2,045
Views
3
CrossRef citations to date
0
Altmetric
Review

Novel therapeutic targets for amyotrophic lateral sclerosis: ribonucleoproteins and cellular autonomy

& ORCID Icon
Pages 971-984 | Received 05 May 2020, Accepted 02 Aug 2020, Published online: 30 Aug 2020

References

  • Balendra R, Patani R. Quo vadis motor neuron disease? World J Methodol. 2016;6:56–64.
  • Pandya VA, Patani R. Decoding the relationship between ageing and amyotrophic lateral sclerosis: a cellular perspective. Brain. 2019. Internet. DOI:10.1093/brain/awz360.
  • Taylor JP, Brown RH Jr, Cleveland DW. Decoding ALS: from genes to mechanism. Nature. 2016;539:197–206.
  • Sreedharan J, Blair IP, Tripathi VB, et al. TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis. Science. 2008;319:1668–1672.
  • Vance C, Rogelj B, Hortobágyi T, et al. Mutations in FUS, an RNA processing protein, cause familial amyotrophic lateral sclerosis type 6. Science. 2009;323:1208–1211.
  • Renton AE, Majounie E, Waite A, et al. A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron. 2011;72:257–268.
  • Mann JR, Gleixner AM, Mauna JC, et al. RNA binding antagonizes neurotoxic phase transitions of TDP-43. Neuron. 2019;102:321–338.e8.
  • Gasset-Rosa F, Lu S, Yu H, et al. Cytoplasmic TDP-43 de-mixing independent of stress granules drives inhibition of nuclear import, loss of nuclear TDP-43, and cell death. Neuron. 2019;102:339–357.e7.
  • Nagai M, Re DB, Nagata T, et al. Astrocytes expressing ALS-linked mutated SOD1 release factors selectively toxic to motor neurons [Internet]. Nat Neurosci. 2007;10:615–622.
  • Yamanaka K, Chun SJ, Boillee S, et al. Astrocytes as determinants of disease progression in inherited amyotrophic lateral sclerosis. Nat Neurosci. 2008;11:251–253.
  • Serio A, Patani R. Concise review: the cellular conspiracy of amyotrophic lateral sclerosis. Stem Cells. 2018;36:293–303.
  • Tyzack G, Lakatos A, Patani R. Human stem cell-derived astrocytes: specification and relevance for neurological disorders. Curr Stem Cell Rep. 2016;2:236–247.
  • Tognatta R, Miller RH. Contribution of the oligodendrocyte lineage to CNS repair and neurodegenerative pathologies. Neuropharmacology. 2016;110:539–547.
  • Ling JP, Pletnikova O, Troncoso JC, et al. TDP-43 repression of nonconserved cryptic exons is compromised in ALS-FTD. Science. 2015;349:650–655.
  • Humphrey J, Emmett W, Fratta P, et al. Quantitative analysis of cryptic splicing associated with TDP-43 depletion. BMC Med Genomics. 2017;10:38.
  • Klim JR, Williams LA, Limone F, et al. ALS-implicated protein TDP-43 sustains levels of STMN2, a mediator of motor neuron growth and repair. Nat Neurosci. 2019;22:167–179.
  • Melamed ZE, López-Erauskin J, Baughn MW, et al. Premature polyadenylation-mediated loss of stathmin-2 is a hallmark of TDP-43-dependent neurodegeneration. Nat Neurosci. 2019;22:180–190.
  • Luisier R, Tyzack GE, Hall CE, et al. Intron retention and nuclear loss of SFPQ are molecular hallmarks of ALS. Nat Commun. 2018;9:2010.
  • Yap K, Lim ZQ, Khandelia P, et al. Coordinated regulation of neuronal mRNA steady-state levels through developmentally controlled intron retention. Genes Dev. 2012;26:1209–1223.
  • Wong JJ-L, Ritchie W, Ebner OA, et al. Orchestrated intron retention regulates normal granulocyte differentiation. Cell. 2013;154:583–595.
  • Buckley PT, Lee MT, Sul J-Y, et al. Cytoplasmic intron sequence-retaining transcripts can be dendritically targeted via ID element retrotransposons. Neuron. 2011;69:877–884.
  • Braunschweig U, Barbosa-Morais NL, Pan Q, et al. Widespread intron retention in mammals functionally tunes transcriptomes. Genome Res. 2014;24:1774–1786.
  • Wadman M. Updated: FDA approves drug that rescues babies with fatal neurodegenerative disease [Internet]. Science. 2016. DOI:10.1126/science.aal0476
  • Finkel RS, Chiriboga CA, Vajsar J, et al. Treatment of infantile-onset spinal muscular atrophy with nusinersen: a phase 2, open-label, dose-escalation study [Internet]. Lancet. 2016;388:3017–3026.
  • Finkel RS, Mercuri E, Darras BT, et al. Nusinersen versus sham control in infantile-onset spinal muscular atrophy. N Engl J Med. 2017;377:1723–1732.
  • Kinane TB, Bernard Kinane T, Mayer OH, et al. Long-term pulmonary function in duchenne muscular dystrophy: comparison of eteplirsen-treated patients to natural history [Internet]. J Neuromuscul Dis. 2018;5:47–58.
  • Miller TM, Pestronk A, David W, et al. An antisense oligonucleotide against SOD1 delivered intrathecally for patients with SOD1 familial amyotrophic lateral sclerosis: a phase 1, randomised, first-in-man study. Lancet Neurol. 2013;12:435–442.
  • Jiang J, Zhu Q, Gendron TF, et al. Gain of toxicity from ALS/FTD-linked repeat expansions in C9ORF72 is alleviated by antisense oligonucleotides targeting GGGGCC-containing RNAs. Neuron. 2016;90:535–550.
  • Neumann M, Sampathu DM, Kwong LK, et al. Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science. 2006;314:130–133.
  • Mackenzie IRA, Bigio EH, Ince PG, et al. Pathological TDP-43 distinguishes sporadic amyotrophic lateral sclerosis from amyotrophic lateral sclerosis with SOD1 mutations. Ann Neurol. 2007;61:427–434.
  • Tyzack GE, Luisier R, Taha DM, et al. Widespread FUS mislocalization is a molecular hallmark of amyotrophic lateral sclerosis. Brain. 2019;142:2572–2580.
  • Chiò A, Restagno G, Brunetti M, et al. Two Italian kindreds with familial amyotrophic lateral sclerosis due to FUS mutation. Neurobiol Aging. 2009;30:1272–1275.
  • Bosco DA, Lemay N, Ko HK, et al. Mutant FUS proteins that cause amyotrophic lateral sclerosis incorporate into stress granules. Hum Mol Genet. 2010;19:4160–4175.
  • Dormann D, Rodde R, Edbauer D, et al. ALS-associated fused in sarcoma (FUS) mutations disrupt Transportin-mediated nuclear import. Embo J. 2010;29:2841–2857.
  • DeJesus-Hernandez M, Kocerha J, Finch N, et al. De novo truncating FUS gene mutation as a cause of sporadic amyotrophic lateral sclerosis. Hum Mutat. 2010;31:E1377–E1389.
  • Savas JN, Toyama BH, Xu T, et al. Extremely long-lived nuclear pore proteins in the rat brain. Science. 2012;335:942.
  • D’Angelo MA, Raices M, Panowski SH, et al. Age-dependent deterioration of nuclear pore complexes causes a loss of nuclear integrity in postmitotic cells. Cell. 2009;136:284–295.
  • Mertens J, Paquola ACM, Ku M, et al. Directly reprogrammed human neurons retain aging-associated transcriptomic signatures and reveal age-related nucleocytoplasmic defects. Cell Stem Cell. 2015;17:705–718.
  • Lu T, Pan Y, Kao S-Y, et al. Gene regulation and DNA damage in the ageing human brain [Internet]. Nature. 2004;429:883–891.
  • Kim HJ, Taylor JP. Lost in transportation: nucleocytoplasmic transport defects in ALS and other neurodegenerative diseases. Neuron. 2017;96:285–297.
  • Taylor RC, Dillin A. Aging as an event of proteostasis collapse. Cold Spring Harb Perspect Biol. [Internet. 2011; 3. DOI:10.1101/cshperspect.a004440.
  • Woerner AC, Frottin F, Hornburg D, et al. Cytoplasmic protein aggregates interfere with nucleocytoplasmic transport of protein and RNA. Science. 2016;351:173–176.
  • van Blitterswijk M, Rademakers R. Neurodegenerative disease: c9orf72 repeats compromise nucleocytoplasmic transport. Nat Rev Neurol. 2015;11:670–672.
  • Vatsavayai SC, Yoon SJ, Gardner RC, et al. Timing and significance of pathological features in C9orf72 expansion-associated frontotemporal dementia. Brain. 2016;139:3202–3216.
  • Proudfoot M, Gutowski NJ, Edbauer D, et al. Early dipeptide repeat pathology in a frontotemporal dementia kindred with C9ORF72 mutation and intellectual disability [Internet]. Acta Neuropathol. 2014;127:451–458.
  • Baborie A, Griffiths TD, Jaros E, et al. Accumulation of dipeptide repeat proteins predates that of TDP-43 in frontotemporal lobar degeneration associated with hexanucleotide repeat expansions in C9ORF72 gene. Neuropathol Appl Neurobiol. 2015;41:601–612.
  • Fujii S, Takanashi K, Kitajo K, et al. Treatment with a global methyltransferase inhibitor induces the intranuclear aggregation of ALS-linked FUS mutant in vitro. Neurochem Res. 2016;41:826–835.
  • Guo L, Kim HJ, Wang H, et al. Nuclear-import receptors reverse aberrant phase transitions of RNA-binding proteins with prion-like domains. Cell. 2018;173:677–692.e20.
  • Yoshizawa T, Ali R, Jiou J, et al. Nuclear import receptor inhibits phase separation of FUS through binding to multiple sites. Cell. 2018;173:693–705.e22.
  • Hofweber M, Hutten S, Bourgeois B, et al. Phase separation of FUS is suppressed by its nuclear import receptor and arginine methylation. Cell. 2018;173:706–719.e13.
  • Qamar S, Wang G, Randle SJ, et al. FUS phase separation is modulated by a molecular chaperone and methylation of arginine cation-π interactions. Cell. 2018;173:720–734.e15.
  • Chou -C-C, Zhang Y, Umoh ME, et al. TDP-43 pathology disrupts nuclear pore complexes and nucleocytoplasmic transport in ALS/FTD. Nat Neurosci. 2018;21:228–239.
  • Archbold HC, Jackson KL, Arora A, et al. TDP43 nuclear export and neurodegeneration in models of amyotrophic lateral sclerosis and frontotemporal dementia. Sci Rep. 2018;8:4606.
  • Ederle H, Funk C, Abou-Ajram C, et al. Nuclear egress of TDP-43 and FUS occurs independently of Exportin-1/CRM1. Sci Rep. 2018;8:7084.
  • Pinarbasi ES, Cağatay T, Fung HYJ, et al. Active nuclear import and passive nuclear export are the primary determinants of TDP-43 localization. Sci Rep. 2018;8:7083.
  • Furukawa Y, Kaneko K, Watanabe S, et al. A seeding reaction recapitulates intracellular formation of Sarkosyl-insoluble transactivation response element (TAR) DNA-binding protein-43 inclusions. J Biol Chem. 2011;286:18664–18672.
  • Smethurst P, Newcombe J, Troakes C, et al. In vitro prion-like behaviour of TDP-43 in ALS. Neurobiol Dis. 2016;96:236–247.
  • Nomura T, Watanabe S, Kaneko K, et al. Intranuclear aggregation of mutant FUS/TLS as a molecular pathomechanism of amyotrophic lateral sclerosis. J Biol Chem. 2014;289:1192–1202.
  • Chia R, Tattum MH, Jones S, et al. Superoxide dismutase 1 and tgSOD1 mouse spinal cord seed fibrils, suggesting a propagative cell death mechanism in amyotrophic lateral sclerosis. PLoS One. 2010;5:e10627.
  • Smethurst P, Risse E, Tyzack GE, et al. Distinct responses of neurons and astrocytes to TDP-43 proteinopathy in amyotrophic lateral sclerosis. Brain. 2020;143:430–440.
  • Kim HJ, Kim NC, Wang Y-D, et al. Mutations in prion-like domains in hnRNPA2B1 and hnRNPA1 cause multisystem proteinopathy and ALS. Nature. 2013;495:467–473.
  • Sun Z, Diaz Z, Fang X, et al. Molecular determinants and genetic modifiers of aggregation and toxicity for the ALS disease protein FUS/TLS. PLoS Biol. 2011;9:e1000614.
  • Johnson BS, Snead D, Lee JJ, et al. TDP-43 is intrinsically aggregation-prone, and amyotrophic lateral sclerosis-linked mutations accelerate aggregation and increase toxicity. J Biol Chem. 2009;284:20329–20339.
  • Couthouis J, Hart MP, Shorter J, et al. A yeast functional screen predicts new candidate ALS disease genes. Proc Natl Acad Sci U S A. 2011;108:20881–20890.
  • Couthouis J, Hart MP, Erion R, et al. Evaluating the role of the FUS/TLS-related gene EWSR1 in amyotrophic lateral sclerosis. Hum Mol Genet. 2012;21:2899–2911.
  • Molliex A, Temirov J, Lee J, et al. Phase separation by low complexity domains promotes stress granule assembly and drives pathological fibrillization. Cell. 2015;163:123–133.
  • Patel A, Lee HO, Jawerth L, et al. A liquid-to-solid phase transition of the ALS protein FUS accelerated by disease mutation. Cell. 2015;162:1066–1077.
  • Mackenzie IR, Nicholson AM, Sarkar M, et al. TIA1 mutations in amyotrophic lateral sclerosis and frontotemporal dementia promote phase separation and alter stress granule dynamics. Neuron. 2017;95:808–816.e9.
  • Chuang E, Hori AM, Hesketh CD, et al. Amyloid assembly and disassembly. J Cell Sci. Internet 2018;131: jcs189928. .
  • Zhang P, Fan B, Yang P, et al. Chronic optogenetic induction of stress granules is cytotoxic and reveals the evolution of ALS-FTD pathology. Elife. 2019;8. Internet. DOI:10.7554/eLife.39578.
  • Gomes E, Shorter J. The molecular language of membraneless organelles. J Biol Chem. 2019;294:7115–7127.
  • Li YR, King OD, Shorter J, et al. Stress granules as crucibles of ALS pathogenesis. J Cell Biol. 2013;201:361–372.
  • Ramaswami M, Taylor JP, Parker R. Altered ribostasis: RNA-protein granules in degenerative disorders. Cell. 2013;154:727–736.
  • Kieran D, Kalmar B, Dick JRT, et al. Treatment with arimoclomol, a coinducer of heat shock proteins, delays disease progression in ALS mice. Nat Med. 2004;10:402–405.
  • Kalmar B, Lu C-H, Greensmith L. The role of heat shock proteins in amyotrophic lateral sclerosis: the therapeutic potential of arimoclomol. Pharmacol Ther. 2014;141:40–54.
  • Wang P, Wander CM, Yuan C-X, et al. Acetylation-induced TDP-43 pathology is suppressed by an HSF1-dependent chaperone program. Nat Commun. 2017;8:82.
  • Chen H-J, Mitchell JC, Novoselov S, et al. The heat shock response plays an important role in TDP-43 clearance: evidence for dysfunction in amyotrophic lateral sclerosis. Brain. 2016;139:1417–1432.
  • Benatar M, Wuu J, Andersen PM, et al. Randomized, double-blind, placebo-controlled trial of arimoclomol in rapidly progressive SOD1 ALS [Internet]. Neurology. 2018;90:e565–e574.
  • Jackrel ME, DeSantis ME, Martinez BA, et al. Potentiated Hsp104 variants antagonize diverse proteotoxic misfolding events. Cell. 2014;156:170–182.
  • Jackrel ME, Shorter J. Potentiated Hsp104 variants suppress toxicity of diverse neurodegenerative disease-linked proteins. Dis Model Mech. 2014;7:1175–1184.
  • Yasuda K, Clatterbuck-Soper SF, Jackrel ME, et al. FUS inclusions disrupt RNA localization by sequestering kinesin-1 and inhibiting microtubule detyrosination. J Cell Biol. 2017;216:1015–1034.
  • Shorter J. Engineering therapeutic protein disaggregases. Mol Biol Cell. 2016;27:1556–1560.
  • Jackrel ME, Shorter J. Engineering enhanced protein disaggregases for neurodegenerative disease. Prion. 2015;9:90–109.
  • Mack KL, Shorter J. Engineering and evolution of molecular chaperones and protein disaggregases with enhanced activity [Internet]. Front Mol Biosci. 2016;3. DOI:10.3389/fmolb.2016.00008
  • Jackrel ME, Shorter J. Reversing deleterious protein aggregation with re-engineered protein disaggregases. Cell Cycle. 2014;13:1379–1383.
  • Shorter J. Designer protein disaggregases to counter neurodegenerative disease [Internet]. Curr Opin Genet Dev. 2017;44:1–8.
  • Wippich F, Bodenmiller B, Trajkovska MG, et al. Dual specificity kinase DYRK3 couples stress granule condensation/dissolution to mTORC1 signaling. Cell. 2013;152:791–805.
  • Rai AK, Chen J-X, Selbach M, et al. Kinase-controlled phase transition of membraneless organelles in mitosis. Nature. 2018;559:211–216.
  • Reineke LC, Tsai W-C, Jain A, et al. Casein kinase 2 is linked to stress granule dynamics through phosphorylation of the stress granule nucleating protein G3BP1. Mol Cell Biol. 2017;37. Internet. DOI:10.1128/MCB.00596-16.
  • Nonaka T, Masuda-Suzukake M, Arai T, et al. Prion-like properties of pathological TDP-43 aggregates from diseased brains. Cell Rep. 2013;4:124–134.
  • Bulawa CE, Connelly S, Devit M, et al. Tafamidis, a potent and selective transthyretin kinetic stabilizer that inhibits the amyloid cascade. Proc Natl Acad Sci U S A. 2012;109:9629–9634.
  • Prasad A, Raju G, Sivalingam V, et al. An acridine derivative, [4,5-bis(N-carboxy methyl imidazolium)methylacridine] dibromide, shows anti-TDP-43 aggregation effect in ALS disease models [Internet]. Sci Rep. 2016;6. DOI:10.1038/srep39490.
  • Boyd JD, Lee P, Feiler MS, et al. A high-content screen identifies novel compounds that inhibit stress-induced TDP-43 cellular aggregation and associated cytotoxicity. J Biomol Screen. 2014;19:44–56.
  • Cassel JA, McDonnell ME, Velvadapu V, et al. Characterization of a series of 4-aminoquinolines that stimulate caspase-7 mediated cleavage of TDP-43 and inhibit its function. Biochimie. 2012;94:1974–1981.
  • Parker SJ, Meyerowitz J, James JL, et al. Inhibition of TDP-43 accumulation by bis(thiosemicarbazonato)-copper complexes. PLoS One. 2012;7:e42277.
  • Becker LA, Huang B, Bieri G, et al. Therapeutic reduction of ataxin-2 extends lifespan and reduces pathology in TDP-43 mice. Nature. 2017;544:367–371.
  • Elden AC, Kim H-J, Hart MP, et al. Ataxin-2 intermediate-length polyglutamine expansions are associated with increased risk for ALS. Nature. 2010;466:1069–1075.
  • McGurk L, Gomes E, Guo L, et al. Poly(ADP-Ribose) prevents pathological phase separation of TDP-43 by promoting liquid demixing and stress granule localization. Mol Cell. 2018;71:703–717.e9.
  • Maharana S, Wang J, Papadopoulos DK, et al. RNA buffers the phase separation behavior of prion-like RNA binding proteins. Science. 2018;360:918–921.
  • Huang Y-C, Lin K-F, He R-Y, et al. Inhibition of TDP-43 aggregation by nucleic acid binding. PLoS One. 2013;8:e64002.
  • Ishiguro T, Sato N, Ueyama M, et al. Regulatory role of RNA chaperone TDP-43 for RNA misfolding and repeat-associated translation in SCA31. Neuron. 2017;94:108–124.e7.
  • Flores BN, Li X, Malik AM, et al. An intramolecular salt bridge linking TDP43 RNA binding, protein stability, and TDP43-dependent neurodegeneration. Cell Rep. 2019;27:1133–1150.e8.
  • Barmada SJ, Serio A, Arjun A, et al. Autophagy induction enhances TDP43 turnover and survival in neuronal ALS models. Nat Chem Biol. 2014;10:677–685.
  • Rusmini P, Cristofani R, Galbiati M, et al. The role of the heat shock protein B8 (HSPB8) in motoneuron diseases. Front Mol Neurosci. 2017;10:176.
  • Mandrioli J, Crippa V, Cereda C, et al. Proteostasis and ALS: protocol for a phase II, randomised, double-blind, placebo-controlled, multicentre clinical trial for colchicine in ALS (Co-ALS). BMJ Open. 2019;9:e028486.
  • Towers CG, Thorburn A. Therapeutic targeting of autophagy [Internet]. EBioMedicine. 2016;14:15–23.
  • Mandrioli J, D’Amico R, Zucchi E, et al. Rapamycin treatment for amyotrophic lateral sclerosis: protocol for a phase II randomized, double-blind, placebo-controlled, multicenter, clinical trial (RAP-ALS trial). Medicine (Baltimore). 2018;97:e11119.
  • Wang I-F, Guo B-S, Liu Y-C, et al. Autophagy activators rescue and alleviate pathogenesis of a mouse model with proteinopathies of the TAR DNA-binding protein 43. Proc Natl Acad Sci U S A. 2012;109:15024–15029.
  • Fox JH, Connor T, Chopra V, et al. The mTOR kinase inhibitor Everolimus decreases S6 kinase phosphorylation but fails to reduce mutant huntingtin levels in brain and is not neuroprotective in the R6/2 mouse model of Huntington’s disease. Mol Neurodegener. 2010;5:26.
  • Tsvetkov AS, Miller J, Arrasate M, et al. A small-molecule scaffold induces autophagy in primary neurons and protects against toxicity in a Huntington disease model. Proc Natl Acad Sci U S A. 2010;107:16982–16987.
  • Fontaine SN, Zheng D, Sabbagh JJ, et al. DnaJ/Hsc70 chaperone complexes control the extracellular release of neurodegenerative-associated proteins. Embo J. 2016;35:1537–1549.
  • Holmes BB, DeVos SL, Kfoury N, et al. Heparan sulfate proteoglycans mediate internalization and propagation of specific proteopathic seeds. Proc Natl Acad Sci U S A. 2013;110:E3138–E3147.
  • Urushitani M, Ezzi SA, Julien J-P. Therapeutic effects of immunization with mutant superoxide dismutase in mice models of amyotrophic lateral sclerosis. Proc Natl Acad Sci U S A. 2007;104:2495–2500.
  • Takeuchi S, Fujiwara N, Ido A, et al. Induction of protective immunity by vaccination with wild-type apo superoxide dismutase 1 in mutant SOD1 transgenic mice [Internet]. J Neuropathol Exp Neurol. 2010;69:1044–1056.
  • Gros-Louis F, Soucy G, Larivière R, et al. Intracerebroventricular infusion of monoclonal antibody or its derived Fab fragment against misfolded forms of SOD1 mutant delays mortality in a mouse model of ALS. J Neurochem. 2010;113:1188–1199.
  • Mehta D, Jackson R, Paul G, et al. Why do trials for Alzheimer’s disease drugs keep failing? A discontinued drug perspective for 2010-2015 [Internet]. Expert Opin Investig Drugs. 2017;26(6):735–739.
  • Huang L-K, Chao S-P, Hu C-J. Clinical trials of new drugs for Alzheimer disease. J Biomed Sci. 2020;27:18.
  • Hall CE, Yao Z, Choi M, et al. Progressive motor neuron pathology and the role of astrocytes in a human stem cell model of VCP-related ALS. Cell Rep. 2017;19:1739–1749.
  • Tyzack GE, Hall CE, Sibley CR, et al. A neuroprotective astrocyte state is induced by neuronal signal EphB1 but fails in ALS models. Nat Commun. 2017;8:1164.
  • Serio A, Bilican B, Barmada SJ, et al. Astrocyte pathology and the absence of non-cell autonomy in an induced pluripotent stem cell model of TDP-43 proteinopathy. Proc Natl Acad Sci U S A. 2013;110:4697–4702.
  • Giorgio FPD, Di Giorgio FP, Carrasco MA, et al. Non–cell autonomous effect of glia on motor neurons in an embryonic stem cell–based ALS model [Internet]. Nat Neurosci. 2007;608–614. DOI:10.1038/nn1885.
  • Giorgio FPD, Di Giorgio FP, Boulting GL, et al. Human embryonic stem cell-derived motor neurons are sensitive to the toxic effect of glial cells carrying an ALS-causing mutation [Internet]. Cell Stem Cell. 2008;3:637–648.
  • Liddelow SA, Guttenplan KA, Clarke LE, et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature. 2017;541:481–487.
  • Liu J, Wang F. Role of neuroinflammation in amyotrophic lateral sclerosis: cellular mechanisms and therapeutic implications [Internet]. Front Immunol. 2017. DOI:10.3389/fimmu.2017.01005
  • Xu Z, Lee A, Nouwens A, et al. Mass spectrometry analysis of plasma from amyotrophic lateral sclerosis and control subjects. Amyotroph Lateral Scler Frontotemporal Degener. 2018;19:362–376.
  • Clarke LE, Liddelow SA, Chakraborty C, et al. Normal aging induces A1-like astrocyte reactivity. Proc Natl Acad Sci U S A. 2018;115:E1896–E1905.
  • Colombo E, Farina C. Astrocytes: key regulators of neuroinflammation. Trends Immunol. 2016;37(9):608–620.
  • Kong Q, Chang L-C, Takahashi K, et al. Small-molecule activator of glutamate transporter EAAT2 translation provides neuroprotection. J Clin Invest. 2014;124:1255–1267.
  • Ganel R, Ho T, Maragakis NJ, et al. Selective up-regulation of the glial Na+-dependent glutamate transporter GLT1 by a neuroimmunophilin ligand results in neuroprotection. Neurobiol Dis. 2006;21:556–567.
  • Fischer A, Sananbenesi F, Mungenast A, et al. Targeting the correct HDAC(s) to treat cognitive disorders. Trends Pharmacol Sci. 2010;31:605–617.
  • Lapucci A, Cavone L, Buonvicino D, et al. Effect of Class II HDAC inhibition on glutamate transporter expression and survival in SOD1-ALS mice. Neurosci Lett. 2017;656:120–125.
  • Cudkowicz ME, Titus S, Kearney M, et al. Safety and efficacy of ceftriaxone for amyotrophic lateral sclerosis: a multi-stage, randomised, double-blind, placebo-controlled trial. Lancet Neurol. 2014;13:1083–1091.
  • Rothstein JD, Patel S, Regan MR, et al. Beta-lactam antibiotics offer neuroprotection by increasing glutamate transporter expression. Nature. 2005;433:73–77.
  • Tanaka K, Kanno T, Yanagisawa Y, et al. Bromocriptine methylate suppresses glial inflammation and moderates disease progression in a mouse model of amyotrophic lateral sclerosis [Internet]. Exp Neurol. 2011;232:41–52.
  • Rodríguez-Cueto C, Santos-García I, García-Toscano L, et al. Neuroprotective effects of the cannabigerol quinone derivative VCE-003.2 in SOD1 transgenic mice, an experimental model of amyotrophic lateral sclerosis. Biochem Pharmacol. 2018;157:217–226.
  • Chiarotto GB, Cartarozzi LP, Perez M, et al. Tempol improves neuroinflammation and delays motor dysfunction in a mouse model (SOD1G93A) of ALS [Internet]. J Neuroinflammation. 2019;16. DOI:10.1186/s12974-019-1598-x.
  • Yun SP, Kam T-I, Panicker N, et al. Block of A1 astrocyte conversion by microglia is neuroprotective in models of Parkinson’s disease [Internet]. Nat Med. 2018;24:931–938.
  • Vallarola A, Sironi F, Tortarolo M, et al. RNS60 exerts therapeutic effects in the SOD1 ALS mouse model through protective glia and peripheral nerve rescue. J Neuroinflammation. 2018;15:65.
  • Paganoni S, Alshikho MJ, Luppino S, et al. A pilot trial of RNS60 in amyotrophic lateral sclerosis. Muscle Nerve. 2019;59:303–308.
  • Mazzini L, Gelati M, Profico DC, et al. Human neural stem cell transplantation in ALS: initial results from a phase I trial. J Transl Med. 2015;13:17.
  • Glass JD, Boulis NM, Johe K, et al. Lumbar intraspinal injection of neural stem cells in patients with amyotrophic lateral sclerosis: results of a phase I trial in 12 patients. Stem Cells. 2012;30:1144–1151.
  • Feldman EL, Boulis NM, Hur J, et al. Intraspinal neural stem cell transplantation in amyotrophic lateral sclerosis: phase 1 trial outcomes. Ann Neurol. 2014;75:363–373.
  • Glass JD, Hertzberg VS, Boulis NM, et al. Transplantation of spinal cord-derived neural stem cells for ALS: analysis of phase 1 and 2 trials. Neurology. 2016;87:392–400.
  • Suzuki M, McHugh J, Tork C, et al. GDNF secreting human neural progenitor cells protect dying motor neurons, but not their projection to muscle, in a rat model of familial ALS. PLoS One. 2007;2:e689.
  • Nichols NL, Gowing G, Satriotomo I, et al. Intermittent hypoxia and stem cell implants preserve breathing capacity in a rodent model of amyotrophic lateral sclerosis [Internet]. Am J Respir Crit Care Med. 2013;187:535–542.
  • Lepore AC, O’Donnell J, Kim AS, et al. Human glial-restricted progenitor transplantation into cervical spinal cord of the SOD1 mouse model of ALS. PLoS One. 2011;6:e25968.
  • Kondo T, Funayama M, Tsukita K, et al. Focal transplantation of human iPSC-derived glial-rich neural progenitors improves lifespan of ALS mice. Stem Cell Reports. 2014;3:242–249.
  • Sareen D, Gowing G, Sahabian A, et al. Human induced pluripotent stem cells are a novel source of neural progenitor cells (iNPCs) that migrate and integrate in the rodent spinal cord. J Comp Neurol. 2014;522:2707–2728.
  • Corcia P, Tauber C, Vercoullie J, et al. Molecular imaging of microglial activation in amyotrophic lateral sclerosis. PLoS One. 2012;7:e52941.
  • Turner MR, Cagnin A, Turkheimer FE, et al. Evidence of widespread cerebral microglial activation in amyotrophic lateral sclerosis: an [11C](R)-PK11195 positron emission tomography study. Neurobiol Dis. 2004;15:601–609.
  • Kawamata T, Akiyama H, Yamada T, et al. Immunologic reactions in amyotrophic lateral sclerosis brain and spinal cord tissue. Am J Pathol. 1992;140:691–707.
  • Engelhardt JI, Appel SH. IgG reactivity in the spinal cord and motor cortex in amyotrophic lateral sclerosis. Arch Neurol. 1990;47:1210–1216.
  • Brettschneider J, Toledo JB, Van Deerlin VM, et al. Microglial activation correlates with disease progression and upper motor neuron clinical symptoms in amyotrophic lateral sclerosis. PLoS One. 2012;7:e39216.
  • Chiu IM, Morimoto ETA, Goodarzi H, et al. A neurodegeneration-specific gene-expression signature of acutely isolated microglia from an amyotrophic lateral sclerosis mouse model. Cell Rep. 2013;4:385–401.
  • Liao B, Zhao W, Beers DR, et al. Transformation from a neuroprotective to a neurotoxic microglial phenotype in a mouse model of ALS [Internet]. Exp Neurol. 2012;237:147–152.
  • Ohgomori T, Yamada J, Takeuchi H, et al. Comparative morphometric analysis of microglia in the spinal cord of SOD1G93Atransgenic mouse model of amyotrophic lateral sclerosis [Internet]. Eur J Neurosci. 2016;43:1340–1351.
  • Gravel M, Béland L-C, Soucy G, et al. IL-10 controls early microglial phenotypes and disease onset in ALS caused by misfolded superoxide dismutase 1. J Neurosci. 2016;36:1031–1048.
  • Beers DR, Henkel JS, Zhao W, et al. Endogenous regulatory T lymphocytes ameliorate amyotrophic lateral sclerosis in mice and correlate with disease progression in patients with amyotrophic lateral sclerosis. Brain. 2011;134:1293–1314.
  • Moisse K, Strong MJ. Innate immunity in amyotrophic lateral sclerosis. Biochim Biophys Acta. 2006;1762:1083–1093.
  • Krasemann S, Madore C, Cialic R, et al. The TREM2-APOE pathway drives the transcriptional phenotype of dysfunctional microglia in neurodegenerative diseases. Immunity. 2017;47:566–581.e9.
  • Boillée S, Yamanaka K, Lobsiger CS, et al. Onset and progression in inherited ALS determined by motor neurons and microglia. Science. 2006;312:1389–1392.
  • Beers DR, Henkel JS, Xiao Q, et al. Wild-type microglia extend survival in PU.1 knockout mice with familial amyotrophic lateral sclerosis. Proc Natl Acad Sci U S A. 2006;103:16021–16026.
  • Spiller KJ, Restrepo CR, Khan T, et al. Microglia-mediated recovery from ALS-relevant motor neuron degeneration in a mouse model of TDP-43 proteinopathy [Internet]. Nat Neurosci. 2018;21:329–340.
  • Frakes AE, Ferraiuolo L, Haidet-Phillips AM, et al. Microglia induce motor neuron death via the classical NF-κB pathway in amyotrophic lateral sclerosis. Neuron. 2014;81:1009–1023.
  • Hu H, Lin H, Duan W. Intrathecal Injection of scAAV9–hIGF1 prolongs the survival of ALS model mice by inhibiting the NF-kB pathway. Neuroscience. 2018;381:1–10.
  • Dong H, Zhang W, Zeng X, et al. Histamine induces upregulated expression of histamine receptors and increases release of inflammatory mediators from microglia. Mol Neurobiol. 2014;49:1487–1500.
  • Apolloni S, Fabbrizio P, Parisi C, et al. Clemastine confers neuroprotection and induces an anti-inflammatory phenotype in SOD1G93A mouse model of amyotrophic lateral sclerosis [Internet]. Mol Neurobiol. 2016;53:518–531.
  • Granucci EJ, Griciuc A, Mueller KA, et al. Cromolyn sodium delays disease onset and is neuroprotective in the SOD1 mouse model of amyotrophic lateral sclerosis. Sci Rep. 2019;9:17728.
  • Meer F, Molawi K, Zychlinsky A. Mutant superoxide dismutase 1-induced IL-1beta accelerates ALS pathogenesis. Proc Natl Acad Sci U S A. 2010;541:13046–13050.
  • Zhou Q, Mareljic N, Michaelsen M, et al. Active poly-GA vaccination prevents microglia activation and motor deficits in a C9orf72 mouse model. EMBO Mol Med. 2020;12:e10919.
  • Martínez-Muriana A, Mancuso R, Francos-Quijorna I, et al. CSF1R blockade slows the progression of amyotrophic lateral sclerosis by reducing microgliosis and invasion of macrophages into peripheral nerves. Sci Rep. 2016;6:25663.
  • Trias E, Ibarburu S, Barreto-Núñez R, et al. Post-paralysis tyrosine kinase inhibition with masitinib abrogates neuroinflammation and slows disease progression in inherited amyotrophic lateral sclerosis. J Neuroinflammation. 2016;13:177.
  • Suzumura A, Ito A, Yoshikawa M, et al. Ibudilast suppresses TNFalpha production by glial cells functioning mainly as type III phosphodiesterase inhibitor in the CNS. Brain Res. 1999;837:203–212.
  • Mizuno T, Kurotani T, Komatsu Y, et al. Neuroprotective role of phosphodiesterase inhibitor ibudilast on neuronal cell death induced by activated microglia. Neuropharmacology. 2004;46:404–411.
  • Kang SH, Li Y, Fukaya M, et al. Degeneration and impaired regeneration of gray matter oligodendrocytes in amyotrophic lateral sclerosis. Nat Neurosci. 2013;16:571–579.
  • Philips T, Bento-Abreu A, Nonneman A, et al. Oligodendrocyte dysfunction in the pathogenesis of amyotrophic lateral sclerosis [Internet]. Brain. 2013;136:471–482.
  • Lee Y, Morrison BM, Li Y, et al. Oligodendroglia metabolically support axons and contribute to neurodegeneration. Nature. 2012;487:443–448.
  • Kim S, Chung A-Y, Na JE, et al. Myelin degeneration induced by mutant superoxide dismutase 1 accumulation promotes amyotrophic lateral sclerosis. Glia. 2019;67:1910–1921.
  • Ferraiuolo L, Meyer K, Sherwood TW, et al. Oligodendrocytes contribute to motor neuron death in ALS via SOD1-dependent mechanism. Proc Natl Acad Sci U S A. 2016;113:E6496–E6505.
  • van Rheenen W, Shatunov A, Dekker AM, et al. Genome-wide association analyses identify new risk variants and the genetic architecture of amyotrophic lateral sclerosis. Nat Genet. 2016;48:1043–1048.
  • Clarke BE, Taha DM, Tyzack GE, et al. Regionally encoded functional heterogeneity of astrocytes in health and disease: A perspective [Internet]. Glia. 2020. DOI:10.1002/glia.23877.
  • Zirra A, Wiethoff S, Patani R. Neural conversion and patterning of human pluripotent stem cells: a developmental perspective. Stem Cells Int. 2016;2016:8291260.
  • Patani R. Generating diverse spinal motor neuron subtypes from human pluripotent stem cells. Stem Cells Int. 2016;2016:1036974.
  • Maffioletti SM, Sarcar S, Henderson ABH, et al. Three-dimensional human iPSC-derived artificial skeletal muscles model muscular dystrophies and enable multilineage tissue engineering. Cell Rep. 2018;23:899–908.
  • Patani R. Human stem cell models of disease and the prognosis of academic medicine. Nat Med. 2020;26:449.
  • Devine H, Patani R. The translational potential of human induced pluripotent stem cells for clinical neurology: the translational potential of hiPSCs in neurology. Cell Biol Toxicol. 2017;33:129–144.
  • Kelley KW, Ben Haim L, Schirmer L, et al. Kir4.1-dependent astrocyte-fast motor neuron interactions are required for peak strength. Neuron. 2018;98:306–319.e7.
  • Simone R, Balendra R, Moens TG, et al. G-quadruplex-binding small molecules ameliorate FTD/ALS pathology and. EMBO Mol Med. 2018;10:22–31.
  • Patani R, Lewis PA, Trabzuni D, et al. Investigating the utility of human embryonic stem cell-derived neurons to model ageing and neurodegenerative disease using whole-genome gene expression and splicing analysis. J Neurochem. 2012;122:738–751.
  • Ziff OJ, Patani R. Harnessing cellular aging in human stem cell models of amyotrophic lateral sclerosis. Aging Cell. 2019;18:e12862.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.