Skip to main navigation menu Skip to main content Skip to site footer

Review article: Biomedical intelligence

Vol. 145 No. 3738 (2015)

New and emerging roles of small RNAs in neurodegeneration, muscle, cardiovascular and inflammatory diseases

  • Marian Hruska-Plochan
  • Bei Li
  • Diego Kyburz
  • Jan Krützfeldt
  • Ulf Landmesser
  • Adriano Aguzzi
  • Magdalini Polymenidou
DOI
https://doi.org/10.4414/smw.2015.14192
Cite this as:
Swiss Med Wkly. 2015;145:w14192
Published
06.09.2015

Summary

Small noncoding RNAs (snRNAs) were discovered more than two decades ago, yet it was not until relatively recently that their important role in genome regulation was recognised. With such a substantial role in genome regulation, it is not surprising that snRNAs are crucial contributors to an ever-increasing number of diseases, as evidenced by the long list of published studies. Currently, microRNAs (miRNAs) represent the most intensively studied snRNAs. Dysregulation of miRNAs has been confirmed in numerous diseases, and changes in their levels could play an essential role in disease onset and progression and could be used for prognosis and potential therapy. Indeed, disease-altered miRNAs may either signify a direct trigger or a consequence of the disease. Therefore, miRNAs represent unique targets for disease intervention through their down- or up-regulation. Importantly, miRNAs may facilitate disease monitoring by detection of disease-altered miRNAs in easily accessible bodily fluids, such as blood or cerebrospinal fluid. Therefore, study of these events is of utmost importance for understanding the molecular mechanisms that drive disease, as well as for diagnosis and therapy. Here we attempted to synthesise a large number of studies to highlight the crucial role of miRNAs in the pathogenesis of neurodegenerative, muscle, cardiovascular and inflammatory diseases.

References

  1. Lander ES, et al, Initial sequencing and analysis of the human genome. Nature. 2001;409(6822):p.860–921.
  2. Venter JC, et al. The sequence of the human genome. Science. 2001;291(5507):p.1304–51.
  3. Derrien T, et al. The GENCODE v7 catalog of human long noncoding RNAs: Analysis of their gene structure, evolution, and expression. Genome Research. 2012;22(9):p.1775–89.
  4. Berretta JMA. Pervasive transcription constitutes a new level of eukaryotic genome regulation. EMBO reports. 2009;10(9):p.973–82.
  5. Clark MB, et al. The Reality of Pervasive Transcription. PLoS Biol. 2011;9(7):p.e1000625.
  6. Jacquier A. The complex eukaryotic transcriptome: unexpected pervasive transcription and novel small RNAs. Nat Rev Genet. 2009;10(12):p.833–44.
  7. Carthew RW. Sontheimer EJ. Origins and Mechanisms of miRNAs and siRNAs. Cell. 2009;136(4) p.642–55.
  8. Barry, G., Integrating the roles of long and small non-coding RNA in brain function and disease. Mol Psychiatry, 2014;19(4):p.410–6.
  9. Ng S-Y, et al. Long noncoding RNAs in development and disease of the central nervous system. Trends in Genetics, 2013;29(8):p.461–8.
  10. Lee RC, Feinbaum RL, Ambros V. The C. elegans heterochronic gene lin-4 encodes small RNAs with antisense complementarity to lin-14. Cell. 1993;75(5):p.843–54.
  11. Geisler S. Coller J. RNA in unexpected places: long non-coding RNA functions in diverse cellular contexts. Nat Rev Mol Cell Biol. 2013;14(11):p.699–712.
  12. Keniry A, et al. The H19 lincRNA is a developmental reservoir of miR-675 that suppresses growth and Igf1r. Nat Cell Biol. 2012;14(7):p.659–65.
  13. Yoon J-H, Abdelmohsen K, Gorospe M. Functional interactions among microRNAs and long noncoding RNAs. Seminars in Cell & Developmental Biology. (0).
  14. Friedlander M, et al. Evidence for the biogenesis of more than 1,000 novel human microRNAs. Genome Biology, 2014;15(4):p.R57.
  15. Liu C, et al. Effects of genetic variations on microRNA:target interactions. Nucleic Acids Research. 2014.
  16. Marti E, et al. A myriad of miRNA variants in control and Huntington's disease brain regions detected by massively parallel sequencing. Nucleic Acids Res. 2010;38(20): p. 7219–35.
  17. Salta E. De Strooper B. Non-coding RNAs with essential roles in neurodegenerative disorders. Lancet Neurol. 2012;11(2):p.189–200.
  18. Bhattacharya A, Ziebarth JD, Cui Y. PolymiRTS Database 3.0: linking polymorphisms in microRNAs and their target sites with human diseases and biological pathways. Nucleic Acids Research. 2013.
  19. Qureshi IA, Mehler MF. Emerging roles of non-coding RNAs in brain evolution, development, plasticity and disease. Nat Rev Neurosci. 2012;13(8):p.528–41.
  20. Bravo JA, Dinan TG. MicroRNAs: a novel therapeutic target for schizophrenia. Curr Pharm Des. 2011;17(2):p.176–88.
  21. Stenvang J, et al. Inhibition of microRNA function by antimiR oligonucleotides. Silence. 2012;3(1):p.1.
  22. Ross CA, Poirier MA. Protein aggregation and neurodegenerative disease. Nat Med. 2004;10Suppl:p.S10–7.
  23. Bernstein E, et al. Dicer is essential for mouse development. Nat Genet. 2003;35(3):p.215–7.
  24. Davis TH, et al. Conditional loss of Dicer disrupts cellular and tissue morphogenesis in the cortex and hippocampus. J Neurosci. 2008;28(17):p.4322–30.
  25. Shin D, et al. Dicer ablation in oligodendrocytes provokes neuronal impairment in mice. Annals of Neurology. 2009;66(6):p.843–57.
  26. Christensen M, Schratt GM. microRNA involvement in developmental and functional aspects of the nervous system and in neurological diseases. Neuroscience Letters. 2009;466(2):p.55–62.
  27. Makeyev EV, et al. The MicroRNA miR-124 Promotes Neuronal Differentiation by Triggering Brain-Specific Alternative Pre-mRNA Splicing. Molecular Cell. 2007;27(3):p.435–48.
  28. Rajasethupathy P, et al. A Role for Neuronal piRNAs in the Epigenetic Control of Memory-Related Synaptic Plasticity. Cell. 2012;149(3):p.693–707.
  29. Rogelj B, et al. Contextual fear conditioning regulates the expression of brain-specific small nucleolar RNAs in hippocampus. European Journal of Neuroscience. 2003;18(11):p. 3089–96.
  30. Bertram L, Tanzi RE. Thirty years of Alzheimer's disease genetics: the implications of systematic meta-analyses. Nat Rev Neurosci. 2008;9(10):p.768–78.
  31. Mayeux R, Stern Y. Epidemiology of Alzheimer disease. Cold Spring Harb Perspect Med. 2012;2(8).
  32. Bateman RJ, et al. Clinical and biomarker changes in dominantly inherited Alzheimer's disease. N Engl J Med. 2012;367(9):p.795–804.
  33. Bertram L, Lill CM, Tanzi RE. The genetics of Alzheimer disease: back to the future. Neuron. 2010;68(2):p.270–81.
  34. Huang Y, Mucke L. Alzheimer mechanisms and therapeutic strategies. Cell. 2012;148(6):p. 04–22.
  35. Nunez-Iglesias J, et al. Joint genome-wide profiling of miRNA and mRNA expression in Alzheimer's disease cortex reveals altered miRNA regulation. PLoS One. 2010;5(2):p.e8898.
  36. Cogswell JP, et al. Identification of miRNA changes in Alzheimer's disease brain and CSF yields putative biomarkers and insights into disease pathways. J Alzheimers Dis. 2008;14(1):p.27–41.
  37. Wang WX, et al. The expression of microRNA miR-107 decreases early in Alzheimer's disease and may accelerate disease progression through regulation of beta-site amyloid precursor protein-cleaving enzyme 1. J Neurosci. 2008;28(5):p.1213–23.
  38. Liu CG, et al. MicroRNA-384 regulates both amyloid precursor protein and beta-secretase expression and is a potential biomarker for Alzheimer's disease. Int J Mol Med. 2014;34(1):p.160–6.
  39. Vilardo E, et al. MicroRNA-101 regulates amyloid precursor protein expression in hippocampal neurons. J Biol Chem. 2010;285(24):p.18344–51.
  40. Long JM, Lahiri DK. MicroRNA-101 downregulates Alzheimer's amyloid-beta precursor protein levels in human cell cultures and is differentially expressed. Biochem Biophys Res Commun. 2011;404(4):p.889–95.
  41. Delay C, et al. Alzheimer-specific variants in the 3'UTR of Amyloid precursor protein affect microRNA function. Mol Neurodegener. 2011;6:p.70.
  42. Smith P, et al. In vivo regulation of amyloid precursor protein neuronal splicing by microRNAs. J Neurochem. 2011;116(2): p. 240–7.
  43. Alexandrov PN, et al. microRNA (miRNA) speciation in Alzheimer's disease (AD) cerebrospinal fluid (CSF) and extracellular fluid (ECF). Int J Biochem Mol Biol. 2012;3(4):p.365–73.
  44. Nakasa T, et al. The inhibitory effect of microRNA-146a expression on bone destruction in collagen-induced arthritis. Arthritis Rheum. 2011;63(6):p.1582–90.
  45. Thai TH, et al. Regulation of the germinal center response by microRNA-155. Science. 2007;316(5824):p.604–8.
  46. Bala S, et al. Up-regulation of microRNA-155 in macrophages contributes to increased tumor necrosis factor {alpha} (TNF{alpha}) production via increased mRNA half-life in alcoholic liver disease. J Biol Chem. 2011;286(2):p.1436–44.
  47. Boldin MP, et al. miR-146a is a significant brake on autoimmunity, myeloproliferation, and cancer in mice. J Exp Med. 2011;208(6):p.1189–201.
  48. Zhu HC, et al. MicroRNA-195 downregulates Alzheimer's disease amyloid-beta production by targeting BACE1. Brain Res Bull. 2012;88(6):p.596–601.
  49. Boissonneault V, et al. MicroRNA-298 and microRNA-328 regulate expression of mouse beta-amyloid precursor protein-converting enzyme 1. J Biol Chem. 2009;284(4):p.1971–81.
  50. Hebert SS, et al. Loss of microRNA cluster miR-29a/b-1 in sporadic Alzheimer's disease correlates with increased BACE1/beta-secretase expression. Proc Natl Acad Sci U S A. 2008;105(17):p.6415–20.
  51. Fang M, et al. The miR-124 regulates the expression of BACE1/beta-secretase correlated with cell death in Alzheimer's disease. Toxicol Lett. 2012;209(1):p.94–105.
  52. Zong Y. et al. miR-29c regulates BACE1 protein expression. Brain Res. 2011;1395:p.108–15.
  53. Shioya M, et al. Aberrant microRNA expression in the brains of neurodegenerative diseases: miR-29a decreased in Alzheimer disease brains targets neurone navigator 3. Neuropathol Appl Neurobiol. 2010;36(4):p.320–30.
  54. Schipper HM, et al. MicroRNA expression in Alzheimer blood mononuclear cells. Gene Regul Syst Bio. 2007.1:p.263–74.
  55. Dickson JR, et al. Alternative polyadenylation and miR-34 family members regulate tau expression. J Neurochem. 2013;127(6):p.739–49.
  56. Absalon S, et al. MiR-26b, upregulated in Alzheimer's disease, activates cell cycle entry, tau-phosphorylation, and apoptosis in postmitotic neurons. J Neurosci. 2013;33(37):p.14645–59.
  57. Lau P, et al. Alteration of the microRNA network during the progression of Alzheimer's disease. EMBO Mol Med. 2013.5(10):p.1613–34.
  58. Lukiw WJ. Micro-RNA speciation in fetal, adult and Alzheimer's disease hippocampus. Neuroreport. 2007;18(3):p.297–300.
  59. Kumar P, et al. Circulating miRNA biomarkers for Alzheimer's disease. PLoS One. 2013;8(7):p.e69807.
  60. Shtilbans A, Henchcliffe C. Biomarkers in Parkinson's disease: an update. Curr Opin Neurol. 2012;25(4):p.460–5.
  61. Dawson TM Dawson VL, Molecular pathways of neurodegeneration in Parkinson's disease. Science. 2003;302(5646):p.819–22.
  62. Spillantini MG, et al. Alpha-synuclein in Lewy bodies. Nature. 1997;388(6645):p.839–40.
  63. Saiki S, Sato S, Hattori N. Molecular pathogenesis of Parkinson's disease: update. J Neurol Neurosurg Psychiatry. 2012;83(4):p.430–6.
  64. Gasser T. Molecular pathogenesis of Parkinson disease: insights from genetic studies. Expert Rev Mol Med. 2009;11:p.e22.
  65. Davis GC, et al. Chronic Parkinsonism secondary to intravenous injection of meperidine analogues. Psychiatry Res. 1979;1(3):p.249–54.
  66. Langston JW, et al. Chronic Parkinsonism in humans due to a product of meperidine-analog synthesis. Science. 1983;219(4587):p.979–80.
  67. Langston JW Ballard Jr. PA, Parkinson's disease in a chemist working with 1-methyl-4-phenyl-1,2,5,6-tetrahydropyridine. N Engl J Med. 1983;309(5):p.310.
  68. Dolezalova D, et al. Pig models of neurodegenerative disorders: Utilization in cell replacement-based preclinical safety and efficacy studies. J Comp Neurol. 2014;522(12):p.2784–801.
  69. Kim J, et al. A MicroRNA feedback circuit in midbrain dopamine neurons. Science. 2007;317(5842):p.1220–4.
  70. Junn E, et al. Repression of alpha-synuclein expression and toxicity by microRNA-7. Proc Natl Acad Sci U S A. 2009;106(31):p.13052–7.
  71. Doxakis E. Post-transcriptional regulation of alpha-synuclein expression by mir-7 and mir-153. J Biol Chem. 2010;285(17):p.12726–34.
  72. Gehrke S, et al. Pathogenic LRRK2 negatively regulates microRNA-mediated translational repression. Nature. 2010;466(7306):p.637–41.
  73. Cho HJ, et al. MicroRNA-205 regulates the expression of Parkinson's disease-related leucine-rich repeat kinase 2 protein. Hum Mol Genet. 201322(3):p.608–20.
  74. Wang G, et al. Variation in the miRNA-433 binding site of FGF20 confers risk for Parkinson disease by overexpression of alpha-synuclein. Am J Hum Genet. 2008;82(2):p.283–9.
  75. van der Walt JM, et al. Fibroblast growth factor 20 polymorphisms and haplotypes strongly influence risk of Parkinson disease. Am J Hum Genet. 2004;74(6):p.1121–7.
  76. de Mena L, et al. FGF20 rs12720208 SNP and microRNA-433 variation: no association with Parkinson's disease in Spanish patients. Neurosci Lett. 2010;479(1):p.22–5.
  77. Wider C, et al. FGF20 and Parkinson's disease: no evidence of association or pathogenicity via alpha-synuclein expression. Mov Disord. 2009;24(3):p.455–9.
  78. Margis R, Margis R, Rieder CR. Identification of blood microRNAs associated to Parkinsonis disease. J Biotechnol. 2011;152(3):p.96–101.
  79. Cardo LF, et al. Profile of microRNAs in the plasma of Parkinson's disease patients and healthy controls. J Neurol. 2013;260(5):p.1420–2.
  80. Aguzzi A, Nuvolone M, Zhu C. The immunobiology of prion diseases. Nat Rev Immunol. 2013;13(12):p.888–902.
  81. Weissmann C. Birth of a prion: Spontaneous generation revisited. Cell. 2005;122(2):p.165–8.
  82. Greenwood AD, et al. Cell line dependent RNA expression profiles of prion-infected mouse neuronal cells. J Mol Biol. 2005;349(3):p.487–500.
  83. Julius C, et al. Transcriptional stability of cultured cells upon prion infection. J Mol Biol. 2008;375(5):p.1222–33.
  84. Mahal SP, et al. Prion strain discrimination in cell culture: The cell panel assay. Proceedings of the National Academy of Sciences of the United States of America. 2007;104(52):p.20908–13.
  85. Saba R, et al. A miRNA signature of prion induced neurodegeneration. PLoS One. 2008;3(11):p.e3652.
  86. Bueler H, et al. Mice Devoid of Prp Are Resistant to Scrapie. Cell. 1993;73(7):p.1339–47.
  87. Manson JC, et al. Prp Gene Dosage Determines the Timing but Not the Final Intensity or Distribution of Lesions in Scrapie Pathology. Neurodegeneration. 1994;3(4):p.331–40.
  88. Daude N, Marella M, Chabry J. Specific inhibition of pathological prion protein accumulation by small interfering RNAs. J Cell Sci. 2003;116(Pt 13):p.2775–9.
  89. Pulford B, et al. Liposome-siRNA-peptide complexes cross the blood-brain barrier and significantly decrease PrP on neuronal cells and PrP in infected cell cultures. PLoS One. 2010;5(6):p.e11085.
  90. Tilly G, et al. Efficient and specific down-regulation of prion protein expression by RNAi. Biochem Biophys Res Commun. 2003;305(3):p.548–51.
  91. Pfeifer A, et al. Lentivector-mediated RNAi efficiently suppresses prion protein and prolongs survival of scrapie-infected mice. J Clin Invest. 2006;116(12):p. 3204–10.
  92. White MD, et al. Single treatment with RNAi against prion protein rescues early neuronal dysfunction and prolongs survival in mice with prion disease. Proc Natl Acad Sci U S A. 2008;105(29):p.10238–43.
  93. Golding MC, et al. Suppression of prion protein in livestock by RNA interference. Proc Natl Acad Sci U S A. 2006;103(14):p.5285–90.
  94. Wongsrikeao P. et al. Combination of the somatic cell nuclear transfer method and RNAi technology for the production of a prion gene-knockdown calf using plasmid vectors harboring the U6 or tRNA promoter. Prion. 2011;5(1):p.39–46.
  95. Kang SG, et al. Establishment and characterization of Prnp knockdown neuroblastoma cells using dual microRNA-mediated RNA interference. Prion. 2011;5(2):p.93–102.
  96. Nazor Friberg K, et al. Intracerebral Infusion of Antisense Oligonucleotides Into Prion-infected Mice. Mol Ther Nucleic Acids. 2012;1:p.e9.
  97. Weiss S, et al. RNA aptamers specifically interact with the prion protein PrP. J Virol. 1997;71(11):p.8790–7.
  98. Proske D, et al. Prion-protein-specific aptamer reduces PrPSc formation. Chembiochem. 2002;3(8):p.717–25.
  99. Karapetyan YE, et al. Unique drug screening approach for prion diseases identifies tacrolimus and astemizole as antiprion agents. Proc Natl Acad Sci U S A. 2013;110(17):p.7044–9.
  100. Silber BM, et al. Novel compounds lowering the cellular isoform of the human prion protein in cultured human cells. Bioorg Med Chem. 2014;22(6):p.1960–72.
  101. Mohr S, Bakal C, Perrimon N. Genomic screening with RNAi: results and challenges. Annu Rev Biochem. 2010;79:p.37–64.
  102. Prusiner SB, et al. Measurement of the Scrapie Agent Using an Incubation-Time Interval Assay. Annals of Neurology. 1982;11(4):p.3538.
  103. Klohn PC, et al. A quantitative, highly sensitive cell-based infectivity assay for mouse scrapie prions. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(20):p.11666–71.
  104. Rothstein JD. Current hypotheses for the underlying biology of amyotrophic lateral sclerosis. Ann Neurol. 2009;65:Suppl1p.S3–9.
  105. Boillee S, Vande Velde C, Cleveland DW. ALS: a disease of motor neurons and their nonneuronal neighbors. Neuron. 2006;52(1):p.39–59.
  106. Robberecht W, Philips T. The changing scene of amyotrophic lateral sclerosis. Nat Rev Neurosci. 2013;14(4):p.248–64.
  107. Kiernan MC, et al. Amyotrophic lateral sclerosis. Lancet. 2011;377(9769):p.942–55.
  108. Pasinelli P, Brown RH. Molecular biology of amyotrophic lateral sclerosis: insights from genetics. Nat Rev Neurosci. 2006;7(9):p.710–23.
  109. Logroscino G, et al. Incidence of amyotrophic lateral sclerosis in southern Italy: a population based study. J Neurol Neurosurg Psychiatry, 2005;76(8):p.1094–8.
  110. Haidet-Phillips AM, et al. Astrocytes from familial and sporadic ALS patients are toxic to motor neurons. Nat Biotechnol. 2011;29(9):p.824–8.
  111. Gitcho MA, et al. TDP-43 A315T mutation in familial motor neuron disease. Ann Neurol. 2008;63(4):p.535–8.
  112. Sreedharan J, et al. TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis. Science. 2008;319(5870):p.1668–72.
  113. Kabashi E, et al. TARDBP mutations in individuals with sporadic and familial amyotrophic lateral sclerosis. Nat Genet. 2008;40(5):p.572–4.
  114. Yokoseki A, et al. TDP-43 mutation in familial amyotrophic lateral sclerosis. Ann Neurol. 2008;63(4):p.538–42.
  115. Van Deerlin VM, et al. TARDBP mutations in amyotrophic lateral sclerosis with TDP-43 neuropathology: a genetic and histopathological analysis. The Lancet Neurology. 2008;7(5):p.409–16.
  116. Kwiatkowski Jr. TJ, et al, Mutations in the FUS/TLS gene on chromosome 16 cause familial amyotrophic lateral sclerosis. Science. 2009;323(5918):p.1205–8.
  117. Vance C, et al. Mutations in FUS, an RNA processing protein, cause familial amyotrophic lateral sclerosis type 6. Science. 2009;323(5918):p.1208–11.
  118. Arai T, et al. TDP-43 is a component of ubiquitin-positive tau-negative inclusions in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Biochem Biophys Res Commun. 2006;351(3):p.602–11.
  119. Neumann M, et al. Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science. 2006;314(5796):p.130–3.
  120. Lagier-Tourenne C, Polymenidou M, Cleveland DW. TDP-43 and FUS/TLS: emerging roles in RNA processing and neurodegeneration. Hum Mol Genet. 2010;19(R1):p.R46–64.
  121. Polymenidou M, et al. Misregulated RNA processing in amyotrophic lateral sclerosis. Brain Res. 2012;1462:p.3–15.
  122. DeJesus-Hernandez, M, et al. Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron. 2011;72(2):p.245–56.
  123. Gijselinck I, et al. A C9orf72 promoter repeat expansion in a Flanders-Belgian cohort with disorders of the frontotemporal lobar degeneration-amyotrophic lateral sclerosis spectrum: a gene identification study. Lancet Neurol. 2012;11(1):p.54–65.
  124. Renton A.E, et al. A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron. 2011;72(2):p.257–68.
  125. Laferriere F, Polymenidou M. Advances and challenges in understanding the multifaceted pathogenesis of amyotrophic lateral sclerosis. Swiss Med Wkly. 2015;145:p.w14054.
  126. Ling SC, Polymenidou M, Cleveland DW. Converging mechanisms in ALS and FTD: disrupted RNA and protein homeostasis. Neuron. 2013;79(3):p.416–38.
  127. Cruts M, et al. Current insights into the C9orf72 repeat expansion diseases of the FTLD/ALS spectrum. Trends Neurosci. 2013;36(8):p.450–9.
  128. Snowden JS, Neary D, Mann DM. Frontotemporal dementia. Br J Psychiatry. 2002;180:p.140–3.
  129. Sleegers K, Cruts M, Van Broeckhoven C. Molecular pathways of frontotemporal lobar degeneration. Annu Rev Neurosci. 2010;33:p.71–88.
  130. Rabinovici GD, Miller BL. Frontotemporal lobar degeneration: epidemiology, pathophysiology, diagnosis and management. CNS Drugs. 2010;24(5):p.375–98.
  131. Piguet O, et al. Behavioural-variant frontotemporal dementia: diagnosis, clinical staging, and management. Lancet Neurol. 2011;10(2):p.162–72.
  132. Haramati S. et al., miRNA malfunction causes spinal motor neuron disease. Proc Natl Acad Sci U S A. 2010;107(29):p.13111–6.
  133. Tao J, et al. Deletion of astroglial Dicer causes non-cell-autonomous neuronal dysfunction and degeneration. J Neurosci. 2011;31(22):p.8306–19.
  134. Zhang Z, et al. Downregulation of microRNA-9 in iPSC-derived neurons of FTD/ALS patients with TDP-43 mutations. PLoS One. 2013;8(10):p.e76055.
  135. Packer AN, et al. The bifunctional microRNA miR-9/miR-9* regulates REST and CoREST and is downregulated in Huntington's disease. J Neurosci. 2008;28(53):p.14341–6.
  136. Campos-Melo D, et al. Altered microRNA expression profile in Amyotrophic Lateral Sclerosis: a role in the regulation of NFL mRNA levels. Mol Brain. 2013;6:p.26.
  137. Ishtiaq M, et al. Analysis of novel NEFL mRNA targeting microRNAs in amyotrophic lateral sclerosis. PLoS One. 2014;9(1):p.e85653.
  138. Zhou FH, et al. miRNA-9 expression is upregulated in the spinal cord of G93A-SOD1 transgenic mice. International Journal of Clinical and Experimental Pathology. 2013;6(9):p.1826–38.
  139. Ling SC, et al. ALS-associated mutations in TDP-43 increase its stability and promote TDP-43 complexes with FUS/TLS. Proc Natl Acad Sci U S A. 2010;107(30):p.13318–23.
  140. Freibaum BD, et al. Global analysis of TDP-43 interacting proteins reveals strong association with RNA splicing and translation machinery. J Proteome Res. 2010;9(2):p.1104–20.
  141. Kawahara Y, Mieda-Sato A. TDP-43 promotes microRNA biogenesis as a component of the Drosha and Dicer complexes. Proc Natl Acad Sci U S A. 2012;109(9):p.3347–52.
  142. Buratti E, et al. Nuclear factor TDP-43 can affect selected microRNA levels. FEBS J. 2010;277(10):p.2268–81.
  143. Fan Z, Chen X, Chen R. Transcriptome-wide analysis of TDP-43 binding small RNAs identifies miR-NID1 (miR-8485), a novel miRNA that represses NRXN1 expression. Genomics. 2014;103(1):p.76–82.
  144. Tollervey JR, et al. Characterizing the RNA targets and position-dependent splicing regulation by TDP-43. Nat Neurosci. 2011;14(4): p. 452–8.
  145. Polymenidou M, et al. Long pre-mRNA depletion and RNA missplicing contribute to neuronal vulnerability from loss of TDP-43. Nat Neurosci. 2011;14(4):p.459–68.
  146. Mitchell JC, et al. Overexpression of human wild-type FUS causes progressive motor neuron degeneration in an age- and dose-dependent fashion. Acta Neuropathol. 2013;125(2):p.273–88.
  147. Hicks GG, et al. Fus deficiency in mice results in defective B-lymphocyte development and activation, high levels of chromosomal instability and perinatal death. Nat Genet. 2000;24(2):p.175–9.
  148. Dini Modigliani S, et al. An ALS-associated mutation in the FUS 3'-UTR disrupts a microRNA-FUS regulatory circuitry. Nat Commun. 2014;5 p 4335.
  149. Morlando M, et al. FUS stimulates microRNA biogenesis by facilitating co-transcriptional Drosha recruitment. EMBO J. 2012;31(24):p.4502–10.
  150. Williams AH, et al. MicroRNA-206 delays ALS progression and promotes regeneration of neuromuscular synapses in mice. Science. 2009;326(5959):p.1549–54.
  151. Russell AP, et al. Disruption of skeletal muscle mitochondrial network genes and miRNAs in amyotrophic lateral sclerosis. Neurobiology of Disease. 2013;49:p.107–17.
  152. Toivonen JM, et al. MicroRNA-206: a potential circulating biomarker candidate for amyotrophic lateral sclerosis. PLoS One. 2014;9(2):p.e89065.
  153. Gascon E, Gao FB. The emerging roles of microRNAs in the pathogenesis of frontotemporal dementia-amyotrophic lateral sclerosis (FTD-ALS) spectrum disorders. J Neurogenet. 2014;28(1–2):p.30–40.
  154. Baker M, et al. Mutations in progranulin cause tau-negative frontotemporal dementia linked to chromosome 17. Nature. 2006;442(7105):p.916–9.
  155. Cruts M, et al. Null mutations in progranulin cause ubiquitin-positive frontotemporal dementia linked to chromosome 17q21. Nature. 2006;442(7105):p.920-4.
  156. Rademakers R, et al. Common variation in the miR-659 binding-site of GRN is a major risk factor for TDP43-positive frontotemporal dementia. Hum Mol Genet. 2008;17(23):p.3631–42.
  157. Kocerha J, et al. Altered microRNA expression in frontotemporal lobar degeneration with TDP-43 pathology caused by progranulin mutations. BMC Genomics. 2011;12:p.527.
  158. Van Deerlin VM, et al. Common variants at 7p21 are associated with frontotemporal lobar degeneration with TDP-43 inclusions. Nat Genet. 2010;42(3):p.234–9.
  159. Chen-Plotkin AS, et al. TMEM106B, the risk gene for frontotemporal dementia, is regulated by the microRNA-132/212 cluster and affects progranulin pathways. J Neurosci. 2012;32(33):p.11213–27.
  160. Finch, N, et al. TMEM106B regulates progranulin levels and the penetrance of FTLD in GRN mutation carriers. Neurology. 2011;76(5):p.467–74.
  161. van Blitterswijk M, et al. TMEM106B protects C9ORF72 expansion carriers against frontotemporal dementia. Acta Neuropathol. 2014;127(3):p.397–406.
  162. Gallagher MD, et al. TMEM106B is a genetic modifier of frontotemporal lobar degeneration with C9orf72 hexanucleotide repeat expansions. Acta Neuropathol. 2014;127(3):p.407;18.
  163. Koval ED, et al. Method for widespread microRNA-155 inhibition prolongs survival in ALS-model mice. Hum Mol Genet. 2013;22(20):p.4127–35.
  164. Butovsky O, et al. Modulating inflammatory monocytes with a unique microRNA gene signature ameliorates murine ALS. J Clin Invest. 2012;122(9):p.3063–87.
  165. Cardoso AL, et al. miR-155 modulates microglia-mediated immune response by down-regulating SOCS-1 and promoting cytokine and nitric oxide production. Immunology. 2012;135(1):p.73–88.
  166. Gaughwin PM, et al. Hsa-miR-34b is a plasma-stable microRNA that is elevated in pre-manifest Huntington's disease. Hum Mol Genet. 2011;20(11):p.2225–37.
  167. De Felice B, et al. A miRNA signature in leukocytes from sporadic amyotrophic lateral sclerosis. Gene. 2012;508(1):p.35–40.
  168. Freischmidt A, et al. Systemic dysregulation of TDP-43 binding microRNAs in amyotrophic lateral sclerosis. Acta Neuropathol Commun. 2013;1(1):p.42.
  169. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington's disease chromosomes. The Huntington's Disease Collaborative Research Group. Cell. 1993;72(6):p.971–83.
  170. Vonsattel JP, DiFiglia M. Huntington disease. J Neuropathol Exp Neurol. 1998;57(5):p.369–84.
  171. Dumas EM, et al. Early changes in white matter pathways of the sensorimotor cortex in premanifest Huntington's disease. Hum Brain Mapp. 2012;33(1):p.203–12.
  172. Tabrizi SJ, et al. Potential endpoints for clinical trials in premanifest and early Huntington's disease in the TRACK-HD study: analysis of 24 month observational data. Lancet Neurol. 2012;11(1):p.42–53.
  173. Ross CA, Tabrizi SJ. Huntington's disease: from molecular pathogenesis to clinical treatment. Lancet Neurol. 2011;10(1):p.83–98.
  174. Tabrizi SJ, et al. Biological and clinical changes in premanifest and early stage Huntington's disease in the TRACK-HD study: the 12-month longitudinal analysis. Lancet Neurol. 2011;10(1):p.31–42.
  175. Langbehn DR, et al. A new model for prediction of the age of onset and penetrance for Huntington's disease based on CAG length. Clin Genet. 2004;65(4):p.267–77.
  176. Wexler NS, Res UVC. Venezuelan kindreds reveal that genetic and environmental factors modulate Huntington's disease age of onset. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(10):p.3498–503.
  177. Novak MJ, Tabrizi SJ. Huntington's disease. BMJ. 2010;340:p.c3109.
  178. Zuccato C, et al. Huntingtin interacts with REST/NRSF to modulate the transcription of NRSE-controlled neuronal genes. Nat Genet. 2003;35(1):p.76–83.
  179. Cattaneo E, Zuccato C, Tartari M. Normal huntingtin function: an alternative approach to Huntington's disease. Nat Rev Neurosci. 2005;6(12):p.919–30.
  180. Conaco C, et al. Reciprocal actions of REST and a microRNA promote neuronal identity. Proc Natl Acad Sci U S A. 2006;103(7):p.2422–7.
  181. Johnson R, et al. A microRNA-based gene dysregulation pathway in Huntington's disease. Neurobiol Dis. 2008;29(3):p.438–45.
  182. Griffiths-Jones S. The microRNA Registry. Nucleic Acids Res. 2004;32(Database issue):p.D109–11.
  183. Soldati C, et al. Dysregulation of REST-regulated coding and non-coding RNAs in a cellular model of Huntington's disease. J Neurochem. 2013;124(3):p.418–30.
  184. Kocerha J, et al. microRNA-128a dysregulation in transgenic Huntington's disease monkeys. Mol Brain. 2014;7:p.46.
  185. Cheng PH, et al. miR-196a ameliorates phenotypes of Huntington disease in cell, transgenic mouse, and induced pluripotent stem cell models. Am J Hum Genet. 2013;93(2):p.306–12.
  186. Tarasov V, et al. Differential regulation of microRNAs by p53 revealed by massively parallel sequencing: miR-34a is a p53 target that induces apoptosis and G1-arrest. Cell Cycle. 2007;6(13):p.1586–93.
  187. Bae BI, et al. p53 mediates cellular dysfunction and behavioral abnormalities in Huntington's disease. Neuron. 2005;47(1):p.29–41.
  188. Sinha M, Mukhopadhyay S, Bhattacharyya NP. Mechanism(s) of alteration of micro RNA expressions in Huntington's disease and their possible contributions to the observed cellular and molecular dysfunctions in the disease. Neuromolecular Med. 2012;4(4):p.221–43.
  189. Hoss AG, et al. MicroRNAs located in the Hox gene clusters are implicated in huntington's disease pathogenesis. PLoS Genet. 2014;10(2):p.e1004188.
  190. Savas JN, et al. A role for huntington disease protein in dendritic RNA granules. J Biol Chem. 2010;285(17):p.13142–53.
  191. Savas JN, et al. Huntington's disease protein contributes to RNA-mediated gene silencing through association with Argonaute and P bodies. Proc Natl Acad Sci U S A. 2008;105(31):p.10820–5.
  192. Andersson MK, et al. The multifunctional FUS, EWS and TAF15 proto-oncoproteins show cell type-specific expression patterns and involvement in cell spreading and stress response. BMC Cell Biol. 2008;9:p.37.
  193. Liu-Yesucevitz L, et al. Tar DNA binding protein-43 (TDP-43) associates with stress granules: analysis of cultured cells and pathological brain tissue. PLoS One. 2010;5(10):p.e13250.
  194. Borovecki F, et al, Genome-wide expression profiling of human blood reveals biomarkers for Huntington's disease. Proc Natl Acad Sci U S A. 2005;102(31):p.11023–8.
  195. Runne, H, et al. Analysis of potential transcriptomic biomarkers for Huntington's disease in peripheral blood. Proc Natl Acad Sci U S A. 2007;104(36):p.14424–9.
  196. Chen JF, et al. The role of microRNA-1 and microRNA-133 in skeletal muscle proliferation and differentiation. Nat Genet. 2006;38(2):p.228–33.
  197. Kim HK, et al. Muscle-specific microRNA miR-206 promotes muscle differentiation. J Cell Biol. 2006;174(5):p.677–87.
  198. Naguibneva I, et al. The microRNA miR-181 targets the homeobox protein Hox-A11 during mammalian myoblast differentiation. Nat Cell Biol. 2006;8(3):p.278–84.
  199. Davies KE, Nowak KJ. Molecular mechanisms of muscular dystrophies: old and new players. Nat Rev Mol Cell Biol. 2006;7(10):p.762–73.
  200. McCarthy JJ, Esser KA, Andrade FH. MicroRNA-206 is overexpressed in the diaphragm but not the hindlimb muscle of mdx mouse. Am J Physiol Cell Physiol. 2007;293(1):p.C451–7.
  201. Eisenberg I, et al. Distinctive patterns of microRNA expression in primary muscular disorders. Proc Natl Acad Sci U S A. 2007;104(43):p.17016–21.
  202. Greco S, et al. Common micro-RNA signature in skeletal muscle damage and regeneration induced by Duchenne muscular dystrophy and acute ischemia. FASEB J. 2009;23(10):p3335–46.
  203. Cacchiarelli D, et al. MicroRNAs involved in molecular circuitries relevant for the Duchenne muscular dystrophy pathogenesis are controlled by the dystrophin/nNOS pathway. Cell Metab. 2010;12(4):p.341–51.
  204. Cacchiarelli D, et al. miRNAs as serum biomarkers for Duchenne muscular dystrophy. EMBO Mol Med. 2011;3(5):p.258–65.
  205. Cacchiarelli D, et al. miR-31 modulates dystrophin expression: new implications for Duchenne muscular dystrophy therapy. EMBO Rep. 2011;12(2):p.136–41.
  206. Liu N, et al. microRNA-206 promotes skeletal muscle regeneration and delays progression of Duchenne muscular dystrophy in mice. J Clin Invest. 2012;122(6):p.2054–65.
  207. Small EM, Olson EN. Pervasive roles of microRNAs in cardiovascular biology. Nature. 2011;469(7330):p.336–42.
  208. Jakob P, Landmesser U. Role of microRNAs in stem/progenitor cells and cardiovascular repair. Cardiovasc Res. 2012;93(4):p.614–22.
  209. van Rooij E, Kauppinen S. Development of microRNA therapeutics is coming of age. EMBO Mol Med. 2014;6(7):p.851–64.
  210. Chen JF, et al. Targeted deletion of Dicer in the heart leads to dilated cardiomyopathy and heart failure. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(6):p.2111–6.
  211. Rao PK, et al. Loss of cardiac microRNA-mediated regulation leads to dilated cardiomyopathy and heart failure. Circulation research. 2009;105(6):p.585–94.
  212. Wahlquist C, et al. Inhibition of miR-25 improves cardiac contractility in the failing heart. Nature. 2014;508(7497):p.531–5.
  213. Boon RA, et al. MicroRNA-34a regulates cardiac ageing and function. Nature. 2013;495(7439):p.107–10.
  214. Wilson KD, et al. Dynamic microRNA expression programs during cardiac differentiation of human embryonic stem cells: role for miR-499. Circulation. Cardiovascular genetics. 2010;3(5):p.426–35.
  215. Sluijter JP, et al. MicroRNA-1 and -499 regulate differentiation and proliferation in human-derived cardiomyocyte progenitor cells. Arteriosclerosis, thrombosis, and vascular biology. 2010;30(4):p.859–68.
  216. Rayner KJ, et al. MiR-33 contributes to the regulation of cholesterol homeostasis. Science. 2010;328(5985):p.1570–3.
  217. Rayner KJ, et al. Antagonism of miR-33 in mice promotes reverse cholesterol transport and regression of atherosclerosis. J Clin Invest. 2011;121(7):p.2921–31.
  218. Rayner KJ, et al. Inhibition of miR-33a/b in non-human primates raises plasma HDL and lowers VLDL triglycerides. Nature. 2011;478(7369):p.404–7.
  219. Jakob P, et al. Loss of angiomiR-126 and 130a in angiogenic early outgrowth cells from patients with chronic heart failure: role for impaired in vivo neovascularization and cardiac repair capacity. Circulation. 2012;126(25):p.2962–75.
  220. Schober A, et al. MicroRNA-126-5p promotes endothelial proliferation and limits atherosclerosis by suppressing Dlk1. Nat Med. 2014;20(4):p.368–76.
  221. Harris TA, et al. MicroRNA-126 regulates endothelial expression of vascular cell adhesion molecule 1. Proc Natl Acad Sci U S A. 2008;105(5):p.1516–21.
  222. Nakasa T, et al. Expression of microRNA-146 in rheumatoid arthritis synovial tissue. Arthritis Rheum. 2008;58(5):p.1284–92.
  223. Stanczyk J, et al. Altered expression of MicroRNA in synovial fibroblasts and synovial tissue in rheumatoid arthritis. Arthritis Rheum. 2008;58(4):p.1001–9.
  224. O'Connell RM, et al. Sustained expression of microRNA-155 in hematopoietic stem cells causes a myeloproliferative disorder. J Exp Med. 2008;205(3):p.585–94.
  225. Chatzikyriakidou A, et al. A polymorphism in the 3'-UTR of interleukin-1 receptor-associated kinase (IRAK1), a target gene of miR-146a, is associated with rheumatoid arthritis susceptibility. Joint Bone Spine. 2010;77(5):p.411–3.
  226. Tang Y, et al. MicroRNA-146A contributes to abnormal activation of the type I interferon pathway in human lupus by targeting the key signaling proteins. Arthritis Rheum. 2009;60(4):p.1065–75.
  227. Luo X, et al. A functional variant in microRNA-146a promoter modulates its expression and confers disease risk for systemic lupus erythematosus. PLoS Genet. 2011;7(6):p.e1002128.
  228. Pan Y, et al. MS2 VLP-based delivery of microRNA-146a inhibits autoantibody production in lupus-prone mice. Int J Nanomedicine. 2012;7:p.5957–67.
  229. Denby L, et al. miR-21 and miR-214 are consistently modulated during renal injury in rodent models. Am J Pathol. 2011;179(2):p.661–72.
  230. Altorok N, et al. Epigenetics, the holy grail in the pathogenesis of systemic sclerosis. Rheumatology (Oxford). 2014.
  231. Bhanji RA, et al. Clinical and serological features of patients with autoantibodies to GW/P bodies. Clinical Immunology. 2007;125(3):p.247–56.
  232. Wu F, et al. MicroRNAs are differentially expressed in ulcerative colitis and alter expression of macrophage inflammatory peptide-2 alpha. Gastroenterology. 2008;135(5):p.1624–35e24.
  233. Nguyen HT, et al. MicroRNA-7 modulates CD98 expression during intestinal epithelial cell differentiation. J Biol Chem. 2010;285(2):p.1479–89.
  234. Brest P, et al. A synonymous variant in IRGM alters a binding site for miR-196 and causes deregulation of IRGM-dependent xenophagy in Crohn's disease. Nat Genet. 2011;43(3):p.242–5.
  235. Zwiers A, et al. Cutting edge: a variant of the IL-23R gene associated with inflammatory bowel disease induces loss of microRNA regulation and enhanced protein production. J Immunol. 2012;188(4):p.1573–7.
  236. Duroux-Richard I, Jorgensen C, and F. Apparailly, miRNAs and rheumatoid arthritis - promising novel biomarkers. Swiss Med Wkly. 2011;141:p.w13175.
  237. Martins M, et al. Convergence of miRNA expression profiling, alpha-synuclein interacton and GWAS in Parkinson's disease. PLoS One. 2011;6(10):p.e25443.

Most read articles by the same author(s)

1 2 > >>