Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Neuron. Author manuscript; available in PMC 2016 May 20.
Published in final edited form as:
PMCID: PMC4551425
NIHMSID: NIHMS677823
PMID: 25937169

Injury-induced decline of intrinsic regenerative ability revealed by quantitative proteomics

Associated Data

Supplementary Materials

Abstract

Neurons differ in their responses to injury but the underlying mechanisms remain poorly understood. Using quantitative proteomics, we characterized the injury-triggered response from purified intact and axotomized retinal ganglion cells (RGCs). Subsequent informatics analyses revealed a network of injury-response signaling hubs. In addition to confirm known players, such as mTOR, this also identified new candidates, such as c-myc, NFkB and Huntingtin. Similar to mTOR, c-myc has been implicated as key regulators of anabolic metabolism and is down-regulated by axotomy. Forced expression of c-myc in RGCs, either before or after injury, promotes dramatic RGCs neuronal survival and axon regeneration after optic nerve injury. Finally, in contrast to RGCs, neither c-myc nor mTOR was down-regulated in injured peripheral sensory neurons. Our studies suggest that c-myc and other injury responsive pathways are critical to the intrinsic regenerative mechanisms and might represent a novel target for developing neural repair strategies in adults.

INTRODUCTION

In the adult mammalian central nervous system (CNS), axotomy often triggers neuronal death, and spontaneous axon regeneration rarely occurs. An implicated mechanism is the diminished intrinsic regenerative ability of mature CNS neurons (Fawcett, 2006; Goldberg et al., 2002b; Moore et al., 2009; Park et al., 2008). Based on the differences in axon growth between immature and mature neurons, it has been proposed that the neuronal intrinsic growth ability is lost over the course of development. Therefore, much effort has been made to seek molecular pathways that are differentially expressed during development and in the adult, resulting in the identification of several axon regeneration regulators. For example, in many types of neurons, cAMP levels appear to be higher in the immature neurons, but decline in the mature neurons (Cai et al., 2001; Filbin, 2003). By analyzing the axon growth of retinal ganglion cells (RGCs) from different developmental stages, Goldberg et al showed a development-dependent decline of axon growth rate, with a dramatic decrease in postnatal RGCs (Goldberg et al., 2002a). Further, recent studies implicated transcription factors, the Krüppel-like family of transcription factors (KLFs) as critical regulators of development-dependent axon growth ability in RGCs (Moore et al., 2009). Interestingly, while KLF7 is down-regulated, other members such as KLF4 are up-regulated during development. Importantly, manipulations of these factors could promote the regrowth of injured optic nerve axons and corticospinal tract axons in the adult (Blackmore et al., 2012).

However, ample evidence indicates that even in the adult CNS, many uninjured neurons possess considerable capacity for structural plasticity and collateral sprouting (Holtmaat and Svoboda, 2009; Raineteau and Schwab, 2001). For example, upon incomplete spinal cord injury in the adult, spared axons have been repetitively shown to elaborate spontaneous sprouting responses (Bareyre et al., 2004; Rosenzweig et al., 2010). Furthermore, such sprouting responses could be further enhanced by rehabilitation training (Harel et al., 2013; van den Brand et al., 2012). This is in contrast to no or limited regrowth from injured axons in the adult CNS (Bradke et al., 2012; Goldberg et al., 2002b; Moore et al., 2009; Park et al., 2008; Rossi et al., 2007). Therefore, it is possible that in addition to development-dependent processes, axonal injury-triggered stress responses might contribute to the impaired intrinsic regenerative ability of mature neurons.

While many previous studies documented gene expression changes in different types of injured neurons (Costigan et al., 2002; Michaelevski et al., 2010; Saul et al., 2010; Tanabe et al., 2003; Fischer et al., 2004), pinpointing key molecular pathways that orchestrate neuronal survival and axon regeneration remains a major challenge. One possible reason is that axotomy may impinge on both gene transcription as well as protein translation and degradation. Thus analyzing the transcriptome may not reflect the full scope of injury-induced changes in neurons. As protein abundance is the final readout of gene expression, translation and degradation, analyzing injury-induced changes on protein abundance may provide a direct assessment of neuronal injury responses. Although proteomic methods have been used extensively for the analyses of cultured cells and whole tissues (Chen and Springer, 2009; Kim et al., 2014; Magharious et al., 2011; Wilhelm et al., 2014), very few studies have utilized procedures to analyze a specific type of primary neurons purified directly from live tissue. In this study, we performed a quantitative proteomics experiment to profile changes in protein abundance in RGCs induced by optic nerve injury relative to an uninjured control. Subsequent informatics analyses identified a set of injury-response pathways that have an impact on the intrinsic growth ability and regenerative responses after injury.

RESULTS

Quantitative proteomic analysis of intact and injured RGCs

As an optic nerve injury transects the axons from all RGCs in the retina, it is an ideal model for isolating homogenously injured neurons and analyzing their injury responses. This is further facilitated by the availability of the transgenic mouse line YFP-17 (Sun et al., 2011), in which the yellow fluorescent protein (YFP) is predominantly expressed in the RGCs and in a few amacrine cells (Figure S1A). Using these YFP-17 mice, we performed an optic nerve injury on one eye of these animals and left the other eye uninjured as control. At 3 days post-injury before extensive apoptotic death has occurred, axotomized and intact retinas were dissected out and dissociated cells were subjected to fluorescence activated cell sorting (FACS) purification.

FACS gating parameters were optimized to select cells based on size, viability (DAPI negative cells), and YFP+ signal (Figure 1A) and resulted in the isolation of 99% viable YFP+ cells (Figure S1B). In this experiment, the isolation of a subset of neurons i.e. predominantly RGCs results in the specific analysis of the injury response to the optic nerve. We collected 3 biological replicate samples, each of which consists of a pool of injured or control RGCs (~15 retinas per pooled replicate corresponding to approximately 1.5 × 106 cells). Samples were lyzed and total protein concentrations were measured for the lysates using BCA. Equal amounts of protein from individual samples were digested using trypsin and the resultant peptides were labeled with one of the six tandem mass tags (TMT). These samples were then mixed and fractionated using isoelectric focusing. Less than 10 ug from each fraction was subjected to LC-MS/MS analysis using the standard methods and Pulsed Q Dissociation to quantify the TMT reporter ions on an Orbitrap Classic instrument (Griffin et al. 2007). The resulting spectra were searched with MASCOT (v2.1) against an annotated mouse protein database (IPI 3.68) to identify spectral matches to annotated peptides. These matched spectra (peptide spectral matches, PSMs) corresponded to 7514 unique peptides and were grouped into 1409 proteins using a 1% false-discovery rate (FDR) cut-off and a minimum of two peptides per protein (Figure 1C table S1). Proteins were identified from a range of subcellular compartments, such as the cytoplasm (60%), nucleus (18%), and plasma membrane (14%), and functional classes, such as kinases/phosphatases, transcriptional regulators, or ion channels. This result indicates that a broad functional depth of the proteome was measured (Figure S2, and Table S2).

An external file that holds a picture, illustration, etc.
Object name is nihms677823f1.jpg
Proteomics analysis of the RGCs before and 3 days after optic nerve crush

(A) Representative FACS plots illustrating the steps of RGC purification. Dissociated retinal cells were gated based on size and surface characteristics (Forward Scatter FSC-A, x axis, Side Scatter SSC-A, y-axis; first graph from left). DAPI positive neurons, staining dead cells depicted in blue, are excluded by sorting (second graph from left). Retinal cells without YFP were used as controls to set up the threshold for YFP+ cells (third graph). The last graph shows the population of YFP+ RGCs that were selected. (B) A schematic of the multiplexed quantitative proteomics workflow. Retinas from uninjured (WT-wild type) and 3 days post optic nerve crush (WTc) YFP17 mice were collected (i). After dissociation, YFP+ neurons were purified by FACS (ii). 3 independent biological replicate samples were collected for each condition (each sample from at least 10 animals), and subjected to cell lysis (iii) and tryptic digestion (iv). Samples were then labeled with specific mass-coded TMT tags (illustrated with different colors), prior to mixture (v), OFFgel fractionation (vi) LC-MS/MS and informatics analyses to identify and quantify peptide spectrum matches (PSMs) (vii–x). (C) Table of the numbers of PSMs (peptide spectrum matches), unique peptides and proteins identified from our analysis. (D) The plot showing the quantitative distribution of the 1409 identified proteins according to their log2 expression ratio (WTc/WT).

Protein abundance was analyzed in a pairwise fashion within biological samples of RGCs with injured optic nerves and their control RGCs from uninjured optic nerves from the same animals and plotted as the ratio of expression in the injured sample divided by expression in the control (log2 scale), revealing that the majority of identified proteins showed similar abundance values before and after injury (Figure 1D). Interestingly, 62 proteins were significantly increased or decreased respectively in the injured RGCs compared to the uninjured control (Figure 1D and Table S2).

Functional annotation of injury-altered proteins

Next, we used Ingenuity Pathway Analysis (IPA) (Park et al., 2013) to assess which cellular functions might be related to these altered 62 proteins that are affected in the injury response. As shown in Fig. 2A and Table S3, most down-regulated proteins appear to be associated with cellular functions such as formation of cellular protrusions, organization of cytoskeleton and microtubule dynamics. Since these processes represent critical steps of axon regeneration, these changes might reflect the reduced intrinsic regenerative ability of injured RGCs.

An external file that holds a picture, illustration, etc.
Object name is nihms677823f2.jpg
Functional annotation and network analysis of injury-altered proteins in RGCs

(A) Protein ontology categories enriched in the proteins affected by injury. IPA analysis indicates the up-regulation (in orange), down-regulation (in blue) or non-predictable (in white) function variation after injury for each cluster. The size of each square is proportional to the number of altered proteins in this functional category. (B) Major hubs from the merged network analysis of injury-altered proteins. Factors and/or pathways of interest are added in this spider-web, based on their targets. (C) Table of major regulators ranked based on their numbers of total interactions (nodes) in the interaction web.

Putative key regulators of injury responses revealed by informatics analysis

We reasoned that such protein abundance alterations might reflect the outcomes of injury-response signaling events. Therefore, we subjected these 62 proteins to an IPA-based network analysis (Staab et al., 2007), in order to find major signaling events occurring in axotomized RGCs. This analysis revealed 9 signaling networks (Figure S2C, Table S4,) and relevant signaling hubs (Figure S2C). The top 12 signaling hubs were listed in Figures 2B and 2C.

Strikingly, this list includes several known regulators of neuronal survival and axon regeneration, such as TP53, Ca2+ (Ghosh-Roy et al., 2010, Chierzi et al., 2005; Ghosh-Roy et al., 2010; Sung et al., 2001; Bradke et al., 2012), MAPK (Chen et al., 2011; Hammarlund et al., 2009; Watkins et al., 2013), JAK (Smith et al., 2009; Sun et al., 2011) and the components of mTOR pathway, such as Rictor, Raptor and mTOR (Dupraz et al., 2013; Liu et al., 2010; Near et al., 1992; Park et al., 2008; Toth et al., 2006; Zhu et al., 2010), validating the use of this approach in analyzing neuronal injury responses. Interestingly, this list also includes several other molecules, such as c-myc, Huntingtin (HTT) and NFkB (Figure 2B and 2C), which might represent new candidates for critical regulators of the neuronal injury response.

c-myc expression is down-regulated in RGCs in response to injury

Our further analysis has been focused on c-myc, which was one of the most connective nodes of the regulator network (Figure 2B, 2C and table S4). c-myc was not among the 1409 proteins identified by proteomics analysis (Table S1 and S2), possibly due to its low protein abundance in adult RGCs. However, among 62 injury-altered proteins identified from our proteomics analysis, 10 proteins (highlighted in red in Table S4) have functional interactions with c-myc in the network. Importantly, similar to mTOR (Howell et al., 2013; Laplante and Sabatini, 2012; Shaw and Cantley, 2006), c-myc has been implicated as a master control gene of cellular metabolism (Dang, 2013), positioning it as an excellent candidate to improve RGCs survival and axon regeneration.

To verify the involvement of c-myc in the injury response, we examined the expression level of c-myc before and after injury. Using real time qPCR analysis of mRNA purified from samples subjected to identical treatments as the proteomics experiments (Figure 1), we observed that the expression of c-myc mRNA is decreased by 70% after injury (Figure 3A, B). This injury-induced c-myc down-regulation was also verified by immunohistochemistry with anti-c-myc antibodies on retinal sections from intact and injured mice (Figure 3C). These results together verified the injury-induced c-myc down-regulation in injured RGCs.

An external file that holds a picture, illustration, etc.
Object name is nihms677823f3.jpg
c-myc expression promotes neuronal survival and axon regeneration after optic nerve injury

(A) Timeline of the experimental procedures. (B) c-myc mRNA levels in intact and injured RGCs detected by real time q-PCR. mRNA samples were extracted from purified RGCs isolated from YFP+ retina at 3 days post injury or from intact control retinas. Results are normalized to GAPDH. **: p < 0.01, T-test. Error bars: S.E.M (C) Representative images of retinal sections stained with anti-c-myc (in red), anti-Tuj1 (in green) antibodies, and DAPI (in blue). Scale bar: 20 μm. (D) Scheme of experimental procedures to assess axonal regeneration. (E) Representative confocal images of CTB-labeled optic nerve sections from Rosamer-c-myc mice with intravitreal injection of AAV-PLAP (n = 5) or AAV-Cre (n =7). Red stars indicate the crush site. Scale Bar: 100um. (F) Quantification of the numbers of regenerative axons counted at different distances distal from the lesion. ***: p < 0.001, **: p < 0.01 and *: p < 0.05, T-Test. Error bars: S.E.M (G) Quantification of survived RGCs as detected by anti-Tuj-1 immunostaining in whole mount retinas from intact Rosamer-c-myc mice (n = 5), injured Rosamer-c-myc mice with injection of AAV2-PLAP (n = 5) or AAV2-Cre (n = 5). ***: p < 0.001, ANOVA with Bonferroni post test.

Over-expression of c-myc in RGCs promotes axon regeneration in vivo

We next assessed the effects of c-myc over-expression on neuronal responses by two different approaches. First, as we demonstrated previously, injection of adeno-associated virus serotype 2 (AAV2) vectors into the vitreous body results in highly selective and efficient transduction to RGCs (Park et al., 2008; Sun et al., 2011), thus we injected AAV2 expressing c-myc or a control placenta alkaline phosphatase (PLAP) into the vitreous body of adult mice. These mice were then subjected to a standard optic nerve injury procedure, followed by a Cholera Toxin B subunit (CTB) injection to analyze neuronal survival and axon regeneration (Park et al., 2008). At 2 weeks post-injury, we observed that 49% of RGCs survived in c-myc over-expressing mice, whereas only 25% survived in the mice injected with AAV2-PLAP (Figure S3A, B). Importantly, a significant axonal growth increase was observed in the optic nerves from c-myc over-expressing mice (Figure S3C, D), suggesting that c-myc over-expression could promote both neuronal survival and axon regeneration after injury.

As an independent approach to assess the effect of c-myc over-expression, we used an inducible c-myc transgenic mouse line, ROSAMER mice (Jager et al., 2004). In these mice, a cassette with the gene encoding a tamoxifen-inducible c-mycERT fusion protein (mycERT) down-stream of a lox-stop-lox sequence was inserted in the ubiquitously active ROSA26 locus. Therefore, mycERT transcription is Cre-dependent and the activity of the fusion protein is tamoxifen-inducible. We then tested the effect of a transient activation of c-myc prior to injury. After the intravitreal injection of AAV2-Cre or AAV2-PLAP, and subsequent tamoxifen induction, these mice were subjected to optic nerve injury (Figure 3D). As shown in Figure 3E–G, both robust neuronal survival and axon regeneration were observed in the mice injected with AAV2-Cre, but not with control AAV2-PLAP. The extent of the regenerative response was comparable to that observed in PTEN-deleted mice (Park et al., 2008). These phenotypes are more robust than what was observed after AAV2-c-myc expression (Figure S3), possibly due to a more efficient expression of c-myc in these transgenic mice. We also monitored the mTOR activity in RGCs under these treatments, three days after injury, by immunohistochemistry using antibodies against phosphorylated-ribosome protein S6 (P-S6), an established indicator of the mTOR activity (Guertin and Sabatini, 2007). As shown in Figure S4, RGCs with c-myc over-expression still exhibited significant injury-induced mTOR down-regulation, although the extent of down-regulation is less than that observed for the controls. These data further support the notion that c-myc over-expression is able to promote RGCs axons regeneration after optic nerve injury.

Delayed c-myc over-expression still promotes optic nerve regeneration

To mimic clinically relevant settings, we assessed the effects of expressing c-myc after a delayed post-injury period. We thus administered tamoxifen to ROSAMER mice starting from one day after optic nerve injury (Figure 4A). At two weeks post-injury, we observed that approximately 55% of RGCs survived, in comparison with 25% in controls (Figure 4D and 4E). In addition, significant numbers of regenerating axons were observed in the animals with delayed c-myc expression with many regenerating axons growing beyond the injury site and extending 1.5mm from the crush site (Figures 4B and 4C). As the apoptotic pathways might be activated at these axotomized RGCs (Figure 2), our results suggest that c-myc can rescue these injured neurons from apoptotic death and promote their axon regeneration.

An external file that holds a picture, illustration, etc.
Object name is nihms677823f4.jpg
Delayed c-myc expression promotes RGC survival and axon regeneration after optic nerve injury

(A) Timeline for delayed activation of c-myc after the optic nerve injury. (B) Representative confocal images of CTB-labeled optic nerve sections from control (AAV2-PLAP; n= 5) and Rosamer-c-myc (c-Myc) mice (n=7). Scale bar 100μm. (C) Quantification of regenerative axons counted at different distances distal from the lesion. T-Test. ***: p <0.001 and *: p <0.05. (D) Representative images of whole mount retinas, stained with anti-Tuj1, from intact, injured control (AAV2-PLAP; n= 5) and injured c-myc mice (AAV2-Cre, n=7). Scale bar: 20 um. Error bars: S.E.M (E) Quantification of Tuj1+ RGCs in different groups. AI: After Injury. BI: Before Injury. ANOVA with Bonferroni post test. ***: P <0,001.

c-myc is essential for regenerative phenotype of PTEN deletion

The experiments above implicated injury-induced c-myc down-regulation as an important mechanism reducing the intrinsic regenerative ability of adult RGCs. We then investigated whether c-myc is required for axon regeneration in adult RGCs induced by other manipulations. Since PTEN inhibition was previously shown to activate mTOR and promote neuronal survival and axon regeneration (Liu et al., 2010; Park et al., 2008; de Lima et al., 2012), we thus examined whether a genetic knockout of c-myc might affect the axon regeneration induced by shRNA-mediated PTEN silencing (Zukor et al., 2013). We co-injected AAV2-ShPTEN with AAV2-Cre or AAV2-PLAP into the vitreal body of the c-mycf/f mice and performed optic nerve injury after 4 weeks. As expected, PTEN silencing by ShRNA resulted in significant axon regeneration as analyzed at 2 weeks post-injury (Figure S5A and B). The knockout of c-myc significantly reduced the numbers of regenerating axons (Figure S5A, B), but did not affect neuronal survival (Figure S5C, D) or the mTOR activity (Figure S5E). However, some residual regenerating axons were still observed after c-myc deletion (Figure S5A, B). It remains unclear whether this result reflects the differential involvement of c-myc in RGCs subtypes or the compensatory effects of other myc members.

Synergistic effects of c-myc over-expression and PTEN and SOCS3 deletion

Our systems biology analyses of the pathways involved in the neuronal response to injury strongly suggest that other major signaling hubs work in concert with c-myc (Figure 2B, C). To better understand their functional interactions, we analyzed the combinatorial effects in the optic nerve injury model by manipulating both c-myc and the mTOR pathway, which was also implicated by the network analysis (Figure 2). Thus we performed an intravitreal co-injection of AAV2-Cre together with AAV2-c-myc or AAV2-PLAP, in PTENf/f mice (Figure 5A). Two weeks after injury, we observed significant increase in neuronal survival and axon regeneration in the PTENf/f mice with c-myc expression with around 75% of RGCs alive (Figure 5B–D).

An external file that holds a picture, illustration, etc.
Object name is nihms677823f5.jpg
Synergistic effect of PTEN deletion with c-myc overexpression

(A) Timeline of experimental procedure. (B) Representative confocal images of the optic nerve sections from PTENf/f (infected with AAV2-Cre), Rosamer-myc (infected with AAV2-Cre and treated with Tamoxifen) and PTENf/f mice infected with AAV2-Cre and AAV2-c-myc. Axons are labeled with CTB. Red stars indicate the crush site. Scale bar: 100 μm. (C) Quantification of regenerative axons in the three groups presented in (B). ANOVA with Bonferroni post test. *p<0.05. Error bars: S.E.M. (D) Quantification of RGC survival as measured by Tuj1 staining ANOVA with Bonferroni post test. ***p-value <0,001 and *p<0.05.

As our previous studies showed robust axon regeneration after co-deletion of PTEN and SOCS3 in RGCs (Sun et al., 2011), we further examined the effects of this co-deletion in combination with c-myc over-expression (Figure 6A). Thus we performed an intravitreal co-injection of AAV2-Cre and AAV2-CNTF, together with AAV2-c-myc or AAV2-PLAP, in PTENf/f/SOCS3f/f mice. As shown in Figure 6B, such a combinatorial treatment resulted in further enhanced axon regeneration. At 4 weeks post-injury, about 5 times more regenerating axons reached the proximal end of the chiasm when compared with PTEN/SOCS3 deletion alone (Figures 6B and C and Figure S6). In the mice with PTEN/SOCS3 co-deletion, the majority of regenerating axons ceased growing in the proximal end of the chiasm (Luo et al., 2013; Sun et al., 2011). However, with additional c-myc expression, while some axons grew ectopically into the contralateral optic nerve, many more axons crossed the chiasm and continued to grow in the optic tracts, with some ipsilaterally and more contralaterally (Figure 6B, E), After crossing the chiasm, many regenerating axons derail from the optic tract, similar to that shown by Luo et al., 2013. At 8 weeks post-injury, while most of regenerating axons disappeared in the mice with PTEN/SOCS3 deletion without c-myc expression, many regenerating axons remain in the mice with triple treatment (Figure S7). In these animals, however, the numbers of regenerating axons are reduced to some extent (Figure. S7), possibly due to the elimination of regenerating axons that have not formed proper synaptic connections. These results suggested that co-manipulation of these injury-response pathways could result in a synergistic effect on promoting neuronal survival and axon regeneration.

An external file that holds a picture, illustration, etc.
Object name is nihms677823f6.jpg
c-myc expression enhances optic nerve regeneration induced by co-deletion of PTEN and SOCS3 in RGCs

(A) Timeline of the experimental procedures. (B) Representative confocal images of CTB-labeled optic nerve whole-mounts from PTENf/f/SOCS3f/f mice with intravitreal injection of AAV2-Cre, AAV2-CNTF and AAV2-c-myc (n=6) or AAV2-Cre, AAV2-CNTF and AAV2-PLAP (n=6). Optic nerve crush was performed on the right optic nerve head and regenerative axons were labeled with CTB-Alexa-555. Red stars indicate the crush site. Scale bar: 100 um. (C, D) Quantification of the numbers of regenerating axon at the proximal end of the chiasm (C) and regenerating axons that passed the chiasm, including those crossing the midline (contra) and navigating ipsilaterally (ipsi) (D). (E) Distribution of regenerating axons projecting in the contralateral optic nerve (gray), the contralateral optic tract (white) and ipsilateral optic tract (black) in the groups of PS (PTENf/f/SOCS3f/f with AAV2-Cre, AAV2-CNTF and AAV2-PLAP) and PSM (PTENf/f/SOCS3f/f with AAV2-Cre, AAV2-CNTF and AAV2-c-myc). T-test, ***: p < 0.001 and *: p < 0.05.

Differential injury responses in RGCs and DRG neurons

As the neurons in the PNS and the CNS differ in their regenerative ability, we examined how these injury regulatory pathways are differentially regulated in regeneration-competent DRG neurons. As shown in Figure 7, the c-myc expression level, measured by quantitative qPCR with the mRNA samples from DRGs, was significantly increased at 1 day post-injury, but returned to the basal levels from 3 days post-injury. Consistent with the regulation of c-myc transcripts, the protein abundance of c-myc as measured using anti-c-myc antibodies showed delayed increase at 3 days post-injury (Figure 7B). In addition, we also monitored the mTOR activation levels by immunohistochemistry with anti-phospho-S6 antibodies. As shown in Figure 7C, the phospho-S6 is significantly increased in DRG neurons with a sciatic nerve injury, in contrast to the injury-induced mTOR down-regulation observed in RGCs (Figure 2 and S6, Park et al., 2008). These results revealed a correlation between the activation levels of these pathways and the neuronal regenerative ability, suggesting the critical roles of these and other identified injury responsive molecules in determining the intrinsic regenerative ability in mature neurons.

An external file that holds a picture, illustration, etc.
Object name is nihms677823f7.jpg
c-myc expression and mTOR activation in injured DRG neurons

(A) The quantification of the c-myc mRNA levels measured by qPCR on the mRNA extracted from intact L4-6 DRGs or those at different time points after sciatic nerve crush. Results are normalized against GAPDH and the level of c-myc expression in intact DRGs was used as references. One day after injury there is an increase in c-myc RNA. ANOVA test ***: p<0.001. (B) DRG sections collected before sciatic nerve crush (no injury), 1, 3, or 7 days post crush (DPC). DRG sections are stained with c-myc (green), NeuN (red) and DAPI (blue). c-myc expression increases after sciatic nerve injury. Scale bar: 50 um. (C) DRG sections from different conditions are stained with Tuj1 (red), anti-phospho-S6 (green) and DAPI (blue). The Phospho-S6 immunoreactivity increases after sciatic nerve injury. Scale bar: 50 um.

DISCUSSION

By using a systems biology analysis of protein abundance changes measured by quantitative mass spectrometry followed by bioinformatics analyses, we identified a network of molecular pathways that are altered in injured RGCs. Manipulations of two of these pathways, myc and mTOR, resulted in dramatically increased neuronal survival and axon regeneration. Although the combined effects with other pathways remain to be tested, these results suggest the possibility that manipulation of these injury-response pathways simultaneously might represent a powerful strategy of promoting neuronal survival and axon regeneration.

While it has been proposed that the intrinsic regenerative ability is a critical determinant of axon regeneration, its molecular signatures and its regulatory mechanisms remain poorly understood. Our results provide mechanistic evidence for the hypothesis that injury-triggered stress responses might be important to the loss of intrinsic regenerative ability observed in the adult RGCs. The major injury-altered signaling pathways identified in this study could be strong potential molecular determinants, which might play different roles in regulating the intrinsic capacity to regenerate upon injury. For example, the components of the MAPK pathway were found on the list of the identified injury-response pathways, consistent with ample evidence supporting MAPK activation as evolutionarily conserved injury signals (Chen et al., 2011; Hammarlund et al., 2009; Watkins et al., 2013). Similarly, the current studies point to a critical role for injury-triggered responses in down-regulating the mTOR activity, an implicated player in regulating neuronal survival and axon regeneration after injury (Liu et al., 2010; Park et al., 2008; Sun et al., 2011). In addition, our analysis also reveals a number of novel putative injury response molecules such as c-myc, TGF-b, NFkb, Huntingtin, which might provide new avenues of exploration to understand the molecular mechanisms of intrinsic regenerative ability. Our further analysis on c-myc provides compelling evidence verifying the role of such novel molecular players in neuronal injury responses. Forced expression of c-myc in RGCs, either before or after injury, promotes neuronal survival and axon regeneration, and this effect is further amplified by the combined deletion of PTEN and SOCS3. Importantly, both c-myc and mTOR are differentially regulated in regeneration-competent peripheral sensory neurons, suggesting a possibility that the role of these molecules in affecting the intrinsic regenerative ability might not limited to a single neuronal types.

How c-myc acts in injury responses remains unclear. c-myc is described as being a master transcription factor that regulates genes involved in different aspects of anabolic metabolism such as ribosome biogenesis, lipid synthesis, nucleic acid synthesis, intermediary metabolism, and cell growth and proliferation (Dang, 2013). Our data suggests that the activity of c-myc is a key step towards achieving successful axon regeneration after an injury to over-ride a shift in metabolic state that usually results in the degeneration of axons and eventual apoptosis. This is especially important for the mature neurons in the adult, because in these resting cells the ATP is mainly generated by catabolic processes and is used to maintain energy-costly homoeostatic processes, such as cytoskeletal functions and ion and nutrient transport. However, for an injured neuron to initiate regenerative growth, it has to shift toward anabolic processes, allowing for the de novo synthesis of macromolecules from available nutrients. In fact, mTOR has also been implicated as an important regulator of cell metabolism, because it can alter cellular metabolism to drive the biosynthesis of building blocks and macromolecules essential for cell growth (Guertin and Sabatini, 2007; Xie and Proud, 2013). Therefore, a plausible model is that c-myc alone or together with mTOR might allow injured neurons to achieve such a shift of metabolic states for axonal regrowth. It would be also interesting to test whether other metabolic modulating methods could alter neuronal regenerative ability. On the other hand, recent studies documented an increase in c-myc protein in the promoter region of active genes as opposed to regulating the expression of specific genes, thus amplifying the overall output of the existing gene expression program in tumor cells (Lin et al., 2012; Nie et al., 2012). Thus, an alternative model is that c-myc might act similarly in neurons by boosting the general gene expression, which may explain its dramatic effect on increasing the regeneration observed in combination with the PTEN/SOCS3 deletion. Understanding how these and other pathways act in neuronal injury responses hold promise in identifying new targets for designing new strategies to promote neuronal survival and axon regeneration.

The identification of injury response network by such systems-wide approaches in this study may also provide an example for investigating the pathogenesis of other types of neurological diseases, because how different types of neurons respond to specific stress conditions remains as an unsolved mystery. For example, the abnormal expression of pathological proteins, such as amyloid-b or mutant superoxide dismutase 1 (SOD1), might be critical for the development of Alzhiemer’s diseases or Amyotrophic lateral sclerosis. In these cases, despite the ubiquitous expression of pathological proteins in all cell types, often only subtypes of neurons exhibit pathological alterations and neuronal loss, suggesting that similar to injury responses, neurons may differ in their responses to other types of stress conditions. It therefore will be valuable to analyze the stress response programs in individual neuronal types and to test whether similar or different stress-response programs are activated in other pathological conditions and manipulating such pathways could be beneficial.

EXPERIMENTAL PROCEDURES

Surgical procedure

All experimental procedures were performed in compliance with animal protocols approved by the IACUC at Boston Children’s Hospital. For all surgery procedures, mice were anaesthetized with ketamine and xylazine. Eye ointment containing atropine sulfate was applied to protect the cornea during surgery. Animals received Buprenorphine for 24 hours as a post-operative analgesic.

Mouse lines

C57BL6/J mice are used as wild type controls. PTENf/f mice are maintained in BALB/cAnNTac genetic background (JAX). SOCS3f/f mice are in C56/129 background (JAX). Mycf/f mice are in C57BL6/J background (JAX). Myc Rosamer mice (Jager et al., 2004) are maintained in 129/SV background. We crossed the PTENf/f and Socs3f/f lines obtained from JAX to generate the PTENf/f/Socs3f/f double mutants. The background of this mouse line is mixed. This mouse line has been previously described (Sun et al., 2011).

AAV2 virus injection

C57BL6/J mice (used as wild type controls), PTENf/f, Rosamer-myc mice were injected intravitreally (to their right eyes) with 1.5ul of AAV2-Cre, AAV2-myc, AAV2-CNTF, AAV2-ShPTEN or AAV2-PLAP (titers around 1 × 1012). PTENf/f/SOCS3f/f mice (Sun et al., 2011) were injected first with AAV2-Cre and AAV2-CNTF and then AAV2-PLAP or AAV2-c-myc. For each injection, a glass micropipette was inserted to peripheral retinal, behind ora serrate to avoid damage to the lens.

Optic nerve injury

Two weeks or 4 weeks for AAV2-ShPTEN following the AAV2 injection, the right optic nerve was exposed intraorbitally and crushed with forceps (Drumont #5 FST) for 5 seconds approximately 1mm behind the eye ball, as previously described (Park et al., 2008; Sun et al., 2011).

RGC anterograde labeling

3 days before termination, 1.5ul of cholera toxin β subunit (CTB) conjugated to Alexa-488 (2ug/ul, Invitrogen) was injected into the vitreous chamber with Hamilton syringe (Hamilton) to label axons in optic nerve. All tissues were processed as described previously (Park et al., 2008; Sun et al., 2011).

Sciatic nerve crush

The right sciatic nerve was exposed and crushed with forceps (drumont #5, FST) for approximately 20 seconds. Quantification was performed as describe previously (Cho et al., 2013)

Axon regeneration quantification

Numbers of CTB labeled axons were estimated by counting the numbers of CTB-labeled axons at different distances distal from the crush site in at least 5 sections per animal as described previously (Park et al., 2008; Sun et al., 2011).

Immunohistochemistry

Sections or whole mount retinas were incubated over night at 4°C with primary antibodies: Tuj-1 (1/400 Covance); P-S6 (1/200-Cell Signaling); NeuN (1/100-Millipore) and c-myc (1/200-Abcam) diluted in PBS-3%BSA-0. 5% Triton-X100 as described in Nawabi et al 2010 (Nawabi et al., 2010). Then, tissues were incubated for 2 hours at room temperature with appropriate Alexa-fluor-conjugate secondary antibodies (Alexa-488; Alexa-594, Alexa-674, Invitrogen) diluted with the same blocking solution as the primary antibodies. Sections or whole mount retinas were mounted with DAPI-Fluoromont-G (Solulink).

RGC survival and phospho-S6 signal

Tuj1-positive cells in the ganglion layer were counted using a fluorescent microscope after immunostaining whole mount retinas with anti-Tuj-1 antibodies. 8 random fields per retina were enumerated. The average number per field was determined and the percentages of viable RGCs were obtained by comparing the values determined from the uninjured contralateral retinas. In the same condition, after phospho-S6 staining, the densities of phopsho-S6 positive RGCs were obtained by comparing the value from the uninjured contralateral retinas.

Whole mount optic nerve preparation

Optic nerves were prepared as described in (Dodt et al., 2007). Briefly optic nerves were post fixed overnight at 4°C in PFA4%, then dehydrated in ethanol (50%, 80% and 95% for 20min each) Optic nerves were incubated for 1h in ethanol 100% and then 2h in Hexan (Sigma). Finally samples were transferred in benzyl benzoate/benzyl alcohol (2:1 in volume-Sigma) clearing solution.

Imaging

Optic nerves from different experiments (sections and whole mount) were imaged using a spinning disc confocal microscope (Zeiss) with Velocity software (Perkin Elmer). 20x objective was used to image stacks of all the optic nerves with 20% of overlap between each image. The motorized stage enables the automation of the imaging procedure. All the images were then stitched together using the Velocity software and z-projected to maximal intensity for quantification. Whole mount retinas were imaged with Nikon 80i epifluorescence microscope with 20x objective. At least 8 random fields were imaged per retina.

Statistical analysis

One-way ANOVAs were performed using PRISM software and specific differences between groups were confirmed with Bonferrroni t-test. When two groups were compared, t-test was used.

FACS analysis

YFP-17 mice were subjected to optic nerve injury. 3 days after injury, animals were euthanized and injured or intact retinas were immediately dissected out for dissociation. Retinas were collected in 800μl HBSS supplemented with trypsin (5mg/ml) and DNase (0.2mg/ml) for 10min at 37°C. Retinas were washed 3 times with DMEM/10% FBS and dissociated mechanically using a P1000 pipetman. Cells were pelleted by centrifugation for 5min at 200g suspended in Neurobasal supplemented with L-glutamine and B27 plus DAPI (1mg/ml) and filtered through 40μm cell strainer (BD Falcon) before FACS sorting. FACS sorting was performed with a BD FACSAria IIu instrument. Prior to each sorting event, a purity test was performed to ensure that the specificity for the YFP signal was higher than 99%. Dissociated retinal cells were separated based on size (forward scatter) and surface characteristics (side scatter) as well as viability (DAPI staining). Aggregated cells were excluded based on the FSC-H vs FSC-A ratio. Retinal cells without YFP expression was used as negative controls to optimize the detection gate prior to each sorting. Sorted cells were immediately subjected to either proteins or RNA extraction.

Quantitative Proteomics analysis of RGC

Cells from each preparation were immediately lysed using Qproteome mammalian buffer and proteins were precipitated using chloroform/methanol. 100 ug protein from each sample was reduced with (tris 20 carboxyethyl) phosphine (TCEP) and alkylated with a 0.05 M iodoacetamide solution in the dark at room temperature (23 °C) for 20 minutes. A proteolytic digestion was carried out with trypsin at 37 °C for 16 hours. Labeling reagents from the TMT sixplex Isobaric Label Reagent Set (PI-90064, Thermo Fischer Scientific) were added to the peptide samples and incubated at room temperature for one hour. The reaction was quenched by adding 8 ml of 5% hydroxylamine to the samples, incubated at room temperature for 1 hour and combined together. Peptides were eluted from the spin columns. After acidification, the peptides were desalted on an OASIS HLB column (Waters) and evaporated to dryness, prior to isoelectric focusing fractionation according to their isoelectric point. Resultant peptide fractions were dried before re-suspension in MS loading buffer (5% acetonitrile, 0.1% formic acid in H2O).

Less than 10ug of each fraction was analyzed by nano-LC-MS/MS using a 60-minute gradient. Pulsed Q-Dissociation (PQD) on the LTQ-Orbitrap Classic was used to fragment the peptide precursor ions. Thermo RAW files were processed using an in-house informatics pipeline (Bigbang/Horizon) and searched with a Mascot (version 2.1) search engine against the IPI mouse protein database (IPI 3.68) using a target-decoy database to calculate the false discovery rate (FDR). TMT 6-plex modifications were set as dynamic modifications in the database search and as static modifications for all other data analysis. This allowed us to calculate labeling efficiencies -the labeling efficiency for our experiments was 98.73%. All PSMs used in this analysis had a MASCOT score above 32 and q value below 0.01. Only proteins identified with a high rate of confidence (1% FDR) were used for analysis. Functional analysis was performed using Ingenuity Pathways Analysis (IPA Ingenuity System). The extended methods are described in Supplemental Information.

RNA extraction

Cells were immediately lysed using a TRIzol reagent (Invitrogen) and RNA was purified using the RNeasy RNA extraction kit (Qiagen). Total RNA (50–100 ng) was reverse transcribed and amplified with real-time PCR assays iQ SYBR green (Bio-Rad). Real-time PCR was performed on a Bio-Rad CFX 96. Each sample was run four times in each assay. GADPH was used as the endogenous control.

Supplementary Material

3

4

Acknowledgments

We thank Katherine Zukor for reading the manuscript. This study was supported by a fellowship from William Randolf Hearst Foundation (SB), NIH grants EY021242 (ZH) and NS066973 (JS) and a grant from Dr. Miriam and Sheldon G. Adelson Medical Research Foundation (ZH). IDDRC and viral cores supported by the grants NIH P30 HD018655 and P30EY012196 were used for this study.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • Bareyre FM, Kerschensteiner M, Raineteau O, Mettenleiter TC, Weinmann O, Schwab ME. The injured spinal cord spontaneously forms a new intraspinal circuit in adult rats. Nat Neurosci. 2004;7:269–277. [PubMed] [Google Scholar]
  • Blackmore MG, Wang Z, Lerch JK, Motti D, Zhang YP, Shields CB, Lee JK, Goldberg JL, Lemmon VP, Bixby JL. Kruppel-like Factor 7 engineered for transcriptional activation promotes axon regeneration in the adult corticospinal tract. Proc Natl Acad Sci U S A. 2012;109:7517–7522. [PMC free article] [PubMed] [Google Scholar]
  • Bradke F, Fawcett JW, Spira ME. Assembly of a new growth cone after axotomy: the precursor to axon regeneration. Nat Rev Neurosci. 2012;13:183–193. [PubMed] [Google Scholar]
  • Cai D, Qiu J, Cao Z, McAtee M, Bregman BS, Filbin MT. Neuronal cyclic AMP controls the developmental loss in ability of axons to regenerate. J Neurosci. 2001;21:4731–4739. [PMC free article] [PubMed] [Google Scholar]
  • Chen A, Springer JE. Neuroproteomic methods in spinal cord injury. Methods Mol Biol. 2009;566:57–67. [PubMed] [Google Scholar]
  • Chen L, Wang Z, Ghosh-Roy A, Hubert T, Yan D, O’Rourke S, Bowerman B, Wu Z, Jin Y, Chisholm AD. Axon regeneration pathways identified by systematic genetic screening in C. elegans. Neuron. 2011;71:1043–1057. [PMC free article] [PubMed] [Google Scholar]
  • Chierzi S, Ratto GM, Verma P, Fawcett JW. The ability of axons to regenerate their growth cones depends on axonal type and age, and is regulated by calcium, cAMP and ERK. Eur J Neurosci. 2005;21:2051–2062. [PubMed] [Google Scholar]
  • Cho Y, Sloutsky R, Naegle KM, Cavalli V. Injury-induced HDAC5 nuclear export is essential for axon regeneration. Cell. 2013;155:894–908. [PMC free article] [PubMed] [Google Scholar]
  • Costigan M, Befort K, Karchewski L, Griffin RS, D’Urso D, Allchorne A, Sitarski J, Mannion JW, Pratt RE, Woolf CJ. Replicate high-density rat genome oligonucleotide microarrays reveal hundreds of regulated genes in the dorsal root ganglion after peripheral nerve injury. BMC Neurosci. 2002;3:16. [PMC free article] [PubMed] [Google Scholar]
  • De Lima S, Koriyama Y, Kurimoto T, Oliveira JT, Yin Y, Li Y, Gilbert HY, Fagiolini M, Martinez AM, Benowitz L. Full-length axon regeneration in the adult mouse optic nerve and partial recovery of simple visual behaviors. Proc Natl Acad Sci USA. 2012;109:9149–9154. [PMC free article] [PubMed] [Google Scholar]
  • Dang CV. MYC, Metabolism, Cell Growth, and Tumorigenesis. Cold Spring Harb Perspect Med. 2013;3 [PMC free article] [PubMed] [Google Scholar]
  • Dodt HU, Leischner U, Schierloh A, Jahrling N, Mauch CP, Deininger K, Deussing JM, Eder M, Zieglgansberger W, Becker K. Ultramicroscopy: three-dimensional visualization of neuronal networks in the whole mouse brain. Nat Methods. 2007;4:331–336. [PubMed] [Google Scholar]
  • Dupraz S, Grassi D, Karnas D, Nieto Guil AF, Hicks D, Quiroga S. The insulin-like growth factor 1 receptor is essential for axonal regeneration in adult central nervous system neurons. PLoS ONE. 2013;8:e54462. [PMC free article] [PubMed] [Google Scholar]
  • Fawcett JW. Overcoming inhibition in the damaged spinal cord. J Neurotrauma. 2006;23:371–383. [PubMed] [Google Scholar]
  • Filbin MT. Myelin-associated inhibitors of axonal regeneration in the adult mammalian CNS. Nat Rev Neurosci. 2003;4:703–713. [PubMed] [Google Scholar]
  • Fischer D, Petkova V, Thanos S, Benowitz LI. Switching mature retinal ganglion cells to robust growth state in vivo: gene expression and synergy with RhoA inactivation. J Neurosci. 2004;40:8726–40. [PMC free article] [PubMed] [Google Scholar]
  • Ghosh-Roy A, Wu Z, Goncharov A, Jin Y, Chisholm AD. Calcium and cyclic AMP promote axonal regeneration in Caenorhabditis elegans and require DLK-1 kinase. J Neurosci. 2010;30:3175–3183. [PMC free article] [PubMed] [Google Scholar]
  • Goldberg JL, Espinosa JS, Xu Y, Davidson N, Kovacs GT, Barres BA. Retinal ganglion cells do not extend axons by default: promotion by neurotrophic signaling and electrical activity. Neuron. 2002a;33:689–702. [PubMed] [Google Scholar]
  • Goldberg JL, Klassen MP, Hua Y, Barres BA. Amacrine-signaled loss of intrinsic axon growth ability by retinal ganglion cells. Science. 2002b;296:1860–1864. [PubMed] [Google Scholar]
  • Griffin TJ, Xie H, Bandhakavi S, Popko J, Mohan A, Carlis JV, Higgins L. iTRAQ reagent-based quantitative proteomic analysis on a linear ion trap mass spectrometer. J Proteome Res. 2007;6:4200–4209. [PMC free article] [PubMed] [Google Scholar]
  • Guertin DA, Sabatini DM. Defining the role of mTOR in cancer. Cancer cell. 2007;12:9–22. [PubMed] [Google Scholar]
  • Hammarlund M, Nix P, Hauth L, Jorgensen EM, Bastiani M. Axon regeneration requires a conserved MAP kinase pathway. Science. 2009;323:802–806. [PMC free article] [PubMed] [Google Scholar]
  • Harel NY, Yigitkanli K, Fu Y, Cafferty WB, Strittmatter SM. Multimodal exercises simultaneously stimulating cortical and brainstem pathways after unilateral corticospinal lesion. Brain Res. 2013;1538:17–25. [PMC free article] [PubMed] [Google Scholar]
  • Holtmaat A, Svoboda K. Experience-dependent structural synaptic plasticity in the mammalian brain. Nat Rev Neurosci. 2009;10:647–658. [PubMed] [Google Scholar]
  • Howell JJ, Ricoult SJ, Ben-Sahra I, Manning BD. A growing role for mTOR in promoting anabolic metabolism. Biochem Soc Trans. 2013;41:906–912. [PubMed] [Google Scholar]
  • Huang da W, Sherman BT, Lempicki RA. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc. 2009;4:44–57. [PubMed] [Google Scholar]
  • Jager R, Maurer J, Jacob A, Schorle H. Cell type-specific conditional regulation of the c-myc proto-oncogene by combining Cre/loxP recombination and tamoxifen-mediated activation. Genesis. 2004;38:145–150. [PubMed] [Google Scholar]
  • Kim MS, Pinto SM, Getnet D, Nirujogi RS, Manda SS, Chaerkady R, Madugundu AK, Kelkar DS, Isserlin R, Jain S, et al. A draft map of the human proteome. Nature. 2014;509:575–581. [PMC free article] [PubMed] [Google Scholar]
  • Laplante M, Sabatini DM. mTOR signaling in growth control and disease. Cell. 2012;149:274–293. [PMC free article] [PubMed] [Google Scholar]
  • Lin CY, Loven J, Rahl PB, Paranal RM, Burge CB, Bradner JE, Lee TI, Young RA. Transcriptional amplification in tumor cells with elevated c-Myc. Cell. 2012;151:56–67. [PMC free article] [PubMed] [Google Scholar]
  • Liu K, Lu Y, Lee JK, Samara R, Willenberg R, Sears-Kraxberger I, Tedeschi A, Park KK, Jin D, Cai B, et al. PTEN deletion enhances the regenerative ability of adult corticospinal neurons. Nat Neurosci. 2010;13:1075–1081. [PMC free article] [PubMed] [Google Scholar]
  • Luo X, Salgueiro Y, Beckerman SR, Lemmon VP, Tsoulfas P, Park KK. Three-dimensional evaluation of retinal ganglion cell axon regeneration and pathfinding in whole mouse tissue after injury. Exp Neurol. 2013;247:653–662. [PMC free article] [PubMed] [Google Scholar]
  • Magharious M, D’Onofrio PM, Hollander A, Zhu P, Chen J, Koeberle PD. Quantitative iTRAQ analysis of retinal ganglion cell degeneration after optic nerve crush. J Proteome Res. 2011;10:3344–3362. [PubMed] [Google Scholar]
  • Michaelevski I, Segal-Ruder Y, Rozenbaum M, Medzihradszky KF, Shalem O, Coppola G, Horn-Saban S, Ben-Yaakov K, Dagan SY, Rishal I, et al. Signaling to transcription networks in the neuronal retrograde injury response. Sci Signal. 2010;3:ra53. [PMC free article] [PubMed] [Google Scholar]
  • Moore DL, Blackmore MG, Hu Y, Kaestner KH, Bixby JL, Lemmon VP, Goldberg JL. KLF family members regulate intrinsic axon regeneration ability. Science. 2009;326:298–301. [PMC free article] [PubMed] [Google Scholar]
  • Nawabi H, Briancon-Marjollet A, Clark C, Sanyas I, Takamatsu H, Okuno T, Kumanogoh A, Bozon M, Takeshima K, Yoshida Y, et al. A midline switch of receptor processing regulates commissural axon guidance in vertebrates. Genes Dev. 2010;24:396–410. [PMC free article] [PubMed] [Google Scholar]
  • Toth C, Brussee V, Martinez JA, McDonald D, Cunningham FA, Zochodne DW. Rescue and regeneration of injured peripheral nerve axons by intrathecal insulin. Neuroscience. 2006;139:429–449. [PubMed] [Google Scholar]
  • Near SL, Whalen LR, Miller JA, Ishii DN. Insulin-like growth factor II stimulates motor nerve regeneration. Proc Natl Acad Sci U S A. 1992;89:11716–11720. [PMC free article] [PubMed] [Google Scholar]
  • Nie Z, Hu G, Wei G, Cui K, Yamane A, Resch W, Wang R, Green DR, Tessarollo L, Casellas R, et al. c-Myc is a universal amplifier of expressed genes in lymphocytes and embryonic stem cells. Cell. 2012;151:68–79. [PMC free article] [PubMed] [Google Scholar]
  • Park J, Al-Ramahi I, Tan Q, Mollema N, Diaz-Garcia JR, Gallego-Flores T, Lu HC, Lagalwar S, Duvick L, Kang H, et al. RAS-MAPK-MSK1 pathway modulates ataxin 1 protein levels and toxicity in SCA1. Nature. 2013;498:325–331. [PMC free article] [PubMed] [Google Scholar]
  • Park KK, Liu K, Hu Y, Smith PD, Wang C, Cai B, Xu B, Connolly L, Kramvis I, Sahin M, et al. Promoting axon regeneration in the adult CNS by modulation of the PTEN/mTOR pathway. Science. 2008;322:963–966. [PMC free article] [PubMed] [Google Scholar]
  • Raineteau O, Schwab ME. Plasticity of motor systems after incomplete spinal cord injury. Nat Rev Neurosci. 2001;2:263–273. [PubMed] [Google Scholar]
  • Rosenzweig ES, Courtine G, Jindrich DL, Brock JH, Ferguson AR, Strand SC, Nout YS, Roy RR, Miller DM, Beattie MS, et al. Extensive spontaneous plasticity of corticospinal projections after primate spinal cord injury. Nat Neurosci. 2010;13:1505–1510. [PMC free article] [PubMed] [Google Scholar]
  • Rossi F, Gianola S, Corvetti L. Regulation of intrinsic neuronal properties for axon growth and regeneration. Prog Neurobiol. 2007;81:1–28. [PubMed] [Google Scholar]
  • Saul KE, Koke JR, Garcia DM. Activating transcription factor 3 (ATF3) expression in the neural retina and optic nerve of zebrafish during optic nerve regeneration. Comp Biochem Physiol A Mol Integr Physiol. 2010;155:172–182. [PubMed] [Google Scholar]
  • Shaw RJ, Cantley LC. Ras, PI(3)K and mTOR signalling controls tumour cell growth. Nature. 2006;441:424–430. [PubMed] [Google Scholar]
  • Singh SA, Winter D, Kirchner M, Chauhan R, Ahmed S, Ozlu N, Tzur A, Steen JA, Steen H. Co-regulation proteomics reveals substrates and mechanisms of APC/C-dependent degradation. Embo J. 2014;33:385–399. [PMC free article] [PubMed] [Google Scholar]
  • Staab CA, Ceder R, Jagerbrink T, Nilsson JA, Roberg K, Jornvall H, Hoog JO, Grafstrom RC. Bioinformatics processing of protein and transcript profiles of normal and transformed cell lines indicates functional impairment of transcriptional regulators in buccal carcinoma. J Proteome Res. 2007;6:3705–3717. [PubMed] [Google Scholar]
  • Sun F, Park KK, Belin S, Wang D, Lu T, Chen G, Zhang K, Yeung C, Feng G, Yankner BA, et al. Sustained axon regeneration induced by co-deletion of PTEN and SOCS3. Nature 2011 [PMC free article] [PubMed] [Google Scholar]
  • Sung YJ, Povelones M, Ambron RT. RISK-1: a novel MAPK homologue in axoplasm that is activated and retrogradely transported after nerve injury. J Neurobiol. 2001;47:67–79. [PubMed] [Google Scholar]
  • Tanabe K, Bonilla I, Winkles JA, Strittmatter SM. Fibroblast growth factor-inducible-14 is induced in axotomized neurons and promotes neurite outgrowth. J Neurosci. 2003;23:9675–9686. [PMC free article] [PubMed] [Google Scholar]
  • Thompson A, Schafer J, Kuhn K, Kienle S, Schwarz J, Schmidt G, Neumann T, Johnstone R, Mohammed AK, Hamon C. Tandem mass tags: a novel quantification strategy for comparative analysis of complex protein mixtures by MS/MS. Anal Chem. 2003;75:1895–1904. [PubMed] [Google Scholar]
  • Toth C, Brussee V, Martinez JA, McDonald D, Cunningham FA, Zochodne DW. Rescue and regeneration of injured peripheral nerve axons by intrathecal insulin. Neuroscience. 2006;139:429–449. [PubMed] [Google Scholar]
  • van den Brand R, Heutschi J, Barraud Q, DiGiovanna J, Bartholdi K, Huerlimann M, Friedli L, Vollenweider I, Moraud EM, Duis S, et al. Restoring voluntary control of locomotion after paralyzing spinal cord injury. Science. 2012;336:1182–1185. [PubMed] [Google Scholar]
  • Watkins TA, Wang B, Huntwork-Rodriguez S, Yang J, Jiang Z, Eastham-Anderson J, Modrusan Z, Kaminker JS, Tessier-Lavigne M, Lewcock JW. DLK initiates a transcriptional program that couples apoptotic and regenerative responses to axonal injury. Proc Natl Acad Sci U S A. 2013;110:4039–4044. [PMC free article] [PubMed] [Google Scholar]
  • Wilhelm M, Schlegl J, Hahne H, Moghaddas Gholami A, Lieberenz M, Savitski MM, Ziegler E, Butzmann L, Gessulat S, Marx H, et al. Mass-spectrometry-based draft of the human proteome. Nature. 2014;509:582–587. [PubMed] [Google Scholar]
  • Xie J, Proud CG. Crosstalk between mTOR complexes. Nat Cell Biol. 2013;15:1263–1265. [PubMed] [Google Scholar]
  • Zhu LQ, Zheng HY, Peng CX, Liu D, Li HL, Wang Q, Wang JZ. Protein phosphatase 2A facilitates axonogenesis by dephosphorylating CRMP2. J Neurosci. 2010;30:3839–3848. [PMC free article] [PubMed] [Google Scholar]
  • Zukor K, Belin S, Wang C, Keelan N, Wang X, He Z. Short hairpin RNA against PTEN enhances regenerative growth of corticospinal tract axons after spinal cord injury. J Neurosci. 2013;33:15350–15361. [PMC free article] [PubMed] [Google Scholar]
-