. Cai G, Salonikidis PS, Fei J, Schwarz W, Schulein R, Reutter W, Fan H. The role of N-glycosylation in the stability, trafficking and GABA-uptake of GABA-transporter 1. Terminal N-glycans facilitate efficient GABA-uptake activity of the GABA transporter. FEBS J. 2005;272:1625–1638. doi: 10.1111/j.1742-4658.2005.04595.x. [DOI] [PubMed] [Google Scholar]
  13. Casiraghi LP, Croci DO, Poirier F, Rabinovich GA, Golombek DA. “Time sweet time”: circadian characterization of galectin-1 null mice. J Circadian Rhythms. 2010;8:4–0. doi: 10.1186/1740-3391-8-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Castillo C, Diaz ME, Balbi D, Thornhill WB, Recio-Pinto E. Changes in sodium channel function during postnatal brain development reflect increases in the level of channel sialidation. Brain Res Dev Brain Res. 1997;104:119–130. doi: 10.1016/s0165-3806(97)00159-4. [DOI] [PubMed] [Google Scholar]
  15. Cha SK, Ortega B, Kurosu H, Rosenblatt KP, Kuro OM, Huang CL. Removal of sialic acid involving Klotho causes cell-surface retention of TRPV5 channel via binding to galectin-1. Proc Natl Acad Sci USA. 2008;105:9805–9810. doi: 10.1073/pnas.0803223105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Chang Q, Hoefs S, van der Kemp AW, Topala CN, Bindels RJ, Hoenderop JG. The beta-glucuronidase klotho hydrolyzes and activates the TRPV5 channel. Science. 2005;310:490–493. doi: 10.1126/science.1114245. [DOI] [PubMed] [Google Scholar]
  17. Chang WP, Sudhof TC. SV2 renders primed synaptic vesicles competent for Ca2+-induced exocytosis. J Neurosci. 2009;29:883–897. doi: 10.1523/JNEUROSCI.4521-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Chen D, Dang H, Patrick JW. Contributions of N-linked glycosylation to the expression of a functional alpha7-nicotinic receptor in Xenopus oocytes. J Neurochem. 1998;70:349–357. doi: 10.1046/j.1471-4159.1998.70010349.x. [DOI] [PubMed] [Google Scholar]
  19. Chen L. In pursuit of the high-resolution structure of nicotinic acetylcholine receptors. J Physiol. 2010;588:557–564. doi: 10.1113/jphysiol.2009.184085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Colley KJ. Structural basis for the polysialylation of the neural cell adhesion molecule. Adv Exp Med Biol. 2010;663:111–126. doi: 10.1007/978-1-4419-1170-4_7. [DOI] [PubMed] [Google Scholar]
  21. Cotella D, Radicke S, Bortoluzzi A, Ravens U, Wettwer E, Santoro C, Sblattero D. Impaired glycosylation blocks DPP10 cell surface expression and alters the electrophysiology of Ito channel complex. Pflugers Arch. 2010;460:87–97. doi: 10.1007/s00424-010-0824-2. [DOI] [PubMed] [Google Scholar]
  22. Cronin NB, O'Reilly A, Duclohier H, Wallace BA. Effects of deglycosylation of sodium channels on their structure and function. Biochemistry. 2005;44:441–449. doi: 10.1021/bi048741q. [DOI] [PubMed] [Google Scholar]
  23. da Costa CJ, Kaiser DE, Baenziger JE. Role of glycosylation and membrane environment in nicotinic acetylcholine receptor stability. Biophys J. 2005;88:1755–1764. doi: 10.1529/biophysj.104.052944. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Dani N, Broadie K. Glycosylated synaptomatrix regulation of trans-synaptic signaling. Dev Neurobiol. 2012;72:2–21. doi: 10.1002/dneu.20891. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Dellisanti CD, Yao Y, Stroud JC, Wang ZZ, Chen L. Crystal structure of the extracellular domain of nAChR alpha1 bound to alpha-bungarotoxin at 1.94 A resolution. Nat Neurosci. 2007;10:953–962. doi: 10.1038/nn1942. [DOI] [PubMed] [Google Scholar]
  26. Dennis JW, Nabi IR, Demetriou M. Metabolism, cell surface organization, and disease. Cell. 2009;139:1229–1241. doi: 10.1016/j.cell.2009.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Dietrich A, Mederos y Schnitzler M, Emmel J, Kalwa H, Hofmann T, Gudermann T. N-linked protein glycosylation is a major determinant for basal TRPC3 and TRPC6 channel activity. J Biol Chem. 2003;278:47842–47852. doi: 10.1074/jbc.M302983200. [DOI] [PubMed] [Google Scholar]
  28. Dityatev A, Schachner M, Sonderegger P. The dual role of the extracellular matrix in synaptic plasticity and homeostasis. Nat Rev Neurosci. 2010;11:735–746. doi: 10.1038/nrn2898. [DOI] [PubMed] [Google Scholar]
  29. Ednie AR, Bennett ES. Modulation of voltage-gated ion channels by sialylation. Compr Physiol. 2012;2:1269–1301. doi: 10.1002/cphy.c110044. [DOI] [PubMed] [Google Scholar]
  30. Elmer LW, O'Brien BJ, Nutter TJ, Angelides KJ. Physicochemical characterization of the alpha-peptide of the sodium channel from rat brain. Biochemistry. 1985;24:8128–8137. doi: 10.1021/bi00348a044. [DOI] [PubMed] [Google Scholar]
  31. Engel AG. Current status of the congenital myasthenic syndromes. Neuromuscul Disord. 2012;22:99–111. doi: 10.1016/j.nmd.2011.10.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Everts I, Petroski R, Kizelsztein P, Teichberg VI, Heinemann SF, Hollmann M. Lectin-induced inhibition of desensitization of the kainate receptor GluR6 depends on the activation state and can be mediated by a single native or ectopic N-linked carbohydrate side chain. J Neurosci. 1999;19:916–927. doi: 10.1523/JNEUROSCI.19-03-00916.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Everts I, Villmann C, Hollmann M. N-Glycosylation is not a prerequisite for glutamate receptor function but Is essential for lectin modulation. Mol Pharmacol. 1997;52:861–873. doi: 10.1124/mol.52.5.861. [DOI] [PubMed] [Google Scholar]
  34. Fay AM, Bowie D. Concanavalin-A reports agonist-induced conformational changes in the intact GluR6 kainate receptor. J Physiol. 2006;572:201–213. doi: 10.1113/jphysiol.2005.103580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Freeman LC, Lippold JJ, Mitchell KE. Glycosylation influences gating and pH sensitivity of I(sK) J Membr Biol. 2000;177:65–79. doi: 10.1007/s002320001100. [DOI] [PubMed] [Google Scholar]
  36. Freeze HH, Eklund EA, Ng BG, Patterson MC. Neurology of inherited glycosylation disorders. Lancet Neurol. 2012;11:453–466. doi: 10.1016/S1474-4422(12)70040-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Gehle VM, Walcott EC, Nishizaki T, Sumikawa K. N-glycosylation at the conserved sites ensures the expression of properly folded functional ACh receptors. Brain Res Mol Brain Res. 1997;45:219–229. doi: 10.1016/s0169-328x(96)00256-2. [DOI] [PubMed] [Google Scholar]
  38. Gill MB, Vivithanaporn P, Swanson GT. Glutamate binding and conformational flexibility of ligand-binding domains are critical early determinants of efficient kainate receptor biogenesis. J Biol Chem. 2009;284:14503–14512. doi: 10.1074/jbc.M900510200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Gong Q, Anderson CL, January CT, Zhou Z. Role of glycosylation in cell surface expression and stability of HERG potassium channels. Am J Physiol Heart Circ Physiol. 2002;283:H77–H84. doi: 10.1152/ajpheart.00008.2002. [DOI] [PubMed] [Google Scholar]
  40. Gottlieb C, Baenziger J, Kornfeld S. Deficient uridine diphosphate-N-acetylglucosamine:glycoprotein N-acetylglucosaminyltransferase activity in a clone of Chinese hamster ovary cells with altered surface glycoproteins. J Biol Chem. 1975;250:3303–3309. [PubMed] [Google Scholar]
  41. Gurba KN, Hernandez CC, Hu N, Macdonald RL. GABRB3 mutation, G32R, associated with childhood absence epilepsy alters alpha1beta3gamma2L gamma-aminobutyric acid type A (GABAA) receptor expression and channel gating. J Biol Chem. 2012;287:12083–12097. doi: 10.1074/jbc.M111.332528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Hahn MK, Mazei-Robison MS, Blakely RD. Single nucleotide polymorphisms in the human norepinephrine transporter gene affect expression, trafficking, antidepressant interaction, and protein kinase C regulation. Mol Pharmacol. 2005;68:457–466. doi: 10.1124/mol.105.011270. [DOI] [PubMed] [Google Scholar]
  43. 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]
  44. 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]
  45. Hall MK, Reutter W, Lindhorst T, Schwalbe RA. Biochemical engineering of the N-acyl side chain of sialic acids alters the kinetics of a glycosylated potassium channel Kv3.1. FEBS Lett. 2011;585:3322–3327. doi: 10.1016/j.febslet.2011.09.021. [DOI] [PubMed] [Google Scholar]
  46. Helenius A, Aebi M. Intracellular functions of N-linked glycans. Science. 2001;291:2364–2369. doi: 10.1126/science.291.5512.2364. [DOI] [PubMed] [Google Scholar]
  47. Henion TR, Faden AA, Knott TK, Schwarting GA. beta3GnT2 maintains adenylyl cyclase-3 signaling and axon guidance molecule expression in the olfactory epithelium. J Neurosci. 2011;31:6576–6586. doi: 10.1523/JNEUROSCI.0224-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Hildebrandt H, Muhlenhoff M, Oltmann-Norden I, Rockle I, Burkhardt H, Weinhold B, Gerardy-Schahn R. Imbalance of neural cell adhesion molecule and polysialyltransferase alleles causes defective brain connectivity. Brain. 2009;132:2831–2838. doi: 10.1093/brain/awp117. [DOI] [PubMed] [Google Scholar]
  49. Hollmann M, Maron C, Heinemann S. N-glycosylation site tagging suggests a three transmembrane domain topology for the glutamate receptor GluR1. Neuron. 1994;13:1331–1343. doi: 10.1016/0896-6273(94)90419-7. [DOI] [PubMed] [Google Scholar]
  50. Hu J, Fei J, Reutter W, Fan H. Involvement of sialic acid in the regulation of gamma-aminobutyric acid uptake activity of gamma-aminobutyric acid transporter 1. Glycobiology. 2011;21:329–339. doi: 10.1093/glycob/cwq166. [DOI] [PubMed] [Google Scholar]
  51. Ioffe E, Stanley P. Mice lacking N-acetylglucosaminyltransferase I activity die at mid-gestation, revealing an essential role for complex or hybrid N-linked carbohydrates. Proc Natl Acad Sci USA. 1994;91:728–732. doi: 10.1073/pnas.91.2.728. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Isaev D, Isaeva E, Shatskih T, Zhao Q, Smits NC, Shworak NW, Khazipov R, Holmes GL. Role of extracellular sialic acid in regulation of neuronal and network excitability in the rat hippocampus. J Neurosci. 2007;27:11587–11594. doi: 10.1523/JNEUROSCI.2033-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Isaev D, Zhao Q, Kleen JK, Lenck-Santini PP, Adstamongkonkul D, Isaeva E, Holmes GL. Neuroaminidase reduces interictal spikes in a rat temporal lobe epilepsy model. Epilepsia. 2011;52:e12–e15. doi: 10.1111/j.1528-1167.2011.02988.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Isaeva E, Lushnikova I, Savrasova A, Skibo G, Holmes GL, Isaev D. Blockade of endogenous neuraminidase leads to an increase of neuronal excitability and activity-dependent synaptogenesis in the rat hippocampus. Eur J Neurosci. 2010;32:1889–1896. doi: 10.1111/j.1460-9568.2010.07468.x. [DOI] [PubMed] [Google Scholar]
  55. Islam R, Nakamura M, Scott H, Repnikova E, Carnahan M, Pandey D, Caster C, Khan S, Zimmermann T, Zoran MJ, et al. The role of Drosophila cytidine monophosphate-sialic acid synthetase in the nervous system. J Neurosci. 2013;33:12306–12315. doi: 10.1523/JNEUROSCI.5220-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. James WM, Agnew WS. Multiple oligosaccharide chains in the voltage-sensitive Na channel from electrophorus electricus: evidence for alpha-2,8-linked polysialic acid. Biochem Biophys Res Commun. 1987;148:817–826. doi: 10.1016/0006-291x(87)90949-1. [DOI] [PubMed] [Google Scholar]
  57. Janz R, Goda Y, Geppert M, Missler M, Sudhof TC. SV2A and SV2B function as redundant Ca2+ regulators in neurotransmitter release. Neuron. 1999;24:1003–1016. doi: 10.1016/s0896-6273(00)81046-6. [DOI] [PubMed] [Google Scholar]
  58. Jing L, Chu XP, Jiang YQ, Collier DM, Wang B, Jiang Q, Snyder PM, Zha XM. N-glycosylation of acid-sensing ion channel 1a regulates its trafficking and acidosis-induced spine remodeling. J Neurosci. 2012;32:4080–4091. doi: 10.1523/JNEUROSCI.5021-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Johnson D, Bennett ES. Isoform-specific effects of the beta(2) subunit on voltage-gated sodium channel gating. J Biol Chem. 2006;281:25875–25881. doi: 10.1074/jbc.M605060200. [DOI] [PubMed] [Google Scholar]
  60. Johnson D, Bennett ES. Gating of the shaker potassium channel is modulated differentially by N-glycosylation and sialic acids. Pflugers Arch. 2008;456:393–405. doi: 10.1007/s00424-007-0378-0. [DOI] [PubMed] [Google Scholar]
  61. Johnson D, Montpetit ML, Stocker PJ, Bennett ES. The sialic acid component of the beta(1) subunit modulates voltage-gated sodium channel function. J Biol Chem. 2004;279:44303–44310. doi: 10.1074/jbc.M408900200. [DOI] [PubMed] [Google Scholar]
  62. Kadurin I, Golubovic A, Leisle L, Schindelin H, Grunder S. Differential effects of N-glycans on surface expression suggest structural differences between the acid-sensing ion channel (ASIC) 1a and ASIC1b. Biochem J. 2008;412:469–475. doi: 10.1042/BJ20071614. [DOI] [PubMed] [Google Scholar]
  63. Kariya Y, Kato R, Itoh S, Fukuda T, Shibukawa Y, Sanzen N, Sekiguchi K, Wada Y, Kawasaki N, Gu J. N-Glycosylation of laminin-332 regulates its biological functions. A novel function of the bisecting GlcNAc. J Biol Chem. 2008;283:33036–33045. doi: 10.1074/jbc.M804526200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Kleene R, Schachner M. Glycans and neural cell interactions. Nat Rev Neurosci. 2004;5:195–208. doi: 10.1038/nrn1349. [DOI] [PubMed] [Google Scholar]
  65. Koles K, Irvine KD, Panin VM. Functional characterization of Drosophila sialyltransferase. J Biol Chem. 2004;279:4346–4357. doi: 10.1074/jbc.M309912200. [DOI] [PubMed] [Google Scholar]
  66. Koles K, Lim JM, Aoki K, Porterfield M, Tiemeyer M, Wells L, Panin V. Identification of N-glycosylated proteins from the central nervous system of Drosophila melanogaster. Glycobiology. 2007;17:1388–1403. doi: 10.1093/glycob/cwm097. [DOI] [PubMed] [Google Scholar]
  67. Koles K, Repnikova E, Pavlova G, Korochkin LI, Panin VM. Sialylation in protostomes: a perspective from Drosophila genetics and biochemistry. Glycoconj J. 2009;26:313–324. doi: 10.1007/s10719-008-9154-4. [DOI] [PubMed] [Google Scholar]
  68. Kristensen AS, Andersen J, Jorgensen TN, Sorensen L, Eriksen J, Loland CJ, Stromgaard K, Gether U. SLC6 neurotransmitter transporters: structure, function, and regulation. Pharmacol Rev. 2011;63:585–640. doi: 10.1124/pr.108.000869. [DOI] [PubMed] [Google Scholar]
  69. Kwon SE, Chapman ER. Glycosylation is dispensable for sorting of synaptotagmin 1 but is critical for targeting of SV2 and synaptophysin to recycling synaptic vesicles. J Biol Chem. 2012;287:35658–35668. doi: 10.1074/jbc.M112.398883. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Lekishvili T, Hesketh S, Brazier MW, Brown DR. Mouse galectin-1 inhibits the toxicity of glutamate by modifying NR1 NMDA receptor expression. Eur J Neurosci. 2006;24:3017–3025. doi: 10.1111/j.1460-9568.2006.05207.x. [DOI] [PubMed] [Google Scholar]
  71. Leonard R, Rendic D, Rabouille C, Wilson IB, Preat T, Altmann F. The Drosophila fused lobes gene encodes an N-acetylglucosaminidase involved in N-glycan processing. J Biol Chem. 2006;281:4867–4875. doi: 10.1074/jbc.M511023200. [DOI] [PubMed] [Google Scholar]
  72. Leunissen EH, Nair AV, Bull C, Lefeber DJ, van Delft FL, Bindels RJ, Hoenderop JG. The epithelial calcium channel TRPV5 is regulated differentially by klotho and sialidase. J Biol Chem. 2013;288:29238–29246. doi: 10.1074/jbc.M113.473520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Li LB, Chen N, Ramamoorthy S, Chi L, Cui XN, Wang LC, Reith ME. The role of N-glycosylation in function and surface trafficking of the human dopamine transporter. J Biol Chem. 2004;279:21012–21020. doi: 10.1074/jbc.M311972200. [DOI] [PubMed] [Google Scholar]
  74. Lowe JB, Marth JD. A genetic approach to Mammalian glycan function. Annu Rev Biochem. 2003;72:643–691. doi: 10.1146/annurev.biochem.72.121801.161809. [DOI] [PubMed] [Google Scholar]
  75. Mah SJ, Cornell E, Mitchell NA, Fleck MW. Glutamate receptor trafficking: endoplasmic reticulum quality control involves ligand binding and receptor function. J Neurosci. 2005;25:2215–2225. doi: 10.1523/JNEUROSCI.4573-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Martinez-Maza R, Poyatos I, Lopez-Corcuera B, Nu E, Gimenez C, Zafra F, Aragon C. The role of N-glycosylation in transport to the plasma membrane and sorting of the neuronal glycine transporter GLYT2. J Biol Chem. 2001;276:2168–2173. doi: 10.1074/jbc.M006774200. [DOI] [PubMed] [Google Scholar]
  77. McKemy DD, Neuhausser WM, Julius D. Identification of a cold receptor reveals a general role for TRP channels in thermosensation. Nature. 2002;416:52–58. doi: 10.1038/nature719. [DOI] [PubMed] [Google Scholar]
  78. Melikian HE, Ramamoorthy S, Tate CG, Blakely RD. Inability to N-glycosylate the human norepinephrine transporter reduces protein stability, surface trafficking, and transport activity but not ligand recognition. Mol Pharmacol. 1996;50:266–276. [PubMed] [Google Scholar]
  79. Messner DJ, Catterall WA. The sodium channel from rat brain. Separation and characterization of subunits. J Biol Chem. 1985;260:10597–10604. [PubMed] [Google Scholar]
  80. Metzler M, Gertz A, Sarkar M, Schachter H, Schrader JW, Marth JD. Complex asparagine-linked oligosaccharides are required for morphogenic events during post-implantation development. EMBO J. 1994;13:2056–2065. doi: 10.1002/j.1460-2075.1994.tb06480.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Miller JA, Agnew WS, Levinson SR. Principal glycopeptide of the tetrodotoxin/saxitoxin binding protein from Electrophorus electricus: isolation and partial chemical and physical characterization. Biochemistry. 1983;22:462–470. doi: 10.1021/bi00271a032. [DOI] [PubMed] [Google Scholar]
  82. Miwa JM, Freedman R, Lester HA. Neural systems governed by nicotinic acetylcholine receptors: emerging hypotheses. Neuron. 2011;70:20–33. doi: 10.1016/j.neuron.2011.03.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Montpetit ML, Stocker PJ, Schwetz TA, Harper JM, Norring SA, Schaffer L, North SJ, Jang-Lee J, Gilmartin T, Head SR, et al. Regulated and aberrant glycosylation modulate cardiac electrical signaling. Proc Natl Acad Sci USA. 2009;106:16517–16522. doi: 10.1073/pnas.0905414106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Moremen KW, Tiemeyer M, Nairn AV. Vertebrate protein glycosylation: diversity, synthesis and function. Nat Rev Mol Cell Biol. 2012;13:448–462. doi: 10.1038/nrm3383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Morita I, Kakuda S, Takeuchi Y, Itoh S, Kawasaki N, Kizuka Y, Kawasaki T, Oka S. HNK-1 glyco-epitope regulates the stability of the glutamate receptor subunit GluR2 on the neuronal cell surface. J Biol Chem. 2009;284:30209–30217. doi: 10.1074/jbc.M109.024208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Muhlenhoff M, Eckhardt M, Gerardy-Schahn R. Polysialic acid: three-dimensional structure, biosynthesis and function. Curr Opin Struct Biol. 1998;8:558–564. doi: 10.1016/s0959-440x(98)80144-9. [DOI] [PubMed] [Google Scholar]
  87. Muhlenhoff M, Oltmann-Norden I, Weinhold B, Hildebrandt H, Gerardy-Schahn R. Brain development needs sugar: the role of polysialic acid in controlling NCAM functions. Biol Chem. 2009;390:567–574. doi: 10.1515/BC.2009.078. [DOI] [PubMed] [Google Scholar]
  88. Muller D, Wang C, Skibo G, Toni N, Cremer H, Calaora V, Rougon G, Kiss JZ. PSA-NCAM is required for activity-induced synaptic plasticity. Neuron. 1996;17:413–422. doi: 10.1016/s0896-6273(00)80174-9. [DOI] [PubMed] [Google Scholar]
  89. Nakamura M, Pandey D, Panin VM. Genetic Interactions Between Drosophila sialyltransferase and beta1,4-N-acetylgalactosaminyltransferase-A Genes Indicate Their Involvement in the Same Pathway. G3 (Bethesda) 2012;2:653–656. doi: 10.1534/g3.112.001974. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Nanao MH, Green T, Stern-Bach Y, Heinemann SF, Choe S. Structure of the kainate receptor subunit GluR6 agonist-binding domain complexed with domoic acid. Proc Natl Acad Sci USA. 2005;102:1708–1713. doi: 10.1073/pnas.0409573102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Nguyen TT, Amara SG. N-linked oligosaccharides are required for cell surface expression of the norepinephrine transporter but do not influence substrate or inhibitor recognition. J Neurochem. 1996;67:645–655. doi: 10.1046/j.1471-4159.1996.67020645.x. [DOI] [PubMed] [Google Scholar]
  92. Nishizaki T. N-glycosylation sites on the nicotinic ACh receptor subunits regulate receptor channel desensitization and conductance. Brain Res Mol Brain Res. 2003;114:172–176. doi: 10.1016/s0169-328x(03)00171-2. [DOI] [PubMed] [Google Scholar]
  93. Noma K, Kimura K, Minatohara K, Nakashima H, Nagao Y, Mizoguchi A, Fujiyoshi Y. Triple N-glycosylation in the long S5-P loop regulates the activation and trafficking of the Kv12.2 potassium channel. J Biol Chem. 2009;284:33139–33150. doi: 10.1074/jbc.M109.021519. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. O'Brien RJ, Xu D, Petralia RS, Steward O, Huganir RL, Worley P. Synaptic clustering of AMPA receptors by the extracellular immediate-early gene product Narp. Neuron. 1999;23:309–323. doi: 10.1016/s0896-6273(00)80782-5. [DOI] [PubMed] [Google Scholar]
  95. Olivares L, Aragon C, Gimenez C, Zafra F. The role of N-glycosylation in the targeting and activity of the GLYT1 glycine transporter. J Biol Chem. 1995;270:9437–9442. doi: 10.1074/jbc.270.16.9437. [DOI] [PubMed] [Google Scholar]
  96. Ozaslan D, Wang S, Ahmed BA, Kocabas AM, McCastlain JC, Bene A, Kilic F. Glycosyl modification facilitates homo- and hetero-oligomerization of the serotonin transporter. A specific role for sialic acid residues. J Biol Chem. 2003;278:43991–44000. doi: 10.1074/jbc.M306360200. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  97. Partin KM, Patneau DK, Winters CA, Mayer ML, Buonanno A. Selective modulation of desensitization at AMPA versus kainate receptors by cyclothiazide and concanavalin A. Neuron. 1993;11:1069–1082. doi: 10.1016/0896-6273(93)90220-l. [DOI] [PubMed] [Google Scholar]
  98. Peier AM, Moqrich A, Hergarden AC, Reeve AJ, Andersson DA, Story GM, Earley TJ, Dragoni I, McIntyre P, Bevan S, et al. A TRP channel that senses cold stimuli and menthol. Cell. 2002;108:705–715. doi: 10.1016/s0092-8674(02)00652-9. [DOI] [PubMed] [Google Scholar]
  99. Pertusa M, Madrid R, Morenilla-Palao C, Belmonte C, Viana F. N-glycosylation of TRPM8 ion channels modulates temperature sensitivity of cold thermoreceptor neurons. J Biol Chem. 2012;287:18218–18229. doi: 10.1074/jbc.M111.312645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Plachta N, Annaheim C, Bissiere S, Lin S, Ruegg M, Hoving S, Muller D, Poirier F, Bibel M, Barde YA. Identification of a lectin causing the degeneration of neuronal processes using engineered embryonic stem cells. Nat Neurosci. 2007;10:712–719. doi: 10.1038/nn1897. [DOI] [PubMed] [Google Scholar]
  101. Rasmussen TN, Plenge P, Bay T, Egebjerg J, Gether U. A single nucleotide polymorphism in the human serotonin transporter introduces a new site for N-linked glycosylation. Neuropharmacology. 2009;57:287–294. doi: 10.1016/j.neuropharm.2009.05.009. [DOI] [PubMed] [Google Scholar]
  102. Recio-Pinto E, Thornhill WB, Duch DS, Levinson SR, Urban BW. Neuraminidase treatment modifies the function of electroplax sodium channels in planar lipid bilayers. Neuron. 1990;5:675–684. doi: 10.1016/0896-6273(90)90221-z. [DOI] [PubMed] [Google Scholar]
  103. Repnikova E, Koles K, Nakamura M, Pitts J, Li H, Ambavane A, Zoran MJ, Panin VM. Sialyltransferase regulates nervous system function in Drosophila. J Neurosci. 2010;30:6466–6476. doi: 10.1523/JNEUROSCI.5253-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Roberts RH, Barchi RL. The voltage-sensitive sodium channel from rabbit skeletal muscle. Chemical characterization of subunits. J Biol Chem. 1987;262:2298–2303. [PubMed] [Google Scholar]
  105. Rushton E, Rohrbough J, Deutsch K, Broadie K. Structure-function analysis of endogenous lectin mind-the-gap in synaptogenesis. Dev Neurobiol. 2012;72:1161–1179. doi: 10.1002/dneu.22006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Rutishauser U. Polysialic acid in the plasticity of the developing and adult vertebrate nervous system. Nat Rev Neurosci. 2008;9:26–35. doi: 10.1038/nrn2285. [DOI] [PubMed] [Google Scholar]
  107. Sarkar M, Iliadi KG, Leventis PA, Schachter H, Boulianne GL. Neuronal expression of Mgat1 rescues the shortened life span of Drosophila Mgat11 null mutants and increases life span. Proc Natl Acad Sci USA. 2010;107:9677–9682. doi: 10.1073/pnas.1004431107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Sarkar M, Leventis PA, Silvescu CI, Reinhold VN, Schachter H, Boulianne GL. Null mutations in Drosophila N-acetylglucosaminyltransferase I produce defects in locomotion and a reduced life span. J Biol Chem. 2006;281:12776–12785. doi: 10.1074/jbc.M512769200. [DOI] [PubMed] [Google Scholar]
  109. Schachter H. Mgat1-dependent N-glycans are essential for the normal development of both vertebrate and invertebrate metazoans. Semin Cell Dev Biol. 2010;21:609–615. doi: 10.1016/j.semcdb.2010.02.010. [DOI] [PubMed] [Google Scholar]
  110. Schwalbe RA, Corey MJ, Cartwright TA. Novel Kv3 glycoforms differentially expressed in adult mammalian brain contain sialylated N-glycans. Biochem Cell Biol. 2008;86:21–30. doi: 10.1139/o07-152. [DOI] [PubMed] [Google Scholar]
  111. Schwetz TA, Norring SA, Bennett ES. N-glycans modulate K(v)1.5 gating but have no effect on K(v)1.4 gating. Biochim Biophys Acta. 2010;1798:367–375. doi: 10.1016/j.bbamem.2009.11.018. [DOI] [PubMed] [Google Scholar]
  112. Sekine SU, Haraguchi S, Chao K, Kato T, Luo L, Miura M, Chihara T. Meigo governs dendrite targeting specificity by modulating ephrin level and N-glycosylation. Nat Neurosci. 2013;16:683–691. doi: 10.1038/nn.3389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Senderek J, Muller JS, Dusl M, Strom TM, Guergueltcheva V, Diepolder I, Laval SH, Maxwell S, Cossins J, Krause S, et al. Hexosamine biosynthetic pathway mutations cause neuromuscular transmission defect. Am J Hum Genet. 2011;88:162–172. doi: 10.1016/j.ajhg.2011.01.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Senn C, Kutsche M, Saghatelyan A, Bosl MR, Lohler J, Bartsch U, Morellini F, Schachner M. Mice deficient for the HNK-1 sulfotransferase show alterations in synaptic efficacy and spatial learning and memory. Mol Cell Neurosci. 2002;20:712–729. doi: 10.1006/mcne.2002.1142. [DOI] [PubMed] [Google Scholar]
  115. Sharon N. Lectins: carbohydrate-specific reagents and biological recognition molecules. J Biol Chem. 2007;282:2753–2764. doi: 10.1074/jbc.X600004200. [DOI] [PubMed] [Google Scholar]
  116. Stanley P, Narasimhan S, Siminovitch L, Schachter H. Chinese hamster ovary cells selected for resistance to the cytotoxicity of phytohemagglutinin are deficient in a UDP-N-acetylglucosamine–glycoprotein N-acetylglucosaminyltransferase activity. Proc Natl Acad Sci USA. 1975;72:3323–3327. doi: 10.1073/pnas.72.9.3323. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Stanley P, Schachter H, Taniguchi N. N-Glycans. In: Varki A, Cummings RD, et al., editors. Essentials of Glycobiology. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press; 2009. [PubMed] [Google Scholar]
  118. Stocker PJ, Bennett ES. Differential sialylation modulates voltage-gated Na+ channel gating throughout the developing myocardium. J General Physiol. 2006;127:253–265. doi: 10.1085/jgp.200509423. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Sumikawa K, Parker I, Miledi R. Effect of tunicamycin on the expression of functional brain neurotransmitter receptors and voltage-operated channels in Xenopus oocytes. Brain Res. 1988;464:191–199. doi: 10.1016/0169-328x(88)90025-3. [DOI] [PubMed] [Google Scholar]
  120. Sutachan JJ, Watanabe I, Zhu J, Gottschalk A, Recio-Pinto E, Thornhill WB. Effects of Kv1.1 channel glycosylation on C-type inactivation and simulated action potentials. Brain Res. 2005;1058:30–43. doi: 10.1016/j.brainres.2005.07.050. [DOI] [PubMed] [Google Scholar]
  121. Tate CG, Blakely RD. The effect of N-linked glycosylation on activity of the Na(+)- and Cl(-)-dependent serotonin transporter expressed using recombinant baculovirus in insect cells. J Biol Chem. 1994;269:26303–26310. [PubMed] [Google Scholar]
  122. Thio LL, Clifford DB, Zorumski CF. Concanavalin A enhances excitatory synaptic transmission in cultured rat hippocampal neurons. Synapse. 1993;13:94–97. doi: 10.1002/syn.890130111. [DOI] [PubMed] [Google Scholar]
  123. Thornhill WB, Wu MB, Jiang X, Wu X, Morgan PT, Margiotta JF. Expression of Kv1.1 delayed rectifier potassium channels in Lec mutant chinese hamster ovary cell lines reveals a role for sialidation in channel function. J Biol Chem. 1996;271:19093–19098. doi: 10.1074/jbc.271.32.19093. [DOI] [PubMed] [Google Scholar]
  124. Traynelis SF, Wollmuth LP, McBain CJ, Menniti FS, Vance KM, Ogden KK, Hansen KB, Yuan H, Myers SJ, Dingledine R. Glutamate receptor ion channels: structure, regulation, and function. Pharmacol Rev. 2010;62:405–496. doi: 10.1124/pr.109.002451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Tyrrell L, Renganathan M, Dib-Hajj SD, Waxman SG. Glycosylation alters steady-state inactivation of sodium channel Nav1.9/NaN in dorsal root ganglion neurons and is developmentally regulated. J Neurosci. 2001;21:9629–9637. doi: 10.1523/JNEUROSCI.21-24-09629.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Ufret-Vincenty CA, Baro DJ, Lederer WJ, Rockman HA, Quinones LE, Santana LF. Role of sodium channel deglycosylation in the genesis of cardiac arrhythmias in heart failure. J Biol Chem. 2001;276:28197–28203. doi: 10.1074/jbc.M102548200. [DOI] [PubMed] [Google Scholar]
  127. Ufret-Vincenty CA, Baro DJ, Santana LF. Differential contribution of sialic acid to the function of repolarizing K(+) currents in ventricular myocytes. Am J Physiol Cell Physiol. 2001;281:C464–C474. doi: 10.1152/ajpcell.2001.281.2.C464. [DOI] [PubMed] [Google Scholar]
  128. Varki A, Etzler ME, Cummings RD, Esko JD. Discovery and classification of glycan-binding proteins. In: Varki A, Cummings RD, et al., editors. Essentials of Glycobiology. Cold Spring Harbor (NY): 2009. [PubMed] [Google Scholar]
  129. Watanabe I, Wang HG, Sutachan JJ, Zhu J, Recio-Pinto E, Thornhill WB. Glycosylation affects rat Kv1.1 potassium channel gating by a combined surface potential and cooperative subunit interaction mechanism. J Physiol. 2003;550:51–66. doi: 10.1113/jphysiol.2003.040337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Watanabe I, Zhu J, Recio-Pinto E, Thornhill WB. Glycosylation affects the protein stability and cell surface expression of Kv1.4 but Not Kv1.1 potassium channels. A pore region determinant dictates the effect of glycosylation on trafficking. J Biol Chem. 2004;279:8879–8885. doi: 10.1074/jbc.M309802200. [DOI] [PubMed] [Google Scholar]
  131. Watanabe I, Zhu J, Sutachan JJ, Gottschalk A, Recio-Pinto E, Thornhill WB. The glycosylation state of Kv1.2 potassium channels affects trafficking, gating, and simulated action potentials. Brain Res. 2007;1144:1–18. doi: 10.1016/j.brainres.2007.01.092. [DOI] [PubMed] [Google Scholar]
  132. Weinhold B, Seidenfaden R, Rockle I, Muhlenhoff M, Schertzinger F, Conzelmann S, Marth JD, Gerardy-Schahn R, Hildebrandt H. Genetic ablation of polysialic acid causes severe neurodevelopmental defects rescued by deletion of the neural cell adhesion molecule. J Biol Chem. 2005;280:42971–42977. doi: 10.1074/jbc.M511097200. [DOI] [PubMed] [Google Scholar]
  133. Weiss N, Black SA, Bladen C, Chen L, Zamponi GW. Surface expression and function of Cav3.2 T-type calcium channels are controlled by asparagine-linked glycosylation. Pflugers Arch. 2013;465:1159–1170. doi: 10.1007/s00424-013-1259-3. [DOI] [PubMed] [Google Scholar]
  134. Wirkner K, Hognestad H, Jahnel R, Hucho F, Illes P. Characterization of rat transient receptor potential vanilloid 1 receptors lacking the N-glycosylation site N604. Neuroreport. 2005;16:997–1001. doi: 10.1097/00001756-200506210-00023. [DOI] [PubMed] [Google Scholar]
  135. Wittwer AJ, Howard SC. Glycosylation at Asn-184 inhibits the conversion of single-chain to two-chain tissue-type plasminogen activator by plasmin. Biochemistry. 1990;29:4175–4180. doi: 10.1021/bi00469a021. [DOI] [PubMed] [Google Scholar]
  136. Woodard-Grice AV, McBrayer AC, Wakefield JK, Zhuo Y, Bellis SL. Proteolytic shedding of ST6Gal-I by BACE1 regulates the glycosylation and function of alpha4beta1 integrins. J Biol Chem. 2008;283:26364–26373. doi: 10.1074/jbc.M800836200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Wormald MR, Dwek RA. Glycoproteins: glycan presentation and protein-fold stability. Structure. 1999;7:R155–R160. doi: 10.1016/s0969-2126(99)80095-1. [DOI] [PubMed] [Google Scholar]
  138. Xu H, Fu Y, Tian W, Cohen DM. Glycosylation of the osmoresponsive transient receptor potential channel TRPV4 on Asn-651 influences membrane trafficking. Am J Physiol Renal Physiol. 2006;290:F1103–F1109. doi: 10.1152/ajprenal.00245.2005. [DOI] [PubMed] [Google Scholar]
  139. Yamamoto S, Oka S, Inoue M, Shimuta M, Manabe T, Takahashi H, Miyamoto M, Asano M, Sakagami J, Sudo K, et al. Mice deficient in nervous system-specific carbohydrate epitope HNK-1 exhibit impaired synaptic plasticity and spatial learning. J Biol Chem. 2002;277:27227–27231. doi: 10.1074/jbc.C200296200. [DOI] [PubMed] [Google Scholar]
  140. Ye Z, Marth JD. N-glycan branching requirement in neuronal and postnatal viability. Glycobiology. 2004;14:547–558. doi: 10.1093/glycob/cwh069. [DOI] [PubMed] [Google Scholar]
  141. Yoshihara T, Sugihara K, Kizuka Y, Oka S, Asano M. Learning/memory impairment and reduced expression of the HNK-1 carbohydrate in beta4-galactosyltransferase-II-deficient mice. J Biol Chem. 2009;284:12550–12561. doi: 10.1074/jbc.M809188200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Yue KT, MacDonald JF, Pekhletski R, Hampson DR. Differential effects of lectins on recombinant glutamate receptors. Eur J Pharmacol. 1995;291:229–235. doi: 10.1016/0922-4106(95)90062-4. [DOI] [PubMed] [Google Scholar]
  143. Zhang Y, Hartmann HA, Satin J. Glycosylation influences voltage-dependent gating of cardiac and skeletal muscle sodium channels. J Membr Biol. 1999;171:195–207. doi: 10.1007/s002329900571. [DOI] [PubMed] [Google Scholar]
  144. Zhao Y, Nakagawa T, Itoh S, Inamori K, Isaji T, Kariya Y, Kondo A, Miyoshi E, Miyazaki K, Kawasaki N, et al. N-acetylglucosaminyltransferase III antagonizes the effect of N-acetylglucosaminyltransferase V on alpha3beta1 integrin-mediated cell migration. J Biol Chem. 2006;281:32122–32130. doi: 10.1074/jbc.M607274200. [DOI] [PubMed] [Google Scholar]
  145. Zhu J, Recio-Pinto E, Hartwig T, Sellers W, Yan J, Thornhill WB. The Kv1.2 potassium channel: The position of an N-glycan on the extracellular linkers affects its protein expression and function. Brain Res. 2009;1251:16–29. doi: 10.1016/j.brainres.2008.11.033. [DOI] [PubMed] [Google Scholar]
  146. Zhu J, Yan J, Thornhill WB. N-glycosylation promotes the cell surface expression of Kv1.3 potassium channels. FEBS J. 2012;279:2632–2644. doi: 10.1111/j.1742-4658.2012.08642.x. [DOI] [PubMed] [Google Scholar]
  147. Zhuo Y, Bellis SL. Emerging role of alpha2,6-sialic acid as a negative regulator of galectin binding and function. J Biol Chem. 2011;286:5935–5941. doi: 10.1074/jbc.R110.191429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Zuber C, Lackie PM, Catterall WA, Roth J. Polysialic acid is associated with sodium channels and the neural cell adhesion molecule N-CAM in adult rat brain. J Biol Chem. 1992;267:9965–9971. [PubMed] [Google Scholar]

Articles from Glycobiology are provided here courtesy of Oxford University Press

RESOURCES