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:

Insights into the ribosome function from the structures of non-arrested ribosome–nascent chain complexes

Abstract

During protein synthesis, the growing polypeptide threads through the ribosomal exit tunnel and modulates ribosomal activity by itself or by sensing various small molecules, such as metabolites or antibiotics, appearing in the tunnel. While arrested ribosome–nascent chain complexes (RNCCs) have been extensively studied structurally, the lack of a simple procedure for the large-scale preparation of peptidyl-tRNAs, intermediates in polypeptide synthesis that carry the growing chain, means that little attention has been given to RNCCs representing functionally active states of the ribosome. Here we report the facile synthesis of stably linked peptidyl-tRNAs through a chemoenzymatic approach based on native chemical ligation and use them to determine several structures of RNCCs in the functional pre-attack state of the peptidyl transferase centre. These structures reveal that C-terminal parts of the growing peptides adopt the same uniform β-strand conformation stabilized by an intricate network of hydrogen bonds with the universally conserved 23S rRNA nucleotides, and explain how the ribosome synthesizes growing peptides containing various sequences with comparable efficiencies.

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: Synthesis of peptidyl-tRNA by NCL.
Fig. 2: Synthesis of full-length non-hydrolysable peptidyl-tRNAs by NCL.
Fig. 3: Structure of the 70S ribosome in complex with fMSEAC-peptidyl-tRNA.
Fig. 4: Tight coordination of the PTC-proximal part of the nascent peptide in the NPET.
Fig. 5: Structure of the 70S ribosome in complex with fMRC- and fMTHSMRC-peptidyl-tRNAs.

Similar content being viewed by others

Data Availability

Coordinates and structure factors were deposited in the RCSB Protein Data Bank with the following accession codes: 8CVJ for the T. thermophilus 70S ribosome in complex with mRNA, aminoacylated A-site Phe-NH-tRNAPhe, peptidyl P-site fMSEAC-NH-tRNAiMet and deacylated E-site tRNAPhe; 8CVK for the T. thermophilus 70S ribosome in complex with mRNA, aminoacylated A-site Phe-NH-tRNAPhe, peptidyl P-site fMRC-NH-tRNAiMet and deacylated E-site tRNAPhe; 8CVL for the T. thermophilus 70S ribosome in complex with mRNA, aminoacylated A-site Phe-NH-tRNAPhe, peptidyl P-site fMTHSMRC-NH-tRNAiMet and deacylated E-site tRNAPhe. All previously published structures that were used in this work for model building and structural comparisons were retrieved from the RCSB Protein Data Bank: PDB entries 6XHW, 6WDD, 6TC3, 5NWY and 5DGV. Source data are provided with this paper.

References

  1. Bhushan, S. et al. α-Helical nascent polypeptide chains visualized within distinct regions of the ribosomal exit tunnel. Nat. Struct. Mol. Biol. 17, 313–317 (2010).

    Article  CAS  Google Scholar 

  2. Ito, K. & Chiba, S. Arrest peptides: cis-acting modulators of translation. Annu. Rev. Biochem. 82, 171–202 (2013).

    Article  CAS  Google Scholar 

  3. Vazquez-Laslop, N. & Mankin, A. S. How macrolide antibiotics work. Trends Biochem. Sci 43, 668–684 (2018).

    Article  CAS  Google Scholar 

  4. Wilson, D. N. Ribosome-targeting antibiotics and mechanisms of bacterial resistance. Nat. Rev. Microbiol. 12, 35–48 (2014).

    Article  CAS  Google Scholar 

  5. Lin, J., Zhou, D., Steitz, T. A., Polikanov, Y. S. & Gagnon, M. G. Ribosome-targeting antibiotics: modes of action, mechanisms of resistance, and implications for drug design. Annu. Rev. Biochem. 87, 451–478 (2018).

  6. Seip, B. & Innis, C. A. How widespread is metabolite sensing by ribosome-arresting nascent peptides? J. Mol. Biol. 428, 2217–2227 (2016).

    Article  CAS  Google Scholar 

  7. Doerfel, L. K. et al. EF-P is essential for rapid synthesis of proteins containing consecutive proline residues. Science 339, 85–88 (2013).

    Article  CAS  Google Scholar 

  8. Arenz, S. et al. A combined cryo-EM and molecular dynamics approach reveals the mechanism of ErmBL-mediated translation arrest. Nat. Commun. 7, 12026 (2016).

    Article  CAS  Google Scholar 

  9. Arenz, S. et al. Drug sensing by the ribosome induces translational arrest via active site perturbation. Mol. Cell 56, 446–452 (2014).

    Article  CAS  Google Scholar 

  10. Arenz, S. et al. Molecular basis for erythromycin-dependent ribosome stalling during translation of the ErmBL leader peptide. Nat. Commun. 5, 3501 (2014).

    Article  Google Scholar 

  11. Beckert, B. et al. Structural and mechanistic basis for translation inhibition by macrolide and ketolide antibiotics. Nat. Commun. 12, 4466 (2021).

    Article  CAS  Google Scholar 

  12. Herrero Del Valle, A. et al. Ornithine capture by a translating ribosome controls bacterial polyamine synthesis. Nat. Microbiol. 5, 554–561 (2020).

    Article  CAS  Google Scholar 

  13. Su, T. et al. The force-sensing peptide VemP employs extreme compaction and secondary structure formation to induce ribosomal stalling. Elife 6, e25642 (2017).

    Article  Google Scholar 

  14. Schaffitzel, C. & Ban, N. Generation of ribosome nascent chain complexes for structural and functional studies. J. Struct. Biol. 158, 463–471 (2007).

    Article  CAS  Google Scholar 

  15. Rutkowska, A. et al. Large-scale purification of ribosome-nascent chain complexes for biochemical and structural studies. FEBS Lett. 583, 2407–2413 (2009).

    Article  CAS  Google Scholar 

  16. Jha, S. S. & Komar, A. A. Using SecM arrest sequence as a tool to isolate ribosome bound polypeptides. J. Vis. Exp. 19, 4027 (2012).

  17. Cabrita, L. D. et al. A structural ensemble of a ribosome–nascent chain complex during cotranslational protein folding. Nat. Struct. Mol. Biol. 23, 278–285 (2016).

    Article  CAS  Google Scholar 

  18. Cassaignau, A. M. et al. A strategy for co-translational folding studies of ribosome-bound nascent chain complexes using NMR spectroscopy. Nat. Protoc. 11, 1492–1507 (2016).

    Article  Google Scholar 

  19. Loveland, A. B., Demo, G. & Korostelev, A. A. Cryo-EM of elongating ribosome with EF-Tu*GTP elucidates tRNA proofreading. Nature 584, 640–645 (2020).

    Article  CAS  Google Scholar 

  20. Polikanov, Y. S., Steitz, T. A. & Innis, C. A. A proton wire to couple aminoacyl-tRNA accommodation and peptide-bond formation on the ribosome. Nat. Struct. Mol. Biol. 21, 787–793 (2014).

    Article  CAS  Google Scholar 

  21. Goto, Y. & Suga, H. Translation initiation with initiator tRNA charged with exotic peptides. J. Am. Chem. Soc. 131, 5040–5041 (2009).

    Article  CAS  Google Scholar 

  22. Fraser, T. H. & Rich, A. Synthesis and aminoacylation of 3′-amino-3′-deoxy transfer RNA and its activity in ribosomal protein synthesis. Proc. Natl Acad. Sci. USA 70, 2671–2675 (1973).

    Article  CAS  Google Scholar 

  23. Graber, D. et al. Reliable semi-synthesis of hydrolysis-resistant 3′-peptidyl-tRNA conjugates containing genuine tRNA modifications. Nucleic Acids Res. 38, 6796–6802 (2010).

    Article  CAS  Google Scholar 

  24. Dawson, P. E., Muir, T. W., Clark-Lewis, I. & Kent, S. B. Synthesis of proteins by native chemical ligation. Science 266, 776–779 (1994).

    Article  CAS  Google Scholar 

  25. Geiermann, A. S., Polacek, N. & Micura, R. Native chemical ligation of hydrolysis-resistant 3′-peptidyl-tRNA mimics. J. Am. Chem. Soc. 133, 19068–19071 (2011).

    Article  CAS  Google Scholar 

  26. Katoh, T. & Suga, H. Flexizyme-catalyzed synthesis of 3′-aminoacyl-NH-tRNAs. Nucleic Acids Res. 47, e54 (2019).

    Article  CAS  Google Scholar 

  27. Gamper, H. & Hou, Y. M. tRNA 3′-amino-tailing for stable amino acid attachment. RNA 24, 1878–1885 (2018).

    Article  CAS  Google Scholar 

  28. Svetlov, M. S. et al. Structure of Erm-modified 70S ribosome reveals the mechanism of macrolide resistance. Nat. Chem. Biol. 17, 412–420 (2021).

    Article  CAS  Google Scholar 

  29. Freist, W., Sternbach, H. & Cramer, F. Isoleucyl-tRNA synthetase from baker’s yeast and from Escherichia coli MRE 600. Discrimination of 20 amino acids in aminoacylation of tRNA(Ile)-C-C-A(3′NH2). Eur. J. Biochem. 169, 33–39 (1987).

    Article  CAS  Google Scholar 

  30. Schmitt, E., Blanquet, S. & Mechulam, Y. Crystallization and preliminary X-ray analysis of Escherichia coli methionyl-tRNAMet(f) formyltransferase complexed with formyl-methionyl-tRNAMet(f). Acta Crystallogr. D 55, 332–334 (1999).

    Article  CAS  Google Scholar 

  31. Vidales, F. J., Bernabeu, C. & Ballesta, J. P. Peptidyl transfer ribonucleic acid hydrolase activity of proteinase K. Biochemistry 18, 4155–4158 (1979).

    Article  CAS  Google Scholar 

  32. Sothiselvam, S. et al. Macrolide antibiotics allosterically predispose the ribosome for translation arrest. Proc. Natl Acad. Sci. USA 111, 9804–9809 (2014).

    Article  CAS  Google Scholar 

  33. Schmeing, T. M., Huang, K. S., Strobel, S. A. & Steitz, T. A. An induced-fit mechanism to promote peptide bond formation and exclude hydrolysis of peptidyl-tRNA. Nature 438, 520–524 (2005).

    Article  CAS  Google Scholar 

  34. Kuhlenkoetter, S., Wintermeyer, W. & Rodnina, M. V. Different substrate-dependent transition states in the active site of the ribosome. Nature 476, 351–354 (2011).

    Article  CAS  Google Scholar 

  35. Hiller, D. A., Singh, V., Zhong, M. & Strobel, S. A. A two-step chemical mechanism for ribosome-catalysed peptide bond formation. Nature 476, 236–239 (2011).

    Article  CAS  Google Scholar 

  36. Sievers, A., Beringer, M., Rodnina, M. V. & Wolfenden, R. The ribosome as an entropy trap. Proc. Natl Acad. Sci. USA 101, 7897–7901 (2004).

    Article  CAS  Google Scholar 

  37. Karimi, R., Pavlov, M. Y., Heurgue-Hamard, V., Buckingham, R. H. & Ehrenberg, M. Initiation factors IF1 and IF2 synergistically remove peptidyl-tRNAs with short polypeptides from the P-site of translating Escherichia coli ribosomes. J. Mol. Biol. 281, 241–252 (1998).

    Article  CAS  Google Scholar 

  38. Heurgue-Hamard, V. et al. Ribosome release factor RF4 and termination factor RF3 are involved in dissociation of peptidyl-tRNA from the ribosome. EMBO J. 17, 808–816 (1998).

    Article  CAS  Google Scholar 

  39. Youngman, E. M., Brunelle, J. L., Kochaniak, A. B. & Green, R. The active site of the ribosome is composed of two layers of conserved nucleotides with distinct roles in peptide bond formation and peptide release. Cell 117, 589–959 (2004).

    Article  CAS  Google Scholar 

  40. Sato, N. S., Hirabayashi, N., Agmon, I., Yonath, A. & Suzuki, T. Comprehensive genetic selection revealed essential bases in the peptidyl-transferase center. Proc. Natl Acad. Sci. USA 103, 15386–15391 (2006).

    Article  CAS  Google Scholar 

  41. d’Aquino, A. E. et al. Mutational characterization and mapping of the 70S ribosome active site. Nucleic Acids Res. 48, 2777–2789 (2020).

    Article  Google Scholar 

  42. Koch, M., Willi, J., Pradere, U., Hall, J. & Polacek, N. Critical 23S rRNA interactions for macrolide-dependent ribosome stalling on the ErmCL nascent peptide chain. Nucleic Acids Res. 45, 6717–6728 (2017).

    Article  CAS  Google Scholar 

  43. Vazquez-Laslop, N., Ramu, H., Klepacki, D., Kannan, K. & Mankin, A. S. The key function of a conserved and modified rRNA residue in the ribosomal response to the nascent peptide. EMBO J. 29, 3108–3117 (2010).

    Article  CAS  Google Scholar 

  44. Syroegin, E. A. et al. Structural basis for the context-specific action of the classic peptidyl transferase inhibitor chloramphenicol. Nat. Struct. Mol. Biol. 29, 152–161 (2022).

    Article  CAS  Google Scholar 

  45. Sothiselvam, S. et al. Binding of macrolide antibiotics leads to ribosomal selection against specific substrates based on their charge and size. Cell Rep. 16, 1789–1799 (2016).

    Article  CAS  Google Scholar 

  46. Lim, V. I. & Spirin, A. S. Stereochemical analysis of ribosomal transpeptidation. Conformation of nascent peptide. J. Mol. Biol. 188, 565–574 (1986).

    Article  CAS  Google Scholar 

  47. Katunin, V. I., Muth, G. W., Strobel, S. A., Wintermeyer, W. & Rodnina, M. V. Important contribution to catalysis of peptide bond formation by a single ionizing group within the ribosome. Mol. Cell 10, 339–346 (2002).

    Article  CAS  Google Scholar 

  48. Wohlgemuth, I., Brenner, S., Beringer, M. & Rodnina, M. V. Modulation of the rate of peptidyl transfer on the ribosome by the nature of substrates. J. Biol. Chem. 283, 32229–32235 (2008).

    Article  CAS  Google Scholar 

  49. Panet, A., de Groot, N. & Lapidot, Y. Substrate specificity of Escherichia coli peptidyl-transferase. Eur. J. Biochem. 15, 222–225 (1970).

    Article  CAS  Google Scholar 

  50. Woolstenhulme, C. J. et al. Nascent peptides that block protein synthesis in bacteria. Proc. Natl Acad. Sci. USA 110, E878–E887 (2013).

    Article  CAS  Google Scholar 

  51. Ude, S. et al. Translation elongation factor EF-P alleviates ribosome stalling at polyproline stretches. Science 339, 82–85 (2013).

    Article  CAS  Google Scholar 

  52. Doerfel, L. K. et al. Entropic contribution of elongation factor P to proline positioning at the catalytic center of the ribosome. J. Am. Chem. Soc. 137, 12997–13006 (2015).

    Article  CAS  Google Scholar 

  53. Melnikov, S. et al. Molecular insights into protein synthesis with proline residues. EMBO Rep. 17, 1776–1784 (2016).

    Article  CAS  Google Scholar 

  54. Bretscher, M. S. & Marcker, K. A. Polypeptidyl-sigma-ribonucleic acid and amino-acyl-sigma-ribonucleic acid binding sites on ribosomes. Nature 211, 380–384 (1966).

    Article  CAS  Google Scholar 

  55. Saini, P., Eyler, D. E., Green, R. & Dever, T. E. Hypusine-containing protein eIF5A promotes translation elongation. Nature 459, 118–121 (2009).

    Article  CAS  Google Scholar 

  56. Schuller, A. P., Wu, C. C., Dever, T. E., Buskirk, A. R. & Green, R. eIF5A functions globally in translation elongation and termination. Mol. Cell 66, 194–205 e195 (2017).

    Article  CAS  Google Scholar 

  57. Blaha, G., Stanley, R. E. & Steitz, T. A. Formation of the first peptide bond: the structure of EF-P bound to the 70S ribosome. Science 325, 966–970 (2009).

    Article  CAS  Google Scholar 

  58. Benne, R. & Hershey, J. W. The mechanism of action of protein synthesis initiation factors from rabbit reticulocytes. J. Biol. Chem. 253, 3078–3087 (1978).

    Article  CAS  Google Scholar 

  59. Kemper, W. M., Berry, K. W. & Merrick, W. C. Purification and properties of rabbit reticulocyte protein synthesis initiation factors M2Balpha and M2Bbeta. J. Biol. Chem. 251, 5551–5557 (1976).

    Article  CAS  Google Scholar 

  60. Schreier, M. H., Erni, B. & Staehelin, T. Initiation of mammalian protein synthesis. I. Purification and characterization of seven initiation factors. J. Mol. Biol. 116, 727–753 (1977).

    Article  CAS  Google Scholar 

  61. Menninger, J. R. Peptidyl transfer RNA dissociates during protein synthesis from ribosomes of Escherichia coli. J. Biol. Chem. 251, 3392–3398 (1976).

    Article  CAS  Google Scholar 

  62. Fujino, T., Goto, Y., Suga, H. & Murakami, H. Reevaluation of the d-amino acid compatibility with the elongation event in translation. J. Am. Chem. Soc. 135, 1830–1837 (2013).

    Article  CAS  Google Scholar 

  63. Tajima, K., Katoh, T. & Suga, H. Drop-off-reinitiation triggered by EF-G-driven mistranslocation and its alleviation by EF-P. Nucleic Acids Res. 50, 2736–2753 (2022).

    Article  CAS  Google Scholar 

  64. Mohapatra, S., Choi, H., Ge, X., Sanyal, S. & Weisshaar, J. C. Spatial distribution and ribosome-binding dynamics of EF-P in live Escherichia coli. mBio 8 (2017).

  65. Katoh, T., Iwane, Y. & Suga, H. Logical engineering of D-arm and T-stem of tRNA that enhances d-amino acid incorporation. Nucleic Acids Res. 45, 12601–12610 (2017).

    Article  CAS  Google Scholar 

  66. Hormoz, S. Amino acid composition of proteins reduces deleterious impact of mutations. Sci Rep. 3, 2919 (2013).

    Article  Google Scholar 

  67. Kulkarni, S. S., Sayers, J., Premdjee, B. & Payne, R. J. Rapid and efficient protein synthesis through expansion of the native chemical ligation concept. Nat. Rev. Chem. 2, 0122 (2018).

    Article  CAS  Google Scholar 

  68. Wan, Q. & Danishefsky, S. J. Free-radical-based, specific desulfurization of cysteine: a powerful advance in the synthesis of polypeptides and glycopolypeptides. Angew. Chem. Int. Ed. Engl. 46, 9248–9252 (2007).

    Article  CAS  Google Scholar 

  69. Jin, K., Li, T., Chow, H. Y., Liu, H. & Li, X. P-B desulfurization: an enabling method for protein chemical synthesis and site-specific deuteration. Angew. Chem. Int. Ed. Engl. 56, 14607–14611 (2017).

    Article  CAS  Google Scholar 

  70. Degut, C., Monod, A., Brachet, F., Crepin, T. & Tisne, C. In vitro/in vivo production of tRNA for X-ray studies. Methods Mol. Biol. 1320, 37–57 (2016).

    Article  Google Scholar 

  71. Selmer, M. et al. Structure of the 70S ribosome complexed with mRNA and tRNA. Science 313, 1935–1942 (2006).

    Article  CAS  Google Scholar 

  72. Polikanov, Y. S., Blaha, G. M. & Steitz, T. A. How hibernation factors RMF, HPF, and YfiA turn off protein synthesis. Science 336, 915–918 (2012).

    Article  CAS  Google Scholar 

  73. Polikanov, Y. S., Melnikov, S. V., Soll, D. & Steitz, T. A. Structural insights into the role of rRNA modifications in protein synthesis and ribosome assembly. Nat. Struct. Mol. Biol. 22, 342–344 (2015).

    Article  CAS  Google Scholar 

  74. Pettersen, E. F. et al. UCSF ChimeraX: structure visualization for researchers, educators, and developers. Protein Sci. 30, 70–82 (2021).

    Article  CAS  Google Scholar 

  75. Vimberg, V., Xiong, L., Bailey, M., Tenson, T. & Mankin, A. Peptide-mediated macrolide resistance reveals possible specific interactions in the nascent peptide exit tunnel. Mol. Microbiol. 54, 376–385 (2004).

    Article  CAS  Google Scholar 

  76. Marks, J. et al. Context-specific inhibition of translation by ribosomal antibiotics targeting the peptidyl transferase center. Proc. Natl Acad. Sci. USA 113, 12150–12155 (2016).

    Article  CAS  Google Scholar 

  77. Florin, T. et al. An antimicrobial peptide that inhibits translation by trapping release factors on the ribosome. Nat. Struct. Mol. Biol. 24, 752–757 (2017).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank R. Micura, A. Mankin, N. Vazquez-Laslop and M. Svetlov for critical reading of the manuscript and valuable suggestions. We thank I. Kamyshko for providing expertise in reagent and material sourcing. We thank the staff at NE-CAT beamlines 24ID-C and 24ID-E for help with data collection and freezing of the crystals, especially M. Capel, F. Murphy, S. Banerjee, I. Kourinov, D. Neau, J. Schuermann, N. Sukumar, A. Lynch, J. Withrow, K. Perry, A. Kaya and C. Salbego. This work is based upon research conducted at the Northeastern Collaborative Access Team beamlines, which are funded by the National Institute of General Medical Sciences from the National Institutes of Health (P30-GM124165 to NE-CAT). The Eiger 16M detector on the 24-ID-E beamline is funded by an NIH-ORIP HEI grant (S10-OD021527 to NE-CAT). This research used resources of the Advanced Photon Source, a US Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Argonne National Laboratory under contract number DE-AC02-06CH11357. This work was supported by the National Institutes of Health (R01-GM132302 and R21-AI163466 to Y.S.P.), the National Science Foundation (MCB-1907273 to Y.S.P.) and Illinois State start-up funds (to Y.S.P.). The funders had no role in study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

E.A.S. with help from E.V.A. developed the NCL-based method for the synthesis of non-hydrolysable full-length peptidyl-tRNAs. E.A.S., E.V.A. and Y.S.P. designed and performed X-ray crystallography experiments. Y.S.P. supervised the experiments. All authors interpreted the results. E.A.S., E.V.A. and Y.S.P. wrote the manuscript.

Corresponding author

Correspondence to Yury S. Polikanov.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemistry thanks the anonymous reviewer(s) for their contribution to the peer review of this work.

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 Native chemical ligation of Cys-NH-tRNAiMet with TBZ-activated peptides of various lengths and sequences.

(a) General information about the model peptides chosen for this study: macrolide-sensing ErmDL stalling peptide11,32; macrolide-resistance peptide MFLRC75; chloramphenicol-sensitive protein H-NS76; proline-rich antimicrobial peptide apidaecin77. ERY, erythromycin or other macrolides; CHL, chloramphenicol; RF, release factor. Peptidyl-tRNA products used in the structural studies are highlighted in grey. (b) Electrophoretic separation of crude NCL reaction mixtures. Indicated TBZ-activated peptides and Cys-NH-tRNAiMet were used as N- and C-terminal reactants, respectively. Electrophoresis was performed in 20-cm long 8% PAAG with 7 M urea and stained with ethidium bromide. Note the slower mobility of tRNA after the NCL that is also proportional to the ligated peptide length (lanes 2–8 vs. 1).

Source data

Extended Data Fig. 2 Native chemical ligation of non-hydrolyzable or native Cys-tRNACys with TBZ-activated fMSEA-peptide.

(a, b) Electrophoretic separations of deacylated, aminoacyl- and peptidyl-tRNACys before and after NCL reactions. Amino-tailed 3’-NH2-tRNACys (a, lane 1) or native 3’-HO-tRNACys (b, lane 1) were first aminoacylated using cysteine-specific aminoacyl-tRNA-synthetase (lanes 2) and then ligated with activated fMSEA-TBZ peptide for various times (lanes 3–5 in a and 3–7 in b) to yield either non-hydrolyzable fMSEAC-NH-tRNACys (a) or native ester-linked fMSEAC-O-tRNACys (b) peptidyl-tRNAs.

Source data

Extended Data Fig. 3 Comparison of the structures of fMSEAC-peptidyl-tRNA with aminoacylated full-length tRNAs.

(a, b) Superpositioning of our 70S ribosome structure carrying Phe-NH-tRNAPhe (green) and fMSEAC-NH-tRNAiMet (blue) in the A and P sites, respectively, with the previously reported structures of ribosome-bound full-length aminoacyl-tRNAs featuring either non-hydrolyzable amide linkages (a, PDB entry 6XHW28) or native ester bonds (b, PDB entry 6WDD19) between the amino acid moieties and the ribose of nucleotide A76 of A- and P-site tRNAs. All structures were aligned based on domain V of the 23S rRNA. (c, d) Comparisons of the positions of key 23S rRNA nucleotides around the PTC in the same structures. Note that there are no significant differences in the positions of A- or P-site substrates or the PTC nucleotides indicating that the amide-linked aminoacyl and peptidyl-tRNAs represent functionally meaningful analogs of native ester-linked tRNAs.

Extended Data Fig. 4 Tightly coordinated water molecules in the pre-attack state of the peptidyl transferase center.

(a-c) Close-up views of the 2Fo-Fc electron difference Fourier map (black mesh) for water molecules W1, W2, and W3 (yellow; nomenclature from20) in the pre-peptidyl-transfer complex structures featuring Phe-NH-tRNAPhe in the A site (omitted for clarity) and fMSEAC-NH-tRNAiMet (a, yellow), fMRC-NH-tRNAiMet (b, orange), or fMTHSMRC-NH-tRNAiMet (c, crimson) in the P site. H-bonds are shown by black dotted lines.

Extended Data Fig. 5 Ramachandran plots for the peptide moieties of the ribosome-bound peptidyl-tRNAs.

(a-c) Diagrams of the phi vs. psi angles for amino acid residues in fMSEAC-NH-tRNAiMet (a, yellow), fMRC-NH-tRNAiMet (b, orange), or fMTHSMRC-NH-tRNAiMet (c, crimson) in the P site. Favored regions are shown in green; allowed regions are light green. Due to the absence of fMet1 residue in the structure of fMSEAC-peptidyl-tRNA, phi angle for the subsequent Ser2 residue cannot be determined.

Extended Data Fig. 6 Comparison of the structures of fMSEAC-peptidyl-tRNA with the structures of other ribosome-bound peptidyl-tRNAs.

Superpositioning of the 70S ribosome structure containing A-site Phe-NH-tRNAPhe (green) and P-site fMSEAC-NH-tRNAiMet (blue) with the previously reported structures of stalled RNCCs carrying full-length peptidyl-tRNAs (a-c) or non-stalled RNCCs carrying short non-hydrolyzable tripeptidyl-tRNA analogs (d-f). Individual panels show comparisons of the fMSEAC-tripeptidyl-tRNA with the following peptides: (a) SpeFL (dark blue, PDB entry 6TC312); (b) ErmDL (teal, PDB entry 7NSO11), (c) VemP (cyan, PDB entry 5NWY13); (d) MAI-tripeptide (yellow, PDB entry 7RQB44); (e) MTI-tripeptide (orange, PDB entry 7RQA44), (f) MFI-tripeptide (red, PDB entry 7RQC44). All structures were aligned based on domain V of the 23S rRNA. Note that the overall path of the fMSEAC-peptide in our structure is similar to the trajectories of the other peptides in the NPET.

Extended Data Fig. 7 Additional interactions of the side chains of fMRC- and fMTHSMRC-peptidyl-tRNAs with the ribosome.

(a, b) Close-up views of the electrostatic interactions between the side chain of the penultimate Arg residue of the P-site fMRC-NH-tRNAiMet (a, blue with peptide highlighted in orange) or fMTHSMRC-NH-tRNAiMet (b, navy with peptide highlighted in crimson) and the phosphate of nucleotide G2505 of the 23S rRNA. H-bonds are shown by black dotted lines. (c) Stacking interactions between the aromatic side chain of His3 of fMTHSMRC-peptidyl-tRNA and A2062 nucleobase of the 23S rRNA.

Extended Data Fig. 8 Proline residues in the nascent peptide are unable to form stabilizing H-bonds in the NPET.

(a-c) In silico modeling of proline residues at ultimate (a, Cys5Pro), penultimate (b, Ala4Pro), or pen-penultimate (c, Glu3Pro) positions of the fMSEAC-peptide chain. Note that besides its inability to form most of the peptide-stabilizing H-bonds, proline in the penultimate and pen-penultimate positions clashes with nucleotides A2062 and U2506 of the 23S rRNA, respectively. Geometrically possible and impossible H-bonds are shown by black and white dotted lines, respectively.

Extended Data Fig. 9 Proline residues alter the path of the nascent peptide in the NPET.

(a) In silico modeling of the two consecutive proline residues at ultimate (Cys5Pro) and penultimate (Ala4Pro) positions of the fMSEAC-peptide chain. Note that, due to the side chains, the diproline-containing peptide cannot adopt a conformation possible for other peptides in the NPET and must re-orient. (b) Comparison of the previous structure of ribosome-bound short diprolyl-tRNA analog (green, PDB entry 5DGV53) with the in silico-modeled diprolyl-containing tRNA based on the structure of fMSEAC-peptidyl-tRNA (red). Note that in order to avoid a steric clash with the A2062, the diprolyl moiety of the nascent peptide deviates to the side (black dashed arrows) and, thus, has an alternative path in the NPET.

Extended Data Fig. 10 Formylation of the first methionine residue provides additional stability to the initiator tRNA substrate in the P site.

Superpositioning of the previous structure of ribosome-bound initiator fMet-NH-tRNAiMet (navy with the fMet moiety in red, PDB entry 6XHW28) with the new structure of fMSEAC-peptidyl-tRNA (blue with the peptide moiety highlighted in yellow) viewed from two opposite sides (a, b). H-bonds are shown by black dotted lines. Note that the positions of carbon and oxygen atoms in the formyl group and those in the carbonyl group of the penultimate residue in the nascent peptide chain are nearly identical, ensuring formation of the same H-bond with the exocyclic amino group of the G2061 residue.

Supplementary information

Source data

Source Data Fig. 2

Unprocessed gels for Fig. 2b and Fig. 2c

Source Data Extended Data Fig. 1

Unprocessed gel for Extended Data Fig. 1b

Source Data Extended Data Fig. 2

Unprocessed gels for Extended Data Fig. 2a and Extended Data Fig. 2b

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Syroegin, E.A., Aleksandrova, E.V. & Polikanov, Y.S. Insights into the ribosome function from the structures of non-arrested ribosome–nascent chain complexes. Nat. Chem. 15, 143–153 (2023). https://doi.org/10.1038/s41557-022-01073-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41557-022-01073-1

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