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:

Sub-3-Å cryo-EM structure of RNA enabled by engineered homomeric self-assembly

Abstract

High-resolution structural studies are essential for understanding the folding and function of diverse RNAs. Herein, we present a nanoarchitectural engineering strategy for efficient structural determination of RNA-only structures using single-particle cryogenic electron microscopy (cryo-EM). This strategy—ROCK (RNA oligomerization-enabled cryo-EM via installing kissing loops)—involves installing kissing-loop sequences onto the functionally nonessential stems of RNAs for homomeric self-assembly into closed rings with multiplied molecular weights and mitigated structural flexibility. ROCK enables cryo-EM reconstruction of the Tetrahymena group I intron at 2.98-Å resolution overall (2.85 Å for the core), allowing de novo model building of the complete RNA, including the previously unknown peripheral domains. ROCK is further applied to two smaller RNAs—the Azoarcus group I intron and the FMN riboswitch, revealing the conformational change of the former and the bound ligand in the latter. ROCK holds promise to greatly facilitate the use of cryo-EM in RNA structural studies.

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: Conceptualization and workflow of ROCK.
Fig. 2: Construct engineering of the Tetrahymena group I intron (TetGI) as a case study for ROCK.
Fig. 3: Sub-3-Å-resolution cryo-EM map of TetGI-DS, a dimeric pre-2SΔ5′ex construct of the TetGI, enabled by ROCK.
Fig. 4: Assembly, activity and cryo-EM analyses of TetGI-D and TetGI-T.
Fig. 5: Structural insights gained from the cryo-EM structures of the TetGI.
Fig. 6: The homomeric self-assembly strategy applied to two smaller structured RNAs.

Similar content being viewed by others

Data availability

The data supporting the findings of this study are principally within the figures and the associated Supplementary Information. Atomic coordinates and cryo-EM maps have been deposited with the Protein Data Bank and the Electron Microscopy Data Bank under the accession codes: 7R6L and EMD-24281 for TetGI-DS, 7R6M and EMD-24282 for TetGI-D, 7R6N and EMD-24283 for TetGI-T, EMD-24284 for AzoGI-T and EMD-24285 for FMNrbsw-T. In this study, the following structures from the PDB were utilized: 2BJ2, 1X8W, 1U6B and 3F2Q.

References

  1. Mortimer, S. A., Kidwell, M. A. & Doudna, J. A. Insights into RNA structure and function from genome-wide studies. Nat. Rev. Genet. 15, 469–479 (2014).

    Article  CAS  PubMed  Google Scholar 

  2. Serganov, A. & Patel, D. J. Ribozymes, riboswitches and beyond: regulation of gene expression without proteins. Nat. Rev. Genet. 8, 776–790 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Hangauer, M. J., Vaughn, I. W. & McManus, M. T. Pervasive transcription of the human genome produces thousands of previously unidentified long intergenic noncoding RNAs. PLoS Genet. 9, e1003569 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Robertson, D. L. & Joyce, G. F. Selection in vitro of an RNA enzyme that specifically cleaves single-stranded DNA. Nature 344, 467–468 (1990).

    Article  CAS  PubMed  Google Scholar 

  5. Ellington, A. D. & Szostak, J. W. In vitro selection of RNA molecules that bind specific ligands. Nature 346, 818–822 (1990).

    Article  CAS  PubMed  Google Scholar 

  6. Tuerk, C. & Gold, L. Systematic evolution of ligands by exponential enrichment: RNA ligands to bacteriophage T4 DNA polymerase. Science 249, 505–510 (1990).

    Article  CAS  PubMed  Google Scholar 

  7. Zhang, J. & Ferre-D’Amare, A. R. New molecular engineering approaches for crystallographic studies of large RNAs. Curr. Opin. Struct. Biol. 26, 9–15 (2014).

    Article  PubMed  CAS  Google Scholar 

  8. Hendrickson, W. A., Horton, J. R. & LeMaster, D. M. Selenomethionyl proteins produced for analysis by multiwavelength anomalous diffraction (MAD): a vehicle for direct determination of three-dimensional structure. EMBO J. 9, 1665–1672 (1990).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Nogales, E. The development of cryo-EM into a mainstream structural biology technique. Nat. Methods 13, 24–27 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Qu, G. et al. Structure of a group II intron in complex with its reverse transcriptase. Nat. Struct. Mol. Biol. 23, 549–557 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Li, S. et al. Structural basis of amino acid surveillance by higher-order tRNA-mRNA interactions. Nat. Struct. Mol. Biol. 26, 1094–1105 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Zhang, K. et al. Cryo-EM structure of a 40 kDa SAM-IV riboswitch RNA at 3.7 A resolution. Nat. Commun. 10, 5511 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Kappel, K. et al. Accelerated cryo-EM-guided determination of three-dimensional RNA-only structures. Nat. Methods 17, 699–707 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Hyeon, C., Dima, R. I. & Thirumalai, D. Size, shape, and flexibility of RNA structures. J. Chem. Phys. 125, 194905 (2006).

    Article  PubMed  CAS  Google Scholar 

  15. Seeman, N. C. Nanomaterials based on DNA. Annu. Rev. Biochem. 79, 65–87 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Seeman, N. C. Structural DNA Nanotechnology (Cambridge University Press, 2016).

  17. Seeman, N. C. & Sleiman, H. F. DNA nanotechnology. Nat. Rev. Mater. 3, 17068 (2017).

    Article  CAS  Google Scholar 

  18. Guo, P. The emerging field of RNA nanotechnology. Nat. Nanotechnol. 5, 833–842 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Grabow, W. W. & Jaeger, L. RNA self-assembly and RNA nanotechnology. Acc. Chem. Res. 47, 1871–1880 (2014).

    Article  CAS  PubMed  Google Scholar 

  20. Kruger, K. et al. Self-splicing RNA: autoexcision and autocyclization of the ribosomal RNA intervening sequence of tetrahymena. Cell 31, 147–157 (1982).

    Article  CAS  PubMed  Google Scholar 

  21. Pujari, N. et al. Engineering crystal packing in RNA structures I: past and future strategies for engineering RNA packing in crystals. Cryst. 11, 8 (2021).

    Google Scholar 

  22. Ferre-D’Amare, A. R., Zhou, K. & Doudna, J. A. A general module for RNA crystallization. J. Mol. Biol. 279, 621–631 (1998).

    Article  PubMed  Google Scholar 

  23. Ferre-D’Amare, A. R. & Doudna, J. A. Crystallization and structure determination of a hepatitis delta virus ribozyme: use of the RNA-binding protein U1A as a crystallization module. J. Mol. Biol. 295, 541–556 (2000).

    Article  PubMed  CAS  Google Scholar 

  24. Ferre-D’Amare, A. R. Use of the spliceosomal protein U1A to facilitate crystallization and structure determination of complex RNAs. Methods 52, 159–167 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  25. Ye, J. D. et al. Synthetic antibodies for specific recognition and crystallization of structured RNA. Proc. Natl Acad. Sci. USA 105, 82–87 (2008).

    Article  CAS  PubMed  Google Scholar 

  26. Koldobskaya, Y. et al. A portable RNA sequence whose recognition by a synthetic antibody facilitates structural determination. Nat. Struct. Mol. Biol. 18, 100–106 (2011).

    Article  CAS  PubMed  Google Scholar 

  27. Lee, A. J. & Crothers, D. M. The solution structure of an RNA loop–loop complex: the ColE1 inverted loop sequence. Structure 6, 993–1007 (1998).

    Article  CAS  PubMed  Google Scholar 

  28. Goodsell, D. S. & Olson, A. J. Structural symmetry and protein function. Annu. Rev. Biophys. Biomol. Struct. 29, 105–153 (2000).

    Article  CAS  PubMed  Google Scholar 

  29. Jones, C. P. & Ferre-D’Amare, A. R. RNA quaternary structure and global symmetry. Trends Biochem. Sci. 40, 211–220 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Bou-Nader, C. & Zhang, J. Structural Insights into RNA dimerization: motifs, interfaces and functions. Molecules 25, 12 (2020).

    Article  CAS  Google Scholar 

  31. Bindewald, E., Grunewald, C., Boyle, B., O’Connor, M. & Shapiro, B. A. Computational strategies for the automated design of RNA nanoscale structures from building blocks using NanoTiler. J. Mol. Graph Model 27, 299–308 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Hougland, J. L., Piccirilli, J. A., Forconi, M., Lee, J. & Herschlag, D. in RNA World 3rd edn (eds Gesteland, R. F., Atkins, J. F. & Cech, T. R.) 133–205 (Cold Spring Harbor Laboratory Press, 2006).

  33. Golden, B. L. in Ribozymes and RNA Catalysis 178–200 (The Royal Society of Chemistry, 2007).

  34. Michel, F. & Westhof, E. Modelling of the three-dimensional architecture of group I catalytic introns based on comparative sequence analysis. J. Mol. Biol. 216, 585–610 (1990).

    Article  CAS  PubMed  Google Scholar 

  35. Lehnert, V., Jaeger, L., Michel, F. & Westhof, E. New loop-loop tertiary interactions in self-splicing introns of subgroup IC and ID: a complete 3D model of the Tetrahymena thermophila ribozyme. Chem. Biol. 3, 993–1009 (1996).

    Article  CAS  PubMed  Google Scholar 

  36. Cate, J. H. et al. Crystal structure of a group I ribozyme domain: principles of RNA packing. Science 273, 1678–1685 (1996).

    Article  CAS  PubMed  Google Scholar 

  37. Juneau, K., Podell, E., Harrington, D. J. & Cech, T. R. Structural basis of the enhanced stability of a mutant ribozyme domain and a detailed view of RNA–solvent interactions. Structure 9, 221–231 (2001).

    Article  CAS  PubMed  Google Scholar 

  38. Golden, B. L., Gooding, A. R., Podell, E. R. & Cech, T. R. A preorganized active site in the crystal structure of the Tetrahymena ribozyme. Science 282, 259–264 (1998).

    Article  CAS  PubMed  Google Scholar 

  39. Guo, F., Gooding, A. R. & Cech, T. R. Structure of the Tetrahymena ribozyme: base triple sandwich and metal ion at the active site. Mol. Cell 16, 351–362 (2004).

    CAS  PubMed  Google Scholar 

  40. Adams, P. L., Stahley, M. R., Kosek, A. B., Wang, J. & Strobel, S. A. Crystal structure of a self-splicing group I intron with both exons. Nature 430, 45–50 (2004).

    Article  CAS  PubMed  Google Scholar 

  41. Zaug, A. J., Been, M. D. & Cech, T. R. The Tetrahymena ribozyme acts like an RNA restriction endonuclease. Nature 324, 429–433 (1986).

    Article  CAS  PubMed  Google Scholar 

  42. Inoue, T., Sullivan, F. X. & Cech, T. R. New reactions of the ribosomal RNA precursor of Tetrahymena and the mechanism of self-splicing. J. Mol. Biol. 189, 143–165 (1986).

    Article  CAS  PubMed  Google Scholar 

  43. Woodson, S. A. Metal ions and RNA folding: a highly charged topic with a dynamic future. Curr. Opin. Chem. Biol. 9, 104–109 (2005).

    Article  CAS  PubMed  Google Scholar 

  44. Liu, D. et al. Branched kissing loops for the construction of diverse RNA homooligomeric nanostructures. Nat. Chem. 12, 249–259 (2020).

    Article  CAS  PubMed  Google Scholar 

  45. Rook, M. S., Treiber, D. K. & Williamson, J. R. An optimal Mg2+ concentration for kinetic folding of the Tetrahymena ribozyme. Proc. Natl Acad. Sci. U. S. A. 96, 12471–12476 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Herschlag, D. & Cech, T. R. Catalysis of RNA cleavage by the Tetrahymena thermophila ribozyme. 2. Kinetic description of the reaction of an RNA substrate that forms a mismatch at the active site. Biochemistry 29, 10172–10180 (1990).

    Article  CAS  PubMed  Google Scholar 

  47. Pyle, A. M. & Cech, T. R. Ribozyme recognition of RNA by tertiary interactions with specific ribose 2′-OH groups. Nature 350, 628–631 (1991).

    Article  CAS  PubMed  Google Scholar 

  48. Herschlag, D., Eckstein, F. & Cech, T. R. Contributions of 2′-hydroxyl groups of the RNA substrate to binding and catalysis by the Tetrahymena ribozyme. An energetic picture of an active site composed of RNA. Biochemistry 32, 8299–8311 (1993).

    Article  CAS  PubMed  Google Scholar 

  49. Strobel, S. A. & Cech, T. R. Tertiary interactions with the internal guide sequence mediate docking of the P1 helix into the catalytic core of the Tetrahymena ribozyme. Biochemistry 32, 13593–13604 (1993).

    Article  CAS  PubMed  Google Scholar 

  50. Golden, B. L., Kim, H. & Chase, E. Crystal structure of a phage Twort group I ribozyme-product complex. Nat. Struct. Mol. Biol. 12, 82–89 (2005).

    Article  CAS  PubMed  Google Scholar 

  51. Barfod, E. T. & Cech, T. R. Deletion of nonconserved helices near the 3′ end of the rRNA intron of Tetrahymena thermophila alters self-splicing but not core catalytic activity. Genes Dev. 2, 652–663 (1988).

    Article  CAS  PubMed  Google Scholar 

  52. Laggerbauer, B., Murphy, F. L. & Cech, T. R. Two major tertiary folding transitions of the Tetrahymena catalytic RNA. EMBO J. 13, 2669–2676 (1994).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Denesyuk, N. A. & Thirumalai, D. How do metal ions direct ribozyme folding? Nat. Chem. 7, 793–801 (2015).

    Article  CAS  PubMed  Google Scholar 

  54. Stahley, M. R. & Strobel, S. A. Structural evidence for a two-metal-ion mechanism of group I intron splicing. Science 309, 1587–1590 (2005).

    Article  CAS  PubMed  Google Scholar 

  55. Piccirilli, J. A., Vyle, J. S., Caruthers, M. H. & Cech, T. R. Metal ion catalysis in the Tetrahymena ribozyme reaction. Nature 361, 85–88 (1993).

    Article  CAS  PubMed  Google Scholar 

  56. Shan, S., Yoshida, A., Sun, S., Piccirilli, J. A. & Herschlag, D. Three metal ions at the active site of the Tetrahymena group I ribozyme. Proc. Natl Acad. Sci. USA 96, 12299–12304 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Yoshida, A., Sun, S. & Piccirilli, J. A. A new metal ion interaction in the Tetrahymena ribozyme reaction revealed by double sulfur substitution. Nat. Struct. Biol. 6, 318–321 (1999).

    Article  CAS  PubMed  Google Scholar 

  58. Kuo, L. Y. & Piccirilli, J. A. Leaving group stabilization by metal ion coordination and hydrogen bond donation is an evolutionarily conserved feature of group I introns. Biochim. Biophys. Acta 1522, 158–166 (2001).

    Article  CAS  PubMed  Google Scholar 

  59. Weinstein, L. B., Jones, B. C., Cosstick, R. & Cech, T. R. A second catalytic metal ion in group I ribozyme. Nature 388, 805–808 (1997).

    Article  CAS  PubMed  Google Scholar 

  60. Lipchock, S. V. & Strobel, S. A. A relaxed active site after exon ligation by the group I intron. Proc. Natl Acad. Sci. USA 105, 5699–5704 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Serganov, A., Huang, L. & Patel, D. J. Coenzyme recognition and gene regulation by a flavin mononucleotide riboswitch. Nature 458, 233–237 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Wilt, H. M., Yu, P., Tan, K., Wang, Y. X. & Stagno, J. R. FMN riboswitch aptamer symmetry facilitates conformational switching through mutually exclusive coaxial stacking configurations. J. Struct. Biol. X 4, 100035 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Michel, F. & Costa, M. Inferring RNA structure by phylogenetic and genetic analyses. Cold Spring Harb. Monogr. Ser. 35, 175–202 (1998). %@ 0270-1847.

    CAS  Google Scholar 

  64. Schön, P. Imaging and force probing RNA by atomic force microscopy. Methods 103, 25–33 (2016).

    Article  PubMed  CAS  Google Scholar 

  65. Chen, Y. & Pollack, L. SAXS studies of RNA: structures, dynamics, and interactions with partners. Wiley Interdiscip. Rev. RNA 7, 512–526 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Lilley, D. M. Analysis of branched nucleic acid structure using comparative gel electrophoresis. Q. Rev. Biophys. 41, 1–39 (2008).

    Article  CAS  PubMed  Google Scholar 

  67. Su, Z. et al. Cryo-EM structures of full-length Tetrahymena ribozyme at 3.1 A resolution. Nature 596, 603–607 2021.

  68. Wang, H. W. & Wang, J. W. How cryo-electron microscopy and X-ray crystallography complement each other. Protein Sci. 26, 32–39 (2017).

    Article  CAS  PubMed  Google Scholar 

  69. Warner, K. D., Hajdin, C. E. & Weeks, K. M. Principles for targeting RNA with drug-like small molecules. Nat. Rev. Drug Discov. 17, 547–558 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Tomizawa, J.-i Control of ColE1 plasmid replication: the process of binding of RNA I to the primer transcript. Cell 38, 861–870 (1984).

    Article  CAS  PubMed  Google Scholar 

  71. Zuker, M. Mfold web server for nucleic acid folding and hybridization prediction. Nucleic Acids Res. 31, 3406–3415 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Schorb, M., Haberbosch, I., Hagen, W. J. H., Schwab, Y. & Mastronarde, D. N. Software tools for automated transmission electron microscopy. Nat. Methods 16, 471–477 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Zheng, S. Q. et al. MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Nat. Methods 14, 331–332 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Rohou, A. & Grigorieff, N. CTFFIND4: fast and accurate defocus estimation from electron micrographs. J. Struct. Biol. 192, 216–221 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  75. Shaikh, T. R. et al. SPIDER image processing for single-particle reconstruction of biological macromolecules from electron micrographs. Nat. Protoc. 3, 1941–1974 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Zivanov, J. et al. New tools for automated high-resolution cryo-EM structure determination in RELION-3. eLife 7, e42166 (2018).

  77. Scheres, S. H. Processing of structurally heterogeneous cryo-EM data in RELION. Methods Enzymol. 579, 125–157 (2016).

    Article  CAS  PubMed  Google Scholar 

  78. Kucukelbir, A., Sigworth, F. J. & Tagare, H. D. Quantifying the local resolution of cryo-EM density maps. Nat. Methods 11, 63–65 (2014).

    Article  CAS  PubMed  Google Scholar 

  79. Terwilliger, T. C., Ludtke, S. J., Read, R. J., Adams, P. D. & Afonine, P. V. Improvement of cryo-EM maps by density modification. Nat. Methods 17, 923–927 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Pettersen, E. F. et al. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004).

    Article  CAS  PubMed  Google Scholar 

  82. Afonine, P. V. et al. Real-space refinement in PHENIX for cryo-EM and crystallography. Acta Crystallogr D. Struct. Biol. 74, 531–544 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  84. McGreevy, R., Teo, I., Singharoy, A. & Schulten, K. Advances in the molecular dynamics flexible fitting method for cryo-EM modeling. Methods 100, 50–60 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Hahn, C. S., Strauss, E. G. & Strauss, J. H. Dideoxy sequencing of RNA using reverse transcriptase. Methods Enzymol. 180, 121–130 (1989).

    Article  CAS  PubMed  Google Scholar 

  86. Johnston, W. K., Unrau, P. J., Lawrence, M. S., Glasner, M. E. & Bartel, D. P. RNA-catalyzed RNA polymerization: accurate and general RNA-templated primer extension. Science 292, 1319–1325 (2001).

    Article  CAS  PubMed  Google Scholar 

  87. Zaug, A. J., Kent, J. R. & Cech, T. R. A labile phosphodiester bond at the ligation junction in a circular intervening sequence RNA. Science 224, 574–578 (1984).

    Article  CAS  PubMed  Google Scholar 

  88. Ennifar, E., Walter, P., Ehresmann, B., Ehresmann, C. & Dumas, P. Crystal structures of coaxially stacked kissing complexes of the HIV-1 RNA dimerization initiation site. Nat. Struct. Biol. 8, 1064–1068 (2001).

    Article  CAS  PubMed  Google Scholar 

  89. Lebars, I. et al. Exploring TAR-RNA aptamer loop-loop interaction by X-ray crystallography, UV spectroscopy and surface plasmon resonance. Nucleic Acids Res. 36, 7146–7156 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Lilley, D. M. Structures of helical junctions in nucleic acids. Q. Rev. Biophys. 33, 109–159 (2000).

    Article  CAS  PubMed  Google Scholar 

  91. Szewczak, A. A. et al. An important base triple anchors the substrate helix recognition surface within the Tetrahymena ribozyme active site. Proc. Natl Acad. Sci. USA 96, 11183–11188 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Suh, S. O., Jones, K. G. & Blackwell, M. A group I intron in the nuclear small subunit rRNA gene of Cryptendoxyla hypophloia, an ascomycetous fungus: evidence for a new major class of group I introns. J. Mol. Evol. 48, 493–500 (1999).

    Article  CAS  PubMed  Google Scholar 

  93. Leontis, N. B. & Westhof, E. A common motif organizes the structure of multi-helix loops in 16S and 23S ribosomal RNAs. J. Mol. Biol. 283, 571–583 (1998).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank E. Westhof for providing the computer model of the complete TetGI and information on the peripheral domains of group I introns, T. Cech and F. Guo for information on the X-ray model of the TetGI core, M. Dai and Y. Shao for helpful discussions, S. Stoilova-McPhie for the help in pilot EM experiments, and M. Ziegler for proofreading. D.L. is a Merck Fellow of the Life Sciences Research Foundation. This work was supported by NSF grants (CMMI-1333215, CMMI-1344915 and CBET-1729397), AFOSR grant (MURI FATE, FA9550-15-1-0514), NIH grant (5DP1GM133052) and Molecular Robotics Initiative fund from the Wyss Institute to P.Y., a NIH grant (R01GM122797) to M.L. and a NIH grant (R01GM102489) to J.A.P.

Author information

Authors and Affiliations

Authors

Contributions

D.L. conceived and designed the study, designed and prepared the constructs, performed the functional assays, conducted pilot EM experiments, built and refined the atomic models and drafted the manuscript. F.A.T. collected the EM data, performed the reconstruction and generated the maps, assisted in building atomic models and drafted the manuscript. J.A.P. advised the experiments and drafted the manuscript. M.L. conceived, designed and supervised the study, performed the reconstruction and generated the maps. P.Y. conceived, designed and supervised the study and drafted the manuscript. All authors analyzed the data and commented on the manuscript.

Corresponding authors

Correspondence to Maofu Liao or Peng Yin.

Ethics declarations

Competing interests

A provisional patent related to this work has been filed with D.L., F.A.T., M.L. and P.Y. listed as coinventors. J.A.P. declares no competing interests.

Peer review

Peer review information

Nature Methods thanks the anonymous reviewers for their contribution to the peer review of this work. Rita Strack was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

Additional information

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

Extended data

Extended Data Fig. 1 The benefits of RNA oligomerization-enabled cryo-EM via installing kissing-loops (ROCK) and the process of screening the lengths of connector helices for ring closure.

a, The folding wild-type monomeric RNA leads to correctly folded RNA and species of misfolded or alternative conformations: the former is a difficult subject for cryo-EM structural determination due to small size and structural flexibility and the latter are difficult to eliminate. b, The engineered RNA construct with kissing-loop sequences installed folds and assembles into the desired oligomer from the correctly folded RNA that has a larger size and mitigated structural flexibility, while the species of misfolded or alternative conformations assembled into other assemblies of undesired oligomeric states that can be readily eliminated by native purification methods such as native PAGE. Therefore, besides offering RNA constructs more amenable to high-resolution cryo-EM reconstruction, the self-assembled system also facilitates the experimental procedures of RNA folding optimization and native RNA purification. This helps eliminate the misfolding and conformational heterogeneity that are well known to complicate the functional and structural studies of RNA. c, The lengths of the connector helices between the RNA core and kissing-loop motif need to be optimized for ring closure. The computer model with optimized lengths of connector helices is shown in the middle, and the removal (left) or addition (right) of one base-pair (bp) would lead to the formation of open spiral structures. The screening process can be performed by NanoTiler as described in Methods.

Extended Data Fig. 2 Designing the 5′ and 3′ fragments for the pre-2SΔ5′ex TetGI complex.

a, Sequences of the RNA mimicking the 5′ fragment of the TetGI (5′ mimic) as the reverse transcription (RT) control and the 56-FAM-labelled RT primer. Underlined nucleotides in 5′ mimic were mutated to facilitate the synthesis by IVT; the nucleotides marked by grey line above are the primer-binding region. b, Analysis of the 5′ sequences of the TetGI by RT. Lanes 1 to 5 are the dideoxynucleotide sequencing results of 5′ mimic with a single dideoxynucleoside triphosphate (ddNTP; indicated under the gel) added (lanes 1 to 4) or no ddNTP added (lane 5); lanes 6 and 7 are the results of the constructs with wild-type 5′ sequence (same as the post-2S complex in Fig. 2c) or mutated 5′ sequence (same as the pre-2SΔ5′ex complex in Fig. 2d). While the RNA used in the post-2S complex is cleaved after U20, the RNA in the pre-2SΔ5′ex complex is not cleaved at this site. c, The likely formation of hairpin structures (two possibilities shown) of the wild-type 5′ sequence accounting for the self-cleavage. We note that the reaction of the wild-type leads to a L-19 product that is cleaved after U19 (ref. 87) and we attribute the difference of cleavage products to the different 5′ sequences. d, Two RNA constructs transcribed with a self-cleaving ribozyme (rbz; VS or HDV) for producing a homogeneous 3′-end. The G14-a+9 RNA was originally designed to have 3′-exon, which was almost completely cleaved in the absence of the 5′-exon. Green and red scissors mark the cleavage sites for the appended ribozyme and the group I intron itself, respectively. e, Structures of the 3′-ends produced by the cleavage of the appended ribozyme or the self-cleavage of the intron. While the latter can be extended by yeast poly(A) polymerase (PAP), the former cannot. f, PAP extension assay demonstrates that the majority products of both G14-A386 (lanes 1 and 2) and G14-a+9 (lanes 3 and 4) RNA preparation are extended by PAP (lanes 2 and 4), indicating that they are mostly the products from the intron’s self-cleavage. g, h, Sequences of the 5′ IVT RNA (g; G14-A386, from nucleotides 14 to 386) and the 3′ fragments (h; dTetCIRC with two deoxynucleotides flanking the 3′ splice site, and rTetCIRC with all-RNA nucleotides) for the pre-2SΔ5′ex TetGI complex. i, TetGI-catalyzed hydrolysis at the 3′ splice site is inhibited by the modification of DNA bases flanking the scissile phosphate of the 3′ splice site. Black arrow points to the band of the 3′ fragments of (dTetCIRC or rTetCIRC); grey arrows point to the bands of the cleaved 3′ fragments. j, Assembly assays for the monomeric (lanes 1 to 3) and dimeric (lanes 4 to 6) constructs of 5′ IVT G14-A386 RNA without 3′ fragment (lanes 1 and 4), with dTetCIRC (lanes 2 and 5) or with rTetCIRC (lanes 3 and 6). The annealing buffer contains 3 mM Mg2+, which is chosen from the assembly assay of the post-2S constructs (TetGI-M, -D, and -T; see Fig. 3a). Lanes 7 and 8 are control assemblies of TetGI-M and -D. Without 3′ fragment (lanes 1 and 4), there are trailing smears for the 5′ IVT RNA, indicating incorrect folding of the intron if P9.2, P9a and P9.0 are not formed. In the presence of dTetCIRC (lanes 2 and 5) or rTetCIRC (lanes 3 and 7), especially for dTetCIRC, some slower-migrating bands emerged, probably due to domain swapping, that is a 3′ fragment simultaneously binds to a 5′ IVT RNA molecule to form P9.2 and another 5′ IVT RNA molecule to form P10.

Extended Data Fig. 3 Cryo-EM imaging, processing and validation for TetGI-DS.

a, Representative cryo-EM image of TetGI-DS. The scale bar represents 20 nm. b, 2D class averages of TetGI-DS. Box size is 264 Å. c, Processing flowchart for the TetGI-DS dataset. d, Angle distribution for the particles included in the final 3D reconstruction. e, Fourier Shell Correlation (FSC) curves of the final TetGI-DS reconstruction. Half map #1 vs. half map #2 for the entire monomer is shown in black. The remaining FSC curves were calculated for the core domains only: half map #1 vs. half map #2 (red), model vs. refined map (blue), model refined in half map #1 vs. half map #1 (green), and model refined in half map #1 vs. half map #2 (orange).

Extended Data Fig. 4 Cryo-EM imaging, processing and validation for TetGI-D.

a, Representative cryo-EM image of TetGI-D. The scale bar represents 20 nm. b, 2D class averages of TetGI-D. Box size is 264 Å. c, Processing flowchart for the TetGI-D dataset. 3D classification of the symmetry-expanded monomers results in classes according to the conformations with double-stranded P1 (green arrows) or with single-stranded IGS (red arrows). The ratio of the two conformations is calculated and shown. For the conformation with double-stranded P1, green and blue boxes show the details of tertiary contacts of P1 on the P4-P6 side and P3-P8 side, respectively. The contacts agree well with the previous biochemical studies48,49,50 and the crystal structure of AzoGI40. The final cryo-EM map for TetGI-D was refined from two classes, and exhibits a stronger map intensity of double-stranded P1 than single-stranded IGS; therefore, the atomic model was built with double-stranded P1. d, Angle distribution for the particles included in the final 3D reconstruction. e, Fourier Shell Correlation (FSC) curves of the final TetGI-D reconstruction. Half map #1 vs. half map #2 for the entire monomer is shown in black. The remaining FSC curves were calculated for the core domains only: half map #1 vs. half map #2 (red), model vs. refined map (blue), model refined in half map #1 vs. half map #1 (green), and model refined in half map #1 vs. half map #2 (orange).

Extended Data Fig. 5 Cryo-EM imaging, processing and validation for TetGI-T.

a, Representative cryo-EM image of TetGI-T. The scale bar represents 20 nm. b, 2D class averages of TetGI-T. Box size is 317 Å. c, Processing flowchart for the TetGI-T dataset. Similar to the case of TetGI-D, 3D classification of the symmetry-expanded monomers of TetGI-T also results in classes according to the conformations with double-stranded P1 (green arrows) or with single-stranded IGS (red arrows). The ratio of the two conformations is calculated and shown, which is close to the dimeric construct TetGI-D shown in Extended Data Fig. 4c. The final cryo-EM map for TetGI-T was refined from two classes, and exhibits a stronger map of single-stranded IGS than double-stranded P1, probably due to the stabilization of the peripheral domains by the trimer construct; therefore, the atomic model was built with the single-stranded IGS. d, Angle distribution for the particles included in the final 3D reconstruction. e, Fourier Shell Correlation (FSC) curves of the final TetGI-T reconstruction. Half map #1 vs. half map #2 for the entire monomer is shown in black. The remaining FSC curves were calculated for the core domains only: half map #1 vs. half map #2 (red), model vs. refined map (blue), model refined in half map #1 vs. half map #1 (green), and model refined in half map #1 vs. half map #2 (orange).

Extended Data Fig. 6 Newly visualized structural elements involving the peripheral domains of the TetGI.

a, Structural details of the tertiary interaction P13. P13 is a 6-bp duplex formed by the base-pairings of U75 through U80 (p13′) and A352 through A347 (p13′), and stacks coaxially between the G73:C81 pair of P2.1 and the G346:C353 pair of P9.1a, bearing a conformational resemblance to some other 6-bp kissing-loop complexes88,89. A 4-nt bulge consisting of A69 through A72 is present near the tip of P2.1, allowing for the bent shape at the junction of P2.1 and P13. b, Structure of the four-way junction (4WJ) at P9a-P9b and P9.1-P9.2. Top-right inset shows the strand directions of the continuous strands of the 4WJ, indicating the flanking helices stacked in a left-handed parallel configuration90. Bottom-right inset shows interaction details at the crossover site. Though the overall connectivity and other long-range tertiary interactions may be the major determinant for the configuration of this 4WJ, the sugar-phosphate interactions of the nucleotides from the exchanging strands at the crossover site may contribute to the configuration of this 4WJ. c, Structure of the complex multiway junction at P1, P3-P8 and P2-P2.1. Insets show the details of two tertiary contacts centered by A95 (top-right) and A97 (bottom-right), respectively, stabilizing the juxtaposition of the pseudo-continuous helices of P2-P2.1 and P3-P8. are two tertiary interactions between the P2-P2.1 and P3-P9 domains. Within the A97-centered tertiary interaction (right), U300 forms a base-triple with A97:U277 pair, corroborating the previous biochemical evidence91. d, A similar contact is observed in the crystal structure of the Twort group I intron51 (TwoGI; PDB code: 1y0q) formed between P7 and the internal loop in P7.2. e, f, Comparing the tertiary interactions observed in TetGI and TwoGI (overlayed by P7). g, Sequence alignment of different class IC1 and class IE group I introns reveals a conserved purine-rich loop (highlighted in yellow) at J9.1/9.1a. The extracts of the alignments of 14 sequences are from Lehnert et al. 35, where all the sequences were regarded as subgroup IC1, but later some of them were categorized as subgroup IE92. TtLSU (intron in the large ribosomal RNA precursor of Tetrahymena thermophila) is the TetGI studied in this work. We note that the internal loop of J9.1/9.1a of the TetGI, which has the sequence of 5′-AUGA-3′/5′-GGAG-3′, is reminiscent of but not exactly the loop E motif93, which is normally 5′-AGUA-3′/5′-GAA-3′, though the sequence of the abovementioned internal loop in P7.2 of the TwoGI is the same as loop E motif.

Extended Data Fig. 7 AzoGI-T sequence design, assembly and activity.

a, Two RNA designs, G4-G206 and G4-a+7, transcribed with a self-cleaving ribozyme (rbz) for producing a homogeneous 3′ end. Green and red scissors mark the cleavage site for the appended ribozyme or the intron itself. AzoLEM is the ligated exon mimic for the AzoGI, and its complexing with G4-G206 RNA forms the post-2S state of the intron. b, The PAP extension assay of various constructs for the AzoGI. The products from the preparation of G4-G206 RNA constructs cannot be extended by PAP (lanes 2 and 8), indicting a 2′,3′-cyclic phosphate generated by rbz at the 3′ end (this is different from the G14-G414 RNA of TetGI, the majority of which can be extended by PAP; see lane 2 of Extended Data Fig. 2f). For G4-a+7 constructs, the products generated by the cleavage of rbz and intron itself can be directly distinguished based on the different electrophoretic mobilities in the presented analytical gel because the former is longer due to the appendage of a 7-nt 3′ exon (the small size difference is not noticeable in the preparative gel for RNA purification). As shown in lanes 3 and 9, about half of G4-a+7 RNA of either construct is cleaved at the 3′ splice site (whereas, in the case of the TetGI G14-a+9 RNA, more than 90% is cleaved at the 3′ splice site, indicating a substantially higher activity of the TetGI; see lane 4 of Extended Data Fig. 2f). Only the shorter products, which is produced by intron cleavage, can be extended by PAP. After folding, the intron-cleaved shorter products increase to ~70% (lanes 5 and 11), and some portion of these shorter products cannot be extended by PAP, probably due to the lower accessibility of the 3′ end after RNA folding. c, Assembly assay of the trimeric construct AzoGI-T (lanes 2 to 5) and the monomer control AzoGI-M (lane 1). Similar to the TetGI constructs, the optimal condition for folding/assembly was determined to be 3 mM Mg2+. Interestingly, monomer control AzoGI-M runs into two major bands, indicating conformational heterogeneity that is likely due to the open and closed conformations of the post-2S construct. d, Activity assay to test the activity of the second step of splicing. The reactions were conducted with 1 µM monomer units of either intron construct and 2 µM substrate (S, reacting as the 5′ exon) at 25 °C. The assay takes the advantage of the fact that there is still about 30% of AzoGI*-M and AzoGI*-T constructs containing 5′ exon after IVT preparation, purification and folding due to presumably lower splicing activity of the AzoGI.

Extended Data Fig. 8 Cryo-EM imaging, processing and validation for AzoGI-T.

a, Representative cryo-EM image of AzoGI-T. The scale bar represents 20 nm. b, 2D class averages of AzoGI-T. Box size is 236 Å. c, Processing flowchart for the AzoGI-T dataset. d, Angle distribution for the particles included in the final 3D reconstruction. e, Gold-standard Fourier Shell Correlation (FSC) for the final AzoGI-T reconstruction.

Extended Data Fig. 9 FMNrbsw-T assembly and activity.

a, b, Assembly assay. FMNrbsw-M is the monomer control. Three factors to increase the trimer yield: the presence of FMN ligand; high Na+; and increasing RNA concentration. c, Native PAGE (6%, in the presence of 2 mM free Mg2+) analysis of preassembled FMNrbsw-T trimer (lane T) under different temperatures ranging from 23 °C to 58 °C. Disassembly of the trimer starts to occur at 45 °C. The high thermal stability of the kissing-loop motif may contribute to the formation of the kinetic assembly of dimer because the kissing-loop formation may precede the complete folding of the riboswitch during annealing as suggested by its weaker ligand binding as shown in e. Lane M contains the low-molecular weight DNA ladder (NEB). d, Derivation of the equation to analyze the ligand binding (1:1 stoichiometry) based on fluorescence quenching. Data were fitted using a two-parameter (Kd and fc) quadratic Equation (4). e, Fluorescent binding assay of FMN (60 nM) with the riboswitch constructs conducted in 100 mM KCl and 2 mM MgCl2. Each data point is represented as mean ± s.d from three independent measurements. FMNrbsw-T dimer is the alternate dimeric assembly of the FMNrbsw-T RNA. The calculated Kd (nM, mean ± s.d.), fc (mean ± s.d.), and R2 are: monomer control, 30.1 ± 2.2, 0.254 ± 0.006, 0.9961; FMNrbsw-T dimer, 91.6 ± 21.9, 0.273 ± 0.017, 0.9944; FMNrbsw-T trimer, 17.4 ± 7.6, 0.197 ± 0.020, 0.9881. The dashed blue line for is fitted FMNrbsw-T trimer taking [L]t as 20 nM, which assumes that each monomeric subunit in the trimer can locally sense only one third of the ligand, and calculated values of Kd (nM, mean ± s.d.), fc (mean ± s.d.), and R2 are: 28.8 ± 6.3, 0.188 ± 0.015, 0.9937. This adjustment results in an improved fitting as reflected by an improved R2.

Source data

Extended Data Fig. 10 Cryo-EM imaging, processing and validation for FMNrbsw-T.

a, Representative cryo-EM image of FMNrbsw-T. The scale bar represents 20 nm. b, 2D class averages of FMNrbsw-T. Box size is 236 Å. c, Processing flowchart for the FMNrbsw-T dataset d, Angle distribution for the particles included in the final 3D reconstruction. e, Gold-standard Fourier Shell Correlation (FSC) for the final FMNrbsw-T reconstruction.

Supplementary information

Supplementary Information

Supplementary Notes 1 and 2, Supplementary Figures 1–10, Supplementary Tables 1–3 and Supplementary References

Reporting Summary

Supplementary Video 1

Two conformations of the TetGI refined from the cryo-EM reconstruction of the construct TetGI-D

Supplementary Video 2

Two conformations of the AzoGI corresponding to the closed and open states after the second step of splicing

Supplementary Video 3

Different classes of the 3D classification of FMNrbsw-T particles showing the relatively larger structural motion of the peripheries

Source data

Source Data Extended Data Fig. 9

Statistical Source Data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, D., Thélot, F.A., Piccirilli, J.A. et al. Sub-3-Å cryo-EM structure of RNA enabled by engineered homomeric self-assembly. Nat Methods 19, 576–585 (2022). https://doi.org/10.1038/s41592-022-01455-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41592-022-01455-w

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