1,290
Views
37
CrossRef citations to date
0
Altmetric
Original

An overview of trafficking and assembly of neurotransmitter receptors and ion channels (Review)

Pages 270-278 | Received 31 Oct 2007, Published online: 09 Jul 2009

Abstract

Ionotropic neurotransmitter receptors and voltage-gated ion channels assemble from several homologous and non-homologous subunits. Assembly of these multimeric membrane proteins is a tightly controlled process subject to primary and secondary quality control mechanisms. An assembly pathway involving a dimerization of dimers has been demonstrated for a voltage-gated potassium channel and for different types of glutamate receptors. While many novel C-terminal assembly domains have been identified in various members of the voltage-gated cation channel superfamily, the assembly pathways followed by these proteins remain largely elusive. Recent progress on the recognition of polar residues in the transmembrane segments of membrane proteins by the retrieval factor Rer1 is likely to be relevant for the further investigation of trafficking defects in channelopathies. This mechanism might also contribute to controlling the assembly of ion channels by retrieving unassembled subunits to the endoplasmic reticulum. The endoplasmic reticulum is a metabolic compartment studded with small molecule transporters. This environment provides ligands that have recently been shown to act as pharmacological chaperones in the biogenesis of ligand-gated ion channels. Future progress depends on the improvement of tools, in particular the antibodies used by the field, and the continued exploitation of genetically tractable model organisms in screens and physiological experiments.

Introduction

Ion channels provide an ion-conducting pore in the hydrophobic environment of the lipid bilayer Citation[1]. The pore can be gated by different molecular events, e.g., the binding of a neurotransmitter ligand or a change in transmembrane potential. Ion channels are polytopic membrane proteins and in many cases complex heteromultimers. The species of heteromultimers present at the cell surface reflect many different (however not all) conceivable combinations of subunits, each arrangement conferring different functional properties. In the fully assembled channel complex, the pore is shielded from the lipid bilayer by the surrounding protein. However, during co-translational integration of the protein into the membrane or before completion of multimeric assembly of the channel, polar residues eventually contributing to the pore may be exposed. It has been shown that exposed polar residues in transmembrane segments can lead to the endoplasmic reticulum-associated degradation (ERAD) of membrane proteins Citation[2]. The existence of energetically unfavourable assembly intermediates and the need to select only some of the possible multimers means that ion channels are demanding substrates of the secretory pathway: in addition to primary quality control processes at the endoplasmic reticulum (ER) they are subject to protein- or cell type-specific secondary quality control mechanisms Citation[3]. Many have to pass trafficking checkpoints before leaving the ER-to-Golgi shuttle Citation[3–5]. Ion channels are particularly interesting cargo proteins of the vesicular transport machinery because their numbers at the cell surface or in subcellular compartments are tightly controlled and sometimes regulated Citation[6–13].

Complex heteromultimers need sophisticated quality control

Like membrane proteins of the immune system, neurotransmitter receptors and ion channels require particularly stringent quality control mechanisms Citation[3] that ensure proper protein folding, subunit assembly, and functionality of the multimeric complex. All biophysical parameters of an ion channel can be influenced by events that occur early in the life of the protein. For instance, combination of alternative AMPA receptor subunits by a variety of mechanisms – e.g., differential expression of homologous and auxiliary subunits, differential splicing, and RNA-editing – gives rise to receptors of different ion selectivity or results in different kinetic properties Citation[14–19]. In the case of ionotropic GABAA receptors, in addition to the subunit composition, the positioning of subunits within the pentameric complex affects the pharmacology of the receptor Citation[20], Citation[21].

Classic approaches in this field combining molecular biology, biochemistry, and electrophysiology (mostly in heterologous expression systems like Xenopus oocytes or immortalized cell lines) still produce interesting results. Typical research in this area asks questions such as: which subunit combinations can reach the cell surface and what are their functional properties? Which domains drive the assembly of a given (hetero)multimer? How does the interaction between multimerization domains confer specificity to the assembly process? Which trafficking determinants contribute to cell surface expression or its regulation? At the same time, there are problems that the field often neglects because they are technically less tractable: how available are alternative subunits for assembly when one considers the timing of gene expression and spatial aspects of protein biosynthesis? Do the non-translated regions in the relevant messenger RNAs target translation of a given subunit to a specific subcellular localization or (a putative) assembly domain of the ER? Are concentrating mechanisms required to make assembly efficient under physiological conditions, e.g., when only small amounts of the assembling subunits are expressed?

This overview to the thematic issue of trafficking, assembly, and regulation of neurotransmitter receptors and ion channels attempts to summarize progress and emerging themes across the superfamilies of ligand-gated and voltage-gated channels that are addressed in the individual reviews.

Assembly is at minimum a bimolecular reaction

Since by definition assembly is at least a bimolecular reaction, the process depends on the concentration of the co-assembling subunits. In the case of neurotransmitter receptors and ion channels, these parameters are usually unknown. Our knowledge of ion channel assembly depends largely on experiments employing in-vitro translation or heterologous expression (in Xenopus oocytes or tissue culture cells). This means that key results have been obtained under vast over-expression conditions. Whilst many conclusions, e.g., about assembly domains and assembly intermediates should not be affected, these experimental conditions may have skewed our picture of the kinetic aspects of assembly Citation[22–30]. In principle, assembly could be a reaction of higher molecular order since many ion channels and ionotropic receptors are trimers, tetramers, or pentamers. Based on their studies of Kv1.3 assembly, Tu and Deutsch have proposed an assembly pathway involving a dimerization of dimers Citation[30] – sequential bimolecular reactions resulting in the final tetrameric assembly.

This sequential dimerization concept has proven useful for understanding the biogenesis of tetrameric AMPA receptors Citation[31], Citation[32]. The dose-response curve obtained for the dominant negative effect of a chimeric subunit on AMPA receptor heteromeric channel assembly suggested dimeric assembly intermediates. This chimeric subunit had been engineered to be compatible only in the N-terminal assembly domain and not in other regions of the protein required for the formation of functional tetramers. The assembly process of AMPA receptors utilizing the chimeric subunit was stalled, presumably as a dimeric intermediate Citation[22]. Subsequent biochemical work Citation[15] was able to detect the proposed dimeric assembly intermediate directly by cross-linking and by blue native polyacrylamide gel electrophoresis of endogenous neuronal AMPA receptors.

A dimer of dimers is also apparent in the structure of a bacterial inward rectifier potassium channel Citation[33]. Thus, these channels might assemble from dimeric intermediates. However, the situation in cyclic-nucleotide gated (CNG) channels is an important reminder that it might be dangerous to deduce assembly pathways from the structure of the fully assembled complex: a low-resolution structure of the retinal cGMP-gated channel suggests arrangement of the four subunits as a pair of dimers Citation[34]. But beautiful biochemical work addressing the subunit stoichiometry of the same native rod CNG channel found that the channel is a tetramer composed of three A1 and one B1 subunit Citation[35], Citation[36]. Interestingly, Zhong et al. identified a trimer-forming multimerization domain that is the molecular determinant of a subunit stoichiometry of 3:1 for A and B type subunits in the assembled tetramer Citation[36], Citation[37]. Therefore, assembly of this channel is unlikely to involve a dimerization of dimers despite the tetrameric nature of the mature channel and the intriguing two-fold rotational symmetry of the intracellular domains observed at low resolution Citation[34]. In conclusion, the symmetry of multimers might change during assembly and a proposal for an assembly pathway is more plausible if based on the detection of assembly intermediates.

Shaker-type voltage-gated potassium channels like Kv1.3 are capable of forming a tetramer via their N-terminal T1 domains (a) while the protein is still being translated and translocated Citation[25], Citation[27]. However, this conclusion is based on in-vitro translation in rabbit reticulocyte lysate. It is unknown if coupled tertiary folding and oligomerization of endogenous Kv channels occurs under physiologically relevant conditions. The question is whether the concentration of nascent channel subunits would suffice to make an encounter during the initial steps of biosynthesis probable.

Figure 1.  N- and C-terminal assembly domains found in the superfamily of voltage-gated cation channels. (a) Schematic representation of a Kv pore-forming subunit. The N-terminal location of the T1 domain is indicated and the corresponding crystal structure obtained for four assembled T1 domains of Kv1.2 Citation[109] is shown. (b) Schematic representation of the KCNQ7 pore-forming subunit. The location of the C-terminal A-domain tail, a self-assembling, parallel, four-stranded coiled-coil, is shown along with the corresponding crystal structure of four assembled coiled-coil forming domains Citation[39]. Structure reproduced from Citation[109] and Citation[39] with permission from Elsevier.

Figure 1.  N- and C-terminal assembly domains found in the superfamily of voltage-gated cation channels. (a) Schematic representation of a Kv pore-forming subunit. The N-terminal location of the T1 domain is indicated and the corresponding crystal structure obtained for four assembled T1 domains of Kv1.2 Citation[109] is shown. (b) Schematic representation of the KCNQ7 pore-forming subunit. The location of the C-terminal A-domain tail, a self-assembling, parallel, four-stranded coiled-coil, is shown along with the corresponding crystal structure of four assembled coiled-coil forming domains Citation[39]. Structure reproduced from Citation[109] and Citation[39] with permission from Elsevier.

For many other channels of the voltage-gated cation channel superfamily including CNG channels Citation[36], Citation[37], ether á go-go (eag) channels Citation[23], intermediate conductance (IK) calcium-activated channels Citation[38], KCNQ channels Citation[24], Citation[26], Citation[28], Citation[39–41], and TRPM channels Citation[29], assembly domains that form coiled-coil structures are found in the C-terminus of the polypeptide chain (b). Based on the recent discovery of these C-terminal assembly domains Tsuruda et al. concluded: ‘In the case of a C-terminally located assembly domain, the temporal order of assembly must be very different’ Citation[29] from the assembly of Kv channels via the T1 domain (). This determination holds true if the physiological concentration of nascent subunits is of sufficient magnitude to allow for co-translational interactions between subunits presenting N-terminal assembly domains. Careful structural and biophysical characterization of assembly domains continues to be a prerequisite for further investigation of assembly pathways in the cellular environment.

Properly assembled – and functional!

Ellgaard and Helenius distinguish between primary (e.g., generic, a) and secondary (e.g., cargo-specific, b) quality control processes Citation[3]. Depending on their topology, channels and receptors interact with different subsets of the chaperones involved in primary quality control Citation[42–44].

Figure 2.  Major mechanisms involved in general and protein-specific quality control. Following the terminology introduced by Ellgaard and Helenius Citation[3] primary (a; closed triangles) and secondary (b; open triangles) quality control processes are shown. The assembling subunits of a heteromultimeric channel are shown as rectangular shapes. Large dashed arrows indicate forward transport and ER retrieval. (a) General chaperones assist the folding of channels with luminal domains and monitor core glycosylation Citation[3]. Assembly intermediates exposing polar residues in the plane of the membrane can be eliminated by ERAD Citation[2]. Alternatively, they can be prevented from leaving the ER-Golgi shuttle by the retrieval factor Rer1 Citation[45], Citation[46]. (b) Herrmann et al. divided protein-specific chaperones into outfitters, escorts and guides Citation[81]. Many auxiliary subunits of ion channels can perform these functions for the specific pore-forming subunit(s) that they interact with. Beyond quality control, they usually contribute to the functional properties of the channel at the cell surface. Adapted with permission from Citation[5].

Figure 2.  Major mechanisms involved in general and protein-specific quality control. Following the terminology introduced by Ellgaard and Helenius Citation[3] primary (a; closed triangles) and secondary (b; open triangles) quality control processes are shown. The assembling subunits of a heteromultimeric channel are shown as rectangular shapes. Large dashed arrows indicate forward transport and ER retrieval. (a) General chaperones assist the folding of channels with luminal domains and monitor core glycosylation Citation[3]. Assembly intermediates exposing polar residues in the plane of the membrane can be eliminated by ERAD Citation[2]. Alternatively, they can be prevented from leaving the ER-Golgi shuttle by the retrieval factor Rer1 Citation[45], Citation[46]. (b) Herrmann et al. divided protein-specific chaperones into outfitters, escorts and guides Citation[81]. Many auxiliary subunits of ion channels can perform these functions for the specific pore-forming subunit(s) that they interact with. Beyond quality control, they usually contribute to the functional properties of the channel at the cell surface. Adapted with permission from Citation[5].

An emerging new aspect of primary quality control is the recognition by the ER retrieval factor Rer1 of membrane-spanning segments that are energetically unstable in the lipid bi-layer (a). This type of quality control is highly relevant to multimeric receptors and ion channels because they may expose polar residues in the plane of the membrane until fully assembled. The mechanism was first studied in Saccharomyces cerevisiae where Sato et al. demonstrated that in the Golgi, the membrane protein Rer1 binds to cargo proteins via transmembrane segments containing polar residues and sorts them back to the ER in coatomer (COPI)-coated vesicles Citation[45], Citation[46]. The same group went on to show that recognition of a polar residue within a transmembrane segment by Rer1 and subsequent retrieval to the ER was also the basis for the retention of an unassembled iron transporter subunit from the cell surface Citation[45], Citation[46]. This work and studies addressing the role of mammalian Rer1 in the multimeric assembly of γ-secretase (a protease involved in intramembrane proteolysis) Citation[47], Citation[48] has raised an interesting possibility: Rer1 might be involved in the biogenesis of multimeric neurotransmitter receptors and ion channels since assembly intermediates of these complexes are likely to expose polar residues.

Insights into the function of Rer1 might also be relevant to the study of channelopathies where mutations in genes coding for ion channels cause human inherited disorders Citation[49], often because the mutated gene gives rise to a channel that cannot reach the cell surface. For example, in familial hyperinsulinism, mutations in the ABCC8 gene encoding the regulatory subunit of ATP-sensitive potassium (KATP) channels, SUR1, can cause trafficking defects where SUR1 is localized to the ER instead of the plasma membrane Citation[50–56]. Some of the relevant mutations introduce a polar residue into a transmembrane segment Citation[54]. Hence one may speculate that the mammalian homologue of Rer1 could be involved in the ER localization of these disease-associated channel variants.

Consistent with the idea that changes in the transmembrane domain might confer a dominant additional retrieval signal (e.g., a Rer1 recognition site) to SUR1, mutation of an arginine (R)-based ER localization motif in SUR1 did not suppress trafficking defects found for variants containing A116P in the third and V187D in the fifth transmembrane segment of SUR1 Citation[54]. However, the same manipulation fully rescued the cell surface expression of another disease-associated SUR1 variant, L1544P Citation[53]. Variants SUR1 A116P and V187D could hypothetically be recognized and retrieved by Rer1 whereas L1544 maps to the distal cytosolic C-terminus of SUR1. Thus, it might be worthwhile to experimentally test the hypothesis that mammalian Rer1 participates in the quality control of disease-associated channel variants. If true, Rer1 would have a role at the intersection of primary, ER-based quality control and protein sorting Citation[3].

The function of R-based ER localization motifs in assembly-dependent forward transport is a well-established example for the interplay between the cellular sorting and quality control machinery Citation[5], Citation[57–67]: the Golgi-localized COPI vesicle coat complex recognizes these peptide-sorting motifs presented by ion channel and neurotransmitter receptor subunits. COPI recognition of the ER localization motifs leads to the ER retrieval of unassembled subunits or partially assembled complexes. Steric masking and binding of 14-3-3 proteins can overcome ER retrieval leading to the cell surface expression of properly assembled complexes. The presence of several ER localization and 14-3-3 binding motifs in multimeric KATP and acid-sensitive potassium TASK channel complexes makes it difficult to propose a detailed mechanistic model for the relevant binding and masking events Citation[58], Citation[63], Citation[68]. In order to understand the interplay between COPI and 14-3-3 protein binding precisely, high-resolution structures for these channels will have to become available.

Both ER localization signals and forward trafficking information on ion channel subunits contribute to the rate with which the fully assembled complex can leave the ER and reach the cell surface Citation[4]. Current studies characterized di-acidic sorting motifs in the acid-sensitive potassium channel TASK-3 and the plant KAT1 potassium channel Citation[69], Citation[70]. It is plausible to think that these sorting motifs are recognized by the COPII vesicle coat involved in the budding of anterograde transport vesicles at the ER. It remains to be shown which cellular machinery recognizes other recently identified forward-trafficking determinants like a RSRYW motif found in the α-subunit of epithelial sodium channels Citation[71].

Unlike most other ion channels, ionotropic neurotransmitter receptors contain a large luminal domain that is stringently monitored to ensure proper folding of this crucial part of the protein. A number of recent reports demonstrate the correlation between the ability of glutamate receptors to leave the ER and their functionality, e.g., ability to bind glutamate and respond to it with the appropriate conformational changes Citation[14], Citation[15], Citation[72–75]. Specifically, mutations that affect both glutamate binding and desensitization kinetics result in receptor subunits that are competent to tetramerize but almost completely fail to pass a post-assembly trafficking checkpoint Citation[74]. This failure results in drastically reduced cell surface expression of the mutant variants although their functional properties can be analyzed in over-expressing systems. Priel et al. suggest that this intracellular retention of variants with an altered response to glutamate is irrespective of channel activity in the ER membrane by combining two mutations: one that blocks ion flow and one that does not desensitize after glutamate binding and causes ER retention Citation[74]. The authors conclude from this experiment that glutamate binding affects the structure of the luminal domain in a way that enables forward transport of the complex. Coupling between the functionally relevant conformational changes and ER exit is an elegant mechanism to prevent the appearance of improperly regulated receptors at the cell surface. (See a recent review by Greger et al. Citation[31] for an excellent discussion of functional quality control in glutamate receptors.)

If glutamate is required for receptor maturation it becomes important to ask whether glutamate is present in the lumen of the ER. The amino acid has been detected in the ER of magnocellular endocrine cells of the supraoptic nucleus in the rat hypothalamus employing an anti-glutamate antibody Citation[76]. Work on metabotropic glutamate receptors in the ER and nuclear envelope of striatal neurons suggests that glutamate can enter the ER via cystine-glutamate exchanger and other secondary active transporters Citation[77]. Gating motions of glutamate receptors in the ER Citation[78] might come as a surprise but are consistent with the picture of the ER as a metabolic compartment Citation[79] studded with transport proteins running on coupled gradients Citation[80]. It is now clear that this transport activity provides small molecules that perform the function of pharmacological chaperones to support the biogenesis of ligand-binding membrane proteins.

Dedicated chaperones or regulatory subunits?

In many cases it is difficult to assign subunits of channels and receptors that are not pore-forming subunits neatly to one category or the other – particularly if one further divides the factors involved in secondary quality control into categories such as outfitters, escorts, and guides Citation[81] to distinguish between chaperones that promote the folding of specific substrates, ER exit factors, and associating proteins that provide targeting information beyond the ER (b).

The well characterized effect of different auxiliary subunits (e.g., beta-subunits of the oxido-reductase fold Citation[82], Ca+ + -binding proteins of the KChIP family Citation[83], or the most recently discovered dipeptidyl aminopeptidase-like protein DPPX Citation[84]) on the cell surface expression of voltage-gated potassium (Kv) channels is a good example to illustrate the blurring of these concepts (without disputing their usefulness): acting as escorts, they promote forward transport in the secretory pathway Citation[84–87] and acting as guides they determine axonal targeting Citation[85], Citation[88]. Each type of auxiliary subunit can also modulate the functional properties of Kv channels at the cell surface. This suggests that trafficking competence and certain functional properties may be different aspects of the same structural states and is reminiscent of the correlation found between trafficking and glutamate binding or desensitization for glutamate receptors (compare above).

The Stargazin or transmembrane AMPA receptor regulatory protein (TARP) family is yet another example of a family of auxiliary subunits with multiple effects: these proteins act as outfitters, escorts, and guides and modify the function of AMPA receptors at the cell surface (reviewed in Citation[89], Citation[90]). Other interesting auxiliary proteins that have recently been identified include RIC-3 Citation[91–99], a membrane protein with cytosolic domains bearing features of a coiled-coil domain, and its homologues. This factor has been shown to enhance the cell surface expression and affect the functional properties of nicotinic acetylcholine and serotonine receptors. Elegant work performed by Eimer et al. in Caenorhabditis elegans Citation[100] identified the conserved UNC-50 protein as a membrane protein involved in the subtype-specific sorting of nicotinic receptors within the Golgi apparatus. Also, two specific β-subunits for members of the CLC family of chloride transport proteins have been identified Citation[101], Citation[102]: Barttin is a β-subunit that is required for the cell surface expression of CLC-Ka and −Kb in the kidney and the inner ear Citation[101] and Ostm1 is a highly glycosylated membrane protein that travels with ClC-7 to lysosomes where it stabilizes the chloride transport protein Citation[102]. The precise molecular mechanism by which any of these auxiliary proteins exert their function in protein trafficking is unclear.

Perspectives – better tools and the power of model organisms

Progress towards understanding the trafficking of neurotransmitter receptors and ion channels in a physiological setting is frequently complicated by the fact that these cargo proteins are difficult to detect and to manipulate in their native environment. This situation can be improved by at least two approaches. One has recently been taken by the NeuroMab facility at the University of California, Davis (supported by a grant from NINDS and NIMH) that pursues the goal ‘to generate high quality mAbs that are validated … and to make these validated NeuroMabs available to the research community’ Citation[103]. This move to end the ubiquitous use of ‘reagents of mass distraction’ Citation[103] can only be met with the greatest enthusiasm because only good antibodies will allow the field to move away from heterologous expression and over-expression into model organisms. It will require the context of the physiological system to understand the sequence and orchestration of the many protein-protein interactions underlying assembly, trafficking, scaffolding, regulation, and internalization of neurotransmitter receptors and ion channels. As there is redundancy in most of these processes, one major challenge is to understand and distinguish the physiological roles of closely related proteins Citation[104]. Elias et al. combined genetic engineering of the mouse with RNA interference methodology to investigate the role of closely related scaffolding proteins of the PSD-95-like membrane-associated guanylate kinase family (PSD-MAGUK) in the targeting of AMPA receptors Citation[104]. They combine these two complementary methods of controlling gene expression with detailed functional analyses and biochemical experiments to show that the two closely related MAGUKs PSD-95 and PSD-93 affect AMPA receptor localization in a synapse-specific and developmentally regulated manner. In addition to synaptic targeting, dendrite growth and local dendritic protein biosynthesis are areas that require physiologically integrated approaches. Using organotypic hippocampal slice and dissociated neuronal primary cultures Raab-Graham et al. discovered a link between dendritic translation and surface expression of the voltage-gated potassium channel Kv1.1 and synaptic activity Citation[105]. The presence and function of secretory organelles in dendrites is a particularly exciting frontier of cellular neuroscience: recent work in mammalian neurons Citation[106] and a revealing genetic screen in Drosophila melanogaster Citation[107] provide the first glimpses of how a spatially distinct secretory pathway dedicated to a specific dendrite might eventually contribute to neuronal plasticity. Another genetically tractable model organism, C. elegans, has allowed for the discovery of three novel auxiliary proteins involved in the biogenesis, post-ER trafficking, and scaffolding of nicotinic acetylcholine receptors, namely RIC-3, UNC-50, and LEV-10 Citation[96], Citation[100], Citation[108]. These examples illustrate convincingly that the classical model organisms continue to contribute original and groundbreaking insights into the trafficking of ionotropic neurotransmitter receptors and ion channels.

Acknowledgements

I thank all members of my group – past and present – for sharing excitement and frustrations whilst working on ion channel trafficking. Work in my laboratory was funded by the ZMBH (Universität Heidelberg) Deutsche Forschungsgemeinschaft, Landesstiftung Baden-Württemberg, EMBO Young Investigator Programme, Fonds der Chemischen Industrie, and is currently funded by The Wellcome Trust. I am indebted to Nikolai Braun, Thomas Mrowiec, and Volker Schmid for valuable discussions and very helpful comments on the manuscript.

References

  • Hille B. Ionic channels of excitable membranes2nd ed. Sinauer Associates Inc, Sunderland 1992
  • Bonifacino JS, Cosson P, Klausner RD. Colocalized transmembrane determinants for ER degradation and subunit assembly explain the intracellular fate of TCR chains. Cell 1990; 63: 503–513
  • Ellgaard L, Helenius A. Quality control in the endoplasmic reticulum. Nat Rev Mol Cell Biol 2003; 4: 181–191
  • Ma D, Jan LY. ER transport signals and trafficking of potassium channels and receptors. Curr Opin Neurobiol 2002; 12: 287–292
  • Michelsen K, Yuan H, Schwappach B. Hide and run. EMBO Rep 2005; 6: 717–722
  • Debonneville C, Flores SY, Kamynina E, Plant PJ, Tauxe C, Thomas MA, Munster C, Chraibi A, Pratt JH, Horisberger JD, Pearce D, Loffing J, Staub O. Phosphorylation of Nedd4-2 by Sgk1 regulates epithelial Na(+) channel cell surface expression. EMBO J 2001; 20: 7052–7059
  • Lai HC, Jan LY. The distribution and targeting of neuronal voltage-gated ion channels. Nat Rev Neurosci 2006; 7: 548–562
  • Lambers TT, Oancea E, de Groot T, Topala CN, Hoenderop JG, Bindels RJ. Extracellular pH dynamically controls cell surface delivery of functional TRPV5 channels. Mol Cell Biol 2007; 27: 1486–1494
  • Misonou H, Trimmer JS. Determinants of voltage-gated potassium channel surface expression and localization in Mammalian neurons. Crit Rev Biochem Mol Biol 2004; 39: 125–145
  • Shibasaki K, Nakahira K, Trimmer JS, Shibata R, Akita M, Watanabe S, Ikenaka K. Mossy fibre contact triggers the targeting of Kv4.2 potassium channels to dendrites and synapses in developing cerebellar granule neurons. J Neurochem 2004; 89: 897–907
  • Varga AW, Yuan LL, Anderson AE, Schrader LA, Wu GY, Gatchel JR, Johnston D, Sweatt JD. Calcium-calmodulin-dependent kinase II modulates Kv4.2 channel expression and upregulates neuronal A-type potassium currents. J Neurosci 2004; 24: 3643–3654
  • Viard P, Butcher AJ, Halet G, Davies A, Nurnberg B, Heblich F, Dolphin AC. PI3K promotes voltage-dependent calcium channel trafficking to the plasma membrane. Nat Neurosci 2004; 7: 939–946
  • Yang JW, Vacher H, Park KS, Clark E, Trimmer JS. Trafficking-dependent phosphorylation of Kv1.2 regulates voltage-gated potassium channel cell surface expression. Proc Natl Acad Sci USA 2007; 104: 20055–20060
  • Greger IH, Akamine P, Khatri L, Ziff EB. Developmentally regulated, combinatorial RNA processing modulates AMPA receptor biogenesis. Neuron 2006; 51: 85–97
  • Greger IH, Khatri L, Kong X, Ziff EB. AMPA receptor tetramerization is mediated by Q/R editing. Neuron 2003; 40: 763–774
  • Greger IH, Khatri L, Ziff EB. RNA editing at arg607 controls AMPA receptor exit from the endoplasmic reticulum. Neuron 2002; 34: 759–772
  • Priel A, Kolleker A, Ayalon G, Gillor M, Osten P, Stern-Bach Y. Stargazin reduces desensitization and slows deactivation of the AMPA-type glutamate receptors. J Neurosci 2005; 25: 2682–2686
  • Tomita S, Adesnik H, Sekiguchi M, Zhang W, Wada K, Howe JR, Nicoll RA, Bredt DS. Stargazin modulates AMPA receptor gating and trafficking by distinct domains. Nature 2005; 435: 1052–1058
  • Turetsky D, Garringer E, Patneau DK. Stargazin modulates native AMPA receptor functional properties by two distinct mechanisms. J Neurosci 2005; 25: 7438–7448
  • Boulineau N, Baur R, Minier F, Sigel E. Consequence of the presence of two different beta subunit isoforms in a GABA(A) receptor. J Neurochem 2005; 95: 1724–1731
  • Sigel E, Baur R, Boulineau N, Minier F. Impact of subunit positioning on GABAA receptor function. Biochem Soc Trans 2006; 34: 868–871
  • Ayalon G, Stern-Bach Y. Functional assembly of AMPA and kainate receptors is mediated by several discrete protein-protein interactions. Neuron 2001; 31: 103–113
  • Jenke M, Sanchez A, Monje F, Stuhmer W, Weseloh RM, Pardo LA. C-terminal domains implicated in the functional surface expression of potassium channels. EMBO J 2003; 22: 395–403
  • Kanki H, Kupershmidt S, Yang T, Wells S, Roden DM. A structural requirement for processing the cardiac K+ channel KCNQ1. J Biol Chem 2004; 279: 33976–33983
  • Lu J, Robinson JM, Edwards D, Deutsch C. T1-T1 interactions occur in ER membranes while nascent Kv peptides are still attached to ribosomes. Biochemistry 2001; 40: 10934–10946
  • Maljevic S, Lerche C, Seebohm G, Alekov AK, Busch AE, Lerche H. C-terminal interaction of KCNQ2 and KCNQ3 K+ channels. J Physiol 2003; 548: 353–360
  • Robinson JM, Deutsch C. Coupled tertiary folding and oligomerization of the T1 domain of Kv channels. Neuron 2005; 45: 223–232
  • Schwake M, Jentsch TJ, Friedrich T. A carboxy-terminal domain determines the subunit specificity of KCNQ K+ channel assembly. EMBO Rep 2003; 4: 76–81
  • Tsuruda PR, Julius D, Minor DL, Jr. Coiled coils direct assembly of a cold-activated TRP channel. Neuron 2006; 51: 201–212
  • Tu L, Deutsch C. Evidence for dimerization of dimers in K+ channel assembly. Biophys J 1999; 76: 2004–2017
  • Greger IH, Ziff EB, Penn AC. Molecular determinants of AMPA receptor subunit assembly. Trends Neurosci 2007; 30: 407–416
  • Stern-Bach Y. AMPA receptor activation; not a square dance. Neuron 2004; 41: 309–311
  • Kuo A, Gulbis JM, Antcliff JF, Rahman T, Lowe ED, Zimmer J, Cuthbertson J, Ashcroft FM, Ezaki T, Doyle DA. Crystal structure of the potassium channel KirBac1.1 in the closed state. Science 2003; 300: 1922–1926
  • Higgins MK, Weitz D, Warne T, Schertler GF, Kaupp UB. Molecular architecture of a retinal cGMP-gated channel: the arrangement of the cytoplasmic domains. EMBO J 2002; 21: 2087–2094
  • Weitz D, Ficek N, Kremmer E, Bauer PJ, Kaupp UB. Subunit stoichiometry of the CNG channel of rod photoreceptors. Neuron 2002; 36: 881–889
  • Zhong H, Molday LL, Molday RS, Yau KW. The heteromeric cyclic nucleotide-gated channel adopts a 3A:1B stoichiometry. Nature 2002; 420: 193–198
  • Zhong H, Lai J, Yau KW. Selective heteromeric assembly of cyclic nucleotide-gated channels. Proc Natl Acad Sci USA 2003; 100: 5509–5513
  • Syme CA, Hamilton KL, Jones HM, Gerlach AC, Giltinan L, Papworth GD, Watkins SC, Bradbury NA, Devor DC. Trafficking of the Ca2 + -activated K+ channel, hIK1, is dependent upon a C-terminal leucine zipper. J Biol Chem 2003; 278: 8476–8486
  • Howard RJ, Clark KA, Holton JM, Minor DL, Jr. Structural insight into KCNQ (Kv7) channel assembly and channelopathy. Neuron 2007; 53: 663–675
  • Schmitt N, Schwarz M, Peretz A, Abitbol I, Attali B, Pongs O. A recessive C-terminal Jervell and Lange-Nielsen mutation of the KCNQ1 channel impairs subunit assembly. EMBO J 2000; 19: 332–340
  • Schwake M, Athanasiadu D, Beimgraben C, Blanz J, Beck C, Jentsch TJ, Saftig P, Friedrich T. Structural determinants of M-type KCNQ (Kv7) K+ channel assembly. J Neurosci 2006; 26: 3757–3766
  • Fukata Y, Tzingounis AV, Trinidad JC, Fukata M, Burlingame AL, Nicoll RA, Bredt DS. Molecular constituents of neuronal AMPA receptors. J Cell Biol 2005; 169: 399–404
  • Khanna R, Lee EJ, Papazian DM. Transient calnexin interaction confers long-term stability on folded K+ channel protein in the ER. J Cell Sci 2004; 117: 2897–2908
  • Wang X, Venable J, LaPointe P, Hutt DM, Koulov AV, Coppinger J, Gurkan C, Kellner W, Matteson J, Plutner H, Riordan JR, Kelly JW, Yates JR, 3rd, Balch WE. Hsp90 cochaperone Aha1 downregulation rescues misfolding of CFTR in cystic fibrosis. Cell 2006; 127: 803–815
  • Sato K, Sato M, Nakano A. Rer1p, a retrieval receptor for ER membrane proteins, recognizes transmembrane domains in multiple modes. Mol Biol Cell 2003; 14: 3605–3616
  • Sato M, Sato K, Nakano A. Endoplasmic reticulum quality control of unassembled iron transporter depends on Rer1p-mediated retrieval from the golgi. Mol Biol Cell 2004; 15: 1417–1424
  • Kaether C, Scheuermann J, Fassler M, Zilow S, Shirotani K, Valkova C, Novak B, Kacmar S, Steiner H, Haass C. Endoplasmic reticulum retention of the gamma-secretase complex component Pen2 by Rer1. EMBO Rep 2007; 8: 743–748
  • Spasic D, Raemaekers T, Dillen K, Declerck I, Baert V, Serneels L, Fullekrug J, Annaert W. Rer1p competes with APH-1 for binding to nicastrin and regulates gamma-secretase complex assembly in the early secretory pathway. J Cell Biol 2007; 176: 629–640
  • Ashcroft FM. From molecule to malady. Nature 2006; 440: 440–447
  • Cartier EA, Conti LR, Vandenberg CA, Shyng SL. Defective trafficking and function of KATP channels caused by a sulfonylurea receptor 1 mutation associated with persistent hyperinsulinemic hypoglycemia of infancy. Proc Natl Acad Sci USA 2001; 98: 2882–2887
  • Crane A, Aguilar-Bryan L. Assembly, maturation, and turnover of KATP channel subunits. J Biol Chem 2003; 279: 9080–9090
  • Partridge CJ, Beech DJ, Sivaprasadarao A. Identification and pharmacological correction of a membrane trafficking defect associated with a mutation in the sulfonylurea receptor causing familial hyperinsulinism. J Biol Chem 2001; 276: 35947–35952
  • Taschenberger G, Mougey A, Shen S, Lester LB, LaFranchi S, Shyng SL. Identification of a familial hyperinsulinism-causing mutation in the sulfonylurea receptor 1 that prevents normal trafficking and function of KATP channels. J Biol Chem 2002; 277: 17139–17146
  • Yan F, Lin CW, Weisiger E, Cartier EA, Taschenberger G, Shyng SL. Sulfonylureas correct trafficking defects of ATP-sensitive potassium channels caused by mutations in the sulfonylurea receptor. J Biol Chem 2004; 279: 11096–11105
  • Yan FF, Casey J, Shyng SL. Sulfonylureas correct trafficking defects of disease-causing ATP-sensitive potassium channels by binding to the channel complex. J Biol Chem 2006; 281: 33403–33413
  • Yan FF, Lin YW, MacMullen C, Ganguly A, Stanley CA, Shyng SL. Congenital hyperinsulinism associated ABCC8 mutations that cause defective trafficking of ATP-sensitive K+ channels: identification and rescue. Diabetes 2007; 56: 2339–2348
  • Brock C, Boudier L, Maurel D, Blahos J, Pin JP. Assembly-dependent surface targeting of the heterodimeric GABAB Receptor is controlled by COPI but not 14-3-3. Mol Biol Cell 2005; 16: 5572–5578
  • Heusser K, Yuan H, Neagoe I, Tarasov AI, Ashcroft FM, Schwappach B. Scavenging of 14-3-3 proteins reveals their involvement in the cell-surface transport of ATP-sensitive K+ channels. J Cell Sci 2006; 119: 4353–4363
  • Margeta-Mitrovic M, Jan YN, Jan LY. A trafficking checkpoint controls GABA(B) receptor heterodimerization. Neuron 2000; 27: 97–106
  • Michelsen K, Mrowiec T, Duderstadt KE, Frey S, Minor DL, Mayer MP, Schwappach B. A multimeric membrane protein reveals 14-3-3 isoform specificity in forward transport in yeast. Traffic 2006; 7: 903–916
  • Michelsen K, Schmid V, Metz J, Heusser K, Liebel U, Schwede T, Spang A, Schwappach B. Novel cargo-binding site in the beta and delta subunits of coatomer. J Cell Biol 2007; 179: 209–217
  • Mrowiec T, Schwappach B. 14-3-3 proteins in membrane protein transport. Biol Chem 2006; 387: 1227–1236
  • O'Kelly I, Butler MH, Zilberberg N, Goldstein SA. Forward transport. 14-3-3 binding overcomes retention in endoplasmic reticulum by dibasic signals. Cell 2002; 111: 577–588
  • Shikano S, Coblitz B, Sun H, Li M. Genetic isolation of transport signals directing cell surface expression. Nat Cell Biol 2005; 7: 985–992
  • Shikano S, Coblitz B, Wu M, Li M. 14-3-3 proteins: regulation of endoplasmic reticulum localization and surface expression of membrane proteins. Trends Cell Biol 2006; 16: 370–375
  • Vivithanaporn P, Yan S, Swanson GT. Intracellular trafficking of KA2 kainate receptors mediated by interactions with coatomer protein complex I (COPI) and 14-3-3 chaperone systems. J Biol Chem 2006; 281: 15475–15484
  • Yuan H, Michelsen K, Schwappach B. 14-3-3 dimers probe the assembly status of multimeric membrane proteins. Curr Biol 2003; 13: 638–646
  • O'Kelly I, Goldstein SA. Forward transport of K(2.1)3.1: mediation by 14-3-3 and COPI, modulation by p11. Traffic 2008; 9: 72–78
  • Mikosch M, Hurst AC, Hertel B, Homann U. Diacidic motif is required for efficient transport of the K+ channel KAT1 to the plasma membrane. Plant Physiol 2006; 142: 923–930
  • Zuzarte M, Rinne S, Schlichthorl G, Schubert A, Daut J, Preisig-Muller R. A di-acidic sequence motif enhances the surface expression of the potassium channel TASK-3. Traffic 2007; 8: 1093–1100
  • Mueller GM, Kashlan OB, Bruns JB, Maarouf AB, Aridor M, Kleyman TR, Hughey RP. Epithelial sodium channel exit from the endoplasmic reticulum is regulated by a signal within the carboxyl cytoplasmic domain of the alpha subunit. J Biol Chem 2007; 282: 33475–33483
  • Grunwald ME, Kaplan JM. Mutations in the ligand-binding and pore domains control exit of glutamate receptors from the endoplasmic reticulum in C. elegans. Neuropharmacology 2003; 45: 768–776
  • 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
  • Priel A, Selak S, Lerma J, Stern-Bach Y. Block of kainate receptor desensitization uncovers a key trafficking checkpoint. Neuron 2006; 52: 1037–1046
  • Valluru L, Xu J, Zhu Y, Yan S, Contractor A, Swanson GT. Ligand binding is a critical requirement for plasma membrane expression of heteromeric kainate receptors. J Biol Chem 2005; 280: 6085–6093
  • Meeker RB, Swanson DJ, Hayward JN. Light and electron microscopic localization of glutamate immunoreactivity in the supraoptic nucleus of the rat hypothalamus. Neuroscience 1989; 33: 157–167
  • Jong YJ, Kumar V, Kingston AE, Romano C, O'Malley KL. Functional metabotropic glutamate receptors on nuclei from brain and primary cultured striatal neurons. Role of transporters in delivering ligand. J Biol Chem 2005; 280: 30469–30480
  • Greger IH, Esteban JA. AMPA receptor biogenesis and trafficking. Curr Opin Neurobiol 2007; 17: 289–297
  • Csala M, Banhegyi G, Benedetti A. Endoplasmic reticulum: a metabolic compartment. FEBS Lett 2006; 580: 2160–2165
  • Csala M, Marcolongo P, Lizak B, Senesi S, Margittai E, Fulceri R, Magyar JE, Benedetti A, Banhegyi G. Transport and transporters in the endoplasmic reticulum. Biochim Biophys Acta 2007; 1768: 1325–1341
  • Herrmann JM, Malkus P, Schekman R. Out of the ER – outfitters, escorts and guides. Trends Cell Biol 1999; 9: 5–7
  • Gulbis JM, Mann S, MacKinnon R. Structure of a voltage-dependent K+ channel beta subunit. Cell 1999; 97: 943–952
  • An WF, Bowlby MR, Betty M, Cao J, Ling HP, Mendoza G, Hinson JW, Mattsson KI, Strassle BW, Trimmer JS, Rhodes KJ. Modulation of A-type potassium channels by a family of calcium sensors. Nature 2000; 403: 553–556
  • Nadal MS, Ozaita A, Amarillo Y, Vega-Saenz de Miera E, Ma Y, Mo W, Goldberg EM, Misumi Y, Ikehara Y, Neubert TA, Rudy B. The CD26-related dipeptidyl aminopeptidase-like protein DPPX is a critical component of neuronal A-type K+ channels. Neuron 2003; 37: 449–461
  • Campomanes CR, Carroll KI, Manganas LN, Hershberger ME, Gong B, Antonucci DE, Rhodes KJ, Trimmer JS. Kv beta subunit oxidoreductase activity and Kv1 potassium channel trafficking. J Biol Chem 2002; 277: 8298–8305
  • Shi G, Nakahira K, Hammond S, Rhodes KJ, Schechter LE, Trimmer JS. Beta subunits promote K+ channel surface expression through effects early in biosynthesis. Neuron 1996; 16: 843–852
  • Shibata R, Misonou H, Campomanes CR, Anderson AE, Schrader LA, Doliveira LC, Carroll KI, Sweatt JD, Rhodes KJ, Trimmer JS. A fundamental role for KChIPs in determining the molecular properties and trafficking of Kv4.2 potassium channels. J Biol Chem 2003; 278: 36445–36454
  • Gu C, Jan YN, Jan LY. A conserved domain in axonal targeting of Kv1 (Shaker) voltage-gated potassium channels. Science 2003; 301: 646–649
  • Nicoll RA, Tomita S, Bredt DS. Auxiliary subunits assist AMPA-type glutamate receptors. Science 2006; 311: 1253–1256
  • Osten P, Stern-Bach Y. Learning from stargazin: the mouse, the phenotype and the unexpected. Curr Opin Neurobiol 2006; 16: 275–280
  • Ben-Ami HC, Yassin L, Farah H, Michaeli A, Eshel M, Treinin M. RIC-3 affects properties and quantity of nicotinic acetylcholine receptors via a mechanism that does not require the coiled-coil domains. J Biol Chem 2005; 280: 28053–28060
  • Castillo M, Mulet J, Gutierrez LM, Ortiz JA, Castelan F, Gerber S, Sala S, Sala F, Criado M. Dual role of the RIC-3 protein in trafficking of serotonin and nicotinic acetylcholine receptors. J Biol Chem 2005; 280: 27062–27068
  • Castillo M, Mulet J, Gutierrez LM, Ortiz JA, Castelan F, Gerber S, Sala S, Sala F, Criado M. Role of the RIC-3 protein in trafficking of serotonin and nicotinic acetylcholine receptors. J Mol Neurosci 2006; 30: 153–156
  • Cheng A, Bollan KA, Greenwood SM, Irving AJ, Connolly CN. Differential subcellular localization of RIC-3 isoforms and their role in determining 5-HT3 receptor composition. J Biol Chem 2007; 282: 26158–26166
  • Cheng A, McDonald NA, Connolly CN. Cell surface expression of 5-hydroxytryptamine type 3 receptors is promoted by RIC-3. J Biol Chem 2005; 280: 22502–22507
  • Halevi S, McKay J, Palfreyman M, Yassin L, Eshel M, Jorgensen E, Treinin M. The C. elegans ric-3 gene is required for maturation of nicotinic acetylcholine receptors. Embo J 2002; 21: 1012–1020
  • Halevi S, Yassin L, Eshel M, Sala F, Sala S, Criado M, Treinin M. Conservation within the RIC-3 gene family. Effectors of mammalian nicotinic acetylcholine receptor expression. J Biol Chem 2003; 278: 34411–34417
  • Lansdell SJ, Gee VJ, Harkness PC, Doward AI, Baker ER, Gibb AJ, Millar NS. RIC-3 enhances functional expression of multiple nicotinic acetylcholine receptor subtypes in mammalian cells. Mol Pharmacol 2005; 68: 1431–1438
  • Williams ME, Burton B, Urrutia A, Shcherbatko A, Chavez-Noriega LE, Cohen CJ, Aiyar J. Ric-3 promotes functional expression of the nicotinic acetylcholine receptor alpha7 subunit in mammalian cells. J Biol Chem 2005; 280: 1257–1263
  • Eimer S, Gottschalk A, Hengartner M, Horvitz HR, Richmond J, Schafer WR, Bessereau JL. Regulation of nicotinic receptor trafficking by the transmembrane Golgi protein UNC-50. Embo J 2007; 26: 4313–4323
  • Estevez R, Boettger T, Stein V, Birkenhager R, Otto E, Hildebrandt F, Jentsch TJ. Barttin is a Cl- channel beta-subunit crucial for renal Cl- reabsorption and inner ear K+ secretion. Nature 2001; 414: 558–561
  • Lange PF, Wartosch L, Jentsch TJ, Fuhrmann JC. ClC-7 requires Ostm1 as a beta-subunit to support bone resorption and lysosomal function. Nature 2006; 440: 220–223
  • Rhodes KJ, Trimmer JS. Antibodies as valuable neuroscience research tools versus reagents of mass distraction. J Neurosci 2006; 26: 8017–8020
  • Elias GM, Funke L, Stein V, Grant SG, Bredt DS, Nicoll RA. Synapse-specific and developmentally regulated targeting of AMPA receptors by a family of MAGUK scaffolding proteins. Neuron 2006; 52: 307–320
  • Raab-Graham KF, Haddick PC, Jan YN, Jan LY. Activity- and mTOR-dependent suppression of Kv1.1 channel mRNA translation in dendrites. Science 2006; 314: 144–148
  • Horton AC, Racz B, Monson EE, Lin AL, Weinberg RJ, Ehlers MD. Polarized secretory trafficking directs cargo for asymmetric dendrite growth and morphogenesis. Neuron 2005; 48: 757–771
  • Ye B, Zhang Y, Song W, Younger SH, Jan LY, Jan YN. Growing dendrites and axons differ in their reliance on the secretory pathway. Cell 2007; 130: 717–729
  • Gally C, Eimer S, Richmond JE, Bessereau JL. A transmembrane protein required for acetylcholine receptor clustering in Caenorhabditis elegans. Nature 2004; 431: 578–582
  • Minor DL, Lin YF, Mobley BC, Avelar A, Jan YN, Jan LY, Berger JM. The polar T1 interface is linked to conformational changes that open the voltage-gated potassium channel. Cell 2000; 102: 657–670

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.