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.

  • Article
  • Published:

Atomic structures of TDP-43 LCD segments and insights into reversible or pathogenic aggregation

An Author Correction to this article was published on 17 September 2019

This article has been updated

Abstract

The normally soluble TAR DNA-binding protein 43 (TDP-43) is found aggregated both in reversible stress granules and in irreversible pathogenic amyloid. In TDP-43, the low-complexity domain (LCD) is believed to be involved in both types of aggregation. To uncover the structural origins of these two modes of β-sheet-rich aggregation, we have determined ten structures of segments of the LCD of human TDP-43. Six of these segments form steric zippers characteristic of the spines of pathogenic amyloid fibrils; four others form LARKS, the labile amyloid-like interactions characteristic of protein hydrogels and proteins found in membraneless organelles, including stress granules. Supporting a hypothetical pathway from reversible to irreversible amyloid aggregation, we found that familial ALS variants of TDP-43 convert LARKS to irreversible aggregates. Our structures suggest how TDP-43 adopts both reversible and irreversible β-sheet aggregates and the role of mutation in the possible transition of reversible to irreversible pathogenic aggregation.

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: Segments from the LCD of TDP-43 form steric zippers.
Fig. 2: Validation of steric-zipper-forming segments.
Fig. 3: The 312NFGAFS317 segments form a kinked β-sheet that is strengthened by familial variants A315T and A315E.
Fig. 4: Speculative model showing the alternative pathways of formation of stress granules with pathogenic amyloid.

Similar content being viewed by others

Change history

  • 17 September 2019

    An amendment to this paper has been published and can be accessed via a link at the top of the paper.

References

  1. 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 

  2. Kato, M. et al. Cell-free formation of RNA granules: low complexity sequence domains form dynamic fibers within hydrogels. Cell 149, 753–767 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Conicella, A. E., Zerze, G. H., Mittal, J. & Fawzi, N. L. ALS mutations disrupt phase separation mediated by α-helical structure in the TDP-43 low-complexity C-terminal domain. Structure 24, 1537–1549 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Kato, M. & McKnight, S. L. A solid-state conceptualization of information transfer from gene to message to protein. Annu. Rev. Biochem. https://doi.org/10.1146/annurev-biochem-061516-044700 (2017).

    CAS  PubMed  Google Scholar 

  5. Marcotte, E. M., Pellegrini, M., Yeates, T. O. & Eisenberg, D. A census of protein repeats. J. Mol. Biol. 293, 151–160 (1999).

    CAS  PubMed  Google Scholar 

  6. Dyson, H. J. & Wright, P. E. Intrinsically unstructured proteins and their functions. Nat. Rev. Mol. Cell Biol. 6, 197–208 (2005).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  8. March, Z. M., King, O. D. & Shorter, J. Prion-like domains as epigenetic regulators, scaffolds for subcellular organization, and drivers of neurodegenerative disease. Brain Res. 1647, 9–18 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Bentmann, E. et al. Requirements for stress granule recruitment of fused in sarcoma (FUS) and TAR DNA-binding protein of 43 kDa (TDP-43). J. Biol. Chem. 287, 23079–23094 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Colombrita, C. et al. TDP-43 is recruited to stress granules in conditions of oxidative insult. J. Neurochem. 111, 1051–1061 (2009).

    CAS  PubMed  Google Scholar 

  11. Dewey, C. M. et al. TDP-43 is directed to stress granules by sorbitol, a novel physiological osmotic and oxidative stressor. Mol. Cell. Biol. 31, 1098–1108 (2011).

    CAS  PubMed  Google Scholar 

  12. 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 

  13. Dang, Y. et al. Eukaryotic initiation factor 2α-independent pathway of stress granule induction by the natural product pateamine A. J. Biol. Chem. 281, 32870–32878 (2006).

    CAS  PubMed  Google Scholar 

  14. 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 

  15. 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 

  16. Reineke, L. C. et al. Casein kinase 2 is linked to stress granule dynamics through phosphorylation of the stress granule nucleating protein G3BP1. Mol. Cell. Biol. 37, e00596–e16 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Murray, D. T. et al. Structure of FUS protein fibrils and its relevance to self-assembly and phase separation of low-complexity domains. Cell 171, 615–627.e16 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Wheeler, J. R., Matheny, T., Jain, S., Abrisch, R. & Parker, R. Distinct stages in stress granule assembly and disassembly. eLife 5, 1–25 (2016).

    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. Lin, Y., Protter, D. S. W., 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 

  21. Hughes, M. P. et al. Atomic structures of low-complexity protein segments reveal kinked β sheets that assemble networks. Science 359, 698–701 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Eisenberg, D. S. & Sawaya, M. R. Structural studies of amyloid proteins at the molecular level. Annu. Rev. Biochem. 86, 69–95 (2017).

    CAS  PubMed  Google Scholar 

  23. Knowles, T. P. J., Vendruscolo, M. & Dobson, C. M. The amyloid state and its association with protein misfolding diseases. Nat. Rev. Mol. Cell Biol. 15, 384–396 (2014).

    CAS  PubMed  Google Scholar 

  24. Meersman, F. & Dobson, C. M. Probing the pressure-temperature stability of amyloid fibrils provides new insights into their molecular properties. Biochim. Biophys. Acta. 1764, 452–460 (2006).

    CAS  PubMed  Google Scholar 

  25. Harrison, A. F. & Shorter, J. RNA-binding proteins with prion-like domains in health and disease. Biochem. J. 474, 1417–1438 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Chang, C. K. et al. The N-terminus of TDP-43 promotes its oligomerization and enhances DNA binding affinity. Biochem. Biophys. Res. Commun. 425, 219–224 (2012).

    CAS  PubMed  Google Scholar 

  27. Lukavsky, P. J. et al. Molecular basis of UG-rich RNA recognition by the human splicing factor TDP-43. Nat. Struct. Mol. Biol. 20, 1443–1449 (2013).

    CAS  PubMed  Google Scholar 

  28. Saini, A. & Chauhan, V. S. Self-assembling properties of peptides derived from TDP-43 C-terminal fragment. Langmuir 30, 3845–3856 (2014).

    CAS  PubMed  Google Scholar 

  29. Buratti, E. & Baralle, F. E. The multiple roles of TDP-43 in pre-mRNA processing and gene expression regulation. RNA Biol. 7, 420–429 (2010).

    CAS  PubMed  Google Scholar 

  30. Lee, E. B., Lee, V. M. Y. & Trojanowski, J. Q. Gains or losses: molecular mechanisms of TDP43-mediated neurodegeneration. Nat. Rev. Neurosci. 13, 38–50 (2011).

    PubMed  PubMed Central  Google Scholar 

  31. Schwab, C., Arai, T., Hasegawa, M., Yu, S. & McGeer, P. L. Colocalization of transactivation-responsive DNA-binding protein 43 and huntingtin in inclusions of Huntington disease. J. Neuropathol. Exp. Neurol. 67, 1159–1165 (2008).

    PubMed  Google Scholar 

  32. Amador-Ortiz, C. et al. TDP-43 immunoreactivity in hippocampal sclerosis and Alzheimer’s disease. Ann. Neurol. 61, 435–445 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Nakashima-Yasuda, H. et al. Co-morbidity of TDP-43 proteinopathy in Lewy body related diseases. Acta Neuropathol. 114, 221–229 (2007).

    CAS  PubMed  Google Scholar 

  34. Higashi, S. et al. Concurrence of TDP-43, tau and α-synuclein pathology in brains of Alzheimer’s disease and dementia with Lewy bodies. Brain Res. 1184, 284–294 (2007).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  36. Kabashi, E. et al. TARDBP mutations in individuals with sporadic and familial amyotrophic lateral sclerosis. Nat. Genet. 40, 572–574 (2008).

    CAS  PubMed  Google Scholar 

  37. Chiang, C.-H. et al. Structural analysis of disease-related TDP-43 D169G mutation: linking enhanced stability and caspase cleavage efficiency to protein accumulation. Sci. Rep. 6, 21581 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Lagier-Tourenne, C., Polymenidou, M. & Cleveland, D. W. TDP-43 and FUS/TLS: emerging roles in RNA processing and neurodegeneration. Hum. Mol. Genet. 19R1, R46–R64 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Pesiridis, G. S., Lee, V. M. Y. & Trojanowski, J. Q. Mutations in TDP-43 link glycine-rich domain functions to amyotrophic lateral sclerosis. Hum. Mol. Genet. 18R1, R156–R162 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Igaz, L. M. et al. Expression of TDP-43 C-terminal fragments in vitro recapitulates pathological features of TDP-43 proteinopathies. J. Biol. Chem. 284, 8516–8524 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Hasegawa, M. et al. Phosphorylated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Ann. Neurol. 64, 60–70 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Ling, S. C., Polymenidou, M. & Cleveland, D. W. Converging mechanisms in ALS and FTD: disrupted RNA and protein homeostasis. Neuron 79, 416–438 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Ling, J. P., Pletnikova, O., Troncoso, J. C. & Wong, P. C. TDP-43 repression of nonconserved cryptic exons is compromised in ALS-FTD. Science 349, 650–655 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Yang, C. et al. The C-terminal TDP-43 fragments have a high aggregation propensity and harm neurons by a dominant-negative mechanism. PLoS One 5, e15878 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Chen, A. K. et al. Induction of amyloid fibrils by the C-terminal fragments of TDP-43 in amyotrophic lateral sclerosis. J. Am. Chem. Soc. 132, 1186–1187 (2010).

    CAS  PubMed  Google Scholar 

  46. Cushman, M., Johnson, B. S., King, O. D., Gitler, A. D. & Shorter, J. Prion-like disorders: blurring the divide between transmissibility and infectivity. J. Cell Sci. 123, 1191–1201 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Guo, W. et al. An ALS-associated mutation affecting TDP-43 enhances protein aggregation, fibril formation and neurotoxicity. Nat. Struct. Mol. Biol. 18, 822–830 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Jiang, L. L. et al. Structural transformation of the amyloidogenic core region of TDP-43 protein initiates its aggregation and cytoplasmic inclusion. J. Biol. Chem. 288, 19614–19624 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Jiang, L. L. et al. Two mutations G335D and Q343R within the amyloidogenic core region of TDP-43 influence its aggregation and inclusion formation. Sci. Rep. 6, 23928 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Mompeán, M. et al. “Structural characterization of the minimal segment of TDP-43 competent for aggregation”. Arch. Biochem. Biophys. 545, 53–62 (2014).

    PubMed  Google Scholar 

  51. Sawaya, M. R. et al. Atomic structures of amyloid cross-beta spines reveal varied steric zippers. Nature 447, 453–457 (2007).

    CAS  PubMed  Google Scholar 

  52. Goldschmidt, L., Teng, P. K., Riek, R. & Eisenberg, D. Identifying the amylome, proteins capable of forming amyloid-like fibrils. Proc. Natl. Acad. Sci. USA 107, 3487–3492 (2010).

    CAS  PubMed  Google Scholar 

  53. Budini, M. et al. Cellular model of TAR DNA-binding protein 43 (TDP-43) aggregation based on its C-terminal Gln/Asn-rich region. J. Biol. Chem. 287, 7512–7525 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Fuentealba, R. A. et al. Interaction with polyglutamine aggregates reveals a Q/N-rich domain in TDP-43. J. Biol. Chem. 285, 26304–26314 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Seidler, P. M. et al. Structure-based inhibitors of tau aggregation. Nat. Chem. 10, 170–176 (2018).

    CAS  PubMed  Google Scholar 

  56. Xiang, S. et al. The LC domain of hnRNPA2 adopts similar conformations in hydrogel polymers, liquid-like droplets, and nuclei. Cell 163, 829–839 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Gitcho, M. A. et al. TDP-43 A315T mutation in familial motor neuron disease. Ann. Neurol. 63, 535–538 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Xu, M. et al. Characterization of β-domains in C-terminal fragments of TDP-43 by scanning tunneling microscopy. J. Struct. Biol. 181, 11–16 (2013).

    CAS  PubMed  Google Scholar 

  59. Lim, L., Wei, Y., Lu, Y. & Song, J. ALS-Causing mutations significantly perturb the self-assembly and interaction with nucleic acid of the intrinsically disordered prion-like domain of TDP-43. PLoS Biol. 14, e1002338 (2016).

    PubMed  PubMed Central  Google Scholar 

  60. Estes, P. S. et al. Wild-type and A315T mutant TDP-43 exert differential neurotoxicity in a Drosophila model of ALS. Hum. Mol. Genet. 20, 2308–2321 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Walker, A. K. et al. ALS-associated TDP-43 induces endoplasmic reticulum stress, which drives cytoplasmic TDP-43 accumulation and stress granule formation. PLoS One 8, e81170 (2013).

    PubMed  PubMed Central  Google Scholar 

  62. Bosco, D. A. et al. Mutant FUS proteins that cause amyotrophic lateral sclerosis incorporate into stress granules. Hum. Mol. Genet. 19, 4160–4175 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Hackman, P. et al. Welander distal myopathy is caused by a mutation in the RNA-binding protein TIA1. Ann. Neurol. 73, 500–509 (2013).

    CAS  PubMed  Google Scholar 

  64. 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 

  65. Schmidt, H. B. & Rohatgi, R. In vivo formation of vacuolated multi-phase compartments lacking membranes. Cell Reports 16, 1228–1236 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Franzmann, T. M. et al. Phase separation of a yeast prion protein promotes cellular fitness. Science 359, eaao5654 (2018).

    PubMed  Google Scholar 

  67. 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 

  68. Soragni, A. et al. A designed inhibitor of p53 aggregation rescues p53 tumor suppression in ovarian carcinomas. Cancer Cell 29, 90–103 (2016).

    CAS  PubMed  Google Scholar 

  69. Winn, M. D. et al. Overview of the CCP4 suite and current developments. Acta Crystallogr. D Biol. Crystallogr. 67, 235–242 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Cheng, P.-N., Liu, C., Zhao, M., Eisenberg, D. & Nowick, J. S. Amyloid β-sheet mimics that antagonize protein aggregation and reduce amyloid toxicity. Nat. Chem. 4, 927–933 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Kabsch, W. Xds. Acta Crystallogr. D Biol. Crystallogr. 66, 125–132 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Wiltzius, J. J. et al. Atomic structure of the cross-beta spine of islet amyloid polypeptide (amylin). Protein Sci. 17, 1467–1474 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr. D Biol. Crystallogr. 60, 2126–2132 (2004).

    PubMed  Google Scholar 

  75. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr. D Biol. Crystallogr. 66, 213–221 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Sheldrick, G. M. A short history of SHELX. Acta Crystallogr. A 64, 112–122 (2008).

    CAS  PubMed  Google Scholar 

  77. Murshudov, G. N., Vagin, A. A. & Dodson, E. J. Refinement of macromolecular structures by the maximum-likelihood method. Acta Crystallogr. D Biol. Crystallogr. 53, 240–255 (1997).

    CAS  PubMed  Google Scholar 

  78. Otwinowski, Z. & Minor, W. Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276, 307–326 (1997).

    CAS  PubMed  Google Scholar 

  79. Kabsch, W. Automatic processing of rotation diffraction data from crystals of initially unknown symmetry and cell constants. J. Appl. Crystallogr. 26, 795–800 (1993).

    CAS  Google Scholar 

  80. Rodriguez, J. A. et al. Structure of the toxic core of α-synuclein from invisible crystals. Nature 525, 486–490 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Hattne, J. et al. MicroED data collection and processing. Acta Crystallogr. A Found. Adv. 71, 353–360 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  82. Shi, D. et al. The collection of MicroED data for macromolecular crystallography. Nat. Protoc. 11, 895–904 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Bricogne, G. et al. BUSTER v.2.9. (Global Phasing Ltd., Cambridge, 2010).

  84. Lawrence, M. C. & Colman, P. M. Shape complementarity at protein/protein interfaces. J. Mol. Biol. 234, 946–950 (1993).

    CAS  PubMed  Google Scholar 

  85. Collaborative Computational Project, Number 4. The CCP4 suite: programs for protein crystallography. Acta Crystallogr. D Biol. Crystallogr. 50, 760–763 (1994).

    Google Scholar 

Download references

Acknowledgements

We thank S. Sangwan and P. Seidler for discussion, D. Shi and T. Gonen at Janelia for microscope support and M. Collazo at UCLA-DOE Macromolecular Crystallization Core Technology Center for crystallization support. This work is based upon research conducted at the Northeastern Collaborative Access Team beamlines, which is funded by the National Institute of General Medical Sciences from the National Institutes of Health (P41 GM103403). The Eiger 16 M detector on 24-ID-E beam line is funded by a NIH-ORIP HEI grant (S10OD021527). This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02-06CH11357. This research was supported in part by grants from the National Institutes of Health NIH NIA AG029430, NIH NIA AG054022, Howard Hughes Medical Institute and the Janelia Research Campus visitor program. This material is based upon work supported by the National Science Foundation under Grant No. NSF 1616265. We acknowledge the use of instruments at the Electron Imaging Center for Nanomachines supported by UCLA and by instrumentation grants from NIH (1S10RR23057 and 1U24GM116792) and NSF (DBI-1338135). D.R.B. was supported by the National Science Foundation Graduate Research Fellowship Program.

Author information

Authors and Affiliations

Authors

Contributions

E.L.G., Q.C. and D.S.E. designed the project and wrote the manuscript with input from all other authors, especially M.R.S. E.L.G. and H.T. conducted fibril growth assays and prepared peptide crystals. E.L.G., Q.C. and J.L. cloned and purified TDP-43 constructs and performed the protein aggregation assay and fibril diffraction. M.P.H. predicted putative LARKS. E.L.G., H.T. and M.R.S. processed and solved 300GNNQGSN306. E.L.G. and M.R.S. processed and solved 312NFGAFS317, 312NFGTFS317, 370GNNSYS375 and 396GFNGGFG402. E.L.G., Q.C., J.L. and M.R.S. processed data and solved the 321AMMAAA326 and 328AALQSS333. J.A.R. collected MicroED data on 312NFGEFS317 and 333SWGMMGMLASQ343, M.R.S., D.C. and E.L.G. processed data and solved the structure of 312NFGEFS317 and 333SWGMMGMLASQ343. D.R.B. collected MicroED data for 312NFGpTFS317. Q.C., M.R.S., D.C. and D.R.B. processed data and solved the structure of 312NFGpTFS317. E.L.G., Q.C., M.R.S., M.P.H. and D.S.E. analyzed structures and designed the model of the LCD in SG formation and pathological aggregation.

Corresponding author

Correspondence to David S. Eisenberg.

Ethics declarations

Competing interests

D.S.E. is an advisor and equity shareholder in ADDRx, Inc.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Integrated Supplementary Information

Supplementary Figure 1 321AMMAAA326 and 333SWGMMGMLASQ343 form antiparallel steric-zipper structures.

Views perpendicular to the fibril axis (vertical) are shown for three steric-zipper structures of Fig. 1. 321AMMAAA326 and 333SWGMMGMLASQ343 are two zippers that adopt anti-parallel packing, as shown by the alternating orientation of arrows representing the direction of strands. 328AALOSS333 is shown as an example of the zippers which adopt parallel packing, as shown by the same orientation of arrows in each sheet.

Supplementary Figure 2 Validation of TDP-CTF and TDP-LCD constructs.

The SUMO tagged TDP-CTF and TDP-LCD were analyzed by SDS-PAGE and stained with coomassie blue and anti-TDP-43 antibody. For TDP-CTF, notice that 1 hour after (1h) ULP1 cleavage, it generates SUMO protein and tag-free TDP-CTF, and both SUMO-tagged and tag-free TDP-CTF can be recognized by anti-TDP-43 antibody. SUMO tagged TDP-LCD also can be recognized by anti-TDP-43 antibody, and generates SUMO protein and tag-free TDP-LCD after cleavage. The molecular weight of SUMO and tag-free TDP-LCD are so close to each other that the two bands are indistinguishable on SDS-PAGE when stained by coomassie blue. However, the existence of tag-free TDP-LCD can be detected by anti-TDP-43 antibody, forming a band (although very dim) at its expected position. Tag-free TDP-LCD can also be detected by SDS-PAGE of the TDP-LCD pellet, showing a band at its expected position, and SUMO protein remains soluble as shown in Fig. 2a.

Supplementary Figure 3 Positive control of denaturing conditions shown in Fig. 3b and structural details of 312NFGEFS317 and 312NFGpTFS317.

(A) Negative stain EM images of Aβ1-42 fibrils, typical amyloid fibrils, before and after treatment with 2% SDS and 70 °C for 15 mins. Notice that Aβ1-42 fibrils are stable and maintain their morphology after treatment in denaturing conditions. Scale bar = 200 nm. (B) Omit map of phosphate in 312NFGpTFS317 structure viewed from two angles. Green mesh shows the electron density of the Fo-Fc map with sigma > 4.5. Notice that the electron density of the same map with sigma <-4.5 is also computed as red mesh, however there is no observable negative density. TPO: phosphorylated threonine. (C) Hydrogen bond networks that stabilizes the packing of 312NFGEFS317 and 312NFGpTFS317. In each structure, three Glu/pThr residues from three adjacent strands of same sheet are shown, as well as the residues hydrogen-bonded to them. The hydrogen bonds formed by all Glu and the central pThr are shown as yellow dash lines, with distances between 2.3 and 3.2 Å.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–3 and Supplementary Tables 1 and 2

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guenther, E.L., Cao, Q., Trinh, H. et al. Atomic structures of TDP-43 LCD segments and insights into reversible or pathogenic aggregation. Nat Struct Mol Biol 25, 463–471 (2018). https://doi.org/10.1038/s41594-018-0064-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41594-018-0064-2

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