49
Epigenomic signatures underpinning the axonal regenerative ability of dorsal root ganglia sensory neurons Ilaria Palmisano 1*§ , Matt C. Danzi 2,3§ , Thomas H.Hutson 1 , Luming Zhou 1 , Eilidh McLachlan 1 , Elisabeth Serger 1 , Kirill Shkura 4 , Prashant K. Srivastava 4,5 , Arnau Hervera 1,.6 , Nick O’Neill 2 , Tong Liu 7 , Hassen Dhrif 7 , Zheng Wang 7 , Miroslav Kubat 8 , Stefan Wuchty 3,7,9,10 , Matthias Merkenschlager 11 , Liron Levi 12 , Evan Elliott 12 , John L. Bixby 2,3 , Vance P. Lemmon 2,3 and Simone Di Giovanni 1* 1. Centre for Restorative Neuroscience, Division of Brain Sciences, Department of Medicine, Imperial College London, London, UK. 2. The Miami Project to Cure Paralysis, University of Miami, Miami, FL, USA. 3. Center for Computational Science, University of Miami, Miami, FL, USA. 4. Integrative Genomics and Medicine, Division of Brain Sciences, Department of Medicine, Imperial College London, London, UK. 5. National Heart & Lung Institute, Imperial College London, London, UK. 6. Molecular and Cellular Neurobiotechnology, Institut for Bioengineering of Catalonia (IBEC), Barcelona, Spain. 7. Department of Computer Science, University of Miami, Miami, FL, USA. 8. Dept. of Electrical & Computer Engineering, University of Miami, Miami, FL, USA. 9. Department of Biology, University of Miami, Miami, FL, USA. 10. Sylvester Comprehensive Cancer Center, University of Miami, Miami, FL, USA. 1

T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

Epigenomic signatures underpinning the axonal regenerative ability of dorsal root ganglia sensory

neurons

Ilaria Palmisano1*§, Matt C. Danzi2,3§, Thomas H.Hutson1, Luming Zhou1, Eilidh McLachlan1, Elisabeth

Serger1, Kirill Shkura4, Prashant K. Srivastava4,5, Arnau Hervera1,.6, Nick O’Neill2, Tong Liu7, Hassen

Dhrif7, Zheng Wang7, Miroslav Kubat8, Stefan Wuchty3,7,9,10, Matthias Merkenschlager11, Liron Levi12,

Evan Elliott12, John L. Bixby2,3, Vance P. Lemmon2,3 and Simone Di Giovanni1*

1. Centre for Restorative Neuroscience, Division of Brain Sciences, Department of Medicine, Imperial

College London, London, UK.

2. The Miami Project to Cure Paralysis, University of Miami, Miami, FL, USA.

3. Center for Computational Science, University of Miami, Miami, FL, USA.

4. Integrative Genomics and Medicine, Division of Brain Sciences, Department of Medicine, Imperial

College London, London, UK.

5. National Heart & Lung Institute, Imperial College London, London, UK.

6. Molecular and Cellular Neurobiotechnology, Institut for Bioengineering of Catalonia (IBEC),

Barcelona, Spain.

7. Department of Computer Science, University of Miami, Miami, FL, USA.

8. Dept. of Electrical & Computer Engineering, University of Miami, Miami, FL, USA.

9. Department of Biology, University of Miami, Miami, FL, USA.

10. Sylvester Comprehensive Cancer Center, University of Miami, Miami, FL, USA.

11. MRC, London Institute of Medical Sciences, Faculty of Medicine, Imperial College London,

London, UK.

12. Faculty of Medicine, Bar Ilan University, Safed, Israel.

§The authors contributed equally to the work

* Correspondence should be addressed to: Simone Di Giovanni, MD, PhD, or Ilaria Palmisano, PhD,

Division of Brain Sciences, Imperial College London, Hammersmith Hospital Campus, Imperial College

London, Du Cane Road, W12 0NN. e.mail: [email protected] or

[email protected]

KEY WORDS

1

Page 2: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

Chromatin remodelling, Epigenetics, Open Chromatin, Transcription Factors, Spinal Cord Injury, Axon

Regeneration, Machine learning, CTCF

ABSTRACT

Axonal injury results in regenerative success or failure depending on whether the axon lies in the

peripheral or the central nervous system respectively. Here, we addressed whether epigenetic

signatures in dorsal root ganglia discriminate between regenerative and non-regenerative axonal

injury. We performed RNA, chromatin immunoprecipitation for H3K9ac, H3K27ac, and H3K27me3

and Assay for Transposase-Accessible Chromatin sequencing in DRG after sciatic nerve or dorsal

column axotomy. Distinct histone acetylation and chromatin accessibility signatures correlated with

gene expression following peripheral but not central axonal injury. DNA footprinting analyses

revealed novel transcriptional regulators associated with regenerative ability. Machine learning

algorithms inferred the direction of the majority of gene expression changes. Neuronal conditional

deletion of the chromatin remodeler CTCF impaired nerve regeneration implicating chromatin

organization in the regenerative capacity. Altogether, we offer the first epigenomic map providing

insight into the transcriptional response to injury and into the differential regenerative ability of

sensory neurons.

2

Page 3: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

INTRODUCTION

A peripheral nerve injury initiates multiple signalling pathways, converging into the activation of

transcription factors (TF) 1-3, and resulting in the fine tuning of a robust regenerative programme

characterized by the expression of regeneration associated genes (RAGs) 4. In contrast, poor

neuronal regenerative gene expression as well as the presence of glial inhibitory signals lead to

regenerative failure upon axonal injury in the central nervous system (CNS) 5.

The dorsal root ganglia (DRG) sensory neurons allow the direct comparison of molecular

events initiated in the same cell body upon a peripheral vs a central axonal lesion. The DRG

pseudounipolar neurons extend a peripheral regeneration-competent axonal branch that lies in the

peripheral nervous system (PNS) and a central regeneration-incompetent axonal branch that enters

the dorsal columns in the spinal cord 6.

Epigenetic mechanisms play a critical role in gene regulation by controlling the accessibility

of gene promoters and enhancers to TFs 7. These include DNA methylation and hydroxymethylation,

histone post-translational modifications (HPTMs), exchange of histone variants, and binding or

displacement of architectural proteins 8-10. Recent evidence has begun to uncover the importance of

epigenetic processes in regulating the neuronal regenerative transcriptional programme 11-14. Cavalli

cell, finelli JN and pcaf radhika’s papers must be added Nevertheless, it remains to be determined (i)

whether chromatin conformation and accessibility are modified after axonal injury, (ii) whether

chromatin accessibility is associated with other epigenomic reprogramming events, such as histone

modification at cis-regulatory elements, and (iii) whether changes in the transcriptional programme

might be predictive of regenerative success vs failure.

Therefore, we systematically investigated epigenomic signatures in response to axonal

injury. Our study provides the first comprehensive map of epigenomic modifications after axonal

injury and it demonstrates that epigenetic and transcriptional changes discriminate between

regenerative peripheral and non-regenerative central axonal injury. Moreover, it provides new

3

Page 4: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

insight into the epigenetic mechanisms that might be harnessed to reprogram neurons into a

regenerative mode.

RESULTS

Overview of the RNAseq, ATACseq, and ChIPseq datasets

Initially, we collected sciatic DRG 24h after an injury of the peripheral branches following a sciatic

nerve axotomy (SNA) vs. sham or after injury of the central branches following a dorsal column

axotomy (DCA) vs. laminectomy (lam). Thereafter, we performed: i) RNAseq 15; ii) ATACseq, in which

the Tn5 transposase exclusively integrates sequencing adapters into open and accessible chromatin,

allowing identification of accessible chromatin regions 16; iii) ChIPseq for H3K9ac 13 and H3K27ac

(marks of active promoters and enhancers) and H3K27me3 (repressive mark) (Figure 1A), which

resulted in a significant DNA enrichment with respect to control IgG (Figure S1A). Pairwise

comparisons of signal within peaks in ATACseq replicate samples indicated high reproducibility

among triplicates (Pearson correlation coefficient threshold >0.7). We also observed strong

reproducibility among replicates in the signal surrounding the differentially accessible (DA)

transcription start sites (TSS+/-1Kb) (Figure S1B). In sham or laminectomy treated animals, ATACseq,

H3K27ac, and H3K9ac signal was identified preferentially at promoters, enhancers, and gene bodies

(Figure 1B and S1C, red), whilst centromeric and telomeric heterochromatin regions were poorly

represented (Figure 1B and S1C, pink and blue). In contrast, H3K27me3 was found selectively

enriched at genomic locations of repressed chromatin, such as centromeric and pericentromeric

regions (red) and depleted at enhancers and promoters of expressed genes (blue). The integration of

chromatin accessibility and ChIPseq signal at the TSS with gene expression data revealed a strong

correlation, with highly expressed genes showing higher levels of chromatin accessibility and

acetylation of H3K27 and H3K9, and lower levels of methylation of H3K27 in comparison to the lower

expressed genes (Figure 1C and D).

Peripheral axotomy induces an increase in chromatin accessibility and in histone acetylation

4

Page 5: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

We quantified the changes in chromatin accessibility and histone occupancy upon the two injury

paradigms. Specifically, SNA resulted in a large gain in chromatin accessibility (2,447 more accessible

and 85 less accessible genes, Pvalue<0.05), while few alterations in chromatin accessibility occurred

after DCA (389 more accessible and 106 less accessible genes, Pvalue<0.05) (Figure 2A-B and

Supplementary Data 1). Gene ontology (GO) analysis of the genes showing a gain in chromatin

accessibility after SNA revealed a significant enrichment (Pvalue< 0.01) for several functional

categories related to calcium channel activity, regulation of transcription, chromatin remodelling,

cell signalling, axon transport, and proteasome regulation (Figure 2C and Supplementary Data 2).

Consistent with a more relaxed chromatin conformation upon peripheral injury, H3K9ac and

H3K27ac occupancy was increased on 1,909 and 1,205 TSS regions (Pvalue<0.05) respectively, while

only 64 and 136 TSS regions showed decreased occupancy; 511 TSS regions showed a decreased

H3K27me3 occupancy (Figure 2D). In contrast, the changes after DCA were more modest: the

greatest was a decreased occupancy of H3K27me3 on 587 TSS regions (Figure 2D). A similar pattern

was found at the gene body level (Figure S1D and E), with several genes showing an increased

occupancy of H3K9ac and H3K27ac at both TSS and gene body (Figure S1F). The level of histone

acetylation at the promoter regions correlated with the degree of chromatin accessibility (Figure 2E

and S1G; genes with increased accessibility showed preferential H3ac occupancy: Pearson corr=

0.039 (H3K9ac_SNA); 0.033 (H3K27ac_SNA); 0.01 (H3K27me3_SNA); 0.024 (H3K9ac_DCA); 0.028

(H3K27ac_DCA); -0.072 (H3K27me3_DCA)).

Taken together, our findings imply that peripheral but not central axonal injury promotes a

more relaxed chromatin conformation that may allow the expression of genes involved in

regenerative-associated molecular processes.

Peripheral axotomy induces robust gene expression changes associated with enhanced chromatin

accessibility

5

Page 6: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

In line with the ATACseq and ChIPseq findings, the transcriptome analysis revealed that a higher

number of genes were differentially regulated after SNA than after DCA (3,082 genes upon SNA and

830 upon DCA, Pvalue<0.05) (Figure 3A). Interestingly, only a minority of genes (16) were co-

regulated between the two injuries, and over 40% of the genes differentially expressed (DE) upon

DCA showed an opposite regulation after SNA (Figure 3B). In agreement with previous findings 17-19,

GO term enrichment analysis (Pvalue<0.01) of the upregulated genes upon SNA identified several

categories related to transcription, inflammation, developmental and biosynthetic processes, and

injury response, while genes that were downregulated after SNA were enriched for categories

related to ion channels, ion transport, axonogenesis, and synapse (Figure S2A and Supplementary

Data 3). In contrast, DCA-upregulated transcripts mainly belonged to energy metabolism, particularly

related to mitochondrial oxidative phosphorylation (Figure S2A and Supplementary Data 3). Pathway

analysis of the DE genes revealed significant enrichment (Pvalue<0.05) for multiple axonal

regenerative signalling pathways upon SNA, including neurotrophin, JAK-STAT, MAPK, insulin, and

mTOR signalling (Figure S2B and Supplementary Data 3). In contrast, pathways enriched upon DCA

were related to neurological disorders and mitochondrial oxidative metabolism.

The enhanced accessibility of chromatin regions upon SNA correlated with changes in gene

expression (Figure 3C, Fisher’s exact Pvalue=2e-183), while decreased or unchanged chromatin

accessibility was associated with non-DE genes (Fisher’s exact Pvalue=2e-18 and 4e-123). This close

correlation was further supported by the Gene Set Enrichment Analysis (GSEA, see Methods)

showing that genes with gain in accessibility after SNA were enriched for upregulated transcripts

(Figure 3D, Normalized Enrichment Score, NES=1.18, FDR=0.06). In contrast, the few less accessible

genes (85) were not significantly enriched (Figure 3E).

Gene clusters functionally associated to axonal regeneration share similar epigenetic signatures

We next combined the chromatin landscape data with the differential gene expression results upon

SNA or DCA. Across the genome, DE genes were marked for ATACseq and H3K9ac and H3K27ac

6

Page 7: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

active histone marks (Figure 4A and S3A), resulting in a correlation between gene expression and

these epigenetic signals (Figure S3B). Specifically, 321 and 305 upregulated genes correlated with an

increased H3K9ac and H3K27ac occupancy at the promoter and gene body regions following SNA

respectively, representing altogether about 35% of the upregulated genes upon SNA; fewer

downregulated genes (252 and 144) showed increased histone acetylation (Figure 4B-C). Following

DCA, we also found 59 upregulated genes that correlated with a decreased occupancy of H3K27me3

(Figure 4B-C). Upregulated genes after SNA with increased histone acetylation occupancy were

enriched for GO categories related to transcription and nucleosome/chromatin assembly,

underlining the importance of chromatin remodelling and gene expression regulations after

regenerative axonal injury (Figure 4D and Supplementary Data 4). Genes known to be associated

with axonal regeneration were well represented among the upregulated transcripts, and they

localised on hyperacetylated chromatin regions more accessible upon SNA (Figure 4E and S3C).

Together, these results indicate that specific gene functional categories share an epigenetic

pattern characterized by a more open and hyperacetylated chromatin configuration.

Enhancers are associated with increased chromatin accessibility following peripheral axonal injury

only

Enhancers are cis-regulatory elements that can mediate gene transcription via long-range

interactions with their target genes. We took advantage of the Encyclopedia of DNA Elements

(ENCODE) collection of more than 200,000 potential mouse enhancers 20, 21 and filtered the list for

those with an overlapping H3K27ac peak in at least one of the tested biological conditions to identify

enhancers active in DRGs (Supplementary Data 5). Differential accessibility analysis revealed that a

higher number of enhancers were more accessible after SNA than DCA (3,677 more accessible after

SNA and 1,620 after DCA, Pvalue<0.05, Figure 5A). Enhancers with gain in chromatin accessibility

upon SNA were about 8 fold more abundant than the ones with decreased accessibility (Figure 5A).

However, while more enhancers had changes in H3K27ac upon SNA than DCA (5,414 after SNA and

7

Page 8: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

1,109 after DCA), a higher number of those differentially acetylated enhancers had decreased than

increased H3K27ac signal after SNA (2,168 enhancers with increased and 3,246 with decreased

signal) (Figure 5B). We also found a higher number of enhancers with a decreased H3K27me3

occupancy upon DCA (1,148) versus SNA (687), which might reflect release from Polycomb complex

2 repression (Figure 5B).

We next exploited the enhancer/promoter unit (EPU) annotation from Shen et al. 20,

retrieving information about predicted enhancer-dependent target genes. While many DE genes

were identified as targets of enhancers with differential histone marks, we did not observe a clear

pattern between the regulation of the enhancer and the direction of gene expression change (Figure

5C).

Many of the axonal regenerative pathways identified so far are also involved in neuronal

plasticity during development 22. Therefore, we used a published enhancer activity profiling in the

developing forebrain across a time course from embryonic day 11 to post-natal day 56 22 to assess

how many of these developmentally regulated enhancers were responsive to injury. We found that a

high proportion of injury-responsive neuronal enhancers were developmentally regulated, including

20% of the enhancers showing increased H3K27ac and about 50% increased chromatin accessibility

(Figure 5D-F).

Distinct repertoires of transcription factors bind to accessible promoters and enhancers after

peripheral or central axonal injury

Since promoters and enhancers become more accessible after a peripheral axotomy, we decided to

identify potential TFs that could operate in proximity of accessible DNA cis-regulatory elements. To

this end, we performed a differential TF footprinting analysis using the Bivariate Genomic

Footprinting (BaGfoot) pipeline, which allows the prediction and identification of putative TF

activators or repressors by quantifying the difference in footprint depth (FPD) and flanking

accessibility (FA) between two experimental conditions 23. The analysis (Supplementary Data 6)

8

Page 9: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

revealed a set of TFs showing increased occupancy with a deeper footprint and with increased

chromatin accessibility at enhancers and promoters following SNA only (Figure 6A and C, I quadrant,

and 6 E). They included several transcriptional activators acting at DNA motifs recognized by the FOS,

JUN, CREB, and CCAAT/Enhancer binding protein (CEBP) families (Figure 6A and C). Many of these

are well-established factors involved in axon regeneration including after SNA 1-3, 17, 24-26.remove refs

24-26 here Other TFs, including members of the Homeobox (HOX) and the ONECUT family, were

released after SNA resulting in reduced chromatin accessibility in the flanking regions (Figure 6A and

C, III quadrant). Interestingly, only PROP Paired-Like Homeobox (PROP), POU Paired Box (PAX), AT-

Rich Interaction Domain (ARID), and HOX factors were identified as repressors, inducing a decrease

in FA (Figure 6A and C, II quadrant). DCA was characterized by the lack of substantial changes in TF

binding with a tendency for TF repressor recruitment (Figure 6B and D, II quadrant, and 6F). This

provides the first evidence that a regenerative peripheral axonal injury triggers robust transcriptional

activation via the establishment of accessible and dynamic chromatin, while a non-regenerative

central injury is mainly associated with a poor recruitment of transcription regulators.

Next, we asked whether HDAC inhibition leading to increased histone acetylation would

mimic enhanced gene accessibility and expression observed following SNA. Immediately following

DCA, we intraperitoneally injected MS275, a HDAC class I inhibitor previously shown to enhance

axon growth in vitro and to reduce axon retraction in vivo following spinal cord injury (SCI) 27. As

expected, MS275 delivery induced increased H3 K9 and K27 acetylation and changes in chromatin

accessibility that correlated with gene expression (Figure S4A-D and Supplementary Data 7).

Interestingly, MS275 treatment resulted in only a 12% overlap with the SNA-dependent

transcriptional programme (Supplementary Data 8). Indeed, the GO analysis revealed that the

chromatin accessibility changes upon MS275 occurred at different genomic locations compared to

SNA. Increased chromatin accessibility was observed for genes mainly related to neuronal

differentiation, axonogenesis, and nucleoside metabolism. Decreased chromatin accessibility was

found in genes mainly related to chromatin organization, mRNA processing and translation (Figure

9

Page 10: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

S4E and Supplementary Data 9). The BaGfoot analysis revealed that MS275 treatment was

associated with the increased or decreased occupancy of several TFs at the level of TSS and

enhancers (Figure S4F and G). However, the TF repertoire was clearly distinct from the one observed

upon SNA (Fig.6A and C). Together these data indicate that SNA-dependent changes in chromatin

accessibility and gene expression are poorly recapitulated by HDAC inhibition following DCA.

To gain further insight into the SNA-dependent genes that are controlled by the identified

TFs, we used the Hidden Markov Model (HMM)-based IdeNtification of Transcription factor

footprints (HINT) pipeline 28. This allows the retrieval of the chromosome coordinates for the

footprints found in the DA regions upon SNA or DCA (Supplementary Data 10). A complementary in

silico enrichment motif analysis performed at the level of the DA regions and of the DE genes

revealed a convergence on similar sets of TFs (Figure S5 and Supplementary Data 11). When

comparing SNA to sham, we found a clear relationship between changes in TF expression and their

footprint signal (Spearman correlation r=0.53; Pvalue=2.523e-8 Figure S6A). However, the

comparison of DCA to control laminectomy did not show any such correlation (Spearman correlation

r=0.02, Pvalue=0.829, Figure S6B), consistently with the BaGfoot analysis. About 60,503 and 33,978

footprints were found at DA regions after SNA and DCA, respectively, and these were associated

with a total of 580 and 486 TFs (Supplementary Data 10), mostly shared between promoters and

enhancers.

We then retrieved the target genes of the footprints found by HINT within DA promoters

and enhancers (see Methods and Supplementary Data 12). About 2,145 and 618 genes were found

associated with footprints identified in DA regions following SNA and DCA respectively. Among

them, a higher number showed an increased occupancy of H3K9ac and H3K27ac, and, as expected,

increased chromatin accessibility upon SNA than DCA, resulting in 478 genes significantly DE

following SNA whilst only 40 genes DE upon DCA (Figure 7A-B). This suggests dissimilar TF

recruitment that differentially regulates the transcriptional programme between peripheral and

central axonal injury.

10

Page 11: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

Since DRG contain both neurons and non-neuronal satellite cells, we took advantage of the

single DRG neuron RNAseq analysis performed by Usoskin and colleagues 29 to measure how many of

the identified TFs and DE genes after SNA or DCA were expressed in neurons. Although we were

unable to determine expression levels in DRG cell types other than neurons, as they have not been

profiled with single-cell resolution yet, the large majority of TFs and genes in our analysis was indeed

expressed in neurons (Figure S6C).

Random forest classifier is able to differentiate between upregulated and downregulated genes

following injury based on TF footprints

We next examined whether epigenetic and transcriptional information collected in our datasets

could be used to predict whether genes are upregulated or downregulated after each axonal injury.

To this end, we used the program TEPIC (see online Methods in Supplementary Information) to

generate a score quantifying the effect of each TF on the expression of each gene using HINT

footprints and chromatin accessibility in each injury condition as input, including additional features

for histone occupancy and DNA content at the TSS and gene body. We next subtracted the scores in

the uninjured from the scores in the injured condition for each DE gene (except for the DNA content

features, as those values do not change in response to injury). Finally, we labelled each gene with a

specific class according to its change in expression after injury.

A random forest classifier trained on this input data as part of a five-fold cross-validation was

then able to make out-of-set predictions on whether a gene is upregulated or downregulated with

74% accuracy following peripheral axotomy (75% sensitivity and 72% specificity) and with 77%

accuracy following central axotomy (84% sensitivity and 68% specificity) (Figure 7C). Such high

performance indicates that the random forest algorithm successfully learned patterns in the input

data that explain whether a gene is upregulated or downregulated after axotomy.

In order to resolve which specific features were most informative to predict the direction of

gene expression changes, we employed a feature selection method based on particle swarm

11

Page 12: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

optimization (PSO) (see On line Methods in Supplementary Information). Besides the CG content,

the top 10 non-DNA based features most frequently selected from the peripheral axotomy dataset

(Figure 7D) were related to chromatin accessibility, H3K27ac level and several TFs. Most of these

bind the same motifs that were highlighted by the BaGfoot analysis as bound to putative recruited

activators (Figure 6A and C): i.e. TFs that bind to cAMP Response Elements (CRE), and TFs of the AP-1

family. Additional TFs frequently selected by the PSO include Zinc finger and BTB domain containing

7C (ZBTB7C) and members of the ETS variant (ETV) and pre B cell leukemia homeobox (PBX) families

as well as interesting novel putative regulators of axon regeneration such as SOX10 and NANOG.

From the DCA dataset, the PSO algorithm retrieved TFs found in the previous BaGfoot analysis,

mainly associated with a reduced accessibility, as well as additional factors (Figure 7E).

The chromatin organiser CTCF is implicated in the conditioning-dependent DRG regenerative

growth and in the regeneration of DRG sensory axons

The ATACseq experiments allowed the identification of footprints of factors involved in chromatin

remodelling, such as FOXA, ARID3A, Nuclear Factor1 (NF1), and the CCCTC-binding factor (CTCF)

(Supplementary Data 10). Given the role of CTCF in neuronal biology 30, and its wide influence on

both transcription and chromatin organisation 31, we investigated if CTCF signatures would mark

axonal injury-responsive genes with chromatin changes. The in silico motif enrichment analysis

revealed an enrichment of CTCF at the promoter of injury-dependent DE genes (Supplementary Data

11). To better identify possible target genomic regions, we combined our datasets with published

and curated databases of CTCF binding sites (BS) 32, 33. When analysed at the TSS level, the genomic

regions containing CTCF BS were characterized by higher chromatin accessibility (1034 genes) and

higher histone acetylation (728 and 302 genes with increased H3K9ac and H3K27ac occupancy,

respectively) after SNA compared to sham, while this did not occur following DCA (Figure S7A). Since

CTCF BS were also found on active enhancers in DRG (Figure S7B), we then identified target genes

having a CTCF BS at promoters (TSS +-/- 1 Kb) or enhancers (+/- 1Kb) (Supplementary Data 13). We

12

Page 13: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

found that 666 and 880 target genes, representing about 50% of the injury-dependent

transcriptional programme, were up and down-regulated upon SNA, respectively (Figure S7C). GO

enrichment analysis of these genes suggests that CTCF might contribute to the regulation of a wide

range of the SNA-dependent transcriptional programme (Figure S7D and Supplementary Data 14).

Interestingly, interrogating a published RNAseq from CTCF-cKO mouse hippocampus (GSE84175)33,

we found that several of these GO categories were affected by the deletion of CTCF, including cell

adhesion, neuron projection, regulation of synaptic transmission, ion transport/binding, and cell

signalling (Figure S7D). CTCF BS were also found in a considerable percentage of the RAGs in our

datasets (compare Figure S7E with Figure 4E). Microccocal nuclease (MNase) digestion assays for

selected RAGs (Cebpd, 2m, Fos, KLf16, and Bdnf) confirmed increased chromatin accessibility upon

SNA (Figure S7F and G).

Next, we assessed whether CTCF was required for conditioning-dependent regenerative

growth of DRG sensory neurons and for axonal regeneration in vivo. CTCF-floxed and WT mice were

injected with an AAV-CreGFP in the sciatic nerve 4 weeks before a sciatic nerve crush (SNC) injury. As

anticipated, CTCF deletion resulted in a strong decrease in the protein expression (Figure 8A and B)

and in a consistent number of neurons with reduced CTCF expression (84.6+14.2% out of the total

number of infected neurons, N=5). Analysis of DRG neurite outgrowth in WT versus CTCF-cKO GFP-

positive neurons plated 24h after injury showed that the conditioning lesion-dependent increase in

neurite outgrowth was significantly reduced in CTCF-cKO neurons (Figure 8C and D). To investigate

the role of CTCF in vivo, sciatic axonal regeneration was assessed at 1, 3 and 7 days following SNC.

We measured the intensity of Stathmin-like 2 (STMN2 or SCG10), marker of regenerating axons 34, at

different distances from the crush site and generated the “regeneration index”, the distance at

which SCG10 intensity is half that at the crush site, as previously reported 34. Data analysis revealed

that the degree of axonal regeneration was significantly reduced up to 7 days (Figure 8E-M).

However, we found no difference between CTCF-cKO and WT in the number of regenerating

retrogradely traced neurons after injection of cholera toxin subunit B (CTB) in the tibialis anterioris

13

Page 14: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

and gastrocneus muscles at 28 days post-injury (Figure 8N and O). Consistently, no significant

difference between control and CTCF-cKO mice was observed in locomotion or mechanoception

upon SNC (Fig. S8). Together these data indicate that CTCF is required for conditioning-dependent

DRG regenerative growth and that CTCF deletion delays axonal regeneration after SNC.

Validation of the dataset

We investigated the concordance of our dataset with previously published datasets of various

models of sciatic nerve injury at several time points 35-38 by performing a Spearman correlation

analysis. This showed a positive significant correlation between the differentially expression

signatures upon SNA in our dataset and the majority of the tested comparisons (Fig. S9A, first

column, in red, Pvalues indicated). As expected, upon DCA a negative correlation was found in most

of the comparisons (Fig. S9A, third column, in blue). These results were supported by Fisher’s exact

test showing a strong overlap between up- and downregulated genes in our dataset upon SNA and

the up- and downregulated genes in the tested datasets (Fig. S9B, Fisher’s exact test, Pvalue

indicated).

Interestingly, by applying False Discovery Rate (FDR)-corrected Pvalues to our data, which reduced

the number of DE genes, we lost most of the significant correlations with published datasets

(Fig.S9A, second and forth columns). While the overall pattern of the transcriptional response to the

peripheral lesion remained detectable, DE genes following the central lesion were lost (Fig. S9C),

likely due to the less pronounced degree of differential gene expression after DCA. We were able to

validate differential chromatin accessibility by MNase assay (Fig. S7F-G), gene expression by RT-PCR

(Fig. S9D-E), and H3K27ac promoter occupancy by ChIP followed by RT-PCR (Fig. S9F). Finally, we

compared all published experimentally validated DE genes identified by microarrays or RNAseq

(validation includes RT-PCR, Northern blot and Western blot) in various models of sciatic nerve injury

to genes present in our dataset (Table S1). Indeed, out of 51 entries, 45 genes (88.2%) were

validated by our dataset.

14

Page 15: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

DISCUSSION

Our genome-wide combinatorial analysis (Figure S10) a revealed a unique epigenetic

landscape characterized by increased chromatin accessibility and increased acetyl H3K27 and H3K9

occupancy at TSS and body regions of genes related to transcription and regenerative signalling

pathways after SNA but not after DCA. This correlated with prominent or weak changes in gene

expression respectively.

Our data suggests that unique sets of enhancers, including developmentally-regulated

enhancers, respond to axonal injury with changes in H3 acetylation and chromatin accessibility and

may be involved in the activation of the transcriptional programme after peripheral nerve axotomy.

Footprinting analysis of the open chromatin regions showed that the response to axonal

injury is mediated by transcriptional activator recruitment at enhancers and promoters with a more

relaxed chromatin conformation, while only few transcriptional regulators, mainly repressors, are

bound at the compact chromatin regions after central injury. Altogether, this evidence raises the

intriguing hypothesis that a “silent” or poorly dynamic chromatin state might limit axonal

regeneration after injury in the CNS. Importantly, machine learning analysis quantitatively

demonstrated that inclusion of the epigenetic and transcriptional regulatory data is sufficient to

predict the majority of the gene expression changes following axonal injury.

The BaGfoot analysis and the PSO together identified several TF families associated with the

transcriptional response to axonal injury. Several of them have been uncovered previously through

either experimental or computational work and are known to be involved in axonal regeneration or

growth 17, 25, 39. The PSO also disclosed a number of additional novel candidates, such as members of

the ETV family, NANOG, ZBTB7C, and PBX3 upon SNA. A striking feature is the presence of TFs

involved in neuronal differentiation. PBX3 and other family members have been shown to interact

with HOX proteins (identified in the BaGfoot analysis), resulting in higher DNA binding activity and

specificity 40. HOX and NANOG are important regulators of embryonic stem cell and neuronal

15

Page 16: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

differentiation during development including in the DRG 41, 42, remove 41 or 42 suggesting that the

axon regeneration programme shares molecular pathways with developmental processes.

HDAC inhibition via MS275 following DCA led to enhanced H3 acetylation and chromatin

accessibility and to differential occupancy of a repertoire of TFs at DA gene locations. However, this

resulted in the activation of a transcriptional programme that only minimally mimics what observed

following SNA.

These data suggest that SNA and HDAC inhibition modulate largely independent molecular

pathways, suggesting that broad HDAC inhibition might not be the best strategy to mimic

conditioning-dependent regeneration of sensory axons.

The identification of BS for the chromatin organiser CTCF at TSS and enhancers of accessible

regenerative genes upon SNA suggested that chromatin organisation might contribute to axonal

regeneration. Indeed, we found that CTCF is required for conditioning lesion-dependent neurite

outgrowth in cultured DRG neurons, and conditional deletion of CTCF in neurons delays axonal

regeneration in vivo following SNC.

The machine learning algorithm suggests that the epigenetic components allow predicting a

significant part but not the entire injury-dependent transcriptional programme. Our data also show

that differential gene expression is not always associated with histone acetylation at gene promoters

and enhancers or with changes in chromatin accessibility. Therefore, it is likely that additional

mechanisms are involved, such as H3K4me2/3, H3K36me3, H3K56ac, H4K8ac, which mark active

promoters/enhancers and gene bodies 43, DNA modifications, such as DNA methylation and

hydroxymethylation 44, 45, as well as RNA-dependent post-transcriptional regulation 46 that have been

described to be partially involved in the regenerative response.

In conclusion, our study provides a map of the chromatin and transcriptional landscape in

DRG after axonal injury offering a novel platform to investigate gene regulatory mechanisms for the

control of axonal signalling and regeneration.

16

Page 17: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

Acknowledgments

We would like to acknowledge the start-up funds-Division of Brain Sciences, Imperial College London

(SDG), Wings for Life WFL-UK-07/16 (SDG); Rosetrees Trust A1438 (SDG), the Leverhulme Trust RPG-

2015-092 (SDG), ISRT (SDG), the MRC R/R005311/1 (SDG); WT 099276/Z/12/Z (MM). The research

was supported by the National Institute for Health Research (NIHR) Imperial Biomedical Research

Centre (SDG). The views expressed are those of the author(s) and not necessarily those of the NHS,

the NIHR or the Department of Health. We thank Christopher Sibley, Imperial College of London, for

carefully reviewing the mansucript and for useful discussion and suggestions.

Author contributions

I.P. designed and performed experiments, analysed the data, wrote and edited the manuscript.

M.C.D. analysed the data, wrote and edited the manuscript.

T.H. performed experiments, analysed the data and edited the manuscript.

L.Z., and E.S. performed experiments

E.M. performed experiments, analysed the data, and edited the manuscript.

K.S. analysed data and edited the manuscript

P.K.S. edited the manuscript

A.H. performed experiments, and edited the manuscript

N.O. assisted with the data analysis and edited the manuscript.

H.D., M.K. and S.W. performed the PSO analysis and edited the manuscript.

T.L. and W.Z. performed the random forest analysis and edited the manuscript.

M.M, and E.E. provided the CTCFfloxflox mice and discussed data analysis

L.L. maintained the CTCFfloxflox mice colony

J.L.B. and V.P.L. directed and interpreted experiments and edited the manuscript

S.D.G. designed experiments, wrote and edited the manuscript

17

Page 18: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

Competing Interest:

No competing Interests

Figure Legends

Figure 1. ATACseq, ChIPseq and RNAseq datasets

(A) Schematic representation of the experimental design. (B) Heatmaps of the genomic distribution

of ATACseq signal, as Log2(Enrichment vs Background) (left) and H3K9ac, H3K27ac, and H3K27me3

histone mark signal, as Log2(IP vs Input) (right) in the sham control sample. (C) Line plot of the

correlation in the sham sample between chromatin accessibility, histone occupancy of the indicated

histone marks with respect to the TSS (percentile epigenetic signal in the TSS+/-1Kb, mean+s.d. of

N=3 for ATACseq, and N=2 for ChIPseq, independent samples), and the gene expression level

(Log2FPKM, mean+s.d. of N=3 independent samples); genes were divided in quartiles according to

their expression level; Pvalue: 2-way ANOVA, Tukey’s multiple comparison test of the histone mark

and ATAC signal in each quartile compared to the preceding one (ns=not significant) (full statistics in

Supplementary Data 15). (D) Profile of the distribution of ATACseq and the indicated histone mark

signal across the TSS and the gene body; genes were divided in quartiles according to their

expression level.

Figure 2. Peripheral but not central axotomy induces an increased chromatin accessibility and

histone H3 acetylation

(A) Bar plot of the number of genes with differential TSS accessibility upon SNA or DCA (TSS +/- 1Kb;

N=3 independent samples; Pvalue<0.05). (B) Area-proportional Venn diagram of the genes with DA

TSS in SNA and in DCA. (C) Bar chart of the first ranked GO categories of the genes with a gained

accessibility in the TSS upon SNA (BP: Biological Process, Modified Fisher Exact Pvalue<0.01). (D)

Stacked bar plot of the number of genes with a differential occupancy of the indicated histone marks

upon SNA or DCA (TSS +/- 1Kb; N=2 independent samples; Pvalue<0.05). (E) Line plot of the histone

occupancy (peak intensity in the TSS+/-1Kb, mean+s.e.m. of N=2 independent samples) in the genes

18

Page 19: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

with higher and lower accessibility after SNA or DCA; Pvalue: 2-way ANOVA, Tukey’s multiple

comparison test (ns= not significant) (full statistics in Supplementary Data 15).

Figure 3. Peripheral but not central axotomy induces a robust transcriptional response in DRG

(A) Bar plot of the number of DE genes in DRG after SNA or DCA (N=3 independent samples;

Pvalue<0.05). (B) Area-proportional Venn diagram of the commonly regulated genes in SNA and in

DCA. (C) Odds ratio analysis of enrichment of genes associated with a higher, lower, or unchanged

chromatin accessibility at the TSS for DE or non-DE genes; the numbers in red represent the Pvalue

given by two-sided Fisher’s exact test (ns= not significant). (D-E) GSEA of genes with gain (D) or loss

(E) of accessibility at the TSS ranked according to the corrected fold change of gene expression (see

Methods), showing the enrichment of the genes with more accessible promoters among

upregulated genes on the left side of the ranked gene list (x axis of the plot) (Kolmogorov Smirnov

test).

Figure 4. Functional gene categories share a similar epigenetic signature

(A) Heatmaps showing ATACseq and ChIPseq signal density for the up and downregulated genes

upon SNA at equivalent genomic regions spanning from 2Kb upstream of the TSS to 2kb downstream

of the Transcription End Site (TES). (B) Bar plot of the number of upregulated and downregulated

genes (N=3 independent samples; Pvalue<0.05) with increased or decreased histone occupancy (TSS

+/- 1Kb, and/or gene body, N=2 independent samples; Pvalue<0.05) in the different conditions. (C)

Odds ratio analysis of the enrichment of genes with differential chromatin accessibility and H3K9ac,

H3K27ac, and H3K27me3 occupancy across each upregulated and downregulated gene cluster in (B);

the numbers in red represent the Pvalue given by two-sided Fisher’s exact test (blank= not

significant). (D) Heatmap showing the GO enrichment analysis of the upregulated and

downregulated genes with differential enrichment for the indicated histone mark (BP:biological

process, Modified Fisher Exact Pvalue<0.05). (E) Heatmap showing the upregulated RAG, ranked by

fold change; in red the ones displaying an increased histone H3K9 or K27 acetylation at the TSS

and/or gene body level are highlighted.

19

Page 20: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

Figure 5. Specific enhancers respond to axonal injury

(A) Bar plot of the number of enhancers with differential accessibility in DRG after SNA or DCA (N=3

independent samples; Pvalue<0.05). (B) Bar plot of the number of enhancers with differential

occupancy of the indicated histone marks upon SNA or DCA (N=2 independent samples;

Pvalue<0.05). (C) Bar plot of the number of up and downregulated genes (N=3 independent samples;

Pvalue<0.05) controlled by enhancers displaying a change in the occupancy of the indicated histone

mark. (D-E) Stacked bar plots of the number of enhancers with a differential occupancy of the

indicated histone marks or differential accessibility upon SNA (D) or DCA (E) (N=2 for ChIPseq, N=3

for ATACseq, independent samples; Pvalue<0.05) that are activated across the indicated

developmental stages. (F) Odds ratio analyses of the enrichment of genes with differential chromatin

accessibility, H3K27ac and H3K27me3 occupancy across enhancer clusters in (D-E) (two-sided

Fisher’s exact test).

Figure 6. Footprinting analysis identifies specific clusters of transcription factors after axonal injury

(A-D) BaGfoot plots of the ΔFA (differential flaking accessibility) and ΔFPD (differential footprint

depth) values of the motifs upon SNA (A, C) and DCA (B, D) at the level of TSS (TSS+/-1Kb) (A, B) and

enhancers (C, D). The population median is marked in orange; the bag area (dark blue) represents

the region where 50% of the motifs are located; the fence area (light blue) is the region where the

remaining non-outlier motifs are located. The plots are divided in 4 quadrants (I-IV), and the red

squares depict the statistically significant motifs outside the fence (N=3 independent samples;

Pvalue<0.05). (E-F) Examples of motif-centered aggregation plots used to score differential

footprints for two transcription factors.

Figure 7. Random forest classification differentiates between upregulated and downregulated

genes based on TF footprints

(A) Bar plot of the number of DE genes in DRG after SNA or DCA that are associated to TF footprints

(see On line Methods in Supplementary Information) (N=3 independent samples; Pvalue<0.05). (B)

Stacked bar plot of the number of genes, associated to TF footprints, with a differential occupancy of

20

Page 21: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

the indicated histone mark or differential chromatin accessibility upon SNA and DCA (TSS +/- 1Kb;

N=2 for ChIPseq, and N=3 for ATAseq, independent samples; Pvalue<0.05). (C) Receiver Operating

Characteristic (ROC) curves for random forest classification of upregulated vs downregulated genes

after SNA (green line) or DCA (red line). (D-E) Selection frequency of the ten most commonly

selected non-DNA content-derived features for SNA (D) or DCA (E) by the particle swarm

optimization algorithm.

Figure 8. CTCF is required for axonal outgrowth and nerve regeneration

(A) Micrographs showing GFP (green) and CTCF (red) expression in DRG 4 weeks after AAV-Cre-GFP

injection in WT and CTCF-cKO mice. Arrowheads mark GFP positive infected neurons showing

presence or loss of CTCF nuclear staining in WT and CTCF-cKO respectively. Scale bar, 100 m. (B)

Bar graphs of the quantification of CTCF nuclear signal as shown in (A) (mean + s.e.m of N=6 DRGs

from 3 mice; two-sided unpaired Student’s t-test; full statistics in Supplementary Data 15). (C)

Micrographs showing neurite outgrowth in DRG neurons extracted from AAV-Cre-GFP injected WT

and CTCF-cKO mice 24 hours after sham or SNC surgery and plated for 16h. Cre-GFP is showed in

green, III-tubulin in red. Scale bar, 100 m. (D) Bar graphs show quantification of neurite

outgrowth as shown in (C) (mean + s.e.m of N=6 for sham, and N=7 for SNC, mice; 2-way Anova,

Tukey’s multiple comparisons test; full statistics in Supplementary Data 15). (E-L) Representative

micrographs and quantification of SCG10 intensity at the indicated distances from the lesion site in

sciatic nerves from WT and CTCF-cKO mice at 1 (E and F), 3 (G and H) and 7 (I and L) days following

SNC. Asterisk marks the lesion site; scale bar, 1mm. Higher magnification insets display regenerating

axons; scale bar, 05 mm (mean ± s.e.m of N=8 nerves from 4 mice for 1 day; N=6 nerves from 3 mice

for 3 days; N= 6 nerves from 4 wt mice for 7 days; N= 7 nerves from 4 CTCF-cKO mice for 7 days; 2-

way Anova, Sidak’s multiple comparisons test; full statistics in Supplementary Data 15). (M) Bar chart

of the regeneration index in WT and CTCF-cKO mice at 1, 3 and 7 days after SNC (mean + s.e.m of N=

12 nerves from 7 wt mice at 1 day; N=10 nerves from 6 wt mice at 3 days; N=5 nerves from 4 wt

mice at 7 days; N=13 nerves from 7 CTCF-cKO mice at 1 days; N=11 nerves from 6 CTCF-cKO mice at

21

Page 22: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

3 days; N=6 nerves from 4 CTCF-cKO mice at 7 days; 2-way Anova, Sidak’s multiple comparisons test;

full statistics in Supplementary Data 15). (N) Micrographs showing GFP (green) and CTB (red) signal

in DRG 4 days after injection of CTB in the tibialis anterioris and gastrocnemius muscle 28 days after

SNC. Arrowheads mark GFP/CTB double positive neurons. Scale bar, 100 m. (O) Bar graphs of the

CTB positive neurons as percentage of GFP expressing neurons (N) (mean + s.e.m of N=4 mice;

ns=not significant; two-sided unpaired Student’s t-test; full statistics in Supplementary Data 15).

Materials and Methods

Animal procedures

Animal work was carried out in accordance to regulations of the UK Home Office. Wild-type C57Bl6/J

(Harlan) and CTCFflox/flox mice (6-10 weeks old) were used. The total number of mice used in the

entire study was 478 WT and 47 genetically modified mice. Always males were used, with the

exception of behaviour tests and the Figure 8N-O where females were used. Animals were

randomized in all experiments, when appropriate.

For all surgeries mice were anesthetized with isofluorane (3% induction, 2% maintenance) and

buprenorphine (0.1mg/kg)/carprophen (5mg/kg) were administered preoperatively as analgesic.

Dorsal column axotomy (DCA), sciatic nerve axotomy (SNA), and sciatic nerve crush (SNC) were

performed as previously reported 15. Briefly, for DCA, the spinal cord was exposed by a T9

laminectomy (~20mm from sciatic DRGs), dura mater was removed and a dorsal hemisection until

the central canal was performed with fine forceps (FST). In selected experiments, MS275 (Bio-

techne, 12.5 mg/Kg) or vehicle (3.2% Tween80, 20% PEG300, 4% DMSO) was injected

intraperitoneally at the time of the injury and 2 h before sacrifice. For the control laminectomy

(LAM) surgery, the dura mater was removed without performing the hemisection. For SNA, the

sciatic nerve was exposed by blunt dissection of the biceps femoris and the gluteus superficialis and

axotomy was carried out with iridectomy scissors (FST) (~20mm from sciatic DRGs). Alternatively,

sciatic nerve was crushed (SNC) by applying pressure in two perpendicular directions for 30sec with

22

Page 23: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

Dumont #5 Forceps (FST). The crush site was marked by carbon powder. In control mice (sham) the

sciatic nerve was exposed without axotomy or crush.

For nerve injection of AAV-Cre-GFP (Tebu-bio), after nerve exposure, 3 microliters of viral particles

were injected with a Hamilton syringe slowly, and mice were used after 4 weeks to allow complete

Cre expression. In selected experiments, twenty-eight days after SNC, 2.5 l of 10% CTB were

injected in the gastrocnemius and tibialis anterior and sacrificed 4 days afterwards to trace

regenerating neurons in the DRG, as a measure of muscle reinnervation.

RNAseq

RNAseq upon SNA vs Sham and upon DCA vs Lam has been already published 15. Sciatic DRGs (3

biological replicates, 2 mice/condition) were extracted 24h after DCA injury and MS275 treatment,

and collected in RNAlater (Qiagen) to prevent RNA degradation. RNA extraction was performed as

reported 15. Briefly, tissue was crushed in RLT Lysis buffer (Qiagen), and RNA was isolated with

RNeasy mini kit and DNase on column digestion (Qiagen) following manufacturer’s guidelines. RNA

quality was assessed with Agilent 2100 Bioanalyzer (Agilent). RNA with RIN factor above 8.0 was

used for library preparation. Libraries were generated using the NEBNext poly(A) mRNA magnetic

isolation module and NEBNext Ultra II Directional RNA Library Prep Kit for Illumina according to the

recommended manufacturer protocol. Ligation and library integrity was verified using a Tapestation

and Glowmax station. The NEBNext Multiplex Oligos for Illumina (Dual Index Primers Set 1) adapter

sequences were ligated to each sample in each group to allow for sequencing multiplexing. Sample

libraries ligated with unique adapter sequences, were multiplexed, and sequencing was performed

on an Illumina HiSeq4000, generating 75bp pair-ended reads at Imperial BRC Genomic Facility. An

average of 59 million read pairs were generated for each sample, with a total of 353 million read

pairs generated.

ChIPseq

Sciatic DRGs (2 biological replicates, 10 mice/condition) were extracted as above, 24 hours after

injuries. Chromatin immunoprecipitation (ChIP) was performed according to a previously published

23

Page 24: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

protocol 13 (see also Supplementary Note 1). Briefly, DRG pellets were crushed using an automatic

pestle and chemically cross-linked with a 1% formaldehyde solution for 15 minutes at room

temperature. Chromatin was sonicated for 30 min using a Bioruptor (Diagenode) to get 200-800 bp

fragments. Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen)

bound to 10 μg of H3K27ac (Ab4729, Abcam), or H3K9ac (Ab10812, Abcam), or H3K27me3

(C15410195, Diagenode) antibodies. After washing, elution and reverse crosslinking, DNA was

treated with RNAse A and Proteinase K and purified using Qiagen PCR purification columns. Equal

amounts (30 ng) of Input and immunoprecipitated DNA were used for library preparation with the

NEBNext Ultra DNA Library Prep Kit from Illumina (New England BioLabs) following the

manufacturer’s protocol, using a dilution 1:10 of adaptors. Samples were cleaned to remove

unligated adaptors and size selection was performed to select and enrich fragments of 200bp in size.

Libraries were amplified by PCR using multiplex oligo following manufacturer’s protocol. Prior to

sequencing, libraries were run on an Agilent Bioanalyser for size and quality control checking,

quantified using Qbit (ThermoFisher), and multiplexed. Sequencing was performed in an Illumina

HiSeq2000, generating 50bp single ended reads, at the Imperial MRC Genomic Facility. On average,

30 million reads per sample were produced for H3K9ac, 17 million reads per sample for H3K27ac,

and 25 million reads per sample for H3K27me3. This gives a total of 242 million reads for H3K9ac,

132 million reads for H3K27ac, and 200 million reads for H3K27me3. ChiPseq for H3K9ac has been

already published 13.

ATACseq

ATACseq tested the accessibility to Tn5 transposase, which integrates and inserts sequencing

adapters into open chromatin regions (see also Supplementary Note 1). ATACseq was performed

according to a previously published protocol 16. Briefly, sciatic DRGs (3 biological replicates, 1

mouse/condition) were extracted 24 hours after injuries and crushed using an automatic pestle in

Lysis buffer containing 0.1% Igepal (Sigma). Transposition reaction with Nextera Tn5 Transposase

(Illumina, FC-121-1030) was performed on 50000 nuclei. After purification with minElute PCR

24

Page 25: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

purification kit (Qiagen), transposed DNA was amplified by 11 cycle PCR using barcoded primers.

Libraries were run on an Agilent Bioanalyser for size and quality control checking, quantified using

Qbit (ThermoFisher), and multiplexed. Sequencing was performed on an Illumina HiSeq4000,

generating 75bp pair-ended reads at Imperial BRC Genomic Facility. On average, 83 million read

pairs were produced per sample for the ATACseq, yielding a total of approximately 1.5 billion read

pairs.

Sequence alignment, differential expression analysis, and identification of ChIP and ATAC peaks

RNAseq analysis was performed as described 15. For ChIPseq analysis, reads were quality checked,

aligned to the mm10 reference genome, and called for peaks using the AQUAS histone ChIPseq

pipeline (https://github.com/kundajelab/chipseq_pipeline) running BWA and MACS2, using Pvalue

<0.01 and then using the IDR naïve overlapping method as described

(https://github.com/kundajelab/chipseq_pipeline). The peak set produced by the AQUAS pipeline

using overlap analysis was used as the peak set for each condition. Genomic bins of 1000bp

upstream and downstream of each TSS for each gene were created using the same gene annotation

as used for the RNAseq data. Read counts per genomic bin (for gene analysis) or peak (for enhancer

analysis) were obtained from the mapped reads using HTSeq-0.6.1 and subsequently, differential

binding testing was conducted in EdgeR-3.8.6. Additional quality examination of aligned reads was

performed using ChIPQC-1.2.2. The cut-off criteria for the differentially occupied regions was set at

Pvalue<0.05. No false discovery rate correction or additional fold change cutoffs were applied in

determining DE genes or differentially occupied regions. The AQUAS pipeline also produced signal

tracks for each sample which plot the –Log10(Pvalue) of enrichment over input control at each

position. ChIP signal distribution plots and heatmaps were generated using NGSplot version 2.47.1 47.

Briefly, the ChIP signal for each condition is plotted along a gene body region either for all genes in

the annotation or a select subset of genes. The enrichment vs input was calculated using the

GenomicAlignment package in R to measure the proportion of reads within each input or ChIP

sample, and then dividing ChIP by input. The expression and differential expression data from the

25

Page 26: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

RNAseq experiment were used to select the groups of genes from which to sample the ChIP signal in

order to correlate the ChIP signal with changes in the gene’s expression between biological

conditions.

For ATACseq analysis, quality checking, read alignment, signal track generation, and peak calling

were performed using the Kundaje lab’s ATACseq processing pipeline

(https://github.com/kundajelab/atac_dnase_pipelines) running Bowtie2 and MACS2. The peak set

produced by the pipeline using overlap analysis was used as the peak set for each condition.

Genomic bins of 1000bp upstream and downstream of each TSS for each gene were created using

the same gene annotation as used for the RNAseq data. Read counts per genomic bin (for gene

analysis) or peak (for enhancer analysis) were obtained from the mapped reads using HTSeq-0.6.1

and subsequently, differential accessibility testing was conducted in EdgeR-3.8.6. The enrichment vs

background was calculated using the GenomicAlignment package in R to measure the proportion of

reads within each sample and then dividing by the proportion of the genome spanned by that type

of region. Signal distribution plots and heatmaps were generated using NGSplot version 2.47.1.

Validation of RNAseq differential expression signatures

We compared the gene expression signatures of our dataset upon SNA or DCA against published

datasets. Data for GSE30165 38, GSE21007 37, GSE26350 35, and GSE33175 36, comprising various

models of sciatic nerve injury at several time points, was downloaded from Gene Expression

Omnibus (GSE). The data was pre-processed as probe-level expression and normalised across the

arrays. Quality control tests were performed to determine within and between cluster similarity.

Differential expression analysis was performed using limma 48 R package remove this reference, may

be just provide a web address. To compare these datasets to ours, the ranks of all genes based on

logFC * -log10(P-value) was first calculated. The datasets were then compared using pairwise

Spearman correlation on genes that were significantly expressed (Pvalue<0.05 or FDR<0.05).

In addition, we tested the overlaps between genes that were significantly up- or downregulated

(Pvalue<0.05) between experiments to determine whether the overlaps can be considered

26

Page 27: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

significant using Fisher exact test

(https://bioconductor.org/packages/release/bioc/html/GeneOverlap.html).

Transcriptional factor motif enrichment analysis

TF motif enrichment analysis was performed using Analysis of Motif Enrichment (AME) 49 remove

this reference, just provide a web address on the enhancer and TSS (+/-1Kb) regions with

differential chromatin accessibility, and at the level of TSS (+/-1Kb) region of DE genes, using a

Fisher's exact test against the Jaspar core 2016. A shuffled input sequence background was used.

Footprints

ATACseq footprints were generated using HINT 28 for all genomic regions within an ATACseq peak in

that condition. The ATAC footprints setting was used and bias correction was estimated. The

scatterplots showing the change in footprint score vs change in gene expression for the footprinted

TFs was performed using the scripts given in 28 and using the Footprint Score statistic 50 remove this

reference, just provide a web address Differential footprinting analysis was performed using

BaGFoot 23 at the level of promoters and enhancers, separately. The motif database used was

generated by FIMO on the complete mm10 genome with the Jaspar core 2016 motif matrices 51.

remove this reference, just provide a web address. To find TF target genes, the footprints found by

HINT at the level of DA regions were input into GREAT 52 in order to identify the closest genes. The

gene regulatory region was set at -2000/+500 from the TSS. Moreover, if the HINT motifs were inside

enhancer regions (enhancer +/- 05Kb), we retrieved their target genes, identified by ENCODE 21.

Gene Ontology and Pathway analysis

GO and KEGG Pathway analysis were performed using DAVID 6.7 (https://david.ncifcrf.gov/) and

setting all the expressed genes in our dataset as background.

Gene set Enrichment analysis

Pre-ranked GSEA was performed using GSEA software

(http://software.broadinstitute.org/gsea/index.jsp). The DE genes were ranked according to their

combined score “–Log10(Pvalue)*Log2(FC)” (x axis of the GSEA plot). The dataset representing the

27

Page 28: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

genes with a more or less accessible promoter region was uploaded to GSEA (with 1000

permutations) to determine if these genes were significantly enriched on either side of the ranked

gene list.

Statistics and Power analysis

All statistical analyses were performed with GraphPad Prism software. Values are presented as

dot blots with individual data points plus bar, with mean ± s.e.m. or s.d., as indicated in figure

legends. All measurements were taken from distinct samples and the same sample was never

measured repeatedly. Statistical analyses were designed using the assumption of normal

distribution, although that assumption was not explicitly tested. Statistical comparisons included

two-sided paired or unpaired Student’s t-test, 1- or 2-way ANOVA followed by Sidak’, Dunnett's or

Tukey's multiple comparison post hoc tests as specified in the figure legends, and one or two-

sided Fisher’s exact test. P values< 0.05 were considered to be statistically significant. The full

statistics is provided in Supplementary Data 15. Sample size was either based upon similar

previously established experimental designs, or calculated using AEEC power calculator, to estimate

the number of replicates required, for a difference of 1.5 and 80% to 90% power assuming a 5%

significance level.

For RNAseq, ATACseq, and ChIPseq sample size calculation and power estimation were carried out

accordingly to an already described pipeline

(http://andre-rendeiro.com/2018/08/10/atacseq_power_analysis) 53. The results are reported in

Supplementary Data 15.

Replication

All attempt of replication were successful.

Blinding

Nerve regeneration was assessed in blind by two independent scientists. VonFrey test was

assessed by 2 experimenters, and one of them was blind to experimental groups.

28

Page 29: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

Data and code availability

RNAseq, ChIPseq and ATACseq data have been deposited at the Gene Expression Omnibus (GEO)

under accession ID GSE97090, GSE108806, and GSE132382. All code for processing and analysing

the data presented in this work are available upon request. For detailed information on

experimental design please see the provided Life Sciences Reporting Summary. Uncropped blots

with molecular weight standards are provided in the Supplementary Figure 10.

Accession codes

All the accession codes used for this study are: GSE97090, GSE108806, GSE132382, GSE30165,

GSE21007, GSE26350, and GSE33175.

1. Tedeschi, A. Tuning the orchestra: transcriptional pathways controlling axon regeneration. Frontiers in molecular neuroscience 4, 60 (2011).

2. Kiryu-Seo, S. & Kiyama, H. The nuclear events guiding successful nerve regeneration. Frontiers in molecular neuroscience 4, 53 (2011).

3. Patodia, S. & Raivich, G. Role of transcription factors in peripheral nerve regeneration. Frontiers in molecular neuroscience 5, 8 (2012).

4. Ma, T.C. & Willis, D.E. What makes a RAG regeneration associated? Frontiers in molecular neuroscience 8, 43 (2015).

5. Baldwin, K.T. & Giger, R.J. Insights into the physiological role of CNS regeneration inhibitors. Frontiers in molecular neuroscience 8, 23 (2015).

6. Neumann, S. & Woolf, C.J. Regeneration of dorsal column fibers into and beyond the lesion site following adult spinal cord injury. Neuron 23, 83-91 (1999).

7. Luger, K., Dechassa, M.L. & Tremethick, D.J. New insights into nucleosome and chromatin structure: an ordered state or a disordered affair? Nature reviews. Molecular cell biology 13, 436-447 (2012).

8. Geiman, T.M. & Robertson, K.D. Chromatin remodeling, histone modifications, and DNA methylation-how does it all fit together? Journal of cellular biochemistry 87, 117-125 (2002).

9. Vignali, M., Hassan, A.H., Neely, K.E. & Workman, J.L. ATP-dependent chromatin-remodeling complexes. Molecular and cellular biology 20, 1899-1910 (2000).

10. Li, E. Chromatin modification and epigenetic reprogramming in mammalian development. Nature reviews. Genetics 3, 662-673 (2002).

11. Palmisano, I. & Di Giovanni, S. Advances and Limitations of Current Epigenetic Studies Investigating Mammalian Axonal Regeneration. Neurotherapeutics : the journal of the American Society for Experimental NeuroTherapeutics 15, 529-540 (2018).

29

Page 30: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

12. Hutson, T.H. et al. Cbp-dependent histone acetylation mediates axon regeneration induced by environmental enrichment in rodent spinal cord injury models. Science translational medicine 11 (2019).

13. Hervera, A. et al. PP4-dependent HDAC3 dephosphorylation discriminates between axonal regeneration and regenerative failure. The EMBO journal (2019).

14. Oh, Y.M. et al. Epigenetic regulator UHRF1 inactivates REST and growth suppressor gene expression via DNA methylation to promote axon regeneration. Proceedings of the National Academy of Sciences of the United States of America 115, E12417-E12426 (2018).

15. Hervera, A. et al. Reactive oxygen species regulate axonal regeneration through the release of exosomal NADPH oxidase 2 complexes into injured axons. Nature cell biology 20, 307-319 (2018).

16. Buenrostro, J.D., Giresi, P.G., Zaba, L.C., Chang, H.Y. & Greenleaf, W.J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nature methods 10, 1213-1218 (2013).

17. Chandran, V. et al. A Systems-Level Analysis of the Peripheral Nerve Intrinsic Axonal Growth Program. Neuron 89, 956-970 (2016).

18. Tedeschi, A. et al. The Calcium Channel Subunit Alpha2delta2 Suppresses Axon Regeneration in the Adult CNS. Neuron 92, 419-434 (2016).

19. Li, S. et al. The transcriptional landscape of dorsal root ganglia after sciatic nerve transection. Scientific reports 5, 16888 (2015).

20. Shen, Y. et al. A map of the cis-regulatory sequences in the mouse genome. Nature 488, 116-120 (2012).

21. Yue, F. et al. A comparative encyclopedia of DNA elements in the mouse genome. Nature 515, 355-364 (2014).

22. Nord, A.S. et al. Rapid and pervasive changes in genome-wide enhancer usage during mammalian development. Cell 155, 1521-1531 (2013).

23. Baek, S., Goldstein, I. & Hager, G.L. Bivariate Genomic Footprinting Detects Changes in Transcription Factor Activity. Cell reports 19, 1710-1722 (2017).

24. Gaub, P. et al. The histone acetyltransferase p300 promotes intrinsic axonal regeneration. Brain : a journal of neurology 134, 2134-2148 (2011).

25. Nadeau, S., Hein, P., Fernandes, K.J., Peterson, A.C. & Miller, F.D. A transcriptional role for C/EBP beta in the neuronal response to axonal injury. Molecular and cellular neurosciences 29, 525-535 (2005).

26. Danzi, M.C. et al. The effect of Jun dimerization on neurite outgrowth and motif binding. Molecular and cellular neurosciences 92, 114-127 (2018).

27. Finelli, M.J., Wong, J.K. & Zou, H. Epigenetic regulation of sensory axon regeneration after spinal cord injury. The Journal of neuroscience : the official journal of the Society for Neuroscience 33, 19664-19676 (2013).

28. Gusmao, E.G., Allhoff, M., Zenke, M. & Costa, I.G. Analysis of computational footprinting methods for DNase sequencing experiments. Nature methods 13, 303-309 (2016).

29. Usoskin, D. et al. Unbiased classification of sensory neuron types by large-scale single-cell RNA sequencing. Nature neuroscience 18, 145-153 (2015).

30. Hirayama, T., Tarusawa, E., Yoshimura, Y., Galjart, N. & Yagi, T. CTCF is required for neural development and stochastic expression of clustered Pcdh genes in neurons. Cell reports 2, 345-357 (2012).

31. Ong, C.T. & Corces, V.G. CTCF: an architectural protein bridging genome topology and function. Nature reviews. Genetics 15, 234-246 (2014).

32. Ziebarth, J.D., Bhattacharya, A. & Cui, Y. CTCFBSDB 2.0: a database for CTCF-binding sites and genome organization. Nucleic acids research 41, D188-194 (2013).

30

Page 31: T - Imperial College London · Web view2019/07/19  · Immunoprecipitation was performed overnight with Protein G Dynabeads (Invitrogen) bound to 10 μg of H3K27ac (Ab4729, Abcam),

33. Sams, D.S. et al. Neuronal CTCF Is Necessary for Basal and Experience-Dependent Gene Regulation, Memory Formation, and Genomic Structure of BDNF and Arc. Cell reports 17, 2418-2430 (2016).

34. Shin, J.E., Geisler, S. & DiAntonio, A. Dynamic regulation of SCG10 in regenerating axons after injury. Experimental neurology 252, 1-11 (2014).

35. Michaelevski, I. et al. Signaling to transcription networks in the neuronal retrograde injury response. Science signaling 3, ra53 (2010).

36. Wang, Y. et al. Gene network revealed involvements of Birc2, Birc3 and Tnfrsf1a in anti-apoptosis of injured peripheral nerves. PloS one 7, e43436 (2012).

37. Geeven, G. et al. LLM3D: a log-linear modeling-based method to predict functional gene regulatory interactions from genome-wide expression data. Nucleic acids research 39, 5313-5327 (2011).

38. Li, S. et al. Differential gene expression profiling and biological process analysis in proximal nerve segments after sciatic nerve transection. PloS one 8, e57000 (2013).

39. Wang, Z., Reynolds, A., Kirry, A., Nienhaus, C. & Blackmore, M.G. Overexpression of Sox11 promotes corticospinal tract regeneration after spinal injury while interfering with functional recovery. The Journal of neuroscience : the official journal of the Society for Neuroscience 35, 3139-3145 (2015).

40. Mann, R.S. & Affolter, M. Hox proteins meet more partners. Current opinion in genetics & development 8, 423-429 (1998).

41. Philippidou, P. & Dasen, J.S. Hox genes: choreographers in neural development, architects of circuit organization. Neuron 80, 12-34 (2013).

42. De Kumar, B. et al. Dynamic regulation of Nanog and stem cell-signaling pathways by Hoxa1 during early neuro-ectodermal differentiation of ES cells. Proceedings of the National Academy of Sciences of the United States of America 114, 5838-5845 (2017).

43. Zhou, V.W., Goren, A. & Bernstein, B.E. Charting histone modifications and the functional organization of mammalian genomes. Nature reviews. Genetics 12, 7-18 (2011).

44. Loh, Y.E. et al. Comprehensive mapping of 5-hydroxymethylcytosine epigenetic dynamics in axon regeneration. Epigenetics 12, 77-92 (2017).

45. Weng, Y.L. et al. An Intrinsic Epigenetic Barrier for Functional Axon Regeneration. Neuron 94, 337-346 e336 (2017).

46. Zhou, S., Ding, F. & Gu, X. Non-coding RNAs as Emerging Regulators of Neural Injury Responses and Regeneration. Neuroscience bulletin 32, 253-264 (2016).

47. Shen, L., Shao, N., Liu, X. & Nestler, E. ngs.plot: Quick mining and visualization of next-generation sequencing data by integrating genomic databases. BMC genomics 15, 284 (2014).

48. Ritchie, M.E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic acids research 43, e47 (2015).

49. McLeay, R.C. & Bailey, T.L. Motif Enrichment Analysis: a unified framework and an evaluation on ChIP data. BMC bioinformatics 11, 165 (2010).

50. He, H.H. et al. Refined DNase-seq protocol and data analysis reveals intrinsic bias in transcription factor footprint identification. Nature methods 11, 73-78 (2014).

51. Grant, C.E., Bailey, T.L. & Noble, W.S. FIMO: scanning for occurrences of a given motif. Bioinformatics 27, 1017-1018 (2011).

52. McLean, C.Y. et al. GREAT improves functional interpretation of cis-regulatory regions. Nature biotechnology 28, 495-501 (2010).

53. Hart, S.N., Therneau, T.M., Zhang, Y., Poland, G.A. & Kocher, J.P. Calculating sample size estimates for RNA sequencing data. Journal of computational biology : a journal of computational molecular cell biology 20, 970-978 (2013).

31