Postmortem Cortex Samples Identify Distinct Molecular Subtypes of ALS: Retrotransposon Activation, Oxidative Stress, and Activated Glia

Oliver H Tam, Nikolay V Rozhkov, Regina Shaw, Duyang Kim, Isabel Hubbard, Samantha Fennessey, Nadia Propp, NYGC ALS Consortium, Delphine Fagegaltier, Brent T Harris, Lyle W Ostrow, Hemali Phatnani, John Ravits, Josh Dubnau, Molly Gale Hammell, Hemali Phatnani, Justin Kwan, Dhruv Sareen, James R Broach, Zachary Simmons, Ximena Arcila-Londono, Edward B Lee, Vivianna M Van Deerlin, Neil A Shneider, Ernest Fraenkel, Lyle W Ostrow, Frank Baas, Noah Zaitlen, James D Berry, Andrea Malaspina, Pietro Fratta, Gregory A Cox, Leslie M Thompson, Steve Finkbeiner, Efthimios Dardiotis, Timothy M Miller, Siddharthan Chandran, Suvankar Pal, Eran Hornstein, Daniel J MacGowan, Terry Heiman-Patterson, Molly G Hammell, Nikolaos A Patsopoulos, Oleg Butovsky, Joshua Dubnau, Avindra Nath, Robert Bowser, Matt Harms, Eleonora Aronica, Mary Poss, Jennifer Phillips-Cremins, John Crary, Nazem Atassi, Dale J Lange, Darius J Adams, Leonidas Stefanis, Marc Gotkine, Robert Baloh, Suma Babu, Towfique Raj, Sabrina Paganoni, Ophir Shalem, Colin Smith, Bin Zhang, Brent T Harris, Oliver H Tam, Nikolay V Rozhkov, Regina Shaw, Duyang Kim, Isabel Hubbard, Samantha Fennessey, Nadia Propp, NYGC ALS Consortium, Delphine Fagegaltier, Brent T Harris, Lyle W Ostrow, Hemali Phatnani, John Ravits, Josh Dubnau, Molly Gale Hammell, Hemali Phatnani, Justin Kwan, Dhruv Sareen, James R Broach, Zachary Simmons, Ximena Arcila-Londono, Edward B Lee, Vivianna M Van Deerlin, Neil A Shneider, Ernest Fraenkel, Lyle W Ostrow, Frank Baas, Noah Zaitlen, James D Berry, Andrea Malaspina, Pietro Fratta, Gregory A Cox, Leslie M Thompson, Steve Finkbeiner, Efthimios Dardiotis, Timothy M Miller, Siddharthan Chandran, Suvankar Pal, Eran Hornstein, Daniel J MacGowan, Terry Heiman-Patterson, Molly G Hammell, Nikolaos A Patsopoulos, Oleg Butovsky, Joshua Dubnau, Avindra Nath, Robert Bowser, Matt Harms, Eleonora Aronica, Mary Poss, Jennifer Phillips-Cremins, John Crary, Nazem Atassi, Dale J Lange, Darius J Adams, Leonidas Stefanis, Marc Gotkine, Robert Baloh, Suma Babu, Towfique Raj, Sabrina Paganoni, Ophir Shalem, Colin Smith, Bin Zhang, Brent T Harris

Abstract

Amyotrophic lateral sclerosis (ALS) is a fatal neurodegenerative disease characterized by the progressive loss of motor neurons. While several pathogenic mutations have been identified, the vast majority of ALS cases have no family history of disease. Thus, for most ALS cases, the disease may be a product of multiple pathways contributing to varying degrees in each patient. Using machine learning algorithms, we stratify the transcriptomes of 148 ALS postmortem cortex samples into three distinct molecular subtypes. The largest cluster, identified in 61% of patient samples, displays hallmarks of oxidative and proteotoxic stress. Another 19% of the samples shows predominant signatures of glial activation. Finally, a third group (20%) exhibits high levels of retrotransposon expression and signatures of TARDBP/TDP-43 dysfunction. We further demonstrate that TDP-43 (1) directly binds a subset of retrotransposon transcripts and contributes to their silencing in vitro, and (2) pathological TDP-43 aggregation correlates with retrotransposon de-silencing in vivo.

Keywords: TDP-43; amyotrophic lateral sclerosis; genetics and genomics of ALS; neurodegeneration; neurodegenerative disease; retrotransposons; transposable elements.

Conflict of interest statement

DECLARATION OF INTERESTS

The authors declare no competing interests.

Copyright © 2019 The Author(s). Published by Elsevier Inc. All rights reserved.

Figures

Figure 1.. ALS Patients in the NYGC…
Figure 1.. ALS Patients in the NYGC ALS Consortium Cohort Show 3 Distinct Subtypes: Retrotransposon Reactivated (ALS-TE), Oxidative Stress Signatures (ALS-Ox), and Glial Dysfunction (ALS-Glial)
(A) A schematic depiction of cortical regions from which post-mortem tissue samples were obtained: 148 from ALS patients, 11 from patients with other neurological disorders, and 17 from subjects. Multiple samples were obtained for most patients, such that these samples represent 77 ALS patients, 6 neurological control subjects, and 12 healthy controls. (B) A heatmap of gene expression in ALS patient samples showing the results of NMF-based clustering, which returned 3 robust and distinct groups of patients. The retrotransposon reactivated set showed high TE expression as a dominant signature and formed its own subgroup (ALS-TE), representing 20% of all patient samples. Two additional groups were identified that showed signatures of oxidative stress (ALS-Ox, 61% of samples), and glial dysfunction (ALS-Glial, 19% of samples). Representative gene sets that describe markers from each subset are identified along the left of the heatmap. (C) Gene Set Enrichment Analysis was performed for each group of patients (ALS-TE, ALS-Ox, and ALS-Glial) versus controls to identify group-specific aberrant pathways. (D) Violin plots of example markers for each group show the LINE retrotransposon L1PA6 marking the ALS-TE group, SOD1 marking the ALS-Ox group, and TREM2 marking the ALS-Glial samples.
Figure 2.. ALS-Ox Patients in the NYGC…
Figure 2.. ALS-Ox Patients in the NYGC ALS Consortium Show Markers of Oxidative Stress
(A) The ALS-Ox group represents the middle 61% of patient samples in the ALS Consortium set (outlined in blue), with ALS-Ox marker genes listed along the left-hand side of the heatmap. (B) Violin plots of example markers for this group show two known ALS-associated genes, MATR3 and NEFL, specifically elevated in the ALS-Ox group (blue violins). (C) Top enriched pathways identified by GSEA in the ALS-Ox group samples compared to controls. (D) Additional ALS patient samples from UCSD are assigned to the NYGC ALS subgroups by position on this PCA plot (UCSD samples marked by patient ID in blue). (E) To confirm the link between ALS-Ox membership and stress pathways, individual PCA plots are shown where each sample is colored by the normalized expression of genes in the Oxidative Stress pathway (OXR1, TXN), ER-linked proteotoxic stress pathways (UBQLN2, BECN1 and downstream genes related to autophagy (ATG5, TBK1).
Figure 3.. ALS-Glia Patients in the NYGC…
Figure 3.. ALS-Glia Patients in the NYGC ALS Consortium Cohort Show Evidence of Disease-Associated Microglia
(A) The ALS-Glia group represents the 19% of patient samples on the right side of the ALS Consortium set (outlined in gold), with additional ALS-Glia marker genes listed along the left side of the heatmap. (B) Violin plots of example markers for this group show two known glial marker genes, AIF1/IBA1 and CD44, specifically elevated in the ALS-Glia group (gold violins). (C) Top enriched pathways identified by GSEA in the ALS-Glia group samples compared to controls. (D) Additional ALS patient samples from UCSD are assigned to the NYGC ALS subgroups by position on this PCA plot (UCSD samples marked by patient ID in gold). (E) To confirm the link between ALS-Glia membership and markers of particular glial subsets, individual PCA plots are shown where each sample is colored by the normalized expression of genes that mark Microglia (IBA1, TREM2), Oligodendrocytes (OLIG1, MOG), and astrocytes (GFAP, CD44). (F) To confirm the link between ALS-Glia membership and activated microglia, UCSD patient samples were stained with antibodies to IBA1. ALS-Glia samples show presence of enlarged, activated microglia not seen in controls or other ALS subgroups. Images are shown at 20 × magnification, with labeled scale bars indicating a size of 50–70μm.
Figure 4.. ALS-TE Patients in the NYGC…
Figure 4.. ALS-TE Patients in the NYGC ALS Consortium Cohort Show Retrotransposon Re-activation
(A) The ALS-TE group represents the left hand 20% of patients in the ALS Consortium sample set (outlined in Red), with ALS-TE markers listed along the left-hand panel. (B) Violin plots of example markers for this group show two relatively young retrotransposons, L1HS and SVA_A, and TARDBP expression. (C) Top enriched pathways identified by GSEA in the ALS-TE versus controls. (D) Additional ALS samples from UCSD are assigned to the defined subgroups by position on this PCA plot (UCSD samples marked by patient ID in red). (E) To confirm the link between ALS-TE membership and TDP-43 pathology, UCSD patient samples were stained with antibodies to the phosphorylated form of TDP-43 (pTDP-43, left) as well as with antibodies that recognize full-length TDP-43 protein (right). ALS-TE samples show evidence of pTDP-43 pathology not present in controls or other ALS subgroups. Images are shown at 20 × magnification, with labeled scale bars indicating a size of 50μm.
Figure 5.. TDP-43 Binds and Regulates Retrotransposon…
Figure 5.. TDP-43 Binds and Regulates Retrotransposon Transcripts
(A) A schematic of the eCLIP-seq based identification (2 biological replicates) of directly bound RNA targets for TDP-43, with a representative pileup of reads where TDP-43 is known to target its own 3′ UTR. (B) A pie chart of annotation categories for TDP-43 targets shows that retrotransposons are a small but important fraction of direct targets (12.9%). Gene intronic sequences represent the largest fraction of all peaks at 44%, while an additional 17.6% represent regions where TEs provide regulatory sequences to the host gene intron. (C) The known TDP-43 binding motif (UGUGU) is present in both gene and retrotransposon targets. (D) Representative browser pileup plots for one retrotransposon from each major repeat class: an L1PA6 LINE element, an AluY SINE element, a HERV3 LTR element, and an SVA_D from the human specific SVA class. (E) Knock down of TDP-43 levels (2 biological replicates) results in the upregulation of retrotransposon transcripts, most of which also show evidence for direct binding in the eCLIP-seq experiments (gray bars). (F) A volcano plot of fold change versus false discovery rate (FDR) q values shows that upregulated TEs (red dots) form a substantial fraction of all TDP-43 dependent genes, and all are upregulated. (G) Knock down efficiency of short hairpin RNAs (shRNAs) targeting TARDBP was validated by western blot analysis and qPCR. For RNA levels, replicates were averaged and SD shown as error bars. For protein levels, western blots of the two biological replicates are shown separately.

References

    1. Antony JM, van Marle G, Opii W, Butterfield DA, Mallet F, Yong VW, Wallace JL, Deacon RM, Warren K, and Power C (2004). Human endogenous retrovirus glycoprotein-mediated induction of redox reactants causes oligodendrocyte death and demyelination. Nat. Neurosci 7, 1088–1095.
    1. Arai T, Hasegawa M, Akiyama H, Ikeda K, Nonaka T, Mori H, Mann D, Tsuchiya K, Yoshida M, Hashizume Y, and Oda T (2006). TDP-43 is a component of ubiquitin-positive tau-negative inclusions in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Biochem. Biophys. Res. Commun 351, 602–611.
    1. Attig J, Agostini F, Gooding C, Chakrabarti AM, Singh A, Haberman N, Zagalak JA, Emmett W, Smith CWJ, Luscombe NM, and Ule J (2018). Heteromeric RNP Assembly at LINEs Controls Lineage-Specific RNA Processing. Cell 174, 1067–1081.
    1. Ayala YM, De Conti L, Avendaño-Vázquez SE, Dhir A, Romano M, D’Ambrogio A, Tollervey J, Ule J, Baralle M, Buratti E, and Baralle FE (2011). TDP-43 regulates its mRNA levels through a negative feedback loop. EMBO J. 30, 277–288.
    1. Bankhead P, Loughrey MB, Fernández JA, Dombrowski Y, McArt DG, Dunne PD, McQuaid S, Gray RT, Murray LJ, Coleman HG, et al. (2017). QuPath: Open source software for digital pathology image analysis. Sci. Rep 7, 16878.
    1. Bao W, Kojima KK, and Kohany O (2015). Repbase Update, a database of repetitive elements in eukaryotic genomes. Mob. DNA 6, 11.
    1. Bruijn LI, Houseweart MK, Kato S, Anderson KL, Anderson SD, Ohama E, Reaume AG, Scott RW, and Cleveland DW (1998). Aggregation and motor neuron toxicity of an ALS-linked SOD1 mutant independent from wild-type SOD1. Science 281, 1851–1854.
    1. Chang Y-H, and Dubnau J (2019). The Gypsy Endogenous Retrovirus Drives Non-Cell-Autonomous Propagation in a Drosophila TDP-43 Model of Neurodegeneration. Curr. Biol, S0960-9822(19)30951-0.
    1. Chia R, Chiò A, and Traynor BJ (2018). Novel genes associated with amyotrophic lateral sclerosis: diagnostic and clinical implications. Lancet Neurol. 17, 94–102.
    1. Chiu IM, Morimoto ETA, Goodarzi H, Liao JT, O’Keeffe S, Phatnani HP, Muratet M, Carroll MC, Levy S, Tavazoie S, et al. (2013). A neurodegeneration-specific gene-expression signature of acutely isolated microglia from an amyotrophic lateral sclerosis mouse model. Cell Rep. 4, 385–401.
    1. Cohen TJ, Lee VM-Y, and Trojanowski JQ (2011). TDP-43 functions and pathogenic mechanisms implicated in TDP-43 proteinopathies. Trends Mol. Med 17, 659–667.
    1. Colombrita C, Onesto E, Megiorni F, Pizzuti A, Baralle FE, Buratti E, Silani V, and Ratti A (2012). TDP-43 and FUS RNA-binding proteins bind distinct sets of cytoplasmic messenger RNAs and differently regulate their post-transcriptional fate in motoneuron-like cells. J. Biol. Chem 287, 15635–15647.
    1. Crow YJ, and Manel N (2015). Aicardi-Goutières syndrome and the type I interferonopathies. Nat. Rev. Immunol 15, 429–440.
    1. Deczkowska A, Keren-Shaul H, Weiner A, Colonna M, Schwartz M, and Amit I (2018). Disease-Associated Microglia: A Universal Immune Sensor of Neurodegeneration. Cell 173, 1073–1081.
    1. Di Giorgio FP, Carrasco MA, Siao MC, Maniatis T, and Eggan K (2007). Non-cell autonomous effect of glia on motor neurons in an embryonic stem cell-based ALS model. Nat. Neurosci 10, 608–614.
    1. Dobin A, Davis CA, Schlesinger F, Drenkow J, Zaleski C, Jha S, Batut P, Chaisson M, and Gingeras TR (2013). STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21.
    1. Evrony GD, Lee E, Park PJ, and Walsh CA (2016). Resolving rates of mutation in the brain using single-neuron genomics. eLife 5, 56.
    1. Guo C, Jeong H-H, Hsieh Y-C, Klein H-U, Bennett DA, De Jager PL, Liu Z, and Shulman JM (2018). Tau Activates Transposable Elements in Alzheimer’s Disease. Cell Rep. 23, 2874–2880.
    1. Ho Y-J, Anaparthy N, Molik D, Mathew G, Aicher T, Patel A, Hicks J, and Hammell MG (2018). Single-cell RNA-seq analysis identifies markers of resistance to targeted BRAF inhibitors in melanoma cell populations. Genome Res. 28, 1353–1363.
    1. Jin Y, Tam OH, Paniagua E, and Hammell M (2015). TEtranscripts: a package for including transposable elements in differential expression analysis of RNA-seq datasets. Bioinformatics 31, 3593–3599.
    1. Kanehisa M, and Goto S (2000). KEGG: kyoto encyclopedia of genes and genomes. Nucleic Acids Res. 28, 27–30.
    1. Kelley DR, Hendrickson DG, Tenen D, and Rinn JL (2014). Transposable elements modulate human RNA abundance and splicing via specific RNA-protein interactions. Genome Biol. 15, 537.
    1. Keren-Shaul H, Spinrad A, Weiner A, Matcovitch-Natan O, Dvir-Szternfeld R, Ulland TK, David E, Baruch K, Lara-Astaiso D, Toth B, et al. (2017). A Unique Microglia Type Associated with Restricting Development of Alzheimer’s Disease. Cell 169, 1276–1290.
    1. Kremer D, Schichel T, Förster M, Tzekova N, Bernard C, van der Valk P, van Horssen J, Hartung H-P, Perron H, and Küry P (2013). Human endogenous retrovirus type W envelope protein inhibits oligodendroglial precursor cell differentiation. Ann. Neurol 74, 721–732.
    1. Krug L, Chatterjee N, Borges-Monroy R, Hearn S, Liao W-W, Morrill K, Prazak L, Rozhkov N, Theodorou D, Hammell M, and Dubnau J (2017). Retrotransposon activation contributes to neurodegeneration in a Drosophila TDP-43 model of ALS. PLoS Genet. 13, e1006635.
    1. Li W, Jin Y, Prazak L, Hammell M, and Dubnau J (2012). Transposable elements in TDP-43-mediated neurodegenerative disorders. PLoS ONE 7, e44099.
    1. Li W, Lee MH, Henderson L, Tyagi R, Bachani M, Steiner J, Campa-nac E, Hoffman DA, von Geldern G, Johnson K, et al. (2015). Human endogenous retrovirus-K contributes to motor neuron disease. Sci. Transl. Med 7, 307ra153.
    1. Liberzon A, Birger C, Thorvaldsdóttir H, Ghandi M, Mesirov JP, and Tamayo P (2015). The Molecular Signatures Database (MSigDB) hallmark gene set collection. Cell Syst. 1, 417–425.
    1. Liu EY, Russ J, Cali CP, Phan JM, Amlie-Wolf A, and Lee EB (2019). Loss of Nuclear TDP-43 Is Associated with Decondensation of LINE Retrotransposons. Cell Rep. 27, 1409–1421.
    1. Love MI, Huber W, and Anders S (2014). Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550.
    1. Macia A, Widmann TJ, Heras SR, Ayllon V, Sanchez L, Benkaddour-Boumzaouad M, Muñoz-Lopez M, Rubio A, Amador-Cubero S, Blanco-Jimenez E, et al. (2017). Engineered LINE-1 retrotransposition in nondividing human neurons. Genome Res. 27, 335–348.
    1. Majounie E, Renton AE, Mok K, Dopper EGP, Waite A, Rollinson S, Chiò A, Restagno G, Nicolaou N, Simon-Sanchez J, et al.; Chromosome 9-ALS/FTD Consortium; French research network on FTLD/FTLD/ALS; ITALSGEN Consortium (2012). Frequency of the C9orf72 hexanucleotide repeat expansion in patients with amyotrophic lateral sclerosis and frontotemporal dementia: a cross-sectional study. Lancet Neurol. 11, 323–330.
    1. Mancarci BO, Toker L, Tripathy SJ, Li B, Rocco B, Sibille E, and Pavlidis P (2017). Cross-Laboratory Analysis of Brain Cell Type Transcriptomes with Applications to Interpretation of Bulk Tissue Data. eNeuro 4, ENEURO.0212–17.2017.
    1. Mayer J, Harz C, Sanchez L, Pereira GC, Maldener E, Heras SR, Ostrow LW, Ravits J, Batra R, Meese E, et al. (2018). Transcriptional profiling of HERV-K(HML-2) in amyotrophic lateral sclerosis and potential implications for expression of HML-2 proteins. Mol. Neurodegener 13, 39.
    1. Mills RE, Bennett EA, Iskow RC, and Devine SE (2007). Which transposable elements are active in the human genome? Trends Genet. 23, 183–191.
    1. Nagai M, Re DB, Nagata T, Chalazonitis A, Jessell TM, Wichterle H, and Przedborski S (2007). Astrocytes expressing ALS-linked mutated SOD1 release factors selectively toxic to motor neurons. Nat. Neurosci 10, 615–622.
    1. Neumann M, Sampathu DM, Kwong LK, Truax AC, Micsenyi MC, Chou TT, Bruce J, Schuck T, Grossman M, Clark CM, et al. (2006). Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science 314, 130–133.
    1. Neumann M, Kwong LK, Lee EB, Kremmer E, Flatley A, Xu Y, Forman MS, Troost D, Kretzschmar HA, Trojanowski JQ, and Lee VM-Y (2009). Phosphorylation of S409/410 of TDP-43 is a consistent feature in all sporadic and familial forms of TDP-43 proteinopathies. Acta Neuropathol. 117, 137–149.
    1. Nicolas A, Kenna KP, Renton AE, Ticozzi N, Faghri F, Chia R, Dominov JA, Kenna BJ, Nalls MA, Keagle P, et al.; ITALSGEN Consortium; Genomic Translation for ALS Care (GTAC) Consortium; ALS Sequencing Consortium; NYGC ALS Consortium; Answer ALS Foundation; Clinical Research in ALS and Related Disorders for Therapeutic Development (CReATe) Consortium; SLAGEN Consortium; French ALS Consortium; Project MinE ALS Sequencing Consortium (2018). Genome-wide Analyses Identify KIF5A as a Novel ALS Gene. Neuron 97, 1268–1283.
    1. Pereira GC, Sanchez L, Schaughency PM, Rubio-Roldán A, Choi JA, Planet E, Batra R, Turelli P, Trono D, Ostrow LW, et al. (2018). Properties of LINE-1 proteins and repeat element expression in the context of amyotrophic lateral sclerosis. Mob. DNA 9, 35.
    1. Polymenidou M, Lagier-Tourenne C, Hutt KR, Huelga SC, Moran J, Liang TY, Ling S-C, Sun E, Wancewicz E, Mazur C, et al. (2011). Long pre-mRNA depletion and RNA missplicing contribute to neuronal vulnerability from loss of TDP-43. Nat. Neurosci 14, 459–468.
    1. Prudencio M, Gonzales PK, Cook CN, Gendron TF, Daughrity LM, Song Y, Ebbert MTW, van Blitterswijk M, Zhang Y-J, Jansen-West K, et al. (2017). Repetitive element transcripts are elevated in the brain of C9orf72 ALS/FTLD patients. Hum. Mol. Genet 26, 3421–3431.
    1. Renton AE, Majounie E, Waite A, Simón-Sánchez J, Rollinson S, Gibbs JR, Schymick JC, Laaksovirta H, van Swieten JC, Myllykangas L, et al.; ITALSGEN Consortium (2011). A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron 72, 257–268.
    1. Robinson JT, Thorvaldsdóttir H, Winckler W, Guttman M, Lander ES, Getz G, and Mesirov JP (2011). Integrative genomics viewer. Nat. Bio-technol 29, 24–26.
    1. Rosen DR, Siddique T, Patterson D, Figlewicz DA, Sapp P, Hentati A, Donaldson D, Goto J, O’Regan JP, Deng HX, et al. (1993). Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature 362, 59–62.
    1. Rudnick ND, Griffey CJ, Guarnieri P, Gerbino V, Wang X, Piersaint JA, Tapia JC, Rich MM, and Maniatis T (2017). Distinct roles for motor neuron autophagy early and late in the SOD1 G93A mouse model of ALS. Proc. Natl. Acad. Sci. USA 114, E8294–E8303.
    1. Saldi TK, Ash PE, Wilson G, Gonzales P, Garrido-Lecca A, Roberts CM, Dostal V, Gendron TF, Stein LD, Blumenthal T, et al. (2014). TDP-1, the Caenorhabditis elegans ortholog of TDP-43, limits the accumulation of double-stranded RNA. EMBO J. 33, 2947–2966.
    1. Spandidos A, Wang X, Wang H, and Seed B (2010). PrimerBank: a resource of human and mouse PCR primer pairs for gene expression detection and quantification. Nucleic Acids Res. 33, D792–D799.
    1. Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, Paulovich A, Pomeroy SL, Golub TR, Lander ES, and Mesirov JP (2005). Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl. Acad. Sci. USA 102, 15545–15550.
    1. Sun W, Samimi H, Gamez M, Zare H, and Frost B (2018). Pathogenic tau-induced piRNA depletion promotes neuronal death through transposable element dysregulation in neurodegenerative tauopathies. Nat. Neurosci 21, 1038–1048.
    1. Taylor JP, Brown RH Jr., and Cleveland DW (2016). Decoding ALS: from genes to mechanism. Nature 539, 197–206.
    1. Tollervey JR, Curk T, Rogelj B, Briese M, Cereda M, Kayikci M, König J, Hortobágyi T, Nishimura AL, Zupunski V, et al. (2011). Characterizing the RNA targets and position-dependent splicing regulation by TDP-43. Nat. Neurosci 14, 452–458.
    1. Upton KR, Gerhardt DJ, Jesuadian JS, Richardson SR, Sánchez-Luque FJ, Bodea GO, Ewing AD, Salvador-Palomeque C, van der Knaap MS, Brennan PM, et al. (2015). Ubiquitous L1 mosaicism in hippocampal neurons. Cell 161, 228–239.
    1. Van Nostrand EL, Pratt GA, Shishkin AA, Gelboin-Burkhart C, Fang MY, Sundararaman B, Blue SM, Nguyen TB, Surka C, Elkins K, et al. (2016). Robust transcriptome-wide discovery of RNA-binding protein binding sites with enhanced CLIP (eCLIP). Nat. Methods 13, 508–514.
    1. van Rheenen W, Shatunov A, Dekker AM, McLaughlin RL, Diekstra FP, Pulit SL, van der Spek RAA, Võsa U, de Jong S, Robinson MR, et al.; PARALS Registry; SLALOM Group; SLAP Registry; FALS Sequencing Consortium; SLAGEN Consortium; NNIPPS Study Group (2016). Genome-wide association analyses identify new risk variants and the genetic architecture of amyotrophic lateral sclerosis. Nat. Genet 48, 1043–1048.
    1. Wong YC, and Holzbaur ELF (2015). Autophagosome dynamics in neurodegeneration at a glance. J. Cell Sci 128, 1259–1267.
    1. Xiao S, Sanelli T, Dib S, Sheps D, Findlater J, Bilbao J, Keith J, Zinman L, Rogaeva E, and Robertson J (2011). RNA targets of TDP-43 identified by UV-CLIP are deregulated in ALS. Mol. Cell. Neurosci 47, 167–180.
    1. Zarnack K, König J, Tajnik M, Martincorena I, Eustermann S, Stévant I, Reyes A, Anders S, Luscombe NM, and Ule J (2013). Direct competition between hnRNP C and U2AF65 protects the transcriptome from the exonization of Alu elements. Cell 152, 453–466.
    1. Zhang Z, and Xing Y (2017). CLIP-seq analysis of multi-mapped reads discovers novel functional RNA regulatory sites in the human transcriptome. Nucleic Acids Res. 45, 9260–9271.
    1. Zhang Y-J, Guo L, Gonzales PK, Gendron TF, Wu Y, Jansen-West K, O’Raw AD, Pickles SR, Prudencio M, Carlomagno Y, et al. (2019). Heterochromatin anomalies and double-stranded RNA accumulation underlie C9orf72 poly(PR) toxicity. Science 363, eaav2606.

Source: PubMed

3
Abonneren