Genome-wide CRISPR screen reveals v-ATPase as a drug target to lower levels of ALS protein ataxin-2

Garam Kim, Lisa Nakayama, Jacob A Blum, Tetsuya Akiyama, Steven Boeynaems, Meenakshi Chakraborty, Julien Couthouis, Eduardo Tassoni-Tsuchida, Caitlin M Rodriguez, Michael C Bassik, Aaron D Gitler, Garam Kim, Lisa Nakayama, Jacob A Blum, Tetsuya Akiyama, Steven Boeynaems, Meenakshi Chakraborty, Julien Couthouis, Eduardo Tassoni-Tsuchida, Caitlin M Rodriguez, Michael C Bassik, Aaron D Gitler

Abstract

Mutations in the ataxin-2 gene (ATXN2) cause the neurodegenerative disorders amyotrophic lateral sclerosis (ALS) and spinocerebellar ataxia type 2 (SCA2). A therapeutic strategy using antisense oligonucleotides targeting ATXN2 has entered clinical trial in humans. Additional ways to decrease ataxin-2 levels could lead to cheaper or less invasive therapies and elucidate how ataxin-2 is normally regulated. Here, we perform a genome-wide fluorescence-activated cell sorting (FACS)-based CRISPR-Cas9 screen in human cells and identify genes encoding components of the lysosomal vacuolar ATPase (v-ATPase) as modifiers of endogenous ataxin-2 protein levels. Multiple FDA-approved small molecule v-ATPase inhibitors lower ataxin-2 protein levels in mouse and human neurons, and oral administration of at least one of these drugs-etidronate-is sufficient to decrease ataxin-2 in the brains of mice. Together, we propose v-ATPase as a drug target for ALS and SCA2 and demonstrate the value of FACS-based screens in identifying genetic-and potentially druggable-modifiers of human disease proteins.

Trial registration: ClinicalTrials.gov NCT04494256.

Keywords: ALS; CP: Neuroscience; FACS; SCA2; TDP-43; ataxin-2; bisphosphonate; etidronate; genetic screens; small-molecule therapy; v-ATPase.

Conflict of interest statement

Declaration of interests A.D.G. is a scientific founder of Maze Therapeutics. Stanford University has filed a provisional patent (63/286,436) on methods described in this manuscript for treatment of neurodegenerative diseases through the inhibition of ataxin-2.

Copyright © 2022 The Author(s). Published by Elsevier Inc. All rights reserved.

Figures

Figure 1.. Genome-wide CRISPR-Cas9 KO screens in…
Figure 1.. Genome-wide CRISPR-Cas9 KO screens in human cells identify regulators of ataxin-2 protein levels
(A) Pooled CRISPR-Cas9 screening paradigm. After transducing HeLa cells expressing Cas9 with a lentiviral whole-genome sgRNA library, we fixed and co-immunostained the cells for ataxin-2 and a control protein (β-actin or GAPDH). We then used FACS to sort top and bottom 20% ataxin-2 expressors relative to control protein levels (duplicate sorts per each control). After isolating genomic DNA from these populations, as well as the unsorted control population, we performed NGS to read sgRNA barcodes. (B) Volcano plots based on effect and confidence scores summarizing genes that modify ataxin-2 protein levels relative to β-actin (left) or GAPDH (right) levels when knocked out (FDR PNISR, PAXBP1, or ATXN2 siRNAs in HeLa cells.
Figure 2.. Schematic of proteins encoded by…
Figure 2.. Schematic of proteins encoded by selected hits (5% FDR), categorized by function and subcellular localization
Figure 3.. Genetically and pharmacologically perturbing lysosomal…
Figure 3.. Genetically and pharmacologically perturbing lysosomal v-ATPase leads to decreased ataxin-2 protein levels in vitro
(A) Left, volcano plot shows confidence score on y axis and effect score on x axis, with gene hits encoding lysosomal v-ATPase subunits highlighted in red. Right, a representation of the lysosomal v-ATPase, with its V0 and V1 domains, as well as individual subunits. (B and C) Immunoblot (B) and quantification (C) of ataxin-2 protein levels after HeLa cells were transfected with siRNAs against various v-ATPase subunits. (D) Quantification of ATXN2 RNA levels after siRNA knockdown of v-ATPase subunits in HeLa cells. Values normalized to β-actin RNA levels. Quantifications for (C) and (D) are normalized to the NT siRNA condition (mean ± SD; analyzed using one-way ANOVA with post hoc Dunnett’s multiple comparisons tests; ****p ATP6V1A Cas9-edited HeLa cell lines. (F) Representative microscopy images of WT or ATP6V1A Cas9-edited HeLa cells, immunostained for ataxin-2 or β-actin (scale bar: 20 μm). Ataxin-2 fluorescence quantifications are shown on the right (lines denote mean ± SD; analyzed using unpaired t test; ****p < 0.0001).
Figure 4.. Small-molecule drug etidronate lowers ataxin-2…
Figure 4.. Small-molecule drug etidronate lowers ataxin-2 protein levels in human iPSC-derived neurons, mouse primary neurons, and in vivo in mice
(A) Timeline of induced neuron differentiation in a human iPSC line with NGN2 stably integrated and drug treatment. (B) Immunoblot on lysates from human iPSC-derived neurons treated with various doses of etidronate. (C) Quantification of Figure 4B, with ataxin-2 protein levels normalized to H2O-treated condition (mean ± SD; analyzed using one-way ANOVA with post hoc Dunnett’s multiple comparisons tests; *p < 0.05). (D) Timeline of primary neuron plating from embryonic mouse cortex and drug treatment. (E) Immunoblot on lysates from mouse primary neurons treated with various doses of etidronate. (F) Quantification of the dose-dependent effect of etidronate on ataxin-2 (normalized to control condition). (G) Representative microscopy images of mouse cortical neurons treated with sham (H2O) or 10 μM etidronate for 24 h, stained for MAP2, ataxin-2, and DAPI (scale bar: 10 μm). (H) Representative microscopy images of mouse cortical neurons treated with H2O or 10 μM etidronate for 24 h, with 0.5 mM sodium arsenite treatment for the final hour. The neurons were stained for PABP, MAP2, and DAPI (scale bar: 10 μm). (I) Quantifications of cells containing PABP-positive stress granules (SGs) (mean ± SEM; analyzed using unpaired t test; ****p 2O-treated condition) using lysates from cortices of mice that received water or drug treatment (mean ± SEM; analyzed using Welch’s t test; **p < 0.01). We performed the experiment two independent times, for a total of n = 16 in the control group and n = 14 in the drug treatment group.

References

    1. Altman RD, Johnston CC, Khairi MR, Wellman H, Serafini AN, and Sankey RR (1973). Influence of disodium etidronate on clinical and laboratory manifestations of Paget’s disease of bone (osteitis deformans). N. Engl. J. Med. 289, 1379–1384. 10.1056/NEJM197312272892601.
    1. Armakola M, Hart MP, and Gitler AD (2011). TDP-43 toxicity in yeast. Methods 53, 238–245. 10.1016/j.ymeth.2010.11.006.
    1. Arnold ES, Ling S-C, Huelga SC, Lagier-Tourenne C, Polymenidou M, Ditsworth D, Kordasiewicz HB, McAlonis-Downes M, Platoshyn O, Parone PA, et al. (2013). ALS-linked TDP-43 mutations produce aberrant RNA splicing and adult-onset motor neuron disease without aggregation or loss of nuclear TDP-43. Proc. Natl. Acad. Sci. USA 110, E736–E745. 10.1073/pnas.1222809110.
    1. Becker LA, Huang B, Bieri G, Ma R, Knowles DA, Jafar-Nejad P, Messing J, Kim HJ, Soriano A, Auburger G, et al. (2017). Therapeutic reduction of ataxin-2 extends lifespan and reduces pathology in TDP-43 mice. Nature 544, 367–371. 10.1038/nature22038.
    1. Bennett CF, Kordasiewicz HB, and Cleveland DW (2021). Antisense drugs make sense for neurological diseases. Annu. Rev. Pharmacol. Toxicol. 61, 831–852. 10.1146/annurev-pharmtox-010919-023738.
    1. Chan C-Y, Prudom C, Raines SM, Charkhzarrin S, Melman SD, De Haro LP, Allen C, Lee SA, Sklar LA, and Parra KJ (2012). Inhibitors of V-ATPase proton transport reveal uncoupling functions of tether linking cytosolic and membrane domains of V0 subunit a (Vph1p). J. Biol. Chem. 287, 10236–10250. 10.1074/jbc.M111.321133.
    1. David P, Nguyen H, Barbier A, and Baron R (1996). The bisphosphonate tiludronate is a potent inhibitor of the osteoclast vacuolar H(+)-ATPase. J. Bone Miner. Res. 11, 1498–1507. 10.1002/jbmr.5650111017.
    1. de Boer EMJ, Orie VK, Williams T, Baker MR, De Oliveira HM, Polvikoski T, Silsby M, Menon P, van den Bos M, Halliday GM, et al. (2020). TDP-43 proteinopathies: a new wave of neurodegenerative diseases. J. Neurol. Neurosurg. Psychiatry 92, 86–95. 10.1136/jnnp-2020-322983.
    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. 10.1093/bioinformatics/bts635.
    1. Drake MT, Clarke BL, and Khosla S (2008). Bisphosphonates: mechanism of action and role in clinical practice. Mayo Clin. Proc. 83, 1032–1045. 10.4065/83.9.1032.
    1. Elden AC, Kim H-J, Hart MP, Chen-Plotkin AS, Johnson BS, Fang X, Armakola M, Geser F, Greene R, Lu MM, et al. (2010). Ataxin-2 intermediate-length polyglutamine expansions are associated with increased risk for ALS. Nature 466, 1069–1075. 10.1038/nature09320.
    1. Forman MS, Trojanowski JQ, and Lee VM-Y (2004). Neurodegenerative diseases: a decade of discoveries paves the way for therapeutic breakthroughs. Nat. Med. 10, 1055–1063. 10.1038/nm1113.
    1. Gitcho MA, Strider J, Carter D, Taylor-Reinwald L, Forman MS, Goate AM, and Cairns NJ (2009). VCP mutations causing frontotemporal lobar degeneration disrupt localization of TDP-43 and induce cell death. J. Biol. Chem. 284, 12384–12398. 10.1074/jbc.M900992200.
    1. Hart MP, and Gitler AD (2012). ALS-associated ataxin 2 polyQ expansions enhance stress-induced caspase 3 activation and increase TDP-43 pathological modifications. J. Neurosci. 32, 9133–9142. 10.1523/JNEUROSCI.0996-12.2012.
    1. Imbert G, Saudou F, Yvert G, Devys D, Trottier Y, Garnier JM,Weber C, Mandel JL, Cancel G, Abbas N, et al. (1996). Cloning of the gene for spinocerebellar ataxia 2 reveals a locus with high sensitivity to expanded CAG/glutamine repeats. Nat. Genet. 14, 285–291. 10.1038/ng1196-285.
    1. Kim G, Gautier O, Tassoni-Tsuchida E, Ma XR, and Gitler AD (2020). ALS genetics: gains, losses, and implications for future therapies. Neuron 108, 822–842. 10.1016/j.neuron.2020.08.022.
    1. Kim H-J, Raphael AR, LaDow ES, McGurk L, Weber RA, Trojanowski JQ, Lee VM-Y, Finkbeiner S, Gitler AD, and Bonini NM (2014). Therapeutic modulation of eIF2α phosphorylation rescues TDP-43 toxicity in amyotrophic lateral sclerosis disease models. Nat. Genet. 46, 152–160. 10.1038/ng.2853.
    1. Kramer NJ, Haney MS, Morgens DW, Jovičić A, Couthouis J, Li A, Ousey J, Ma R, Bieri G, Tsui CK, et al. (2018). CRISPR–Cas9 screens in human cells and primary neurons identify modifiers of C9ORF72 dipeptide-repeat-protein toxicity. Nat. Genet. 50, 603–612. 10.1038/s41588-018-0070-7.
    1. Li YR, King OD, Shorter J, and Gitler AD (2013). Stress granules as crucibles of ALS pathogenesis. J. Cell Biol. 201, 361–372. 10.1083/jcb.201302044.
    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. 10.1186/s13059-014-0550-8.
    1. Lu B, Al-Ramahi I, Valencia A, Wang Q, Berenshteyn F, Yang H, Gallego-Flores T, Ichcho S, Lacoste A, Hild M, et al. (2013). Identification of NUB1 as a suppressor of mutant Huntington toxicity via enhanced protein clearance. Nat. Neurosci. 16, 562–570. 10.1038/nn.3367.
    1. Ma XR, Prudencio M, Koike Y, Vatsavayai SC, Kim G, Harbinski F, Briner A, Rodriguez CM, Guo C, Akiyama T, et al. (2022). TDP-43 represses cryptic exon inclusion in the FTD-ALS gene UNC13A. Nature 603, 124–130.
    1. Maxson ME, and Grinstein S (2014). The vacuolar-type H+-ATPase at a glance - more than a proton pump. J. Cell Sci. 127, 4987–4993. 10.1242/jcs.158550.
    1. Morgens DW, Deans RM, Li A, and Bassik MC (2016). Systematic comparison of CRISPR/Cas9 and RNAi screens for essential genes. Nat. Biotechnol. 34, 634–636. 10.1038/nbt.3567.
    1. Morgens DW, Wainberg M, Boyle EA, Ursu O, Araya CL, Tsui CK, Haney MS, Hess GT, Han K, Jeng EE, et al. (2017). Genome-scale measurement of off-target activity using Cas9 toxicity in high-throughput screens. Nat. Commun. 8, 15178. 10.1038/ncomms15178.
    1. Neumann M, Mackenzie IR, Cairns NJ, Boyer PJ, Markesbery WR, Smith CD, Taylor JP, Kretzschmar HA, Kimonis VE, and Forman MS (2007). TDP-43 in the ubiquitin pathology of frontotemporal dementia with VCP gene mutations. J. Neuropathol. Exp. Neurol. 66, 152–157. 10.1097/nen.0b013e31803020b9.
    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. 10.1126/science.1134108.
    1. Park J, Al-Ramahi I, Tan Q, Mollema N, Diaz-Garcia JR, Gallego-Flores T, Lu H-C, Lagalwar S, Duvick L, Kang H, et al. (2013). RAS-MAPK-MSK1 pathway modulates ataxin 1 protein levels and toxicity in SCA1. Nature 498, 325–331. 10.1038/nature12204.
    1. Pulst SM, Nechiporuk A, Nechiporuk T, Gispert S, Chen XN, LopesCendes I, Pearlman S, Starkman S, Orozco-Diaz G, Lunkes A, et al. (1996). Moderate expansion of a normally biallelic trinucleotide repeat in spinocerebellar ataxia type 2. Nat. Genet. 14, 269–276. 10.1038/ng1196-269.
    1. Rodriguez CM, Bechek SC, Jones GL, Nakayama L, Akiyama T, Kim G, Solow-Cordero DE, Strittmatter SM, and Gitler AD (2022). Targeting RTN4/NoGo-Receptor reduces levels of ALS protein ataxin-2. Cell Rep. 41. 10.1101/2021.12.20.473562.
    1. Rossi A, Kontarakis Z, Gerri C, Nolte H, Hölper S, Krüger M, and Stainier DYR (2015). Genetic compensation induced by deleterious mutations but not gene knockdowns. Nature 524, 230–233. 10.1038/nature14580.
    1. Rousseaux MWC, Vázquez-Vélez GE, Al-Ramahi I, Jeong H-H, Bajić A, Revelli J-P, Ye H, Phan ET, Deger JM, Perez AM, et al. (2018). A druggable genome screen identifies modifiers of α-synuclein levels via a tiered cross-species validation approach. J. Neurosci. 38, 9286–9301. 10.1523/JNEUROSCI.0254-18.2018.
    1. Sanpei K, Takano H, Igarashi S, Sato T, Oyake M, Sasaki H, Wakisaka A, Tashiro K, Ishida Y, Ikeuchi T, et al. (1996). Identification of the spinocerebellar ataxia type 2 gene using a direct identification of repeat expansion and cloning technique, DIRECT. Nat. Genet. 14, 277–284. 10.1038/ng1196-277.
    1. Scoles DR, Meera P, Schneider MD, Paul S, Dansithong W, Figueroa KP, Hung G, Rigo F, Bennett CF, Otis TS, and Pulst SM (2017). Antisense oligonucleotide therapy for spinocerebellar ataxia type 2. Nature 544, 362–366. 10.1038/nature22044.
    1. Scoles DR, and Pulst SM (2018). Spinocerebellar ataxia type 2. Adv. Exp. Med. Biol. 1049, 175–195. 10.1007/978-3-319-71779-1_8.
    1. Watts GDJ, Wymer J, Kovach MJ, Mehta SG, Mumm S, Darvish D, Pestronk A, Whyte MP, and Kimonis VE (2004). Inclusion body myopathy associated with Paget disease of bone and frontotemporal dementia is caused by mutant valosin-containing protein. Nat. Genet. 36, 377–381. 10.1038/ng1332.
    1. Wils H, Kleinberger G, Janssens J, Pereson S, Joris G, Cuijt I, Smits V, Ceuterick-de Groote C, Van Broeckhoven C, and Kumar-Singh S (2010). TDP-43 transgenic mice develop spastic paralysis and neuronal inclusions characteristic of ALS and frontotemporal lobar degeneration. Proc. Natl. Acad. Sci. USA 107, 3858–3863. 10.1073/pnas.0912417107.
    1. Yin J-A, Frick L, Scheidmann MC, Trevisan C, Dhingra A, Spinelli A, Wu Y, Yao L, Vena DL, De Cecco E, et al. (2022). Robust and versatile arrayed libraries for human genome-wide CRISPR activation, deletion and silencing. Preprint at bioRxiv. 10.1101/2022.05.25.493370.
    1. Zhang K, Daigle JG, Cunningham KM, Coyne AN, Ruan K, Grima JC, Bowen KE, Wadhwa H, Yang P, Rigo F, et al. (2018). Stress granule assembly disrupts nucleocytoplasmic transport. Cell 173, 958–971.e17. 10.1016/j.cell.2018.03.025.

Source: PubMed

3
Prenumerera