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:

Structural basis of odorant recognition by a human odorant receptor

Abstract

Our sense of smell enables us to navigate a vast space of chemically diverse odour molecules. This task is accomplished by the combinatorial activation of approximately 400 odorant G protein-coupled receptors encoded in the human genome1,2,3. How odorants are recognized by odorant receptors remains unclear. Here we provide mechanistic insight into how an odorant binds to a human odorant receptor. Using cryo-electron microscopy, we determined the structure of the active human odorant receptor OR51E2 bound to the fatty acid propionate. Propionate is bound within an occluded pocket in OR51E2 and makes specific contacts critical to receptor activation. Mutation of the odorant-binding pocket in OR51E2 alters the recognition spectrum for fatty acids of varying chain length, suggesting that odorant selectivity is controlled by tight packing interactions between an odorant and an odorant receptor. Molecular dynamics simulations demonstrate that propionate-induced conformational changes in extracellular loop 3 activate OR51E2. Together, our studies provide a high-resolution view of chemical recognition of an odorant by a vertebrate odorant receptor, providing insight into how this large family of G protein-coupled receptors enables our olfactory sense.

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: Structure of the human OR OR51E2.
Fig. 2: Odorant-binding pocket in OR51E2.
Fig. 3: Tuning OR51E2 odorant selectivity.
Fig. 4: Activation mechanism of OR51E2.
Fig. 5: Structural dynamics of ECL3 in OR function.

Similar content being viewed by others

Data availability

Coordinates for the propionate OR51E2–Gs complex have been deposited in the RCSB Protein Data Bank under accession code 8F76. EM density maps for OR51E2–Gs and the 7TM domain of OR51E2 have been deposited in the Electron Microscopy Data Bank under accession codes EMD-28896 and EMD-28900, respectively. The MD simulation trajectories for apo OR51E2, OR51E2 bound to propionate, and OR51E2–Q18145×53D mutant have been deposited in the GPCRmd database under accession codes 1244, 1245, and 1246, respectively. This paper makes use of RCSB Protein Data Bank accession codes 3SN6, 4LDO and 6FUF.

References

  1. Buck, L. & Axel, R. A novel multigene family may encode odorant receptors: a molecular basis for odor recognition. Cell 65, 175–187 (1991).

    Article  CAS  PubMed  Google Scholar 

  2. Malnic, B., Hirono, J., Sato, T. & Buck, L. B. Combinatorial receptor codes for odors. Cell 96, 713–723 (1999).

    Article  CAS  PubMed  Google Scholar 

  3. Zhao, H. et al. Functional expression of a mammalian odorant receptor. Science 279, 237–242 (1998).

    Article  ADS  CAS  PubMed  Google Scholar 

  4. Mayhew, E. J. et al. Transport features predict if a molecule is odorous. Proc. Natl Acad. Sci. USA 119, e2116576119 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Niimura, Y., Matsui, A. & Touhara, K. Extreme expansion of the olfactory receptor gene repertoire in African elephants and evolutionary dynamics of orthologous gene groups in 13 placental mammals. Genome Res. 24, 1485–1496 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Malnic, B., Godfrey, P. A. & Buck, L. B. The human olfactory receptor gene family. Proc. Natl Acad. Sci. USA 101, 2584–2589 (2004).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  7. Bjarnadóttir, T. K. et al. Comprehensive repertoire and phylogenetic analysis of the G protein-coupled receptors in human and mouse. Genomics 88, 263–273 (2006).

    Article  PubMed  Google Scholar 

  8. Glusman, G., Yanai, I., Rubin, I. & Lancet, D. The complete human olfactory subgenome. Genome Res. 11, 685–702 (2001).

    Article  CAS  PubMed  Google Scholar 

  9. Jones, D. T. & Reed, R. R. Golf: an olfactory neuron specific-G protein involved in odorant signal transduction. Science 244, 790–795 (1989).

    Article  ADS  CAS  PubMed  Google Scholar 

  10. Pourmorady, A. & Lomvardas, S. Olfactory receptor choice: a case study for gene regulation in a multi-enhancer system. Curr. Opin. Genet. Dev. 72, 101–109 (2022).

    Article  CAS  PubMed  Google Scholar 

  11. Butterwick, J. A. et al. Cryo-EM structure of the insect olfactory receptor Orco. Nature 560, 447–452 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  12. Del Mármol, J., Yedlin, M. A. & Ruta, V. The structural basis of odorant recognition in insect olfactory receptors. Nature 597, 126–131 (2021).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  13. Ikegami, K. et al. Structural instability and divergence from conserved residues underlie intracellular retention of mammalian odorant receptors. Proc. Natl Acad. Sci. USA 117, 2957–2967 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  14. Saito, H., Kubota, M., Roberts, R. W., Chi, Q. & Matsunami, H. RTP family members induce functional expression of mammalian odorant receptors. Cell 119, 679–691 (2004).

    Article  CAS  PubMed  Google Scholar 

  15. Cook, B. L. et al. Large-scale production and study of a synthetic G protein-coupled receptor: human olfactory receptor 17-4. Proc. Natl Acad. Sci. USA 106, 11925–11930 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  16. Katada, S., Tanaka, M. & Touhara, K. Structural determinants for membrane trafficking and G protein selectivity of a mouse olfactory receptor. J. Neurochem. 90, 1453–1463 (2004).

    Article  CAS  PubMed  Google Scholar 

  17. Lee, S. J., Depoortere, I. & Hatt, H. Therapeutic potential of ectopic olfactory and taste receptors. Nat. Rev. Drug Discov. 18, 116–138 (2019).

    Article  CAS  PubMed  Google Scholar 

  18. Freitag, J., Ludwig, G., Andreini, I., Rössler, P. & Breer, H. Olfactory receptors in aquatic and terrestrial vertebrates. J. Comp. Physiol. A 183, 635–650 (1998).

    Article  CAS  PubMed  Google Scholar 

  19. Shayya, H. J. et al. ER stress transforms random olfactory receptor choice into axon targeting precision. Cell 185, 3896–3912.e22 (2022).

    Article  CAS  PubMed  Google Scholar 

  20. Saito, H., Chi, Q., Zhuang, H., Matsunami, H. & Mainland, J. D. Odor coding by a mammalian receptor repertoire. Sci. Signal. 2, ra9 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Xu, L. L. et al. PSGR, a novel prostate-specific gene with homology to a G protein-coupled receptor, is overexpressed in prostate cancer. Cancer Res. 60, 6568–6572 (2000).

    CAS  PubMed  Google Scholar 

  22. Gelis, L. et al. Functional characterization of the odorant receptor 51E2 in human melanocytes. J. Biol. Chem. 291, 17772–17786 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Kotlo, K. et al. The olfactory G protein-coupled receptor (Olfr-78/OR51E2) modulates the intestinal response to colitis. Am. J. Physiol. Cell Physiol. 318, C502–C513 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Vadevoo, S. M. P. et al. The macrophage odorant receptor Olfr78 mediates the lactate-induced M2 phenotype of tumor-associated macrophages. Proc. Natl Acad. Sci. USA 118, e2102434118 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Pluznick, J. L. et al. Olfactory receptor responding to gut microbiota-derived signals plays a role in renin secretion and blood pressure regulation. Proc. Natl Acad. Sci. USA 110, 4410–4415 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  26. Flegel, C., Manteniotis, S., Osthold, S., Hatt, H. & Gisselmann, G. Expression profile of ectopic olfactory receptors determined by deep sequencing. PLoS ONE 8, e55368 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. Nakashima, A. et al. Agonist-independent GPCR activity regulates anterior–posterior targeting of olfactory sensory neurons. Cell 154, 1314–1325 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Rasmussen, S. G. F. et al. Crystal structure of the β2 adrenergic receptor–Gs protein complex. Nature 477, 549–555 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Nehmé, R. et al. Mini-G proteins: novel tools for studying GPCRs in their active conformation. PLoS ONE 12, e0175642 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  30. Ring, A. M. et al. Adrenaline-activated structure of β2-adrenoceptor stabilized by an engineered nanobody. Nature 502, 575–579 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  31. Tsai, C. J. et al. Crystal structure of rhodopsin in complex with a mini-Go sheds light on the principles of G protein selectivity. Sci. Adv. 4, eaat7052 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  32. Ballesteros, J. A. & Weinstein, H. [19] Integrated methods for the construction of three-dimensional models and computational probing of structure–function relations in G protein-coupled receptors. Methods Neurosci. 25, 366–428 (1995).

    Article  CAS  Google Scholar 

  33. de March, C. A., Kim, S. K., Antonczak, S., Goddard, W. A. 3rd & Golebiowski, J. G protein-coupled odorant receptors: from sequence to structure. Protein Sci. 24, 1543–1548 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  35. Cichy, A., Shah, A., Dewan, A., Kaye, S. & Bozza, T. Genetic depletion of class I odorant receptors impacts perception of carboxylic acids. Curr. Biol. 29, 2687–2697.e4 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Pronin, A. & Slepak, V. Ectopically expressed olfactory receptors OR51E1 and OR51E2 suppress proliferation and promote cell death in a prostate cancer cell line. J. Biol. Chem. 296, 100475 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Manglik, A. & Kruse, A. C. Structural basis for G protein-coupled receptor activation. Biochemistry 56, 5628–5634 (2017).

    Article  CAS  PubMed  Google Scholar 

  38. Zhou, Q. et al. Common activation mechanism of class A GPCRs. eLife 8, e50279 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Hauser, A. S. et al. GPCR activation mechanisms across classes and macro/microscales. Nat. Struct. Mol. Biol. 28, 879–888 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. de March, C. A. et al. Conserved residues control activation of mammalian G protein-coupled odorant receptors. J. Am. Chem. Soc. 137, 8611–8616 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Dror, R. O. et al. Activation mechanism of the β2-adrenergic receptor. Proc. Natl Acad. Sci. USA 108, 18684–18689 (2011).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  42. Bushdid, C. et al. Mammalian class I odorant receptors exhibit a conserved vestibular-binding pocket. Cell. Mol. Life Sci. 76, 995–1004 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Shim, T. et al. The third extracellular loop of mammalian odorant receptors is involved in ligand binding. Int. J. Mol. Sci. 23, 12501 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  45. Staus, D. P. et al. Sortase ligation enables homogeneous GPCR phosphorylation to reveal diversity in β-arrestin coupling. Proc. Natl Acad. Sci. USA 115, 3834–3839 (2018).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  46. Mastronarde, D. N. SerialEM: a program for automated tilt series acquisition on Tecnai microscopes using prediction of specimen position. Microsc. Microanal. 9, 1182–1183 (2003).

    Article  ADS  Google Scholar 

  47. 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 

  48. Punjani, A., Rubinstein, J. L., Fleet, D. J. & Brubaker, M. A. cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 14, 290–296 (2017).

    Article  CAS  PubMed  Google Scholar 

  49. Asarnow, D., Palovcak, E. & Cheng, Y. Asarnow/pyem: UCSF Pyem v0.5. Zenodo https://doi.org/10.5281/zenodo.3576630 (2019).

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

    Article  CAS  PubMed  Google Scholar 

  51. Scheres, S. H. W. RELION: implementation of a Bayesian approach to cryo-EM structure determination. J. Struct. Biol. 180, 519–530 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Croll, T. I. ISOLDE: a physically realistic environment for model building into low-resolution electron-density maps. Acta Crystallogr. D Struct. Biol. 74, 519–530 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  Google Scholar 

  55. Schüttelkopf, A. W. & van Aalten, D. M. F. PRODRG: a tool for high-throughput crystallography of protein–ligand complexes. Acta Crystallogr. D Biol. Crystallogr. 60, 1355–1363 (2004).

    Article  PubMed  Google Scholar 

  56. Bushdid, C., de March, C. A., Matsunami, H. & Golebiowski, J. Numerical models and in vitro assays to study odorant receptors. Methods Mol. Biol. 1820, 77–93 (2018).

    Article  CAS  PubMed  Google Scholar 

  57. Zhang, Y., Pan, Y., Matsunami, H. & Zhuang, H. Live-cell measurement of odorant receptor activation using a real-time cAMP assay. J. Vis. Exp. 128, 55831 (2017).

    Google Scholar 

  58. Zhuang, H. & Matsunami, H. Evaluating cell-surface expression and measuring activation of mammalian odorant receptors in heterologous cells. Nat. Protoc. 3, 1402–1413 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Krautwurst, D., Yau, K. W. & Reed, R. R. Identification of ligands for olfactory receptors by functional expression of a receptor library. Cell 95, 917–926 (1998).

    Article  CAS  PubMed  Google Scholar 

  60. Berendsen, H. J. C., van der Spoel, D. & van Drunen, R. GROMACS: a message-passing parallel molecular dynamics implementation. Comput. Phys. Commun. 91, 43–56 (1995).

    Article  ADS  CAS  Google Scholar 

  61. Huang, J. et al. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nat. Methods 14, 71–73 (2017).

    Article  CAS  PubMed  Google Scholar 

  62. Madhavi Sastry, G., Adzhigirey, M., Day, T., Annabhimoju, R. & Sherman, W. Protein and ligand preparation: parameters, protocols, and influence on virtual screening enrichments. J. Comput. Aided Mol. Des. 27, 221–234 (2013).

    Article  ADS  CAS  PubMed  Google Scholar 

  63. Jo, S., Kim, T., Iyer, V. G. & Im, W. CHARMM-GUI: a web-based graphical user interface for CHARMM. J. Comput. Chem. 29, 1859–1865 (2008).

    Article  CAS  PubMed  Google Scholar 

  64. Lomize, M. A., Pogozheva, I. D., Joo, H., Mosberg, H. I. & Lomize, A. L. OPM database and PPM web server: resources for positioning of proteins in membranes. Nucleic Acids Res. 40, D370–D376 (2012).

    Article  CAS  PubMed  Google Scholar 

  65. Evans, D. J. & Holian, B. L. The Nose–Hoover thermostat. J. Chem. Phys. 83, 4069–4074 (1985).

    Article  ADS  CAS  Google Scholar 

  66. Parrinello, M. & Rahman, A. Polymorphic transitions in single crystals: a new molecular dynamics method. J. Appl. Phys. 52, 7182–7190 (1981).

    Article  ADS  CAS  Google Scholar 

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

    Article  ADS  CAS  Google Scholar 

  68. Halgren, T. New method for fast and accurate binding-site identification and analysis. Chem. Biol. Drug Des. 69, 146–148 (2007).

    Article  CAS  PubMed  Google Scholar 

  69. Halgren, T. A. Identifying and characterizing binding sites and assessing druggability. J. Chem. Inf. Model. 49, 377–389 (2009).

    Article  CAS  PubMed  Google Scholar 

  70. Friesner, R. A. et al. Extra precision glide: docking and scoring incorporating a model of hydrophobic enclosure for protein–ligand complexes. J. Med. Chem. 49, 6177–6196 (2006).

    Article  CAS  PubMed  Google Scholar 

  71. 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  PubMed  Google Scholar 

  72. Friesner, R. A. et al. Glide: a new approach for rapid, accurate docking and scoring. 1. Method and assessment of docking accuracy. J. Med. Chem. 47, 1739–1749 (2004).

    Article  CAS  PubMed  Google Scholar 

  73. Larkin, M. A. et al. Clustal W and Clustal X version 2.0. Bioinformatics 23, 2947–2948 (2007).

    Article  CAS  PubMed  Google Scholar 

  74. Waterhouse, A. M., Procter, J. B., Martin, D. M. A., Clamp, M. & Barton, G. J. Jalview version 2—a multiple sequence alignment editor and analysis workbench. Bioinformatics 25, 1189–1191 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Pándy-Szekeres, G. et al. GPCRdb in 2023: state-specific structure models using AlphaFold2 and new ligand resources. Nucleic Acids Res. 51, D395–D402 (2022).

    Article  PubMed Central  Google Scholar 

  76. Pagès, H., Aboyoun, P., Gentleman, R. & DebRoy, S. Biostrings: efficient manipulation of biological strings. R package version 2.66.0 (Bioconductor, 2022).

  77. Charif, D. & Lobry, J. R. in Structural Approaches to Sequence Evolution: Molecules, Networks, Populations (eds Bastolla, U. et al.) 207–232 (Springer, 2007).

  78. Paradis, E. & Schliep, K. ape 5.0: An environment for modern phylogenetics and evolutionary analyses in R. Bioinformatics 35, 526–528 (2019).

    Article  CAS  PubMed  Google Scholar 

  79. Xu, S. et al. Ggtree: a serialized data object for visualization of a phylogenetic tree and annotation data. iMeta 1, e56 (2022).

    Article  Google Scholar 

  80. Dang, S. et al. Cryo-EM structures of the TMEM16A calcium-activated chloride channel. Nature 552, 426–429 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank D. Toso at Cal-Cryo at QB3-Berkeley for help in microscope operation and data collection; and H.M., C.A.d.M. and J.T. thank M. J. Ni and H.-Y. Lu for their technical support. This work was supported by the US NIH grant R01DC020353 (to H.M., N.V. and A.M.) and K99DC018333 (to C.A.d.M.). Cryo-EM equipment at UCSF is partially supported by NIH grants S10OD020054 and S10OD021741. This project was funded by the UCSF Program for Breakthrough Biomedical Research, funded in part by the Sandler Foundation. A.M. acknowledges support from the Edward Mallinckrodt Jr Foundation and the Vallee Foundation, and is a Chan Zuckerberg Biohub Investigator. H.M. acknowledges support from NSF/CIHR/DFG/FRQ/UKRI-MRC Next Generation Networks for Neuroscience Program (award #2014217).

Author information

Authors and Affiliations

Authors

Contributions

C.B.B., C.A.d.M., W.J.C.v.d.V., N.V., H.M. and A.M. designed the study. C.B.B. cloned constructs, prepared baculoviruses, expressed and purified G protein-complexing reagents, and optimized large-scale production of OR51E2. C.B.B. worked out conditions to biochemically purify and stabilize the propionate-bound OR51E2–Gs complex, and identified optimal cryo-EM grid preparation procedures following screening, collection and processing of 200-kV cryo-EM data. B.F. and A.M. performed 300-kV cryo-EM data collection. C.B.B. determined high-resolution cryo-EM maps by extensive image processing with input from A.M. A.M. and C.B.B. built, refined models of propionate-bound OR51E2 in complex with Gs and Nb35. C.B.B. and A.M. analysed cryo-EM data and models, and prepared figures and tables. C.A.d.M. and J.T. analysed OR models and sequences to design and clone OR mutants, performed Glosensor signalling experiments for OR functional activity and generated OR cell-surface expression data by flow cytometry with input from H.M. C.A.d.M. and J.T. analysed and prepared figures and tables for signalling and flow cytometry data. C.A.d.M. built the phylogenetic tree of ORs and non-olfactory class A GPCRs. N.M. set up and performed MD simulations and ligand docking, and performed binding pocket volume calculations. W.J.C.v.d.V. analysed simulation trajectories and prepared figures describing simulation data. W.J.C.v.d.V., N.M. and N.V. provided mechanistic insight from simulation data. C.L.d.T. performed bioinformatic analysis of OR and non-olfactory class A GPCR conservation. L.L. and C.B.B. performed pilot GloSensor signalling studies in suspension cells. C.B.B., C.A.d.M. and A.M. wrote an initial draft of the manuscript and generated figures with contributions from all authors. Further edits to the manuscript were provided by W.J.C.v.d.V., N.M., N.V. and H.M. The overall project was supervised by N.V., H.M. and A.M.

Corresponding authors

Correspondence to Nagarajan Vaidehi, Hiroaki Matsunami or Aashish Manglik.

Ethics declarations

Competing interests

H.M. has received royalties from Chemcom, research grants from Givaudan and consultant fees from Kao.

Peer review

Peer review information

Nature thanks David Gloriam and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

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 Alignment of OR51E2, rhodopsin and β2 adrenergic receptor (β2AR) amino acid sequences as described in part by de March et al.33 and implemented on GPCRdb75.

Conservation is highlighted from low (white) to high (dark blue) and the consensus amino acid is shown. Transmembrane domains are boxed in yellow. The most conserved residue in class A GPCRs for each transmembrane domain is boxed and labeled in orange. Residues used to align OR and Class A GPCR sequences are highlighted by asterisks, which are colored orange when the residue is common to all Class A GPCRs and black when it is specific to ORs. The most conserved residues used for numbering of the intracellular and extracellular loops are also indicated in italic when available. Generic numbers follow the revised Ballesteros-Weinstein numbering for Class A GPCRs32,34.

Extended Data Fig. 2 Biochemical preparation of OR51E2-Gs complex bound to propionate.

a) Schematic outlining the strategy for stabilization and purification of the activated OR51E2-Gs complex bound to propionate. b) GloSensor cAMP assay demonstrating that fusion of miniGs to OR51E2 blocks activation of endogenous Gs in response to treatment with propionate, suggesting that miniGs couples to the OR51E2 transmembrane core. Data points are the mean of analytical replicates from a representative experiment. Error bars represent the standard deviation between replicates (n = 4). c) Size-exclusion chromatogram of purified OR51E2-Gs-Nb35 complex used for structure determinations shown together with a representative SDS-PAGE gel analysis of the collected fraction containing the OR51E2-Gs-Nb35 complex. We observe two bands for OR51E2, likely due to heterogeneous glycosylation of the receptor N-terminus.

Extended Data Fig. 3 Cryo-EM data processing for OR51E2-Gs.

a) A representative cryo-EM micrograph from the curated OR51E2-Gs dataset (n = 8,010) obtained from a Titan Krios microscope. b) A subset of highly populated, reference-free 2D-class averages are shown. Scale bar is 50 Å. c) Schematic showing the image processing workflow for OR51E1-Gs. Initial processing was performed using UCSF MotionCor2 and cryoSPARC. Particles were then transferred using the pyem script package49 to RELION for alignment-free 3D classification. Finally, particles were processed in cryoSPARC using the non-uniform and local refinement tools. Dashed boxes indicate selected classes, and 3D volumes of classes and refinements are shown along with global Gold-standard Fourier Shell Correlation (GSFSC) resolutions. d,e) Map validation for the OR51E2-Gs (d) globally refined, and (e) locally refined cryo-EM maps. GSFSC curves are calculated in cryoSPARC, and shown together with directional FSC (dFSC) curves generated with dfsc.0.0.1.py as previously described80. Map-model correlations calculated in the Phenix suite are also shown. Arrows indicate map and map-model resolution estimates at 0.143 and 0.5 correlation respectively. Euler angle distributions calculated in cryoSPARC are also provided for each map.

Extended Data Fig. 4 Cryo-EM density and atomic model.

a) Orthogonal views of local resolution for the globally refined map of OR51E2-Gs calculated with the local resolution estimation tool in cryoSPARC. b) Close-up view showing the local resolution of the propionate binding site. c) Representative cryo-EM densities from the 3D reconstruction of OR51E2 from a sharpened, globally refined map of OR51E2-Gs at a map threshold of 0.635. Shown are the transmembrane helices and loop regions of OR51E2 as well as the C-terminal helix of miniGαs. d) Close-up view of cryo-EM density (yellow sticks and density) supporting propionate binding pose using a sharpened map locally refined around only the 7TM domain of OR51E2 at map threshold of 1.0.

Extended Data Fig. 5 Interactions between propionate and OR51E2 in molecular dynamics simulations.

a) Minimum distance plot between R2626×59 and propionate from 5 independent runs at different velocities (top to bottom). Minimum distance was measured between guanidinium nitrogens of R2626×59 and oxygens of propionate. Thick trace represents smoothed values with an averaging window of 8 nanoseconds; thin trace represents unsmoothed values. b) Root-mean-square deviation (RMSD) values of production simulation runs for propionate calculated with reference to the equilibrated structure of OR51E2 prior to 1 µs production simulation from 5 independent runs at different velocities (top to bottom). c) Minimum distances (Ȧ) between ligand heavy atoms and residue side chain heavy atoms (hydrogen bond and van der Waals contacts combined) are shown in gray. Gray dashed arrows highlight the interactions made between a certain receptor residue and ligand atom(s). All distances are shown as means from n = 5 independent runs (at different velocities) each 1 μs long. Standard deviation of measurement for each of the residue-ligand distance are as follows; 0.03 Å (R2626×59), 0.10 Å(S2586×55), 0.16 Å (I2025×43), 0.12 Å (G1985×39), 0.23 Å (Q18145×53), 0.23 Å (H18045×52), 0.25 Å (L1584×60), and 0.14 Å (H1043×33).

Extended Data Fig. 6 Conservation of residues within the odorant binding pocket.

a) View of propionate-contacting residues. Conservation weblogo of key residues in Class I (b) and Class II ORs (c). d) The percentage of receptors harboring a given amino acid at each position are shown for all human Class I and Class II ORs. OR51E2 residues at each position are indicated by a black box.

Extended Data Fig. 7 Analysis of active state structure of OR51E2.

a) Structural comparison of G protein interaction for OR51E2 (green) and β2-adrenergic receptor (β2AR in blue, PDB code: 3SN6). b) Close-up views of intracellular loop 2 (ICL2) interaction with the Gαs subunit shown in surface representation. c) interactions between residues in ICL2 and the αΝ and α5 helices of the Gαs subunit. d) G protein-coupling region of OR51E2 is shown along with a weblogo (right) highlighting conservation of key residues for all human ORs. e) Residues that participate in the extended interaction hydrogen bonding network between TM3, TM4, TM5, and TM6 are conserved in human Class I ORs, but not in Class II ORs. f,g) The percentage of receptors harboring a given amino acid at each position are shown for all human Class I and Class II ORs at the G protein-coupling region and connector regions. OR51E2 residues at each position are indicated by a black box.

Extended Data Fig. 8 OR51E2 molecular dynamics simulation trajectories.

ac) Simulation trajectories for WT and Q18145×53D OR51E2 are shown in a–c. Five independent runs at different velocities are shown for each condition (top to bottom). a) F2506×47 χ1 angle over replicate simulations. b) Minimum distance between oxygen atoms of the hydroxyl groups in the side chains of S111 and Y2516×48 over replicate simulations. c) Minimum distance between R2626×59 sidechain atoms and G1985×39 mainchain atoms (excluding the hydrogens) for replicate simulations. d) Root-mean-square deviation (RMSD) values for TM backbone atoms in the transmembrane helices (see Methods) calculated with reference to the equilibrated structure of the no ligand and propionate bound OR51E2 simulations, as well as for simulations of Q18145×53D OR51E2 from 5 independent MD simulation replicates (top to bottom). Thick traces represent smoothed values with an averaging window of 8 nanoseconds; thin traces represent unsmoothed values. ef) Aggregate frequency distributions are shown for F2506×47 χ1 angle (e), minimum distance between heavy atoms of the hydroxyl groups of S1113×40 and Y2516×48 (f), and minimum distance between R2626×59 sidechain heavy atoms and G1985×39 main chain heavy atoms (excluding hydrogens) (g) using all five simulation replicates for each condition.

Extended Data Fig. 9 Molecular dynamics snapshots of OR51E2.

a) Comparison of cryo-EM structure of propionate-bound OR51E2 with representative snapshots from simulations of WT OR51E2 with propionate, WT OR51E2 without ligand, and Q18145×53D OR51E2 without ligand. Notably, OR51E2 does not transition to the inactive conformation in any of these simulations. b) Close-up views of OR51E2 binding site and ECL3 region in the cryo-EM structure and simulations. In propionate-bound MD simulations of WT OR51E2, R2626×59 persistently forms an ionic interaction with propionate. In simulations of WT OR51E2 with propionate removed, R2626×59 is flexible. Introduction of Asp in position 45x53 (Q18145×53D) stabilizes R2626×59 in an active-like state by a direct ionic interaction. c) Close-up views of OR51E2 connector region shows increased flexibility of WT OR51E2 simulated without propionate. This flexibility is decreased for the Q18145×53D mutant. In a-c, displayed snapshots are the last 1000th ns snapshots from each simulation replicate. d,e) Molecular dynamics trajectories from representative simulations to highlight structural organization of connector region. d) Minimum distance between S1113×40 and Y2516×47 hydroxyl groups is comparable for Q18145×53D and propionate-bound WT OR51E2. e) Rotamer angle of F2506×47is comparable for Q18145×53D and propionate-bound WT OR51E2. Simulations were performed with or without propionate over the course of 1000 ns (see Extended Data Fig. 8 for replicates of simulation trajectories). Thick traces represent smoothed values with an averaging window of 8 nanoseconds; thin traces represent unsmoothed values.

Extended Data Fig. 10 AlphaFold2 model of OR51E2.

a) AlphaFold2 predicted structure of OR51E2. The pLDDT confidence metric is shown highlighting relatively high confidence in the transmembrane regions and extracellular loops. b) AlphaFold2 predicted structure of unbound OR51E2 (gray) superimposed onto the experimentally determined structure of propionate-bound OR51E2 in the active state (green cartoon and yellow spheres). In the AlphaFold2 model, TM6 is inwardly displaced compared to the active structure. Closeup views of (c) the Connector region and (d) the G protein-coupling region are provided. e) Slice through surface representation of AlphaFold2 predicted OR51E2, suggests solvent accessibility of the ligand binding site in the inactive state.

Extended Data Table 1 Cryo-EM data collection, refinement, and validation statistics
Extended Data Table 2 Expression and pharmacodynamic constants for OR51E2 variants
Extended Data Table 3 Pharmacodynamic constants for fatty acid series at OR51E2 variants

Supplementary information

Supplementary Information

This file contains Supplementary Tables 1–8, which includes summary statistics for molecular dynamics simulations and a structural comparison of OR51E2 to other Class A GPCRs. Also included are Supplementary Figs. 1 and 2, which include snapshots of molecular dynamics simulations and an uncropped SDS-PAGE gel for Extended Data Fig. 2, respectively, and Supplementary Fig. 3, which shows the flow cytometry protocol and gating strategy.

Reporting Summary

Peer Review File

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) 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

Billesbølle, C.B., de March, C.A., van der Velden, W.J.C. et al. Structural basis of odorant recognition by a human odorant receptor. Nature 615, 742–749 (2023). https://doi.org/10.1038/s41586-023-05798-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-023-05798-y

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