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:

Ultra-large library docking for discovering new chemotypes

Abstract

Despite intense interest in expanding chemical space, libraries containing hundreds-of-millions to billions of diverse molecules have remained inaccessible. Here we investigate structure-based docking of 170 million make-on-demand compounds from 130 well-characterized reactions. The resulting library is diverse, representing over 10.7 million scaffolds that are otherwise unavailable. For each compound in the library, docking against AmpC β-lactamase (AmpC) and the D4 dopamine receptor were simulated. From the top-ranking molecules, 44 and 549 compounds were synthesized and tested for interactions with AmpC and the D4 dopamine receptor, respectively. We found a phenolate inhibitor of AmpC, which revealed a group of inhibitors without known precedent. This molecule was optimized to 77 nM, which places it among the most potent non-covalent AmpC inhibitors known. Crystal structures of this and other AmpC inhibitors confirmed the docking predictions. Against the D4 dopamine receptor, hit rates fell almost monotonically with docking score, and a hit-rate versus score curve predicted that the library contained 453,000 ligands for the D4 dopamine receptor. Of 81 new chemotypes discovered, 30 showed submicromolar activity, including a 180-pM subtype-selective agonist of the D4 dopamine receptor.

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: Make-on-demand compounds are diverse and have increased exponentially.
Fig. 2: Structural fidelity between docked-predicted and crystallographically determined poses of the new β-lactamase inhibitors.
Fig. 3: Testing 549 molecules at different docking ranks against the D4 receptor.
Fig. 4: Estimating the number of active D4 receptor ligands in the 138 million compound library.

Similar content being viewed by others

Data availability

Active molecules reported here are available from B.K.S. or directly from Enamine. The four structures of AmpC determined with the new docking hits are available from the PDB with accession numbers 6DPZ, 6DPY, 6DPX and 6DPT. The compounds docked in this study are freely available from our ZINC lead-like make-on-demand library (http://zinc15.docking.org). All active compounds are available either from the authors or may be purchased from Enamine. Figures with associated raw data include: Fig. 2, for which electron density and reflection files are deposited with the PDB; Figs. 3, 4 and Extended Data Fig. 5, for which Source Data are available in the online version of the paper; Extended Data Fig. 1, for which the data are included in Supplementary Table 1; Extended Data Fig. 6, for which raw clustering or no-clustering rank numbers are included in Supplementary Tables 8, 9. Further data are provided in Supplementary Tables 3, 5 (aggregation assays for AmpC inhibitors and D4 ligands); Extended Data Table 1 (crystallographic data collection and refinement); Supplementary Tables 9, 10 and Supplementary Data 1215 (chemical purity of active ligands and their spectra); Supplementary Data 11 and 14 (synthetic routes to compounds). All other data are available from the authors on request.

References

  1. Bohacek, R. S., McMartin, C. & Guida, W. C. The art and practice of structure-based drug design: a molecular modeling perspective. Med. Res. Rev. 16, 3–50 (1996).

    Article  CAS  Google Scholar 

  2. Ertl, P. Cheminformatics analysis of organic substituents: identification of the most common substituents, calculation of substituent properties, and automatic identification of drug-like bioisosteric groups. J. Chem. Inf. Comput. Sci. 43, 374–380 (2003).

    Article  CAS  Google Scholar 

  3. Fink, T., Bruggesser, H. & Reymond, J. L. Virtual exploration of the small-molecule chemical universe below 160 Daltons. Angew. Chem. Int. Ed. 44, 1504–1508 (2005).

    Article  CAS  Google Scholar 

  4. Chevillard, F. & Kolb, P. SCUBIDOO: a large yet screenable and easily searchable database of computationally created chemical compounds optimized toward high likelihood of synthetic tractability. J. Chem. Inf. Model. 55, 1824–1835 (2015).

    Article  CAS  Google Scholar 

  5. Keserü, G. M. & Makara, G. M. The influence of lead discovery strategies on the properties of drug candidates. Nat. Rev. Drug Discov. 8, 203–212 (2009).

    Article  Google Scholar 

  6. McGovern, S. L., Caselli, E., Grigorieff, N. & Shoichet, B. K. A common mechanism underlying promiscuous inhibitors from virtual and high-throughput screening. J. Med. Chem. 45, 1712–1722 (2002).

    Article  CAS  Google Scholar 

  7. Brenner, S. & Lerner, R. A. Encoded combinatorial chemistry. Proc. Natl Acad. Sci. USA 89, 5381–5383 (1992).

    Article  ADS  CAS  Google Scholar 

  8. Ahn, S. et al. Allosteric “beta-blocker” isolated from a DNA-encoded small molecule library. Proc. Natl Acad. Sci. USA 114, 1708–1713 (2017).

    Article  CAS  Google Scholar 

  9. Goodnow, R. A. Jr, Dumelin, C. E. & Keefe, A. D. DNA-encoded chemistry: enabling the deeper sampling of chemical space. Nat. Rev. Drug Discov. 16, 131–147 (2017).

    Article  CAS  Google Scholar 

  10. Jorgensen, W. L. The many roles of computation in drug discovery. Science 303, 1813–1818 (2004).

    Article  ADS  CAS  Google Scholar 

  11. de Graaf, C. et al. Crystal structure-based virtual screening for fragment-like ligands of the human histamine H1 receptor. J. Med. Chem. 54, 8195–8206 (2011).

    Article  ADS  Google Scholar 

  12. Katritch, V. et al. Structure-based discovery of novel chemotypes for adenosine A2A receptor antagonists. J. Med. Chem. 53, 1799–1809 (2010).

    Article  CAS  Google Scholar 

  13. Manglik, A. et al. Structure-based discovery of opioid analgesics with reduced side effects. Nature 537, 185–190 (2016).

    Article  ADS  CAS  Google Scholar 

  14. Wang, S. et al. D4 dopamine receptor high-resolution structures enable the discovery of selective agonists. Science 358, 381–386 (2017).

    Article  ADS  CAS  Google Scholar 

  15. Negri, A. et al. Discovery of a novel selective kappa-opioid receptor agonist using crystal structure-based virtual screening. J. Chem. Inf. Model. 53, 521–526 (2013).

    Article  CAS  Google Scholar 

  16. Jazayeri, A., Andrews, S. P. & Marshall, F. H. Structurally enabled discovery of adenosine A2A receptor antagonists. Chem. Rev. 117, 21–37 (2017).

    Article  CAS  Google Scholar 

  17. Lane, J. R. et al. Structure-based ligand discovery targeting orthosteric and allosteric pockets of dopamine receptors. Mol. Pharmacol. 84, 794–807 (2013).

    Article  CAS  Google Scholar 

  18. Langmead, C. J. et al. Identification of novel adenosine A2A receptor antagonists by virtual screening. J. Med. Chem. 55, 1904–1909 (2012).

    Article  CAS  Google Scholar 

  19. Becker, O. M. et al. G protein-coupled receptors: in silico drug discovery in 3D. Proc. Natl Acad. Sci. USA 101, 11304–11309 (2004).

    Article  ADS  CAS  Google Scholar 

  20. Kooistra, A. J. et al. Function-specific virtual screening for GPCR ligands using a combined scoring method. Sci. Rep. 6, 28288 (2016).

    Article  ADS  CAS  Google Scholar 

  21. Congreve, M. et al. Discovery of 1,2,4-triazine derivatives as adenosine A2A antagonists using structure based drug design. J. Med. Chem. 55, 1898–1903 (2012).

    Article  CAS  Google Scholar 

  22. Kiss, R. et al. Discovery of novel human histamine H4 receptor ligands by large-scale structure-based virtual screening. J. Med. Chem. 51, 3145–3153 (2008).

    Article  CAS  Google Scholar 

  23. Oprea, T. I. & Gottfries, J. Chemography: the art of navigating in chemical space. J. Comb. Chem. 3, 157–166 (2001).

    Article  CAS  Google Scholar 

  24. Mysinger, M. M., Carchia, M., Irwin, J. J. & Shoichet, B. K. Directory of useful decoys, enhanced (DUD-E): better ligands and decoys for better benchmarking. J. Med. Chem. 55, 6582–6594 (2012).

    Article  CAS  Google Scholar 

  25. Bemis, G. W. & Murcko, M. A. The properties of known drugs. 1. Molecular frameworks. J. Med. Chem. 39, 2887–2893 (1996).

    Article  CAS  Google Scholar 

  26. Gaulton, A. et al. The ChEMBL database in 2017. Nucleic Acids Res. 45, D945–D954 (2017).

    Article  CAS  Google Scholar 

  27. Katz, B. A. et al. A novel serine protease inhibition motif involving a multi-centered short hydrogen bonding network at the active site. J. Mol. Biol. 307, 1451–1486 (2001).

    Article  CAS  Google Scholar 

  28. Congreve, M., Langmead, C. J., Mason, J. S. & Marshall, F. H. Progress in structure based drug design for G protein-coupled receptors. J. Med. Chem. 54, 4283–4311 (2011).

    Article  CAS  Google Scholar 

  29. Vaidehi, N. Dynamics and flexibility of G-protein-coupled receptor conformations and their relevance to drug design. Drug Discov. Today 15, 951–957 (2010).

    Article  CAS  Google Scholar 

  30. Irwin, J. J. & Shoichet, B. K. Docking screens for novel ligands conferring new biology. J. Med. Chem. 59, 4103–4120 (2016).

    Article  CAS  Google Scholar 

  31. Vass, M. et al. in Computational Methods for GPCR Drug Discovery (Heifetz, A.) Ch. 4, 73–113 (Humana, Springer, New Jersey, 2018).

  32. Isberg, V. et al. Generic GPCR residue numbers – aligning topology maps while minding the gaps. Trends Pharmacol. Sci. 36, 22–31 (2015).

    Article  CAS  Google Scholar 

  33. McCorvy, J. D. et al. Structural determinants of 5-HT2B receptor activation and biased agonism. Nat. Struct. Mol. Biol. 25, 787–796 (2018).

    Article  CAS  Google Scholar 

  34. Powers, R. A., Morandi, F. & Shoichet, B. K. Structure-based discovery of a novel, noncovalent inhibitor of AmpC β-lactamase. Structure 10, 1013–1023 (2002).

    Article  CAS  Google Scholar 

  35. Feng, B. Y. et al. A high-throughput screen for aggregation-based inhibition in a large compound library. J. Med. Chem. 50, 2385–2390 (2007).

    Article  CAS  Google Scholar 

  36. Babaoglu, K. et al. Comprehensive mechanistic analysis of hits from high-throughput and docking screens against β-lactamase. J. Med. Chem. 51, 2502–2511 (2008).

    Article  CAS  Google Scholar 

  37. Rowley, M. et al. 5-(4-chlorophenyl)-4-methyl-3-(1-(2-phenylethyl)piperidin-4-yl)isoxazole: a potent, selective antagonist at human cloned dopamine D4 receptors. J. Med. Chem. 39, 1943–1945 (1996).

    Article  CAS  Google Scholar 

  38. Enguehard-Gueiffier, C. et al. 2-[(4-phenylpiperazin-1-yl)methyl]imidazo(di)azines as selective D4-ligands. Induction of penile erection by 2-[4-(2-methoxyphenyl)piperazin-1-ylmethyl]imidazo[1,2-a]pyridine (PIP3EA), a potent and selective D4 partial agonist. J. Med. Chem. 49, 3938–3947 (2006).

    Article  CAS  Google Scholar 

  39. Löber, S., Hübner, H. & Gmeiner, P. Synthesis and biological investigations of dopaminergic partial agonists preferentially recognizing the D4 receptor subtype. Bioorg. Med. Chem. Lett. 16, 2955–2959 (2006).

    Article  Google Scholar 

  40. Lindsley, C. W. & Hopkins, C. R. Return of D4 dopamine receptor antagonists in drug discovery. J. Med. Chem. 60, 7233–7243 (2017).

    Article  CAS  Google Scholar 

  41. Tirado-Rives, J. & Jorgensen, W. L. Contribution of conformer focusing to the uncertainty in predicting free energies for protein-ligand binding. J. Med. Chem. 49, 5880–5884 (2006).

    Article  CAS  Google Scholar 

  42. Abagyan, R., Totrov, M. & Kuznetsov, D. ICM—a new method for protein modeling and design: applications to docking and structure prediction from the distorted native conformation. J. Comput. Chem. 15, 488–506 (1994).

    Article  CAS  Google Scholar 

  43. Halgren, T. A. et al. Glide: a new approach for rapid, accurate docking and scoring. 2. Enrichment factors in database screening. J. Med. Chem. 47, 1750–1759 (2004).

    Article  CAS  Google Scholar 

  44. Goodsell, D. S. & Olson, A. J. Automated docking of substrates to proteins by simulated annealing. Proteins 8, 195–202 (1990).

    Article  CAS  Google Scholar 

  45. Kufareva, I., Katritch, V., Stevens, R. C. & Abagyan, R. Advances in GPCR modeling evaluated by the GPCR Dock 2013 assessment: meeting new challenges. Structure 22, 1120–1139 (2014).

    Article  CAS  Google Scholar 

  46. Kramer, B., Rarey, M. & Lengauer, T. Evaluation of the FLEXX incremental construction algorithm for protein-ligand docking. Proteins 37, 228–241 (1999).

    Article  CAS  Google Scholar 

  47. McGann, M. FRED pose prediction and virtual screening accuracy. J. Chem. Inf. Model. 51, 578–596 (2011).

    Article  CAS  Google Scholar 

  48. Jones, G., Willett, P., Glen, R. C., Leach, A. R. & Taylor, R. Development and validation of a genetic algorithm for flexible docking. J. Mol. Biol. 267, 727–748 (1997).

    Article  CAS  Google Scholar 

  49. Corbeil, C. R., Williams, C. I. & Labute, P. Variability in docking success rates due to dataset preparation. J. Comput. Aided Mol. Des. 26, 775–786 (2012).

    Article  ADS  CAS  Google Scholar 

  50. Hawkins, P. C., Skillman, A. G., Warren, G. L., Ellingson, B. A. & Stahl, M. T. Conformer generation with OMEGA: algorithm and validation using high quality structures from the Protein Databank and Cambridge Structural Database. J. Chem. Inf. Model. 50, 572–584 (2010).

    Article  CAS  Google Scholar 

  51. Hawkins, G. D. et al. AMSOL version 7.1 https://comp.chem.umn.edu/amsol/ (2004).

  52. Wei, B. Q., Baase, W. A., Weaver, L. H., Matthews, B. W. & Shoichet, B. K. A model binding site for testing scoring functions in molecular docking. J. Mol. Biol. 322, 339–355 (2002).

    Article  CAS  Google Scholar 

  53. Mysinger, M. M. & Shoichet, B. K. Rapid context-dependent ligand desolvation in molecular docking. J. Chem. Inf. Model. 50, 1561–1573 (2010).

    Article  CAS  Google Scholar 

  54. Sterling, T. & Irwin, J. J. ZINC 15—ligand discovery for everyone. J. Chem. Inf. Model. 55, 2324–2337 (2015).

    Article  CAS  Google Scholar 

  55. Barelier, S. et al. Increasing chemical space coverage by combining empirical and computational fragment screens. ACS Chem. Biol. 9, 1528–1535 (2014).

    Article  CAS  Google Scholar 

  56. Gray, D. L. et al. Impaired β-arrestin recruitment and reduced desensitization by non-catechol agonists of the D1 dopamine receptor. Nat. Commun. 9, 674 (2018).

    Article  ADS  Google Scholar 

  57. Carlsson, J. et al. Ligand discovery from a dopamine D3 receptor homology model and crystal structure. Nat. Chem. Biol. 7, 769–778 (2011).

    Article  CAS  Google Scholar 

  58. Meng, E. C., Shoichet, B. K. & Kuntz, I. D. Automated docking with gridb-based energy evaluation. J. Comput. Chem. 13, 505–524 (1992).

    Article  CAS  Google Scholar 

  59. Sharp, K. A., Friedman, R. A., Misra, V., Hecht, J. & Honig, B. Salt effects on polyelectrolyte-ligand binding: comparison of Poisson–Boltzmann, and limiting law/counterion binding models. Biopolymers 36, 245–262 (1995).

    Article  CAS  Google Scholar 

  60. Gallagher, K. & Sharp, K. Electrostatic contributions to heat capacity changes of DNA-ligand binding. Biophys. J. 75, 769–776 (1998).

    Article  ADS  CAS  Google Scholar 

  61. Coleman, R. G., Carchia, M., Sterling, T., Irwin, J. J. & Shoichet, B. K. Ligand pose and orientational sampling in molecular docking. PLoS ONE 8, e75992 (2013).

    Article  ADS  CAS  Google Scholar 

  62. Tolmachev, A. et al. Expanding synthesizable space of disubstituted 1,2,4-oxadiazoles. ACS Comb. Sci. 18, 616–624 (2016).

    Article  CAS  Google Scholar 

  63. Eidam, O. et al. Design, synthesis, crystal structures, and antimicrobial activity of sulfonamide boronic acids as β-lactamase inhibitors. J. Med. Chem. 53, 7852–7863 (2010).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  68. Eidam, O. et al. Fragment-guided design of subnanomolar β-lactamase inhibitors active in vivo. Proc. Natl Acad. Sci. USA 109, 17448–17453 (2012).

    Article  ADS  CAS  Google Scholar 

  69. Feng, B. Y. & Shoichet, B. K. A detergent-based assay for the detection of promiscuous inhibitors. Nat. Protoc. 1, 550–553 (2006).

    Article  CAS  Google Scholar 

  70. Allen, J. A. et al. Discovery of β-arrestin-biased dopamine D2 ligands for probing signal transduction pathways essential for antipsychotic efficacy. Proc. Natl Acad. Sci. USA 108, 18488–18493 (2011).

    Article  ADS  CAS  Google Scholar 

  71. Carpenter, B. et al. Stan: a probabilistic programming language. J. Stat. Softw. 1, 1–32 (2017).

    Google Scholar 

  72. Ryan, E. G., Drovandi, C. C., McGree, J. M. & Pettitt, A. N. A review of modern computational algorithms for Bayesian optimal design. Int. Stat. Rev. 84, 128–154 (2016).

    Article  MathSciNet  Google Scholar 

  73. Rainforth, T., Cornish, R., Yang, H., Warrington, A. & Wood, F. On Nesting Monte Carlo Estimators. In Proc. 35th International Conference on Machine Learning PMLR 80 (eds Dy, J. & Krause, A.) 4267–4276 (2018).

Download references

Acknowledgements

This research was supported by GM71896 (to J.J.I.); R35 GM122481 and a UCSF PBBR New Frontier Award (to B.K.S.); R01 MH112205, U24DK1169195 and the NIMH Psychoactive Drug Screening Contract (to B.L.R.); Strategic Priority Research Program of the Chinese Academy of Sciences, grant number XDB19000000 (to S.W.). We thank R. Stein and I. Fish for help with AmpC preparation, H. Torosyan for aggregation assays, R. H. J. Olsen for developing the D4 receptor BRET assay, B. Wong and C. Dandarchuluun for computer support, and M. Korczynska and J. Pottel for reading this manuscript; ChemAxon for a license to JChem, OpenEye Scientific software for a license to OEChem and Omega2, Molecular Networks for a license to Corina, and Molinspriation for a license to Mitools.

Reviewer information

Nature thanks M. M. Babu, D. E. Gloriam and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

J.J.I. and B.K.S. conceived the study. J.J.I. created the enlarged docking libraries. J.L., T.E.B. and A.L. performed the docking. S.W., T.C. and B.L.R. performed D4 receptor assays and analysis. I.S. performed all crystallography and assays for AmpC. Y.S.M., K.T. and A.A.T. directed the compound synthesis, purification and characterization. M.J.O. performed Bayesian modelling. E.A., T.E.B. and J.L. contributed enabling computer code. B.K.S., J.L., J.J.I., B.L.R., T.E.B., I.S., A.L., S.W., T.C. and M.J.O. wrote the paper.

Corresponding authors

Correspondence to Brian K. Shoichet, Bryan L. Roth or John J. Irwin.

Ethics declarations

Competing interests

B.K.S. and J.J.I. are founders of a company, BlueDolphin LLC, that works in the area of molecular docking. All other authors declare no competing interests.

Additional information

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 Simulating the effect of library size on ligand enrichment among the top 1,000 docked molecules.

a, b, The energy distribution of ligands (a) and decoys (b) from docking enrichment calculations against AmpC. The skewed normal fitting curves are plotted in red lines. The fitting parameters (shape (α), location (loc) and scale values) are shown. c, Heat maps of the number of active molecules in the top 1,000 docked molecules for 6 targets. The number of ligands in the top 1,000 docked molecules for a given library size and the ratio between ligands and decoys is coloured using a log10(number of ligands) scale ranging from 1 (blue) to 1,000 (red). Cells with zero ligands are shown in white. d, Large-library docking screens of AmpC (top, n = 99 million molecules) and D4 (bottom, n = 138 million molecules). Molecules that are known to bind to AmpC and D4, as well as close analogues, are treated as ligands and the rest of the molecules are treated as decoys. Left, the energy distributions of decoys (grey), ligands defined by ECFP4 Tc similarity ≥0.5 (blue), 0.6 (green) and 0.7 (orange) to ligands from ChEMBL. Middle, heat maps of the number of ligands in the top 1,000 docked molecules based on fit to full-library docking with the ligands (AmpC, Tc ≥ 0.5, green; D4, Tc ≥ 0.6, orange) and decoys (grey) distributions. Right, number of ligands in the top 1,000 docked molecules as the library grows based on actual distributions plotted in the left panel. Data are mean ± s.d. of 20 samples. See Supplementary Table 1 for retrospective performance on three more targets.

Extended Data Fig. 2 Initial hits and selected analogues against AmpC.

The five initial hits are shown in the left column. Under each compound, the first row includes the ZINC identifier; the second row is the cluster rank (position in cluster head list sorted by DOCK score) with global rank (position in unclustered hit list sorted by DOCK score) shown in brackets; the third row is the Tc value (Tanimoto coefficient to known AmpC inhibitors in ChEMBL); the fourth row is the Ki value. Five selected analogues for the corresponding hits are shown in the right column. Under each compound, the first row includes the ZINC identifier; the second row is the Tc value; and the third row is the Ki value.

Extended Data Fig. 3 Lineweaver–Burk plot and Ki analysis for analogues of each of the five series of AmpC inhibitors.

af, Lineweaver–Burk plots for ZINC776666294 (a), 275579920 (b), ZINC548592534 (c), ZINC1187516987 (d), 339204163 (e) and 549719643 (f), indicating competitive inhibition. IC50 values were determined by nonlinear regression fit in GraphPad Prism, and Ki values calculated by a replot of the slope of each Lineweaver–Burk plot versus the corresponding inhibitor concentration.

Extended Data Fig. 4 Electron density maps for AmpC–inhibitor complexes.

The initial Fo − Fc electron density map contoured at 2.5σ around the inhibitor (density in cyan) with refined 2Fo − Fc electron density contoured at 1σ for enzyme residues for the complexes with the following compounds. a, 547933290. b, 275579920. c, 339204163. d, 549719643. Inhibitor carbons are shown in cyan and enzyme carbons are shown in grey, oxygens in red, nitrogens in blue, sulfurs in yellow and chlorides in green.

Extended Data Fig. 5 Selected D4 hits from docking 138 million make-on-demand molecules.

Six ligands with docked poses (first column), cAMP Gαi/o activities (second column), Tango β-arrestin activities (third column) and 3H-N-methylspiperone displacement and chemical drawing (fourth column) are shown. The receptor structure is in grey and ligand carbons are in teal. Ballesteros–Weinstein residue numbers are included as superscripts. Functional assays represent normalized concentration–response curves of the ligands in cloned human D4-mediated activation of Gαi/o and β-arrestin translocation. Data are mean ± s.e.m. of three assays. The first row shows an example of an antagonist identified among the D4 hits. Both agonist (teal curve) and antagonist (purple curve) modes are shown for ZINC130532671 in the third panel; the concentration of quinpirole in the antagonist mode was 100 nM.

Source Data

Extended Data Fig. 6 Pre-clustering the docking library yields much worse scores of scaffold representatives compared to full library docking.

a, b, Comparison of energy distributions of scaffold representatives between full library docking (orange) and pre-clustered library docking for D4 (a) and AmpC (b) using four strategies: the closest member to the centroid of molecular masses and clogP (blue), the closest member to the centroid of molecular masses (pink), the member with the largest molecular masses (magenta) and the member with the smallest molecular masses (green). The inset shows the ratio of the number of molecules at a given docking score for full library docking divided by the number at that score when only cluster representatives are docked (coloured by clustering method). For each target, two examples illustrate the effect on our experimentally active scaffold families. c, D4. d, AmpC. The scaffold for each molecule is highlighted in red. The ZINC identifier, post-cluster rank and pre-cluster rank are labelled for each pair. The arrow colour is as for the pre-clustering methods in a and b.

Extended Data Fig. 7 Comparison of hit rates achieved by combined docking score and human prioritization compared to the rates achieved by docking score alone.

a, The hit rates for selecting compounds at different scoring ranges by each strategy: human prioritization and docking score (orange), or docking score alone (blue). Hit rate is the ratio of active compounds/tested compounds; the raw numbers appear at the top of each bar. b, Distribution of the binding affinity level among the hits from a. There are 32 hits from human prioritization and docking score, and 26 hits from the docking score alone. These are divided into three affinity ranges: <100 nM (pale blue); 100 nM–1 μM (blue); 1–10 μM (dark blue). c, Functional activity distribution among the hits from b. There are 22 molecules from human prioritization and docking score, and 7 molecules from the docking score alone. These are divided into five activity ranges: <10 nM (pale green); 10 nM–1 μM (light green); 1–10 μM (olive); 10–50 μM (forest green); and not determined (dark green).

Extended Data Fig. 8 Bayesian prior modelling for balancing information gain and ligand discovery in molecule-selection design and error estimation.

a, Sigmoid functional form for the hit-rate model. bd, Marginal Bayesian prior (teal) and posterior (red) distributions (n = 200,000) for each model parameter. b, Top. c, Dock50. d, Slope. e, Estimated hit rate based on evaluation by the authors of the docked poses before any molecules were tested. Brown, mean ± s.d.; n = 200, 220, 230, 230, 285, 235, 210, 230 and 200 compounds; n = 5, 4, 4, 4, 4, 4, 4, 4 and 4 experts. The prior mean (green) and samples (n = 200) from the prior (blue) are shown. f, Candidate (blue) and chosen (orange) experimental designs (inset, designs 1–6), with expected number of hits and information gain for each design. g, Expected number of active scaffolds (orange, mean; grey, posterior draws n = 200,000) superimposed on the total number of scaffold cluster heads (black). h, i, Marginal distribution of the number of active compounds (h) and scaffolds (i) over the posterior distributions (n = 200,000).

Extended Data Table 1 Data collection and refinement statistics of AmpC inhibitors
Extended Data Table 2 The highest-affinity direct docking hits for the D4 receptor

Supplementary information

Supplementary Information

This file contains Supplementary Tables and Data 1-15, except Supplementary Tables 2, 4, 7 and 8, which are provided as separate files.

Reporting Summary

Supplementary Table

This file contains Supplementary Table 2: Molecules tested against β-lactamase AmpC. We report the zinc id, smiles, and indicator of binding or non-binding. (supplied as a separate file. See Extended Data Fig. 2 for affinities and chemical drawings of potent binders).

Supplementary Table

This file contains Supplementary Table 4: All molecules tested against D4.

Supplementary Table

This file contains Supplementary Table 7: Full-library vs pre-clustering library docking for D4.

Supplementary Table

This file contains Supplementary Table 8: Full-library vs pre-clustering library docking for AmpC.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lyu, J., Wang, S., Balius, T.E. et al. Ultra-large library docking for discovering new chemotypes. Nature 566, 224–229 (2019). https://doi.org/10.1038/s41586-019-0917-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-0917-9

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