Skip to main content
UKPMC Funders Author Manuscripts logoLink to UKPMC Funders Author Manuscripts
. Author manuscript; available in PMC: 2013 Jun 17.
Published in final edited form as: Curr Opin Genet Dev. 2011 Apr 27;21(4):458–464. doi: 10.1016/j.gde.2011.04.004

Translational regulation in growth cones

Hosung Jung 1,*, Catherine M O’Hare 1,*, Christine E Holt 1
PMCID: PMC3683644  EMSID: EMS49271  PMID: 21530230

Abstract

Axonal growth cones (GCs) steer in response to extrinsic cues using mechanisms that include local protein synthesis. This adaptive form of gene regulation occurs with spatial precision and depends on subcellular mRNA localisation. Recent genome-wide studies have shown unexpectedly complex and dynamically changing mRNA repertoires in growing axons and GCs. Axonal targeting of some transcripts seems to be highly selective and involves sequence diversity in non-coding regions generated by transcriptional and/or posttranscriptional mechanisms. New evidence reports direct coupling of a guidance receptor to the protein synthesis machinery and other findings demonstrate that some guidance cues can repress translation. The recent findings shed further light on the exquisitely regulated process that enables distant cellular compartments to respond to local stimuli.

Introduction

The axonal growth cone (GC) represents a unique signaling compartment, existing for the purpose of guiding an axon to its postsynaptic target. On reaching the target, the GC transforms into the developing axonal arbor with presynaptic terminals. Since the first observation of β-actin mRNA in the GC [1], mounting evidence has shown that this transient structure uses local mRNA translation to respond directionally to stimuli, contributing to its autonomous function [210]. For example, the chemotropic responses of GCs to Netrin-1, BDNF, Sema3A and Slit2 require local mRNA translation [47,9,11], and axonal mRNA translation is necessary for efficient GC regeneration [12]. A requirement for axonal protein synthesis (PS) for cue-induced responses was not seen in one study [13], although it may have been masked by the high cue concentrations used. Studies in recent years have indicated that remarkable complexity exists in the regulation of the GC’s proteome and, moreover, that growing axons have adopted some specialised mechanisms for processing newly synthesised membrane and secreted proteins. Although clear evidence exists that axons can locally synthesise transmembrane and secreted proteins (e.g. snail egg-laying hormone [14], CGRP [15], kappa opioid receptor [16], and EphA2 [3]), puzzlingly, rough ER (RER) and Golgi necessary for the processing and secretion of these types of proteins have rarely been detected ultrastructurally in axons [17]. A recent study helps to solve this mystery by providing immunocytochemical and functional evidence for RER and Golgi in axons and GCs [17], and suggests that these trafficking ‘outposts’ have evolved non-canonical morphology to handle the dynamic demands of growing axons.

The focus of this review is the GC but, due to its small size, most experimental studies use entire axons. Therefore, we will discuss recent results on local translation in both axons and GCs. It should be kept in mind, however, that the axon shaft and the GC are functionally distinct compartments (e.g. gradient sensing and turning occurs exclusively in GCs) that can employ specific RNA-based mechanisms.

Axonal mRNA repertoire

The number of identified axonally localised mRNAs has grown considerably by the use of more sensitive detection techniques and improved methods for obtaining isolated axons. Recent microarray studies identified around 2000 transcripts in murine retinal axonal GCs [18••], primary sensory axons [19••], and in cortical and hippocampal neuronal axons [20••], and up to 11,000 mRNAs were identified in sympathetic neuronal axons by SAGE analysis [21••]. Despite this remarkable complexity, different axonal mRNA repertoires show a surprising similarity as a group, representing 6–10% of the total cellular transcripts and, reassuringly, are composed of functionally similar mRNAs [18••,19••,20••]. For example, mRNAs encoding proteins involved in PS, molecular transport and mitochondrial maintenance invariably represent major categories in four independent screens using different neurons [18••,19••,20••,21••]. Conversely, there are distinct differences that point to cell type-specific roles for particular mRNAs. For example, mRNA encoding Impa1, a key enzyme in the inositol cycle, is the most abundant mRNA found in sympathetic axons [21••], but it is not reported in other axonal profiling studies. Similarly, CREB mRNA is present in dorsal root ganglion neuronal axons where its translation helps to promote neuronal survival [22] but is absent in sympathetic neuronal axons [21••].

Many axonally localised mRNAs are highly enriched in the axon compared to the cell body, as revealed by comparative bioinformatics analyses, suggesting that anterograde transport, rather than overflow from the cell body, accounts for their axonal localisation [21••]. Moreover, using laser capture microdissection to isolate the GC compartment or the axon compartment specifically, Zivraj et al. showed that certain mRNAs are enriched in the GC over the axon, suggesting that the GC is a distinct subcellular compartment rather than a simple extension of the axon [18••]. Interestingly, GC mRNA repertoires showed functionally relevant developmental changes. For example, mRNAs encoding presynaptic machinery reside in the GCs of target-arriving, but not pathfinding, axons suggesting that the composition of mRNAs changes dynamically to meet the changing demands of the GC [18••]. In support of this, presynaptic protein-encoding mRNAs show increased axonal localisation after axotomy in cultured cortical neurons [20••]. These results provide a clear example of how axonal PS could be used to regulate context-dependent responses during development and regeneration, in accordance with the notion that axonal PS confers plasticity. A recent study identified over 300 transcripts in uninjured mature CNS neuronal axons [20••], lending support to early reports of PS in adult axons [2326], and suggesting a requirement for PS in fully mature axons.

Mechanisms regulating general translational activity

In addition to altering its mRNA repertoire during development, the GC must possess mechanisms to regulate local PS on a rapid timescale in response to guidance cues. Work in the last two years has begun to reveal how guidance cue receptors are linked to PS machinery in the GC. Most PS-inducing guidance cues identified so far activate various signaling cascades converging on the mTORC1-mediated activation of cap-dependent mRNA translation [9]. Sahin’s group recently showed that EphrinA, a non-PS-inducing repulsive guidance cue, represses mTORC1 activity [27••]. tsc2 mutant mice exhibit defects in topographic mapping, similar to ephrinA knockout mice, and retinal GCs are less responsive to EphrinA in vitro. Furthermore, EphrinA normally increases the activity of Tsc2 through its receptor EphA, resulting in a decrease in downstream mTORC1 activity and decreased axonal PS (Figure 1b). Previously, Ephrins were shown to induce GC collapse in a PS inhibitor-insensitive manner, and therefore have been regarded as non-PS-inducing cues [28]. This new evidence, however, suggests the interesting possibility that some non-PS-inducing cues may, in fact, repress axonal PS. Intriguingly, Sahin’s group also showed that BDNF, an attractive PS-inducing cue, resulted in decreased Tsc2 activity [27••], suggesting the possibility that mTORC1 signaling may be inversely regulated to mediate some attractive versus repulsive responses. Because pathfinding axons receive numerous guidance cues simultaneously in vivo, it is conceptually appealing to speculate that multiple gradients of PS-inducing and PS-repressing cues exert concerted actions on the GC, which then integrates these signals to fine-tune local PS.

Figure 1.

Figure 1

Mechanisms regulating overall translational activity. (a) Under attractive conditions, Netrin binding causes release of ribosome subunits from DCC receptors, allowing assembly of polysomes and localised translation necessary for an attractive GC response. (b) In response to repulsive EphrinA, EphA receptor activation increases Tsc2 activity via decreased phosphorylation at the ERK1/2 site, inhibiting mTORC1 signaling and downstream PS to mediate a repulsive collapse response. Diagram of repulsive turning is speculatively based on demonstrated collapse response.

A more direct link between guidance cue receptors and PS machinery was revealed by Flanagan’s group [29••]. They showed that DCC, a Netrin receptor, co-localises with ribosomes at the EM level in axons and GCs of spinal commissural neurons, and provided evidence that DCC and ribosomes form biochemical complexes when co-expressed in cell lines. Intriguingly, DCC appears to interact with translationally inactive PS machinery, as DCC co-purified with ribosomal subunits and monosomes, but not with polysomes. This interaction was negatively regulated by the binding of Netrin-1 to DCC, suggesting an interesting mechanism in which Netrin-1 induces local mRNA translation by releasing ribosomes from DCC (Figure 1a). This study provides not only a novel mechanism for direct coupling of an extracellular cue to PS machinery, but also uncovers a potential mechanism for spatially restricting PS to ‘microdomains’ within the GC at the site of cue binding. As these experiments were done under conditions in which Netrin-1 is an attractive cue, it will be interesting to know whether DCC-ribosome coupling differs under conditions when Netrin-1 is repulsive. It will be important in future to determine whether other guidance cue receptors interact directly with translational machinery, and specifically whether PS-inhibiting cues such as EphrinA [27••] increase receptor-ribosome association to sequester PS machinery away from mRNAs.

Mechanisms regulating mRNA-specific translation

Different PS-inducing cues regulate translation of distinct sets of mRNAs. For example, attractive cues such as Netrin-1 and BDNF induce β-Actin synthesis [6,7,30], whereas repulsive cues such as Slit2b and Sema3A induce local synthesis of actin depolymerising molecules such as Cofilin and RhoA [5,11]. All of these guidance cues, however, increase general translational activity in the GC. How, then, is mRNA-specific translation achieved? One way would be to control the activity of RNA-binding proteins (RBPs), which regulate a specific subset of mRNAs. Fragile X mental retardation protein (FMRP) is an RBP whose role in dendritic mRNA transport and translation in the context of long-term synaptic plasticity is well characterised [31]. Evidence that FMRP also localises to axons and GCs suggests that it might play a role on the presynaptic side as well [3235]. Indeed, Li and colleagues report that hippocampal neurons cultured from fmr1 knockout mice have defects in Sema3A-induced axonal PS and GC collapse response [36]. They propose that local translation of map1b mRNA in axons and GCs via FMRP may mediate Sema3A-induced GC collapse.

The best known example of an RBP that regulates mRNA translation in axons is zipcode-binding protein (ZBP), which controls the transport, stability, and translation of β-actin mRNA by directly binding to a cis-element in the 3′-UTR. The zipcode, a 54-nt segment in the 3′-UTR, is necessary and sufficient for local translation of β-actin mRNA and GC turning in response to guidance cues such as Netrin-1 and BDNF [6,7]. Interestingly, the core sequence of the zipcode that directly participates in ZBP binding is present in other mRNAs, such as Arp2/3, which are also found enriched in the axon and are functionally related to β-Actin [37]. Bassell’s group recently uncovered a mechanism by which the attractive guidance cue, BDNF, regulates mRNA-specific translation by altering the function of ZBP1 [38] (Figure 2b). They showed that BDNF activates a cascade of signaling events leading to Src-mediated phosphorylation of ZBP1 at Y396. When this phosphorylation was blocked by overexpression of a nonphosphorylatable version of ZBP1 (Y396F), both BDNF-induced β-actin mRNA translation and GC turning responses were attenuated. Therefore, it could be speculated that different guidance cues activate a distinct set of RBPs, which then bind a cohort of functionally related mRNAs to co-ordinately regulate their translation. RBP-mRNA binding may induce a conformational change conducive to the translation of a given mRNA, as was recently shown to be the case for ZBP1-β-actin mRNA interaction [39] (Figure 2b).

Figure 2.

Figure 2

Mechanisms regulating mRNA-specific translation. (a) In the nucleus, diverse species of mRNAs encoding the same protein can be generated by using different transcription initiation/termination sites (1) [21••] or by alternative splicing (2). RBPs (green oval) regulate axonal transport of their target mRNAs by binding to specific cis-elements (green rectangle). (b) In the GC, cue-RBP-mRNA specificity regulates mRNA-specific translation. BDNF induces Src-mediated phosphorylation of ZBP1 [38] activating translation of β-actin mRNA, at least in part by RNA looping [39], and GC turning. NT3 induces CamKII-mediated phosphorylation of CPEB1, which activates translation of β-catenin mRNA by cytoplasmic polyadenylation [46], and axonal elongation and branching [45,46].

Another recent example of mRNA-specific regulation by RBPs in axons comes from Okano and colleagues, who reported a novel function of the RBP Musashi1 (Msi1) to control the translation of robo3 mRNA [40]. They observed that precerebellar inferior olivary neurons in msi1 knockout mice showed a midline-crossing defect similar to robo3 knockout mice [41]. Furthermore, they reported that Msi1 positively regulates Robo3 expression under normal circumstances by binding to and increasing the translation of robo3 mRNA. Interestingly, the cis-element responsible for this regulation resides in the protein coding sequence rather than in the predicted Msi1-binding motif in the 3′-UTR. Msi1 binding to this motif is likely to displace unknown translational repressors because the RNA-binding domain of Msi1 alone functions as a weak activator rather than a dominant negative. Previously, Msi1 was shown to inhibit translation of other target mRNAs such as m-numb [42], providing an example of an RBP that can both enhance and repress mRNA translation depending on the cis-element. This could represent a particularly efficient mechanism for a single RBP to control diverse groups of mRNAs, and it will be interesting to determine whether mRNAs encoding proteins with antagonistic functions could be inversely regulated by a single RBP in this manner.

A better understanding of axonal mRNA translation awaits molecular identification of additional RNA regulatory elements. Novel cis-elements have been identified in recent years. Riccio and colleagues showed that impa1 mRNA is transported into axons in response to NGF stimulation and that this axonal targeting is mediated by a newly identified 150-nt sequence in the 3′-UTR [21••]. Like the zipcode, this sequence is necessary and sufficient for axonal impa1 mRNA transport and axonal survival, although the responsible RBP is not known. Interestingly, different species of impa1 mRNA are generated from alternative transcriptional initiation and termination, generating NGF-responsive and NGF-nonresponsive species of mRNAs (Figure 2a). Considering most mRNAs are produced with diverse UTRs by differential transcriptional and/or posttranscriptional regulation, regulating UTRs would be an efficient way to control axonal mRNA translation without altering protein structures [43]. It is also plausible that similar mechanisms are used to control the responsiveness of mRNAs to microRNA regulation as microRNA-mediated translational inhibition and disinhibition continue to be identified as common mechanisms to control mRNA-specific translation in neuronal processes [44].

Another way an RBP can control mRNA translation is by regulating the length of its poly(A) tail. Cytoplasmic polyadenylation element binding proteins (CPEBs) control translation through this mechanism by directly binding to the CPE present in 3′-UTRs. Two recent papers provide evidence that this mechanism is indeed used to regulate mRNA translation in the axon. The first paper showed that Sema3A-induced local PS involves CPEB function, and that the translation of CPEB-regulated mRNAs is required for Sema3A-induced GC collapse in Xenopus retinal axons [45]. Moreover, inhibiting the function of multiple CPEBs (e.g. CPEB1–4) by over-expressing a dominant negative mutant (i.e. RNA-binding domain alone) of CPEB1 (CPEB1-RBD) disrupted axonal growth in vivo. This is likely mediated by CPEBs other than CPEB1 (e.g. CPEB2–4), because knocking down CPEB1 itself did not interfere with axonal growth. Evidence for how CPEB function is controlled by guidance cues comes from Wells and colleagues [46], who showed that NT3 treatment of cultured rat hippocampal neurons activates local translation of β-catenin mRNA in the GC. As was shown in Xenopus, CPEB1-RBD was used to interfere with CPE-mediated mRNA regulation, and its overexpression disrupted NT3-induced β-catenin synthesis in the GC as well as axonal outgrowth (and branching). Furthermore, NT3 induces a rise in intracellular Ca2+ by the activation of IP3 receptors, which then activates CamKII to phosphorylate and activate CPEB1, providing a mechanistic link between cue binding and mRNA translation in the GC (Figure 2b).

As next generation sequencing is becoming the most powerful tool to analyse mRNA diversity [47], we expect more complete information on the UTRs of axonally localised mRNAs will emerge in the near future, helping us to identify more axon-resident RBPs and leading us to a better understanding of how those RBPs regulate mRNA-specific translation in axons and GCs. Remarkably, Eberwine’s group recently reported that the targeting of certain dendritically localised mRNAs is dependent on a sequence within retained introns, rather than the 3′-UTR [48], indicating that unbiased cataloguing of axonally localised mRNAs may be needed to uncover regulatory mechanisms that include cytoplasmic splicing (Figure 2b).

Future prospects

Functional studies have so far concentrated on only a handful of axonal mRNAs yet thousands of axonal mRNAs have now been identified. This presents a new challenge commonly encountered in the post-genomic era: how to determine the most functionally relevant candidates? Moreover, advances in next generation RNA sequencing (RNA-seq) technology will likely add even more to this ever-growing list of axonal mRNAs. Careful characterization of dynamic spatiotemporal changes in repertoires (e.g. young versus old; axon versus GC) will be one way to gain insight into the functional aspects of axonal PS. On the other hand, the growing lists compiled from the work of different groups will allow for systematic bioinformatics analysis and identification of common elements from these mRNAs that may help uncover novel RBPs and regulatory mechanisms. Sensitive RNA-Seq technologies may eventually allow profiling of single axons and GCs enabling the question to be addressed of whether mRNA diversity is equally represented across an axonal population, or whether it reflects diversity between individual axons of a given subtype.

In terms of GC signaling, it will be important to test whether guidance receptors other than DCC [29••], such as Neuropilin and Robo, are also coupled to the translation machinery in order to determine whether this represents a general mechanism of stimulus-induced translation regulation. Future studies should explore the interplay between factors that activate (e.g. Sema3A) versus those that repress (e.g. EphrinA) local PS to build a better understanding of how the GC integrates guidance signals at the level of translation. Finally, it will prove critical in the next few years to determine what role axonal PS plays in vivo. Developing new strategies to inhibit translation of selected mRNAs in axons without affecting the cell bodies in vivo will be an important challenge that may be met with photo-inducible technologies [49].

References and Recommended Reading

Papers of particular interest published within the period of review have been highlighted as:

• of special interest

•• of outstanding interest

  • 1.Bassell GJ, Zhang H, Byrd AL, Femino AM, Singer RH, Taneja KL, Lifshitz LM, Herman IM, Kosik KS. Sorting of beta-actin mRNA and protein to neurites and growth cones in culture. J Neurosci. 1998;18:251–265. doi: 10.1523/JNEUROSCI.18-01-00251.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Ming GL, Wong ST, Henley J, Yuan XB, Song HJ, Spitzer NC, Poo MM. Adaptation in the chemotactic guidance of nerve growth cones. Nature. 2002;417:411–418. doi: 10.1038/nature745. [DOI] [PubMed] [Google Scholar]
  • 3.Brittis PA, Lu Q, Flanagan JG. Axonal protein synthesis provides a mechanism for localized regulation at an intermediate target. Cell. 2002;110:223–235. doi: 10.1016/s0092-8674(02)00813-9. [DOI] [PubMed] [Google Scholar]
  • 4.Campbell DS, Holt CE. Chemotropic responses of retinal growth cones mediated by rapid local protein synthesis and degradation. Neuron. 2001;32:1013–1026. doi: 10.1016/s0896-6273(01)00551-7. [DOI] [PubMed] [Google Scholar]
  • 5.Wu KY, Hengst U, Cox LJ, Macosko EZ, Jeromin A, Urquhart ER, Jaffrey SR. Local translation of RhoA regulates growth cone collapse. Nature. 2005;436:1020–1024. doi: 10.1038/nature03885. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Yao J, Sasaki Y, Wen Z, Bassell GJ, Zheng JQ. An essential role for beta-actin mRNA localization and translation in Ca2+-dependent growth cone guidance. Nat Neurosci. 2006;9:1265–1273. doi: 10.1038/nn1773. [DOI] [PubMed] [Google Scholar]
  • 7.Leung KM, van Horck FP, Lin AC, Allison R, Standart N, Holt CE. Asymmetrical beta-actin mRNA translation in growth cones mediates attractive turning to netrin-1. Nat Neurosci. 2006;9:1247–1256. doi: 10.1038/nn1775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Piper M, Holt C. RNA translation in axons. Annu Rev Cell Dev Biol. 2004;20:505–523. doi: 10.1146/annurev.cellbio.20.010403.111746. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Lin AC, Holt CE. Local translation and directional steering in axons. EMBO J. 2007;26:3729–3736. doi: 10.1038/sj.emboj.7601808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Hengst U, Jaffrey SR. Function and translational regulation of mRNA in developing axons. Semin Cell Dev Biol. 2007;18:209–215. doi: 10.1016/j.semcdb.2007.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Piper M, Anderson R, Dwivedy A, Weinl C, van Horck F, Leung KM, Cogill E, Holt C. Signaling mechanisms underlying Slit2-induced collapse of Xenopus retinal growth cones. Neuron. 2006;49:215–228. doi: 10.1016/j.neuron.2005.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Verma P, Chierzi S, Codd AM, Campbell DS, Meyer RL, Holt CE, Fawcett JW. Axonal protein synthesis and degradation are necessary for efficient growth cone regeneration. J Neurosci. 2005;25:331–342. doi: 10.1523/JNEUROSCI.3073-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Roche FK, Marsick BM, Letourneau PC. Protein synthesis in distal axons is not required for growth cone responses to guidance cues. J Neurosci. 2009;29:638–652. doi: 10.1523/JNEUROSCI.3845-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Van Minnen J, Bergman JJ, Van Kesteren ER, Smit AB, Geraerts WP, Lukowiak K, Hasan SU, Syed NI. De novo protein synthesis in isolated axons of identified neurons. Neuroscience. 1997;80:1–7. doi: 10.1016/s0306-4522(97)00137-1. [DOI] [PubMed] [Google Scholar]
  • 15.Denis-Donini S, Branduardi P, Campiglio S, Carnevali MD. Localization of calcitonin gene-related peptide mRNA in developing olfactory axons. Cell Tissue Res. 1998;294:81–91. doi: 10.1007/s004410051158. [DOI] [PubMed] [Google Scholar]
  • 16.Bi J, Tsai NP, Lin YP, Loh HH, Wei LN. Axonal mRNA transport and localized translational regulation of kappa-opioid receptor in primary neurons of dorsal root ganglia. Proc Natl Acad Sci USA. 2006;103:19919–19924. doi: 10.1073/pnas.0607394104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17 •.Merianda TT, Lin AC, Lam JS, Vuppalanchi D, Willis DE, Karin N, Holt CE, Twiss JL. A functional equivalent of endoplasmic reticulum and Golgi in axons for secretion of locally synthesized proteins. Mol Cell Neurosci. 2009;40:128–142. doi: 10.1016/j.mcn.2008.09.008. [This paper identifies ER and Golgi components in growing axons. The authors show that posttranslational trafficking occurs in axons and that inhibiting Golgi function attenuates translation-dependent axonal growth responses.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18 ••.Zivraj KH, Tung YC, Piper M, Gumy L, Fawcett JW, Yeo GS, Holt CE. Subcellular profiling reveals distinct and developmentally regulated repertoire of growth cone mRNAs. J Neurosci. 2010;30:15464–15478. doi: 10.1523/JNEUROSCI.1800-10.2010. [This study, for the first time, identifies mRNAs localised to the growth cone compared to the axon shaft. The authors also identify developmental stage-specific changes in mRNA repertoires in the growth cone.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19 ••.Gumy LF, Yeo GS, Tung YC, Zivraj KH, Willis D, Coppola G, Lam BY, Twiss JL, Holt CE, Fawcett JW. Transcriptome analysis of embryonic and adult sensory axons reveals changes in mRNA repertoire localization. RNA. 2011;17:85–98. doi: 10.1261/rna.2386111. [This study provides the first complete catalog of embryonic and adult sensory axonal mRNAs, and shows that axonal mRNA repertoires change significantly with age.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20 ••.Taylor AM, Berchtold NC, Perreau VM, Tu CH, Li Jeon N, Cotman CW. Axonal mRNA in uninjured and regenerating cortical mammalian axons. J Neurosci. 2009;29:4697–4707. doi: 10.1523/JNEUROSCI.6130-08.2009. [This study identifies for the first time >300 mRNAs localised to naïve mature mammalian CNS axons, and adds supporting evidence to early reports of protein synthesis in adult axons [23–26].] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21 ••.Andreassi C, Zimmermann C, Mitter R, Fusco S, Devita S, Saiardi A, Riccio A. An NGF-responsive element targets myoinositol monophosphatase-1 mRNA to sympathetic neuron axons. Nat Neurosci. 2010;13:291–301. doi: 10.1038/nn.2486. [This study identifies >11,000 mRNAs localised to sympathetic neuronal axons and shows that extracellular stimuli such as NGF promote axonal mRNA transport through specific cis-elements in 3′-UTRs. The authors also report impa1 mRNA as the most abundant axonal mRNA whose local translation plays a critical role in axonal survival.] [DOI] [PubMed] [Google Scholar]
  • 22.Cox LJ, Hengst U, Gurskaya NG, Lukyanov KA, Jaffrey SR. Intra-axonal translation and retrograde trafficking of CREB promotes neuronal survival. Nat Cell Biol. 2008;10:149–159. doi: 10.1038/ncb1677. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Koenig E. Synthetic mechanisms in the axon. IV. In vitro incorporation of [3H]precursors into axonal protein and RNA. J Neurochem. 1967;14:437–446. doi: 10.1111/j.1471-4159.1967.tb09542.x. [DOI] [PubMed] [Google Scholar]
  • 24.Twiss JL, Smith DS, Chang B, Shooter EM. Translational control of ribosomal protein L4 mRNA is required for rapid neurite regeneration. Neurobiol Dis. 2000;7:416–428. doi: 10.1006/nbdi.2000.0293. [DOI] [PubMed] [Google Scholar]
  • 25.Zheng JQ, Kelly TK, Chang B, Ryazantsev S, Rajasekaran AK, Martin KC, Twiss JL. A functional role for intra-axonal protein synthesis during axonal regeneration from adult sensory neurons. J Neurosci. 2001;21:9291–9303. doi: 10.1523/JNEUROSCI.21-23-09291.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Giuditta A, Metafora S, Felsani A, Del Rio A. Factors for protein synthesis in the axoplasm of squid giant axons. J Neurochem. 1977;28:1393–1395. doi: 10.1111/j.1471-4159.1977.tb12339.x. [DOI] [PubMed] [Google Scholar]
  • 27 ••.Nie D, Di Nardo A, Han JM, Baharanyi H, Kramvis I, Huynh T, Dabora S, Codeluppi S, Pandolfi PP, Pasquale EB, et al. Tsc2-Rheb signaling regulates EphA-mediated axon guidance. Nat Neurosci. 2010;13:163–172. doi: 10.1038/nn.2477. [This paper shows that EphrinA inhibits the mTORC1 pathway by activating Tsc2. It suggests an interesting possibility that cues previously thought to be protein synthesis-independent may indeed repress translation.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Mann F, Miranda E, Weinl C, Harmer E, Holt CE. B-type Eph receptors and ephrins induce growth cone collapse through distinct intracellular pathways. J Neurobiol. 2003;57:323–336. doi: 10.1002/neu.10303. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29 ••.Tcherkezian J, Brittis PA, Thomas F, Roux PP, Flanagan JG. Transmembrane receptor DCC associates with protein synthesis machinery and regulates translation. Cell. 2010;141:632–644. doi: 10.1016/j.cell.2010.04.008. [This paper shows that DCC, a Netrin-1 receptor, directly associates with ribosomes to sequester them. The authors propose a novel mechanism to regulate local translational activity in which Netrin-1 induces local mRNA translation in the vicinity of DCC by releasing ribosomes.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Zhang HL, Eom T, Oleynikov Y, Shenoy SM, Liebelt DA, Dictenberg JB, Singer RH, Bassell GJ. Neurotrophin-induced transport of a beta-actin mRNP complex increases beta-actin levels and stimulates growth cone motility. Neuron. 2001;31:261–275. doi: 10.1016/s0896-6273(01)00357-9. [DOI] [PubMed] [Google Scholar]
  • 31.Bassell GJ, Warren ST. Fragile X syndrome: loss of local mRNA regulation alters synaptic development and function. Neuron. 2008;60:201–214. doi: 10.1016/j.neuron.2008.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Murashov AK, Chintalgattu V, Islamov RR, Lever TE, Pak ES, Sierpinski PL, Katwa LC, Van Scott MR. RNAi pathway is functional in peripheral nerve axons. FASEB J. 2007;21:656–670. doi: 10.1096/fj.06-6155com. [DOI] [PubMed] [Google Scholar]
  • 33.Hengst U, Cox LJ, Macosko EZ, Jaffrey SR. Functional and selective RNA interference in developing axons and growth cones. J Neurosci. 2006;26:5727–5732. doi: 10.1523/JNEUROSCI.5229-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Price TJ, Flores CM, Cervero F, Hargreaves KM. The RNA binding and transport proteins staufen and fragile X mental retardation protein are expressed by rat primary afferent neurons and localize to peripheral and central axons. Neuroscience. 2006;141:2107–2116. doi: 10.1016/j.neuroscience.2006.05.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Antar LN, Li C, Zhang H, Carroll RC, Bassell GJ. Local functions for FMRP in axon growth cone motility and activity-dependent regulation of filopodia and spine synapses. Mol Cell Neurosci. 2006;32:37–48. doi: 10.1016/j.mcn.2006.02.001. [DOI] [PubMed] [Google Scholar]
  • 36 •.Li C, Bassell GJ, Sasaki Y. Fragile X mental retardation protein is involved in protein synthesis-dependent collapse of growth cones induced by Semaphorin-3A. Front Neural Circuits. 2009;3:11. doi: 10.3389/neuro.04.011.2009. [This study identifies a presynaptic role of FMRP, a regulator of dendritic mRNA translation. The authors show that Sema3A-induced growth cone collapse requires FMRP function, suggesting that FMRP may play a critical role in axon guidance during development.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Condeelis J, Singer RH. How and why does beta-actin mRNA target? Biol Cell. 2005;97:97–110. doi: 10.1042/BC20040063. [DOI] [PubMed] [Google Scholar]
  • 38 •.Sasaki Y, Welshhans K, Wen Z, Yao J, Xu M, Goshima Y, Zheng JQ, Bassell GJ. Phosphorylation of zipcode binding protein 1 is required for brain-derived neurotrophic factor signaling of local beta-actin synthesis and growth cone turning. J Neurosci. 2010;30:9349–9358. doi: 10.1523/JNEUROSCI.0499-10.2010. [This study provides a missing link in the existing model of ZBP1-mediated β-actin mRNA translation. The authors show that BNDF induces Src-mediated phosphorylation of ZBP1, which is a crucial step in β-actin mRNA local translation and growth cone turning.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Chao JA, Patskovsky Y, Patel V, Levy M, Almo SC, Singer RH. ZBP1 recognition of beta-actin zipcode induces RNA looping. Genes Dev. 2010;24:148–158. doi: 10.1101/gad.1862910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40 •.Kuwako K, Kakumoto K, Imai T, Igarashi M, Hamakubo T, Sakakibara S, Tessier-Lavigne M, Okano HJ, Okano H. Neural RNA-binding protein Musashi1 controls midline crossing of precerebellar neurons through posttranscriptional regulation of Robo3/Rig-1 expression. Neuron. 2010;67:407–421. doi: 10.1016/j.neuron.2010.07.005. [This paper shows that Musashi1, an RNA binding protein known to inhibit m-numb mRNA translation, increases translation of robo-3 mRNA by directly binding to a non-canonical cis-element within the coding sequence. It suggests that the same RBP can increase or decrease mRNA translation depending on the cis-element.] [DOI] [PubMed] [Google Scholar]
  • 41.Marillat V, Sabatier C, Failli V, Matsunaga E, Sotelo C, Tessier-Lavigne M, Chedotal A. The slit receptor Rig-1/Robo3 controls midline crossing by hindbrain precerebellar neurons and axons. Neuron. 2004;43:69–79. doi: 10.1016/j.neuron.2004.06.018. [DOI] [PubMed] [Google Scholar]
  • 42.Okabe M, Imai T, Kurusu M, Hiromi Y, Okano H. Translational repression determines a neuronal potential in Drosophila asymmetric cell division. Nature. 2001;411:94–98. doi: 10.1038/35075094. [DOI] [PubMed] [Google Scholar]
  • 43.Holt CE, Bullock SL. Subcellular mRNA localization in animal cells and why it matters. Science. 2009;326:1212–1216. doi: 10.1126/science.1176488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Abdelmohsen K, Hutchison ER, Lee EK, Kuwano Y, Kim MM, Masuda K, Srikantan S, Subaran SS, Marasa BS, Mattson MP, et al. miR-375 inhibits differentiation of neurites by lowering HuD levels. Mol Cell Biol. 2010;30:4197–4210. doi: 10.1128/MCB.00316-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45 •.Lin AC, Tan CL, Lin CL, Strochlic L, Huang YS, Richter JD, Holt CE. Cytoplasmic polyadenylation and cytoplasmic polyadenylation element-dependent mRNA regulation are involved in Xenopus retinal axon development. Neural Dev. 2009;4:8. doi: 10.1186/1749-8104-4-8. [This paper identifies a critical role of cytoplasmic polyadenylation in axon elongation in vivo] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46 •.Kundel M, Jones KJ, Shin CY, Wells DG. Cytoplasmic polyadenylation element-binding protein regulates neurotrophin-3-dependent beta-catenin mRNA translation in developing hippocampal neurons. J Neurosci. 2009;29:13630–13639. doi: 10.1523/JNEUROSCI.2910-08.2009. [This paper shows that cytoplasmic polyadenylation is required for NT-3-induced axonal translation of β-catenin mRNA and axon elongation in cultured murine hippocampal neurons. The authors also show that phosphorylation of CPEB1 by CamKII is a critical step in this process.] [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Licatalosi DD, Mele A, Fak JJ, Ule J, Kayikci M, Chi SW, Clark TA, Schweitzer AC, Blume JE, Wang X, et al. HITS-CLIP yields genome-wide insights into brain alternative RNA processing. Nature. 2008;456:464–469. doi: 10.1038/nature07488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Buckley PT, Lee MT, Sul JY, Miyashiro KY, Bell TJ, Fisher SA, Kim J, Eberwine J. Cytoplasmic intron sequence-retaining transcripts can be dendritically targeted via ID element retrotransposons. Neuron. 2011;69:877–884. doi: 10.1016/j.neuron.2011.02.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Tomasini AJ, Schuler AD, Zebala JA, Mayer AN. PhotoMorphs: a novel light-activated reagent for controlling gene expression in zebrafish. Genesis. 2009;47:736–743. doi: 10.1002/dvg.20554. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES