Skip to main content
Log in

Recent Advances in RNA Interference Therapeutics for CNS Diseases

  • Published:
Neurotherapeutics

Abstract

Over the last decade, RNA interference technology has shown therapeutic promise in rodent models of dominantly inherited brain diseases, including those caused by polyglutamine repeat expansions in the coding region of the affected gene. For some of these diseases, proof-of concept studies in model organisms have transitioned to safety testing in larger animal models, such as the nonhuman primate. Here, we review recent progress on RNA interference-based therapies in various model systems. We also highlight outstanding questions or concerns that have emerged as a result of an improved (and ever advancing) understanding of the technologies employed.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

References

  1. Zu T, Duvick LA, Kaytor MD, et al. Recovery from polyglutamine-induced neurodegeneration in conditional SCA1 transgenic mice. J Neurosci 2004;24:8853–8861.

    PubMed  CAS  Google Scholar 

  2. Yamamoto A, Lucas JJ, Hen R. Reversal of neuropathology and motor dysfunction in a conditional model of Huntington's disease. Cell 2000;101:57–66.

    PubMed  CAS  Google Scholar 

  3. Diaz-Hernandez M, Torres-Peraza J, Salvatori-Abarca A, et al. Full motor recovery despite striatal neuron loss and formation of irreversible amyloid-like inclusions in a conditional mouse model of Huntington's disease. J Neurosci 2005;25:9773–9781.

    PubMed  CAS  Google Scholar 

  4. Engel T, Hernandez F, Avila J, Lucas JJ. Full reversal of Alzheimer's disease-like phenotype in a mouse model with conditional overexpression of glycogen synthase kinase-3. J Neurosci 2006;26:5083–5090.

    PubMed  CAS  Google Scholar 

  5. Carrington JC, Ambros V. Role of microRNAs in plant and animal development. Science 2003;301:336–338.

    PubMed  CAS  Google Scholar 

  6. Krol J, Loedige I, Filipowicz W. The widespread regulation of microRNA biogenesis, function and decay. Nat Rev Genet 2010;11:597–610.

    PubMed  CAS  Google Scholar 

  7. Monteys AM, Spengler RM, Wan J, et al. Structure and activity of putative intronic miRNA promoters. RNA 2010;16:495–505.

    PubMed  Google Scholar 

  8. Lee Y, Ahn C, Han J, et al. The nuclear RNase III Drosha initiates microRNA processing. Nature 2003;425:415–419.

    PubMed  CAS  Google Scholar 

  9. Gregory RI, Yan KP, Amuthan G, et al. The Microprocessor complex mediates the genesis of microRNAs. Nature 2004;432:235–240.

    PubMed  CAS  Google Scholar 

  10. Yi R, Qin Y, Macara IG, Cullen BR. Exportin-5 mediates the nuclear export of pre-microRNAs and short hairpin RNAs. Genes Dev 2003;17:3011–3016.

    PubMed  CAS  Google Scholar 

  11. Lee Y, Jeon K, Lee JT, Kim S, Kim VN. MicroRNA maturation: stepwise processing and subcellular localization. EMBO J 2002;21:4663–4670.

    PubMed  CAS  Google Scholar 

  12. Ding SW. RNA-based antiviral immunity. Nat Rev Immunol 2010;10:632–644.

    PubMed  CAS  Google Scholar 

  13. Khvorova A, Reynolds A, Jayasena SD. Functional siRNAs and miRNAs exhibit strand bias. Cell 2003;115:209–216.

    PubMed  CAS  Google Scholar 

  14. Guo H, Ingolia NT, Weissman JS, Bartel DP. Mammalian microRNAs predominantly act to decrease target mRNA levels. Nature 2010;466:835–840.

    PubMed  CAS  Google Scholar 

  15. Eulalio A, Huntzinger E, Nishihara T, et al. Deadenylation is a widespread effect of miRNA regulation. RNA 2009;15:21–32.

    PubMed  CAS  Google Scholar 

  16. Lewis BP, Burge CB, Bartel DP. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 2005;120:15–20.

    PubMed  CAS  Google Scholar 

  17. Kim DH, Behlke MA, Rose SD, et al. Synthetic dsRNA Dicer substrates enhance RNAi potency and efficacy. Nat Biotechnol 2005;23:222–226.

    PubMed  CAS  Google Scholar 

  18. Boudreau RL, Martins I, Davidson BL. Artificial microRNAs as siRNA shuttles: improved safety as compared to shRNAs in vitro and in vivo. Mol Ther 2009;17:169–175.

    PubMed  CAS  Google Scholar 

  19. McBride JL, Boudreau RL, Harper SQ, et al. Artificial miRNAs mitigate shRNA-mediated toxicity in the brain: implications for the therapeutic development of RNAi. Proc Natl Acad Sci U S A 2008;105:5868–5873.

    PubMed  CAS  Google Scholar 

  20. Han Y, Khodr CE, Sapru MK, Pedapati J, Bohn MC. A microRNA embedded AAV alpha-synuclein gene silencing vector for dopaminergic neurons. Brain Res 2011;1386:15–24.

    PubMed  CAS  Google Scholar 

  21. Martin JN, Wolken N, Brown T, et al. Lethal toxicity caused by expression of shRNA in the mouse striatum: implications for therapeutic design. Gene Ther 2011;18:666–673.

    PubMed  CAS  Google Scholar 

  22. Khodr CE, Sapru MK, Pedapati J, et al. An alpha-synuclein AAV gene silencing vector ameliorates a behavioral deficit in a rat model of Parkinson's disease, but displays toxicity in dopamine neurons. Brain Res 2011;1395:94–107.

    PubMed  CAS  Google Scholar 

  23. Ehlert EM, Eggers R, Niclou SP, Verhaagen J. Cellular toxicity following application of adeno-associated viral vector-mediated RNA interference in the nervous system. BMC Neurosci 2010;11:20.

    PubMed  Google Scholar 

  24. Birmingham A, Anderson E, Sullivan K, et al. A protocol for designing siRNAs with high functionality and specificity. Nat Protoc 2007;2:2068–2078.

    PubMed  CAS  Google Scholar 

  25. Jackson AL, Linsley PS. Recognizing and avoiding siRNA off-target effects for target identification and therapeutic application. Nat Rev Drug Discov 2010;9:57–67.

    PubMed  CAS  Google Scholar 

  26. Matveeva O, Nechipurenko Y, Rossi L, et al. Comparison of approaches for rational siRNA design leading to a new efficient and transparent method. Nucleic Acids Res 2007;35:e63.

    PubMed  Google Scholar 

  27. Gong W, Ren Y, Zhou H, et al. siDRM: an effective and generally applicable online siRNA design tool. Bioinformatics 2008;24:2405–2406.

    Google Scholar 

  28. Lu ZJ, Mathews DH. OligoWalk: an online siRNA design tool utilizing hybridization thermodynamics. Nucleic Acids Res 2008;36:W104-108.

    PubMed  CAS  Google Scholar 

  29. Naito Y, Yamada T, Ui-Tei K, Morishita S, Saigo K. siDirect: highly effective, target-specific siRNA design software for mammalian RNA interference. Nucleic Acids Res 2004;32:W124-129.

    PubMed  CAS  Google Scholar 

  30. Wang L, Mu FY. A Web-based design center for vector-based siRNA and siRNA cassette. Bioinformatics 2004;20:1818–1820.

    PubMed  CAS  Google Scholar 

  31. Boudreau RL, Spengler RM, Hylock RH, et al. siSPOTR: a tool for designing highly specific and potent siRNAs for human and mouse. Nucleic Acids Res 2013;41:e9.

    Google Scholar 

  32. Naito Y, Ui-Tei K. Designing functional siRNA with reduced off-target effects. Methods Mol Biol 2013;942:57–68.

    PubMed  CAS  Google Scholar 

  33. Naito Y, Yoshimura J, Morishita S, Ui-Tei K. siDirect 2.0: updated software for designing functional siRNA with reduced seed-dependent off-target effect. BMC Bioinformatics 2009;10:392.

    PubMed  Google Scholar 

  34. Fedorov Y, Anderson EM, Birmingham A, et al. Off-target effects by siRNA can induce toxic phenotype. RNA 2006;12:1188–1196.

    PubMed  CAS  Google Scholar 

  35. Birmingham A, Anderson EM, Reynolds A, et al. 3' UTR seed matches, but not overall identity, are associated with RNAi off-targets. Nat Methods 2006;3:199–204.

    PubMed  CAS  Google Scholar 

  36. Jackson AL, Burchard J, Schelter J, et al. Widespread siRNA "off-target" transcript silencing mediated by seed region sequence complementarity. RNA 2006;12:1179–1187.

    PubMed  CAS  Google Scholar 

  37. Boudreau RL, Davidson BL. Generation of hairpin-based RNAi vectors for biological and therapeutic application. Methods Enzymol 2012;507:275–296.

    PubMed  CAS  Google Scholar 

  38. Lentz TB, Gray SJ, Samulski RJ. Viral vectors for gene delivery to the central nervous system. Neurobiol Dis 2012;48:179–188.

    PubMed  CAS  Google Scholar 

  39. Rettig GR, Behlke MA. Progress toward in vivo use of siRNAs-II. Mol Ther 2012;20:483–512.

    PubMed  CAS  Google Scholar 

  40. Yu D, Pendergraff H, Liu J, et al. Single-Stranded RNAs Use RNAi to Potently and Allele-Selectively Inhibit Mutant Huntingtin Expression. Cell 2012;150:895–908.

    PubMed  CAS  Google Scholar 

  41. Lima WF, Prakash TP, Murray HM, et al. Single-stranded siRNAs activate RNAi in animals. Cell 2012;150:883–894.

    PubMed  CAS  Google Scholar 

  42. Davidson BL, Monteys AM. Singles engage the RNA interference pathway. Cell 2012;150:873–875.

    PubMed  CAS  Google Scholar 

  43. Vergoni AV, Tosi G, Tacchi R, et al. Nanoparticles as drug delivery agents specific for CNS: in vivo biodistribution. Nanomedicine 2009;5:369–377.

    PubMed  CAS  Google Scholar 

  44. Tahara K, Miyazaki Y, Kawashima Y, Kreuter J, Yamamoto H. Brain targeting with surface-modified poly(D,L-lactic-co-glycolic acid) nanoparticles delivered via carotid artery administration. Eur J Pharm Biopharm 2011;77:84–88.

    PubMed  CAS  Google Scholar 

  45. Gao H, Qian J, Yang Z, et al. Whole-cell SELEX aptamer-functionalised poly(ethyleneglycol)-poly(epsilon-caprolactone) nanoparticles for enhanced targeted glioblastoma therapy. Biomaterials 2012;33:6264–6272.

    PubMed  CAS  Google Scholar 

  46. Liu C, Zhang N. Nanoparticles in gene therapy principles, prospects, and challenges. Prog Mol Biol Transl Sci 2011;104:509–562.

    PubMed  CAS  Google Scholar 

  47. Stein CS, Martins I, Davidson BL. The lymphocytic choriomeningitis virus envelope glycoprotein targets lentiviral gene transfer vector to neural progenitors in the murine brain. Mol Ther 2005;11:382–389.

    PubMed  CAS  Google Scholar 

  48. Lattanzi A, Neri M, Maderna C, et al. Widespread enzymatic correction of CNS tissues by a single intracerebral injection of therapeutic lentiviral vector in leukodystrophy mouse models. Hum Mol Genet 2010;19:2208–2227.

    PubMed  CAS  Google Scholar 

  49. Di Domenico C, Villani GR, Di Napoli D, et al. Gene therapy for a mucopolysaccharidosis type I murine model with lentiviral-IDUA vector. Hum Gene Ther 2005;16:81–90.

    PubMed  Google Scholar 

  50. Drouet V, Perrin V, Hassig R, et al. Sustained effects of nonallele-specific Huntingtin silencing. Ann Neurol 2009;65:276–285.

    PubMed  CAS  Google Scholar 

  51. Raoul C, Abbas-Terki T, Bensadoun JC, et al. Lentiviral-mediated silencing of SOD1 through RNA interference retards disease onset and progression in a mouse model of ALS. Nat Med 2005;11:423–428.

    PubMed  CAS  Google Scholar 

  52. Ralph GS, Radcliffe PA, Day DM, et al. Silencing mutant SOD1 using RNAi protects against neurodegeneration and extends survival in an ALS model. Nat Med 2005;11:429–433.

    PubMed  CAS  Google Scholar 

  53. Gonzalez-Alegre P, Bode N, Davidson BL, Paulson HL. Silencing primary dystonia: lentiviral-mediated RNA interference therapy for DYT1 dystonia. J Neurosci 2005;25:10502–10509.

    PubMed  CAS  Google Scholar 

  54. Harper SQ, Staber PD, Beck CR, et al. Optimization of feline immunodeficiency virus vectors for RNA interference. J Virol 2006;80:9371–9380.

    PubMed  CAS  Google Scholar 

  55. Hacein-Bey-Abina S, von Kalle C, Schmidt M, et al. A serious adverse event after successful gene therapy for X-linked severe combined immunodeficiency. N Engl J Med 2003;348:255–256.

    PubMed  Google Scholar 

  56. Hacein-Bey-Abina S, Von Kalle C, Schmidt M, et al. LMO2-associated clonal T cell proliferation in two patients after gene therapy for SCID-X1. Science 2003;302:415–419.

    PubMed  CAS  Google Scholar 

  57. Vandenberghe LH, Xiao R, Lock M, et al. Efficient serotype-dependent release of functional vector into the culture medium during adeno-associated virus manufacturing. Hum Gene Ther 2010;21:1251–1257.

    PubMed  CAS  Google Scholar 

  58. Anderson RD, Haskell RE, Xia H, Roessler BJ, Davidson BL. A simple method for the rapid generation of recombinant adenovirus vectors. Gene Ther 2000;7:1034–1038.

    PubMed  CAS  Google Scholar 

  59. Mittermeyer G, Christine CW, Rosenbluth KH, et al. Long-term evaluation of a phase 1 study of AADC gene therapy for Parkinson's disease. Hum Gene Ther 2012;23:377–381.

    PubMed  CAS  Google Scholar 

  60. Hadaczek P, Eberling JL, Pivirotto P, et al. Eight years of clinical improvement in MPTP-lesioned primates after gene therapy with AAV2-hAADC. Mol Ther 2010;18:1458–1461.

    PubMed  CAS  Google Scholar 

  61. Petit L, Lheriteau E, Weber M, et al. Restoration of vision in the pde6beta-deficient dog, a large animal model of rod-cone dystrophy. Mol Ther 2012;20:2019–2030.

    PubMed  CAS  Google Scholar 

  62. Xia H, Anderson B, Mao Q, Davidson BL. Recombinant human adenovirus: targeting to the human transferrin receptor improves gene transfer to brain microcapillary endothelium. J Virol 2000;74:11359–11366.

    PubMed  CAS  Google Scholar 

  63. Alisky JM, Hughes SM, Sauter SL, et al. Transduction of murine cerebellar neurons with recombinant FIV and AAV5 vectors. Neuroreport 2000;11:2669–2673.

    PubMed  CAS  Google Scholar 

  64. Davidson BL, Stein CS, Heth JA, et al. Recombinant adeno-associated virus type 2, 4, and 5 vectors: transduction of variant cell types and regions in the mammalian central nervous system. Proc Natl Acad Sci U S A 2000;97:3428–3432.

    PubMed  CAS  Google Scholar 

  65. Wang C, Wang CM, Clark KR, Sferra TJ. Recombinant AAV serotype 1 transduction efficiency and tropism in the murine brain. Gene Ther 2003;10:1528–1534.

    PubMed  CAS  Google Scholar 

  66. Liu G, Martins IH, Chiorini JA, Davidson BL. Adeno-associated virus type 4 (AAV4) targets ependyma and astrocytes in the subventricular zone and RMS. Gene Ther 2005;12:1503–1508.

    PubMed  CAS  Google Scholar 

  67. Gray SJ, Blake BL, Criswell HE, et al. Directed evolution of a novel adeno-associated virus (AAV) vector that crosses the seizure-compromised blood–brain barrier (BBB). Mol Ther 2010;18:570–578.

    PubMed  CAS  Google Scholar 

  68. Koerber JT, Klimczak R, Jang JH, et al. Molecular evolution of adeno-associated virus for enhanced glial gene delivery. Mol Ther 2009;17:2088–2095.

    PubMed  CAS  Google Scholar 

  69. Koerber JT, Jang JH, Schaffer DV. DNA shuffling of adeno-associated virus yields functionally diverse viral progeny. Mol Ther 2008;16:1703–1709.

    PubMed  CAS  Google Scholar 

  70. Li W, Asokan A, Wu Z, et al. Engineering and selection of shuffled AAV genomes: a new strategy for producing targeted biological nanoparticles. Mol Ther 2008;16:1252–1260.

    PubMed  CAS  Google Scholar 

  71. Grimm D, Lee JS, Wang L, et al. in vitro and in vivo gene therapy vector evolution via multispecies interbreeding and retargeting of adeno-associated viruses. J Virol 2008;82:5887–5911.

    PubMed  CAS  Google Scholar 

  72. Munch RC, Janicki H, Volker I, et al. Displaying high-affinity ligands on adeno-associated viral vectors enables tumor cell-specific and safe gene transfer. Mol Ther 2013;21:109–118.

    PubMed  Google Scholar 

  73. Shi W, Bartlett JS. RGD inclusion in VP3 provides adeno-associated virus type 2 (AAV2)-based vectors with a heparan sulfate-independent cell entry mechanism. Mol Ther 2003;7:515–525.

    PubMed  CAS  Google Scholar 

  74. Stachler MD, Chen I, Ting AY, Bartlett JS. Site-specific modification of AAV vector particles with biophysical probes and targeting ligands using biotin ligase. Mol Ther 2008;16:1467–1473.

    PubMed  CAS  Google Scholar 

  75. Liu G, Martins I, Wemmie JA, Chiorini JA, Davidson BL. Functional correction of CNS phenotypes in a lysosomal storage disease model using adeno-associated virus type 4 vectors. J Neurosci 2005;25:9321–9327.

    PubMed  CAS  Google Scholar 

  76. Gray SJ, Nagabhushan Kalburgi S, McCown TJ, Jude Samulski R. Global CNS gene delivery and evasion of anti-AAV-neutralizing antibodies by intrathecal AAV administration in non-human primates. Gene Ther 2013;20:450–459.

    Google Scholar 

  77. Dehay B, Dalkara D, Dovero S, Li Q, Bezard E. Systemic scAAV9 variant mediates brain transduction in newborn rhesus macaques. Sci Rep 2012;2:253.

    PubMed  Google Scholar 

  78. Foust KD, Nurre E, Montgomery CL, et al. Intravascular AAV9 preferentially targets neonatal neurons and adult astrocytes. Nat Biotechnol 2009;27:59–65.

    PubMed  CAS  Google Scholar 

  79. Duque S, Joussemet B, Riviere C, et al. Intravenous administration of self-complementary AAV9 enables transgene delivery to adult motor neurons. Mol Ther 2009;17:1187–1196.

    PubMed  CAS  Google Scholar 

  80. Boudreau RL, McBride JL, Martins I, et al. Nonallele-specific silencing of mutant and wild-type huntingtin demonstrates therapeutic efficacy in Huntington's disease mice. Mol Ther 2009;17:1053–1063.

    PubMed  CAS  Google Scholar 

  81. Grondin R, Kaytor MD, Ai Y, et al. Six-month partial suppression of Huntingtin is well tolerated in the adult rhesus striatum. Brain 2012;135:1197–1209.

    PubMed  Google Scholar 

  82. McBride JL, Pitzer MR, Boudreau RL, et al. Preclinical safety of RNAi-mediated HTT suppression in the rhesus macaque as a potential therapy for Huntington's disease. Mol Ther 2011;19:2152–2162.

    PubMed  CAS  Google Scholar 

  83. Nasir J, Floresco SB, O'Kusky JR, et al. Targeted disruption of the Huntington's disease gene results in embryonic lethality and behavioral and morphological changes in heterozygotes. Cell 1995;81:811–823.

    PubMed  CAS  Google Scholar 

  84. Ho LW, Brown R, Maxwell M, Wyttenbach A, Rubinsztein DC. Wild type Huntingtin reduces the cellular toxicity of mutant Huntingtin in mammalian cell models of Huntington's disease. J Med Genet 2001;38:450–452.

    PubMed  CAS  Google Scholar 

  85. Dragatsis I, Levine MS, Zeitlin S. Inactivation of Hdh in the brain and testis results in progressive neurodegeneration and sterility in mice. Nat Genet 2000;26:300–306.

    PubMed  CAS  Google Scholar 

  86. Trushina E, Dyer RB, Badger JD, 2nd, et al. Mutant huntingtin impairs axonal trafficking in mammalian neurons in vivo and in vitro. Mol Cell Biol 2004;24:8195–8209.

    PubMed  CAS  Google Scholar 

  87. Zuccato C, Tartari M, Crotti A, et al. Huntingtin interacts with REST/NRSF to modulate the transcription of NRSE-controlled neuronal genes. Nat Genet 2003;35:76–83.

    PubMed  CAS  Google Scholar 

  88. Rigamonti D, Bauer JH, De-Fraja C, et al. Wild-type huntingtin protects from apoptosis upstream of caspase-3. J Neurosci 2000;20:3705–3713.

    PubMed  CAS  Google Scholar 

  89. Hu J, Liu J, Corey DR. Allele-selective inhibition of huntingtin expression by switching to an miRNA-like RNAi mechanism. Chem Biol 2010;17:1183–1188.

    PubMed  CAS  Google Scholar 

  90. Carroll JB, Warby SC, Southwell AL, et al. Potent and selective antisense oligonucleotides targeting single-nucleotide polymorphisms in the Huntington disease gene / allele-specific silencing of mutant huntingtin. Mol Ther 2011;19:2178–2185.

    PubMed  CAS  Google Scholar 

  91. Gagnon KT, Pendergraff HM, Deleavey GF, et al. Allele-selective inhibition of mutant huntingtin expression with antisense oligonucleotides targeting the expanded CAG repeat. Biochemistry 2010;49:10166–10178.

    PubMed  CAS  Google Scholar 

  92. Pfister EL, Kennington L, Straubhaar J, et al. Five siRNAs targeting three SNPs may provide therapy for three-quarters of Huntington's disease patients. Curr Biol 2009;19:774–778.

    PubMed  CAS  Google Scholar 

  93. Zoghbi HY, Orr HT. Spinocerebellar ataxia type 1. Semin Cell Biol 1995;6:29–35.

    PubMed  CAS  Google Scholar 

  94. Xia H, Mao Q, Eliason SL, et al. RNAi suppresses polyglutamine-induced neurodegeneration in a model of spinocerebellar ataxia. Nat Med 2004;10:816–820.

    PubMed  CAS  Google Scholar 

  95. Auburger GW. Spinocerebellar ataxia type 2. Handb Clin Neurol 2012;103:423–436.

    PubMed  Google Scholar 

  96. Kiehl TR, Nechiporuk A, Figueroa KP, et al. Generation and characterization of Sca2 (ataxin-2) knockout mice. Biochem Biophys Res Commun 2006;339:17–24.

    PubMed  CAS  Google Scholar 

  97. Kasumu AW, Liang X, Egorova P, Vorontsova D, Bezprozvanny I. Chronic suppression of inositol 1,4,5-triphosphate receptor-mediated calcium signaling in cerebellar purkinje cells alleviates pathological phenotype in spinocerebellar ataxia 2 mice. J Neurosci 2012;32:12786–12796.

    PubMed  CAS  Google Scholar 

  98. Liu J, Tang TS, Tu H, et al. Deranged calcium signaling and neurodegeneration in spinocerebellar ataxia type 2. J Neurosci 2009;29:9148–62.

    PubMed  CAS  Google Scholar 

  99. Chen X, Wu J, Lvovskaya S, et al. Dantrolene is neuroprotective in Huntington's disease transgenic mouse model. Mol Neurodegener 2011;6:81.

    PubMed  Google Scholar 

  100. Chen X, Tang TS, Tu H, et al. Deranged calcium signaling and neurodegeneration in spinocerebellar ataxia type 3. J Neurosci 2008;28:12713–12724.

    PubMed  CAS  Google Scholar 

  101. Kasumu A, Bezprozvanny I. Deranged calcium signaling in Purkinje cells and pathogenesis in spinocerebellar ataxia 2 (SCA2) and other ataxias. Cerebellum 2012;11:630–639.

    PubMed  CAS  Google Scholar 

  102. Supnet C, Bezprozvanny I. The dysregulation of intracellular calcium in Alzheimer disease. Cell Calcium 2010;47:183–189.

    PubMed  CAS  Google Scholar 

  103. Paulson HL. Dominantly inherited ataxias: lessons learned from Machado-Joseph disease/spinocerebellar ataxia type 3. Semin Neurol 2007;27:133–142.

    PubMed  Google Scholar 

  104. Alves S, Nascimento-Ferreira I, Auregan G, et al. Allele-specific RNA silencing of mutant ataxin-3 mediates neuroprotection in a rat model of Machado-Joseph disease. PLoS One 2008;3:e3341.

    PubMed  Google Scholar 

  105. Alves S, Nascimento-Ferreira I, Dufour N, et al. Silencing ataxin-3 mitigates degeneration in a rat model of Machado-Joseph disease: no role for wild-type ataxin-3? Hum Mol Genet 2010;19:2380–2394.

    PubMed  CAS  Google Scholar 

  106. Hu J, Gagnon KT, Liu J, et al. Allele-selective inhibition of ataxin-3 (ATX3) expression by antisense oligomers and duplex RNAs. Biol Chem 2011;392:315–325.

    PubMed  CAS  Google Scholar 

  107. Zhuchenko O, Bailey J, Bonnen P, et al. Autosomal dominant cerebellar ataxia (SCA6) associated with small polyglutamine expansions in the alpha 1A-voltage-dependent calcium channel. Nat Genet 1997;15:62–69.

    PubMed  CAS  Google Scholar 

  108. Gomez CM, Thompson RM, Gammack JT, et al. Spinocerebellar ataxia type 6: gaze-evoked and vertical nystagmus, Purkinje cell degeneration, and variable age of onset. Ann Neurol 1997;42:933–950.

    PubMed  CAS  Google Scholar 

  109. Jodice C, Mantuano E, Veneziano L, et al. Episodic ataxia type 2 (EA2) and spinocerebellar ataxia type 6 (SCA6) due to CAG repeat expansion in the CACNA1A gene on chromosome 19p. Hum Mol Genet 1997;6:1973–1978.

    PubMed  CAS  Google Scholar 

  110. Kordasiewicz HB, Gomez CM. Molecular pathogenesis of spinocerebellar ataxia type 6. Neurotherapeutics 2007;4:285–294.

    PubMed  CAS  Google Scholar 

  111. Bourinet E, Soong TW, Sutton K, et al. Splicing of alpha 1A subunit gene generates phenotypic variants of P- and Q-type calcium channels. Nat Neurosci 1999;2:407–415.

    PubMed  CAS  Google Scholar 

  112. Tsunemi T, Ishikawa K, Jin H, Mizusawa H. Cell-type-specific alternative splicing in spinocerebellar ataxia type 6. Neurosci Lett 2008;447:78–81.

    PubMed  CAS  Google Scholar 

  113. Tsou WL, Soong BW, Paulson HL, Rodriguez-Lebron E. Splice isoform-specific suppression of the Cav2.1 variant underlying spinocerebellar ataxia type 6. Neurobiol Dis 2011;43:533–542.

    PubMed  CAS  Google Scholar 

  114. Helmlinger D, Hardy S, Sasorith S, et al. Ataxin-7 is a subunit of GCN5 histone acetyltransferase-containing complexes. Hum Mol Genet 2004;13:1257–1265.

    PubMed  CAS  Google Scholar 

  115. Palhan VB, Chen S, Peng GH, et al. Polyglutamine-expanded ataxin-7 inhibits STAGA histone acetyltransferase activity to produce retinal degeneration. Proc Natl Acad Sci U S A 2005;102:8472–8477.

    PubMed  CAS  Google Scholar 

  116. Helmlinger D, Hardy S, Abou-Sleymane G, et al. Glutamine-expanded ataxin-7 alters TFTC/STAGA recruitment and chromatin structure leading to photoreceptor dysfunction. PLoS Biol 2006;4:e67.

    PubMed  Google Scholar 

  117. Scholefield J, Greenberg LJ, Weinberg MS, et al. Design of RNAi hairpins for mutation-specific silencing of ataxin-7 and correction of a SCA7 phenotype. PLoS One 2009;4:e7232.

    PubMed  Google Scholar 

  118. LaFerla FM, Oddo S. Alzheimer's disease: Abeta, tau and synaptic dysfunction. Trends Mol Med 2005;11:170–176.

    PubMed  CAS  Google Scholar 

  119. Hardy J, Selkoe DJ. The amyloid hypothesis of Alzheimer's disease: progress and problems on the road to therapeutics. Science 2002;297:353–356.

    PubMed  CAS  Google Scholar 

  120. Arriagada PV, Growdon JH, Hedley-Whyte ET, Hyman BT. Neurofibrillary tangles but not senile plaques parallel duration and severity of Alzheimer's disease. Neurology 1992;42:631–639.

    PubMed  CAS  Google Scholar 

  121. Wolfe MS. The gamma-secretase complex: membrane-embedded proteolytic ensemble. Biochemistry 2006;45:7931–7939.

    PubMed  CAS  Google Scholar 

  122. Singer O, Marr RA, Rockenstein E, et al. Targeting BACE1 with siRNAs ameliorates Alzheimer disease neuropathology in a transgenic model. Nat Neurosci 2005;8:1343–1349.

    PubMed  CAS  Google Scholar 

  123. Rodriguez-Lebron E, Gouvion CM, Moore SA, Davidson BL, Paulson HL. Allele-specific RNAi mitigates phenotypic progression in a transgenic model of Alzheimer's disease. Mol Ther 2009;17:1563–1573.

    PubMed  CAS  Google Scholar 

  124. Kandimalla RJ, Wani WY, Binukumar BK, Gill KD. siRNA against presenilin 1 (PS1) down regulates amyloid beta42 production in IMR-32 cells. J Biomed Sci 2012;19:2.

    Google Scholar 

  125. Sierant M, Paduszynska A, Kazmierczak-Baranska J, et al. Specific silencing of L392V PSEN1 mutant allele by RNA interference. Int J Alzheimers Dis 2011;2011:809218.

    PubMed  Google Scholar 

  126. Piedrahita D, Hernandez I, Lopez-Tobon A, et al. Silencing of CDK5 reduces neurofibrillary tangles in transgenic alzheimer's mice. J Neurosci 2010;30:13966–13976.

    PubMed  CAS  Google Scholar 

  127. Miller VM, Gouvion CM, Davidson BL, Paulson HL. Targeting Alzheimer's disease genes with RNA interference: an efficient strategy for silencing mutant alleles. Nucleic Acids Res 2004;32:661–668.

    PubMed  CAS  Google Scholar 

  128. Guthrie CR, Greenup L, Leverenz JB, Kraemer BC. MSUT2 is a determinant of susceptibility to tau neurotoxicity. Hum Mol Genet 2011;20:1989–1999.

    PubMed  CAS  Google Scholar 

  129. Song B, Davis K, Liu XS, et al. Inhibition of Polo-like kinase 1 reduces beta-amyloid-induced neuronal cell death in Alzheimer's disease. Aging (Albany NY) 2011;3:846–851.

    CAS  Google Scholar 

  130. Singleton AB, Farrer M, Johnson J, et al. alpha-Synuclein locus triplication causes Parkinson's disease. Science 2003;302:841.

    Google Scholar 

  131. Shen J. Protein kinases linked to the pathogenesis of Parkinson's disease. Neuron 2004;44:575–577.

    PubMed  CAS  Google Scholar 

  132. Sapru MK, Yates JW, Hogan S, et al. Silencing of human alpha-synuclein in vitro and in rat brain using lentiviral-mediated RNAi. Exp Neurol 2006;198:382–390.

    PubMed  CAS  Google Scholar 

  133. Sibley CR, Wood MJ. Identification of allele-specific RNAi effectors targeting genetic forms of Parkinson's disease. PLoS One 2011;6:e26194.

    PubMed  CAS  Google Scholar 

  134. Sibley CR, Seow Y, Curtis H, Weinberg MS, Wood MJ. Silencing of Parkinson's disease-associated genes with artificial mirtron mimics of miR-1224. Nucleic Acids Res 2012;40:9863–9875.

    PubMed  CAS  Google Scholar 

  135. Ozelius LJ, Hewett J, Kramer P, et al. Fine localization of the torsion dystonia gene (DYT1) on human chromosome 9q34: YAC map and linkage disequilibrium. Genome Res 1997;7:483–494.

    PubMed  CAS  Google Scholar 

  136. Hanson PI, Whiteheart SW. AAA+ proteins: have engine, will work. Nat Rev Mol Cell Biol 2005;6:519–529.

    PubMed  CAS  Google Scholar 

  137. Breakefield XO, Kamm C, Hanson PI. TorsinA: movement at many levels. Neuron 2001;31:9–12.

    PubMed  CAS  Google Scholar 

  138. Torres GE, Sweeney AL, Beaulieu JM, Shashidharan P, Caron MG. Effect of torsinA on membrane proteins reveals a loss of function and a dominant-negative phenotype of the dystonia-associated DeltaE-torsinA mutant. Proc Natl Acad Sci U S A 2004;101:15650–15655.

    PubMed  CAS  Google Scholar 

  139. La Spada AR, Wilson EM, Lubahn DB, Harding AE, Fischbeck KH. Androgen receptor gene mutations in X-linked spinal and bulbar muscular atrophy. Nature 1991;352:77–79.

    PubMed  Google Scholar 

  140. Sobue G, Hashizume Y, Mukai E, et al. X-linked recessive bulbospinal neuronopathy. A clinicopathological study. Brain 1989;112:209–232.

    Google Scholar 

  141. Miyazaki Y, Adachi H, Katsuno M, et al. Viral delivery of miR-196a ameliorates the SBMA phenotype via the silencing of CELF2. Nat Med 2012;18:1136–1141.

    PubMed  CAS  Google Scholar 

  142. Strong MJ. The evidence for altered RNA metabolism in amyotrophic lateral sclerosis (ALS). J Neurol Sci 2010;288:1–12.

    PubMed  CAS  Google Scholar 

  143. Bruijn LI, Miller TM, Cleveland DW. Unraveling the mechanisms involved in motor neuron degeneration in ALS. Annu Rev Neurosci 2004;27:723–749.

    PubMed  CAS  Google Scholar 

  144. Bendotti C, Atzori C, Piva R, et al. Activated p38MAPK is a novel component of the intracellular inclusions found in human amyotrophic lateral sclerosis and mutant SOD1 transgenic mice. J Neuropathol Exp Neurol 2004;63:113–119.

    PubMed  CAS  Google Scholar 

  145. Shefner JM, Reaume AG, Flood DG, et al. Mice lacking cytosolic copper/zinc superoxide dismutase display a distinctive motor axonopathy. Neurology 1999;53:1239–1246.

    PubMed  CAS  Google Scholar 

  146. Flood DG, Reaume AG, Gruner JA, et al. Hindlimb motor neurons require Cu/Zn superoxide dismutase for maintenance of neuromuscular junctions. Am J Pathol 1999;155:663–672.

    PubMed  CAS  Google Scholar 

  147. Xia X, Zhou H, Huang Y, Xu Z. Allele-specific RNAi selectively silences mutant SOD1 and achieves significant therapeutic benefit in vivo. Neurobiol Dis 2006;23:578–586.

    PubMed  CAS  Google Scholar 

  148. Geng CM, Ding HL. Design of functional small interfering RNAs targeting amyotrophic lateral sclerosis-associated mutant alleles. Chin Med J (Engl) 2011;124:106–110.

    CAS  Google Scholar 

  149. Del Bo R, Tiloca C, Pensato V, et al. Novel optineurin mutations in patients with familial and sporadic amyotrophic lateral sclerosis. J Neurol Neurosurg Psychiatry 2011;82:1239–1243.

    PubMed  Google Scholar 

  150. Lee EB, Lee VM, Trojanowski JQ, Neumann M. TDP-43 immunoreactivity in anoxic, ischemic and neoplastic lesions of the central nervous system. Acta Neuropathol 2008;115:305–311.

    PubMed  CAS  Google Scholar 

  151. Yokoseki A, Shiga A, Tan CF, et al. TDP-43 mutation in familial amyotrophic lateral sclerosis. Ann Neurol 2008;63:538–542.

    PubMed  CAS  Google Scholar 

  152. Gitcho MA, Baloh RH, Chakraverty S, et al. TDP-43 A315T mutation in familial motor neuron disease. Ann Neurol 2008;63:535–538.

    PubMed  CAS  Google Scholar 

  153. Kabashi E, Valdmanis PN, Dion P, et al. TARDBP mutations in individuals with sporadic and familial amyotrophic lateral sclerosis. Nat Genet 2008;40:572–574.

    PubMed  CAS  Google Scholar 

  154. Winton MJ, Van Deerlin VM, Kwong LK, et al. A90V TDP-43 variant results in the aberrant localization of TDP-43 in vitro. FEBS Lett 2008;582:2252–2256.

    PubMed  CAS  Google Scholar 

  155. Van Deerlin VM, Leverenz JB, Bekris LM, et al. TARDBP mutations in amyotrophic lateral sclerosis with TDP-43 neuropathology: a genetic and histopathological analysis. Lancet Neurol 2008;7:409–416.

    PubMed  Google Scholar 

  156. Rutherford NJ, Zhang YJ, Baker M, et al. Novel mutations in TARDBP (TDP-43) in patients with familial amyotrophic lateral sclerosis. PLoS Genet 2008;4:e1000193.

    PubMed  Google Scholar 

  157. Kwiatkowski TJ, Jr., Bosco DA, Leclerc AL, et al. Mutations in the FUS/TLS gene on chromosome 16 cause familial amyotrophic lateral sclerosis. Science 2009;323:1205–1208.

    PubMed  CAS  Google Scholar 

  158. Vance C, Rogelj B, Hortobagyi T, et al. Mutations in FUS, an RNA processing protein, cause familial amyotrophic lateral sclerosis type 6. Science 2009;323:1208–1211.

    PubMed  CAS  Google Scholar 

  159. Abalkhail H, Mitchell J, Habgood J, Orrell R, de Belleroche J. A new familial amyotrophic lateral sclerosis locus on chromosome 16q12.1-16q12.2. Am J Hum Genet 2003;73:383–389.

    PubMed  CAS  Google Scholar 

  160. Ruddy DM, Parton MJ, Al-Chalabi A, et al. Two families with familial amyotrophic lateral sclerosis are linked to a novel locus on chromosome 16q. Am J Hum Genet 2003;73:390–396.

    PubMed  CAS  Google Scholar 

  161. Yamasaki S, Ivanov P, Hu GF, Anderson P. Angiogenin cleaves tRNA and promotes stress-induced translational repression. J Cell Biol 2009;185:35–42.

    PubMed  CAS  Google Scholar 

  162. Greenway MJ, Andersen PM, Russ C, et al. ANG mutations segregate with familial and 'sporadic' amyotrophic lateral sclerosis. Nat Genet 2006;38:411–413.

    PubMed  CAS  Google Scholar 

  163. Gellera C, Colombrita C, Ticozzi N, et al. Identification of new ANG gene mutations in a large cohort of Italian patients with amyotrophic lateral sclerosis. Neurogenetics 2008;9:33–40.

    PubMed  CAS  Google Scholar 

  164. Corrado L, Battistini S, Penco S, et al. Variations in the coding and regulatory sequences of the angiogenin (ANG) gene are not associated to ALS (amyotrophic lateral sclerosis) in the Italian population. J Neurol Sci 2007;258:123–127.

    PubMed  CAS  Google Scholar 

  165. Lambrechts D, Storkebaum E, Morimoto M, et al. VEGF is a modifier of amyotrophic lateral sclerosis in mice and humans and protects motoneurons against ischemic death. Nat Genet 2003;34:383–394.

    PubMed  CAS  Google Scholar 

  166. Matilla A, Roberson ED, Banfi S, et al. Mice lacking ataxin-1 display learning deficits and decreased hippocampal paired-pulse facilitation. J Neurosci 1998;18:5508–5516.

    PubMed  CAS  Google Scholar 

  167. Switonski PM, Fiszer A, Kazmierska K, et al. Mouse ataxin-3 functional knock-out model. Neuromolecular Med 2011;13:54–65.

    PubMed  CAS  Google Scholar 

  168. Stiles DK, Zhang Z, Ge P, et al. Widespread suppression of huntingtin with convection-enhanced delivery of siRNA. Exp Neurol 2012;233:463–471.

    PubMed  CAS  Google Scholar 

  169. McCormack AL, Mak SK, Henderson JM, et al. Alpha-synuclein suppression by targeted small interfering RNA in the primate substantia nigra. PLoS One 2010;5:e12122.

    PubMed  Google Scholar 

  170. Sarin H, Kanevsky AS, Wu H, et al. Effective transvascular delivery of nanoparticles across the blood–brain tumor barrier into malignant glioma cells. J Transl Med 2008;6:80.

    PubMed  Google Scholar 

  171. Weber W, Fux C, Daoud-el Baba M, et al. Macrolide-based transgene control in mammalian cells and mice. Nat Biotechnol 2002;20:901–907.

    PubMed  CAS  Google Scholar 

Download references

Acknowledgments

We thank members of the Davidson Lab for critical comments, and the NIH, NAF, HDF, and Roy J Carver Trust for funding this work.

Required Author Forms

Disclosure forms provided by the authors are available with the online version of this article.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Beverly L. Davidson.

Additional information

P. S. Ramachandran and M. S. Keiser contributed equally to this work.

Electronic supplementary material

Below is the link to the electronic supplementary material.

ESM 1

(PDF 499 kb)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Ramachandran, P.S., Keiser, M.S. & Davidson, B.L. Recent Advances in RNA Interference Therapeutics for CNS Diseases. Neurotherapeutics 10, 473–485 (2013). https://doi.org/10.1007/s13311-013-0183-8

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13311-013-0183-8

Keywords

Navigation