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:

Stress granules and neurodegeneration

Abstract

Recent advances suggest that the response of RNA metabolism to stress has an important role in the pathophysiology of neurodegenerative diseases, particularly amyotrophic lateral sclerosis, frontotemporal dementias and Alzheimer disease. RNA-binding proteins (RBPs) control the utilization of mRNA during stress, in part through the formation of membraneless organelles termed stress granules (SGs). These structures form through a process of liquid–liquid phase separation. Multiple biochemical pathways regulate SG biology. The major signalling pathways regulating SG formation include the mammalian target of rapamycin (mTOR)–eukaryotic translation initiation factor 4F (eIF4F) and eIF2α pathways, whereas the pathways regulating SG dispersion and removal are mediated by valosin-containing protein and the autolysosomal cascade. Post-translational modifications of RBPs also strongly contribute to the regulation of SGs. Evidence indicates that SGs are supposed to be transient structures, but the chronic stresses associated with ageing lead to chronic, persistent SGs that appear to act as a nidus for the aggregation of disease-related proteins. We suggest a model describing how intrinsic vulnerabilities within the cellular RNA metabolism might lead to the pathological aggregation of RBPs when SGs become persistent. This process might accelerate the pathophysiology of many neurodegenerative diseases and myopathies, and it suggests new targets for disease intervention.

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

Fig. 1: Types of membraneless organelles present in neurons.
Fig. 2: Formation of stress granules from the mRNA–ribosomal complex.
Fig. 3: Regulation of stress granule assembly.
Fig. 4: Phases in the stress granule cycle.
Fig. 5: The RBP cascade hypothesis.

Similar content being viewed by others

References

  1. Nott, T. J. et al. Phase transition of a disordered nuage protein generates environmentally responsive membraneless organelles. Mol. Cell 57, 936–947 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Boeynaems, S. et al. Protein phase separation: a new phase in cell biology. Trends Cell Biol. 28, 420–435 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Taylor, J. P., Brown, R. H., Jr. & Cleveland, D. W. Decoding ALS: from genes to mechanism. Nature 539, 197–206 (2016).

    PubMed  PubMed Central  Google Scholar 

  4. Irwin, D. J. et al. Frontotemporal lobar degeneration: defining phenotypic diversity through personalized medicine. Acta Neuropathol. 129, 469–491 (2015).

    PubMed  Google Scholar 

  5. Kim, H. J. et al. Therapeutic modulation of eIF2α phosphorylation rescues TDP-43 toxicity in amyotrophic lateral sclerosis disease models. Nat. Genet. 46, 152–160 (2014).

    CAS  PubMed  Google Scholar 

  6. Radford, H., Moreno, J. A., Verity, N., Halliday, M. & Mallucci, G. R. PERK inhibition prevents tau-mediated neurodegeneration in a mouse model of frontotemporal dementia. Acta Neuropathol. 130, 633–642 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Chia, R., Chio, A. & Traynor, B. J. Novel genes associated with amyotrophic lateral sclerosis: diagnostic and clinical implications. Lancet Neurol. 17, 94–102 (2018).

    CAS  PubMed  Google Scholar 

  8. Jack, C. R., Jr. et al. NIA-AA research framework: toward a biological definition of alzheimer’s disease. Alzheimers Dement. 14, 535–562 (2018).

    PubMed  PubMed Central  Google Scholar 

  9. Jin, M. et al. Soluble amyloid beta-protein dimers isolated from Alzheimer cortex directly induce Tau hyperphosphorylation and neuritic degeneration. Proc. Natl Acad. Sci. USA 108, 5819–5824 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Bang, J., Spina, S. & Miller, B. L. Frontotemporal dementia. Lancet 386, 1672–1682 (2015).

    PubMed  PubMed Central  Google Scholar 

  11. Lin, Y., Protter, D. S., Rosen, M. K. & Parker, R. Formation and maturation of phase-separated liquid droplets by RNA-binding proteins. Mol. Cell 60, 208–219 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Liu-Yesucevitz, L. et al. Local RNA translation at the synapse and in disease. J. Neurosci. 31, 16086–16093 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Parker, R. & Sheth, U. P bodies and the control of mRNA translation and degradation. Mol. Cell 25, 635–646 (2007).

    CAS  PubMed  Google Scholar 

  14. Mao, Y. S., Zhang, B. & Spector, D. L. Biogenesis and function of nuclear bodies. Trends Genet. 27, 295–306 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Shpargel, K. B. & Matera, A. G. Gemin proteins are required for efficient assembly of Sm-class ribonucleoproteins. Proc. Natl Acad. Sci. USA 102, 17372–17377 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Zhang, H. et al. Multiprotein complexes of the survival of motor neuron protein SMN with Gemins traffic to neuronal processes and growth cones of motor neurons. J. Neurosci. 26, 8622–8632 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. DeJesus-Hernandez, M. et al. Expanded GGGGCC hexanucleotide repeat in noncoding region of C9ORF72 causes chromosome 9p-linked FTD and ALS. Neuron 72, 245–256 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Renton, A. E. et al. A hexanucleotide repeat expansion in C9ORF72 is the cause of chromosome 9p21-linked ALS-FTD. Neuron 72, 257–268 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Jain, A. & Vale, R. D. RNA phase transitions in repeat expansion disorders. Nature 546, 243–247 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Fay, M. M., Anderson, P. J. & Ivanov, P. ALS/FTD-associated C9ORF72 repeat RNA promotes phase transitions in vitro and in cells. Cell Rep. 21, 3573–3584 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  21. Van Treeck, B. et al. RNA self-assembly contributes to stress granule formation and defining the stress granule transcriptome. Proc. Natl Acad. Sci. USA 115, 2734–2739 (2018).

    PubMed  PubMed Central  Google Scholar 

  22. Maharana, S. et al. RNA buffers the phase separation behavior of prion-like RNA binding proteins. Science 360, 918–921 (2018). This study elegantly shows the important role that RNA plays in regulating the phase separation dynamics of membraneless organelles.

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Guo, L. et al. Nuclear-import receptors reverse aberrant phase transitions of RNA-binding proteins with prion-like domains. Cell 173, 677–692.e620 (2018). This paper demonstrates that nuclear import receptors can act as disaggregates, thereby providing a novel mechanism for regulating SGs and identifying a functional link between the nuclear transport machinery and disease pathology.

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Hofweber, M. et al. Phase separation of FUS is suppressed by its nuclear import receptor and arginine methylation. Cell 173, 706–719.e713 (2018).

    CAS  PubMed  Google Scholar 

  25. Qamar, S. et al. FUS phase separation is modulated by a molecular chaperone and methylation of arginine cation–pi interactions. Cell 173, 720–734.e715 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. 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 5, e13250 (2010). This report shows that TDP43 associates with SGs and also provides immunohistochemical evidence that TDP43 pathology associates with SG markers in ALS brain tissue.

    PubMed  PubMed Central  Google Scholar 

  27. Gupta, N. et al. Stress granule-associated protein G3BP2 regulates breast tumor initiation. Proc. Natl Acad. Sci. USA 114, 1033–1038 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Valentin-Vega, Y. A. et al. Cancer-associated DDX3X mutations drive stress granule assembly and impair global translation. Sci. Rep. 6, 25996 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Anderson, P., Kedersha, N. & Ivanov, P. Stress granules, P-bodies and cancer. Biochim. Biophys. Acta 1849, 861–870 (2015).

    CAS  PubMed  Google Scholar 

  30. Kedersha, N. et al. Dynamic shuttling of TIA-1 accompanies the recruitment of mRNA to mammalian stress granules. J. Cell Biol. 151, 1257–1268 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Kedersha, N. L., Gupta, M., Li, W., Miller, I. & Anderson, P. RNA-binding proteins TIA-1 and TIAR link the phosphorylation of eIF-2α to the assembly of mammalian stress granules. J. Cell Biol. 147, 1431–1442 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Anderson, P. & Kedersha, N. Stress granules: the Tao of RNA triage. Trends Biochem. Sci. 33, 141–150 (2008).

    CAS  PubMed  Google Scholar 

  33. Kedersha, N. et al. Evidence that ternary complex (eIF2–GTP–tRNAi Met)-deficient preinitiation complexes are core constituents of mammalian stress granules. Mol. Biol. Cell 13, 195–210 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Markmiller, S. et al. Context-dependent and disease-specific diversity in protein interactions within stress granules. Cell 172, 590–604.e513 (2018). This paper provides a comprehensive characterization of SG proteomes in cells and also highlights the compositional differences among different types of SGs.

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Lastres-Becker, I. et al. Mammalian ataxin-2 modulates translation control at the pre-initiation complex via PI3K/mTOR and is induced by starvation. Biochim. Biophys. Acta 1862, 1558–1569 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Kedersha, N. et al. G3BP–caprin1–USP10 complexes mediate stress granule condensation and associate with 40S subunits. J. Cell Biol. 212, 845–860 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Ohn, T., Kedersha, N., Hickman, T., Tisdale, S. & Anderson, P. A functional RNAi screen links O-GlcNAc modification of ribosomal proteins to stress granule and processing body assembly. Nat. Cell Biol. 10, 1224–1231 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Jain, S. et al. ATPase-modulated stress granules contain a diverse proteome and substructure. Cell 164, 487–498 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Wheeler, J. R., Jain, S., Khong, A. & Parker, R. Isolation of yeast and mammalian stress granule cores. Methods 126, 12–17 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Wheeler, J. R., Matheny, T., Jain, S., Abrisch, R. & Parker, R. Distinct stages in stress granule assembly and disassembly. eLife 5, e18413 (2016).

    PubMed  PubMed Central  Google Scholar 

  41. Antar, L. N., Afroz, R., Dictenberg, J. B., Carroll, R. C. & Bassell, G. J. Metabotropic glutamate receptor activation regulates fragile X mental retardation protein and FMR1 mRNA localization differentially in dendrites and at synapses. J. Neurosci. 24, 2648–2655 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Jackson, R. J., Hellen, C. U. & Pestova, T. V. The mechanism of eukaryotic translation initiation and principles of its regulation. Nat. Rev. Mol. Cell Biol. 11, 113–127 (2010). This review provides an outstanding and detailed overview of the machinery carrying out translational initiation and the pathways of translational regulation.

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Fujimura, K., Sasaki, A. T. & Anderson, P. Selenite targets eIF4E-binding protein-1 to inhibit translation initiation and induce the assembly of non-canonical stress granules. Nucleic Acids Res. 40, 8099–8110 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Emara, M. M. et al. Hydrogen peroxide induces stress granule formation independent of eIF2α phosphorylation. Biochem. Biophys. Res. Commun. 423, 763–769 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Szaflarski, W. et al. Vinca alkaloid drugs promote stress-induced translational repression and stress granule formation. Oncotarget 7, 30307–30322 (2016).

    PubMed  PubMed Central  Google Scholar 

  46. Saxton, R. A. & Sabatini, D. M. mTOR signaling in growth, metabolism, and disease. Cell 168, 960–976 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Thoreen, C. C. et al. A unifying model for mTORC1-mediated regulation of mRNA translation. Nature 485, 109–113 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Frydryskova, K. et al. Distinct recruitment of human eIF4E isoforms to processing bodies and stress granules. BMC Mol. Biol. 17, 21 (2016).

    PubMed  PubMed Central  Google Scholar 

  49. Wek, R. C. Role of eif2α kinases in translational control and adaptation to cellular stress. Cold Spring Harb. Perspect. Biol. 10, a032870 (2018).

    PubMed  PubMed Central  Google Scholar 

  50. Wek, S. A., Zhu, S. & Wek, R. C. The histidyl-tRNA synthetase-related sequence in the eIF-2α protein kinase GCN2 interacts with tRNA and is required for activation in response to starvation for different amino acids. Mol. Cell Biol. 15, 4497–4506 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Harding, H. P., Zhang, Y., Bertolotti, A., Zeng, H. & Ron, D. Perk is essential for translational regulation and cell survival during the unfolded protein response. Mol. Cell 5, 897–904 (2000).

    CAS  PubMed  Google Scholar 

  52. Srivastava, S. P., Kumar, K. U. & Kaufman, R. J. Phosphorylation of eukaryotic translation initiation factor 2 mediates apoptosis in response to activation of the double-stranded RNA-dependent protein kinase. J. Biol. Chem. 273, 2416–2423 (1998).

    CAS  PubMed  Google Scholar 

  53. McEwen, E. et al. Heme-regulated inhibitor kinase-mediated phosphorylation of eukaryotic translation initiation factor 2 inhibits translation, induces stress granule formation, and mediates survival upon arsenite exposure. J. Biol. Chem. 280, 16925–16933 (2005).

    CAS  PubMed  Google Scholar 

  54. Sudhakar, A. et al. Phosphorylation of serine 51 in initiation factor 2α (eIF2α) promotes complex formation between eIF2α(P) and eIF2B and causes inhibition in the guanine nucleotide exchange activity of eIF2B. Biochemistry 39, 12929–12938 (2000).

    CAS  PubMed  Google Scholar 

  55. Kedersha, N. & Anderson, P. Stress granules: sites of mRNA triage that regulate mRNA stability and translatability. Biochem. Soc. Trans. 30, 963–969 (2002).

    CAS  PubMed  Google Scholar 

  56. Maziuk, B. F. et al. RNA binding proteins co-localize with small tau inclusions in tauopathy. Acta Neuropathol. Commun. 6, 71 (2018). This study provides key methods for detecting RBPs associated with pathological SGs.

    PubMed  PubMed Central  Google Scholar 

  57. Li, Y. et al. Immunoprecipitation and mass spectrometry defines an extensive RBM45 protein-protein interaction network. Brain Res. 1647, 79–93 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Umoh, M. E. et al. A proteomic network approach across the ALS–FTD disease spectrum resolves clinical phenotypes and genetic vulnerability in human brain. EMBO Mol. Med. 10, 48–62 (2018).

    CAS  PubMed  Google Scholar 

  59. Gunawardana, C. G. et al. The human tau interactome: binding to the ribonucleoproteome, and impaired binding of the proline-to-leucine mutant at position 301 (p301l) to chaperones and the proteasome. Mol. Cell Proteom. 14, 3000–3014 (2015).

    CAS  Google Scholar 

  60. Dhungel, N. et al. Parkinson’s disease genes VPS35 and EIF4G1 interact genetically and converge on α-synuclein. Neuron 85, 76–87 (2015).

    CAS  PubMed  Google Scholar 

  61. Nichols, N. et al. EIF4G1 mutations do not cause Parkinson’s disease. Neurobiol. Aging 36, 2444.e1–4 (2015).

    CAS  Google Scholar 

  62. Kim, W. J., Kim, J. H. & Jang, S. K. Anti-inflammatory lipid mediator 15d-PGJ2 inhibits translation through inactivation of eIF4A. EMBO J. 26, 5020–5032 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Bordeleau, M. E. et al. Functional characterization of IRESes by an inhibitor of the RNA helicase eIF4A. Nat. Chem. Biol. 2, 213–220 (2006).

    CAS  PubMed  Google Scholar 

  64. Cencic, R. et al. Antitumor activity and mechanism of action of the cyclopenta[b]benzofuran, silvestrol. PLOS ONE 4, e5223 (2009).

    PubMed  PubMed Central  Google Scholar 

  65. Ivanov, P., Emara, M. M., Villen, J., Gygi, S. P. & Anderson, P. Angiogenin-induced tRNA fragments inhibit translation initiation. Mol. Cell 43, 613–623 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Lyons, S. M., Achorn, C., Kedersha, N. L., Anderson, P. J. & Ivanov, P. YB-1 regulates tiRNA-induced stress granule formation but not translational repression. Nucleic Acids Res. 44, 6949–6960 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Apicco, D. J. et al. Reducing the RNA binding protein TIA1 protects against tau-mediated neurodegeneration in vivo. Nat. Neurosci. 21, 72–80 (2018). This article demonstrates that RBPs, such as TIA1, regulate the pathophysiology of tau in vivo, and also shows differential regulation of tau oligomers versus fibrils.

    CAS  PubMed  Google Scholar 

  68. Armstrong, R. A., Carter, D. & Cairns, N. J. A quantitative study of the neuropathology of 32 sporadic and familial cases of frontotemporal lobar degeneration with TDP-43 proteinopathy (FTLD–TDP). Neuropathol. Appl. Neurobiol. 38, 25–38 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Janssens, J. et al. Overexpression of ALS-associated p.M337V human TDP-43 in mice worsens disease features compared to wild-type human TDP-43 mice. Mol. Neurobiol. 48, 22–35 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Figley, M. D., Bieri, G., Kolaitis, R. M., Taylor, J. P. & Gitler, A. D. Profilin 1 associates with stress granules and als-linked mutations alter stress granule dynamics. J. Neurosci. 34, 8083–8097 (2014).

    PubMed  PubMed Central  Google Scholar 

  71. Gilks, N. et al. Stress granule assembly is mediated by prion-like aggregation of TIA-1. Mol. Biol. Cell 15, 5383–5398 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Bounedjah, O. et al. Free mRNA in excess upon polysome dissociation is a scaffold for protein multimerization to form stress granules. Nucleic Acids Res. 42, 8678–8691 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Bley, N. et al. Stress granules are dispensable for mRNA stabilization during cellular stress. Nucleic Acids Res. 43, e26 (2015).

    PubMed  Google Scholar 

  74. Vogler, T. O. et al. TDP-43 and RNA form amyloid-like myo-granules in regenerating muscle. Nature 563, 508–513 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Lefebvre, S. et al. Identification and characterization of a spinal muscular atrophy-determining gene. Cell 80, 155–165 (1995).

    CAS  PubMed  Google Scholar 

  76. Bassell, G. J. & Warren, S. T. Fragile X syndrome: loss of local mRNA regulation alters synaptic development and function. Neuron 60, 201–214 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Kenneson, A., Zhang, F., Hagedorn, C. H. & Warren, S. T. Reduced FMRP and increased FMR1 transcription is proportionally associated with CGG repeat number in intermediate-length and premutation carriers. Hum. Mol. Genet. 10, 1449–1454 (2001).

    CAS  PubMed  Google Scholar 

  78. Hagerman, P. J. & Hagerman, R. J. Fragile X-associated tremor/ataxia syndrome (FXTAS). Ment. Retard. Dev. Disabil. Res. Rev. 10, 25–30 (2004).

    PubMed  Google Scholar 

  79. Neumann, M. et al. Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Science 314, 130–133 (2006).

    CAS  PubMed  Google Scholar 

  80. Vance, C. et al. Mutations in FUS, an RNA processing protein, cause familial amyotrophic lateral sclerosis type 6. Science 323, 1208–1211 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Van Deerlin, V. M. et al. TARDBP mutations in amyotrophic lateral sclerosis with TDP-43 neuropathology: a genetic and histopathological analysis. Lancet Neurol. 7, 409–416 (2008).

    PubMed  PubMed Central  Google Scholar 

  82. Mackenzie, I. R. et al. TIA1 mutations in amyotrophic lateral sclerosis and frontotemporal dementia promote phase separation and alter stress granule dynamics. Neuron 95, 808–816 e809 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Kim, H. J. et al. Mutations in prion-like domains in hnRNPA2B1 and hnRNPA1 cause multisystem proteinopathy and ALS. Nature 495, 467–473 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Sreedharan, J. et al. TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis. Science 319, 1668–1672 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Kwiatkowski, T. J., Jr. et al. Mutations in the FUS/TLS gene on chromosome 16 cause familial amyotrophic lateral sclerosis. Science 323, 1205–1208 (2009).

    CAS  PubMed  Google Scholar 

  86. Elden, A. C. et al. Ataxin-2 intermediate-length polyglutamine expansions are associated with increased risk for ALS. Nature 466, 1069–1075 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Emara, M. M. et al. Angiogenin-induced tRNA-derived stress-induced RNAs promote stress-induced stress granule assembly. J. Biol. Chem. 285, 10959–10968 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  88. Greenway, M. J. et al. ANG mutations segregate with familial and ‘sporadic’ amyotrophic lateral sclerosis. Nat. Genet. 38, 411–413 (2006).

    CAS  PubMed  Google Scholar 

  89. Dormann, D. & Haass, C. TDP-43 and FUS: a nuclear affair. Trends Neurosci. 34, 339–348 (2011).

    CAS  PubMed  Google Scholar 

  90. Bosco, D. A. et al. Mutant FUS proteins that cause amyotrophic lateral sclerosis incorporate into stress granules. Hum. Mol. Genet. 19, 4160–4175 (2011).

    Google Scholar 

  91. Phillips, K., Kedersha, N., Shen, L., Blackshear, P. J. & Anderson, P. Arthritis suppressor genes TIA-1 and TTP dampen the expression of tumor necrosis factor α, cyclooxygenase 2, and inflammatory arthritis. Proc. Natl Acad. Sci. USA 101, 2011–2016 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Garaigorta, U., Heim, M. H., Boyd, B., Wieland, S. & Chisari, F. V. Hepatitis C virus (HCV) induces formation of stress granules whose proteins regulate HCV RNA replication and virus assembly and egress. J. Virol. 86, 11043–11056 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  93. Wolozin, B. Regulated protein aggregation: stress granules and neurodegeneration. Mol. Neurodegener. 7, 56 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Kiernan, M. C. et al. Amyotrophic lateral sclerosis. Lancet 377, 942–955 (2011).

    CAS  PubMed  Google Scholar 

  95. Scheltens, P. et al. Alzheimer’s disease. Lancet 388, 505–517 (2016).

    CAS  PubMed  Google Scholar 

  96. Banani, S. F. et al. Compositional control of phase-separated cellular bodies. Cell 166, 651–663 (2016). This article outlines the different cellular factors controlling the formation of membraneless organelles, such as stress granules and P-bodies.

    CAS  PubMed  PubMed Central  Google Scholar 

  97. Lee, Y. et al. TIA1 variant drives myodegeneration in multisystem proteinopathy with SQSTM1 mutations. J. Clin. Invest. 128, 1164–1177 (2018).

    PubMed  PubMed Central  Google Scholar 

  98. Patel, A. et al. A liquid-to-solid phase transition of the ALS protein FUS accelerated by disease mutation. Cell 162, 1066–1077 (2015).

    CAS  PubMed  Google Scholar 

  99. Molliex, A. et al. Phase separation by low complexity domains promotes stress granule assembly and drives pathological fibrillization. Cell 163, 123–133 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Hughes, M. P. et al. Atomic structures of low-complexity protein segments reveal kinked beta sheets that assemble networks. Science 359, 698–701 (2018). This study shows how different types of amino acid interactions, such as π–π and π–cation binding, control the formation of membraneless organelles and insoluble amyloids.

    CAS  PubMed  PubMed Central  Google Scholar 

  101. Ash, P. E. A. et al. Heavy metal neurotoxicants induce ALS-linked TDP-43 pathology. Toxicol. Sci. 167, 105–115, https://doi.org/10.1093/toxsci/kfy267 (2019).

    Article  CAS  PubMed  Google Scholar 

  102. Bai, B. et al. U1 small nuclear ribonucleoprotein complex and RNA splicing alterations in Alzheimer’s disease. Proc. Natl Acad. Sci. USA 110, 16562–16567 (2013). This paper shows that many RBPs accumulate as protein aggregates in AD, often in cells lacking histologically evident tau aggregates.

    CAS  PubMed  PubMed Central  Google Scholar 

  103. Vanderweyde, T. et al. Interaction of tau with the RNA-binding protein TIA1 regulates tau pathophysiology and toxicity. Cell Rep. 15, 1–12 (2016).

    Google Scholar 

  104. Vanderweyde, T. et al. Contrasting pathology of stress granule proteins TIA-1 and G3BP in tauopathies. J. Neurosci. 32, 8270–8283 (2012). This article demonstrates how tau and TIA1 interact to regulate SG biology and demonstrates that reducing TIA1 inhibits tau-mediated neurodegeneration in cultured neurons.

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Zhang, Y. J. et al. Poly(GR) impairs protein translation and stress granule dynamics in C9orf72-associated frontotemporal dementia and amyotrophic lateral sclerosis. Nat. Med. 24, 1136–1142 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Ash, P. E. et al. Unconventional translation of C9ORF72 GGGGCC expansion generates insoluble polypeptides specific to c9FTD/ALS. Neuron 77, 639–646 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Green, K. M. et al. RAN translation at C9orf72-associated repeat expansions is selectively enhanced by the integrated stress response. Nat. Commun. 8, 2005 (2017).

    PubMed  PubMed Central  Google Scholar 

  108. Boeynaems, S. et al. Phase separation of c9orf72 dipeptide repeats perturbs stress granule dynamics. Mol. Cell 65, 1044–1055.e1045 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Lee, K. H. et al. C9orf72 dipeptide repeats impair the assembly, dynamics, and function of membrane-less organelles. Cell 167, 774–788.e717 (2016). This study shows how an extended dipeptide repeat can disrupt SG biology.

    CAS  PubMed  PubMed Central  Google Scholar 

  110. Gasset-Rosa, F. et al. Cytoplasmic TDP-43 de-mixing independent of stress granules drives inhibition of nuclear import, loss of nuclear TDP-43, and cell death. Neuron 102, 339–357.e7 (2019). This article demonstrates that, in cells, TDP43 aggregates pass through a SG phase transiently before evolving into pathological aggregates.

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Mann, J. R. et al. RNA binding antagonizes neurotoxic phase transitions of TDP-43. Neuron 102, 321–338.e8 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  112. McGurk, L. et al. Poly(ADP–ribose) prevents pathological phase separation of TDP-43 by promoting liquid demixing and stress granule localization. Mol. Cell 71, 703–717 e709 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  113. Brettschneider, J. et al. TDP-43 pathology and neuronal loss in amyotrophic lateral sclerosis spinal cord. Acta Neuropathol. 128, 423–437 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  114. Zhang, K. et al. The C9orf72 repeat expansion disrupts nucleocytoplasmic transport. Nature 525, 56–61 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Chou, C. C. et al. TDP-43 pathology disrupts nuclear pore complexes and nucleocytoplasmic transport in ALS/FTD. Nat. Neurosci. 21, 228–239 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Afroz, T. et al. Functional and dynamic polymerization of the ALS-linked protein TDP-43 antagonizes its pathologic aggregation. Nat. Commun. 8, 45 (2017).

    PubMed  PubMed Central  Google Scholar 

  117. Hanger, D. P., Hughes, K., Woodgett, J. R., Brion, J. P. & Anderton, B. H. Glycogen synthase kinase-3 induces Alzheimer’s disease-like phosphorylation of tau: generation of paired helical filament epitopes and neuronal localisation of the kinase. Neurosci. Lett. 147, 58–62 (1992).

    CAS  PubMed  Google Scholar 

  118. Baumann, K., Mandelkow, E., Biernat, J., Piwnica-Worms, H. & Mandelkow, E. Abnormal Alzheimer-like phosphorylation of tau-protein by cyclin-dependent kinases cdk2 and cdk5. FEBS Lett. 336, 417–424 (1993).

    CAS  PubMed  Google Scholar 

  119. Timm, T., Marx, A., Panneerselvam, S., Mandelkow, E. & Mandelkow, E. M. Structure and regulation of MARK, a kinase involved in abnormal phosphorylation of Tau protein. BMC Neurosci. 9, S9 (2008).

    PubMed  PubMed Central  Google Scholar 

  120. Trinczek, B., Biernat, J., Baumann, K., Mandelkow, E. M. & Mandelkow, E. Domains of tau protein, differential phosphorylation, and dynamic instability of microtubules. Mol. Biol. Cell 6, 1887–1902 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  121. Zempel, H. et al. Amyloid-beta oligomers induce synaptic damage via Tau-dependent microtubule severing by TTLL6 and spastin. EMBO J. 32, 2920–2937 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  122. Kampers, T., Friedhoff, P., Biernat, J., Mandelkow, E. M. & Mandelkow, E. RNA stimulates aggregation of microtubule-associated protein tau into Alzheimer-like paired helical filaments. FEBS Lett. 399, 344–349 (1996).

    CAS  PubMed  Google Scholar 

  123. Fitzpatrick, A. W. P. et al. Cryo-EM structures of tau filaments from Alzheimer’s disease. Nature 547, 185–190 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Friedhoff, P., Schneider, A., Mandelkow, E. M. & Mandelkow, E. Rapid assembly of Alzheimer-like paired helical filaments from microtubule-associated protein tau monitored by fluorescence in solution. Biochemistry 37, 10223–10230 (1998).

    CAS  PubMed  Google Scholar 

  125. Wilson, D. M. & Binder, L. I. Free fatty acids stimulate the polymerization of tau and amyloid beta peptides: in vitro evidence for a common effector of pathogenesis in Alzheimer’s disease. Am. J. Pathol. 150, 2181–2195 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Zhang, X. et al. RNA stores tau reversibly in complex coacervates. PLOS Biol. 15, e2002183 (2017).

    PubMed  PubMed Central  Google Scholar 

  127. Wegmann, S. et al. Tau protein liquid–liquid phase separation can initiate tau aggregation. EMBO J. 37, e98049 (2018). This article provides a detailed characterization of the biophysical factors controlling tau phase separation.

    PubMed  PubMed Central  Google Scholar 

  128. Meier, S. et al. Pathological tau promotes neuronal damage by impairing ribosomal function and decreasing protein synthesis. J. Neurosci. 36, 1001–1007 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  129. Wang, P. et al. Tau interactome mapping based identification of Otub1 as Tau deubiquitinase involved in accumulation of pathological Tau forms in vitro and in vivo. Acta Neuropathol. 133, 731–749 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Eftekharzadeh, B. et al. Tau protein disrupts nucleocytoplasmic transport in Alzheimer’s disease. Neuron 99, 925–940 e927 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  131. Silva, J. M. et al. Dysregulation of autophagy and stress granule-related proteins in stress-driven Tau pathology. Cell Death Differ. 171, 1411–1427 (2019). This paper demonstrates that behaviourally stressful experiences can elicit the biochemical changes associated with the translational stress response in the brain.

    Google Scholar 

  132. Cairns, N. J. et al. TDP-43 in familial and sporadic frontotemporal lobar degeneration with ubiquitin inclusions. Am. J. Pathol. 171, 227–240 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Moujalled, D. et al. Phosphorylation of hnRNP K by cyclin-dependent kinase 2 controls cytosolic accumulation of TDP-43. Hum. Mol. Genet. 24, 1655–1669 (2015).

    CAS  PubMed  Google Scholar 

  134. Jiang, L. et al. TIA1 regulates the generation and response to toxic tau oligomers. Acta Neuropathol. 137, 259–277 (2019). This article demonstrates that tau oligomers and fibrils both propagate pathology in vivo, but that only tau oligomers stimulate the accumulation of tau-associated SGs and cause neurodegeneration.

    CAS  PubMed  Google Scholar 

  135. Sanders, D. W. et al. Distinct tau prion strains propagate in cells and mice and define different tauopathies. Neuron 82, 1271–1288 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  136. Liu, L. et al. Trans-synaptic spread of tau pathology in vivo. PLOS ONE 7, e31302 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  137. Smethurst, P. et al. In vitro prion-like behaviour of TDP-43 in ALS. Neurobiol Dis. 96, 236–247 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  138. Asai, H. et al. Depletion of microglia and inhibition of exosome synthesis halt tau propagation. Nat. Neurosci. 18, 1584–1593 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  139. Ghosh, S. & Geahlen, R. L. Stress granules modulate SYK to cause microglial cell dysfunction in Alzheimer’s disease. EBioMedicine 2, 1785–1798 (2015).

    PubMed  PubMed Central  Google Scholar 

  140. Guerreiro, R. et al. TREM2 variants in Alzheimer’s disease. N. Engl. J. Med. 368, 117–127 (2013).

    CAS  PubMed  Google Scholar 

  141. Pimenova, A. A., Raj, T. & Goate, A. M. Untangling genetic risk for Alzheimer’s disease. Biol. Psychiatry 83, 300–310 (2018).

    CAS  PubMed  Google Scholar 

  142. Roberson, E. D. et al. Reducing endogenous tau ameliorates amyloid beta-induced deficits in an Alzheimer’s disease mouse model. Science 316, 750–754 (2007).

    CAS  PubMed  Google Scholar 

  143. Vossel, K. A. et al. Tau reduction prevents Aβ-induced defects in axonal transport. Science 330, 198 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  144. Cruts, M. et al. Null mutations in progranulin cause ubiquitin-positive frontotemporal dementia linked to chromosome 17q21. Nature 442, 920–924 (2006).

    CAS  PubMed  Google Scholar 

  145. Melamed, Z. et al. Premature polyadenylation-mediated loss of stathmin-2 is a hallmark of TDP-43-dependent neurodegeneration. Nat. Neurosci. 22, 180–190 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  146. Klim, J. R. et al. ALS-implicated protein TDP-43 sustains levels of STMN2, a mediator of motor neuron growth and repair. Nat. Neurosci. 22, 167–179 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  147. Nguyen, H. P., Van Broeckhoven, C. & van der Zee, J. ALS genes in the genomic era and their implications for FTD. Trends Genet. 34, 404–423 (2018).

    CAS  PubMed  Google Scholar 

  148. Deng, H. X. et al. Mutations in UBQLN2 cause dominant X-linked juvenile and adult-onset ALS and ALS/dementia. Nature 477, 211–215 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  149. Watts, G. D. et al. Inclusion body myopathy associated with Paget disease of bone and frontotemporal dementia is caused by mutant valosin-containing protein. Nat. Genet. 36, 377–381 (2004).

    CAS  PubMed  Google Scholar 

  150. Fecto, F. et al. SQSTM1 mutations in familial and sporadic amyotrophic lateral sclerosis. Arch. Neurol. 68, 1440–1446 (2011).

    PubMed  Google Scholar 

  151. Cirulli, E. T. et al. Exome sequencing in amyotrophic lateral sclerosis identifies risk genes and pathways. Science 347, 1436–1441 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  152. Freischmidt, A. et al. Haploinsufficiency of TBK1 causes familial ALS and fronto-temporal dementia. Nat. Neurosci. 18, 631–636 (2015).

    CAS  PubMed  Google Scholar 

  153. Weidberg, H. & Elazar, Z. TBK1 mediates crosstalk between the innate immune response and autophagy. Sci. Signal. 4, pe39 (2011).

    PubMed  Google Scholar 

  154. Skibinski, G. et al. Mutations in the endosomal ESCRTIII-complex subunit CHMP2B in frontotemporal dementia. Nat. Genet. 37, 806–808 (2005).

    CAS  PubMed  Google Scholar 

  155. Gilpin, K. M., Chang, L. & Monteiro, M. J. ALS-linked mutations in ubiquilin-2 or hnRNPA1 reduce interaction between ubiquilin-2 and hnRNPA1. Hum. Mol. Genet. 24, 2565–2577 (2015).

    CAS  PubMed  Google Scholar 

  156. Cassel, J. A. & Reitz, A. B. Ubiquilin-2 (UBQLN2) binds with high affinity to the C-terminal region of TDP-43 and modulates TDP-43 levels in H4 cells: characterization of inhibition by nucleic acids and 4-aminoquinolines. Biochim. Biophys. Acta 1834, 964–971 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  157. Buchan, J. R., Kolaitis, R. M., Taylor, J. P. & Parker, R. Eukaryotic stress granules are cleared by autophagy and CDC48/VCP function. Cell 153, 1461–1474 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  158. Hutton, M. et al. Association of missense and 5′-splice-site mutations in tau with the inherited dementia FTDP-17. Nature 393, 702–705 (1998).

    CAS  PubMed  Google Scholar 

  159. Nicolas, A. et al. Genome-wide analyses identify KIF5A as a novel ALS gene. Neuron 97, 1268–1283.e1266 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  160. Wu, C. H. et al. Mutations in the profilin 1 gene cause familial amyotrophic lateral sclerosis. Nature 488, 499–503 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  161. Puls, I. et al. Mutant dynactin in motor neuron disease. Nat. Genet. 33, 455–456 (2003).

    CAS  PubMed  Google Scholar 

  162. Smith, B. N. et al. Exome-wide rare variant analysis identifies TUBA4A mutations associated with familial ALS. Neuron 84, 324–331 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  163. Munch, C. et al. Point mutations of the p150 subunit of dynactin (DCTN1) gene in ALS. Neurology 63, 724–726 (2004).

    CAS  PubMed  Google Scholar 

  164. Landers, J. E. et al. Reduced expression of the kinesin-associated protein 3 (KIFAP3) gene increases survival in sporadic amyotrophic lateral sclerosis. Proc. Natl Acad. Sci. USA 106, 9004–9009 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  165. Magnani, E. et al. Interaction of tau protein with the dynactin complex. EMBO J. 26, 4546–4554 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  166. Ransohoff, R. M. How neuroinflammation contributes to neurodegeneration. Science 353, 777–783 (2016).

    CAS  PubMed  Google Scholar 

  167. Bemiller, S. M. et al. TREM2 deficiency exacerbates tau pathology through dysregulated kinase signaling in a mouse model of tauopathy. Mol. Neurodegener. 12, 74 (2017).

    PubMed  PubMed Central  Google Scholar 

  168. Broce, I. et al. Immune-related genetic enrichment in frontotemporal dementia: an analysis of genome-wide association studies. PLOS Med. 15, e1002487 (2018).

    PubMed  PubMed Central  Google Scholar 

  169. Becker, L. A. et al. Therapeutic reduction of ataxin-2 extends lifespan and reduces pathology in TDP-43 mice. Nature 544, 367–371 (2017). This study demonstrates that therapeutic reduction of ATXN2 delays the progression of TDP43-induced neurodegeneration in vivo.

    CAS  PubMed  PubMed Central  Google Scholar 

  170. Zhang, K. et al. Stress granule assembly disrupts nucleocytoplasmic transport. Cell 173, 958–971.e917 (2018). This article demonstrates a closely linked relationship between the accumulation of SGs and disruption of nuclear pore biology.

    CAS  PubMed  PubMed Central  Google Scholar 

  171. Smith, H. L. & Mallucci, G. R. The unfolded protein response: mechanisms and therapy of neurodegeneration. Brain 139, 2113–2121 (2016).

    PubMed  PubMed Central  Google Scholar 

  172. Moreno, J. A. et al. Sustained translational repression by eIF2α-P mediates prion neurodegeneration. Nature 485, 507–511 (2012). This study demonstrates that reducing eIF2α phosphorylation in vivo decreases neurodegeneration mediated by prions.

    CAS  PubMed  PubMed Central  Google Scholar 

  173. Ma, T. et al. Suppression of eIF2α kinases alleviates Alzheimer’s disease-related plasticity and memory deficits. Nat. Neurosci. 16, 1299–1305 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  174. Khurana, V. et al. TOR-mediated cell-cycle activation causes neurodegeneration in a Drosophila tauopathy model. Curr. Biol. 16, 230–241 (2006).

    CAS  PubMed  Google Scholar 

  175. Wang, I. F. et al. Autophagy activators rescue and alleviate pathogenesis of a mouse model with proteinopathies of the TAR DNA-binding protein 43. Proc. Natl Acad. Sci. USA 109, 15024–15029 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  176. Martinez, F. J. et al. Protein–RNA networks regulated by normal and ALS-associated mutant HnRNPA2B1 in the nervous system. Neuron 92, 780–795 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  177. Walter, H. & Brooks, D. E. Phase separation in cytoplasm, due to macromolecular crowding, is the basis for microcompartmentation. FEBS Lett. 361, 135–139 (1995).

    CAS  PubMed  Google Scholar 

  178. Toombs, J. A. et al. De novo design of synthetic prion domains. Proc. Natl Acad. Sci. USA 109, 6519–6524 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  179. Prilusky, J. et al. FoldIndex: a simple tool to predict whether a given protein sequence is intrinsically unfolded. Bioinformatics 21, 3435–3438 (2005).

    CAS  PubMed  Google Scholar 

  180. Wei, M. T. et al. Phase behaviour of disordered proteins underlying low density and high permeability of liquid organelles. Nat. Chem. 9, 1118–1125 (2017).

    CAS  PubMed  Google Scholar 

  181. Wang, J. et al. A molecular grammar governing the driving forces for phase separation of prion-like RNA binding proteins. Cell 174, 688–699.e616 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  182. Bergeron-Sandoval, L. P., Safaee, N. & Michnick, S. W. Mechanisms and consequences of macromolecular phase separation. Cell 165, 1067–1079 (2016).

    CAS  PubMed  Google Scholar 

  183. Feric, M. et al. Coexisting liquid phases underlie nucleolar subcompartments. Cell 165, 1686–1697 (2016). This article elegantly demonstrates that the nucleolus exists as a phase-separated structure, as well as demonstrating many key methods used in the field.

    CAS  PubMed  PubMed Central  Google Scholar 

  184. Prum, R. O., Dufresne, E. R., Quinn, T. & Waters, K. Development of colour-producing β-keratin nanostructures in avian feather barbs. J. R. Soc. Interface 6 (Suppl. 2), S253–S265 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  185. Milovanovic, D., Wu, Y., Bian, X. & De Camilli, P. A liquid phase of synapsin and lipid vesicles. Science 361, 604–607 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  186. Li, P. et al. Phase transitions in the assembly of multivalent signalling proteins. Nature 483, 336–340 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  187. Hnisz, D., Shrinivas, K., Young, R. A., Chakraborty, A. K. & Sharp, P. A. A phase separation model for transcriptional control. Cell 169, 13–23 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  188. Youn, J. Y. et al. High-density proximity mapping reveals the subcellular organization of mRNA-associated granules and bodies. Mol. Cell 69, 517–532.e511 (2018).

    CAS  PubMed  Google Scholar 

  189. Wolozin, B. The evolution of phase-separated TDP-43 in stress. Neuron 102, 265–267 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Contributions

B.W. researched data for the article. B.W. and P.I. provided substantial contributions to discussion of the article’s content, wrote the article, and reviewed and edited the manuscript before submission.

Corresponding author

Correspondence to Benjamin Wolozin.

Ethics declarations

Competing interests

B.W. is co-founder and chief scientific officer for Aquinnah Pharmaceuticals Inc. P.I. declares no competing interests.

Additional information

Peer review information

Nature Reviews Neuroscience thanks E. Buratti and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Glossary

Membraneless organelles

Macromolecular complexes composed of a large group of proteins that carry out specific functions and typically are observable via microscopy. Classic membraneless organelles are RNA granules and nuclear bodies that contain RNA-binding proteins and RNA.

Frontotemporal dementia

(FTD). An age-related degenerative disease in which neurons of the frontal cortex degenerate, commonly producing behavioural dysinhibition, followed ultimately by death. FTD most commonly occurs sporadically, but genetic forms are most frequently caused by mutations causing progranulin haploinsufficiency, mutations in MAPT (which encodes tau) or expansions of the G4C2 hexanucleotide repeat domain of C9ORF72.

Intrinsically disordered protein regions

Regions in proteins that have a low propensity to form secondary structures, such as α-helices or β-sheets. These regions often contain hydrophobic or low-complexity domains, and have a high propensity to aggregate.

Amyloid-β

(Aβ). A 4-kD peptide that is generated by cleavage of the amyloid precursor protein and accumulates as neuritic plaques in Alzheimer disease.

RNA recognition motifs

Domains that are present in RBPs that recognize a consensus sequence of RNA. The consensus sequence is typically single-stranded and about six to eight bases long. RBPs typically have one to three RNA recognition motifs.

Pre-initiation complexes

(PICs). PICs contain mRNA, the 40S ribosomal complex and associated translation initiation factors. Under optimal conditions, the PIC combines with the 60S ribosomal subunit to produce 80S ribosomes, which enables the translation of nascent proteins. Under stress, the PIC is bound by other RNA-binding proteins, promoting stress granule formation.

Polysomes

Complexes (also known as polyribosomes) on mRNA molecules that are formed by two or more ribosomes that are synchronously producing a new protein. The polysome represents the most actively translating fraction of a translational complex.

40S ribosomal subunit

A small subunit of the eukaryotic 80S ribosome that contains multiple ribosomal proteins designated by the term RPS, and that is a fundamental component of the pre-initiation complex, which binds mRNAs and interacts with translation initiation factors. The 40S subunit plays a key role in recognition of the start AUG codon on mRNA.

tRNA-derived stress-induced RNAs

(tiRNAs). Small non-coding RNAs produced by the ribonuclease angiogenin in response to stress. They represent the 5′ and 3′ halves of mature cytoplasmic tRNAs and regulate multiple aspects of RNA metabolism, including protein synthesis and stress granule formation.

Low-complexity domains

(LCDs). Protein domains that contain a small number of different types of amino acids. The LCDs that characterize RNA-binding proteins tend to contain alanine, glycine, glutamine and proline residues.

Chronic traumatic encephalopathy

A type of neurodegeneration that appears years after exposure to brain trauma, typically resulting from repetitive insults; it is characterized by the presence of neurofibrillary tangles or aggregated TAR DNA-binding protein 43. The pathology tends to begin at the bases of neuronal sulci where the physical force of the trauma was concentrated.

Oligomers

Small complexes composed of 2 to ~10 subunits of the same protein that are tightly associated in a repetitive manner. Oligomers appear to be more toxic than fibrils, perhaps because oligomers are smaller and more mobile than fibrils.

Fibrils

A large macromolecular complex of proteins that stack in a regular array of repetitive oligomers. Multiple fibrils commonly coalesce to form the hallmark structures of aggregated proteins that are observed in individuals with neurodegenerative diseases.

β-sheet

A highly stable protein structure that can stack to form large macromolecular fibrils.

Low-complexity aromatic-rich kinked segments

These segments exhibit a moderate affinity for like regions and might be a chemical structure that enables liquid–liquid phase separation. Their moderate affinity promotes a dynamic association that advances the coalescence of like proteins in granules that still exhibit extensive protein movement.

60S ribosomal subunit

The large subunit of the eukaryotic 80S ribosome, which contains ribosomal proteins that are designated by the term RPL and attaches to translationally competent PICs. It contains a peptidyl transferase centre, catalysing the addition of amino acids onto the nascent peptides during translation.

Propagation

The process by which disease pathology spreads among neurons in a cell-dependent manner. With propagation, a pathological aggregate secreted by a cell (classically a neuron or microglial cell) or injected into the neuropile seeds a similar aggregate in adjacent cells. These cells can then form new aggregated protein from endogenous stores of the same protein, secrete the newly aggregated protein and seed aggregation in yet another cell. In this manner, disease pathology can propagate via multiple stages of progressive seeding.

Templating

A process in which a particular conformation of a protein acts to induce a similar conformation in like proteins with which it comes in contact.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wolozin, B., Ivanov, P. Stress granules and neurodegeneration. Nat Rev Neurosci 20, 649–666 (2019). https://doi.org/10.1038/s41583-019-0222-5

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41583-019-0222-5

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