Dendritic trafficking for neuronal growth and plasticity Michael D. Ehlers*1

Biochemical Society Transactions

www.biochemsoctrans.org

*Neuroscience Research Unit, Pfizer Worldwide Research and Development, 700 Main Street, Cambridge, MA 02139, U.S.A.

Thudichum Medal Lecture Delivered at the University of Bath on 24 June 2013, as part of the 5th Conference on Advances in Molecular Mechanisms Underlying Neurological Disorders Focused Meeting Michael Ehlers

Abstract Among the largest cells in the body, neurons possess an immense surface area and intricate geometry that poses many unique cell biological challenges. This morphological complexity is critical for neural circuit formation and enables neurons to compartmentalize cell–cell communication and local intracellular signalling to a degree that surpasses other cell types. The adaptive plastic properties of neurons, synapses and circuits have been classically studied by measurement of electrophysiological properties, ionic conductances and excitability. Over the last 15 years, the field of synaptic and neural electrophysiology has collided with neuronal cell biology to produce a more integrated understanding Key words: dendrite, neuron, organelle, plasticity, transport. Abbreviations used: AMPA, α-amino-3-hydroxy-5-methylisoxazole-4-propionic acid; AP-2, adaptor protein 2; Arf1, ADP-ribosylation factor 1; BDNF, brain-derived neurotrophic factor; BFA, brefeldin A; CLIMP63, cytoskeleton-linking membrane protein of 63 kDa; CNS, central nervous system; DCV, dense-core vesicle; ER, endoplasmic reticulum; ERES, ER exit site; ERGIC, ER–Golgi intermediate compartment; EZ, endocytic zone; GA, Golgi apparatus; GABA, γ -aminobutyric acid; GluA1, glutamate A1 receptor; GluA2, glutamate A2 receptor; GluN2B, glutamate N2B receptor; GM130, cis-Golgi matrix protein of 130 kDa; GO, Golgi outpost; GPCR, G-protein-coupled receptor; GRASP65, Golgi reassembly stacking protein of 65 kDa; LDCV, large dense-core vesicle; LTD, longterm depression; LTP, long-term potentiation; mCh, mCherry; mGluR, metabotropic glutamate receptor; MVB, multivesicular body; MyoV, myosin V; Neep21, neuron-enriched endosomal protein of 21 kDa; NMDA, N-methyl-d-aspartate; PKC, protein kinase C; PKD, protein kinase D; PM, plasma membrane; PSD, postsynaptic density; Rab11-FIP2, Rab11 family-interacting protein 2; RE, recycling endosome; SEP, superecliptic pHluorin; SER, smooth ER; SN, substantia nigra; SNAP, synaptosome-associated protein; SNARE, soluble N-ethylmaleimide-sensitive fusion protein-attachment protein receptor; Stx4, syntaxin 4; Stx13, syntaxin 13; Syt4, synaptotagmin IV; TeNT, tetanus toxin; TfR, transferrin receptor; VAMP, vesicle-associated membrane protein; VTA, ventral tegmental area.. 1 email michael.ehlers@pfizer.com

Biochem. Soc. Trans. (2013) 41, 1365–1382; doi:10.1042/BST20130081

of how these remarkable highly differentiated cells utilize common eukaryotic cellular machinery to decode, integrate and propagate signals in the nervous system. The present article gives a very brief and personal overview of the organelles and trafficking machinery of neuronal dendrites and their role in dendritic and synaptic plasticity.

Thudichum Medal Lecture

Thudichum Medal Lecture

Introduction A major challenge for all cells is to traffic integral membrane proteins, secreted factors and lipids to appropriate subcellular domains in the right amount at the right time. With surface areas up to 10 000-fold greater than typical mammalian cells, and dendrites and axons that extend for hundreds of microns to metres from the cell body, neurons are at the pinnacle of complexity among metazoan cells. Pathways for membrane traffic originally emerged in simpler and smaller cells, and a central question in neuronal cell biology is how such evolutionarily conserved pathways became specialized to subserve the unique dimensions and segregated signalling that is the hallmark of neurons. Although intracellular organelles for membrane trafficking were observed in neurons more than a century ago [1], the mechanisms governing membrane transport to local dendritic domains and the role such transport plays in neuronal development, signalling, plasticity and morphology have emerged only recently. Most neurons possess highly branched dendrites contacted by hundreds of synapses. This complex geometry allows for spatial segregation and integration of diverse signals [2], and it is well established that the functional properties of dendrites are highly compartmentalized. Individual dendritic segments possess distinct molecular composition, synapatic inputs and electrophysiological properties [2–5]. The partitioning of dendrites into functionally distinct segments requires exquisite control over receptor and ion channel distribution [6,7], in large part by control over local cargo transport to and from the dendritic membrane. In the present article, I review a select set of work that has revealed the presence of specific organelles and trafficking machinery in neuronal dendrites that mediate dendritic growth and synapse plasticity. I  C The

Authors Journal compilation

 C 2013

Biochemical Society

1365

1366

Biochemical Society Transactions (2013) Volume 41, part 6

concentrate on the organelles of the early secretory pathway, endocytic pathway and exocytosis machinery that together regulate the molecular composition of dendritic domains and the function of synapses. I discuss the relationship between dendritic trafficking and forms of neuronal plasticity involved in learning, memory and brain development.

Early secretory trafficking in dendrites As in non-neuronal cells, the neuronal ER (endoplasmic reticulum) is distributed in the cytoplasm and extends throughout dendrites and the soma [8–11] where it functions both as an intracellular Ca2 + signalling compartment [12] and as the site of membrane protein and lipid synthesis [13]. Although SER (smooth ER) appears most abundant in dendrites, ER-bound ribosomes and translocon subunits are likewise found in distal dendrites [8,11,14,15], indicating the presence of RER (rough ER). The latter is significant, as local dendritic synthesis of membrane proteins such as neurotransmitter receptors has been proposed to mediate spatially delimited modification of dendritic segments [16– 18], but, unlike the case for cytosolic proteins, such local synthesis and secretory processing requires a local complement of secretory organelles. Progression of cargo through the secretory pathway requires proper folding of newly synthesized proteins in the ER and concomitant post-translational modifications, including N-glycosylation and disulfide bond formation. Following quality control checks, cargo concentrates at specialized ERESs (ER exit sites) [19] present on the ER membrane throughout the neuronal soma and dendrites [20,21]. Cargo emerges from the ER as COPII (coatamer protein II)-coated vesicles which then merge with ERGICs (ER–Golgi intermediate compartments) [22] that are also widely distributed throughout dendrites [9,10,23]. ER exit is typically the rate-limiting step for the biosynthesis of integral membrane proteins, including critical postsynaptic neurotransmitter receptors such as AMPA (α-amino-3-hydroxy-5methylisoxazole-4-propionic acid)-type glutamate receptors [24,25] and NMDA (N-methyl-D-aspartate)-type glutamate receptors, which contain ER retention and ER export signals that regulate surface expression [26–31]. Furthermore, by upregulating surface expression, chaperoning and ER exit has been proposed as a mechanism for the long-term action of drugs on their pharmacological targets, including the CNS (central nervous system) actions of nicotine on neuronal nicotinic acetylcholine receptors [32–35]. As secretory cargo progresses through the cis, medial and trans compartments of the GA (Golgi apparatus), further glycosylation and proteolysis occurs [36]. Such modification requires multiple enzymes whose localization and enzymatic activity in the GA is regulated by the low intraluminal pH of the Golgi lumen [37]. Indeed, loss of GA acidification disrupts GA morphology [38,39], impairs glycosylation and processing of lipids and protein [39–42], and alters cargo sorting [43,44]. Recent work has shown that the intraluminal pH of the GA is regulated by the ubiquitin ligase Ube3a [45],  C The

C 2013 Biochemical Society Authors Journal compilation 

whose expression in the brain is required for dendrite growth [46], spine formation [47,48] and synaptic plasticity [48–50]. Accordingly, brain-specific loss of Ube3a occurs in the severe neurodevelopmental disorder Angelman syndrome [51]. In the absence of Ube3a, neurons display dramatically distended Golgi cisternae because of marked underacidification of the GA [45]. Elevation of Golgi pH leads to cisternal swelling [39,52–55] and disrupted protein glycosylation [54,56–61]. Regarding the latter, the enzymatic activity, trans-Golgi accumulation and heteromeric assembly of sialyltransferases are highly pH-sensitive [59,61,62]. Correspondingly, surface sialylation is significantly reduced in Ube3a-deficient cells and in the brains of Angelman syndrome model mice [45]. Sialylation modulates interactions among proteins and lipids, thereby regulating cellular adhesion and cell signalling [63–65], including regulation of nervous system development and function [63,66–71]. Golgi ion homoeostasis and maintenance of pH-dependent functions of the neuronal GA, including sialylation, may thus contribute to dendritic growth and neuronal plasticity. Indeed, Golgi dysfunction and sialylation deficiency is associated with disorders of brain development and function including language impairment [72] and epilepsy [73]. In most mammalian cells, the ER, ERESs and ERGIC elements are distributed throughout the cell, whereas the GA is situated in the perinuclear area close to the MTOC (microtubule-organizing centre) [74]. This distribution of intracellular compartments necessitates an inward then outward directionality to pre- and post-Golgi trafficking respectively. This arrangement is quite different in neurons as the neuronal GA consists of both classic perinuclear Golgi stacks and discrete structures termed GOs (Golgi outposts) dispersed in select dendrites [9,10,20,75–79]. The monitoring of early secretory cargo emerging from the dendritic ER has shown that many post-ER carriers traffic from the dendrite to the somatic GA, whereas a subset of post-ER carriers utilize dendritic GOs [20,79]. Both modes of early secretory trafficking are biologically notable: the former for the long-range transport of post-ER carriers to reach the GA, and the latter for the use of non-canonical Golgi compartments separated spatially from the somatic GA (Figure 1). In dendrites, GOs have been defined by classic cisternal stack ultrastructure visible by EM as well as by immunoreactivity for markers such as GM130 (cis-Golgi matrix protein of 130 kDa) [9,20,77]. However, not all dendrites contain morphological or molecular markers of the GA [9,10,20,77]. In the absence of bona fide GOs, how is early secretory cargo trafficked and processed to reach the PM (plasma membrane)? One possibility is long-range transport of post-ER carriers to the soma [20]. Another possibility is that local dendritic trafficking of nascent secretory cargo bypasses classical Golgi [79], perhaps utilizing distinct post-ER carriers [80]. In PC12 cells, Rab1-containing ERGICs and ERESs, but not Golgi membranes, are selectively distributed to the neurites and growth cones of polarized cells [81], indicating that satellite secretory systems devoid of generic Golgi membranes

Dendritic trafficking for neuronal growth and plasticity

Figure 1 ER–Golgi trafficking in neuronal dendrites (A) Dendrites contain core machinery for early secretory trafficking including ER, ERESs, ERGICs and GOs. Whereas ER, ERESs and ERGICs are distributed throughout the somatodendritic compartment and in all dendrites, GOs localize to a subset of dendrites. As a consequence, post-ER carriers originating in dendrites lacking GOs must either be transported long distances back to the somatic Golgi or bypass the GA by utilizing ERGICs for subsequent processing and sorting. (B) Secretory compartments and biosynthesis in spines. Shown is a large mushroom spine containing ER with an associated spine apparatus. Vesicles may originate from the spine apparatus, although this has not been demonstrated directly. Ribosomes (r) and polysomes (p) are present at the base of spines and can be actively transported into spines, where they are thought to translate proteins that locally comprise the PSD. F-actin, filamentous actin.

are formed during neuritogenesis. One possibility is that dendrites utilize ERGIC rather than classical GA for cargo maturation before local PM delivery (Figure 1). Given the known impact of Golgi-associated glycosylation on ion channel distribution [82], stability [83] and biophysical properties [84], it is tempting to speculate that bypassing the GA may diversify the properties of dendritically synthesized neurotransmitter receptors and voltage-gated ion channels, and hence affect dendritic excitability. At this point, it is not known whether or how cargo carriers emerging from the dendritic ER are selected for long-range or local trafficking to subsequent secretory organelles. Before exiting the ER in discrete carriers, newly translated polypeptides reside in the lipid bilayer of the ER membrane. The ER itself exists as a continuous anastomosing network in dendrites [8,11,85,86]. For numerous synaptic receptors, ER export is rate-limiting for PM delivery, and multisubunit channels and receptors can dwell in the ER for various periods up to several hours [24,87–91]. This suggests that lateral diffusion within the ER determines the ultimate location of ER exit and hence the spatial range of new cargo transport to the PM [87]. Consistent with this notion, quantitative photobleaching, super-resolution imaging and ultrastructural analyses indicate that newly synthesized membrane proteins, including AMPA receptors, rapidly diffuse within the continuous network of the dendritic ER [11]. Nascent cargo diffusion is inhomogeneous in

dendrites, with zones of diffusional confinement corresponding to local regions of increased ER complexity at dendritic branch points, at dendritic spines, and near GOs [11] (Figure 2). In non-neuronal cells, ER morphology is controlled by diverse mechanisms including atlastins [92,93], reticulons [94] and Rab10 [95]. In dendrites, the spatial range of receptor lateral mobility in the ER becomes progressively limited over neuronal development and is rapidly restricted by group I mGluR (metabotropic glutamate receptor) signalling through a mechanism involving PKC (protein kinase C) and the microtubule-binding ER protein CLIMP63 (cytoskeletonlinking membrane protein of 63 kDa) [11] (Figure 2B). CLIMP63 is an ER integral membrane protein whose binding to microtubules is regulated by PKC-dependent phosphorylation [96–98]. When bound to microtubules, CLIMP63 promotes elongation of ER tubules to the cell periphery along microtubule tracts [96,98]. Upon mGluR– PKC activation, site-specific phosphorylation of one or more N-terminal serine residues disrupts CLIMP63 microtubule binding, and thus increases ER complexity, thereby trapping nascent membrane proteins within smaller domains of diffusional confinement. Manipulations that increase ER complexity promote AMPA receptor surface expression and increase synaptic strength [11], perhaps by directing a higher proportion of newly synthesized receptors for local forward trafficking. Together, these findings suggest tight coupling  C The

C 2013 Biochemical Society Authors Journal compilation 

1367

1368

Biochemical Society Transactions (2013) Volume 41, part 6

Figure 2 Local zones of ER complexity confine secretory cargo and direct dendrite growth (A) The ER exists as a continuous anastomosing network of tubules and sheets throughout dendrites. Nascent integral membrane secretory cargo diffuses laterally and rapidly within the lipid bilayer of the dendritic ER. The distance and length scales over which cargo diffuses is determined by the topological complexity of the ER, which is greatest at dendritic branch points and subjacent to large dendritic spines. (B) Geometric complexity of the dendritic ER is controlled by interaction of the ER-resident protein CLIMP63 with microtubules (MT) and this interaction is in turn disrupted by PKC phosphorylation of CLIMP63 stimulated upon activation of mGluRs. Upon activation of the mGluR/PKC/CLIMP63 pathway, the ER uncouples from microtubules, bunches up and adopts a more complex topology, thereby confining cargo to smaller length scales of effective diffusion in the dendrite. (C) Local zones of ER complexity predict sites of new dendritic branch formation and growth, probably due to localized spatial direction of new lipid and membrane cargo.

between the local exit of ER cargo and insertion at the PM.

Post-Golgi trafficking and asymmetric dendrite growth The secretory pathway is the primary site of lipid biosynthesis and membrane addition and thus contributes to polarized cell growth. In neurons, secretory trafficking is required for dendrite growth and maintenance. Disrupting Golgi function with BFA (brefeldin A) or by overexpressing inhibitory mutants of PKD (protein kinase D) or Arf1 (ADPribosylation factor 1) to block post- and intra-Golgi trafficking respectively blocks dendrite growth [77] (Figure 3). Moreover, such manipulations cause a dramatic simplification of established dendrites within hours, indicating a large-scale requirement for ongoing secretory trafficking in maintaining dendritic morphology [77]. In Drosophila, mutations in genes encoding the critical early secretory trafficking proteins Rab1, Sar1 and Sec23 were identified in a screen for dendritic morphology defects, and the dendritic defects seen in these mutants correlated with disrupted Golgi membranes [78]. In mammalian cortical and hippocampal pyramidal neurons, the somatic Golgi is oriented towards the large apical dendrite, and post-Golgi cargo flux is strongly biased towards the large apical dendrites [77]. Golgi polarity precedes dendrite polarity and predicts the formation of the longest dendrites [77,99]. Fragmentation and dispersion of the somatic Golgi by overexpressing GRASP65 (Golgi reassembly stacking protein of 65 kDa) prevents the specification of the apical dendrite without affecting overall total dendrite growth [77]. Notably, in the reeler mutant mouse lacking reelin expression, neurons in the cerebral cortex display misoriented apical dendrites. The secreted factor reelin promotes translocation  C The

C 2013 Biochemical Society Authors Journal compilation 

of the GA into developing apical dendrites via a mechanism requiring an LKB1/Stk25/GM130 signalling pathway [100], and the Cdc42/Rac1 guanine-nucleotide-exchange factor αPIX/Arhgef6 [101]. Thus polarized post-Golgi trafficking determines the generation of an apical dendrite and sustains asymmetric dendrite growth (Figure 3). Organelles of the secretory pathway concentrate at dendritic branch points. GOs are stably positioned at a subset of branch points, where they both receive and transmit secretory cargo [77]. In Drosophila sensory neurons, focal laser disruption of cellular domains containing GOs arrests local branch extension and retraction [78]. Even earlier in the secretory pathway, local zones of ER complexity and ER export localize to dendritic branch points, and such zones predict the site of new dendritic branch formation [11]. Manipulations that increase ER complexity promote dendritic branching, whereas simplification of the ER reduces dendritic branching [11], suggesting that local post-ER trafficking directs membrane cargo for dendrite formation and growth (Figures 2 and 3). Together, these observations indicate that dendrites contain satellite secretory systems that provide an elegant means by which cargo can be directed towards one dendritic branch or another. By spatially targeting the addition of lipid membrane and integral membrane proteins, such distributed biosynthetic machinery allows neurons to locally regulate dendritic morphology and molecular composition, ensuring compartmentalized function of dendritic segments.

Exocytosis at the dendritic PM Following processing in the secretory pathway, intracellular cargo must be exocytosed at the dendritic PM to reach the

Dendritic trafficking for neuronal growth and plasticity

Figure 3 Asymmetric dendritic growth is controlled by polarized secretory trafficking Pyramidal neurons possess a single apical dendrite whose formation and growth requires polarization of the somatic Golgi, imposing a directional bias to post-Golgi traffic. Signalling by factors such as reelin regulates the polarized distribution of the GA and subsequent asymmetric growth. Disruption of pre- or post-Golgi trafficking by manipulations including expression of dn (dominant-negative) PKD, dn Arf1, RNAi knockdown of the ER-associated GTPase Sar1 or BFA prevents dendritic growth. Dispersion and fragmentation of the somatic Golgi by manipulations including overexpression of the Golgi matrix protein GRASP65 prevents asymmetric growth of a single apical dendrite without affecting overall growth of the dendritic arbour.

surface of dendrites. Fusion of intracellular membrane stores with the PM governs cell-surface composition, regulates cell shape and allows the release of soluble factors. Whereas constitutive exocytosis maintains the complement of PM proteins and lipids, many forms of exocytosis are regulated by molecular or electrical stimuli. The most intensely studied form of regulated exocytosis is depolarizationinduced neurotransmitter release at presynaptic terminals, for which the molecular machinery for vesicle docking and fusion has been delineated in exquisite detail [102,103]. From the number and positioning of voltage-gated Ca2 + channels [104,105], to the full molecular composition of synaptic vesicles [106], to the biophysics and dynamics of SNARE (soluble N-ethylmaleimide-sensitive fusion proteinattachment protein receptor)-dependent fusion [107], to the Ca2 + sensing of synaptotagmins [108], few subcellular events in all of cell biology have been defined in as much detail. In contrast, much less is known about exocytosis on the other side of the synapse, and it is only in the last 15 years or so that it has been recognized that exocytosis occurs from postsynaptic compartments. It has become increasingly clear that specific forms of dendritic exocytosis regulate synapse function, plasticity and neuronal morphology (Figure 4).

Dendritic release of neurotransmitters and neuropeptides EM studies of several neuron types have indicated the presence of both classic and dense core secretory vesicles in dendrites that contain glutamate, GABA (γ -aminobutyric acid), dopamine and neuropeptides. In size, shape and

clustered distribution at presumptive fusion sites, these vesicles bear close resemblance to presynaptic vesicles [109– 115]. In brain areas such as the olfactory bulb, dendritic release sites are in contact with other dendrites that likewise contain clustered vesicles for neurotransmitter release. The reciprocal nature of these connections has been confirmed by electrophysiological analysis, demonstrating functional release of both excitatory and inhibitory neurotransmitters at dendrodendritic synapses, with a classic example being the synapse between granule and mitral cell dendrites in the main olfactory bulb [116–119]. In the olfactory bulb, dendritic release of GABA from granule cells requires intracellular Ca2 + , similar to release from axons and suggesting the presence of Ca2 + - regulated exocytic release machinery [116,120]. Whereas granule cell dendrites release GABA, contacting mitral cell dendrites release glutamate. Immunolabelling studies have shown that glutamatergic postsynaptic scaffolding proteins PSD (postsynaptic density)-93 and PSD-95 are present at mitral– granule cell synapses [121]. Notably, activation of NMDA receptors is sufficient to trigger dendritic GABA release from granule cells [122–124] by coupling depolarizing NMDA receptor-induced current to activation of voltage-gated Nor P/Q-type Ca2 + channels [120]. Much like in presynaptic terminals, fast Ca2 + chelators [e.g. BAPTA (1,2-bis-(oaminophenoxy)ethane-N,N,N ,N -tetra-acetic acid)] block GABA release from granule neurons, whereas slow Ca2 + chelators (e.g. EGTA) do not [120], consistent with tight coupling of Ca2 + entry to vesicle release machinery [125]. Dendrodendritic synapses represent an exception to the  C The

C 2013 Biochemical Society Authors Journal compilation 

1369

1370

Biochemical Society Transactions (2013) Volume 41, part 6

Figure 4 Organelles for exocytosis in dendrites Model of a neuron showing exocytosis occurring from both dendrites and axons. Dendrites exocytose soluble factors including neurotransmitters, neuropeptides and neurotrophins, as well as transmembrane proteins including ion channels and neurotransmitter receptors.

classic features of neuronal polarity, but highlight the ability of dendrites to exhibit highly regulated spatially restricted exocytosis. In midbrain dopamine neurons of the SN (substantia nigra) and VTA (ventral tegmental area), axonal release of dopamine occurs in the projection fields of the striatum, whereas release from dendrites occurs locally in the SN and VTA. Initial experiments measured [3 H]dopamine release in the SN after KCl-induced depolarization [126]. Microdialysis and voltammetry measurements in the SN and VTA confirmed local midbrain release of dopamine, presumably from dendrites [127–129]. Dopamine is also released from dissociated dopamine neurons in primary culture [130] and exogenous expression of VMAT2 (vesicular monoamine transporter 2) in cultured hippocampal neurons permits dendritic dopamine release from these non-dopaminergic cells [131]. The precise mechanism of dendritic dopamine release has not yet been fully elucidated. Some experiments have suggested that dendritic release occurs by reversal of the dopamine transporter rather than vesicular fusion [132]. However, other studies have found that SNARE proteindependent membrane fusion is required for exocytosis of dendritic dopamine [130,133,134]. As release from axons and dendrites occurs on to different neurons in different brain areas, these modes of neurotransmitter release have  C The

C 2013 Biochemical Society Authors Journal compilation 

very different functions, and may reflect release from different organelles. Indeed, dopamine has been observed in LDCVs (large dense-core vesicles), small synaptic vesicles and tubulovesicular structures resembling SER or endosomes [135,136]. Many neurons contain LDCVs that store neuropeptides which are released from dendrites. A well-studied example comes from magnocellular neurons of the supraoptic nucleus of the hypothalamus. These neurons have dendrites situated in and receiving input from the CNS, whereas their axons project into the peripheral hypophyseal portal circulation outside the blood–brain barrier. This unique anatomical arrangement means that axonal exocytosis confers peripheral release of neuropeptide which, for oxytocin, mediates uterine contractions and milk ejection associated with reproductive physiology, whereas central release from dendrites mediates adaptive social behaviours associated with reproductive and affiliative states. Presumptive fusion of LDCVs at the dendritic membrane of magnocellular neurons has been directly observed by EM [137]. The dendritic release of oxytocin is Ca2 + -dependent and can be triggered by NMDA receptor activation in the absence of action potentials [138]. In granule neurons of the hippocampus, dynorphin peptides are exocytosed from dendrites where they retrogradely inhibit neurotransmitter release from perforant path terminals through presynaptic κ-opioid receptors [139]. Whereas axonal release of dynorphin is blocked by inhibitors of N-type Ca2 + channels, dynorphin-mediated depression at perforant path synapses is blocked by antagonists of both Ntype and L-type Ca2 + channels [140], once again suggesting the presence of distinct dendritic release machinery that responds to Ca2 + from different sources to mediate exocytic release.

Retrograde synaptic signalling by dendritic exocytosis One of the most obvious functions of dendritic exocytosis is retrograde signalling to presynaptic terminals. Among such retrograde signals are released growth factors and neurotrophins including BDNF (brain-derived neurotrophic factor). Secreted BDNF controls a variety of cellular functions including synaptic transmission, synapse differentiation, morphological plasticity and gene transcription by binding and activating TrkB (tropomyosin receptor kinase B) receptors [141–145]. In addition to its well-known release from axon terminals [146–148], it has become clear that BDNF is also released from dendrites in an activitydependent manner. When expressed in cultured hippocampal neurons, GFP-tagged BDNF localizes to punctate vesicular structures in dendrites [149–153]. Following electrical stimulation or K + -induced depolarization, BDNF–GFP puncta disappear over several seconds [150]. Dendritic BDNF–GFP vesicle fusion requires CaMKIIα (Ca2 + /calmodulindependent protein kinase type IIα) [151] and occurs upon dendritic depolarization by back-propagating action potentials [152]. The specific vesicular structures containing BDNF are not well-defined, and the degree to which the

Dendritic trafficking for neuronal growth and plasticity

sorting of endogenous BDNF differs from heterologous expressed BDNF–GFP remains an area in need of additional study. DCVs (dense-core vesicles) are thought to harbour BDNF at presynaptic terminals, but ultrastructural analysis reveals only sparse DCVs in the dendrites of many CNS neurons thought to release BDNF [85,150,152]. An alternative possibility is that BDNF is directly secreted to the PM in post-Golgi carriers or through a Golgi–endosome pathway. In presynaptic terminals, synaptic vesicle fusion is triggered by the influx of Ca2 + ions, which bind to C2 domains on synaptotagmin 1, resulting in conformational rearrangement of core SNARE machinery that directly couples elevated Ca2 + to SNARE-mediated exocytosis [108]. A distinct synaptotagmin family member, Syt4 (synaptotagmin IV), regulates activity-dependent release of BDNF from cultured hippocampal neurons and localizes to BDNF-containing vesicles in dendrites [149]. Exocytosis of pHluroin-tagged Syt4 was visualized directly in hippocampal neuron dendrites and increased significantly upon depolarization [149]. Neurons from Syt4-knockout mice display increased BDNF release relative to wildtype neurons, suggesting, perhaps counterintuitively, that Syt4 plays a negative role in postsynaptic exocytosis. Syt4 also regulates retrograde signal-mediated plasticity at the Drosophila neuromuscular synapse. Specifically, highfrequency stimulation of muscle cells increases the release probability of contacting presynaptic boutons. In the absence of Syt4, this form of retrograde plasticity is absent, but can be rescued by expression of Syt4 in muscle, suggesting that Ca2 + influx is coupled to postsynaptic vesicular trafficking via Syt4 [154,155]. Syt4 may thus form part of an evolutionarily conserved machinery for postsynaptic exocytosis.

Exocytosis in dendrite growth and morphological plasticity In addition to the release of neurotransmitters and neuropeptides for interneuronal signalling, exocytosis contributes to the growth and morphogenesis of elaborate dendritic arbours. Originating as small roughly spherical cells upon terminal asymmetric division of neuronal precursors, neurons increase their total membrane area by many orders of magnitude, a process requiring huge amounts of membrane synthesis directed to growing dendrites and axons. As discussed above, disruption of the ER–Golgi secretory pathway in developing neurons prevents dendritic outgrowth in both mammals and fruitflies [77,78] (Figure 3). Interestingly, TeNT (tetanus toxin), a protease that cleaves most VAMPs (vesicle-associated membrane proteins), has limited impact on dendritic arbour development when added to cultured hippocampal neurons, even for days or weeks [156]. Moreover, neurons from animals lacking VAMP2 or SNAP-25 [25 kDa SNAP (synaptosome-associated protein)] have normal morphology [157,158], raising the question of which SNARE proteins are mediating membrane fusion that contributes to dendrite growth. One possibility is that growth-related membrane fusion occurs via the

toxin-insensitive VAMP family member Ti-VAMP (TeNTinsensitive VAMP)/VAMP7, which is required for neurite outgrowth in differentiating PC12 cells [159,160]. Although it is clear that local zones of ER complexity and ER export predict the site of new dendritic branch formation and areas of dendritic exocytosis [11], it is not yet known how spatial positioning of these early secretory compartments is coupled to specific exocytic machinery at the dendritic PM. It is now clear that many forms of synaptic plasticity require exocytosis at or near the postsynaptic membrane (Figures 5 and 6). Initial studies found that postsynaptic infusion of various agents that disrupt SNARE-mediated membrane fusion via a recording pipette rapidly block LTP (long-term potentiation) at Schaffer collateral-CA1 synapses [161]. Emerging from such observations has been a prevailing model where intradendritic vesicles harbouring AMPA receptors fuse with the PM upon LTP induction, and this increase in AMPA receptor number in turn increases synaptic strength [162,163]. In addition to such functional plasticity, postsynaptic exocytosis mediates morphological plasticity at glutamatergic synapses [164–168]. Specifically, dendritic spines increase their volume ∼2-fold following stimuli that induce LTP [3]. A variety of manipulations that prevent SNARE-mediated membrane fusion prevent LTPinduced spine growth [165,167,168], suggesting that spine exocytic events couple functional plasticity (more AMPA receptors), structural plasticity (more membrane and other molecules) and retrograde signalling (released factors).

Membrane compartments for dendritic exocytosis What is the identity of intracellular membrane stores involved in postsynaptic vesicle fusion? Serial EM reconstruction studies have demonstrated the presence of numerous membranebound organelles, including REs (recycling endosomes), in dendrites and dendritic spines [85,167]. These studies, together with the early observations that AMPA receptors are endocytosed and reinserted upon synaptic activation [169], suggested dendritic recycling endosomes as the primary internal membrane compartment mobilized to the PM in response to LTP-inducing stimuli [169–172] (Figures 5 and 7). Indeed, blocking RE transport using dominant-negative versions of Stx13 (syntaxin 13) and the Eps15 homology domain protein Rme1/EHD1 blocked LTP, providing direct evidence for RE involvement in synaptic plasticity [166]. A noteworthy early observation was that stimuli that trigger synaptic potentiation not only increase the surface levels of AMPA receptors, but also increase the release of previously endocytosed transferrin, the classic recycling endosome cargo [166]. This observation suggests that LTPstimulated membrane fusion is not selective for specific cargo, such as specific subtypes of AMPA receptors, but instead may act through conserved endosomal machinery independent of receptor content present on the exocytosing vesicle. Such a model is consistent with recent observations identifying synaptic potentiation in response to LTPinducing stimuli independent of specific receptor subtype  C The

C 2013 Biochemical Society Authors Journal compilation 

1371

1372

Biochemical Society Transactions (2013) Volume 41, part 6

Figure 5 Spine machinery for local exocytosis and endocytosis Depicted are postsynaptic trafficking events mediating synaptic plasticity. The exocytic domain (XD) is demarcated by clusters of the t-SNARE (target SNARE) Stx4 where REs fuse with the spine membrane. As with the XD, the EZ is positioned just lateral to the PSD, and contains a stable assembly of clathrin attached to the PSD through interactions involving Homer and dynamin-3. The EZ functions to recapture cargo diffusing laterally in the spine membrane for transport into a recycling pool required for synaptic potentiation. Endocytosis within spines is enhanced by up-regulation of Arc/Arg3.1 which binds the endocytic adaptor endophilin-2 and stimulates endocytosis. Spine endosomes are transported into or stabilized within spines through Ca2 + -dependent association of the actin-based motor MyoVb.

[173]. The promiscuity of LTP with respect to receptor subtypes suggests that signalling pathways may act on conserved cellular transport and exocytosis machinery rather than specific cargo molecules. The first optical demonstration of activity-triggered exocytosis in dendrites relied on the styryl dye FM1-43 and showed destaining of FM1-43-labelled compartments in minutes upon neuronal stimulation [174,175]. More recently, exocytosis of cargo tagged with the pH-sensitive SEP (superecliptic pHluorin) has been visualized directly. SEP-labelled AMPA receptors have been used in several studies to visualize postsynaptic exocytosis [164,165,176– 183]. To selectively visualize newly exocytosed SEP–GluA1 (glutamate A1 receptor), Petrini et al. [183] monitored SEP–GluA1 signal after bleaching a large dendritic region. Small dendritic regions at the boundary of the bleached region were repeatedly bleached to create an optical barrier to exclude fluorescent signal from SEP–GluA1 diffusing laterally in the PM. In this paradigm, approximately 20 % of the SEP–GluA1 signal recovered over 20 min indicating constitutive endocytic cycling of AMPA receptors within dendritic segments [183]. Whereas a chemical LTP stimulus of 0 Mg2 + /glycine increases the frequency of SEP–GluA1 exocytic events [182], glutamate receptor blockers and TTX (tetrodotoxin) decreases the frequency of SEP–GluA1 exocytic events [179]. Simultaneous imaging of SEP– GluA1 and the recycling endosome marker TfR (transferrin receptor)–mCh (mCherry) has shown the abrupt appearance  C The

C 2013 Biochemical Society Authors Journal compilation 

of SEP–GluA1 signal in TfR–mCh-positive spines after stimulation with bicuculline/glycine [178]. Newly inserted SEP–GluA1 either quickly diffused out of the spine (38 % of events) or remained in the spine head (62 % of events). In all cases, co-localized TfR–mCh signal diffused out of the spine immediately following the rapid burst of SEP–GluA1 fluorescence even if newly inserted SEP–GluA1 was retained near the site of spine fusion [178]. Thus, although multiple classes of cargo can be delivered to the spine membrane by all-or-none endosomal fusion events, the fate of cargo after exocytosis can vary on the basis of the molecular identity of the cargo and the state of the synapse. Taken together, a suite of strong evidence supports the notion that paradigms of synaptic potentiation trigger exocytosis of AMPA receptors.

Exocytic domains and molecular machinery for cargo delivery to dendrites Where is cargo exocytosed in relation to the postsynaptic specialization? In the case of SEP–GluA1, studies have reported exocytic events throughout the somatodendritic compartment, but not in dendritic spines proper [179,180,182]. SEP–GluA1 inserted at the subjacent dendritic shaft can diffuse into spines [182], and may become trapped at the PSD by scaffold interactions [184–186] or local endocytic recapture [183,187]. More localized activity manipulations such as two-photon glutamate uncaging have demonstrated exocytosis of SEP–GluA1 in dendrites near and

Dendritic trafficking for neuronal growth and plasticity

Figure 6 Endosomal sorting of dendritic cargo Schematic diagram of endosomal compartments and subdomains involved in cargo sorting and trafficking. AMPA receptors are among the best characterized cargo undergoing endosomal sorting in dendrites. Upon endocytosis, AMPA receptors can be sorted towards recycling or degradative pathways on the basis of signalling events and subunit composition. Endocytosis is accompanied by PP (protein phosphatase) 1- and PP2B-dependent dephosphorylation of GluA1 subunits at Ser845 , whereas recycling and synaptic potentiation are accompanied by rephosphorylation of Ser845 by PKA (protein kinase A). Specific domains and subtypes of endosomes are present in dendrites that transport distinct cargo such as TfRs, Stx13 and Neep21 which in turn can direct traffic to the dendritic membrane or to the axonal compartment. MVBs probably serve to target endosomal cargo for degradation, but may also contribute to exocytic release of exosomes. Although less abundant in dendrites, lysosomes receive cargo for degradation via MVBs as well as autophagosomes whose formation can be regulated by metabolic state and synaptic signalling. Endosomal SNARE proteins including Stx13 and the associated regulator Neep21 define endosomal subcompartments that control transport to the dendritic PM or to the axonal compartment. AMPAR, AMPA receptor; NMDAR, NMDA receptor.

within activated spines [180,181]. Notably, the amplitude of uEPSCs (uncaging-induced excitatory postsynaptic currents) increases first in spines and then in the adjacent dendritic shaft following LTP induction [180], consistent with insertion of glutamate receptors directly in activated spines. One limitation of using a specific receptor cargo as an optical reporter is that the cargo molecule may occupy more than one defined intracellular membrane compartment. Visualization of a tagged fusion of the TfR (TfR–SEP), a classic RE-specific marker, showed that exocytosis of AMPA receptor-containing endosomes occurs within spines, immediately adjacent to the PSD [178]. SEP–GluA1 cargo within TfR-positive endosomes is retained in spines for at least several minutes, whereas co-exocytosed TfR from the same endosome quickly diffuses out of the spine within seconds [178], demonstrating selective AMPA receptor retention at or near synapses [184–186]. Within spines, exocytosis of REs occurs at microdomains enriched for Stx4 (syntaxin 4) [178], one of the four members of the syntaxin

family that localizes to the PM. Disruption of Stx4 blocks spine RE fusion and impairs LTP [178], indicating that Stx4 defines an exocytic domain in dendritic spines for synaptic plasticity and forms part of the core SNARE machinery for spine exocytosis (Figure 5). The other SNARE proteins and adaptors that partner with Stx4 for spine exocytosis have not been determined. Interestingly, SNAP-25 participates in exocytosis of NMDA receptor in dendrites [188,189] and GluA1 insertion is sensitive to botulinum toxin A which selectively cleaves SNAP-25 [180]. A different SNAP family member, SNAP-23 (23 kDa SNAP), is enriched in dendritic spines and localizes at or near the PSD [190], much like Stx4 [178], suggesting that SNAP-23 contributes to the exocytic machinery of dendrites with its known SNARE partner Stx4 [191]. Furthermore, a VAMP family member is likely to be involved in spine exocytosis since postsynaptic infusion of botulinum toxin B or tetanus toxin blocks LTP [161,192]. As discussed above, the synaptotagmin family member Syt4  C The

C 2013 Biochemical Society Authors Journal compilation 

1373

1374

Biochemical Society Transactions (2013) Volume 41, part 6

Figure 7 REs containing AMPA receptors are mobilized by MyoVb for postsynaptic plasticity The Ca2 + -sensitive motor MyoVb associates with REs upon activation of NMDA receptors during LTP, leading to rapid spine recruitment and local exocytosis in spines. Upon Ca2 + influx, MyoVb adopts an open conformation capable of associating with the RE adaptor Rab11-FIP2. Local recruitment and anchoring of REs in spines serves to focus exocytosis locally in spines for delivery of additional AMPA receptors and membrane components that enable synaptic potentiation and spine growth respectively. AMPAR, AMPA receptor; NMDAR, NMDA receptor.

is an attractive candidate for participation in spine-localized SNARE-mediated exocytosis, although this has not been formally demonstrated. Finally, components of the exocyst have been found in spines, and disrupting Sec8 or Exo70 prevents surface trafficking of AMPA receptors, which instead accumulate in intracellular compartments within spines [193]. Before frank membrane fusion, REs are mobilized into spines by the actin-based motor protein MyoVb (MyoV is myosin V) [194] (Figure 7). MyoVb, and the related MyoVa and MyoVc, are Ca2 + -sensitive motors whose threedimensional conformation changes significantly with elevated Ca2 + [195,196], such that in low Ca2 + MyoVb adopts a closed conformation bound to actin, but disassociated from endosomes. As the Ca2 + concentration rises, MyoVb adopts a more extended conformation, allowing physical association with the endosomal adaptor protein Rab11-FIP2 (Rab11 family-interacting protein 2) [194–199]. Hence, upon NMDA receptor-induced Ca2 + influx, MyoVb is recruited to REs and mobilizes them into actin-rich spines where they are positioned to efficiently undergo fusion at Stx4positive exocytic domains (Figures 5 and 7). In hippocampal slices, LTP is acutely blocked by selective chemical-genetic inhibition of MyoVb, demonstrating a requirement for MyoVb-based mobilization of REs for synaptic potentiation [194]. Expression of a dominant-negative version of the related MyoVa or siRNA against MyoVa similarly blocks LTP [200], although synaptic transmission and plasticity was found to be normal in mice lacking MyoVa [201], suggesting differential effects of class V myosins. In summary, it is now clear that exocytosis is a central aspect of the cell biology of dendrites and plays a critical role in both functional and structural plasticity of synapses.  C The

C 2013 Biochemical Society Authors Journal compilation 

A combination of molecular, biochemical, imaging and electrophysiological approaches have provided a much more detailed view of the unique features, spatial regulation, protein machinery and function of exocytosis in dendrites and spines.

Postsynaptic endocytosis and endosomal transport After delivery to specific surface domains of dendrites, the removal of dendritic membrane cargo relies on endocytosis. Endocytosis is, of course, the ubiquitous mechanism by which all cells internalize nutrients, remove membrane from the cell surface and control the level of cell-surface constituents. For many ion channels and postsynaptic receptors, endocytosis and post-endocytic trafficking in dendrites is a primary mechanism controlling their levels and location. In recent years, endocytosis and endosomal sorting of glutamate receptors have emerged as central mechanisms responsible for diverse forms of synaptic plasticity [169–172,183,187,202–207]. Endocytic machinery physically associates with protein complexes of the PSD at excitatory glutamatergic synapses, a finding that initially suggested that endocytosis takes place near the PSD. For example, the GluA2 (glutamate A2 receptor) AMPA receptor subunits and the GluN2B (glutamate N2B receptor) NMDA receptor subunit both interact with the AP-2 (adaptor protein 2) clathrin adaptor complex [207–209], and the PSD adaptor Homer directly binds dynamin-3 [187,210], a member of the dynamin family of large GTPase that mediates vesicle fission [211]. Interactions between postsynaptic receptors and endocytic machinery are tightly regulated. GluA2 forms a Ca2 + -dependent complex with AP-2 through the

Dendritic trafficking for neuronal growth and plasticity

Ca2 + -binding protein hippocalcin required for LTD (longterm depression) [212]. On the other hand, AP-2 binding to GluN2B is prevented by the binding of PSD-95 to an adjacent domain of the receptor [207] as well as by tyrosine phosphorylation of GluN2B by Fyn [206]. This phosphorylation-dependent stabilization can be reversed via tyrosine dephosphorylation by STEP (striatum-enriched phosphatase), which promotes the endocytosis of Glu2Bcontaining NMDA receptors [213] and is itself subject to complex regulation via phosphorylation, ubiquitination and proteolytic cleavage [214]. Initial evidence for dynamic glutamate receptor trafficking demonstrated that AMPA receptors undergo continual endocytosis and exocytosis [169–172,205,215,216]. Endocytosis of glutamate receptors occurs primarily through a dynaminand clathrin-dependent pathway [171,172,187] and most dendritic spines contain the full assembly of clathrin coat proteins and adaptors that capture, remove and recycle glutamate receptors from the cell surface [183,187,217–219]. In dendritic spines, clathrin puncta are found adjacent to, but not overlapping, the PSD where clathrin coats repeatedly assemble and disassemble during ongoing endocytosis [217] (Figure 5). EM has shown the presence of coated structures representing all phases of endocytosis in dendritic spines [218,220], with immunolabelling for AP-2, clathrin and dynamin localized just lateral to the edge of the PSD [218]. The stable lateral positioning of clathrin EZs (endocytic zones) in spines provides a mechanism to fine-tune glutamate receptor levels on a spine by spine basis. How does the endocytic machinery target to spines and what does it do? Studies in hippocampal neurons has shown that dynamin-3 is selectively enriched at the postsynaptic specialization and is the only dynamin isoform known to directly bind to the PSD adaptor protein Homer [187,210,221–224]. Homer proteins assemble as tetramers and form extended protein networks with members of the Shank family of multidomain PSD scaffolding proteins [225– 228]. Disrupting dynamin-3 or preventing its interaction with Homer uncouples the EZ from the PSD, leading to a decrease in the number of PSDs that possess an adjacent EZ [187]. This spatial uncoupling of the EZ depletes synaptic AMPA receptors and decreases synaptic strength, supporting a model whereby the EZ ensures spine-localized endocytic cycling and serves as a trap to recapture and maintain the extrasynaptic pool of AMPA receptors, thereby limiting the escape of AMPA receptors from spines [187]. Indeed, singleparticle tracking and high-resolution live-cell imaging have indicated that both EZs and local receptor recycling maintain a mobile pool of AMPA receptors that is required for synaptic potentiation [183]. Endocytosis in dendrites is regulated by the activityregulated gene product Arc/Arg3.1. Transcription of the Arc gene, transport of its mRNA into dendrites and local translation of Arc protein are all strongly up-regulated by neuronal activity [229,230]. Once up-regulated, Arc binds to the endocytic adaptor endophilin-2 and stimulates

the endocytosis of dendritic and postsynaptic cargo, most notably AMPA receptors and the amyloid precursor protein [231,232] (Figure 5). This activity-induced internalization of AMPA receptors provides a homoeostatic mechanism to dampen excitatory synaptic transmission following bouts of strong neuronal activity [233]. The precise mechanisms by which Arc stimulates endocytosis are not clear, but one possibility is that Arc directs subsets of cargo at specific endocytic sites for specific endosomal sorting pathways [232]. Although Arc expression is limited to neurons, specialization of endocytic sites occurs for specific cargo and controls receptor desensitization and signalling in diverse cell types. For GPCRs (G-protein-coupled receptors), arrestin recruitment to activated receptors is well known to link cargo to clathrin-mediated endocytosis and downstream signalling [234,235]. Interestingly, dendrites possess specific modes of endocytic recycling for GPCRs that differ in spatial and kinetic features [236,237], and clathrin puncta themselves become increasingly stable on dendrites as neurons mature in culture [238]. Following endocytosis, dendritic cargo such as AMPA receptors and NMDA receptors are sorted in endosomes for either recycling or lysosomal degradation (Figure 6). For AMPA receptors, agonist binding alone leads to rapid downregulation and lysosomal degradation, whereas activation of NMDA receptors initially causes AMPA receptor endocytosis that can be followed by subsequent recycling back to the cell surface [169]. Such endosomal sorting corresponds with the phosphorylation state of the GluA1 subunit. Dephosphorylation of Ser845 of GluA1 occurs during LTD [239] and may be the sorting signal that leads to lysosomal degradation [169], whereas phosphorylation of GluA1 at Ser845 accompanies reinsertion at the dendritic membrane via REs [169,240,241]. Coincident with dephosphorylation of GluA1 at Ser845 , LTD-inducing stimuli trigger Ca2 + dependent activation of Rab5 which leads to the endocytosis and early endosomal transport of GluA1 [242]. In the case of NMDA receptors, internalized receptors are routed along degradative and recycling pathways on the basis of specific motifs in GluN1 (glutamate N1 receptor), GluN2A (glutamate N2A receptor) and GluN2B subunits [203,243]. Internal membrane compartments resembling endosomes have been observed in dendrites. These include coated and uncoated vesicles, tubules and MVBs (multivesicular bodies) [85,167]. Approximately 70 % of endosome-like structures are within or at the base of dendritic spines, and approximately one-third of CA1 hippocampal spines are associated with endosomes, which suggests that the same endocytic organelles are shared by multiple spines [85]. Initial studies showed that Stx13, an endosomal SNARE protein, is found in dendritic tubulovesicular compartments, where it co-localizes with TfR, a protein known to be recycled through the endosomal pathway [244]. Similarly, Neep21 (neuron-enriched endosomal protein of 21 kDa) is a binding partner of Stx13 that localizes to and defines a population of early dendritic endosomes [245–247]. Disruption of either

 C The

C 2013 Biochemical Society Authors Journal compilation 

1375

1376

Biochemical Society Transactions (2013) Volume 41, part 6

Stx13 or Neep21 decreases the recycling of GluA1 and GluA2 after endocytosis and impairs synaptic potentiation [166,248]. Whereas early and recycling endosomes are abundant in dendrites, degradative compartments such as late endosomes, MVBs and lysosomes are more sparse [85,169,249]. MVBs have been noted in association with postsynaptic sites [85,148,250–253]. Lysosomes generally concentrate in the soma and proximal dendrites of neurons [254,255], although ultrastructural and biochemical studies have long supported the presence of lysosomes and lysosomal degradation at synapses [256]. In non-neuronal cells, the peripheral versus the perinuclear positioning of lysosomes is controlled by regulated association with kinesin and dynein motors [257– 262], and destabilization of microtubules causes accumulation of lysosomes in dendrites [254,255], suggesting active microtubule-based transport from the somatodendritic compartment into axons. Interestingly, the proximity of lysosomes to early endosomes can determine the propensity of internalized cargo to transit recycling versus degradative pathways [263], suggesting that the presence or absence of lysosomes can control dendritic signalling. Indeed, internalized AMPA receptors can be degraded by lysosomes [169,204] and stimuli that induce LTD up-regulate autophagy and presumably subsequent autophagolysosome-mediated degradation [264] (Figure 6). Many neuronal membrane proteins are long-lived with half-lives of days to weeks [265]. It remains an open question how endocytic cargo destined for destruction in distal dendrites is transported to lysosomes, or, conversely, how lysosomes are generated and distributed upon demand through the dendritic arbour. Given the important role of lysosomes in protein clearance, and their dysfunction in neurodegenerative disorders [266], the local regulation of dendritic lysosomes remains an important topic for future study.

Conclusions and future directions It has been over 100 years since Golgi first described intracellular organelles in neurons [1]. The last 15 years have witnessed a remarkable convergence between neuronal membrane trafficking, brain development and synaptic plasticity. Previously separated as classic disciplines of anatomy and neurophysiology, it is now clear that a full understanding of neural circuits in health and disease requires dissection of the subcellular and organellar mechanisms of neuronal trafficking. For, at their core, many questions of synapse and circuit plasticity are issues of cell biology. From the modification and maintenance of individual synapses, to the integrative properties of dendritic branches, to the distribution of channels and receptors that determines the computational features of neurons, emergent properties of cells and circuits rely on the remarkable ability of neurons to compartmentalize. As a result, neurons have adapted conserved cellular trafficking machinery for highly specialized functions. Beyond a role in normal brain development and physiology, many of the most prevalent and devastating  C The

C 2013 Biochemical Society Authors Journal compilation 

neurological disorders are now understood to be disorders of neuronal membrane trafficking. From the production, release and clearance of β-amyloid and tau in Alzheimer’s disease, to the abnormal accumulation and transport of α-synuclein in Parkinson’s disease, to the trafficking defects associated with abnormal huntingtin in Huntington’s disease, core aspects of neurodegeneration impinge directly on the organelles for dendritic trafficking. In addition, it is now clear that the most common psychiatric disorders, including schizophrenia, autism, bipolar disorder and ADHD (attention-deficit hyperactivity disorder), originate during brain development where the growth and development of dendrites and synapses requires exquisite regulation of dendritic trafficking. Targeting these disorders with next-generation therapeutics will require a better understanding of how receptors and ion channels are trafficked, how synapses are modified and how deficits in neuronal trafficking can be corrected. Many questions remain. What are the full set of molecular sensors that connect neural activity to dendritic membrane trafficking? What controls the biogenesis, localization and transport of dendritic organelles themselves? How do neurons maintain their size, shape and composition in the face of ongoing macromolecular turnover? How do genetic programmes that direct neuronal subtype differentiation impinge on trafficking machinery? How does local dendritic trafficking influence post-translational modification of signalling molecules, channels and receptors? What are the molecular mechanisms by which development, aging and disease alter dendritic transport? Although much work remains, the fusion of fields at the interface of membrane trafficking and neuronal plasticity continues to bud new insight into the inner workings of Nature’s most complex cells.

Acknowledgements I thank Cyril Hanus, Matthew Kennedy, Mikyoung Park and Zhiping Wang for assistance with the Figures. I thank Soheil Aghamohammedzadeh, John Allen, Cyril Hanus, Tom Helton, Juliet Hernandez and Matthew Kennedy for a critical review of the paper before submission. I thank all members of the Ehlers laboratory past and present who made receipt of the Thudichum Medal possible.

Funding Work in my laboratory is supported by Pfizer, Inc.

References 1 2 3

Golgi, C. (1898) Intorno alla struttura delle cellule nervose. Boll. Soc. Med.-Chir. Pavia 13, 3–16 Spruston, N. (2008) Pyramidal neurons: dendritic structure and synaptic integration. Nat. Rev. Neurosci. 9, 206–221 Matsuzaki, M., Honkura, N., Ellis-Davies, G.C. and Kasai, H. (2004) Structural basis of long-term potentiation in single dendritic spines. Nature 429, 761–766

Dendritic trafficking for neuronal growth and plasticity

4

5

6

7 8

9 10

11

12

13

14

15

16

17

18

19 20

21

22

23

24

25 26

Losonczy, A. and Magee, J.C. (2006) Integrative properties of radial oblique dendrites in hippocampal CA1 pyramidal neurons. Neuron 50, 291–307 Losonczy, A., Makara, J.K. and Magee, J.C. (2008) Compartmentalized dendritic plasticity and input feature storage in neurons. Nature 452, 436–441 Nicholson, D.A., Trana, R., Katz, Y., Kath, W.L., Spruston, N. and Geinisman, Y. (2006) Distance-dependent differences in synapse number and AMPA receptor expression in hippocampal CA1 pyramidal neurons. Neuron 50, 431–442 Shah, M.M., Hammond, R.S. and Hoffman, D.A. (2010) Dendritic ion channel trafficking and plasticity. Trends Neurosci. 33, 307–316 Spacek, J. and Harris, K.M. (1997) Three-dimensional organization of smooth endoplasmic reticulum in hippocampal CA1 dendrites and dendritic spines of the immature and mature rat. J. Neurosci. 17, 190–203 Gardiol, A., Racca, C. and Triller, A. (1999) Dendritic and postsynaptic protein synthetic machinery. J. Neurosci. 19, 168–179 Torre, E.R. and Steward, O. (1996) Protein synthesis within dendrites: glycosylation of newly synthesized proteins in dendrites of hippocampal neurons in culture. J. Neurosci. 16, 5967–5978 Cui-Wang, T., Hanus, C., Cui, T., Helton, T., Bourne, J., Watson, D., Harris, K.M. and Ehlers, M.D. (2012) Local zones of endoplasmic reticulum complexity confine cargo in neuronal dendrites. Cell 148, 309–321 Rose, C.R. and Konnerth, A. (2001) Stores not just for storage: intracellular calcium release and synaptic plasticity. Neuron 31, 519–522 Borgese, N., Francolini, M. and Snapp, E. (2006) Endoplasmic reticulum architecture: structures in flux. Curr. Opin. Cell Biol. 18, 358–364 Pierce, J.P., Mayer, T. and McCarthy, J.B. (2001) Evidence for a satellite secretory pathway in neuronal dendritic spines. Curr. Biol. 11, 351–355 Pierce, J.P., van Leyen, K. and McCarthy, J.B. (2000) Translocation machinery for synthesis of integral membrane and secretory proteins in dendritic spines. Nat. Neurosci. 3, 311–313 Sutton, M.A., Ito, H.T., Cressy, P., Kempf, C., Woo, J.C. and Schuman, E.M. (2006) Miniature neurotransmission stabilizes synaptic function via tonic suppression of local dendritic protein synthesis. Cell 125, 785–799 Mameli, M., Balland, B., Lujan, R. and Luscher, C. (2007) Rapid synthesis and synaptic insertion of GluR2 for mGluR-LTD in the ventral tegmental area. Science 317, 530–533 Ju, W., Morishita, W., Tsui, J., Gaietta, G., Deerinck, T.J., Adams, S.R., Garner, C.C., Tsien, R.Y., Ellisman, M.H. and Malenka, R.C. (2004) Activity-dependent regulation of dendritic synthesis and trafficking of AMPA receptors. Nat. Neurosci. 7, 244–253 Gillon, A.D., Latham, C.F. and Miller, E.A. (2012) Vesicle-mediated ER export of proteins and lipids. Biochim. Biophys. Acta 1821, 1040–1049 Horton, A.C. and Ehlers, M.D. (2003) Dual modes of endoplasmic reticulum-to-Golgi transport in dendrites revealed by live-cell imaging. J. Neurosci. 23, 6188–6199 Aridor, M., Guzik, A.K., Bielli, A. and Fish, K.N. (2004) Endoplasmic reticulum export site formation and function in dendrites. J. Neurosci. 24, 3770–3776 Saraste, J., Dale, H.A., Bazzocco, S. and Marie, M. (2009) Emerging new roles of the pre-Golgi intermediate compartment in biosynthetic-secretory trafficking. FEBS Lett. 583, 3804–3810 Krijnse-Locker, J., Parton, R.G., Fuller, S.D., Griffiths, G. and Dotti, C.G. (1995) The organization of the endoplasmic reticulum and the intermediate compartment in cultured rat hippocampal neurons. Mol. Biol. Cell 6, 1315–1332 Greger, I.H., Khatri, L. and Ziff, E.B. (2002) RNA editing at Arg607 controls AMPA receptor exit from the endoplasmic reticulum. Neuron 34, 759–772 Greger, I.H. and Esteban, J.A. (2007) AMPA receptor biogenesis and trafficking. Curr. Opin. Neurobiol. 17, 289–297 Hawkins, L.M., Prybylowski, K., Chang, K., Moussan, C., Stephenson, F.A. and Wenthold, R.J. (2004) Export from the endoplasmic reticulum of assembled N-methyl-d-aspartic acid receptors is controlled by a motif in the c terminus of the NR2 subunit. J. Biol. Chem. 279, 28903–28910

27

28

29

30

31

32

33

34

35

36

37

38 39

40

41

42 43

44 45

46

47

48

49

Horak, M., Chang, K. and Wenthold, R.J. (2008) Masking of the endoplasmic reticulum retention signals during assembly of the NMDA receptor. J. Neurosci. 28, 3500–3509 Standley, S., Roche, K.W., McCallum, J., Sans, N. and Wenthold, R.J. (2000) PDZ domain suppression of an ER retention signal in NMDA receptor NR1 splice variants. Neuron 28, 887–898 Scott, D.B., Blanpied, T.A., Swanson, G.T., Zhang, C. and Ehlers, M.D. (2001) An NMDA receptor ER retention signal regulated by phosphorylation and alternative splicing. J. Neurosci. 21, 3063–3072 Mu, Y., Otsuka, T., Horton, A.C., Scott, D.B. and Ehlers, M.D. (2003) Activity-dependent mRNA splicing controls ER export and synaptic delivery of NMDA receptors. Neuron 40, 581–594 Scott, D.B., Blanpied, T.A. and Ehlers, M.D. (2003) Coordinated PKA and PKC phosphorylation suppresses RXR-mediated ER retention and regulates the surface delivery of NMDA receptors. Neuropharmacology 45, 755–767 Lester, H.A., Miwa, J.M. and Srinivasan, R. (2012) Psychiatric drugs bind to classical targets within early exocytotic pathways: therapeutic effects. Biol. Psychiatry 72, 907–915 Srinivasan, R., Richards, C.I., Xiao, C., Rhee, D., Pantoja, R., Dougherty, D.A., Miwa, J.M. and Lester, H.A. (2012) Pharmacological chaperoning of nicotinic acetylcholine receptors reduces the endoplasmic reticulum stress response. Mol. Pharmacol. 81, 759–769 Srinivasan, R., Pantoja, R., Moss, F.J., Mackey, E.D., Son, C.D., Miwa, J. and Lester, H.A. (2011) Nicotine up-regulates α4β2 nicotinic receptors and ER exit sites via stoichiometry-dependent chaperoning. J. Gen. Physiol. 137, 59–79 Lester, H.A., Xiao, C., Srinivasan, R., Son, C.D., Miwa, J., Pantoja, R., Banghart, M.R., Dougherty, D.A., Goate, A.M. and Wang, J.C. (2009) Nicotine is a selective pharmacological chaperone of acetylcholine receptor number and stoichiometry: implications for drug discovery. AAPS J. 11, 167–177 de Graffenried, C.L. and Bertozzi, C.R. (2004) The roles of enzyme localisation and complex formation in glycan assembly within the Golgi apparatus. Curr. Opin. Cell Biol. 16, 356–363 Maeda, Y. and Kinoshita, T. (2010) The acidic environment of the Golgi is critical for glycosylation and transport. Methods Enzymol. 480, 495–510 Tartakoff, A., Vassalli, P. and Detraz, M. (1978) Comparative studies of intracellular transport of secretory proteins. J. Cell Biol. 79, 694–707 Thorens, B. and Vassalli, P. (1986) Chloroquine and ammonium chloride prevent terminal glycosylation of immunoglobulins in plasma cells without affecting secretion. Nature 321, 618–620 Palokangas, H., Metsikko, K. and Vaananen, K. (1994) Active vacuolar H + ATPase is required for both endocytic and exocytic processes during viral infection of BHK-21 cells. J. Biol. Chem. 269, 17577–17585 Rivinoja, A., Kokkonen, N., Kellokumpu, I. and Kellokumpu, S. (2006) Elevated Golgi pH in breast and colorectal cancer cells correlates with the expression of oncofetal carbohydrate T-antigen. J. Cell. Physiol. 208, 167–174 Rivinoja, A., Pujol, F.M., Hassinen, A. and Kellokumpu, S. (2012) Golgi pH, its regulation and roles in human disease. Ann. Med. 44, 542–554 Chanat, E. and Huttner, W.B. (1991) Milieu-induced, selective aggregation of regulated secretory proteins in the trans-Golgi network. J. Cell Biol. 115, 1505–1519 Huang, C. and Chang, A. (2011) pH-dependent cargo sorting from the Golgi. J. Biol. Chem. 286, 10058–10065 Condon, K.H., Ho, J., Robinson, C.G., Hanus, C. and Ehlers, M.D. (2013) The Angelman syndrome protein Ube3a/E6AP is required for Golgi acidification and surface protein sialylation. J. Neurosci. 33, 3799–3814 Miao, S., Chen, R., Ye, J., Tan, G.H., Li, S., Zhang, J., Jiang, Y.H. and Xiong, Z.Q. (2013) The Angelman syndrome protein Ube3a is required for polarized dendrite morphogenesis in pyramidal neurons. J. Neurosci. 33, 327–333 Dindot, S.V., Antalffy, B.A., Bhattacharjee, M.B. and Beaudet, A.L. (2008) The Angelman syndrome ubiquitin ligase localizes to the synapse and nucleus, and maternal deficiency results in abnormal dendritic spine morphology. Hum. Mol. Genet. 17, 111–118 Yashiro, K., Riday, T.T., Condon, K.H., Roberts, A.C., Bernardo, D.R., Prakash, R., Weinberg, R.J., Ehlers, M.D. and Philpot, B.D. (2009) Ube3a is required for experience-dependent maturation of the neocortex. Nat. Neurosci. 12, 777–783 Jiang, Y.H., Armstrong, D., Albrecht, U., Atkins, C.M., Noebels, J.L., Eichele, G., Sweatt, J.D. and Beaudet, A.L. (1998) Mutation of the Angelman ubiquitin ligase in mice causes increased cytoplasmic p53 and deficits of contextual learning and long-term potentiation. Neuron 21, 799–811  C The

C 2013 Biochemical Society Authors Journal compilation 

1377

1378

Biochemical Society Transactions (2013) Volume 41, part 6

50

51

52

53

54

55

56

57

58

59

60

61

62

63 64 65 66

67

68

69

70

71

72

 C The

Wallace, M.L., Burette, A.C., Weinberg, R.J. and Philpot, B.D. (2012) Maternal loss of Ube3a produces an excitatory/inhibitory imbalance through neuron type-specific synaptic defects. Neuron 74, 793–800 Mabb, A.M., Judson, M.C., Zylka, M.J. and Philpot, B.D. (2011) Angelman syndrome: insights into genomic imprinting and neurodevelopmental phenotypes. Trends Neurosci. 34, 293–303 Ledger, P.W., Uchida, N. and Tanzer, M.L. (1980) Immunocytochemical localization of procollagen and fibronectin in human fibroblasts: effects of the monovalent ionophore, monensin. J. Cell Biol. 87, 663–671 Boss, W.F., Morre, D.J. and Mollenhauer, H.H. (1984) Monensin-induced swelling of Golgi apparatus cisternae mediated by a proton gradient. Eur. J. Cell Biol. 34, 1–8 Kellokumpu, S., Sormunen, R. and Kellokumpu, I. (2002) Abnormal glycosylation and altered Golgi structure in colorectal cancer: dependence on intra-Golgi pH. FEBS Lett. 516, 217–224 Lazaro-Dieguez, F., Jimenez, N., Barth, H., Koster, A.J., Renau-Piqueras, J., Llopis, J.L., Burger, K.N. and Egea, G. (2006) Actin filaments are involved in the maintenance of Golgi cisternae morphology and intra-Golgi pH. Cell Motil. Cytoskeleton 63, 778–791 Niemann, H., Boschek, B., Evans, D., Rosing, M., Tamura, T. and Klenk, H.D. (1982) Post-translational glycosylation of coronavirus glycoprotein E1: inhibition by monensin. EMBO J. 1, 1499–1504 Alonso-Caplen, F.V. and Compans, R.W. (1983) Modulation of glycosylation and transport of viral membrane glycoproteins by a sodium ionophore. J. Cell Biol. 97, 659–668 Ledger, P.W., Nishimoto, S.K., Hayashi, S. and Tanzer, M.L. (1983) Abnormal glycosylation of human fibronectin secreted in the presence of monensin. J. Biol. Chem. 258, 547–554 Axelsson, M.A., Karlsson, N.G., Steel, D.M., Ouwendijk, J., Nilsson, T. and Hansson, G.C. (2001) Neutralization of pH in the Golgi apparatus causes redistribution of glycosyltransferases and changes in the O-glycosylation of mucins. Glycobiology 11, 633–644 Maeda, Y., Ide, T., Koike, M., Uchiyama, Y. and Kinoshita, T. (2008) GPHR is a novel anion channel critical for acidification and functions of the Golgi apparatus. Nat. Cell Biol. 10, 1135–1145 Rivinoja, A., Hassinen, A., Kokkonen, N., Kauppila, A. and Kellokumpu, S. (2009) Elevated Golgi pH impairs terminal N-glycosylation by inducing mislocalization of Golgi glycosyltransferases. J. Cell. Physiol. 220, 144–154 Hassinen, A., Pujol, F.M., Kokkonen, N., Pieters, C., Kihlstrom, M., Korhonen, K. and Kellokumpu, S. (2011) Functional organization of Golgi N- and O-glycosylation pathways involves pH-dependent complex formation that is impaired in cancer cells. J. Biol. Chem. 286, 38329–38340 Kleene, R. and Schachner, M. (2004) Glycans and neural cell interactions. Nat. Rev. Neurosci. 5, 195–208 Varki, A. (2007) Glycan-based interactions involving vertebrate sialic-acid-recognizing proteins. Nature 446, 1023–1029 Varki, A. (2008) Sialic acids in human health and disease. Trends Mol. Med. 14, 351–360 Castillo, C., Diaz, M.E., Balbi, D., Thornhill, W.B. and Recio-Pinto, E. (1997) Changes in sodium channel function during postnatal brain development reflect increases in the level of channel sialidation. Brain Res. 104, 119–130 Tyrrell, L., Renganathan, M., Dib-Hajj, S.D. and Waxman, S.G. (2001) Glycosylation alters steady-state inactivation of sodium channel Nav1.9/NaN in dorsal root ganglion neurons and is developmentally regulated. J. Neurosci. 21, 9629–9637 Isaev, D., Isaeva, E., Shatskih, T., Zhao, Q., Smits, N.C., Shworak, N.W., Khazipov, R. and Holmes, G.L. (2007) Role of extracellular sialic acid in regulation of neuronal and network excitability in the rat hippocampus. J. Neurosci. 27, 11587–11594 Rutishauser, U. (2008) Polysialic acid in the plasticity of the developing and adult vertebrate nervous system. Nat. Rev. Neurosci. 9, 26–35 Yu, R.K., Nakatani, Y. and Yanagisawa, M. (2009) The role of glycosphingolipid metabolism in the developing brain. J. Lipid Res. 50 (Suppl.), S440–S445 Repnikova, E., Koles, K., Nakamura, M., Pitts, J., Li, H., Ambavane, A., Zoran, M.J. and Panin, V.M. (2010) Sialyltransferase regulates nervous system function in Drosophila. J. Neurosci. 30, 6466–6476 Newbury, D.F., Winchester, L., Addis, L., Paracchini, S., Buckingham, L.L., Clark, A., Cohen, W., Cowie, H., Dworzynski, K., Everitt, A. et al. (2009) CMIP and ATP2C2 modulate phonological short-term memory in language impairment. Am. J. Hum. Genet. 85, 264–272 C 2013 Biochemical Society Authors Journal compilation 

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

Simpson, M.A., Cross, H., Proukakis, C., Priestman, D.A., Neville, D.C., Reinkensmeier, G., Wang, H., Wiznitzer, M., Gurtz, K., Verganelaki, A. et al. (2004) Infantile-onset symptomatic epilepsy syndrome caused by a homozygous loss-of-function mutation of GM3 synthase. Nat. Genet. 36, 1225–1229 Lee, M.C., Miller, E.A., Goldberg, J., Orci, L. and Schekman, R. (2004) Bi-directional protein transport between the ER and Golgi. Annu. Rev. Cell Dev. Biol. 20, 87–123 De Camilli, P., Moretti, M., Donini, S.D., Walter, U. and Lohmann, S.M. (1986) Heterogeneous distribution of the cAMP receptor protein RII in the nervous system: evidence for its intracellular accumulation on microtubules, microtubule-organizing centers, and in the area of the Golgi complex. J. Cell Biol. 103, 189–203 Lowenstein, P.R., Morrison, E.E., Bain, D., Shering, A.F., Banting, G., Douglas, P. and Castro, M.G. (1994) Polarized distribution of the trans-Golgi network marker TGN38 during the in vitro development of neocortical neurons: effects of nocodazole and brefeldin A. Eur. J. Neurosci. 6, 1453–1465 Horton, A.C., Racz, B., Monson, E.E., Lin, A.L., Weinberg, R.J. and Ehlers, M.D. (2005) Polarized secretory trafficking directs cargo for asymmetric dendrite growth and morphogenesis. Neuron 48, 757–771 Ye, B., Zhang, Y., Song, W., Younger, S.H., Jan, L.Y. and Jan, Y.N. (2007) Growing dendrites and axons differ in their reliance on the secretory pathway. Cell 130, 717–729 Jeyifous, O., Waites, C.L., Specht, C.G., Fujisawa, S., Schubert, M., Lin, E.I., Marshall, J., Aoki, C., de Silva, T., Montgomery, J.M. et al. (2009) SAP97 and CASK mediate sorting of NMDA receptors through a previously unknown secretory pathway. Nat. Neurosci. 12, 1011–1019 Cutrona, M.B., Beznoussenko, G.V., Fusella, A., Martella, O., Moral, P. and Mironov, A.A. (2013) Silencing of mammalian Sar1 isoforms reveals COPII-independent protein sorting and transport. Traffic 14, 691–708 Sannerud, R., Marie, M., Nizak, C., Dale, H.A., Pernet-Gallay, K., Perez, F., Goud, B. and Saraste, J. (2006) Rab1 defines a novel pathway connecting the pre-Golgi intermediate compartment with the cell periphery. Mol. Biol. Cell 17, 1514–1526 Storey, G.P., Opitz-Araya, X. and Barria, A. (2011) Molecular determinants controlling NMDA receptor synaptic incorporation. J. Neurosci. 31, 6311–6316 Watanabe, I., Zhu, J., Recio-Pinto, E. and Thornhill, W.B. (2004) 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. 279, 8879–8885 Schwetz, T.A., Norring, S.A., Ednie, A.R. and Bennett, E.S. (2011) Sialic acids attached to O-glycans modulate voltage-gated potassium channel gating. J. Biol. Chem. 286, 4123–4132 Cooney, J.R., Hurlburt, J.L., Selig, D.K., Harris, K.M. and Fiala, J.C. (2002) Endosomal compartments serve multiple hippocampal dendritic spines from a widespread rather than a local store of recycling membrane. J. Neurosci. 22, 2215–2224 Terasaki, M., Slater, N.T., Fein, A., Schmidek, A. and Reese, T.S. (1994) Continuous network of endoplasmic reticulum in cerebellar Purkinje neurons. Proc. Natl. Acad. Sci. U.S.A. 91, 7510–7514 Herpers, B. and Rabouille, C. (2004) mRNA localization and ER-based protein sorting mechanisms dictate the use of transitional endoplasmic reticulum–Golgi units involved in gurken transport in Drosophila oocytes. Mol. Biol. Cell 15, 5306–5317 Richard, M., Boulin, T., Robert, V.J., Richmond, J.E. and Bessereau, J.L. (2013) Biosynthesis of ionotropic acetylcholine receptors requires the evolutionarily conserved ER membrane complex. Proc. Natl. Acad. Sci. U.S.A. 110, E1055–E1063 Gorrie, G.H., Vallis, Y., Stephenson, A., Whitfield, J., Browning, B., Smart, T.G. and Moss, S.J. (1997) Assembly of GABAA receptors composed of α1 and β2 subunits in both cultured neurons and fibroblasts. J. Neurosci. 17, 6587–6596 Jensen, T.J., Loo, M.A., Pind, S., Williams, D.B., Goldberg, A.L. and Riordan, J.R. (1995) Multiple proteolytic systems, including the proteasome, contribute to CFTR processing. Cell 83, 129–135 Merlie, J.P. and Lindstrom, J. (1983) Assembly in vivo of mouse muscle acetylcholine receptor: identification of an α subunit species that may be an assembly intermediate. Cell 34, 747–757 Hu, J., Shibata, Y., Zhu, P.P., Voss, C., Rismanchi, N., Prinz, W.A., Rapoport, T.A. and Blackstone, C. (2009) A class of dynamin-like GTPases involved in the generation of the tubular ER network. Cell 138, 549–561

Dendritic trafficking for neuronal growth and plasticity

93

94

95 96

97

98

99

100

101

102 103 104

105

106

107 108 109 110

111

112 113 114

115

116 117 118 119

Orso, G., Pendin, D., Liu, S., Tosetto, J., Moss, T.J., Faust, J.E., Micaroni, M., Egorova, A., Martinuzzi, A., McNew, J.A. and Daga, A. (2009) Homotypic fusion of ER membranes requires the dynamin-like GTPase atlastin. Nature 460, 978–983 Voeltz, G.K., Prinz, W.A., Shibata, Y., Rist, J.M. and Rapoport, T.A. (2006) A class of membrane proteins shaping the tubular endoplasmic reticulum. Cell 124, 573–586 English, A.R. and Voeltz, G.K. (2013) Rab10 GTPase regulates ER dynamics and morphology. Nat. Cell Biol. 15, 169–178 Klopfenstein, D.R., Kappeler, F. and Hauri, H.P. (1998) A novel direct interaction of endoplasmic reticulum with microtubules. EMBO J. 17, 6168–6177 Schweizer, A., Ericsson, M., Bachi, T., Griffiths, G. and Hauri, H.P. (1993) Characterization of a novel 63 kDa membrane protein: implications for the organization of the ER-to-Golgi pathway. J. Cell Sci. 104, 671–683 Vedrenne, C., Klopfenstein, D.R. and Hauri, H.P. (2005) Phosphorylation controls CLIMP-63-mediated anchoring of the endoplasmic reticulum to microtubules. Mol. Biol. Cell 16, 1928–1937 Horton, A.C., Yi, J.J. and Ehlers, M.D. (2006) Cell type-specific dendritic polarity in the absence of spatially organized external cues. Brain Cell Biol. 35, 29–38 Matsuki, T., Matthews, R.T., Cooper, J.A., van der Brug, M.P., Cookson, M.R., Hardy, J.A., Olson, E.C. and Howell, B.W. (2010) Reelin and stk25 have opposing roles in neuronal polarization and dendritic Golgi deployment. Cell 143, 826–836 Meseke, M., Rosenberger, G. and Forster, E. (2013) Reelin and the Cdc42/Rac1 guanine nucleotide exchange factor αPIX/Arhgef6 promote dendritic Golgi translocation in hippocampal neurons. Eur. J. Neurosci. 37, 1404–1412 Jahn, R. and Fasshauer, D. (2012) Molecular machines governing exocytosis of synaptic vesicles. Nature 490, 201–207 Sorensen, J.B. (2009) Conflicting views on the membrane fusion machinery and the fusion pore. Annu. Rev. Cell Dev. Biol. 25, 513–537 Rettig, J., Heinemann, C., Ashery, U., Sheng, Z.H., Yokoyama, C.T., Catterall, W.A. and Neher, E. (1997) Alteration of Ca2 + dependence of neurotransmitter release by disruption of Ca2 + channel/syntaxin interaction. J. Neurosci. 17, 6647–6656 Rettig, J., Sheng, Z.H., Kim, D.K., Hodson, C.D., Snutch, T.P. and Catterall, W.A. (1996) Isoform-specific interaction of the α1A subunits of brain Ca2 + channels with the presynaptic proteins syntaxin and SNAP-25. Proc. Natl. Acad. Sci. U.S.A. 93, 7363–7368 Takamori, S., Holt, M., Stenius, K., Lemke, E.A., Gronborg, M., Riedel, D., Urlaub, H., Schenck, S., Brugger, B., Ringler, P. et al. (2006) Molecular anatomy of a trafficking organelle. Cell 127, 831–846 Sudhof, T.C. and Rothman, J.E. (2009) Membrane fusion: grappling with SNARE and SM proteins. Science 323, 474–477 Chapman, E.R. (2008) How does synaptotagmin trigger neurotransmitter release? Annu. Rev. Biochem. 77, 615–641 Famiglietti, Jr, E.V. (1970) Dendro-dendritic synapses in the lateral geniculate nucleus of the cat. Brain Res. 20, 181–191 Hirata, Y. (1964) Some observations on the fine structure of the synapses in the olfactory bulb of the mouse, with particular reference to the atypical synaptic configurations. Arch. Histol. Jpn. 24, 293–302 Lagier, S., Panzanelli, P., Russo, R.E., Nissant, A., Bathellier, B., Sassoe-Pognetto, M., Fritschy, J.M. and Lledo, P.M. (2007) GABAergic inhibition at dendrodendritic synapses tunes gamma oscillations in the olfactory bulb. Proc. Natl. Acad. Sci. U.S.A. 104, 7259–7264 Price, J.L. and Powell, T.P. (1970) The morphology of the granule cells of the olfactory bulb. J. Cell Sci. 7, 91–123 Price, J.L. and Powell, T.P. (1970) The synaptology of the granule cells of the olfactory bulb. J. Cell Sci. 7, 125–155 Rall, W., Shepherd, G.M., Reese, T.S. and Brightman, M.W. (1966) Dendrodendritic synaptic pathway for inhibition in the olfactory bulb. Exp. Neurol. 14, 44–56 Shanks, M.F. and Powell, T.P. (1981) An electron microscopic study of the termination of thalamocortical fibres in areas 3b, 1 and 2 of the somatic sensory cortex in the monkey. Brain Res. 218, 35–47 Isaacson, J.S. and Strowbridge, B.W. (1998) Olfactory reciprocal synapses: dendritic signaling in the CNS. Neuron 20, 749–761 Jahr, C.E. and Nicoll, R.A. (1980) Dendrodendritic inhibition: demonstration with intracellular recording. Science 207, 1473–1475 Jahr, C.E. and Nicoll, R.A. (1982) Noradrenergic modulation of dendrodendritic inhibition in the olfactory bulb. Nature 297, 227–229 Phillips, C.G., Powell, T.P. and Shepherd, G.M. (1963) Responses of mitral cells to stimulation of the lateral olfactory tract in the rabbit. J. Physiol. 168, 65–88

120

121

122

123

124

125

126

127 128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

Isaacson, J.S. (2001) Mechanisms governing dendritic γ -aminobutyric acid (GABA) release in the rat olfactory bulb. Proc. Natl. Acad. Sci. U.S.A. 98, 337–342 Sassoe-Pognetto, M., Utvik, J.K., Camoletto, P., Watanabe, M., Stephenson, F.A., Bredt, D.S. and Ottersen, O.P. (2003) Organization of postsynaptic density proteins and glutamate receptors in axodendritic and dendrodendritic synapses of the rat olfactory bulb. J. Comp. Neurol. 463, 237–248 Chen, W.R., Shen, G.Y., Shepherd, G.M., Hines, M.L. and Midtgaard, J. (2002) Multiple modes of action potential initiation and propagation in mitral cell primary dendrite. J. Neurophysiol. 88, 2755–2764 Halabisky, B., Friedman, D., Radojicic, M. and Strowbridge, B.W. (2000) Calcium influx through NMDA receptors directly evokes GABA release in olfactory bulb granule cells. J. Neurosci. 20, 5124–5134 Schoppa, N.E., Kinzie, J.M., Sahara, Y., Segerson, T.P. and Westbrook, G.L. (1998) Dendrodendritic inhibition in the olfactory bulb is driven by NMDA receptors. J. Neurosci. 18, 6790–6802 Adler, E.M., Augustine, G.J., Duffy, S.N. and Charlton, M.P. (1991) Alien intracellular calcium chelators attenuate neurotransmitter release at the squid giant synapse. J. Neurosci. 11, 1496–1507 Geffen, L.B., Jessell, T.M., Cuello, A.C. and Iversen, L.L. (1976) Release of dopamine from dendrites in rat substantia nigra. Nature 260, 258–260 Cheramy, A., Leviel, V. and Glowinski, J. (1981) Dendritic release of dopamine in the substantia nigra. Nature 289, 537–542 Jaffe, E.H., Marty, A., Schulte, A. and Chow, R.H. (1998) Extrasynaptic vesicular transmitter release from the somata of substantia nigra neurons in rat midbrain slices. J. Neurosci. 18, 3548–3553 Kalivas, P.W. and Duffy, P. (1988) Effects of daily cocaine and morphine treatment on somatodendritic and terminal field dopamine release. J. Neurochem. 50, 1498–1504 Fortin, G.D., Desrosiers, C.C., Yamaguchi, N. and Trudeau, L.E. (2006) Basal somatodendritic dopamine release requires snare proteins. J. Neurochem. 96, 1740–1749 Li, H., Waites, C.L., Staal, R.G., Dobryy, Y., Park, J., Sulzer, D.L. and Edwards, R.H. (2005) Sorting of vesicular monoamine transporter 2 to the regulated secretory pathway confers the somatodendritic exocytosis of monoamines. Neuron 48, 619–633 Falkenburger, B.H., Barstow, K.L. and Mintz, I.M. (2001) Dendrodendritic inhibition through reversal of dopamine transport. Science 293, 2465–2470 Bergquist, F., Niazi, H.S. and Nissbrandt, H. (2002) Evidence for different exocytosis pathways in dendritic and terminal dopamine release in vivo. Brain Res. 950, 245–253 John, C.E. and Jones, S.R. (2006) Exocytotic release of dopamine in ventral tegmental area slices from C57BL/6 and dopamine transporter knockout mice. Neurochem. Int. 49, 737–745 Bjorklund, A. and Lindvall, O. (1975) Dopamine in dendrites of substantia nigra neurons: suggestions for a role in dendritic terminals. Brain Res. 83, 531–537 Nirenberg, M.J., Chan, J., Liu, Y., Edwards, R.H. and Pickel, V.M. (1996) Ultrastructural localization of the vesicular monoamine transporter-2 in midbrain dopaminergic neurons: potential sites for somatodendritic storage and release of dopamine. J. Neurosci. 16, 4135–4145 Pow, D.V. and Morris, J.F. (1989) Dendrites of hypothalamic magnocellular neurons release neurohypophysial peptides by exocytosis. Neuroscience 32, 435–439 de Kock, C.P., Burnashev, N., Lodder, J.C., Mansvelder, H.D. and Brussaard, A.B. (2004) NMDA receptors induce somatodendritic secretion in hypothalamic neurones of lactating female rats. J. Physiol. 561, 53–64 Drake, C.T., Terman, G.W., Simmons, M.L., Milner, T.A., Kunkel, D.D., Schwartzkroin, P.A. and Chavkin, C. (1994) Dynorphin opioids present in dentate granule cells may function as retrograde inhibitory neurotransmitters. J. Neurosci. 14, 3736–3750 Simmons, M.L., Terman, G.W., Gibbs, S.M. and Chavkin, C. (1995) L-type calcium channels mediate dynorphin neuropeptide release from dendrites but not axons of hippocampal granule cells. Neuron 14, 1265–1272 Lessmann, V., Gottmann, K. and Heumann, R. (1994) BDNF and NT-4/5 enhance glutamatergic synaptic transmission in cultured hippocampal neurones. NeuroReport 6, 21–25 Lohof, A.M., Ip, N.Y. and Poo, M.M. (1993) Potentiation of developing neuromuscular synapses by the neurotrophins NT-3 and BDNF. Nature 363, 350–353  C The

C 2013 Biochemical Society Authors Journal compilation 

1379

1380

Biochemical Society Transactions (2013) Volume 41, part 6

143

144 145

146

147

148

149

150

151

152

153

154

155

156

157

158

159

160

161

162 163

164

 C The

Tanaka, J., Horiike, Y., Matsuzaki, M., Miyazaki, T., Ellis-Davies, G.C. and Kasai, H. (2008) Protein synthesis and neurotrophin-dependent structural plasticity of single dendritic spines. Science 319, 1683–1687 Minichiello, L. (2009) TrkB signalling pathways in LTP and learning. Nat. Rev. Neurosci. 10, 850–860 Yoshii, A. and Constantine-Paton, M. (2010) Postsynaptic BDNF–TrkB signaling in synapse maturation, plasticity, and disease. Dev. Neurobiol. 70, 304–322 Altar, C.A., Cai, N., Bliven, T., Juhasz, M., Conner, J.M., Acheson, A.L., Lindsay, R.M. and Wiegand, S.J. (1997) Anterograde transport of brain-derived neurotrophic factor and its role in the brain. Nature 389, 856–860 Conner, J.M., Lauterborn, J.C., Yan, Q., Gall, C.M. and Varon, S. (1997) Distribution of brain-derived neurotrophic factor (BDNF) protein and mRNA in the normal adult rat CNS: evidence for anterograde axonal transport. J. Neurosci. 17, 2295–2313 von Bartheld, C.S., Byers, M.R., Williams, R. and Bothwell, M. (1996) Anterograde transport of neurotrophins and axodendritic transfer in the developing visual system. Nature 379, 830–833 Dean, C., Liu, H., Dunning, F.M., Chang, P.Y., Jackson, M.B. and Chapman, E.R. (2009) Synaptotagmin-IV modulates synaptic function and long-term potentiation by regulating BDNF release. Nat. Neurosci. 12, 767–776 Hartmann, M., Heumann, R. and Lessmann, V. (2001) Synaptic secretion of BDNF after high-frequency stimulation of glutamatergic synapses. EMBO J. 20, 5887–5897 Kolarow, R., Brigadski, T. and Lessmann, V. (2007) Postsynaptic secretion of BDNF and NT-3 from hippocampal neurons depends on calcium calmodulin kinase II signaling and proceeds via delayed fusion pore opening. J. Neurosci. 27, 10350–10364 Kuczewski, N., Porcher, C., Ferrand, N., Fiorentino, H., Pellegrino, C., Kolarow, R., Lessmann, V., Medina, I. and Gaiarsa, J.L. (2008) Backpropagating action potentials trigger dendritic release of BDNF during spontaneous network activity. J. Neurosci. 28, 7013–7023 Matsuda, N., Lu, H., Fukata, Y., Noritake, J., Gao, H., Mukherjee, S., Nemoto, T., Fukata, M. and Poo, M.M. (2009) Differential activity-dependent secretion of brain-derived neurotrophic factor from axon and dendrite. J. Neurosci. 29, 14185–14198 Yoshihara, M., Adolfsen, B., Galle, K.T. and Littleton, J.T. (2005) Retrograde signaling by Syt 4 induces presynaptic release and synapse-specific growth. Science 310, 858–863 Barber, C.F., Jorquera, R.A., Melom, J.E. and Littleton, J.T. (2009) Postsynaptic regulation of synaptic plasticity by synaptotagmin 4 requires both C2 domains. J. Cell Biol. 187, 295–310 Harms, K.J. and Craig, A.M. (2005) Synapse composition and organization following chronic activity blockade in cultured hippocampal neurons. J. Comp. Neurol. 490, 72–84 Schoch, S., Deak, F., Konigstorfer, A., Mozhayeva, M., Sara, Y., Sudhof, T.C. and Kavalali, E.T. (2001) SNARE function analyzed in synaptobrevin/VAMP knockout mice. Science 294, 1117–1122 Washbourne, P., Thompson, P.M., Carta, M., Costa, E.T., Mathews, J.R., Lopez-Bendito, G., Molnar, Z., Becher, M.W., Valenzuela, C.F., Partridge, L.D. and Wilson, M.C. (2002) Genetic ablation of the t-SNARE SNAP-25 distinguishes mechanisms of neuroexocytosis. Nat. Neurosci. 5, 19–26 Burgo, A., Sotirakis, E., Simmler, M.C., Verraes, A., Chamot, C., Simpson, J.C., Lanzetti, L., Proux-Gillardeaux, V. and Galli, T. (2009) Role of Varp, a Rab21 exchange factor and TI-VAMP/VAMP7 partner, in neurite growth. EMBO Rep. 10, 1117–1124 Martinez-Arca, S., Coco, S., Mainguy, G., Schenk, U., Alberts, P., Bouille, P., Mezzina, M., Prochiantz, A., Matteoli, M., Louvard, D. and Galli, T. (2001) A common exocytotic mechanism mediates axonal and dendritic outgrowth. J. Neurosci. 21, 3830–3838 Lledo, P.M., Zhang, X., Sudhof, T.C., Malenka, R.C. and Nicoll, R.A. (1998) Postsynaptic membrane fusion and long-term potentiation. Science 279, 399–403 Anggono, V. and Huganir, R.L. (2012) Regulation of AMPA receptor trafficking and synaptic plasticity. Curr. Opin. Neurobiol. 22, 461–469 Shepherd, J.D. and Huganir, R.L. (2007) The cell biology of synaptic plasticity: AMPA receptor trafficking. Annu. Rev. Cell Dev. Biol. 23, 613–643 Kopec, C.D., Li, B., Wei, W., Boehm, J. and Malinow, R. (2006) Glutamate receptor exocytosis and spine enlargement during chemically induced long-term potentiation. J. Neurosci. 26, 2000–2009 C 2013 Biochemical Society Authors Journal compilation 

165

166

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184 185

186

Kopec, C.D., Real, E., Kessels, H.W. and Malinow, R. (2007) GluR1 links structural and functional plasticity at excitatory synapses. J. Neurosci. 27, 13706–13718 Park, M., Penick, E.C., Edwards, J.G., Kauer, J.A. and Ehlers, M.D. (2004) Recycling endosomes supply AMPA receptors for LTP. Science 305, 1972–1975 Park, M., Salgado, J.M., Ostroff, L., Helton, T.D., Robinson, C.G., Harris, K.M. and Ehlers, M.D. (2006) Plasticity-induced growth of dendritic spines by exocytic trafficking from recycling endosomes. Neuron 52, 817–830 Yang, Y., Wang, X.B., Frerking, M. and Zhou, Q. (2008) Spine expansion and stabilization associated with long-term potentiation. J. Neurosci. 28, 5740–5751 Ehlers, M.D. (2000) Reinsertion or degradation of AMPA receptors determined by activity-dependent endocytic sorting. Neuron 28, 511–525 Beattie, E.C., Carroll, R.C., Yu, X., Morishita, W., Yasuda, H., von Zastrow, M. and Malenka, R.C. (2000) Regulation of AMPA receptor endocytosis by a signaling mechanism shared with LTD. Nat. Neurosci. 3, 1291–1300 Carroll, R.C., Beattie, E.C., Xia, H., Luscher, C., Altschuler, Y., Nicoll, R.A., Malenka, R.C. and von Zastrow, M. (1999) Dynamin-dependent endocytosis of ionotropic glutamate receptors. Proc. Natl. Acad. Sci. U.S.A. 96, 14112–14117 Luscher, C., Xia, H., Beattie, E.C., Carroll, R.C., von Zastrow, M., Malenka, R.C. and Nicoll, R.A. (1999) Role of AMPA receptor cycling in synaptic transmission and plasticity. Neuron 24, 649–658 Granger, A.J., Shi, Y., Lu, W., Cerpas, M. and Nicoll, R.A. (2013) LTP requires a reserve pool of glutamate receptors independent of subunit type. Nature 493, 495–500 Maletic-Savatic, M. and Malinow, R. (1998) Calcium-evoked dendritic exocytosis in cultured hippocampal neurons. Part I: trans-Golgi network-derived organelles undergo regulated exocytosis. J. Neurosci. 18, 6803–6813 Maletic-Savatic, M., Koothan, T. and Malinow, R. (1998) Calcium-evoked dendritic exocytosis in cultured hippocampal neurons. Part II: mediation by calcium/calmodulin-dependent protein kinase II. J. Neurosci. 18, 6814–6821 Araki, Y., Lin, D.T. and Huganir, R.L. (2010) Plasma membrane insertion of the AMPA receptor GluA2 subunit is regulated by NSF binding and Q/R editing of the ion pore. Proc. Natl. Acad. Sci. U.S.A. 107, 11080–11085 Jaskolski, F., Mayo-Martin, B., Jane, D. and Henley, J.M. (2009) Dynamin-dependent membrane drift recruits AMPA receptors to dendritic spines. J. Biol. Chem. 284, 12491–12503 Kennedy, M.J., Davison, I.G., Robinson, C.G. and Ehlers, M.D. (2010) Syntaxin-4 defines a domain for activity-dependent exocytosis in dendritic spines. Cell 141, 524–535 Lin, D.T., Makino, Y., Sharma, K., Hayashi, T., Neve, R., Takamiya, K. and Huganir, R.L. (2009) Regulation of AMPA receptor extrasynaptic insertion by 4.1N, phosphorylation and palmitoylation. Nat. Neurosci. 12, 879–887 Makino, H. and Malinow, R. (2009) AMPA receptor incorporation into synapses during LTP: the role of lateral movement and exocytosis. Neuron 64, 381–390 Patterson, M.A., Szatmari, E.M. and Yasuda, R. (2010) AMPA receptors are exocytosed in stimulated spines and adjacent dendrites in a Ras-ERK-dependent manner during long-term potentiation. Proc. Natl. Acad. Sci. U.S.A. 107, 15951–15956 Yudowski, G.A., Puthenveedu, M.A., Leonoudakis, D., Panicker, S., Thorn, K.S., Beattie, E.C. and von Zastrow, M. (2007) Real-time imaging of discrete exocytic events mediating surface delivery of AMPA receptors. J. Neurosci. 27, 11112–11121 Petrini, E.M., Lu, J., Cognet, L., Lounis, B., Ehlers, M.D. and Choquet, D. (2009) Endocytic trafficking and recycling maintain a pool of mobile surface AMPA receptors required for synaptic potentiation. Neuron 63, 92–105 Borgdorff, A.J. and Choquet, D. (2002) Regulation of AMPA receptor lateral movements. Nature 417, 649–653 Tardin, C., Cognet, L., Bats, C., Lounis, B. and Choquet, D. (2003) Direct imaging of lateral movements of AMPA receptors inside synapses. EMBO J. 22, 4656–4665 Ehlers, M.D., Heine, M., Groc, L., Lee, M.C. and Choquet, D. (2007) Diffusional trapping of GluR1 AMPA receptors by input-specific synaptic activity. Neuron 54, 447–460

Dendritic trafficking for neuronal growth and plasticity

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

207

Lu, J., Helton, T.D., Blanpied, T.A., Racz, B., Newpher, T.M., Weinberg, R.J. and Ehlers, M.D. (2007) Postsynaptic positioning of endocytic zones and AMPA receptor cycling by physical coupling of dynamin-3 to Homer. Neuron 55, 874–889 Lan, J.Y., Skeberdis, V.A., Jover, T., Grooms, S.Y., Lin, Y., Araneda, R.C., Zheng, X., Bennett, M.V. and Zukin, R.S. (2001) Protein kinase C modulates NMDA receptor trafficking and gating. Nat. Neurosci. 4, 382–390 Lau, C.G., Takayasu, Y., Rodenas-Ruano, A., Paternain, A.V., Lerma, J., Bennett, M.V. and Zukin, R.S. (2010) SNAP-25 is a target of protein kinase C phosphorylation critical to NMDA receptor trafficking. J. Neurosci. 30, 242–254 Suh, Y.H., Terashima, A., Petralia, R.S., Wenthold, R.J., Isaac, J.T., Roche, K.W. and Roche, P.A. (2010) A neuronal role for SNAP-23 in postsynaptic glutamate receptor trafficking. Nat. Neurosci. 13, 338–343 Paumet, F., Le Mao, J., Martin, S., Galli, T., David, B., Blank, U. and Roa, M. (2000) Soluble NSF attachment protein receptors (SNAREs) in RBL-2H3 mast cells: functional role of syntaxin 4 in exocytosis and identification of a vesicle-associated membrane protein 8-containing secretory compartment. J. Immunol. 164, 5850–5857 Lu, W., Man, H., Ju, W., Trimble, W.S., MacDonald, J.F. and Wang, Y.T. (2001) Activation of synaptic NMDA receptors induces membrane insertion of new AMPA receptors and LTP in cultured hippocampal neurons. Neuron 29, 243–254 Gerges, N.Z., Backos, D.S., Rupasinghe, C.N., Spaller, M.R. and Esteban, J.A. (2006) Dual role of the exocyst in AMPA receptor targeting and insertion into the postsynaptic membrane. EMBO J. 25, 1623–1634 Wang, Z., Edwards, J.G., Riley, N., Provance, Jr, D.W., Karcher, R., Li, X.D., Davison, I.G., Ikebe, M., Mercer, J.A., Kauer, J.A. and Ehlers, M.D. (2008) Myosin Vb mobilizes recycling endosomes and AMPA receptors for postsynaptic plasticity. Cell 135, 535–548 Li, X.D., Jung, H.S., Mabuchi, K., Craig, R. and Ikebe, M. (2006) The globular tail domain of myosin Va functions as an inhibitor of the myosin Va motor. J. Biol. Chem. 281, 21789–21798 Li, X.D., Jung, H.S., Wang, Q., Ikebe, R., Craig, R. and Ikebe, M. (2008) The globular tail domain puts on the brake to stop the ATPase cycle of myosin Va. Proc. Natl. Acad. Sci. U.S.A. 105, 1140–1145 Hales, C.M., Griner, R., Hobdy-Henderson, K.C., Dorn, M.C., Hardy, D., Kumar, R., Navarre, J., Chan, E.K., Lapierre, L.A. and Goldenring, J.R. (2001) Identification and characterization of a family of Rab11-interacting proteins. J. Biol. Chem. 276, 39067–39075 Lindsay, A.J. and McCaffrey, M.W. (2002) Rab11-FIP2 functions in transferrin recycling and associates with endosomal membranes via its COOH-terminal domain. J. Biol. Chem. 277, 27193–27199 Lapierre, L.A., Kumar, R., Hales, C.M., Navarre, J., Bhartur, S.G., Burnette, J.O., Provance, Jr, D.W., Mercer, J.A., Bahler, M. and Goldenring, J.R. (2001) Myosin Vb is associated with plasma membrane recycling systems. Mol. Biol. Cell 12, 1843–1857 Correia, S.S., Bassani, S., Brown, T.C., Lise, M.F., Backos, D.S., El-Husseini, A., Passafaro, M. and Esteban, J.A. (2008) Motor protein-dependent transport of AMPA receptors into spines during long-term potentiation. Nat. Neurosci. 11, 457–466 Schnell, E. and Nicoll, R.A. (2001) Hippocampal synaptic transmission and plasticity are preserved in myosin Va mutant mice. J. Neurophysiol. 85, 1498–1501 Carroll, R.C., Beattie, E.C., von Zastrow, M. and Malenka, R.C. (2001) Role of AMPA receptor endocytosis in synaptic plasticity. Nat. Rev. Neurosci. 2, 315–324 Lavezzari, G., McCallum, J., Dewey, C.M. and Roche, K.W. (2004) Subunit-specific regulation of NMDA receptor endocytosis. J. Neurosci. 24, 6383–6391 Lee, S.H., Simonetta, A. and Sheng, M. (2004) Subunit rules governing the sorting of internalized AMPA receptors in hippocampal neurons. Neuron 43, 221–236 Lin, J.W., Ju, W., Foster, K., Lee, S.H., Ahmadian, G., Wyszynski, M., Wang, Y.T. and Sheng, M. (2000) Distinct molecular mechanisms and divergent endocytotic pathways of AMPA receptor internalization. Nat. Neurosci. 3, 1282–1290 Prybylowski, K., Chang, K., Sans, N., Kan, L., Vicini, S. and Wenthold, R.J. (2005) The synaptic localization of NR2B-containing NMDA receptors is controlled by interactions with PDZ proteins and AP-2. Neuron 47, 845–857 Roche, K.W., Standley, S., McCallum, J., Dune Ly, C., Ehlers, M.D. and Wenthold, R.J. (2001) Molecular determinants of NMDA receptor internalization. Nat. Neurosci. 4, 794–802

208

209

210

211 212

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

228

Lavezzari, G., McCallum, J., Lee, R. and Roche, K.W. (2003) Differential binding of the AP-2 adaptor complex and PSD-95 to the C-terminus of the NMDA receptor subunit NR2B regulates surface expression. Neuropharmacology 45, 729–737 Lee, S.H., Liu, L., Wang, Y.T. and Sheng, M. (2002) Clathrin adaptor AP2 and NSF interact with overlapping sites of GluR2 and play distinct roles in AMPA receptor trafficking and hippocampal LTD. Neuron 36, 661–674 Gray, N.W., Fourgeaud, L., Huang, B., Chen, J., Cao, H., Oswald, B.J., Hemar, A. and McNiven, M.A. (2003) Dynamin 3 is a component of the postsynapse, where it interacts with mGluR5 and Homer. Curr. Biol. 13, 510–515 Campelo, F. and Malhotra, V. (2012) Membrane fission: the biogenesis of transport carriers. Annu. Rev. Biochem. 81, 407–427 Palmer, C.L., Lim, W., Hastie, P.G., Toward, M., Korolchuk, V.I., Burbidge, S.A., Banting, G., Collingridge, G.L., Isaac, J.T. and Henley, J.M. (2005) Hippocalcin functions as a calcium sensor in hippocampal LTD. Neuron 47, 487–494 Snyder, E.M., Nong, Y., Almeida, C.G., Paul, S., Moran, T., Choi, E.Y., Nairn, A.C., Salter, M.W., Lombroso, P.J., Gouras, G.K. and Greengard, P. (2005) Regulation of NMDA receptor trafficking by amyloid-β. Nat. Neurosci. 8, 1051–1058 Goebel-Goody, S.M., Baum, M., Paspalas, C.D., Fernandez, S.M., Carty, N.C., Kurup, P. and Lombroso, P.J. (2012) Therapeutic implications for striatal-enriched protein tyrosine phosphatase (STEP) in neuropsychiatric disorders. Pharmacol. Rev. 64, 65–87 Shi, S.H., Hayashi, Y., Petralia, R.S., Zaman, S.H., Wenthold, R.J., Svoboda, K. and Malinow, R. (1999) Rapid spine delivery and redistribution of AMPA receptors after synaptic NMDA receptor activation. Science 284, 1811–1816 Man, H.Y., Lin, J.W., Ju, W.H., Ahmadian, G., Liu, L., Becker, L.E., Sheng, M. and Wang, Y.T. (2000) Regulation of AMPA receptor-mediated synaptic transmission by clathrin-dependent receptor internalization. Neuron 25, 649–662 Blanpied, T.A., Scott, D.B. and Ehlers, M.D. (2002) Dynamics and regulation of clathrin coats at specialized endocytic zones of dendrites and spines. Neuron 36, 435–449 Racz, B., Blanpied, T.A., Ehlers, M.D. and Weinberg, R.J. (2004) Lateral organization of endocytic machinery in dendritic spines. Nat. Neurosci. 7, 917–918 Yao, P.J., Bushlin, I. and Petralia, R.S. (2006) Partially overlapping distribution of epsin1 and HIP1 at the synapse: analysis by immunoelectron microscopy. J. Comp. Neurol. 494, 368–379 Petralia, R.S., Wang, Y.X. and Wenthold, R.J. (2003) Internalization at glutamatergic synapses during development. Eur. J. Neurosci. 18, 3207–3217 Brakeman, P.R., Lanahan, A.A., O’Brien, R., Roche, K., Barnes, C.A., Huganir, R.L. and Worley, P.F. (1997) Homer: a protein that selectively binds metabotropic glutamate receptors. Nature 386, 284–288 Tu, J.C., Xiao, B., Yuan, J.P., Lanahan, A.A., Leoffert, K., Li, M., Linden, D.J. and Worley, P.F. (1998) Homer binds a novel proline-rich motif and links group 1 metabotropic glutamate receptors with IP3 receptors. Neuron 21, 717–726 Xiao, B., Tu, J.C., Petralia, R.S., Yuan, J.P., Doan, A., Breder, C.D., Ruggiero, A., Lanahan, A.A., Wenthold, R.J. and Worley, P.F. (1998) Homer regulates the association of group 1 metabotropic glutamate receptors with multivalent complexes of Homer-related, synaptic proteins. Neuron 21, 707–716 Gray, N.W., Kruchten, A.E., Chen, J. and McNiven, M.A. (2005) A dynamin-3 spliced variant modulates the actin/cortactin-dependent morphogenesis of dendritic spines. J. Cell Sci. 118, 1279–1290 Hayashi, M.K., Ames, H.M. and Hayashi, Y. (2006) Tetrameric hub structure of postsynaptic scaffolding protein Homer. J. Neurosci. 26, 8492–8501 Hayashi, M.K., Tang, C., Verpelli, C., Narayanan, R., Stearns, M.H., Xu, R.M., Li, H., Sala, C. and Hayashi, Y. (2009) The postsynaptic density proteins Homer and Shank form a polymeric network structure. Cell 137, 159–171 Grabrucker, A.M., Knight, M.J., Proepper, C., Bockmann, J., Joubert, M., Rowan, M., Nienhaus, G.U., Garner, C.C., Bowie, J.U., Kreutz, M.R., Gundelfinger, E.D. and Boeckers, T.M. (2011) Concerted action of zinc and ProSAP/Shank in synaptogenesis and synapse maturation. EMBO J. 30, 569–581 Baron, M.K., Boeckers, T.M., Vaida, B., Faham, S., Gingery, M., Sawaya, M.R., Salyer, D., Gundelfinger, E.D. and Bowie, J.U. (2006) An architectural framework that may lie at the core of the postsynaptic density. Science 311, 531–535  C The

C 2013 Biochemical Society Authors Journal compilation 

1381

1382

Biochemical Society Transactions (2013) Volume 41, part 6

229

230

231

232

233

234

235 236

237

238

239

240

241

242

243

244

245

246

247

248

 C The

Lyford, G.L., Yamagata, K., Kaufmann, W.E., Barnes, C.A., Sanders, L.K., Copeland, N.G., Gilbert, D.J., Jenkins, N.A., Lanahan, A.A. and Worley, P.F. (1995) Arc, a growth factor and activity-regulated gene, encodes a novel cytoskeleton-associated protein that is enriched in neuronal dendrites. Neuron 14, 433–445 Bramham, C.R., Worley, P.F., Moore, M.J. and Guzowski, J.F. (2008) The immediate early gene arc/arg3.1: regulation, mechanisms, and function. J. Neurosci. 28, 11760–11767 Chowdhury, S., Shepherd, J.D., Okuno, H., Lyford, G., Petralia, R.S., Plath, N., Kuhl, D., Huganir, R.L. and Worley, P.F. (2006) Arc/Arg3.1 interacts with the endocytic machinery to regulate AMPA receptor trafficking. Neuron 52, 445–459 Wu, J., Petralia, R.S., Kurushima, H., Patel, H., Jung, M.Y., Volk, L., Chowdhury, S., Shepherd, J.D., Dehoff, M., Li, Y. et al. (2011) Arc/Arg3.1 regulates an endosomal pathway essential for activity-dependent β-amyloid generation. Cell 147, 615–628 Shepherd, J.D., Rumbaugh, G., Wu, J., Chowdhury, S., Plath, N., Kuhl, D., Huganir, R.L. and Worley, P.F. (2006) Arc/Arg3.1 mediates homeostatic synaptic scaling of AMPA receptors. Neuron 52, 475–484 Reiter, E., Ahn, S., Shukla, A.K. and Lefkowitz, R.J. (2012) Molecular mechanism of β-arrestin-biased agonism at seven-transmembrane receptors. Annu. Rev. Pharmacol. Toxicol. 52, 179–197 Shenoy, S.K. and Lefkowitz, R.J. (2011) β-Arrestin-mediated receptor trafficking and signal transduction. Trends Pharmacol. Sci. 32, 521–533 Yudowski, G.A., Puthenveedu, M.A. and von Zastrow, M. (2006) Distinct modes of regulated receptor insertion to the somatodendritic plasma membrane. Nat. Neurosci. 9, 622–627 Yu, Y.J., Dhavan, R., Chevalier, M.W., Yudowski, G.A. and von Zastrow, M. (2010) Rapid delivery of internalized signaling receptors to the somatodendritic surface by sequence-specific local insertion. J. Neurosci. 30, 11703–11714 Blanpied, T.A., Scott, D.B. and Ehlers, M.D. (2003) Age-related regulation of dendritic endocytosis associated with altered clathrin dynamics. Neurobiol. Aging 24, 1095–1104 Lee, H.K., Kameyama, K., Huganir, R.L. and Bear, M.F. (1998) NMDA induces long-term synaptic depression and dephosphorylation of the GluR1 subunit of AMPA receptors in hippocampus. Neuron 21, 1151–1162 Esteban, J.A., Shi, S.H., Wilson, C., Nuriya, M., Huganir, R.L. and Malinow, R. (2003) PKA phosphorylation of AMPA receptor subunits controls synaptic trafficking underlying plasticity. Nat. Neurosci. 6, 136–143 Lee, H.K., Barbarosie, M., Kameyama, K., Bear, M.F. and Huganir, R.L. (2000) Regulation of distinct AMPA receptor phosphorylation sites during bidirectional synaptic plasticity. Nature 405, 955–959 Brown, T.C., Tran, I.C., Backos, D.S. and Esteban, J.A. (2005) NMDA receptor-dependent activation of the small GTPase Rab5 drives the removal of synaptic AMPA receptors during hippocampal LTD. Neuron 45, 81–94 Scott, D.B., Michailidis, I., Mu, Y., Logothetis, D. and Ehlers, M.D. (2004) Endocytosis and degradative sorting of NMDA receptors by conserved membrane-proximal signals. J. Neurosci. 24, 7096–7109 Prekeris, R., Foletti, D.L. and Scheller, R.H. (1999) Dynamics of tubulovesicular recycling endosomes in hippocampal neurons. J. Neurosci. 19, 10324–10337 Steiner, P., Sarria, J.C., Glauser, L., Magnin, S., Catsicas, S. and Hirling, H. (2002) Modulation of receptor cycling by neuron-enriched endosomal protein of 21 kD. J. Cell Biol. 157, 1197–1209 Yap, C.C., Wisco, D., Kujala, P., Lasiecka, Z.M., Cannon, J.T., Chang, M.C., Hirling, H., Klumperman, J. and Winckler, B. (2008) The somatodendritic endosomal regulator NEEP21 facilitates axonal targeting of L1/NgCAM. J. Cell Biol. 180, 827–842 Lasiecka, Z.M., Yap, C.C., Caplan, S. and Winckler, B. (2010) Neuronal early endosomes require EHD1 for L1/NgCAM trafficking. J. Neurosci. 30, 16485–16497 Steiner, P., Alberi, S., Kulangara, K., Yersin, A., Sarria, J.C., Regulier, E., Kasas, S., Dietler, G., Muller, D., Catsicas, S. and Hirling, H. (2005) Interactions between NEEP21, GRIP1 and GluR2 regulate sorting and recycling of the glutamate receptor subunit GluR2. EMBO J. 24, 2873–2884

C 2013 Biochemical Society Authors Journal compilation 

249

250

251 252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

Von Bartheld, C.S. and Altick, A.L. (2011) Multivesicular bodies in neurons: distribution, protein content, and trafficking functions. Prog. Neurobiol. 93, 313–340 Chicurel, M.E. and Harris, K.M. (1992) Three-dimensional analysis of the structure and composition of CA3 branched dendritic spines and their synaptic relationships with mossy fiber boutons in the rat hippocampus. J. Comp. Neurol. 325, 169–182 Pappas, G.D. and Purpura, D.P. (1961) Fine structure of dendrites in the superficial neocortical neuropil. Exp. Neurol. 4, 507–530 Paula-Barbosa, M.M., Mota Cardoso, R., Faria, R. and Cruz, C. (1978) Multivesicular bodies in cortical dendrites of two patients with Alzheimer’s disease. J. Neurol. Sci. 36, 259–264 Rind, H.B., Butowt, R. and von Bartheld, C.S. (2005) Synaptic targeting of retrogradely transported trophic factors in motoneurons: comparison of glial cell line-derived neurotrophic factor, brain-derived neurotrophic factor, and cardiotrophin-1 with tetanus toxin. J. Neurosci. 25, 539–549 Gorenstein, C., Bundman, M.C., Lew, P.J., Olds, J.L. and Ribak, C.E. (1985) Dendritic transport. I. Colchicine stimulates the transport of lysosomal enzymes from cell bodies to dendrites. J. Neurosci. 5, 2009–2017 Gorenstein, C. and Ribak, C.E. (1985) Dendritic transport. II. Somatofugal movement of neuronal lysosomes induced by colchicine: evidence for a novel transport system in dendrites. J. Neurosci. 5, 2018–2027 Gordon, M.K., Bench, K.G., Deanin, G.G. and Gordon, M.W. (1968) Histochemical and biochemical study of synaptic lysosomes. Nature 217, 523–527 Caviston, J.P., Zajac, A.L., Tokito, M. and Holzbaur, E.L. (2011) Huntingtin coordinates the dynein-mediated dynamic positioning of endosomes and lysosomes. Mol. Biol. Cell 22, 478–492 Loubery, S., Wilhelm, C., Hurbain, I., Neveu, S., Louvard, D. and Coudrier, E. (2008) Different microtubule motors move early and late endocytic compartments. Traffic 9, 492–509 Jordens, I., Fernandez-Borja, M., Marsman, M., Dusseljee, S., Janssen, L., Calafat, J., Janssen, H., Wubbolts, R. and Neefjes, J. (2001) The Rab7 effector protein RILP controls lysosomal transport by inducing the recruitment of dynein–dynactin motors. Curr. Biol. 11, 1680–1685 Brown, C.L., Maier, K.C., Stauber, T., Ginkel, L.M., Wordeman, L., Vernos, I. and Schroer, T.A. (2005) Kinesin-2 is a motor for late endosomes and lysosomes. Traffic 6, 1114–1124 Matsushita, M., Tanaka, S., Nakamura, N., Inoue, H. and Kanazawa, H. (2004) A novel kinesin-like protein, KIF1Bβ3 is involved in the movement of lysosomes to the cell periphery in non-neuronal cells. Traffic 5, 140–151 Dodson, M.W., Zhang, T., Jiang, C., Chen, S. and Guo, M. (2012) Roles of the Drosophila LRRK2 homolog in Rab7-dependent lysosomal positioning. Hum. Mol. Genet. 21, 1350–1363 Hoepfner, S., Severin, F., Cabezas, A., Habermann, B., Runge, A., Gillooly, D., Stenmark, H. and Zerial, M. (2005) Modulation of receptor recycling and degradation by the endosomal kinesin KIF16B. Cell 121, 437–450 Shehata, M., Matsumura, H., Okubo-Suzuki, R., Ohkawa, N. and Inokuchi, K. (2012) Neuronal stimulation induces autophagy in hippocampal neurons that is involved in AMPA receptor degradation after chemical long-term depression. J. Neurosci. 32, 10413–10422 Ehlers, M.D. (2003) Activity level controls postsynaptic composition and signaling via the ubiquitin–proteasome system. Nat. Neurosci. 6, 231–242 Harris, H. and Rubinsztein, D.C. (2011) Control of autophagy as a therapy for neurodegenerative disease. Nat. Rev. Neurol. 8, 108–117

Received 14 May 2013 doi:10.1042/BST20130081

Copyright of Biochemical Society Transactions is the property of Portland Press Ltd. and its content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written permission. However, users may print, download, or email articles for individual use.

Dendritic trafficking for neuronal growth and plasticity.

Among the largest cells in the body, neurons possess an immense surface area and intricate geometry that poses many unique cell biological challenges...
1004KB Sizes 0 Downloads 0 Views