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:

Synthetic group A streptogramin antibiotics that overcome Vat resistance

Abstract

Natural products serve as chemical blueprints for most antibiotics in clinical use. The evolutionary process by which these molecules arise is inherently accompanied by the co-evolution of resistance mechanisms that shorten the clinical lifetime of any given class of antibiotics1. Virginiamycin acetyltransferase (Vat) enzymes are resistance proteins that provide protection against streptogramins2, potent antibiotics against Gram-positive bacteria that inhibit the bacterial ribosome3. Owing to the challenge of selectively modifying the chemically complex, 23-membered macrocyclic scaffold of group A streptogramins, analogues that overcome the resistance conferred by Vat enzymes have not been previously developed2. Here we report the design, synthesis, and antibacterial evaluation of group A streptogramin antibiotics with extensive structural variability. Using cryo-electron microscopy and forcefield-based refinement, we characterize the binding of eight analogues to the bacterial ribosome at high resolution, revealing binding interactions that extend into the peptidyl tRNA-binding site and towards synergistic binders that occupy the nascent peptide exit tunnel. One of these analogues has excellent activity against several streptogramin-resistant strains of Staphylococcus aureus, exhibits decreased rates of acetylation in vitro, and is effective at lowering bacterial load in a mouse model of infection. Our results demonstrate that the combination of rational design and modular chemical synthesis can revitalize classes of antibiotics that are limited by naturally arising resistance mechanisms.

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: Modular synthesis enables access to more than 60 fully synthetic group A streptogramins.
Fig. 2: Antibiotic activity and in vivo efficacy of selected group A streptogramins.
Fig. 3: In vitro acetylation, VatA binding, and ribosome binding of highly active analogues.

Similar content being viewed by others

Data availability

Models and maps generated during this study are available in the EMDB and PDB (accessions are listed in Extended Data Tables 4 and 5). Source data are provided with this paper.

Code availability

Forcefield-based refinement is available in PHENIX (versions 1.15 and later) using beta features available in Schrӧdinger 2019-3. Python code for analysing IVT data and VatA kinetics data are available on github: https://github.com/fraser-lab/streptogramin.

References

  1. Wright, P. M., Seiple, I. B. & Myers, A. G. The evolving role of chemical synthesis in antibacterial drug discovery. Angew. Chem. Int. Edn Engl. 53, 8840–8869 (2014).

    Article  CAS  Google Scholar 

  2. Stogios, P. J. et al. Potential for reduction of streptogramin A resistance revealed by structural analysis of acetyltransferase VatA. Antimicrob. Agents Chemother. 58, 7083–7092 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Vazquez, D. in Mechanism of Action (eds Gottlieb, D. & Shaw, P. D.) 387–403 (Springer Berlin Heidelberg, 1967).

  4. Waglechner, N. & Wright, G. D. Antibiotic resistance: it’s bad, but why isn’t it worse? BMC Biol. 15, 84 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  5. Seiple, I. B. et al. A platform for the discovery of new macrolide antibiotics. Nature 533, 338–345 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  6. Charest, M. G., Lerner, C. D., Brubaker, J. D., Siegel, D. R. & Myers, A. G. A convergent enantioselective route to structurally diverse 6-deoxytetracycline antibiotics. Science 308, 395–398 (2005).

    Article  ADS  CAS  PubMed  Google Scholar 

  7. Vidaillac, C., Parra-Ruiz, J., Winterfield, P. & Rybak, M. J. In vitro pharmacokinetic/pharmacodynamic activity of NXL103 versus clindamycin and linezolid against clinical Staphylococcus aureus and Streptococcus pyogenes isolates. Int. J. Antimicrob. Agents 38, 301–306 (2011).

    Article  CAS  PubMed  Google Scholar 

  8. Wilson, D. N. The A-Z of bacterial translation inhibitors. Crit. Rev. Biochem. Mol. Biol. 44, 393–433 (2009).

    Article  CAS  PubMed  Google Scholar 

  9. Noeske, J. et al. Synergy of streptogramin antibiotics occurs independently of their effects on translation. Antimicrob. Agents Chemother. 58, 5269–5279 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Hershberger, E., Donabedian, S., Konstantinou, K. & Zervos, M. J. Quinupristin-dalfopristin resistance in Gram-positive bacteria: mechanism of resistance and epidemiology. Clin. Infect. Dis. 38, 92–98 (2004).

    Article  CAS  PubMed  Google Scholar 

  11. Sharkey, L. K. R. & O’Neill, A. J. Antibiotic resistance ABC-F proteins: bringing target protection into the limelight. ACS Infect. Dis. 4, 239–246 (2018).

    Article  CAS  PubMed  Google Scholar 

  12. Leclercq, R. & Courvalin, P. Bacterial resistance to macrolide, lincosamide, and streptogramin antibiotics by target modification. Antimicrob. Agents Chemother. 35, 1267–1272 (1991).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Haroche, J. et al. Clonal diversity among streptogramin A-resistant Staphylococcus aureus isolates collected in French hospitals. J. Clin. Microbiol. 41, 586–591 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Werner, G., Cuny, C., Schmitz, F. J. & Witte, W. Methicillin-resistant, quinupristin-dalfopristin-resistant Staphylococcus aureus with reduced sensitivity to glycopeptides. J. Clin. Microbiol. 39, 3586–3590 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Valour, F. et al. Pristinamycin in the treatment of MSSA bone and joint infection. J. Antimicrob. Chemother. 71, 1063–1070 (2016).

    Article  CAS  PubMed  Google Scholar 

  16. Delgado, G. Jr, Neuhauser, M. M., Bearden, D. T. & Danziger, L. H. Quinupristin-dalfopristin: an overview. Pharmacotherapy 20, 1469–1485 (2000).

    Article  CAS  PubMed  Google Scholar 

  17. Politano, A. D. & Sawyer, R. G. NXL-103, a combination of flopristin and linopristin, for the potential treatment of bacterial infections including community-acquired pneumonia and MRSA. Curr. Opin. Investig. Drugs 11, 225–236 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Li, Q. & Seiple, I. B. Modular, scalable synthesis of group A Streptogramin antibiotics. J. Am. Chem. Soc. 139, 13304–13307 (2017).

    Article  CAS  PubMed  Google Scholar 

  19. Li, Q. & Seiple, I. B. A concise route to virginiamycin M2. Tetrahedron 75, 3309–3318 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Schlessinger, R. H. & Li, Y.-J. Total synthesis of (−)-virginiamycin M2 using second-generation vinylogous urethane chemistry. J. Am. Chem. Soc. 118, 3301–3302 (1996).

    Article  CAS  Google Scholar 

  21. Entwistle, D. A., Jordan, S. I., Montgomery, J. & Pattenden, G. Total synthesis of the virginiamycin antibiotic 14,15-anhydropristinamycin IIB. J. Chem. Soc. Perkin Trans. I 1315–1317 (1996).

  22. Tavares, F., Lawson, J. P. & Meyers, A. I. Total synthesis of streptogramin antibiotics. (−)-Madumycin II. J. Am. Chem. Soc. 118, 3303–3304 (1996).

    Article  CAS  Google Scholar 

  23. Ghosh, A. K. & Liu, W. A convergent, enantioselective total synthesis of streptogramin antibiotic (−)-madumycin II. J. Org. Chem. 62, 7908–7909 (1997).

    Article  CAS  PubMed  Google Scholar 

  24. Breuilles, P. & Uguen, D. Total synthesis of pristinamycin IIB. Tetrahedr. Lett. 39, 3149–3152 (1998).

    CAS  Google Scholar 

  25. Entwistle, D. A. Total synthesis of oxazole-based virginiamycin antibiotics: 14,15-anhydropristinamycin IIB. Synthesis 1998, 603–612 (1998).

    Article  Google Scholar 

  26. Dvorak, C. A. et al. The synthesis of streptogramin antibiotics: (−)-griseoviridin and its C-8 epimer. Angew. Chem. Int. Ed. Engl. 39, 1664–1666 (2000).

    Article  CAS  PubMed  Google Scholar 

  27. Wu, J. & Panek, J. S. Total synthesis of (−)-virginiamycin M2. Angew. Chem. Int. Edn Engl. 49, 6165–6168 (2010).

    Article  CAS  Google Scholar 

  28. Wu, J. & Panek, J. S. Total synthesis of (−)-virginiamycin M2: application of crotylsilanes accessed by enantioselective Rh(II) or Cu(I) promoted carbenoid Si–H insertion. J. Org. Chem. 76, 9900–9918 (2011).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  Google Scholar 

  30. Li, J. et al. The VSGB 2.0 model: a next generation energy model for high resolution protein structure modeling. Proteins 79, 2794–2812 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Harms, J. M., Schlünzen, F., Fucini, P., Bartels, H. & Yonath, A. Alterations at the peptidyl transferase centre of the ribosome induced by the synergistic action of the streptogramins dalfopristin and quinupristin. BMC Biol. 2, 4 (2004).

    Article  PubMed  PubMed Central  Google Scholar 

  32. Osterman, I. A. et al. Madumycin II inhibits peptide bond formation by forcing the peptidyl transferase center into an inactive state. Nucleic Acids Res. 45, 7507–7514 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Hansen, J. L., Moore, P. B. & Steitz, T. A. Structures of five antibiotics bound at the peptidyl transferase center of the large ribosomal subunit. J. Mol. Biol. 330, 1061–1075 (2003).

    Article  CAS  PubMed  Google Scholar 

  34. 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  PubMed  Google Scholar 

  35. Hoang, N. H. et al. Regio-selectively reduced streptogramin A analogue, 5,6-dihydrovirginiamycin M1 exhibits improved potency against MRSA. Lett. Appl. Microbiol. 57, 393–398 (2013).

    Article  CAS  PubMed  Google Scholar 

  36. Kingston, D. G. I., Kolpak, M. X., LeFevre, J. W. & Borup-Grochtmann, I. Biosynthesis of antibiotics of the virginiamycin family. 3. Biosynthesis of virginiamycin M1. J. Am. Chem. Soc. 105, 5106–5110 (1983).

    Article  CAS  Google Scholar 

  37. Richter, M. F. et al. Predictive compound accumulation rules yield a broad-spectrum antibiotic. Nature 545, 299–304 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  38. Sharkey, L. K. R., Edwards, T. A. & O’Neill, A. J. ABC-F proteins mediate antibiotic resistance through ribosomal protection. MBio 7, e01975 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Radika, K. & Northrop, D. B. Correlation of antibiotic resistance with Vmax/Km ratio of enzymatic modification of aminoglycosides by kanamycin acetyltransferase. Antimicrob. Agents Chemother. 25, 479–482 (1984).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Knies, J. L., Cai, F. & Weinreich, D. M. Enzyme efficiency but not thermostability drives cefotaxime resistance evolution in TEM-1 β-lactamase. Mol. Biol. Evol. 34, 1040–1054 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. 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  PubMed  PubMed Central  Google Scholar 

  42. Renaud, J.-P. et al. Cryo-EM in drug discovery: achievements, limitations and prospects. Nat. Rev. Drug Discov. 17, 471–492 (2018).

    Article  CAS  PubMed  Google Scholar 

  43. Wong, W. et al. Mefloquine targets the Plasmodium falciparum 80S ribosome to inhibit protein synthesis. Nat. Microbiol. 2, 17031 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  44. 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  PubMed  PubMed Central  Google Scholar 

  45. Tropea, J. E., Cherry, S. & Waugh, D. S. Expression and purification of soluble His6-tagged TEV protease. Methods Mol. Biol. 498, 297–307 (2009).

    Article  CAS  PubMed  Google Scholar 

  46. Kuhn, M. L., Majorek, K. A., Minor, W. & Anderson, W. F. Broad-substrate screen as a tool to identify substrates for bacterial Gcn5-related N-acetyltransferases with unknown substrate specificity. Protein Sci. 22, 222–230 (2013).

    Article  CAS  PubMed  Google Scholar 

  47. Winter, G. xia2: an expert system for macromolecular crystallography data reduction. J. Appl. Crystallogr. 43, 186–190 (2010).

    Article  CAS  Google Scholar 

  48. Kabsch, W. XDS. Acta Crystallogr. D 66, 125–132 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Liebschner, D. et al. Macromolecular structure determination using X-rays, neutrons and electrons: recent developments in Phenix. Acta Crystallogr. D 75, 861–877 (2019).

    Article  CAS  Google Scholar 

  50. Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D 66, 486–501 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Schuwirth, B. S. et al. Structures of the bacterial ribosome at 3.5 A resolution. Science 310, 827–834 (2005).

    Article  ADS  CAS  PubMed  Google Scholar 

  52. Passmore, L. A. & Russo, C. J. Specimen preparation for high-resolution cryo-EM. Methods Enzymol. 579, 51–86 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Mastronarde, D. N. Automated electron microscope tomography using robust prediction of specimen movements. J. Struct. Biol. 152, 36–51 (2005).

    Article  PubMed  Google Scholar 

  54. Suloway, C. et al. Automated molecular microscopy: the new Leginon system. J. Struct. Biol. 151, 41–60 (2005).

    Article  CAS  PubMed  Google Scholar 

  55. Cheng, A. et al. High resolution single particle cryo-electron microscopy using beam-image shift. J. Struct. Biol. 204, 270–275 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. 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 

  57. Grant, T., Rohou, A. & Grigorieff, N. cisTEM, user-friendly software for single-particle image processing. eLife 7, e35383 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  58. 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 

  59. Noeske, J. et al. High-resolution structure of the Escherichia coli ribosome. Nat. Struct. Mol. Biol. 22, 336–341 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Moriarty, N. W., Grosse-Kunstleve, R. W. & Adams, P. D. electronic Ligand Builder and Optimization Workbench (eLBOW): a tool for ligand coordinate and restraint generation. Acta Crystallogr. D 65, 1074–1080 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Roos, K. et al. OPLS3e: extending force field coverage for drug-like small molecules. J. Chem. Theory Comput. 15, 1863–1874 (2019).

    Article  CAS  PubMed  Google Scholar 

  62. Sindhikara, D. et al. Improving accuracy, diversity, and speed with prime macrocycle conformational sampling. J. Chem. Inf. Model. 57, 1881–1894 (2017).

    Article  CAS  PubMed  Google Scholar 

  63. Bochevarov, A. D. et al. Jaguar: A high-performance quantum chemistry software program with strengths in life and materials sciences. Int. J. Quantum Chem. 113, 2110–2142 (2013).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank F. Ward and J. Cate for initial advice on ribosome purifications and translation assays, E.Nogales and a UCSF-UCB Sackler Sabbatical Exchange Fellowship (J.S.F.) for initial cryo-EM access and training. A.A.T. and J.P. were supported by the National Science Foundation Graduate Research Fellowship Program under Grant No. 1650113. D.J.L. was supported by a Postdoctoral Individual National Research Award NIH AI148120. H.A.C. was supported by a National Institute on Minority Health and Health Disparities (NIMHD) research diversity supplement under NIH GM123159. This project was funded by the UCSF Program for Breakthrough Biomedical Research, funded in part by the Sandler Foundation (J.S.F. and I.B.S.), a Sangvhi-Agarwal Innovation Award (J.S.F.), Packard Fellowships from the David and Lucile Packard Foundation (J.S.F. and I.B.S.), NIH GM123159 (J.S.F.), and NIH GM128656 (I.B.S.). We thank G. Meigs and J. Holton at Beamline 8.3.1 at the Advanced Light Source, which is operated by the University of California Office of the President, Multicampus Research Programs and Initiatives grant MR-15-328599, the National Institutes of Health (R01 GM124149 and P30 GM124169), Plexxikon Inc., and the Integrated Diffraction Analysis Technologies program of the US Department of Energy Office of Biological and Environmental Research. The Advanced Light Source (Berkeley, CA) is a national user facility operated by Lawrence Berkeley National Laboratory on behalf of the US Department of Energy under contract number DE-AC02-05CH11231, Office of Basic Energy Sciences. We thank M. Thompson for comments on the crystallography methods. We thank A. Myasnikov and D. Bulkley for technical support at the UCSF Center for Advanced CryoEM, which is supported by NIH grants S10OD020054 and S10OD021741 and the Howard Hughes Medical Institute (HHMI). We thank E. Eng and E. Kopylov for technical support at the National Center for CryoEM Access and Training (NCCAT) and the Simons Electron Microscopy Center located at the New York Structural Biology Center, which is supported by the NIH Common Fund Transformative High Resolution Cryo-Electron Microscopy program (U24 GM129539) and by grants from the Simons Foundation (SF349247) and NY State. We thank W. Weiss at the University of North Texas Health Science Center for conducting the animal study.

Author information

Authors and Affiliations

Authors

Contributions

Q.L. and I.B.S. determined analogues for synthesis and designed the synthetic routes; Q.L. executed and optimized the syntheses of analogues, with assistance from A.A.T. (analogues 2932), R.W. (analogue 21), K.J. (analogues 27 and 28), and D.C. (analogue 26); J.P. and D.J.L. prepared samples and collected cryo-EM data; J.P. and A.F.B. calculated cryo-EM reconstructions; J.P. and J.E.P. performed the VatA acetylation assay; J.P. performed the in vitro translation experiments; G.v.Z. and K.B. developed new tools for cryo-EM model refinement; J.P., K.B. and J.T.B. performed cryo-EM model refinements; G.v.Z., J.P., H.A.C., N.Z. and M.P.J. determined relative energies of macrocycle confirmations; H.A.C. collected X-ray crystallographic data and performed X-ray model refinements; D.S., C.W., B.M., E.M, and O.C. designed and executed the MIC assays; Q.L., J.P., I.B.S. and J.S.F. wrote the manuscript. All authors discussed the results and commented on the manuscript.

Corresponding author

Correspondence to Ian B. Seiple.

Ethics declarations

Competing interests

K.B. and G.v.Z. are employees of Schrodinger Inc. D.S., C.W. and B.M. are employees of Micromyx.

Additional information

Peer review information Nature thanks Martin Burke, Gerard Wright and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data figures and tables

Extended Data Fig. 1 Natural and semisynthetic streptogramins and their molecular mechanisms of action and resistance.

a, Selected natural and semisynthetic streptogramin analogues. Modifications installed by semisynthesis are highlighted in blue. b, 2.5-Å cryo-EM structure of VM2 bound to the 50S subunit of the E. coli ribosome. Coulomb potential density is contoured in dark blue at 4.0σ and light grey at 1.0σ. Atom colouring of VM2 mirrors the building blocks used in its synthesis (see Fig. 2). c, Binding interactions between VM2 and residues in the ribosomal binding site. d, X-ray crystal structure VM1 bound to the resistance protein VatA (PDB ID: 4HUS). e, Binding interactions between VM1 and VatA, highlighting the extensive hydrophobic interactions at C3–C6. Acetylation occurs at the C14 alcohol. f, g, Conformational energy of VM2 showing contributions on a per atom basis when refined with standard CIF-based restraints generated by ‘phenix.eLBOW’ (f) and when refined with OPLS3e/VSGB2.1 force field (g). Colour indicates low strain (green, −14 kcal mol−1) up to high strain (red, 10 kcal mol−1), with total conformational energy of 39.5 kcal mol−1 (f) and −88.3 (g). Hydrogens were added and optimized with fixed heavy atoms for the CIF-based refined conformation using ‘prepwizard’; the PHENIX-OPLS3e/VSGB2.1 refined conformation was taken as is. Energies were calculated using Prime and per atom contribution visualized using Maestro’s prime energy visualization.

Extended Data Fig. 2 List of streptogramins tested for inhibitory activity.

Fully synthetic group A streptogramins tested for inhibitory activity against 20 strains of bacteria (see Extended Data Figs. 3, 4).

Extended Data Fig. 3 Inhibitory activity against Gram-positive organisms.

MIC values for selected analogues against an expanded panel of Gram-positive pathogens.

Source data

Extended Data Fig. 4 Inhibitory activity against Gram-negative organisms.

MIC values for selected analogues against an expanded panel of Gram-negative pathogens.

Source data

Extended Data Fig. 5 Cryo-EM density for all compounds bound to the E. coli ribosome.

a, 2.6-Å cryo-EM structure of VM2 bound to the 50S subunit of the E. coli ribosome. Coulomb potential density is contoured in dark blue at 4.0σ and light grey at 1.0σ for the entire figure. b, 2.8-Å cryo-EM structure of 21 bound to the 50S subunit of the E. coli ribosome. c, 2.8-Å cryo-EM structure of 40e bound to the 50S subunit of the E. coli ribosome. d, 2.5-Å cryo-EM structure of 40o bound to the 50S subunit of the E. coli ribosome. e, 2.8-Å cryo-EM structure of 40q bound to the 50S subunit of the E. coli ribosome. f, 2.6-Å cryo-EM structure of 41q bound to the 50S subunit of the E. coli ribosome. g, 2.5-Å cryo-EM structure of 46 bound to the 50S subunit of the E. coli ribosome. h, 2.5-Å cryo-EM structure of 47 bound to the 50S subunit of the E. coli ribosome. i, 2.7-Å cryo-EM structure of 46/VS1 bound to the 50S subunit of the E. coli ribosome. j, 2.8-Å cryo-EM structure of 47/VS1 bound to the 50S subunit of the E. coli ribosome.

Extended Data Fig. 6 Gold standard and map to model Fourier shell correlation plots.

aj, The particle Fourier shell correlation (FSC) curves for reconstructions obtained by cisTEM using a molecular mass of 1.8 MDa are shown in blue with unmasked map–model FSC curves obtained from ‘phenix.mtriage’ shown in orange. Dashed lines indicate FSC of 0.143 for estimating gold standard resolution and FSC of 0.5 for estimating map–model resolution.

Extended Data Fig. 7 Conformations of 46 and 47 in the ribosome and in VatA.

a, The conformation of 46 minimized by quantum mechanical methods in low dielectric, shows how the isoquinoline side chain packs over the macrocycle. b, By contrast, the ribosome-bound conformations of 46 determined by cryo-EM show that the side chain extends away from the macrocycle due to interactions formed in the binding site. c, Model of 47 in the conformation bound to the ribosome modelled into the active site of VatA (shown in surface). d, Model of 46 in the conformation bound to the ribosome modelled into the active site of VatA. e, Low energy model of 46 modelled into the active site of VatA. f, Overlay of VatA-bound (marine), ribosome-bound (violet), and ribosome with VS1-bound (light pink) conformations of 47. g, X-ray crystal structures of VM1 bound to VatA (PDB code 4HUS; 2.4 Å) and 46 bound to VatA at 2.8-Å resolution.

Extended Data Table 1 Ligand energies by different refinement schemes
Extended Data Table 2 Comparative energies of ligands bound to VatA and to the ribosome
Extended Data Table 3 Statistical analyses of mouse thigh in vivo data, MIC assays and VatA kinetics data
Extended Data Table 4 X-Ray data collection, processing, and model refinement statistics
Extended Data Table 5 Cryo-EM data collection, processing, and model refinement statistics

Supplementary information

Supplementary Information

Procedures for MIC determination and chemical synthesis.

Reporting Summary

Supplementary Data

Excel file containing MIC data featured in the manuscript and extended data.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, Q., Pellegrino, J., Lee, D.J. et al. Synthetic group A streptogramin antibiotics that overcome Vat resistance. Nature 586, 145–150 (2020). https://doi.org/10.1038/s41586-020-2761-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-2761-3

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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