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:

Nascent peptide assists the ribosome in recognizing chemically distinct small molecules

Abstract

Regulation of gene expression in response to the changing environment is critical for cell survival. For instance, binding of macrolide antibiotics to the ribosome promotes translation arrest at the leader open reading frames ermCL and ermBL, which is necessary for inducing the antibiotic resistance genes ermC and ermB. Cladinose-containing macrolides such as erythromycin (ERY), but not ketolides such as telithromycin (TEL), arrest translation of ermCL, whereas either ERY or TEL stall ermBL translation. How the ribosome distinguishes between chemically similar small molecules is unknown. We show that single amino acid changes in the leader peptide switch the specificity of recognition of distinct molecules, triggering gene activation in response to ERY alone, to TEL alone or to both antibiotics or preventing stalling altogether. Thus, the ribosomal response to chemical signals can be modulated by minute changes in the nascent peptide, suggesting that protein sequences could have been optimized for rendering translation sensitive to environmental cues.

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

Figure 1: The C-terminal segment of the stalled ErmBL is critical for antibiotic cofactor specificity.
Figure 2: Arg7 in ErmBL modulates recognition of macrolide stalling cofactors.
Figure 3: The C-terminal amino acid of the stalled ErmBL directs differentiation between chemically distinct molecules and determines selectivity of antibiotic-mediated gene activation.
Figure 4: Mutations in the ErmCL peptide can change antibiotic specificity of ribosome stalling.

Similar content being viewed by others

References

  1. Weisblum, B. Erythromycin resistance by ribosome modification. Antimicrob. Agents Chemother. 39, 577–585 (1995).

    Article  CAS  Google Scholar 

  2. Ramu, H., Mankin, A. & Vazquez-Laslop, N. Programmed drug-dependent ribosome stalling. Mol. Microbiol. 71, 811–824 (2009).

    Article  CAS  Google Scholar 

  3. Vazquez-Laslop, N., Thum, C. & Mankin, A.S. Molecular mechanism of drug-dependent ribosome stalling. Mol. Cell 30, 190–202 (2008).

    Article  CAS  Google Scholar 

  4. 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 

  5. Schlünzen, F. et al. Structural basis for the interaction of antibiotics with the peptidyl transferase centre in eubacteria. Nature 413, 814–821 (2001).

    Article  Google Scholar 

  6. Tu, D., Blaha, G., Moore, P.B. & Steitz, T.A. Structures of MLSBK antibiotics bound to mutated large ribosomal subunits provide a structural explanation for resistance. Cell 121, 257–270 (2005).

    Article  CAS  Google Scholar 

  7. Bulkley, D., Innis, C.A., Blaha, G. & Steitz, T.A. Revisiting the structures of several antibiotics bound to the bacterial ribosome. Proc. Natl. Acad. Sci. USA 107, 17158–17163 (2010).

    Article  CAS  Google Scholar 

  8. Dunkle, J.A., Xiong, L., Mankin, A.S. & Cate, J.H. Structures of the Escherichia coli ribosome with antibiotics bound near the peptidyl transferase center explain spectra of drug action. Proc. Natl. Acad. Sci. USA 107, 17152–17157 (2010).

    Article  CAS  Google Scholar 

  9. Tenson, T., Lovmar, M. & Ehrenberg, M. The mechanism of action of macrolides, lincosamides and streptogramin B reveals the nascent peptide exit path in the ribosome. J. Mol. Biol. 330, 1005–1014 (2003).

    Article  CAS  Google Scholar 

  10. Kannan, K., Vázquez-Laslop, N. & Mankin, A.S. Selective protein synthesis by ribosomes with a drug-obstructed exit tunnel. Cell 151, 508–520 (2012).

    Article  CAS  Google Scholar 

  11. Davis, A.R., Gohara, D.W. & Yap, M.N. Sequence selectivity of macrolide-induced translational attenuation. Proc. Natl. Acad. Sci. USA 111, 15379–15384 (2014).

    Article  CAS  Google Scholar 

  12. Kannan, K. et al. The general mode of translation inhibition by macrolide antibiotics. Proc. Natl. Acad. Sci. USA 111, 15958–15963 (2014).

    Article  CAS  Google Scholar 

  13. Weisblum, B. Insights into erythromycin action from studies of its activity as inducer of resistance. Antimicrob. Agents Chemother. 39, 797–805 (1995).

    Article  CAS  Google Scholar 

  14. Gupta, P., Sothiselvam, S., Vázquez-Laslop, N. & Mankin, A.S. Deregulation of translation due to post-transcriptional modification of rRNA explains why erm genes are inducible. Nat. Commun. 4, 1984 (2013).

    Article  Google Scholar 

  15. Horinouchi, S. & Weisblum, B. Posttranscriptional modification of mRNA conformation: mechanism that regulates erythromycin-induced resistance. Proc. Natl. Acad. Sci. USA 77, 7079–7083 (1980).

    Article  CAS  Google Scholar 

  16. Gryczan, T.J., Grandi, G., Hahn, J., Grandi, R. & Dubnau, D. Conformational alteration of mRNA structure and the posttranscriptional regulation of erythromycin-induced drug resistance. Nucleic Acids Res. 8, 6081–6097 (1980).

    Article  CAS  Google Scholar 

  17. 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 

  18. Mayford, M. & Weisblum, B. The ermC leader peptide: amino acid alterations leading to differential efficiency of induction by macrolide-lincosamide-streptogramin B antibiotics. J. Bacteriol. 172, 3772–3779 (1990).

    Article  CAS  Google Scholar 

  19. Kamimiya, S. & Weisblum, B. Induction of ermSV by 16-membered-ring macrolide antibiotics. Antimicrob. Agents Chemother. 41, 530–534 (1997).

    Article  CAS  Google Scholar 

  20. Vázquez-Laslop, N. et al. Role of antibiotic ligand in nascent peptide-dependent ribosome stalling. Proc. Natl. Acad. Sci. USA 108, 10496–10501 (2011).

    Article  Google Scholar 

  21. Xiong, L., Shah, S., Mauvais, P. & Mankin, A.S. A ketolide resistance mutation in domain II of 23S rRNA reveals the proximity of hairpin 35 to the peptidyl transferase centre. Mol. Microbiol. 31, 633–639 (1999).

    Article  CAS  Google Scholar 

  22. Hansen, L.H., Mauvais, P. & Douthwaite, S. The macrolide-ketolide antibiotic binding site is formed by structures in domains II and V of 23S ribosomal RNA. Mol. Microbiol. 31, 623–631 (1999).

    Article  CAS  Google Scholar 

  23. Llano-Sotelo, B. et al. Binding and action of CEM-101, a new fluoroketolide antibiotic that inhibits protein synthesis. Antimicrob. Agents Chemother. 54, 4961–4970 (2010).

    Article  CAS  Google Scholar 

  24. Subramanian, S.L., Ramu, H. & Mankin, A.S. in Antibiotic Drug Discovery and Development (eds. Dougherty, T.J. & Pucci, M.J.) 455–484 (Springer, 2011).

  25. Bailey, M., Chettiath, T. & Mankin, A.S. Induction of erm(C) expression by noninducing antibiotics. Antimicrob. Agents Chemother. 52, 866–874 (2008).

    Article  CAS  Google Scholar 

  26. Ramu, H. et al. Nascent peptide in the ribosome exit tunnel affects functional properties of the A-site of the peptidyl transferase center. Mol. Cell 41, 321–330 (2011).

    Article  CAS  Google Scholar 

  27. 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 

  28. Voorhees, R.M., Weixlbaumer, A., Loakes, D., Kelley, A.C. & Ramakrishnan, V. Insights into substrate stabilization from snapshots of the peptidyl transferase center of the intact 70S ribosome. Nat. Struct. Mol. Biol. 16, 528–533 (2009).

    Article  CAS  Google Scholar 

  29. Zaher, H.S., Shaw, J.J., Strobel, S.A. & Green, R. The 2′-OH group of the peptidyl-tRNA stabilizes an active conformation of the ribosomal PTC. EMBO J. 30, 2445–2453 (2011).

    Article  CAS  Google Scholar 

  30. Englander, M.T. et al. The ribosome can discriminate the chirality of amino acids within its peptidyl-transferase center. Proc. Natl. Acad. Sci. USA 112, 6038–6043 (2015).

    Article  CAS  Google Scholar 

  31. Bonnefoy, A., Girard, A.M., Agouridas, C. & Chantot, J.F. Ketolides lack inducibility properties of MLS(B) resistance phenotype. J. Antimicrob. Chemother. 40, 85–90 (1997).

    Article  CAS  Google Scholar 

  32. Sutcliffe, J. & Leclercq, R. in Macrolide Antibiotics (eds. W. Schönfeld & H.A. Kirst) 281–318 (Birkhäuser, 2002).

  33. Oh, T.G., Kwon, A.R. & Choi, E.C. Induction of ermAMR from a clinical strain of Enterococcus faecalis by 16-membered-ring macrolide antibiotics. J. Bacteriol. 180, 5788–5791 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Gupta, P., Kannan, K., Mankin, A.S. & Vázquez-Laslop, N. Regulation of gene expression by macrolide-induced ribosomal frameshifting. Mol. Cell 52, 629–642 (2013).

    Article  CAS  Google Scholar 

  35. Vázquez-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  Google Scholar 

  36. Johnston, T.C., Thompson, R.B. & Baldwin, T.O. Nucleotide sequence of the luxB gene of Vibrio harveyi and the complete amino acid sequence of the beta subunit of bacterial luciferase. J. Biol. Chem. 261, 4805–4811 (1986).

    CAS  PubMed  Google Scholar 

  37. Trabuco, L.G., Villa, E., Mitra, K., Frank, J. & Schulten, K. Flexible fitting of atomic structures into electron microscopy maps using molecular dynamics. Structure 16, 673–683 (2008).

    Article  CAS  Google Scholar 

  38. Trabuco, L.G., Villa, E., Schreiner, E., Harrison, C.B. & Schulten, K. Molecular dynamics flexible fitting: a practical guide to combine cryo-electron microscopy and X-ray crystallography. Methods 49, 174–180 (2009).

    Article  CAS  Google Scholar 

  39. Trabuco, L.G. et al. The role of L1 stalk-tRNA interaction in the ribosome elongation cycle. J. Mol. Biol. 402, 741–760 (2010).

    Article  CAS  Google Scholar 

  40. Jühling, F. et al. tRNAdb 2009: compilation of tRNA sequences and tRNA genes. Nucleic Acids Res. 37, D159–D162 (2009).

    Article  Google Scholar 

  41. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph. 14, 33–38, 27–28 (1996).

    Article  CAS  Google Scholar 

  42. Phillips, J.C. et al. Scalable molecular dynamics with NAMD. J. Comput. Chem. 26, 1781–1802 (2005).

    Article  CAS  Google Scholar 

  43. Cornell, W.D. et al. A second generation force field for the simulation of proteins, nucleic acids, and organic molecules. J. Am. Chem. Soc. 117, 5179–5197 (1995).

    Article  CAS  Google Scholar 

  44. Aduri, R. et al. AMBER force field parameters for the naturally occurring modified nucleosides in RNA. J. Chem. Theory Comput. 3, 1464–1475 (2007).

    Article  CAS  Google Scholar 

  45. Darden, Y., York, D. & Pedersen, L. Particle mesh Ewald: an N·log(N) method for Ewald sums in large systems. J. Chem. Phys. 98, 10089–10092 (1993).

    Article  CAS  Google Scholar 

  46. Essmann, U. et al. A smooth particle mesh ewald method. J. Chem. Phys. 103, 8577–8593 (1995).

    Article  CAS  Google Scholar 

  47. Daura, X. et al. Peptide folding: when simulation meets experiment. Angew. Chem. Int. Ed. 38, 236–240 (1999).

    Article  CAS  Google Scholar 

  48. Pronk, S. et al. GROMACS 4.5: a high-throughput and highly parallel open source molecular simulation toolkit. Bioinformatics 29, 845–854 (2013).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank D. Wilson and S. Arenz (Gene Center, Munich) for helpful discussions, R. Bovee for help with some experiments and P. Fernandes (Cempra Pharmaceuticals) for providing telithromycin. This work was supported by a grant from the National Science Foundation (MCB 1244455) to A.S.M. and N.V.-L. and 9P41GM104601 (to K.S.) and the National Science Foundation (PHY0822613) (to K.S.). MD modeling was facilitated by a grant from the Great Lakes Consortium for Petascale Computation on the Blue Waters Computer, financed by the National Science Foundation (OCI 07-25070) and the Oak Ridge Leadership Computing Facility at Oak Ridge National Laboratory, which is supported by the Office of Science of the Department of Energy under Contract DE-AC05-00OR22725.

Author information

Authors and Affiliations

Authors

Contributions

P.G., A.S.M. and N.V.-L. designed research; P.G., B.L., D.K. and V.G. performed research; P.G., B.L., K.S., A.S.M. and N.V.-L. analyzed data; P.G., N.V.-L. and A.S.M. wrote the paper.

Corresponding authors

Correspondence to Alexander S Mankin or Nora Vázquez-Laslop.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–9 and Supplementary Table 1. (PDF 6737 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gupta, P., Liu, B., Klepacki, D. et al. Nascent peptide assists the ribosome in recognizing chemically distinct small molecules. Nat Chem Biol 12, 153–158 (2016). https://doi.org/10.1038/nchembio.1998

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nchembio.1998

This article is cited by

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing: Microbiology