首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
2.
3.
4.
Reductions in hippocampal neurite complexity and synaptic plasticity are believed to contribute to the progressive impairment in episodic memory and the mild cognitive decline that occur particularly in the early stages of Alzheimer''s disease (AD). Despite the functional and therapeutic importance for patients with AD, intervention to rescue or normalize dendritic elaboration and synaptic plasticity is scarcely provided. Here we show that overexpression of neuritin, an activity-dependent protein, promoted neurite outgrowth and maturation of synapses in parallel with enhanced basal synaptic transmission in cultured hippocampal neurons. Importantly, exogenous application of recombinant neuritin fully restored dendritic complexity as well as spine density in hippocampal neurons prepared from Tg2576 mice, whereas it did not affect neurite branching of neurons from their wild-type littermates. We also showed that soluble recombinant neuritin, when chronically infused into the brains of Tg2576 mice, normalized synaptic plasticity in acute hippocampal slices, leading to intact long-term potentiation. By revealing the protective actions of soluble neuritin against AD-related neural defects, we provide a potential therapeutic approach for patients with AD.Efficient neuronal communications through synapses are crucial for normal brain functions, whereas alterations in synapse numbers, dendritic spine morphology, and dendritic complexity are thought to be reflected by different forms of synaptic plasticity and are also causally associated with a variety of neurological disorders.1, 2, 3, 4, 5 For example, synapse loss and neurite atrophy are the major neurobiological substrates underlying memory impairment in neurodegenerative diseases such as Alzheimer''s disease (AD).6, 7 The increased dendritic mislocalization of hyperphosphorylated tau protein, a microtubule-associated protein enriched at axons of mature neurons,8 and abundance of soluble oligomeric forms of β-amyloid (Aβ) appear to cause the synaptic defects and disruption of synaptic plasticity involving the progression of AD pathology.6, 9, 10 The apparent decreases in neurotrophic factors observed in brains of patients with AD11 have prompted several trials for administration of neurotrophic factors, such as brain-derived neurotrophic factor (BDNF), to attenuate and possibly reverse synaptic defects.11, 12, 13 However, the truncation or decreased expression of its cognate receptors in AD brains have limited their potential usage as AD therapeutics.12, 14, 15Neuritin, also known as the candidate plasticity gene 15, was originally identified in a screening study for activity-regulated genes and was subsequently found to be one of the signaling molecules downstream to BDNF and its receptor tropomyosin-related kinase receptor type B.16, 17 Ensuing studies indicated that neuritin could also be induced by experimental seizure or by normal life experiences, such as sensory stimulation and exercise.17, 18, 19, 20, 21, 22 Located in the 6p24-p25 interval on chromosome 6,23 the neuritin gene encodes a small, highly conserved protein containing a secretory signal sequence at the N-terminus and a consensus sequence for glycosylphosphatidylinositol (GPI) at the C-terminus.16 This GPI linkage enables neuritin to anchor at cell surfaces, and upon cleavage of GPI by phospholipase the resultant soluble neuritin is released into the extracellular space.16, 20, 24, 25, 26During embryonic neural development, neuritin is mainly expressed in brain regions that undergo a rapid proliferation of neuronal progenitor pools, suggesting a protective role of neuritin for differentiated neurons.26, 27 Interestingly, the expression level of neuritin remains elevated after birth or even increases, especially in brain regions presumably exhibiting high neural activity and synaptic plasticity, such as the hippocampus, visual cortex, and external granular layer of the cerebellum.16, 19, 20, 26 In addition, neuritin promotes neuritic arbor growth and synaptic formation.16, 20, 24, 25, 28, 29, 30, 31 Although various studies have suggested these potent neuritogenetic activities of neuritin, the contribution of neuritin expression to or its effectiveness against neurodegenerative diseases that display neurite atrophy and synapse loss has been largely unexplored.Here we determined that neuritin expression increased neurite complexity and promoted the maturation of individual spines in cultured hippocampal neurons. Consistent with these findings, basal synaptic transmission was enhanced by transient expression of neuritin. Importantly, when exogenously applied, the soluble neuritin peptide rescued the dendrite complexity of neurons prepared from Tg2576 mice, a transgenic mouse model of AD, such that the complexity was comparable to that in wild-type (WT) mice and also normalized synaptic plasticity in the hippocampus of the Tg2576 mice. Taken together, these results suggest that neuritin, particularly a soluble form of neuritin, reverses synaptic defects manifest in Tg2576 mice and that manipulations to increase neuritin levels may be beneficial therapeutic approaches in AD.  相似文献   

5.
Elucidating the temporal order of silencing   总被引:1,自引:0,他引:1  
Izaurralde E 《EMBO reports》2012,13(8):662-663
  相似文献   

6.
Synapse formation is a highly regulated process that requires the coordination of many cell biological events. Decades of research have identified a long list of molecular components involved in assembling a functioning synapse. Yet how the various steps, from transporting synaptic components to adhering synaptic partners and assembling the synaptic structure, are regulated and precisely executed during development and maintenance is still unclear. With the improvement of imaging and molecular tools, recent work in vertebrate and invertebrate systems has provided important insight into various aspects of presynaptic development, maintenance, and trans-synaptic signals, thereby increasing our understanding of how extrinsic organizers and intracellular mechanisms contribute to presynapse formation.Chemical synapses are highly specialized, asymmetric intercellular junction structures that are the basic units of neuronal communication. Proper development of synapses determines appropriate connectivity for the assembly of functional neuronal circuits. Synaptic circuits arise during development through a series of intricate steps (Waites et al., 2005; McAllister, 2007; Jin and Garner, 2008). First, spatiotemporal cues guide axons through complex cellular environments to contact their appropriate postsynaptic targets. At their destination, synapse formation is specified and initiated through adhesive interactions between synaptic partner cells or by local diffusible signaling molecules. Stabilization of intercellular contacts and assembly into functional synapses involves cytoskeletal rearrangements, aggregation, and insertion of pre- and postsynaptic components at nascent synaptic sites. Maturation and modulation of these newly formed synapses can then occur by altering the organization or composition of synaptic proteins and post-translational modifications to achieve its required physiological responsiveness (Budnik, 1996; Lee and Sheng, 2000). Conversely, retraction of contacts and elimination of inappropriate synaptic proteins help to refine the neuronal circuitry (Goda and Davis, 2003; Sanes and Yamagata, 2009).Over the last decade, new insights have furthered our understanding of synapse development through the identification of new molecular players and by advanced imaging technology that has allowed for high-resolution inspection of the dynamics and relative positions of synaptic proteins. This review will highlight recent results on the development of presynaptic specializations, and the roles of trans-synaptic organizers, intracellular synaptic proteins, and the cytoskeleton during the formation and maintenance of synapses.

Axonal transport of synaptic vesicle and active zone proteins

After cell fate determination and morphogenesis, neurons continue to differentiate by entering the phase of synapse formation. Most synaptic material required for this process is synthesized in the cell body of neurons and transported to synapses by microtubule (MT)-based molecular motors (Fig. 1). MTs are intrinsically polarized filaments with a plus and a minus end (Fig. 1 B). MT-based molecular motors use this polarity to transport cargoes to specific cellular locations. Examination of MTs by electron microscopy in dissociated cultured neurons showed that the organizations of MTs is different in axon and dendrite (Baas et al., 1988, 2006). In axons, all microtubules have their minus ends oriented toward the cell body and their plus ends extend distally. On the contrary, the MT polarity in dendrites is mixed. Recent studies tracking the movement of end-binding MT-capping proteins confirmed these results in vivo. Specifically, axonal MTs are uniformly organized with their plus ends pointing distally in all organisms. Dendrites of vertebrate neurons show more plus end–out MTs in vivo, whereas flies and worms have more minus end–out MTs in dendrites (Stepanova et al., 2003; Rolls et al., 2007; Stone et al., 2008).Open in a separate windowFigure 1.Regulatory steps during polarized motor-based transport of synaptic material. (A) At the Golgi apparatus, synaptic proteins have to be sorted into appropriate vesicles. These vesicles and other cargo such as mitochondria get loaded onto specific motor proteins. (B) Establishment of proper microtubule polarity along the axon determines anterograde and retrograde trafficking by plus end– and minus end–directed motor proteins such as kinesins and dynein. (C) At the appropriate destination, motor-cargo unloading occurs in a regulated fashion to achieve the appropriate distribution of synaptic boutons. At synapses, synaptic vesicle precursors give rise to mature synaptic vesicles. Proteins required for the SV cycle and trans-synaptic adhesion coalesce into the active zone (AZ) underneath the plasma membrane juxtaposed against the postsynaptic membrane.Does the difference in microtubule organization and polarity help to segregate synaptic cargoes between axons and dendrites? Recent studies have started to identify some molecules that create these differences in MT polarity in different neuronal subcellular compartments and show how disruption of their function affects synapse formation. For example, a recent paper showed that kinesin-1 is required to establish the predominantly minus end–out organization in the dendrites of Caenorhabditis elegans motor neurons (Yan et al., 2013). In kinesin-1/unc-116 mutants, dendrites adopt the axon-like MT polarity causing presynaptic cargoes to mislocalize into dendrites (Seeger and Rice, 2010; Yan et al., 2013). Similarly, loss of the MT-binding CRMP protein UNC-33 or the actin–spectrin adaptor protein ankyrin/UNC-44 in worms also results in MT polarity defects, which also results in ectopic localization of synaptic vesicles and active zone proteins into dendrites (Maniar et al., 2012). These results support the idea that MT polarity ensures the faithful targeting of presynaptic components to the axon. However, another way motors can distinguish between axons and dendrites is through MT-associated proteins (MAPs). In a recent study, Banker and colleagues showed that plus end–orienting kinesins can differentiate axon and dendrite, likely due to specific MT-binding proteins in these compartments (Huang and Banker, 2012).The direct regulation of motor activity by MTs or synaptic vesicle–associated proteins is likely to contribute to the trafficking of synaptic cargoes. Doublecortin, a MAP, binds to kinesin-3/KIF1A to affect the trafficking of the synaptic vesicle protein, synaptobrevin, in hippocampal neurons by altering the affinity of ADP-bound KIF1A to MTs (Liu et al., 2012). The Rab3 guanine nucleotide exchange factor, DENN/MADD, functions as an adaptor between kinesin-3 and GTP-Rab3–containing synaptic vesicles to promote the trafficking of synaptic vesicles in the axon (Niwa et al., 2008).Precise regulation of motor-based transport ensures that synaptic cargoes are delivered to and maintained at synapses. Several recent studies have provided evidence that two postmitotic cyclin-dependent kinases are important regulators of anterograde and retrograde trafficking of presynaptic cargoes. The kinase CDK-5 is required in many aspects of nervous system function. In the context of presynaptic development and function, CDK-5 has been shown to regulate the transport of synaptic vesicles and dense core vesicles, which contain neuropeptides, by inhibiting a dynein-mediated pathway that mobilizes presynaptic components to the somatodendritic compartments in C. elegans neurons (Ou et al., 2010; Goodwin et al., 2012). A paralogue of CDK-5, the PCT-1 kinase acts in a partially redundant pathway to prevent the mislocalization of presynaptic material to dendrites. In animals lacking both kinases or their activators, synaptic cargoes completely mislocalize to the dendrites, leaving an “empty” axon (Ou et al., 2010). Vertebrate CDK-5 also plays profound roles in the regulation of synaptic vesicle pools by modifying Ca2+ channels. Genetic ablation or pharmacological inhibition of CDK-5 increases the pool of synaptic vesicles that are docked at the active zone, termed the readily releasable pool, and potentiates synaptic function (Kim and Ryan, 2010, 2013). These results suggest that CDK-5 and its paralogue control local and global vesicle pools. Regulation of the exchange between these pools can affect membrane trafficking at presynaptic terminals as well as the overall polarity of neurons.To form synapses at defined locations, cargoes not only need to know how to “get on” the transport system but also need to know where to precisely “get off” at their destination (Fig. 1 C). Loss of a conserved small G-protein of the Arf-like family, ARL-8, in C. elegans, resulted in premature exit of synaptic cargoes during transport and showed ectopic aggregations of synaptic vesicles in the proximal axon. This causes a reduction in the number but an increase in the size of synapses (Klassen et al., 2010). ARL-8 localizes to both stable and trafficking synaptic vesicles and promotes trafficking by increasing kinesin-3 activity and suppressing aggregation-induced stoppage of synaptic cargoes along the axon (Wu et al., 2013). Hence, the balance between motor activity and aggregation propensity of trafficking cargoes may determine the number, size, and location of presynaptic terminals. Interestingly, the small GTPase Rab3, which normally associates with synaptic vesicles, has recently been shown to affect the distribution of active zone proteins at fly neuromuscular junction (NMJ) synapses, further suggesting that the trafficking of synaptic vesicles and formation of active zones are linked (Graf et al., 2009).Besides synaptic material, another major organelle cargo that is often present at the presynaptic terminal is mitochondria. The Milton–Miro complex functions as an adaptor between kinesin-1 and mitochondria to support axonal transport of mitochondria. Interestingly, the coupling of the Milton–Miro complex to kinesin is regulated by Ca2+ (Macaskill et al., 2009; Wang and Schwarz, 2009), providing a mechanism for neuronal activity controlling transport of mitochondria along the axon.Previous studies have suggested that components of the presynaptic active zone are transported in a preassembled form by Piccolo-Bassoon transport vesicles (PTVs) that may contain multiple components required to build a synapse (Zhai et al., 2001; Shapira et al., 2003). Recent studies found that Golgi-derived PTVs contain many active zone proteins including Piccolo, Bassoon, RIM1α, and ELKS2/CAST, but lack another active zone component, Munc-13, which may exit the Golgi on separate vesicles (Maas et al., 2012). Packing of various active zone components that have the propensity to self-assemble into separate vesicles may contribute a way to control synaptogenesis. This is interesting in light of the finding that Munc-13 can function as a protein scaffold for Bassoon and ELKS2 (Wang et al., 2009). The link between trafficking of synaptic vesicle and active zone components is not well understood. In vivo time-lapse imaging of synaptic vesicle and active zone trafficking showed that these components, possibly in the form of dense core vesicles, could be trafficked together in C. elegans neurons, suggestive of prepackaged presynaptic material during transport (Wu et al., 2013). Taken together, axonal transport of synaptic components is a necessary step for synapse formation and maintenance. The regulation of MTs, molecular motors, and synaptic cargoes ensure the targeting of appropriate proteins to synapses.

Role of the actin cytoskeleton in presynaptic assembly

Although MT-mediated transport is critical for long-range trafficking, actin-based mechanisms often organize local protein complexes in subcellular domains. A large body of work has described the role of the actin cytoskeleton in postsynaptic structure and function (Schubert and Dotti, 2007; Hotulainen and Hoogenraad, 2010). We will focus on more recent work that has highlighted the importance of the actin cytoskeleton in presynaptic formation.F-actin is required for presynaptic assembly during the early stages of synaptogenesis. Depolymerization of F-actin in young hippocampal neuronal cultures results in a reduction in the size and number of synapses. This effect was not seen with older cultures when synapses are more mature (Zhang and Benson, 2001). This observation correlates with an increase in both pre- and postsynaptic F-actin levels in newly formed synapses compared with mature synapses (Zhang and Benson, 2002).F-actin has been implicated in many steps of synapse assembly and function (Fig. 2; Cingolani and Goda, 2008). One of the roles that has been proposed for F-actin is to act as a scaffold for other presynaptic proteins (Sankaranarayanan et al., 2003). A recent study identified an F-actin–binding active zone molecule Neurabin/NAB-1 that is recruited by a presynaptic F-actin network (Chia et al., 2012). In addition, knockdown of Rac/Cdc42 GTPase exchange factor β-Pix resulted in a decrease in actin at synapses with a concomitant loss of synaptic vesicle clustering (Sun and Bamji, 2011). These studies demonstrate that F-actin at presynaptic sites can recruit and stabilize presynaptic components.Open in a separate windowFigure 2.Assembling the presynaptic active zone. Scaffolding proteins including Liprin, SYD-1, ELKS, Neurabin, Piccolo, and Bassoon form the dense protein network in the presynaptic cytomatrix that facilitates synaptic vesicle docking and fusion. The presynaptic F-actin networks are required for presynaptic assembly and maintenance.Studies of Drosophila NMJs have found that the presynaptic spectrin–actin cytoskeleton is important for synapse stability. Loss of presynaptic spectrin led to retraction of synapses (Pielage et al., 2005). Intriguingly, loss of postsynaptic spectrin increased the total number of the active zone specializations, termed T-bars, and affected the size and distribution of presynaptic sites. Thus, the spectrin cytoskeleton can impose a trans-synaptic influence on synapse development (Pielage et al., 2006).Given the importance of F-actin at synapses, it is crucial to understand the signaling pathways that instruct F-actin organization. Multiple studies have shown that signaling from synaptic cell adhesion molecules can lead to cytoskeletal rearrangements at synapses. Adhesion of hippocampal neurons to syndecan-2–coated beads is sufficient to induce F-actin clustering and downstream formation of presynaptic boutons (Lucido et al., 2009). In mice, the adhesion molecule L1CAM may bind to spectrin–actin adaptor ankyrin to mediate GABAergic synapse formation (Guan and Maness, 2010). Another adhesion molecule of the immunoglobulin superfamily SYG-1 in C. elegans has also been shown to be necessary and sufficient to recruit F-actin to synapses (Chia et al., 2012). In a recent study, secreted bone morphogenetic protein (BMP) can signal in a retrograde fashion to regulate Rac-GEF Trio expression in presynaptic neurons, which is important for controlling synaptic growth (Ball et al., 2010).Interestingly, presynaptic active zone proteins can also affect F-actin assembly (Fig. 2). Knockdown of Piccolo reduced activity-dependent assembly of F-actin at synapses and enhanced dispersion of Synapsin1a and synaptic vesicles in hippocampal neurons. Loss of Piccolo also resulted in a loss of Profilin 2, a regulator of actin polymerization (Waites et al., 2011).Various studies have begun to shed light on the actin regulators required for synaptic F-actin establishment and maintenance. Diaphanous, a formin-related gene that associates with barbed ends of F-actin, was found to function downstream of presynaptic receptor Dlar at fly NMJs. Spectrin–actin capping protein, Adducin, is enriched at presynaptic sites and is required to prevent synapse retraction and elimination (Bednarek and Caroni, 2011; Pielage et al., 2011). Activators of the Arp2/3 complex, WASP and WAVE, have also been implicated in the regulation of F-actin at synapses (Coyle et al., 2004; Stavoe et al., 2012; Zhao et al., 2013). This diversity of F-actin modulators suggests that there are probably different F-actin structures at different stages of development or even in subcellular domains within the synapse. This is supported by observations that F-actin can localize with synaptic vesicles, at the active zone and in the perisynaptic region (Bloom et al., 2003; Sankaranarayanan et al., 2003; Waites et al., 2011; Chia et al., 2012). Thus, much remains to be done in our understanding how distinct F-actin structures are formed and regulated to mediate various processes during synapse assembly and maintenance.

Assembly of the molecular network at presynaptic terminals

Although F-actin might help to initiate the presynaptic assembly process, many other ensuing molecular interactions are required to form the mature presynaptic apparatus (Fig. 2). The presynaptic active zone is comprised of a framework of scaffolding proteins that function as protein-binding hubs for other presynaptic components. Piccolo and Bassoon are important vertebrate multidomain proteins that traditionally have been widely used as active zone markers. Recent electrophysiology data on Piccolo mutant and Bassoon knockdown neurons showed that these molecules are dispensable for synaptic transmission but affect synaptic vesicle clustering (Mukherjee et al., 2010). Furthermore, Piccolo and Bassoon were found to be required for maintaining synapse integrity by regulating ubiquitination and degradation of presynaptic components (Waites et al., 2013).Forward genetic approaches in worms and flies have made important contributions to our understanding of the presynaptic cytomatrix. Studies have found that two active zone scaffolding molecules, SYD-1 and Liprin-α/SYD-2, are required for proper synapse formation (Zhen and Jin, 1999; Patel et al., 2006; Astigarraga et al., 2010; Owald et al., 2010; Stigloher et al., 2011). Interestingly, at fly NMJs, SYD-1 is necessary for clustering presynaptic neurexin that in turn clusters postsynaptic neuroligin (Owald et al., 2012). The presynaptic assembly function of SYD-1 and SYD-2 appears to be conserved because mutation analysis of mammalian SYD-1 and knockdown of Liprin-α both caused defects in presynaptic development and function (Spangler et al., 2013; Wentzel et al., 2013). In flies, the active zone T-bar structure is comprised of ERC/CAST family protein bruchpilot (brp) as the major active zone organizing protein (Fouquet et al., 2009). Brp is not only present at the active zone but also plays important scaffolding roles in localizing Ca2+ channels. In C. elegans, the Brp homologue ELKS-1 is also localized to the active zone; however, the importance of ELKS-1 during development of synapses was only revealed in sensitized genetic backgrounds (Dai et al., 2006; Patel and Shen, 2009), suggesting that there are likely redundant molecular pathways for presynaptic assembly. In the vertebrate system, loss of one of the three ELKS genes, surprisingly, caused an increase in the inhibitory synaptic transmission (Kaeser et al., 2009). Besides Brp, Rab3-interacting molecule (RIM) binding protein (RBP) was found to be important for active zone structural integrity in flies. Using super-resolution microscopy, RBP was found to surround Ca2+ channels at T-bars and loss of RBP resulted in defective Ca2+ channel clustering and reduced evoked neurotransmitter release (Liu et al., 2011).Assembly of the presynaptic active zone is subjected to several layers of regulation. The assembly process is balanced by inhibitory mechanisms that control the number and size of synapses. Loss of the E3 ubiquitin ligase Highwire/RPM-1 results in an increased number of synaptic boutons in flies and multiple active zones in worms (Wan et al., 2000; Zhen et al., 2000). Working together with F-box protein FSN-1, RPM-1 down-regulates the DLK MAP kinase signaling pathway (Liao et al., 2004; Nakata et al., 2005; Yan et al., 2009). Another E3 ubiquitin ligase, the SKP complex, has been shown to eliminate transient synapses during development in worms (Ding et al., 2007). Therefore, ubiquitin-mediated mechanisms play important roles in controlling the presynaptic assembly program.Other inhibitory mechanisms include SRPK79D, a serine–arginine protein kinase discovered in flies that represses T-bar formation (Johnson et al., 2009). In the mutant, the T-bar component Brp is ectopically accumulated in the axonal shaft. Regulator of synaptogenesis, RSY-1, limits the extent of presynaptic assembly by directly binding to active zone scaffold molecule Liprin-α/SYD-2 and SYD-1 (Patel and Shen, 2009). In addition, Liprin-α/SYD-2 may inhibit its own activity via intramolecular interactions (Taru and Jin, 2011; Chia et al., 2013).Taken together, the presynaptic assembly process driven by scaffolding molecules is controlled by complex inhibitory mechanisms to achieve the appropriate extent of aggregation in the process of synapse formation.

Trans-synaptic signals orchestrate pre- and postsynaptic formation

Coordinated pre- and postsynaptic development requires the precise apposition of presynaptic components to postsynaptic specializations. It is conceivable that signals from pre and postsynaptic sides functioning across the synaptic cleft coordinate synaptic differentiation reciprocally. Although a vast assortment of factors have been identified as synaptic organizers, the fact that genetic ablation of some synaptic organizers in vivo fails to elicit dramatic synaptic defects suggests the incomplete view of the trans-synaptic signaling. Moreover, the underlying mechanisms and the cross talk of these signaling pathways are still unclear. In recent years, an emerging body of literature has begun to shed light on trans-synaptic signaling and the importance of environmental cues in synapse formation.

Adhesion proteins instruct synaptic differentiation

A large body of literature suggests that trans-synaptic interactions between synaptic adhesion molecules function bi-directionally for synapse formation and maturation (Fig. 3). Neurexin–neuroligin is the first pair to be shown to induce pre- and postsynapse formation (Scheiffele et al., 2000; Graf et al., 2004; Chih et al., 2005; Nam and Chen, 2005; Chubykin et al., 2007). Recent in vitro studies have unveiled more components interacting with neurexin or neuroligin in specific synaptic differentiation events (Fig. 3, B and C). In early developmental stages, a secreted synaptic organizer, thrombospondin 1 (TSP1, see next section) increases the speed of synaptogenesis through neuroligin 1 (Xu et al., 2010). At excitatory synapses, a retrograde signaling controls synaptic vesicle clustering, neurotransmitter release, and presynaptic maturation by cooperation of neuroligin and N-cadherin (Wittenmayer et al., 2009; Stan et al., 2010; Aiga et al., 2011). A leucine-rich repeat transmembrane (LRRTM) protein family was also identified as an organizer of the function of excitatory synapses through interactions with neurexin (Linhoff et al., 2009). Further studies showed that binding of LRRTMs and neuroligins to neurexin acts redundantly to maintain excitatory synapses by preventing activity and Ca2+-dependent synapse elimination during early development, while performing divergent functions upon synapse maturation (de Wit et al., 2009; Ko et al., 2009, 2011; Soler-Llavina et al., 2011).Open in a separate windowFigure 3.Adhesive trans-synaptic signalings orchestrate excitatory and inhibitory synaptic assembly. Multiple pairs of trans-synaptic adhesion molecules organize synaptic differentiation and function on both pre- and postsynaptic sites. Note that different adhesion molecules are used at excitatory and inhibitory synapses. LPH1, latrophilin 1; α-DG, α-dystroglycan; β-DG, β-dystroglycan; S-SCAM, synaptic scaffolding molecule; Lasso, LPH1-associated synaptic surface organizer; IL-1RAcp, interleukin-1 receptor accessory protein.The function of neurexin and neuroligin in mediating synaptic differentiation has also been shown at Drosophila NMJs and mammalian CNS. In mammalian, although neither compound knockout of three neurexins nor two individual neuroligin knockout mice display severe defects in the number or morphology of synapses (Missler et al., 2003), the deletion of either neurexin or neuroligin affects the neurotransmitter release and in turn impairs the relevant behavior (Zhang et al., 2005; Blundell et al., 2009, 2010; Etherton et al., 2009; Jedlicka et al., 2011). Neurexin loss of function in fly leads to reduced number and defective morphology of synaptic boutons and active zones from early developmental stages (Li et al., 2007; Chen et al., 2010). In contrast, deletion of either neuroligin 1 or 2 causes NMJ defects and alternations of active zones only in the larval stage, indicating that they function mainly in the expansion of NMJs during development (Banovic et al., 2010; Sun et al., 2011). These abnormalities further impair synaptic transmission at the NMJs (Li et al., 2007; Banovic et al., 2010; Chen et al., 2010; Sun et al., 2011). Moreover, these phenotypes are enhanced when the Teneurin family of adhesion molecules is deleted, suggestive of functional redundancy between adhesion molecules (Mosca et al., 2012). Recently, it has been reported that an active zone protein, SYD-1, is required for the formation and function of the neurexin–neuroligin complex in flies (Fig. 2; Owald et al., 2012), providing an example of how trans-synaptic neurexin–neuroligin signaling orchestrates synaptic assembly bi-directionally. Interestingly, at postsynaptic sites, the NMDA receptor activity-triggered Ca2+-dependent cleavage of neuroligin 1 was found to destabilize presynaptic neurexin, reduce presynaptic release probability, and depress synaptic transmission (Peixoto et al., 2012). This observation raises a possibility that neurexin and neuroligin could fine-tune synaptogenesis both positively and negatively.Although Drosophila neuroligin and neurexin mutants share many phenotypes in synaptic differentiation, there are some unique features for each mutant, suggesting that they play distinct roles. For example, some aspects of synaptic specificity are achieved by different pairs of neurexin–neuroligin interactions. Neuroligin 1 promotes the growth and differentiation of excitatory synapses by binding to PSD-95, whose amount balances the ratio of excitatory-to-inhibitory synaptic specializations (Prange et al., 2004; Banovic et al., 2010). Neuroligin 2, on the contrary, binds to a scaffold protein gephyrin at inhibitory synapses, instructing inhibitory postsynaptic assembly (Fig. 3, B and C; Poulopoulos et al., 2009). Different isoforms of neurexin also contribute to the differentiation of excitatory and inhibitory synapses (Fig. 3, B and C; Chih et al., 2006; Graf et al., 2006; Kang et al., 2008).Other novel trans-synaptic interactions have also been identified to organize synaptic differentiation (Fig. 3, B and C). For example, Netrin-G ligand 3 (NGL-3), localized at postsynaptic region, induces excitatory synaptic differentiation by interacting with the receptor tyrosine phosphatase LAR family proteins, including PTPδ and PTPσ (Woo et al., 2009; Kwon et al., 2010). PTPδ can also trans-interact with Slitrk3 and IL-1 receptor accessory protein (IL-1RAcP) to promote presynaptic formation (Takahashi et al., 2012; Yoshida et al., 2012). Molecules that function in other neuronal developmental processes have also been shown to regulate synaptic differentiation. Farp1, essential for the dynamics of dendritic filopodia, regulates postsynaptic development and triggers a retrograde signal promoting active zone assembly by binding to SynCAM 1 (Cheadle and Biederer, 2012). Teneurins, instructing synaptic partner selection in fly olfactory system (Hong et al., 2012), act in synaptogenesis through trans-synaptic interaction at NMJs (Mosca et al., 2012). Another splice variant of a postsynaptic Teneurin-2 in rat, Lasso, binding with presynaptic Latrophilin 1 (LPH1), induces presynaptic Ca2+ signals and regulates synaptic function (Silva et al., 2011). Neural activity is also involved in controlling the growth of the presynapse. Conditioning or BDNF application induces presynaptic bouton development via an ephrin-B–dependent manner (Li et al., 2011), suggesting the role of EphB/ephrin-B signaling in activity-dependent synaptic modification.

Secreted molecules organize synapse differentiation

In addition to adhesion molecules, some secreted molecules also serve as synaptic organizers (Fig. 4). For example, the motor neuron–derived ligand agrin, which was the first identified secreted organizing molecule for postsynaptic differentiation, activates MuSK, a postsynaptic receptor tyrosine kinase, to regulate NMJ specialization (Glass et al., 1996; Zhou et al., 1999). Recently, a low-density lipoprotein receptor–related protein, LRP4, was identified as the co-receptor of agrin, forming a complex with MuSK and mediating MuSK signaling (Kim et al., 2008; Zhang et al., 2008). Several Wnts appears to act together with agrin to activate the LRP4–MuSK receptor complex to promote postsynaptic differentiation (Jing et al., 2009; Zhang et al., 2012). LRP4 also acts as a direct retrograde signal, functioning independently of MuSK for presynaptic differentiation (Yumoto et al., 2012), demonstrating that LRP4 acts as a bi-directional synaptic organizer (Fig. 4, left).Open in a separate windowFigure 4.Secreted trans-synaptic signaling at NMJs and CNS synapses. (Left) At Drosophila neuromuscular junctions (NMJs), Wnts are secreted from presynaptic terminals in association with Evi in the form of exosomes. In vertebrate NMJs, Wnt binds to the Agrin–LRP4–MuSK complex to regulate synapse formation. (Right) At CNS synapses, glia-derived thrombospondins (TSPs) and presynaptic neuron–derived cerebellin (Cbln) organize synapse differentiation and formation bi-directionally through binding to GluD2 and an isoform of neurexin (S4+) on the postsynaptic and presynaptic membranes, respectively. LTCC, L-type Ca2+ channel complex; AChR, acetyl choline receptor.Wnt is another well-characterized signaling molecule regulating many developmental processes including synaptic differentiation bi-directionally. Wnt regulates synaptic assembly both positively and negatively. For example, Wnt3 collaborates with agrin to promote the clustering of acetyl choline receptor (AChR) at the vertebrate NMJs (Henriquez et al., 2008), while Wnt3a inhibits AChR aggregation through β-catenin signaling (Wang et al., 2008). In the C. elegans NMJ, a Wnt molecule, CWN-2, stimulates the delivery and insertion of AchR to the postsynaptic membrane through the activation of a Frizzled–CAM-1 receptor complex (Jensen et al., 2012). Local Wnt gradient can suppress synapse formation in both C. elegans and Drosophila (Inaki et al., 2007; Klassen and Shen, 2007). Interestingly, in these contexts, Wnts are secreted from nonneuronal or nonsynaptic partner cells, suggesting that environmental factors can shape synaptic connections. Wnt can also be secreted from presynaptic neurons. A recent study demonstrated the trans-synaptic transmission of Wnt by exosome-like vesicles containing the Wnt-binding protein Evi at Drosophila NMJs (Fig. 4, left; Korkut et al., 2009; Koles et al., 2012). Presynaptic vesicular release of Evi is required for the secretion of Wnt. Intriguingly, different Wnt ligands regulate synapse formation in distinct cellular contexts. Wnt3a promotes excitatory synaptic assembly through CaMKII, whereas Wnt5a mediates inhibitory synapse formation by stabilizing GABAA receptors (Cuitino et al., 2010; Ciani et al., 2011). This functional diversity indicates that different Wnts, receptors, and downstream pathways, as well as cell-specific contexts dictate the action of extracellular cues. Another conserved secreted molecule, netrin/UNC-6, can also pattern synapses by either promoting or inhibiting synapse formation (Colón-Ramos et al., 2007; Poon et al., 2008). Because Wnt and netrin often exist in gradients, these observations suggest that the localization of synapses can be specified by the gradient of extrinsic cues.In mammalian, several glia-derived cues have been shown to play important roles in regulating synapse formation or elimination. Thrombospondins (TSPs) are trans-synaptic organizers secreted from immature astrocytes (Christopherson et al., 2005). Both in vitro and in vivo data demonstrate the capacity of TSPs to increase synapse number, promote the localization of synaptic molecules, and refine the pre- and postsynaptic alignment (Christopherson et al., 2005; Eroglu et al., 2009). Recently, two transmembrane molecules were uncovered in mediating TSP-induced synaptogenesis (Fig. 4, right). Neuroligin 1 interacts with TSP1 with its extracellular domain mediating the acceleration of synaptogenesis in hippocampal neurons (Xu et al., 2010). α2δ-1, a subunit of the L-type Ca2+ channel complex (LTCC), was also identified as the postsynaptic receptor of TSP in excitatory CNS neurons (Eroglu et al., 2009). Interaction between TSP and α2δ–1 triggers the conformational changes and sequentially recruits synaptic scaffolding molecules and initiates synapse formation (Eroglu et al., 2009). Interestingly, TSP-induced synapses, although structurally normal and presynaptically active, are postsynaptically silent due to the lack of AMPA receptors (Christopherson et al., 2005), indicating the existence of other glia-derived signals involved in synapse formation. In fact, in cultured hippocampal neurons, a glia-derived neurotrophic factor GDNF enhances the pre- and postsynaptic adhesion by triggering the trans-homophilic interaction of its receptors GFRα1 localized at both pre- and postsynaptic sites (Ledda et al., 2007). Several other glia-derived factors have been shown to play critical roles in synaptogenesis. Astrocytes secrete extracellular molecules hevin and SPARC to regulate synapse formation in vitro and in vivo (Kucukdereli et al., 2011). Astrocytes also express a transmembrane adhesion protein, protocadherin-γ, serving as a local cue to promote synapse formation (Garrett and Weiner, 2009). TGF-β secreted from the NMJ glia acts together with the muscle-derived TGF-β to control synaptic growth (Fuentes-Medel et al., 2012). In a similar fashion, secretion of BDNF by vestibular supporting cells is required for synapse formation between hair cells and sensory organs (Gómez-Casati et al., 2010).Another important synaptic organizer is cerebellin (Cbln), a presynapse-derived complement protein, C1q-like family protein. In cbln1-null mice the number of parallel fibers (PF)–Purkinje synapses is dramatically reduced; the postsynaptic densities in the remaining synapses are larger than the apposite active zones (Hirai et al., 2005). Cbln was also found to regulate synaptic plasticity, as cbln1-null mice show impaired long-term depression in cerebellum (Hirai et al., 2005). These defects precisely resemble those in mice lacking a putative glutamate receptor, GluD2 (Kashiwabuchi et al., 1995; Kurihara et al., 1997), suggesting that Cbln1 and GluD2 function in synaptic differentiation through a common pathway. Interestingly, the C-terminal domain and N-terminal domains of GluD2 are indispensable for cerebella LTD and PF–Purkinje synaptic morphology, respectively (Kohda et al., 2007; Uemura et al., 2007; Kakegawa et al., 2008, 2009). Further studies suggested that Cbln1 directly binds to the N-terminal domain of GluD2 and recruits postsynaptic proteins by clustering GluD2 (Matsuda et al., 2010). Neurexin was recently reported as the presynaptic receptor of Cbln in promoting synaptogenesis (Uemura et al., 2010), which reinforces the understanding of Cbln-mediated trans-synaptic signaling: Cbln serves as a bi-directional synapse organizer by linking presynaptic neurexin and postsynaptic GluD2 (Fig. 4, right).Besides being required for synapse formation at early stages, genetic ablation of GluD2 in adult cerebellum leads to loss of PF–Purkinje synapses (Takeuchi et al., 2005), indicating that Cbln1–GluD2 signaling is also important for the maintenance of PF–Purkinje synapses. Chronic stimulation of neural activity decreases Cbln1 expression and diminishes the number of PF–Purkinje synapses (Iijima et al., 2009), suggesting the importance of Cbln1–GluD2 signaling for synaptic plasticity and homeostasis.Cbln subfamily proteins are widely expressed throughout the brain (Miura et al., 2006), suggesting that their synaptogenic roles may be wide spread in other regions of the brain. Cbln2 and 4 are also secreted proteins, whereas Cbln3 is retained in the cellular endomembrane system (Iijima et al., 2007). Cbln1 and 2, interacting with an isoform of presynaptic neurexin, induce synaptogenesis (Joo et al., 2011; Matsuda and Yuzaki, 2011). Notably, the cortical synapses induced by neurexin–Cbln signaling are preferentially inhibitory (Joo et al., 2011), distinguishing the effects of Cbln from neuroligin. GluD1 was recently found to be the postsynaptic receptor of Cbln1 and 2 in cortical neurons, mediating the differentiation of inhibitory presynapses (Yasumura et al., 2012). On the other side, Cbln4 selectively binds to the netrin receptor DCC in a netrin-displaceable manner (Fig. 4, right), suggesting a potential function of Cbln4 through DCC signaling pathway (Iijima et al., 2007). Intriguingly, C1q, although sharing similar structure with Cbln, serves an opposite role by regulating the synapse elimination: C1q released from retinal ganglion cells refines the retinogeniculate connections by eliminating unneeded synapses (Stevens et al., 2007).

Concluding remarks

Synapse development is regulated in multiple steps. Research over the last few years have uncovered many regulatory mechanisms on how trafficking of synaptic material is regulated and how scaffold proteins act with cytoskeleton networks and trans-synaptic signaling to instruct the synapse formation. Nevertheless, our understanding of the cellular and molecular mechanisms regulating synapse development is still incomplete. For example, how is the direction, speed, and amount of synaptic material being transported specified? How is a synapse’s size determined? How is synapse type and strength specified through adhesive and secreted trans-synaptic signaling? How do the redundant synapse-inducing pathways interact with each other? Given the rapidly emerging improvements of technologies, especially super-resolution microscopy and high-throughput genomics and proteomics, the synapse development field will likely rapidly evolve in the near future.  相似文献   

7.
In this issue, Matteoli and colleagues show that SNAP-25 levels regulate the efficacy of presynaptic glutamate release and thereby alter short-term plasticity, with potential relevance for psychiatric diseases.EMBO reports(2013) 14 7, 645–651 doi:10.1038/embor.2013.75Control of exocytotic neurotransmitter release is essential for communication in the nervous system and for preventing synaptic abnormalities. The function of synaptosomal-associated protein of 25 kDa (SNAP-25) as a crucial component of the core machinery required for synaptic vesicle fusion is well established, but evidence is growing to suggest an additional modulatory role in neurotransmission. In this issue of EMBO reports, Antonucci et al show that the efficacy of evoked glutamate release is modulated by the expression levels of SNAP-25—a function that might relate to the ability of SNAP-25 to modulate voltage-gated calcium channels and presynaptic calcium ion concentration [1]. Altered synaptic transmission and short-term plasticity due to changes in SNAP-25 expression might have direct consequences for brain function and for the development of neuropsychiatric disorders.Communication between neurons is essential for brain function and occurs through chemical neurotransmission at specialized cell–cell contacts termed ‘synapses''. Within the nerve terminal of the presynaptic neuron electrical stimuli cause the opening of voltage-gated calcium channels (VGCCs), which results in the influx of calcium ions. This triggers the exocytic release of neurotransmitter by fusion of synaptic vesicles with the presynaptic membrane. Released neurotransmitter molecules are detected by specific receptors expressed by the postsynaptic neuron.Calcium-induced synaptic vesicle fusion requires complex assembly between the soluble N-ethylmaleimide-sensitive factor (NSF) attachment protein receptor (SNARE) synaptobrevin 2, located on the synaptic vesicle, and the abundant plasma membrane SNAREs SNAP-25 and syntaxin 1, on the opposing presynaptic plasma membrane. SNARE complex assembly is tightly regulated by Sec1/Munc18-like proteins [2]. Further regulatory factors such as the synaptic vesicle calcium-sensing protein synaptotagmin 1 couple the SNARE machinery to presynaptic calcium influx. SNARE-mediated neurotransmitter release occurs preferentially at the active zone—a presynaptic membrane domain specialized for exocytosis within which VGCCs are positioned close to docked synaptic vesicles through a proteinaceous cytomatrix and associated cell adhesion molecules [3,4].Altered short-term plasticity due to changes in SNAP-25 expression might have direct consequences for brain function and for the development of neuropsychiatric disordersAn unresolved conundrum in synaptic transmission remains—the observation that SNARE proteins, such as SNAP-25, are among the most highly expressed, in copy number, presynaptic proteins, whilst only a handful of SNARE complexes are needed to drive the fusion of a single synaptic vesicle [5]. Why, then, are SNAREs such as SNAP-25 so abundant? One possible explanation might be that SNARE proteins, in addition to forming trans-SNARE complexes, assemble with other proteins, and such partitioning might regulate neurotransmission. For example, SNAP-25 has been shown to negatively regulate VGCCs in glutamatergic but not in GABAergic neurons [6]. A secondary regulatory function of SNAP-25 is also supported by its genetic association with synaptic abnormalities such as schizophrenia and attention deficit hyperactivity disorder (ADHD) in humans [7]. SNAP-25 expression is reduced twofold in the hippocampus and frontal lobe from schizophrenic patients [8] and in animal models for ADHD [9]. Thus, SNAP-25 expression levels might crucially regulate normal synaptic function.A new study in this issue of EMBO reports by Antonucci and colleagues investigates the consequences of reduced SNAP-25 expression on synaptic function in SNAP-25+/− heterozygous (Het) mutant mice. By using patch clamp electrophysiology, Antonucci et al revealed a selective enhancement of glutamatergic but not GABAergic neurotransmission as a result of reduced SNAP-25 expression. Several other parameters including the amplitude and frequency of miniature excitatory and inhibitory currents were unaffected. These data indicate that reduced levels of SNAP-25, an essential component of the fusion machinery, selectively enhance evoked release of glutamate whilst synaptic connectivity and postsynaptic glutamate receptor sensitivity remain unaltered. Further electrophysiological experiments in hippocampal neurons in culture showed that elevated glutamatergic transmission was probably due to increased release probability rather than changes in the number of fusion-prone, so-called ‘readily releasable synaptic vesicles''. This effect was occluded by pharmacologically induced calcium entry bypassing VGCCs, suggesting that altered calcium influx might underlie the differences in evoked glutamate release between wild-type and SNAP-25 Het neurons. As schizophrenia and ADHD are associated with changes in short-term plasticity, a paradigm reflecting presynaptic function, Antonucci et al analysed neurotransmission by paired-pulse stimulation—a protocol whereby two closely paired stimuli are applied within a 50 ms time interval. Wild-type neurons showed significant short-term facilitation, that is, a stronger response to the second stimulus as a result of increased calcium levels in the presynaptic compartment. By contrast, Het neurons had a reduced response to the second stimulus. Such paired-pulse depression is commonly viewed as a sign of increased release probability, which occurs when the first stimulus induces a partial depletion of release-ready synaptic vesicles during paired stimulation. As a consequence, the second stimulus evokes a comparably reduced response [3]. The switch from paired-pulse facilitation to depression was not fully reproduced in hippocampal slices from wild-type and Het mice, although facilitation seemed to be attenuated in SNAP-25 Het slices. One possible explanation for the apparent discrepancy between cultured neurons taken from newborn animals and acute slices from adult mice is the constant postnatal increase in SNAP-25 expression in SNAP-25 Het mice [10], which might partly counteract the defects caused by heterozygosity. Consistent with this explanation are data from rescue experiments by Antonucci et al, which showed that altered neurotransmission and defects in short-term plasticity in Het neurons can be gradually recovered in parallel with increased SNAP-25 expression. Moreover, cultured neurons show substantially higher levels of endogenous activity compared with acute slice preparations, leading to possible changes in the partitioning of SNAP-25 between SNARE complexes and association with VGCCs. Further experiments are clearly required to resolve these issues. Irrespective of these potential caveats, the combined data support the hypothesis that alterations in SNAP-25 expression underlie regulatory changes in neurotransmission, resulting in altered short-term plasticity and possibly disease.Many open questions remain. In particular, the precise mechanisms underlying elevated glutamatergic transmission and presynaptic plasticity under conditions of reduced SNAP-25 expression remain elusive. It has been shown before that free SNAP-25 inhibits Cav2.1-type VGCCs [6], an effect reversed by overexpression of synaptotagmin 1, which might associate with SNAP-25. Conversely, SNAP-25 occludes negative regulation of Cav2.2 VGCCs by free syntaxin 1 [3]. Hence, it is tempting to speculate that differential partitioning of SNAP-25 between free, SNARE-, synaptotagmin 1- and VGCC-complexed forms could regulate evoked neurotransmission (Fig 1). In this scenario, reduced SNAP-25 expression in Het animals and in schizophrenic and ADHD patients would be sufficient to sustain SNARE-mediated synaptic vesicle fusion but partially releases VGCCS from SNAP-25-mediated inhibition. This would result in elevated calcium influx and facilitated neurotransmission. Additional levels of regulation could be imposed by developmental switching between alternatively spliced ‘a'' and ‘b'' isoforms of SNAP-25 [11], age-dependent alterations in presynaptic protein turnover and post-translational modifications.Open in a separate windowFigure 1Effect of presynaptic SNAP-25 levels on calcium-induced glutamate release. Top: in wild-type (WT) neurons, SNARE-mediated calcium-triggered synaptic vesicle fusion is negatively regulated by complex formation between SNAP-25 and VGCCs. Bottom: reduced SNAP-25 expression in heterozygotes (Het;+/−) partly releases VGCCs from SNAP-25-mediated clamping, resulting in elevated calcium influx through VGCCs and increased glutamate release through SNARE-mediated calcium-triggered synaptic vesicle fusion. Note that many key exocytotic proteins have been omitted for clarity. SNAP-25, synaptosomal-associated protein of 25 kDa; SNARE, soluble NSF attachment protein receptor; VGCC. voltage-gated calcium channel.Future studies need to address these possibilities, and their relationship to cognitive impairments and synaptic diseases, such as schizophrenia and ADHD.  相似文献   

8.
The acid-sensing ion channel 1a (ASIC1a) is widely expressed in central and peripheral neurons where it generates transient cation currents when extracellular pH falls. ASIC1a confers pH-dependent modulation on postsynaptic dendritic spines and has critical effects in neurological diseases associated with a reduced pH. However, knowledge of the proteins that interact with ASIC1a and influence its function is limited. Here, we show that α-actinin, which links membrane proteins to the actin cytoskeleton, associates with ASIC1a in brain and in cultured cells. The interaction depended on an α-actinin-binding site in the ASIC1a C terminus that was specific for ASIC1a versus other ASICs and for α-actinin-1 and -4. Co-expressing α-actinin-4 altered ASIC1a current density, pH sensitivity, desensitization rate, and recovery from desensitization. Moreover, reducing α-actinin expression altered acid-activated currents in hippocampal neurons. These findings suggest that α-actinins may link ASIC1a to a macromolecular complex in the postsynaptic membrane where it regulates ASIC1a activity.Acid-sensing ion channels (ASICs)2 are H+-gated members of the DEG/ENaC family (13). Members of this family contain cytosolic N and C termini, two transmembrane domains, and a large cysteine-rich extracellular domain. ASIC subunits combine as homo- or heterotrimers to form cation channels that are widely expressed in the central and peripheral nervous systems (14). In mammals, four genes encode ASICs, and two subunits, ASIC1 and ASIC2, have two splice forms, a and b. Central nervous system neurons express ASIC1a, ASIC2a, and ASIC2b (57). Homomeric ASIC1a channels are activated when extracellular pH drops below 7.2, and half-maximal activation occurs at pH 6.5–6.8 (810). These channels desensitize in the continued presence of a low extracellular pH, and they can conduct Ca2+ (9, 1113). ASIC1a is required for acid-evoked currents in central nervous system neurons; disrupting the gene encoding ASIC1a eliminates H+-gated currents unless extracellular pH is reduced below pH 5.0 (5, 7).Previous studies found ASIC1a enriched in synaptosomal membrane fractions and present in dendritic spines, the site of excitatory synapses (5, 14, 15). Consistent with this localization, ASIC1a null mice manifested deficits in hippocampal long term potentiation, learning, and memory, which suggested that ASIC1a is required for normal synaptic plasticity (5, 16). ASICs might be activated during neurotransmission when synaptic vesicles empty their acidic contents into the synaptic cleft or when neuronal activity lowers extracellular pH (1719). Ion channels, including those at the synapse often interact with multiple proteins in a macromolecular complex that incorporates regulators of their function (20, 21). For ASIC1a, only a few interacting proteins have been identified. Earlier work indicated that ASIC1a interacts with another postsynaptic scaffolding protein, PICK1 (15, 22, 23). ASIC1a also has been reported to interact with annexin II light chain p11 through its cytosolic N terminus to increase cell surface expression (24) and with Ca2+/calmodulin-dependent protein kinase II to phosphorylate the channel (25). However, whether ASIC1a interacts with additional proteins and with the cytoskeleton remain unknown. Moreover, it is not known whether such interactions alter ASIC1a function.In analyzing the ASIC1a amino acid sequence, we identified cytosolic residues that might bind α-actinins. α-Actinins cluster membrane proteins and signaling molecules into macromolecular complexes and link membrane proteins to the actincytoskeleton (for review, Ref. 26). Four genes encode α-actinin-1, -2, -3, and -4 isoforms. α-Actinins contain an N-terminal head domain that binds F-actin, a C-terminal region containing two EF-hand motifs, and a central rod domain containing four spectrin-like motifs (2628). The C-terminal portion of the rod segment appears to be crucial for binding to membrane proteins. The α-actinins assemble into antiparallel homodimers through interactions in their rod domain. α-Actinins-1, -2, and -4 are enriched in dendritic spines, concentrating at the postsynaptic membrane (2935). In the postsynaptic membrane of excitatory synapses, α-actinin connects the NMDA receptor to the actin cytoskeleton, and this interaction is key for Ca2+-dependent inhibition of NMDA receptors (3638). α-Actinins can also regulate the membrane trafficking and function of several cation channels, including L-type Ca2+ channels, K+ channels, and TRP channels (3941).To better understand the function of ASIC1a channels in macromolecular complexes, we asked if ASIC1a associates with α-actinins. We were interested in the α-actinins because they and ASIC1a, both, are present in dendritic spines, ASIC1a contains a potential α-actinin binding sequence, and the related epithelial Na+ channel (ENaC) interacts with the cytoskeleton (42, 43). Therefore, we hypothesized that α-actinin interacts structurally and functionally with ASIC1a.  相似文献   

9.
Gong B  Wang YT 《The EMBO journal》2012,31(4):783-785
EMBO J 31 4, 805–816 (2012); published online December202011Synaptic plasticity, the activity-dependent modification of synaptic strength, plays a fundamental role in learning and memory as well as in developmental maturation of neuronal circuitry. However, how synaptic plasticity is induced and regulated remains poorly understood. In this issue of The EMBO Journal, Yang and colleagues present sets of exciting data, suggesting that G-protein-coupled receptors (GPCRs) selectively execute distinct signalling pathways to differentially regulate induction thresholds of hippocampal long-term potentiation (LTP) and long-term depression (LTD), thereby governing the direction of synaptic plasticity. These results shed significant light on our current understanding of how bidirectional synaptic plasticity is regulated.Synaptic plasticity has been demonstrated at synapses in various brain regions; the most well-characterized forms are LTP and LTD at hippocampal CA1 glutamatergic synapses (Collingridge et al, 2004). In experimental models, LTP and LTD can be, respectively, induced by high-frequency stimulation (HFS) and low-frequency stimulation (LFS) via activation of the N-methyl-D-aspartic acid (NMDA) subtype ionotropic glutamate receptor (NMDAR). However, how HFS and LFS activate NMDARs and thereby lead to synaptic plasticity remains poorly understood and highly controversial. It is even more unclear how the bidirectional synaptic plasticity is produced and regulated in response to physiological or pathological changes.Functional NMDARs consist primarily of two GluN1 subunits and two GluN2 subunits, with GluN2A and GluN2B subunits being the most common NMDAR subunits found in the cortical and hippocampal regions of the adult brain (Cull-Candy et al, 2001). GluN2A and GluN2B subunits may confer distinct gating and pharmacological properties to NMDARs and couple them to distinct intracellular signalling machineries (Cull-Candy et al, 2001). Moreover, the ratio of these two subpopulations of NMDARs at the glutamatergic synapse is dynamically regulated in an activity-dependent manner (Bellone and Nicoll, 2007; Cho et al, 2009; Xu et al, 2009). Although controversial, GluN2A- and GluN2B-containing NMDARs have been suggested to have differential roles in regulating the direction of synaptic plasticity (Collingridge et al, 2004; Morishita et al, 2007). Among the factors shown to regulate NMDAR function, Src family tyrosine kinases may be the best characterized, with both Src and Fyn able to upregulate NMDAR function, and thus LTP induction (Salter and Kalia, 2004). However, if these kinases modulate NMDAR function in a NMDAR subunit-specific manner remains unknown. To explore this concept, Yang et al (2012) investigated the potential subunit-specific regulation of NMDARs by Src and Fyn using whole-cell patch clamp recording of NMDAR-mediated currents from acutely dissociated CA1 hippocampal neurons or from rat hippocampal slices. They found that intracellular perfusion of recombinant Src or Fyn increased the NMDAR-mediated currents. By applying subunit-preferential antagonists of GluN2A- or GluN2B-containing NMDARs, or by using neurons obtained from GluN2A knockout mice, they discovered that Src and Fyn differentially enhanced currents gated through GluN2A- and GluN2B-containing NMDARs, respectively.Can physiological or pathological factors differentially activate Src or Fyn, thereby exerting subunit-specific regulation of NMDAR function? To answer this question, Yang et al focused their investigation on the role of GPCRs, specifically pituitary adenylate cyclase activating peptide receptor (PAC1R) and dopamine D1 receptor (D1R), both of which have recently been shown to potentiate NMDARs through Src family kinases (Macdonald et al, 2005; Hu et al, 2010). Indeed, they found that activation of PAC1R specifically increased GluN2A-NMDAR-mediated currents without affecting currents gated through GluN2B-NMDARs, and this potentiation was prevented by the Src-specific inhibitory peptide Src(40–58) (Salter and Kalia, 2004). To rule out the contribution of Fyn, the authors developed a novel-specific Fyn inhibitory peptide Fyn(39–57), and demonstrated that it had little effect on PAC1R potentiation. In contrast, activation of D1R potentiated GluN2B- (but not GluN2A-) NMDAR-mediated currents, and this potentiation was specifically eliminated by Fyn(39–57), but not by Src(40–58). The authors further demonstrated that stimulation of PAC1Rs resulted in a selective activation of Src kinase and consequent tyrosine phosphorylation of the GluN2A subunit, whereas activation of D1Rs led to a specific increase in Fyn-mediated tyrosine phosphorylation of the GluN2B subunit. To provide convincing evidence that these subunit-differential modulations are indeed the result of tyrosine phosphorylation of the respective NMDAR subunits, the authors then performed electrophysiological experiments using neurons from two knockin mouse lines GluN2A(Y1325F) and GluN2B(Y1472F), in which the tyrosine phosphorylation residues in native GluN2A and GluN2B subunits were, respectively, replaced with non-phosphorylatable phenylalanine residues. As expected, the authors found that PAC1R-mediated potentiation of NMDA currents was lost in neurons from GluN2A(Y1325F) mice (but maintained in neurons from GluN2B(Y1472F) mice), while D1R-mediated enhancement of NMDA currents was only observed in neurons from GluN2A(Y1325F) mice. Together, as illustrated in Figure 1, the authors have made a very convincing case that PAC1R and D1R, respectively, enhance function of GluN2A- and GluN2B-containing NMDARs by differentially activating Src- and Fyn-mediated phosphorylation of respective NMDAR subunits.Open in a separate windowFigure 1GPCRs regulate the direction of synaptic plasticity via activating distinct signalling pathways. Synaptic NMDA receptors, both GluN2A- and GluN2B-containing, play key roles in the induction of various forms of synaptic plasticity at the hippocampal CA1 glutamatergic synapse. Under the basal level of GluN2A and GluN2B ratio, stimulation with a train of pulses at frequencies from 1 to 100 Hz produces a frequency and plasticity (LTD–LTP) curve, with maximum LTD and LTP being, respectively, induced at 1 and 100 Hz. Activation of PAC1R with its agonist PACAP38 activates Src and thereby results in tyrosine phosphorylation and consequent functional upregulation of GluN2A-containing NMDARs, resulting in an increase in the ratio of functional GluN2A and GluN2B. The increased ratio in turn causes a left shift of frequency and plasticity curve, favouring LTP induction. In contrast, activation of D1R by the receptor agonist SKF81297 triggers Fyn-specific tyrosine phosphorylation and functional upregulation of GluN2B, causing a reduction of GluN2A and GluN2B ratio. This decreased ratio results in a right shift of the curve, favouring LTD induction. The ability of GPCRs to differentially activate distinct downstream signalling pathways involved in synaptic plasticity suggests the potential roles of GPCRs in governing the direction of synaptic plasticity.Given the coupling of NMDARs to the induction of synaptic plasticity, it is then reasonable to ask if activation of the two GPCRs can selectively affect the induction of LTP or LTD at CA1 synapses. Yang and colleagues investigated the effects of pharmacological activation of PAC1R and D1R on the induction of LTP and LTD by recording the field excitatory postsynaptic potentials from hippocampal slices. Consistent with differential roles of NMDAR subunits in governing directions of synaptic plasticity, the authors observed that activation of PAC1Rs reduces the induction threshold of LTP, while stimulation of D1Rs favours LTD induction (Figure 1). Facilitation of LTP by PAC1R and LTD by D1R were, respectively, prevented in the brain slices obtained from GluN2A(Y1325F) and GluN2B(Y1472F) knockin mice, supporting the differential involvements of Src-mediated GluN2A phosphorylation and Fyn-mediated GluN2B phosphorylation.Taken together, the authors'' results have demonstrated that activation of PAC1R and D1R can control the direction of synaptic plasticity at the hippocampal CA1 synapse by differentially regulating NMDAREPSCs in a subunit-specific fashion (Figure 1). Specifically, PAC1R enhances the function of GluN2A-containing NMDARs by increasing Src phosphorylation of GluN2A subunit at Y1325, whereas D1R upregulates GluN2B-containing NMDARs through increased Fyn phosphorylation of GluN2B at Y1472. Moreover, by regulating the ratio of functional GluN2A- and GluN2B-containing NMDARs, PAC1R and D1R in turn modulate the direction of synaptic plasticity, favouring the production of LTP and LTD, respectively.While consistent with the recently proposed hypothesis that GluN2A and GluN2B may have preferential roles in the induction of hippocampal CA1 LTP and LTD (Collingridge et al, 2004; but see also Morishita et al, 2007), the current study further emphasizes the importance of GluN2A/GluN2B ratios in regulating LTP and LTD thresholds: increased ratio favours LTP, while reduced ratio promotes LTD. However, this seems to contradict some recent studies where the reduction and increase in the GluN2A/GluN2B ratio appeared to, respectively, favour LTP (Cho et al, 2009; Xu et al, 2009) and LTD (Xu et al, 2009). Therefore, the direction of plasticity change is likely modulated not only by the GluN2A/GluN2B ratio, but also by additional factors such as experimental conditions, developmental stages, and brain regions.Under many experimental conditions, LTP and LTD are usually induced by HFS and LFS stimulating protocols, respectively, but it remains essentially unknown how LTP and LTD are physiologically or pathologically generated in animals. To this end, the identification of different GPCRs as the endogenous upstream regulators of NMDA receptor subpopulations, and hence regulators of synaptic plasticity, is the major novelty of Yang and colleagues'' work. Future studies are needed to investigate if and how PAC1R and/or D1R are critically involved in the production of LTP or LTD in animals under physiological or pathological conditions. Given the fact that Src family kinases may be required for LTP induced by HFS in hippocampal slices (Salter and Kalia, 2004), an equally intriguing question would be whether these GPCRs are actually required for LTP/LTD induced by HFS/LFS experimental paradigms. In line with this conjecture, it would be interesting to determine if ligands for various GPCRs co-exist in the glutamatergic presynaptic terminals and, if so, can be differentially co-released with glutamate in a frequency-dependent manner, thereby contributing to either HFS-induced LTP or LFS-induced LTD.The findings by Yang and colleagues establish an exciting mechanistic model by which GPCRs can govern the direction of synaptic plasticity by determining the contributions of GluN2A- and GluN2B-NMDARs through differential tyrosine phosphorylation of respective NMDA receptor subtypes. Additional studies further validating this model under physiological and pathological conditions will greatly improve our understanding of the molecular mechanisms underlying synaptic plasticity and cognitive brain functions. In addition, NMDARs, depending on their subunit composition and/or subcellular localization, may also have complex roles in mediating neuronal survival and death (Lai et al, 2011). Considering that neurotoxicity produced by over-activation of NMDARs is widely accepted to be a common mechanism for neuronal loss in a number of acute brain injuries and chronic neurodegenerative diseases, Yang and colleagues'' finding of the differential regulation of NMDAR subunits by different GPCRs could have wider implications beyond synaptic plasticity.  相似文献   

10.
Many forms of synaptic plasticity are triggered by biochemical signaling that occurs in small postsynaptic compartments called dendritic spines, each of which typically houses the postsynaptic terminal associated with a single glutamatergic synapse. Recent advances in optical techniques allow investigators to monitor biochemical signaling in single dendritic spines and thus reveal the signaling mechanisms that link synaptic activity and the induction of synaptic plasticity. This is mostly in the study of Ca2+-dependent forms of synaptic plasticity for which many of the steps between Ca2+ influx and changes to the synapse are now known. This article introduces the new techniques used to investigate signaling in single dendritic spines and the neurobiological insights that they have produced.Each neuron typically receives 1000–10,000 synaptic inputs and sends information to an axon, which branches to produce a similar number of synaptic outputs. Most excitatory postsynaptic terminals are associated with dendritic spines, small protrusions emanating from the dendritic surface (Nimchinsky et al. 2002; Alvarez and Sabatini 2007). Each spine has a volume of ∼0.1 femtoliter, and connects to the parent dendrite through a narrow neck, which acts as a diffusion barrier and compartmentalizes biochemical reactions. Ca2+ influx into spines initiates a cascade of biochemical signals leading to various forms of synaptic plasticity including long-term potentiation (LTP).Because LTP in hippocampal CA1 pyramidal neurons is a cellular mechanism that may underlie long-term memory formation, the signal transduction underlying LTP has been extensively studied by pharmacological and genetic methods (Bliss and Collingridge 1993; Derkach et al. 2007). It is now well established that LTP is induced by Ca2+ influx into dendritic spines through NMDA-type glutamate receptors (NMDARs), which induces the insertion of AMPA-type glutamate receptors (AMPARs) into the synapse, thereby increasing the sensitivity of the postsynaptic terminal to glutamate (Derkach et al. 2007; Kessels and Malinow 2009). An increase of release probability during LTP has also been reported (Enoki et al. 2009), and thus both pre- and postsynaptic mechanisms may contribute to LTP (Lisman and Raghavachari 2006).Manipulations of signal transduction using specific pharmacological inhibitors or genetic perturbations have identified many signaling pathways that connect Ca2+ to LTP induction. For example, LTP requires the activation of many signaling proteins, including Ca2+/calmodulin-dependent kinase II (CaMKII), extracellular signal-related kinase (ERK), Phoshoinositide 3 kinase (PI3K), protein kinase A and C, and GTPases such as Ras, Rab, and Rho (Kennedy et al. 2005). The list is continually growing, and the hundreds of implicated proteins form a complex signaling network whose contribution to LTP is still unclear (Bromberg et al. 2008).Signaling dynamics in neurons have traditionally been measured using biochemical analyses (Bromberg et al. 2008). However, the spatiotemporal resolution of conventional biochemistry is limited, restricting analysis to the time scale of many minutes and requiring the homogenization of tissue containing millions of synapses and other cellular elements. Furthermore, resolving synaptically induced changes in signaling by biochemical analysis typically requires stimulating many synapses at the same time, which may produce unintended effects, for instance, excitotoxicity or homeostatic plasticity.The size of dendritic spines is similar to the resolution of an optical microscope, permitting the optical analysis of biochemical signaling in each dendritic spine (Svoboda and Yasuda 2006). In particular, the advent of two-photon-based FRET techniques and the development of appropriate fluorescent reporters of specific biochemical reactions (see below) have provided readouts for signal transduction with high spatiotemporal resolution in live brain tissue (Svoboda and Yasuda 2006; Yasuda 2006). This has provided detailed information about the dynamics of signal transduction in spines and dendrites, and insights into the molecular mechanisms of synaptic plasticity.  相似文献   

11.
12.
Background:Otitis media with effusion is a common problem that lacks an evidence-based nonsurgical treatment option. We assessed the clinical effectiveness of treatment with a nasal balloon device in a primary care setting.Methods:We conducted an open, pragmatic randomized controlled trial set in 43 family practices in the United Kingdom. Children aged 4–11 years with a recent history of ear symptoms and otitis media with effusion in 1 or both ears, confirmed by tympanometry, were allocated to receive either autoinflation 3 times daily for 1–3 months plus usual care or usual care alone. Clearance of middle-ear fluid at 1 and 3 months was assessed by experts masked to allocation.Results:Of 320 children enrolled, those receiving autoinflation were more likely than controls to have normal tympanograms at 1 month (47.3% [62/131] v. 35.6% [47/132]; adjusted relative risk [RR] 1.36, 95% confidence interval [CI] 0.99 to 1.88) and at 3 months (49.6% [62/125] v. 38.3% [46/120]; adjusted RR 1.37, 95% CI 1.03 to 1.83; number needed to treat = 9). Autoinflation produced greater improvements in ear-related quality of life (adjusted between-group difference in change from baseline in OMQ-14 [an ear-related measure of quality of life] score −0.42, 95% CI −0.63 to −0.22). Compliance was 89% at 1 month and 80% at 3 months. Adverse events were mild, infrequent and comparable between groups.Interpretation:Autoinflation in children aged 4–11 years with otitis media with effusion is feasible in primary care and effective both in clearing effusions and improving symptoms and ear-related child and parent quality of life. Trial registration: ISRCTN, No. 55208702.Otitis media with effusion, also known as glue ear, is an accumulation of fluid in the middle ear, without symptoms or signs of an acute ear infection. It is often associated with viral infection.13 The prevalence rises to 46% in children aged 4–5 years,4 when hearing difficulty, other ear-related symptoms and broader developmental concerns often bring the condition to medical attention.3,5,6 Middle-ear fluid is associated with conductive hearing losses of about 15–45 dB HL.7 Resolution is clinically unpredictable,810 with about a third of cases showing recurrence.11 In the United Kingdom, about 200 000 children with the condition are seen annually in primary care.12,13 Research suggests some children seen in primary care are as badly affected as those seen in hospital.7,9,14,15 In the United States, there were 2.2 million diagnosed episodes in 2004, costing an estimated $4.0 billion.16 Rates of ventilation tube surgery show variability between countries,1719 with a declining trend in the UK.20Initial clinical management consists of reasonable temporizing or delay before considering surgery.13 Unfortunately, all available medical treatments for otitis media with effusion such as antibiotics, antihistamines, decongestants and intranasal steroids are ineffective and have unwanted effects, and therefore cannot be recommended.2123 Not only are antibiotics ineffective, but resistance to them poses a major threat to public health.24,25 Although surgery is effective for a carefully selected minority,13,26,27 a simple low-cost, nonsurgical treatment option could benefit a much larger group of symptomatic children, with the purpose of addressing legitimate clinical concerns without incurring excessive delays.Autoinflation using a nasal balloon device is a low-cost intervention with the potential to be used more widely in primary care, but current evidence of its effectiveness is limited to several small hospital-based trials28 that found a higher rate of tympanometric resolution of ear fluid at 1 month.2931 Evidence of feasibility and effectiveness of autoinflation to inform wider clinical use is lacking.13,28 Thus we report here the findings of a large pragmatic trial of the clinical effectiveness of nasal balloon autoinflation in a spectrum of children with clinically confirmed otitis media with effusion identified from primary care.  相似文献   

13.
HNK-1 (human natural killer-1) glyco-epitope, a sulfated glucuronic acid attached to N-acetyllactosamine on the nonreducing termini of glycans, is highly expressed in the nervous system. Our previous report showed that mice lacking a glucuronyltransferase (GlcAT-P), a key enzyme for biosynthesis of the HNK-1 epitope, showed reduced long term potentiation at hippocampal CA1 synapses. In this study, we identified an α-amino-3-hydroxy-5-methylisoxazole propionate (AMPA)-type glutamate receptor subunit, GluR2, which directly contributes to excitatory synaptic transmission and synaptic plasticity, as a novel HNK-1 carrier molecule. We demonstrated that the HNK-1 epitope is specifically expressed on the N-linked glycan(s) on GluR2 among the glutamate receptors tested, and the glycan structure, including HNK-1 on GluR2, was determined using liquid chromatography-tandem mass spectrometry. As for the function of HNK-1 on GluR2, we found that the GluR2 not carrying HNK-1 was dramatically endocytosed and expressed less on the cell surface compared with GluR2 carrying HNK-1 in both cultured hippocampal neurons and heterologous cells. These results suggest that HNK-1 stabilizes GluR2 on neuronal surface membranes and regulates the number of surface AMPA receptors. Moreover, we showed that the expression of the HNK-1 epitope enhanced the interaction between GluR2 and N-cadherin, which has important roles in AMPA receptor trafficking. Our findings suggest that the HNK-1 epitope on GluR2 regulates cell surface stability of GluR2 by modulating the interaction with N-cadherin.HNK-1 glyco-epitope (HSO3-3GlcAβ1–3Galβ1–4GlcNAc) is characteristically expressed on some cell adhesion molecules (NCAM, L1, and MAG, etc.) and extracellular matrix molecules (tenascin-R and phosphacan, etc.) in the nervous system (1). It has been reported that HNK-1 mediates the interaction of these adhesion molecules, thereby controlling their functions, including cell-to-cell adhesion (2), migration (3), and neurite extension (4). The unique structural feature of the HNK-1 epitope is the sulfated glucuronic acid, because sialic acids are usually attached to the terminal galactose residue of the inner N-acetyllactosamine structure (Galβ1–4GlcNAc) on various glycoproteins. HNK-1 is sequentially biosynthesized by one of two glucuronyltransferases (GlcAT-P or GlcAT-S)3 (5, 6) and a sulfotransferase (HNK-1ST) (7). These enzymes are thought to localize and function in the Golgi apparatus, especially the trans-Golgi to trans-Golgi network, like most sialyltransferases and galactosyltransferases (8).We previously demonstrated that mice deficient in GlcAT-P showed an almost complete loss of HNK-1 expression in the brain and exhibited reduced LTP in hippocampal CA1 synapses (9). Similarly, HNK-1ST-deficient mice also exhibited a reduction of LTP, and several other studies also revealed that HNK-1 is associated with neural plasticity (1012). A recent study showed that β4-galactosyltransferase-2 synthesizes the glycan backbone structure of HNK-1, Galβ1–4GlcNAc. The mice lacking β4-galactosyltransferase-2 showed decreased HNK-1 expression in their brains and also exhibited impaired learning and memory (13). Overall, these studies suggest that HNK-1 plays unique functional roles in some types of neuronal plasticity, but the molecular mechanisms of HNK-1 remain unclear.AMPA-type glutamate receptors mediate most of the fast excitatory synaptic transmissions in the mammalian brain and control synaptic strength. The regulated trafficking of AMPA receptors to the postsynaptic membrane is thought to be a major mechanism contributing to long lasting changes in synaptic strength, including LTP and long term depression (14, 15). AMPA receptors are mainly heterotetrameric channels assembled from the subunits GluR1 to GluR4, and all subunits have 4–6 potential N-glycosylation sites in their extracellular domains (16). Few studies have focused on the function of N-glycosylation in AMPA receptors. Some investigations showed that AMPA receptor subunits expressed at both the cell surface and synaptic sites possess the mature glycosylated form (17, 18), but it is generally accepted that N-glycosylation is not essential for their channel function or ligand binding (19, 20).In this study, we searched for a candidate molecule(s) responsible for the defects in synaptic plasticity seen in GlcAT-P-deficient mice. We found that the HNK-1 epitope is mainly expressed on a specific molecule in the hippocampal postsynaptic density (PSD) fraction. We focused on the molecule and identified a subunit of AMPA-type glutamate receptors, GluR2, as a novel HNK-1 carrier protein. Furthermore, we showed that the loss of HNK-1 epitope on GluR2 greatly increases both constitutive and regulated endocytosis of GluR2, resulting in a decrease in the amount of surface GluR2 in cultured hippocampal neurons and CHO cells. This is the first report demonstrating that the N-glycan on GluR2 regulates its protein function, and our results suggest that HNK-1 epitope on GluR2 is an important factor for synaptic plasticity.  相似文献   

14.
The cysteine protease caspase-3, best known as an executioner of cell death in apoptosis, also plays a non-apoptotic role in N-methyl-d-aspartate receptor-dependent long-term depression of synaptic transmission (NMDAR-LTD) and α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptor endocytosis in neurons. The mechanism by which caspase-3 regulates LTD and AMPA receptor endocytosis, however, remains unclear. Here, we addressed this question by using an enzymatic N-terminal peptide enrichment method and mass spectrometry to identify caspase-3 substrates in neurons. Of the many candidates revealed by this proteomic study, we have confirmed BASP1, Dbn1, and Gap43 as true caspase-3 substrates. Moreover, in hippocampal neurons, Gap43 mutants deficient in caspase-3 cleavage inhibit AMPA receptor endocytosis and LTD. We further demonstrated that Gap43, a protein well-known for its functions in axons, is also localized at postsynaptic sites. Our study has identified Gap43 as a key caspase-3 substrate involved in LTD and AMPA receptor endocytosis, uncovered a novel postsynaptic function for Gap43 and provided new insights into how long-term synaptic depression is induced.Synaptic plasticity (the ability of synapses to change in strength) plays an important role in brain development and cognitive function, including learning and memory. N-methyl-d-aspartate receptor (NMDAR)1-dependent long-term depression of synaptic transmission (LTD) is a major form of synaptic plasticity that leads to long-lasting decreases in synaptic strength. In NMDAR-LTD, synaptic depression is mainly mediated by removal of α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors from the postsynaptic membrane through endocytosis (13).Caspases are cysteine-dependent proteases that cleave after an aspartate residue (46). Primary specificity for aspartate at the cleavage site is so rare in mammalian proteases that only granzyme B, a serine protease derived from lymphocytes, is known to also have such a property (7, 8). Caspases are best known for their pro-apoptotic function in programmed cell death, or apoptosis (9). Caspases activated at the end of a caspase cascade are “effector” caspases, and among them caspase-3 is the dominant one (10). In addition to apoptosis, caspases also play non-apoptotic roles such as in cell differentiation, dendritic development, and memory consolidation (1114). For instance, our earlier studies show that NMDAR-LTD requires moderate and transient caspase-3 activation, which does not induce cell death (15, 16).In LTD, caspase-3 is activated by the mitochondrial pathway (15, 16). With the opening of NMDA receptors, calcineurin and protein phosphatase 1 are activated to dephosphorylate the Bcl-2 family protein BAD. Dephosphorylated BAD then translocates to the mitochondria, activating BAX, which is also a member of the Bcl-2 family protein. The subsequent release of cytochrome-c from mitochondria leads to the sequential activation of caspase-9 and caspase-3. Active caspase-3 induces AMPA receptor endocytosis, and therefore depression of synaptic strength. The mechanism by which caspase-3 promotes AMPA receptor endocytosis, however, remains to be determined.Caspases'' cellular functions are primarily mediated by the proteolysis of caspase substrates, resulting in change of their functions. Identification of specific proteins cleaved by caspases is therefore key to understanding the mechanisms mediating their biological functions. Several proteomic approaches, developed specifically for this purpose, have led to the identification of more than 2000 caspase cleavage sites (1722). In particular, the recently developed subtiligase-based method for isolating proteolytic products has led to the identification of >1,000 putative caspase substrates in human samples (21, 23, 24). Subtiligase is an engineered peptide ligase that ligates esterified peptides onto the N termini of proteins or peptides through free α-amines (25). Because the majority of eukaryotic proteins are N-terminally acetylated—and therefore blocked from subtiligase labeling (26)—subtiligase can couple synthetic tagged peptides selectively to the free N-terminal α-amines of proteins derived from proteolysis. These peptide-conjugated proteolytic products can then be affinity-purified, digested with trypsin and sequenced by mass spectrometry. Identification of peptides ligated to the tagged peptide by subtiligase allows researchers to determine cleavage sites within the substrates.In this study, the subtiligase-based proteomic method was used to find capase-3 substrates in rat neurons, resulting in the identification of 81 putative aspartate cleavage sites in 56 proteins. Of these, 37 proteins (human and a single rat orthologs) were not previously reported in the CASBAH database (20), and 13 (human orthologs) are not included in the DegraBase data set compiled using the subtiligase methodology in non-neuronal tissue (21). Using complimentary methods, we further confirmed that, both in vivo and in vitro, caspase-3 cleaves three of these candidate substrates: growth associated protein 43 (Gap43), drebrin (Dbn1), and brain acid soluble protein 1 (BASP1). Surprisingly, we also found that AMPA receptor endocytosis and LTD induction both require caspase-3 to cleave Gap43, a protein well known for its presynaptic functions, at the sites identified by our study.  相似文献   

15.
Many biological processes involve the mechanistic/mammalian target of rapamycin complex 1 (mTORC1). Thus, the challenge of deciphering mTORC1-mediated functions during normal and pathological states in the central nervous system is challenging. Because mTORC1 is at the core of translation, we have investigated mTORC1 function in global and regional protein expression. Activation of mTORC1 has been generally regarded to promote translation. Few but recent works have shown that suppression of mTORC1 can also promote local protein synthesis. Moreover, excessive mTORC1 activation during diseased states represses basal and activity-induced protein synthesis. To determine the role of mTORC1 activation in protein expression, we have used an unbiased, large-scale proteomic approach. We provide evidence that a brief repression of mTORC1 activity in vivo by rapamycin has little effect globally, yet leads to a significant remodeling of synaptic proteins, in particular those proteins that reside in the postsynaptic density. We have also found that curtailing the activity of mTORC1 bidirectionally alters the expression of proteins associated with epilepsy, Alzheimer''s disease, and autism spectrum disorder—neurological disorders that exhibit elevated mTORC1 activity. Through a protein–protein interaction network analysis, we have identified common proteins shared among these mTORC1-related diseases. One such protein is Parkinson protein 7, which has been implicated in Parkinson''s disease, yet not associated with epilepsy, Alzheimers disease, or autism spectrum disorder. To verify our finding, we provide evidence that the protein expression of Parkinson protein 7, including new protein synthesis, is sensitive to mTORC1 inhibition. Using a mouse model of tuberous sclerosis complex, a disease that displays both epilepsy and autism spectrum disorder phenotypes and has overactive mTORC1 signaling, we show that Parkinson protein 7 protein is elevated in the dendrites and colocalizes with the postsynaptic marker postsynaptic density-95. Our work offers a comprehensive view of mTORC1 and its role in regulating regional protein expression in normal and diseased states.The mechanistic/mammalian target of rapamycin complex 1 (mTORC1)1 is a serine/threonine protein kinase that is highly expressed in many cell types (1). In the brain, mTORC1 tightly coordinates different synaptic plasticities — long-term potentiation (LTP) and long-term depression (LTD) — the molecular correlates of learning and memory (25). Because mTORC1 is at the core of many synaptic signaling pathways downstream of glutamate and neurotrophin receptors, many hypothesize that dysregulated mTORC1 signaling underlies cognitive deficits observed in several neurodegenerative diseases (3, 617). For example, mTORC1 and its downstream targets are hyperactive in human brains diagnosed with Alzheimer''s disease (AD) (1820). Additionally in animal models of autism spectrum disorder (ASD), altered mTORC1 signaling contributes to the observed synaptic dysfunction and aberrant network connectivity (13, 15, 2127). Furthermore, epilepsy, which is common in AD and ASD, has enhanced mTORC1 activity (2832).Phosphorylation of mTORC1, considered the active form, is generally regarded to promote protein synthesis (33). Thus, many theorize that diseases with overactive mTORC1 arise from excessive protein synthesis (14). Emerging data, however, show that suppressing mTORC1 activation can trigger local translation in neurons (34, 35). Pharmacological antagonism of N-methyl-d-aspartate (NMDA) receptors, a subtype of glutamate receptors that lies upstream of mTOR activation, promotes the synthesis of the voltage-gated potassium channel, Kv1.1, in dendrites (34, 35). Consistent with these results, in models of temporal lobe epilepsy there is a reduction in the expression of voltage-gated ion channels including Kv1.1 (30, 31, 36). Interestingly in a model of focal neocortical epilepsy, overexpression of Kv1.1 blocked seizure activity (37). Because both active and inactive mTORC1 permit protein synthesis, we sought to determine the proteins whose expression is altered when mTORC1 phosphorylation is reduced in vivo.Rapamycin is an FDA-approved, immunosuppressive drug that inhibits mTORC1 activity (38). We capitalized on the ability of rapamycin to reduce mTORC1 activity in vivo and the unbiased approach of mass spectrometry to identify changes in protein expression. Herein, we provide evidence that mTORC1 activation bidirectionally regulates protein expression, especially in the PSD where roughly an equal distribution of proteins dynamically appear and disappear. Remarkably, using protein–protein interaction networks facilitated the novel discovery that PARK7, a protein thus far only implicated in Parkinson''s disease, (1) is up-regulated by increased mTORC1 activity, (2) resides in the PSD only when mTORC1 is active, and (3) is aberrantly expressed in a rodent model of TSC, an mTORC1-related disease that has symptoms of epilepsy and autism. Collectively, these data provide the first comprehensive list of proteins whose abundance or subcellular distributions are altered with acute changes in mTORC1 activity in vivo.  相似文献   

16.

Background:

The gut microbiota is essential to human health throughout life, yet the acquisition and development of this microbial community during infancy remains poorly understood. Meanwhile, there is increasing concern over rising rates of cesarean delivery and insufficient exclusive breastfeeding of infants in developed countries. In this article, we characterize the gut microbiota of healthy Canadian infants and describe the influence of cesarean delivery and formula feeding.

Methods:

We included a subset of 24 term infants from the Canadian Healthy Infant Longitudinal Development (CHILD) birth cohort. Mode of delivery was obtained from medical records, and mothers were asked to report on infant diet and medication use. Fecal samples were collected at 4 months of age, and we characterized the microbiota composition using high-throughput DNA sequencing.

Results:

We observed high variability in the profiles of fecal microbiota among the infants. The profiles were generally dominated by Actinobacteria (mainly the genus Bifidobacterium) and Firmicutes (with diverse representation from numerous genera). Compared with breastfed infants, formula-fed infants had increased richness of species, with overrepresentation of Clostridium difficile. Escherichia–Shigella and Bacteroides species were underrepresented in infants born by cesarean delivery. Infants born by elective cesarean delivery had particularly low bacterial richness and diversity.

Interpretation:

These findings advance our understanding of the gut microbiota in healthy infants. They also provide new evidence for the effects of delivery mode and infant diet as determinants of this essential microbial community in early life.The human body harbours trillions of microbes, known collectively as the “human microbiome.” By far the highest density of commensal bacteria is found in the digestive tract, where resident microbes outnumber host cells by at least 10 to 1. Gut bacteria play a fundamental role in human health by promoting intestinal homeostasis, stimulating development of the immune system, providing protection against pathogens, and contributing to the processing of nutrients and harvesting of energy.1,2 The disruption of the gut microbiota has been linked to an increasing number of diseases, including inflammatory bowel disease, necrotizing enterocolitis, diabetes, obesity, cancer, allergies and asthma.1 Despite this evidence and a growing appreciation for the integral role of the gut microbiota in lifelong health, relatively little is known about the acquisition and development of this complex microbial community during infancy.3Two of the best-studied determinants of the gut microbiota during infancy are mode of delivery and exposure to breast milk.4,5 Cesarean delivery perturbs normal colonization of the infant gut by preventing exposure to maternal microbes, whereas breastfeeding promotes a “healthy” gut microbiota by providing selective metabolic substrates for beneficial bacteria.3,5 Despite recommendations from the World Health Organization,6 the rate of cesarean delivery has continued to rise in developed countries and rates of breastfeeding decrease substantially within the first few months of life.7,8 In Canada, more than 1 in 4 newborns are born by cesarean delivery, and less than 15% of infants are exclusively breastfed for the recommended duration of 6 months.9,10 In some parts of the world, elective cesarean deliveries are performed by maternal request, often because of apprehension about pain during childbirth, and sometimes for patient–physician convenience.11The potential long-term consequences of decisions regarding mode of delivery and infant diet are not to be underestimated. Infants born by cesarean delivery are at increased risk of asthma, obesity and type 1 diabetes,12 whereas breastfeeding is variably protective against these and other disorders.13 These long-term health consequences may be partially attributable to disruption of the gut microbiota.12,14Historically, the gut microbiota has been studied with the use of culture-based methodologies to examine individual organisms. However, up to 80% of intestinal microbes cannot be grown in culture.3,15 New technology using culture-independent DNA sequencing enables comprehensive detection of intestinal microbes and permits simultaneous characterization of entire microbial communities. Multinational consortia have been established to characterize the “normal” adult microbiome using these exciting new methods;16 however, these methods have been underused in infant studies. Because early colonization may have long-lasting effects on health, infant studies are vital.3,4 Among the few studies of infant gut microbiota using DNA sequencing, most were conducted in restricted populations, such as infants delivered vaginally,17 infants born by cesarean delivery who were formula-fed18 or preterm infants with necrotizing enterocolitis.19Thus, the gut microbiota is essential to human health, yet the acquisition and development of this microbial community during infancy remains poorly understood.3 In the current study, we address this gap in knowledge using new sequencing technology and detailed exposure assessments20 of healthy Canadian infants selected from a national birth cohort to provide representative, comprehensive profiles of gut microbiota according to mode of delivery and infant diet.  相似文献   

17.

Background:

Polymyalgia rheumatica is one of the most common inflammatory rheumatologic conditions in older adults. Other inflammatory rheumatologic disorders are associated with an excess risk of vascular disease. We investigated whether polymyalgia rheumatica is associated with an increased risk of vascular events.

Methods:

We used the General Practice Research Database to identify patients with a diagnosis of incident polymyalgia rheumatica between Jan. 1, 1987, and Dec. 31, 1999. Patients were matched by age, sex and practice with up to 5 patients without polymyalgia rheumatica. Patients were followed until their first vascular event (cardiovascular, cerebrovascular, peripheral vascular) or the end of available records (May 2011). All participants were free of vascular disease before the diagnosis of polymyalgia rheumatica (or matched date). We used Cox regression models to compare time to first vascular event in patients with and without polymyalgia rheumatica.

Results:

A total of 3249 patients with polymyalgia rheumatica and 12 735 patients without were included in the final sample. Over a median follow-up period of 7.8 (interquartile range 3.3–12.4) years, the rate of vascular events was higher among patients with polymyalgia rheumatica than among those without (36.1 v. 12.2 per 1000 person-years; adjusted hazard ratio 2.6, 95% confidence interval 2.4–2.9). The increased risk of a vascular event was similar for each vascular disease end point. The magnitude of risk was higher in early disease and in patients younger than 60 years at diagnosis.

Interpretation:

Patients with polymyalgia rheumatica have an increased risk of vascular events. This risk is greatest in the youngest age groups. As with other forms of inflammatory arthritis, patients with polymyalgia rheumatica should have their vascular risk factors identified and actively managed to reduce this excess risk.Inflammatory rheumatologic disorders such as rheumatoid arthritis,1,2 systemic lupus erythematosus,2,3 gout,4 psoriatic arthritis2,5 and ankylosing spondylitis2,6 are associated with an increased risk of vascular disease, especially cardiovascular disease, leading to substantial morbidity and premature death.26 Recognition of this excess vascular risk has led to management guidelines advocating screening for and management of vascular risk factors.79Polymyalgia rheumatica is one of the most common inflammatory rheumatologic conditions in older adults,10 with a lifetime risk of 2.4% for women and 1.7% for men.11 To date, evidence regarding the risk of vascular disease in patients with polymyalgia rheumatica is unclear. There are a number of biologically plausible mechanisms between polymyalgia rheumatica and vascular disease. These include the inflammatory burden of the disease,12,13 the association of the disease with giant cell arteritis (causing an inflammatory vasculopathy, which may lead to subclinical arteritis, stenosis or aneurysms),14 and the adverse effects of long-term corticosteroid treatment (e.g., diabetes, hypertension and dyslipidemia).15,16 Paradoxically, however, use of corticosteroids in patients with polymyalgia rheumatica may actually decrease vascular risk by controlling inflammation.17 A recent systematic review concluded that although some evidence exists to support an association between vascular disease and polymyalgia rheumatica,18 the existing literature presents conflicting results, with some studies reporting an excess risk of vascular disease19,20 and vascular death,21,22 and others reporting no association.2326 Most current studies are limited by poor methodologic quality and small samples, and are based on secondary care cohorts, who may have more severe disease, yet most patients with polymyalgia rheumatica receive treatment exclusively in primary care.27The General Practice Research Database (GPRD), based in the United Kingdom, is a large electronic system for primary care records. It has been used as a data source for previous studies,28 including studies on the association of inflammatory conditions with vascular disease29 and on the epidemiology of polymyalgia rheumatica in the UK.30 The aim of the current study was to examine the association between polymyalgia rheumatica and vascular disease in a primary care population.  相似文献   

18.
The erythropoietin receptor (EpoR) was discovered and described in red blood cells (RBCs), stimulating its proliferation and survival. The target in humans for EpoR agonists drugs appears clear—to treat anemia. However, there is evidence of the pleitropic actions of erythropoietin (Epo). For that reason, rhEpo therapy was suggested as a reliable approach for treating a broad range of pathologies, including heart and cardiovascular diseases, neurodegenerative disorders (Parkinson’s and Alzheimer’s disease), spinal cord injury, stroke, diabetic retinopathy and rare diseases (Friedreich ataxia). Unfortunately, the side effects of rhEpo are also evident. A new generation of nonhematopoietic EpoR agonists drugs (asialoEpo, Cepo and ARA 290) have been investigated and further developed. These EpoR agonists, without the erythropoietic activity of Epo, while preserving its tissue-protective properties, will provide better outcomes in ongoing clinical trials. Nonhematopoietic EpoR agonists represent safer and more effective surrogates for the treatment of several diseases such as brain and peripheral nerve injury, diabetic complications, renal ischemia, rare diseases, myocardial infarction, chronic heart disease and others.In principle, the erythropoietin receptor (EpoR) was discovered and described in red blood cell (RBC) progenitors, stimulating its proliferation and survival. Erythropoietin (Epo) is mainly synthesized in fetal liver and adult kidneys (13). Therefore, it was hypothesized that Epo act exclusively on erythroid progenitor cells. Accordingly, the target in humans for EpoR agonists drugs (such as recombinant erythropoietin [rhEpo], in general, called erythropoiesis-simulating agents) appears clear (that is, to treat anemia). However, evidence of a kaleidoscope of pleitropic actions of Epo has been provided (4,5). The Epo/EpoR axis research involved an initial journey from laboratory basic research to clinical therapeutics. However, as a consequence of clinical observations, basic research on Epo/EpoR comes back to expand its clinical therapeutic applicability.Although kidney and liver have long been considered the major sources of synthesis, Epo mRNA expression has also been detected in the brain (neurons and glial cells), lung, heart, bone marrow, spleen, hair follicles, reproductive tract and osteoblasts (617). Accordingly, EpoR was detected in other cells, such as neurons, astrocytes, microglia, immune cells, cancer cell lines, endothelial cells, bone marrow stromal cells and cells of heart, reproductive system, gastrointestinal tract, kidney, pancreas and skeletal muscle (1827). Conversely, Sinclair et al.(28) reported data questioning the presence or function of EpoR on nonhematopoietic cells (endothelial, neuronal and cardiac cells), suggesting that further studies are needed to confirm the diversity of EpoR. Elliott et al.(29) also showed that EpoR is virtually undetectable in human renal cells and other tissues with no detectable EpoR on cell surfaces. These results have raised doubts about the preclinical basis for studies exploring pleiotropic actions of rhEpo (30).For the above-mentioned data, a return to basic research studies has become necessary, and many studies in animal models have been initiated or have already been performed. The effect of rhEpo administration on angiogenesis, myogenesis, shift in muscle fiber types and oxidative enzyme activities in skeletal muscle (4,31), cardiac muscle mitochondrial biogenesis (32), cognitive effects (31), antiapoptotic and antiinflammatory actions (3337) and plasma glucose concentrations (38) has been extensively studied. Neuro- and cardioprotection properties have been mainly described. Accordingly, rhEpo therapy was suggested as a reliable approach for treating a broad range of pathologies, including heart and cardiovascular diseases, neurodegenerative disorders (Parkinson’s and Alzheimer’s disease), spinal cord injury, stroke, diabetic retinopathy and rare diseases (Friedreich ataxia).Unfortunately, the side effects of rhEpo are also evident. Epo is involved in regulating tumor angiogenesis (39) and probably in the survival and growth of tumor cells (25,40,41). rhEpo administration also induces serious side effects such as hypertension, polycythemia, myocardial infarction, stroke and seizures, platelet activation and increased thromboembolic risk, and immunogenicity (4246), with the most common being hypertension (47,48). A new generation of nonhematopoietic EpoR agonists drugs have hence been investigated and further developed in animals models. These compounds, namely asialoerythropoietin (asialoEpo) and carbamylated Epo (Cepo), were developed for preserving tissue-protective properties but reducing the erythropoietic activity of native Epo (49,50). These drugs will provide better outcome in ongoing clinical trials. The advantage of using nonhematopoietic Epo analogs is to avoid the stimulation of hematopoiesis and thereby the prevention of an increased hematocrit with a subsequent procoagulant status or increased blood pressure. In this regard, a new study by van Rijt et al. has shed new light on this topic (51). A new nonhematopoietic EpoR agonist analog named ARA 290 has been developed, promising cytoprotective capacities to prevent renal ischemia/reperfusion injury (51). ARA 290 is a short peptide that has shown no safety concerns in preclinical and human studies. In addition, ARA 290 has proven efficacious in cardiac disorders (52,53), neuropathic pain (54) and sarcoidosis-induced chronic neuropathic pain (55). Thus, ARA 290 is a novel nonhematopoietic EpoR agonist with promising therapeutic options in treating a wide range of pathologies and without increased risks of cardiovascular events.Overall, this new generation of EpoR agonists without the erythropoietic activity of Epo while preserving tissue-protective properties of Epo will provide better outcomes in ongoing clinical trials (49,50). Nonhematopoietic EpoR agonists represent safer and more effective surrogates for the treatment of several diseases, such as brain and peripheral nerve injury, diabetic complications, renal ischemia, rare diseases, myocardial infarction, chronic heart disease and others.  相似文献   

19.
20.
We purified from rat brain a novel actin filament (F-actin)–binding protein of ∼180 kD (p180), which was specifically expressed in neural tissue. We named p180 neurabin (neural tissue–specific F-actin– binding protein). We moreover cloned the cDNA of neurabin from a rat brain cDNA library and characterized native and recombinant proteins. Neurabin was a protein of 1,095 amino acids with a calculated molecular mass of 122,729. Neurabin had one F-actin–binding domain at the NH2-terminal region, one PSD-95, DlgA, ZO-1–like domain at the middle region, a domain known to interact with transmembrane proteins, and domains predicted to form coiled-coil structures at the COOH-terminal region. Neurabin bound along the sides of F-actin and showed F-actin–cross-linking activity. Immunofluorescence microscopic analysis revealed that neurabin was highly concentrated in the synapse of the developed neurons. Neurabin was also concentrated in the lamellipodia of the growth cone during the development of neurons. Moreover, a study on suppression of endogenous neurabin in primary cultured rat hippocampal neurons by treatment with an antisense oligonucleotide showed that neurabin was involved in the neurite formation. Neurabin is a candidate for key molecules in the synapse formation and function.During the development of the nervous system, the distal tip of the elongating axon—the growth cone—actively migrates toward its target cell in response to the combined actions of attractive and repulsive guidance molecules in the extracellular environment (Garrity and Zipursky, 1995; Keynes and Cook, 1995; Chiba and Keshishian, 1996; Culotti and Kolodkin, 1996; Friedman and O''Leary, 1996; Tessier-Lavigne and Goodman, 1996). When the growth cone contacts with the target cell, it is transformed into the functional presynaptic terminal (Garrity and Zipursky, 1995; Chiba and Kishishian, 1996). The actin cytoskeleton has been shown to play crucial roles in these processes of the synapse formation (Mitchison and Kirschner, 1988; Smith, 1988; Bentley and O''Connor, 1994; Lin et al., 1994; Mackay et al., 1995; Tanaka and Sabry, 1995).In the developing nervous system, the actin cytoskeleton is prominent in two structural domains of the growth cone, filopodia and lamellipodia (Mitchison and Kirschner, 1988; Smith, 1988; Bentley and O''Connor, 1994; Lin et al., 1994; Mackay et al., 1995; Tanaka and Sabry, 1995). In these domains, actin filament (F-actin)1 assembled at the leading edge are transported into the center of the growth cone and disassembled there. It has been suggested that this retrograde flow of F-actin is crucial for the growth cone motility. Drugs that disrupt F-actin have also been shown to cause the lamellipodial and filopodial collapse and block the ability of neurons to extend the growth cone in the correct direction (Marsh and Letourneau, 1984; Forscher and Smith, 1988; Bentley and Toroian-Raymond, 1986; Chien et al., 1993). These results suggest that the actin cytoskeleton regulates not only the growth cone motility but also the growth cone directionality. Recently, a variety of guidance molecules and their receptors have been identified (Garrity and Zipursky, 1995; Keynes and Cook, 1995; Chiba and Keshishian, 1996; Culotti and Kolodkin, 1996; Friedman and O''Leary, 1996; Tessier-Lavigne and Goodman, 1996). However, which molecules of the actin cytoskeleton are essential for the growth cone motility and directionality is not well understood.When the growth cone contacts with the target cell, the target cell regulates the development of the presynaptic nerve terminal and the formation of the functional synapse (Bowe and Fallon, 1995; Chiba and Keshishian, 1996). In the established nervous system, the presynaptic and postsynaptic membranes get aligned in space and constitute the synaptic junction (Burns and Augustine, 1995; Garner and Kindler, 1996). Electron microscopic studies have revealed the ultrastructural features of the synaptic junction (Burns and Augustine, 1995; Garner and Kindler, 1996). The presynaptic cytoplasm is characterized by synaptic vesicles (SVs). SVs are not distributed uniformly; SVs cluster together in the vicinity of the presynaptic plasma membrane, where F-actin forms a network and is associated with the presynaptic plasma membrane (Hirokawa et al., 1989). Most SVs within the cluster are linked through thin strands to each other, to F-actin, or to both (Hirokawa et al., 1989). A subset of SVs within the cluster are attached by fine filamentous threads to neurotransmitter release zone at the presynaptic plasma membrane (Hirokawa et al., 1989). The presynaptic submembranous cytoskeleton is assumed to be involved in recruiting Ca2+ channels and the components of the SV fusion complex, delivering SVs to the neurotransmitter release zone, and keeping them in place (Burns and Augustine, 1995; Garner and Kindler, 1996). At the inner surface of the post-synaptic plasma membrane, there is an electron dense thickening, called postsynaptic density. The postsynaptic density is assumed to be involved in the selective targeting and accumulation of ion channels and receptors (Burns and Augustine, 1995; Garner and Kindler, 1996). It is also assumed that the presynaptic and postsynaptic submembranous cytoskeleton elements are linked to cell adhesion molecules to regulate the synaptic stabilization and plasticity (Fields and Itoh, 1996; Garner and Kindler, 1996). The presynaptic and postsynaptic submembranous cytoskeleton elements are thought to be composed of spectrin/fodrin, ankyrin, α-adducin, and protein 4.1 isoforms and to be linked to F-actin through these cytoskeleton proteins (Garner and Kindler, 1996). However, little is known about which molecules of the submembranous cytoskeleton are essential for the synaptic transmission and/or the synaptic stabilization.To understand the regulation of the actin cytoskeleton during and after the development of the nervous system, it is of crucial importance to identify F-actin–binding proteins implicated in the synapse formation and function. Therefore, we attempted here to isolate neural tissue–specific F-actin–binding proteins. We isolated a novel neural tissue–specific F-actin–binding protein from rat brain, which may be involved in neurite formation, and named it neurabin (neural tissue–specific F-actin–binding protein).  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号