Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Diseases of Unstable Repeat Expansion: Mechanisms and Common Principles

Key Points

  • Diseases of unstable repeat expansion are a diverse group of disorders that are caused by the expansion of trinucleotide, tetranucleotide, or pentanucleotide repeat sequences. The position of the repeat sequence determines whether the pathogenic mechanism leads to loss of protein function, altered or enhanced function, or abnormal RNA–protein interactions.

  • Disorders that are caused by loss-of-function mechanisms include the neurodevelopmental fragile X disorders FRAXA and FRAXE and the degenerative disorder Friedreich ataxia. In these disorders repeat expansion results in transcriptional silencing and loss of the gene product.

  • Understanding the roles of FMRP in FRAXA (translational regulation at the synapse), FMR2 in FRAXE (transcription/signalling), and frataxin in FRDA (mitochondrial protein) is providing an insight into pathogenesis and avenues for therapy.

  • Repeat expansion disorders that are caused by altered protein function — the polyglutamine diseases — include several spinocerebellar ataxias, Huntington disease, spinal and bulbar muscular atrophy, and dentatorubral-pallidoluysian atrophy. These disorders share many points of pathogenic convergence, such as protein misfolding and accumulation, but also unique aspects that are determined by the normal function of the disease protein.

  • Mounting evidence highlights the importance of protein context — the protein sequence in which the polygutamine expansion occurs — as a key determinant of pathogenesis, with modulation of function being driven by the expanded polyglutamine tract.

  • More recently, altered RNA function and/or interactions have been recognized as pathogenic mechanisms in repeat expansion disorders. These alterations underlie dystrophia myotonica 1 and 2 (DM1 and DM2) and fragile X tremor/ataxia syndrome.

  • In DM1 and DM2, the evidence indicates that expanded RNA transcripts lead to the dysregulation of specific RNA-binding proteins; this results in aberrant splicing of several transcripts and a broad, multi-systemic phenotype.

  • In addition to the commonalities within classes of expansion disorders certain similarities also exist between classes. These include importance of normal protein function in protein-mediated disorders, and evidence of protein misfolding in diseases that are caused by both expanded protein and RNA.

  • The convergence of both RNA and protein-mediated diseases on components of the protein quality-control machinery indicates that repeat expansions, whether in the RNA or protein, could elicit a common cellular-response mechanism.

  • Increasing insight into pathogenic mechanisms in these disorders is providing the framework for developing targeted therapies. The late onset of many of the neurodegenerative repeat expansion disorders provides a window of therapeutic opportunity.

Abstract

The list of developmental and degenerative diseases that are caused by expansion of unstable repeats continues to grow, and is now approaching 20 disorders. The pathogenic mechanisms that underlie these disorders involve either loss of protein function or gain of function at the protein or RNA level. Common themes have emerged within and between these different classes of disease; for example, among disorders that are caused by gain-of-function mechanisms, altered protein conformations are central to pathogenesis, leading to changes in protein activity or abundance. In all these diseases, the context of the expanded repeat and the abundance, subcellular localization and interactions of the proteins and RNAs that are affected have key roles in disease-specific phenotypes.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Mechanisms of pathogenesis in fragile X syndrome.
Figure 2: Molecular and biochemical basis of Friedreich ataxia.
Figure 3: Mechanisms of pathogenesis in polyglutamine diseases.
Figure 4: Mechanisms of pathogenesis in dystrophia myotonica 1 and 2.

Similar content being viewed by others

References

  1. Verkerk, A. J. et al. Identification of a gene (FMR-1) containing a CGG repeat coincident with a breakpoint cluster region exhibiting length variation in fragile X syndrome. Cell 65, 905–914 (1991).

    Article  CAS  PubMed  Google Scholar 

  2. La Spada, A. R., Wilson, E. M., Lubahn, D. B., Harding, A. E. & Fischbeck, K. H. Androgen receptor gene mutations in X-linked spinal and bulbar muscular atrophy. Nature 352, 77–79 (1991).

    Article  CAS  PubMed  Google Scholar 

  3. Liquori, C. L. et al. Myotonic dystrophy type 2 caused by a CCTG expansion in intron 1 of ZNF9. Science 293, 864–867 (2001).

    Article  CAS  PubMed  Google Scholar 

  4. Matsuura, T. et al. Large expansion of the ATTCT pentanucleotide repeat in spinocerebellar ataxia type 10. Nature Genet. 26, 191–194 (2000).

    Article  CAS  PubMed  Google Scholar 

  5. Bennetto, L. & Pennington, B. F. in Fragile X Syndrome Diagnosis, Treatment, and Research (eds Hagerman, R. J. & Cronister, A) 210–248 (Johns Hopkins Univ. Press, Baltimore, 1996).

    Google Scholar 

  6. Merenstein, S. A. et al. Molecular-clinical correlations in males with an expanded FMR1 mutation. Am. J. Med. Genet. 64, 388–394 (1996).

    Article  CAS  PubMed  Google Scholar 

  7. Warren, S. T. & Sherman, S. L. in The Metabolic and Molecular Basis of Inherited Disease Vol. 8 (eds Scriver, C. R. & Sly, W. S.) 1257–1289 (McGraw Hill, New York, 2001).

    Google Scholar 

  8. Hagerman, R. J. et al. Intention tremor, parkinsonism, and generalized brain atrophy in male carriers of fragile X. Neurology 57, 127–130 (2001).

    Article  CAS  PubMed  Google Scholar 

  9. Corbin, F. et al. The fragile X mental retardation protein is associated with poly(A)+ mRNA in actively translating polyribosomes. Hum. Mol. Genet. 6, 1465–1472 (1997).

    Article  CAS  PubMed  Google Scholar 

  10. Li, Z. et al. The fragile X mental retardation protein inhibits translation via interacting with mRNA. Nucleic Acids Res. 29, 2276–2283 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Laggerbauer, B., Ostareck, D., Keidel, E. M., Ostareck-Lederer, A. & Fischer, U. Evidence that fragile X mental retardation protein is a negative regulator of translation. Hum. Mol. Genet. 10, 329–338 (2001).

    Article  CAS  PubMed  Google Scholar 

  12. De Boulle, K. et al. A point mutation in the FMR-1 gene associated with fragile X mental retardation. Nature Genet. 3, 31–35 (1993).

    Article  CAS  PubMed  Google Scholar 

  13. Feng, Y. et al. FMRP associates with polyribosomes as an mRNP, and the I304N mutation of severe fragile X syndrome abolishes this association. Mol. Cell 1, 109–118 (1997).

    Article  CAS  PubMed  Google Scholar 

  14. Bear, M. F., Huber, K. M. & Warren, S. T. The mGluR theory of fragile X mental retardation. Trends Neurosci. 27, 370–377 (2004).

    Article  CAS  PubMed  Google Scholar 

  15. Huber, K. M., Gallagher, S. M., Warren, S. T. & Bear, M. F. Altered synaptic plasticity in a mouse model of fragile X mental retardation. Proc. Natl Acad. Sci. USA 99, 7746–7750 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Irwin, S. A. et al. Abnormal dendritic spine characteristics in the temporal and visual cortices of patients with fragile-X syndrome: a quantitative examination. Am. J. Med. Genet. 98, 161–167 (2001).

    Article  CAS  PubMed  Google Scholar 

  17. McBride, S. M. et al. Pharmacological rescue of synaptic plasticity, courtship behavior, and mushroom body defects in a Drosophila model of fragile X syndrome. Neuron 45, 753–764 (2005). The investigators use a Drosophila model of fragile X syndrome to show the therapeutic benefit of treatment with mGluR antagonists, providing strong in vivo support of the mGluR theory of fragile X syndrome.

    Article  CAS  PubMed  Google Scholar 

  18. Brown, V. et al. Microarray identification of FMRP-associated brain mRNAs and altered mRNA translational profiles in fragile X syndrome. Cell 107, 477–487 (2001).

    Article  CAS  PubMed  Google Scholar 

  19. Darnell, J. C. et al. Fragile X mental retardation protein targets G quartet mRNAs important for neuronal function. Cell 107, 489–499 (2001). References 18 and 19 describe innovative approaches to identify FMRP targets.

    Article  CAS  PubMed  Google Scholar 

  20. Schaeffer, C. et al. The fragile X mental retardation protein binds specifically to its mRNA via a purine quartet motif. EMBO J. 20, 4803–4813 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Zhang, Y. Q. et al. Drosophila fragile X-related gene regulates the MAP1B homolog Futsch to control synaptic structure and function. Cell 107, 591–603 (2001).

    Article  CAS  PubMed  Google Scholar 

  22. Xu, K. et al. The fragile X-related gene affects the crawling behavior of Drosophila larvae by regulating the mRNA level of the DEG/ENaC protein pickpocket1. Curr. Biol. 14, 1025–1034 (2004).

    Article  CAS  PubMed  Google Scholar 

  23. Lee, A. et al. Control of dendritic development by the Drosophila fragile X-related gene involves the small GTPase Rac1. Development 130, 5543–5552 (2003).

    Article  CAS  PubMed  Google Scholar 

  24. Darnell, J. C. et al. Kissing complex RNAs mediate interaction between the fragile-X mental retardation protein KH2 domain and brain polyribosomes. Genes Dev. 19, 903–918 (2005). The authors identified an RNA target motif for the KH2 domain of FMRP (termed the FMRP kissing complex), showing that the interaction of FMRP with brain polyribosomes occurs through its association with this motif rather than G-quartet RNA motifs.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Caudy, A. A., Myers, M., Hannon, G. J. & Hammond, S. M. Fragile X-related protein and VIG associate with the RNA interference machinery. Genes Dev. 16, 2491–2496 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Ishizuka, A., Siomi, M. C. & Siomi, H. A Drosophila fragile X protein interacts with components of RNAi and ribosomal proteins. Genes Dev. 16, 2497–2508 (2002). References 25 and 26 provide some of the first evidence for an association of Drosophila FMRP with the RNAi machinery, supporting the importance of this association for FMRP's role in translational control.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Jin, P. et al. Biochemical and genetic interaction between the fragile X mental retardation protein and the microRNA pathway. Nature Neurosci. 7, 113–117 (2004).

    Article  CAS  PubMed  Google Scholar 

  28. Jin, P., Alisch, R. S. & Warren, S. T. RNA and microRNAs in fragile X mental retardation. Nature Cell Biol. 6, 1048–1053 (2004).

    Article  CAS  PubMed  Google Scholar 

  29. Mulley, J. C. et al. FRAXE and mental retardation. J. Med. Genet. 32, 162–169 (1995).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Knight, S. J. et al. Trinucleotide repeat amplification and hypermethylation of a CpG island in FRAXE mental retardation. Cell 74, 127–134 (1993).

    Article  CAS  PubMed  Google Scholar 

  31. Gecz, J., Gedeon, A. K., Sutherland, G. R. & Mulley, J. C. Identification of the gene FMR2, associated with FRAXE mental retardation. Nature Genet. 13, 105–108 (1996).

    Article  CAS  PubMed  Google Scholar 

  32. Gu, Y., Shen, Y., Gibbs, R. A. & Nelson, D. L. Identification of FMR2, a novel gene associated with the FRAXE CCG repeat and CpG island. Nature Genet. 13, 109–113 (1996).

    Article  CAS  PubMed  Google Scholar 

  33. Nilson, I. et al. Exon/intron structure of the human AF-4 gene, a member of the AF-4/LAF-4/FMR-2 gene family coding for a nuclear protein with structural alterations in acute leukaemia. Br. J. Haematol. 98, 157–169 (1997).

    Article  CAS  PubMed  Google Scholar 

  34. Gecz, J., Bielby, S., Sutherland, G. R. & Mulley, J. C. Gene structure and subcellular localization of FMR2, a member of a new family of putative transcription activators. Genomics 44, 201–213 (1997).

    Article  CAS  PubMed  Google Scholar 

  35. Miller, W. J., Skinner, J. A., Foss, G. S. & Davies, K. E. Localization of the fragile X mental retardation 2 (FMR2) protein in mammalian brain. Eur. J. Neurosci. 12, 381–384 (2000).

    Article  CAS  PubMed  Google Scholar 

  36. Gu, Y. et al. Impaired conditioned fear and enhanced long-term potentiation in Fmr2 knock-out mice. J. Neurosci. 22, 2753–2763 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Su, M. A., Wisotzkey, R. G. & Newfeld, S. J. A screen for modifiers of decapentaplegic mutant phenotypes identifies lilliputian, the only member of the Fragile-X/Burkitt's Lymphoma family of transcription factors in Drosophila melanogaster. Genetics 157, 717–725 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Wittwer, F., van der Straten, A., Keleman, K., Dickson, B. J. & Hafen, E. Lilliputian: an AF4/FMR2-related protein that controls cell identity and cell growth. Development 128, 791–800 (2001).

    CAS  PubMed  Google Scholar 

  39. Campuzano, V. et al. Friedreich's ataxia: autosomal recessive disease caused by an intronic GAA triplet repeat expansion. Science 271, 1423–1427 (1996).

    Article  CAS  PubMed  Google Scholar 

  40. Campuzano, V. et al. Frataxin is reduced in Friedreich ataxia patients and is associated with mitochondrial membranes. Hum. Mol. Genet. 6, 1771–1780 (1997).

    Article  CAS  PubMed  Google Scholar 

  41. Babcock, M. et al. Regulation of mitochondrial iron accumulation by Yfh1p, a putative homolog of frataxin. Science 276, 1709–1712 (1997).

    Article  CAS  PubMed  Google Scholar 

  42. Foury, F. & Cazzalini, O. Deletion of the yeast homologue of the human gene associated with Friedreich's ataxia elicits iron accumulation in mitochondria. FEBS Lett. 411, 373–377 (1997).

    Article  CAS  PubMed  Google Scholar 

  43. Foury, F. Low iron concentration and aconitase deficiency in a yeast frataxin homologue deficient strain. FEBS Lett. 456, 281–284 (1999).

    Article  CAS  PubMed  Google Scholar 

  44. Muhlenhoff, U., Richhardt, N., Ristow, M., Kispal, G. & Lill, R. The yeast frataxin homolog Yfh1p plays a specific role in the maturation of cellular Fe/S proteins. Hum. Mol. Genet. 11, 2025–2036 (2002).

    Article  PubMed  Google Scholar 

  45. Yoon, T. & Cowan, J. A. Iron-sulfur cluster biosynthesis. Characterization of frataxin as an iron donor for assembly of [2Fe-2S] clusters in ISU-type proteins. J. Am. Chem. Soc. 125, 6078–6084 (2003).

    Article  CAS  PubMed  Google Scholar 

  46. Yoon, T. & Cowan, J. A. Frataxin-mediated iron delivery to ferrochelatase in the final step of heme biosynthesis. J. Biol. Chem. 279, 25943–25946 (2004).

    Article  CAS  PubMed  Google Scholar 

  47. Delatycki, M. B. et al. Direct evidence that mitochondrial iron accumulation occurs in Friedreich ataxia. Ann. Neurol. 45, 673–675 (1999).

    Article  CAS  PubMed  Google Scholar 

  48. Lamarche, J. B., Shapcott, D., Côté, M. & Lemieux, B. in Handbook of Cerebellar Diseases (ed. Lechtenberg, R.) 453–458 (Marcel Dekker, New York, 1993).

    Google Scholar 

  49. Rotig, A. et al. Aconitase and mitochondrial iron-sulphur protein deficiency in Friedreich ataxia. Nature Genet. 17, 215–217 (1997).

    Article  CAS  PubMed  Google Scholar 

  50. Cossee, M. et al. Inactivation of the Friedreich ataxia mouse gene leads to early embryonic lethality without iron accumulation. Hum. Mol. Genet. 9, 1219–1226 (2000).

    Article  CAS  PubMed  Google Scholar 

  51. Puccio, H. et al. Mouse models for Friedreich ataxia exhibit cardiomyopathy, sensory nerve defect and Fe–S enzyme deficiency followed by intramitochondrial iron deposits. Nature Genet. 27, 181–186 (2001). This study reports the generation and characterization of conditional FRDA mouse models, which recapitulate many features of the human disease. Importantly, it also provides in vivo evidence that mitochondrial iron accumulation only takes place well after the onset of pathology.

    Article  CAS  PubMed  Google Scholar 

  52. Voncken, M., Ioannou, P. & Delatycki, M. B. Friedreich ataxia-update on pathogenesis and possible therapies. Neurogenetics 5, 1–8 (2004).

    Article  PubMed  Google Scholar 

  53. Wong, A. et al. The Friedreich's ataxia mutation confers cellular sensitivity to oxidant stress which is rescued by chelators of iron and calcium and inhibitors of apoptosis. Hum. Mol. Genet. 8, 425–430 (1999).

    Article  CAS  PubMed  Google Scholar 

  54. Chantrel-Groussard, K. et al. Disabled early recruitment of antioxidant defenses in Friedreich's ataxia. Hum. Mol. Genet. 10, 2061–2067 (2001).

    Article  CAS  PubMed  Google Scholar 

  55. Emond, M., Lepage, G., Vanasse, M. & Pandolfo, M. Increased levels of plasma malondialdehyde in Friedreich ataxia. Neurology 55, 1752–1753 (2000).

    Article  CAS  PubMed  Google Scholar 

  56. Seznec, H. et al. Friedreich ataxia: the oxidative stress paradox. Hum. Mol. Genet. 14, 463–474 (2005).

    Article  CAS  PubMed  Google Scholar 

  57. Zoghbi, H. Y. & Orr, H. T. Glutamine repeats and neurodegeneration. Annu. Rev. Neurosci. 23, 217–247 (2000).

    Article  CAS  PubMed  Google Scholar 

  58. Nakamura, K. et al. SCA17, a novel autosomal dominant cerebellar ataxia caused by an expanded polyglutamine in TATA-binding protein. Hum. Mol. Genet. 10, 1441–1448 (2001).

    Article  CAS  PubMed  Google Scholar 

  59. Watase, K. et al. A long CAG repeat in the mouse Sca1 locus replicates SCA1 features and reveals the impact of protein solubility on selective neurodegeneration. Neuron 34, 905–919 (2002).

    Article  CAS  PubMed  Google Scholar 

  60. Yoo, S. Y. et al. SCA7 knockin mice model human SCA7 and reveal gradual accumulation of mutant ataxin-7 in neurons and abnormalities in short-term plasticity. Neuron 37, 383–401 (2003).

    Article  CAS  PubMed  Google Scholar 

  61. Matilla, A. et al. Mice lacking ataxin-1 display learning deficits and decreased hippocampal paired-pulse facilitation. J. Neurosci. 18, 5508–5516 (1998).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Dragatsis, I., Levine, M. S. & Zeitlin, S. Inactivation of Hdh in the brain and testis results in progressive neurodegeneration and sterility in mice. Nature Genet. 26, 300–306 (2000).

    Article  CAS  PubMed  Google Scholar 

  63. Zuccato, C. et al. Loss of huntingtin-mediated BDNF gene transcription in Huntington's disease. Science 293, 493–498 (2001).

    Article  CAS  PubMed  Google Scholar 

  64. Ciechanover, A. & Brundin, P. The ubiquitin proteasome system in neurodegenerative diseases: sometimes the chicken, sometimes the egg. Neuron 40, 427–446 (2003).

    Article  CAS  PubMed  Google Scholar 

  65. Chan, H. Y., Warrick, J. M., Gray-Board, G. L., Paulson, H. L. & Bonini, N. M. Mechanisms of chaperone suppression of polyglutamine disease: selectivity, synergy and modulation of protein solubility in Drosophila. Hum. Mol. Genet. 9, 2811–2820 (2000).

    Article  CAS  PubMed  Google Scholar 

  66. Cummings, C. J. et al. Over-expression of inducible HSP70 chaperone suppresses neuropathology and improves motor function in SCA1 mice. Hum. Mol. Genet. 10, 1511–1518 (2001).

    Article  CAS  PubMed  Google Scholar 

  67. Venkatraman, P., Wetzel, R., Tanaka, M., Nukina, N. & Goldberg, A. L. Eukaryotic proteasomes cannot digest polyglutamine sequences and release them during degradation of polyglutamine-containing proteins. Mol. Cell 14, 95–104 (2004).

    Article  CAS  PubMed  Google Scholar 

  68. Bence, N. F., Sampat, R. M. & Kopito, R. R. Impairment of the ubiquitin–proteasome system by protein aggregation. Science 292, 1552–1555 (2001).

    Article  CAS  PubMed  Google Scholar 

  69. Bennett, E. J., Bence, N. F., Jayakumar, R. & Kopito, R. R. Global impairment of the ubiquitin–proteasome system by nuclear or cytoplasmic protein aggregates precedes inclusion body formation. Mol. Cell 17, 351–365 (2005).

    Article  CAS  PubMed  Google Scholar 

  70. Verhoef, L. G., Lindsten, K., Masucci, M. G. & Dantuma, N. P. Aggregate formation inhibits proteasomal degradation of polyglutamine proteins. Hum. Mol. Genet. 11, 2689–2700 (2002).

    Article  CAS  PubMed  Google Scholar 

  71. Michalik, A. & Van Broeckhoven, C. Proteasome degrades soluble expanded polyglutamine completely and efficiently. Neurobiol. Dis. 16, 202–211 (2004).

    Article  CAS  PubMed  Google Scholar 

  72. Kaytor, M. D., Wilkinson, K. D. & Warren, S. T. Modulating huntingtin half-life alters polyglutamine-dependent aggregate formation and cell toxicity. J. Neurochem. 89, 962–973 (2004).

    Article  CAS  PubMed  Google Scholar 

  73. Bowman, A. B., Yoo, S. Y., Dantuma, N. P. & Zoghbi, H. Y. Neuronal dysfunction in a polyglutamine disease model occurs in the absence of ubiquitin–proteasome system impairment and inversely correlates with the degree of nuclear inclusion formation. Hum. Mol. Genet. 14, 679–691 (2005). The first in vivo study of UPS activity in vulnerable neurons, arguing against a role for UPS impairment in polyglutamine diseases and providing evidence for a protective role of inclusions.

    Article  CAS  PubMed  Google Scholar 

  74. Arrasate, M., Mitra, S., Schweitzer, E. S., Segal, M. R. & Finkbeiner, S. Inclusion body formation reduces levels of mutant huntingtin and the risk of neuronal death. Nature 431, 805–810 (2004). This is an elegant in vitro study showing that inclusions have a protective role against pathology induced by an N-terminal fragment of huntingtin carrying a polyglutamine repeat expansion.

    Article  CAS  PubMed  Google Scholar 

  75. Slow, E. J. et al. Absence of behavioral abnormalities and neurodegeneration in vivo despite widespread neuronal huntingtin inclusions. Proc. Natl Acad. Sci. USA 102, 11402–11407 (2005). Using a transgenic HD mouse model that expresses a short fragment of huntingtin, this study provides in vivo evidence against the toxicity of inclusions and emphasizes the importance of protein context in pathogenesis.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Ravikumar, B. et al. Inhibition of mTOR induces autophagy and reduces toxicity of polyglutamine expansions in fly and mouse models of Huntington disease. Nature Genet. 36, 585–595 (2004).

    Article  CAS  PubMed  Google Scholar 

  77. Jiang, H., Nucifora, F. C. Jr, Ross, C. A. & DeFranco, D. B. Cell death triggered by polyglutamine-expanded huntingtin in a neuronal cell line is associated with degradation of CREB-binding protein. Hum. Mol. Genet. 12, 1–12 (2003).

    Article  PubMed  Google Scholar 

  78. Tsuda, H. et al. The AXH domain in mammalian/Drosophila ataxin-1 mediates neurodegeneration in spinocerebellar ataxia 1 through its interaction with Gfi-1/Senseless proteins. Cell 122, 1–12 (2005). This study uses genetics and biochemistry to highlight the importance of domains that are outside the polyglutamine tract as key determinants of pathogenesis. It also suggests that the polyglutamine tract exerts toxicity by modulating the function and interactions of the wild-type protein.

    Article  CAS  Google Scholar 

  79. Klement, I. A. et al. Ataxin-1 nuclear localization and aggregation: role in polyglutamine-induced disease in SCA1 transgenic mice. Cell 95, 41–53 (1998).

    Article  CAS  PubMed  Google Scholar 

  80. Lin, X., Antalffy, B., Kang, D., Orr, H. T. & Zoghbi, H. Y. Polyglutamine expansion down-regulates specific neuronal genes before pathologic changes in SCA1. Nature Neurosci. 3, 157–163 (2000).

    Article  CAS  PubMed  Google Scholar 

  81. Nucifora, F. C. Jr et al. Interference by huntingtin and atrophin-1 with CBP-mediated transcription leading to cellular toxicity. Science 291, 2423–2428 (2001).

    Article  CAS  PubMed  Google Scholar 

  82. Steffan, J. S. et al. Histone deacetylase inhibitors arrest polyglutamine-dependent neurodegeneration in Drosophila. Nature 413, 739–743 (2001).

    Article  CAS  PubMed  Google Scholar 

  83. Dunah, A. W. et al. Sp1 and TAFII130 transcriptional activity disrupted in early Huntington's disease. Science 296, 2238–2243 (2002).

    Article  CAS  PubMed  Google Scholar 

  84. Li, S. H. et al. Interaction of Huntington disease protein with transcriptional activator Sp1. Mol. Cell. Biol. 22, 1277–1287 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Steffan, J. S. et al. The Huntington's disease protein interacts with p53 and CREB-binding protein and represses transcription. Proc. Natl Acad. Sci. USA 97, 6763–6768 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. McCampbell, A. et al. CREB-binding protein sequestration by expanded polyglutamine. Hum. Mol. Genet. 9, 2197–2202 (2000).

    Article  CAS  PubMed  Google Scholar 

  87. Yu, Z. X., Li, S. H., Nguyen, H. P. & Li, X. J. Huntingtin inclusions do not deplete polyglutamine-containing transcription factors in HD mice. Hum. Mol. Genet. 11, 905–914 (2002).

    Article  CAS  PubMed  Google Scholar 

  88. Schaffar, G. et al. Cellular toxicity of polyglutamine expansion proteins: mechanism of transcription factor deactivation. Mol. Cell. 15, 95–105 (2004).

    Article  CAS  PubMed  Google Scholar 

  89. Boutell, J. M. et al. Aberrant interactions of transcriptional repressor proteins with the Huntington's disease gene product, huntingtin. Hum. Mol. Genet. 8, 1647–1655 (1999).

    Article  CAS  PubMed  Google Scholar 

  90. Zhang, S., Xu, L., Lee, J. & Xu, T. Drosophila atrophin homolog functions as a transcriptional corepressor in multiple developmental processes. Cell 108, 45–56 (2002).

    Article  CAS  PubMed  Google Scholar 

  91. Tsai, C. C. et al. Ataxin 1, a SCA1 neurodegenerative disorder protein, is functionally linked to the silencing mediator of retinoid and thyroid hormone receptors. Proc. Natl Acad. Sci. USA 101, 4047–4052 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Palhan, V. B. et al. Polyglutamine-expanded ataxin-7 inhibits STAGA histone acetyltransferase activity to produce retinal degeneration. Proc. Natl Acad. Sci. USA 102, 8472–8477 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. McMahon, S. J., Pray-Grant, M. G., Schieltz, D., Yates, J. R. 3rd & Grant, P. A. Polyglutamine-expanded spinocerebellar ataxia-7 protein disrupts normal SAGA and SLIK histone acetyltransferase activity. Proc. Natl Acad. Sci. USA 102, 8478–8482 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  94. McCampbell, A. et al. Histone deacetylase inhibitors reduce polyglutamine toxicity. Proc. Natl Acad. Sci. USA 98, 15179–15184 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Hockly, E. et al. Suberoylanilide hydroxamic acid, a histone deacetylase inhibitor, ameliorates motor deficits in a mouse model of Huntington's disease. Proc. Natl Acad. Sci. USA 100, 2041–2046 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Ferrante, R. J. et al. Histone deacetylase inhibition by sodium butyrate chemotherapy ameliorates the neurodegenerative phenotype in Huntington's disease mice. J. Neurosci. 23, 9418–9427 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Minamiyama, M. et al. Sodium butyrate ameliorates phenotypic expression in a transgenic mouse model of spinal and bulbar muscular atrophy. Hum. Mol. Genet. 13, 1183–1192 (2004). References 94–97 provide evidence for pathogenic polyglutamine-dependent chromatin remodelling and the potential therapeutic use of histone deacetylase inhibitors to reverse this.

    Article  CAS  PubMed  Google Scholar 

  98. Fernandez-Funez, P. et al. Identification of genes that modify ataxin-1-induced neurodegeneration. Nature 408, 101–106 (2000).

    Article  CAS  PubMed  Google Scholar 

  99. Shibata, H., Huynh, D. P. & Pulst, S. M. A novel protein with RNA-binding motifs interacts with ataxin-2. Hum. Mol. Genet. 9, 1303–1313 (2000).

    Article  CAS  PubMed  Google Scholar 

  100. Gunawardena, S. et al. Disruption of axonal transport by loss of huntingtin or expression of pathogenic polyQ proteins in Drosophila. Neuron 40, 25–40 (2003).

    Article  CAS  PubMed  Google Scholar 

  101. Szebenyi, G. et al. Neuropathogenic forms of huntingtin and androgen receptor inhibit fast axonal transport. Neuron 40, 41–52 (2003).

    Article  CAS  PubMed  Google Scholar 

  102. Panov, A. V. et al. Early mitochondrial calcium defects in Huntington's disease are a direct effect of polyglutamines. Nature Neurosci. 5, 731–736 (2002).

    Article  CAS  PubMed  Google Scholar 

  103. Tang, T. S. et al. Disturbed Ca2+ signaling and apoptosis of medium spiny neurons in Huntington's disease. Proc. Natl Acad. Sci. USA 102, 2602–2607 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Chen, H. K. et al. Interaction of Akt-phosphorylated ataxin-1 with 14-3-3 mediates neurodegeneration in spinocerebellar ataxia type 1. Cell 113, 457–468 (2003).

    Article  CAS  PubMed  Google Scholar 

  105. Emamian, E. S. et al. Serine 776 of ataxin-1 is critical for polyglutamine-induced disease in SCA1 transgenic mice. Neuron 38, 375–387 (2003).

    Article  CAS  PubMed  Google Scholar 

  106. Steffan, J. S. et al. SUMO modification of huntingtin and Huntington's disease pathology. Science 304, 100–104 (2004). References 104–106 highlight the importance of modifications of the host protein in polyglutamine pathogenesis.

    Article  CAS  PubMed  Google Scholar 

  107. Frontali, M. Spinocerebellar ataxia type 6: channelopathy or glutamine repeat disorder? Brain Res. Bull. 56, 227–231 (2001).

    Article  CAS  PubMed  Google Scholar 

  108. Warrick, J. M. et al. Ataxin-3 suppresses polyglutamine neurodegeneration in Drosophila by a ubiquitin-associated mechanism. Mol. Cell 18, 37–48 (2005).

    Article  CAS  PubMed  Google Scholar 

  109. Schmidt, B. J., Greenberg, C. R., Allingham-Hawkins, D. J. & Spriggs, E. L. Expression of X-linked bulbospinal muscular atrophy (Kennedy disease) in two homozygous women. Neurology 59, 770–772 (2002).

    Article  PubMed  Google Scholar 

  110. Takeyama, K. et al. Androgen-dependent neurodegeneration by polyglutamine-expanded human androgen receptor in Drosophila. Neuron 35, 855–864 (2002).

    Article  CAS  PubMed  Google Scholar 

  111. Katsuno, M. et al. Testosterone reduction prevents phenotypic expression in a transgenic mouse model of spinal and bulbar muscular atrophy. Neuron 35, 843–854 (2002).

    Article  CAS  PubMed  Google Scholar 

  112. Brook, J. D. et al. Molecular basis of myotonic dystrophy: expansion of a trinucleotide (CTG) repeat at the 3′ end of a transcript encoding a protein kinase family member. Cell 69, 385 (1992).

  113. Fu, Y. H. et al. An unstable triplet repeat in a gene related to myotonic muscular dystrophy. Science 255, 1256–1258 (1992).

    Article  CAS  PubMed  Google Scholar 

  114. Mahadevan, M. et al. Myotonic dystrophy mutation: an unstable CTG repeat in the 3′ untranslated region of the gene. Science 255, 1253–1255 (1992).

    Article  CAS  PubMed  Google Scholar 

  115. Ranum, L. P., Rasmussen, P. F., Benzow, K. A., Koob, M. D. & Day, J. W. Genetic mapping of a second myotonic dystrophy locus. Nature Genet. 19, 196–198 (1998).

    Article  CAS  PubMed  Google Scholar 

  116. Harper, P. Myotonic Dystrophy Vol. 37 (W.B. Saunders, London, 2001).

    Google Scholar 

  117. Jansen, G. et al. Abnormal myotonic dystrophy protein kinase levels produce only mild myopathy in mice. Nature Genet. 13, 316–324 (1996).

    Article  CAS  PubMed  Google Scholar 

  118. Reddy, S. et al. Mice lacking the myotonic dystrophy protein kinase develop a late onset progressive myopathy. Nature Genet. 13, 325–335 (1996).

    Article  CAS  PubMed  Google Scholar 

  119. Klesert, T. R. et al. Mice deficient in Six5 develop cataracts: implications for myotonic dystrophy. Nature Genet. 25, 105–109 (2000).

    Article  CAS  PubMed  Google Scholar 

  120. Ranum, L. P. & Day, J. W. Pathogenic RNA repeats: an expanding role in genetic disease. Trends Genet. 20, 506–512 (2004).

    Article  CAS  PubMed  Google Scholar 

  121. Mankodi, A. et al. Myotonic dystrophy in transgenic mice expressing an expanded CUG repeat. Science 289, 1769–1773 (2000).

    Article  CAS  PubMed  Google Scholar 

  122. Mankodi, A. et al. Expanded CUG repeats trigger aberrant splicing of ClC-1 chloride channel pre-mRNA and hyperexcitability of skeletal muscle in myotonic dystrophy. Mol. Cell 10, 35–44 (2002).

    Article  CAS  PubMed  Google Scholar 

  123. Fardaei, M. et al. Three proteins, MBNL, MBLL and MBXL, co-localize in vivo with nuclear foci of expanded-repeat transcripts in DM1 and DM2 cells. Hum. Mol. Genet. 11, 805–814 (2002).

    Article  CAS  PubMed  Google Scholar 

  124. Miller, J. W. et al. Recruitment of human muscleblind proteins to (CUG)n expansions associated with myotonic dystrophy. EMBO J. 19, 4439–4448 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Kanadia, R. N. et al. A muscleblind knockout model for myotonic dystrophy. Science 302, 1978–1980 (2003). The disruption of the mouse Mbnl1 gene reported in this study replicates many of the splicing defects and disease phenotypes of DM1, which clearly implicates the MBNL family in pathogenesis.

    Article  CAS  PubMed  Google Scholar 

  126. Timchenko, N. A. et al. RNA CUG repeats sequester CUGBP1 and alter protein levels and activity of CUGBP1. J. Biol. Chem. 276, 7820–7826 (2001).

    Article  CAS  PubMed  Google Scholar 

  127. Ho, T. H., Bundman, D., Armstrong, D. L. & Cooper, T. A. Transgenic mice expressing CUG-BP1 reproduce splicing mis-regulation observed in myotonic dystrophy. Hum. Mol. Genet. 14, 1539–1547 (2005). This is an elegant in vitro study that provides evidence that formation of CUG RNA foci in DM1 is separable from MBNL-mediated splicing alterations, and therefore might not be the primary determinant of pathogenesis.

    Article  CAS  PubMed  Google Scholar 

  128. Philips, A. V., Timchenko, L. T. & Cooper, T. A. Disruption of splicing regulated by a CUG-binding protein in myotonic dystrophy. Science 280, 737–741 (1998).

    Article  CAS  PubMed  Google Scholar 

  129. Savkur, R. S., Philips, A. V. & Cooper, T. A. Aberrant regulation of insulin receptor alternative splicing is associated with insulin resistance in myotonic dystrophy. Nature Genet. 29, 40–47 (2001).

    Article  CAS  PubMed  Google Scholar 

  130. Charlet, B. N. et al. Loss of the muscle-specific chloride channel in type 1 myotonic dystrophy due to misregulated alternative splicing. Mol. Cell 10, 45–53 (2002).

    Article  Google Scholar 

  131. Jiang, H., Mankodi, A., Swanson, M. S., Moxley, R. T. & Thornton, C. A. Myotonic dystrophy type 1 is associated with nuclear foci of mutant RNA, sequestration of muscleblind proteins and deregulated alternative splicing in neurons. Hum. Mol. Genet. 13, 3079–3088 (2004). This paper provides some of the first evidence that the RNA-mediated pathogenesis in dystrophia myotonica extends to the CNS through effects on splicing. It also highlights the potential involvement of protein misfolding in diseases that are mediated by aberrant RNA–protein interactions.

    Article  CAS  PubMed  Google Scholar 

  132. Ho, T. H. et al. Muscleblind proteins regulate alternative splicing. EMBO J. 23, 3103–3112 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Ebralidze, A., Wang, Y., Petkova, V., Ebralidse, K. & Junghans, R. P. RNA leaching of transcription factors disrupts transcription in myotonic dystrophy. Science 303, 383–387 (2004).

    Article  CAS  PubMed  Google Scholar 

  134. Di Prospero, N. A. & Fischbeck, K. H. Therapeutics development for triplet repeat expansion diseases. Nature Rev. Genet. 6, 756–765 (2005).

    Article  CAS  PubMed  Google Scholar 

  135. Jacquemont, S. et al. Fragile X premutation tremor/ataxia syndrome: molecular, clinical, and neuroimaging correlates. Am. J. Hum. Genet. 72, 869–878 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Greco, C. M. et al. Neuronal intranuclear inclusions in a new cerebellar tremor/ataxia syndrome among fragile X carriers. Brain 125, 1760–1771 (2002).

    Article  CAS  PubMed  Google Scholar 

  137. Tassone, F. et al. Elevated levels of FMR1 mRNA in carrier males: a new mechanism of involvement in the fragile-X syndrome. Am. J. Hum. Genet. 66, 6–15 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Tassone, F., Hagerman, R. J., Chamberlain, W. D. & Hagerman, P. J. Transcription of the FMR1 gene in individuals with fragile X syndrome. Am. J. Med. Genet. 97, 195–203 (2000).

    Article  CAS  PubMed  Google Scholar 

  139. Primerano, B. et al. Reduced FMR1 mRNA translation efficiency in fragile X patients with premutations. RNA 8, 1482–1488 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  140. Willemsen, R. et al. The FMR1 CGG repeat mouse displays ubiquitin-positive intranuclear neuronal inclusions; implications for the cerebellar tremor/ataxia syndrome. Hum. Mol. Genet. 12, 949–959 (2003).

    Article  CAS  PubMed  Google Scholar 

  141. Van Dam, D. et al. Cognitive decline, neuromotor and behavioural disturbances in a mouse model for fragile-X-associated tremor/ataxia syndrome (FXTAS). Behav. Brain Res. 162, 233–239 (2005).

    Article  CAS  PubMed  Google Scholar 

  142. Jin, P. et al. RNA-mediated neurodegeneration caused by the fragile X premutation rCGG repeats in Drosophila. Neuron 39, 739–747 (2003).

    Article  CAS  PubMed  Google Scholar 

  143. Ho, T. H. et al. Colocalization of muscleblind with RNA foci is separable from mis-regulation of alternative splicing in myotonic dystrophy. J. Cell Sci. 118, 2923–2933 (2005).

    Article  CAS  PubMed  Google Scholar 

  144. Koob, M. D. et al. An untranslated CTG expansion causes a novel form of spinocerebellar ataxia (SCA8). Nature Genet. 21, 379–384 (1999).

    Article  CAS  PubMed  Google Scholar 

  145. Nemes, J. P., Benzow, K. A., Moseley, M. L., Ranum, L. P. & Koob, M. D. The SCA8 transcript is an antisense RNA to a brain-specific transcript encoding a novel actin-binding protein (KLHL1). Hum. Mol. Genet. 9, 1543–1551 (2000).

    Article  CAS  PubMed  Google Scholar 

  146. Mutsuddi, M., Marshall, C. M., Benzow, K. A., Koob, M. D. & Rebay, I. The spinocerebellar ataxia 8 noncoding RNA causes neurodegeneration and associates with staufen in Drosophila. Curr. Biol. 14, 302–308 (2004).

    Article  CAS  PubMed  Google Scholar 

  147. Holmes, S. E. et al. Expansion of a novel CAG trinucleotide repeat in the 5′ region of PPP2R2B is associated with SCA12. Nature Genet. 23, 391–392 (1999).

    Article  CAS  PubMed  Google Scholar 

  148. Holmes, S. E., O'Hearn, E. & Margolis, R. L. Why is SCA12 different from other SCAs? Cytogenet. Genome Res. 100, 189–197 (2003).

    Article  CAS  PubMed  Google Scholar 

  149. Holmes, S. E. et al. A repeat expansion in the gene encoding junctophilin-3 is associated with Huntington disease-like 2. Nature Genet. 29, 377–378 (2001).

    Article  CAS  PubMed  Google Scholar 

  150. Zu, T. et al. Recovery from polyglutamine-induced neurodegeneration in conditional SCA1 transgenic mice. J. Neurosci. 24, 8853–8861 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Yamamoto, A., Lucas, J. J. & Hen, R. Reversal of neuropathology and motor dysfunction in a conditional model of Huntington's disease. Cell 101, 57–66 (2000).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors wish to thank members of the Zoghbi laboratory and Thai Ho (T. Cooper's laboratory) for critical comments. The authors apologize to anyone whose work they did not cite owing to space constraints. The citations are representative works, and are not inclusive of the many important contributing works for a given topic.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Huda Y. Zoghbi.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Related links

Related links

DATABASES

Entrez Gene

AGO2

AR

ATXN7

CREBBP

DMPK

FMR1

FMR2

futsch

FXN

lilliputian

MAP1B

pickpocket

rac1

SP1

TAF4

TBP

TP53

OMIM

dentatorubral-pallidoluysian atrophy

dystrophia myotonica 1

dystrophia myotonica 2

fragile X syndrome

Friedreich ataxia

Huntington disease

Huntington disease-like 2

spinal and bulbar muscular atrophy

spinocerebellar ataxia 1

Glossary

POLYRIBOSOME

A string of 80S ribosomes that are bound to an mRNA molecule.

DENDRITIC SPINES

Mushroom-shaped structures on neuronal dendrites that receive synaptic input and have postsynaptic densities. Changes in spine shape are thought to be important for modulating synaptic strength.

LONG-TERM DEPRESSION

A long-lasting decrease in synaptic strength to below baseline levels. It is generally induced in the hippocampus by repetitive low-frequency stimulation. It is thought to result from changes in postsynaptic receptor density, although changes in presynaptic release might also have a role.

GROUP 1 mGluRs

Postsynaptic G-protein coupled receptors for the excitatory neurotransmitter glutamate. These receptors have many roles in synaptic plasticity.

SYNAPTIC PRUNING

The selective elimination of weak synapses during brain development.

SYNAPTIC PLASTICITY

The capacity for alterations in synaptic connections between neurons, including changes in the nature, strength or number of interneuronal connections, which subserves learning and memory.

G QUARTET

A secondary structure observed in G-rich RNA in which four consecutive guanosine residues bind to each other to form an intramolecular stem-loop structure.

RNA-INDUCED SILENCING COMPLEX

A multi-protein complex that functions in RNAi pathways to mediate the cleavage of target RNAs.

ARGONAUTE PROTEINS

A family of proteins that are essential for diverse RNA-silencing pathways.

DICER PROTEINS

A highly conserved family of RNaseIII enzymes that mediate dsRNA cleavage. This produces the small RNAs that direct target silencing in RNAi pathways.

MICRORNA

A form of ssRNA typically 20–25 nucleotides long that is thought to regulate the expression of other genes, either through inhibiting protein translation or degrading a target mRNA transcript through a process that is similar to RNA interference.

CpG ISLANDS

Sequences of at least 200 bp that contain both a GC content that is greater than 50% and a high CpG frequency.

FERRITIN

An iron-storage protein.

ATAXIA

The inability to coordinate movement.

ACONITASE

An enzyme that catalyses the reversible hydration of cis-aconitase to yield citrate or isocitrate. Aconitases are involved in the citric acid cycle.

IRON–SULPHUR (FE–S) CLUSTERS

Co-factor-like species that are common in nature and are involved in many cellular processes, such as electron transfer, catalysis, gene regulation and the sensing of iron and oxygen.

UBIQUITIN–PROTEASOME SYSTEM

A key pathway that is involved in the degradation of intracellular proteins, especially of short-lived regulatory proteins. The system uses an enzymatic cascade that involves conjugation of multiple ubiquitin moieties to a protein substrate and degradation of the tagged protein by the 26S proteasome complex.

PURKINJE CELLS

The neurons that convey the output signals of the cerebellar cortex.

AUTOPHAGY

An intracellular pathway that leads to bulk protein degradation and involves the sequestering of cytosol into vesicles for delivery to a degradative organelle.

AXOPLASM

The intracellular fluid of the axon that facilitates vesicle transport.

CASPASE

One of a family of proteases, activated by proteolytic cleavage, that have an important role in apoptosis and inflammation.

CALPAINS

A family of highly conserved and widely expressed cysteine proteases that are activated by a rise in intracellular calcium levels. They have an important role in apoptosis and in modulating intracellular signals in response to stressful conditions.

SUMOYLATION

The post-translational modification of proteins that involves the covalent attachment of small ubiquitin-like modifier (SUMO) and regulates the interactions of those proteins with other macromolecules.

P/Q-TYPE CALCIUM CHANNEL

A high voltage activated calcium channel found in cerebellar Purkinje cells that is responsible for most of the calcium current in these neurons.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Gatchel, J., Zoghbi, H. Diseases of Unstable Repeat Expansion: Mechanisms and Common Principles. Nat Rev Genet 6, 743–755 (2005). https://doi.org/10.1038/nrg1691

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrg1691

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing