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:

Bitter taste receptor activation by cholesterol and an intracellular tastant

Abstract

Bitter taste sensing is mediated by type 2 taste receptors (TAS2Rs (also known as T2Rs)), which represent a distinct class of G-protein-coupled receptors1. Among the 26 members of the TAS2Rs, TAS2R14 is highly expressed in extraoral tissues and mediates the responses to more than 100 structurally diverse tastants2,3,4,5,6, although the molecular mechanisms for recognizing diverse chemicals and initiating cellular signalling are still poorly understood. Here we report two cryo-electron microscopy structures for TAS2R14 complexed with Ggust (also known as gustducin) and Gi1. Both structures have an orthosteric binding pocket occupied by endogenous cholesterol as well as an intracellular allosteric site bound by the bitter tastant cmpd28.1, including a direct interaction with the α5 helix of Ggust and Gi1. Computational and biochemical studies validate both ligand interactions. Our functional analysis identified cholesterol as an orthosteric agonist and the bitter tastant cmpd28.1 as a positive allosteric modulator with direct agonist activity at TAS2R14. Moreover, the orthosteric pocket is connected to the allosteric site via an elongated cavity, which has a hydrophobic core rich in aromatic residues. Our findings provide insights into the ligand recognition of bitter taste receptors and suggest activities of TAS2R14 beyond bitter taste perception via intracellular allosteric tastants.

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: Overall structures of TAS2R14–Gi1–scFv16 and TAS2R14–Ggust–scFv16 complexes.
Fig. 2: Activation of TAS2R14 by cholesterol binding in the orthosteric site.
Fig. 3: Distinct activation of TAS2R14 induced by cmpd28.1 in intracellular allosteric site.
Fig. 4: Distinct structural features facilitate binding of Gαi1 and Ggust.

Similar content being viewed by others

Data availability

Atomic coordinates and cryo-EM maps have been deposited in the Protein Data Bank and Electron Microscopy Data Bank under the accession codes PDB 8VY7 (EMDB accession EMD-43647 for the globally refined map and EMD-43656 for the locally refined map) and PDB 8VY9 (EMD-43650 for the globally refined map and EMD-43657 for the locally refined map) for the TAS2R14–Gi1–scFv16 and TAS2R14–Ggust–scFv16 complexes, respectively. Source data are provided with this paper.

References

  1. Gilbertson, T. A., Damak, S. & Margolskee, R. F. The molecular physiology of taste transduction. Curr. Opin. Neurobiol. 10, 519–527 (2000).

    Article  CAS  PubMed  Google Scholar 

  2. Ji, M. et al. Identification of novel compounds for human bitter taste receptors. Chem. Biol. Drug Des. 84, 63–74 (2014).

    Article  CAS  PubMed  Google Scholar 

  3. Levit, A. et al. The bitter pill: clinical drugs that activate the human bitter taste receptor TAS2R14. FASEB J. 28, 1181–1197 (2014).

    Article  CAS  PubMed  Google Scholar 

  4. Jaggupilli, A. et al. Analysis of the expression of human bitter taste receptors in extraoral tissues. Mol. Cell. Biochem. 426, 137–147 (2017).

    Article  CAS  PubMed  Google Scholar 

  5. Mueller, K. L. et al. The receptors and coding logic for bitter taste. Nature 434, 225–229 (2005).

    Article  ADS  CAS  PubMed  Google Scholar 

  6. Lang, T., Di Pizio, A., Risso, D., Drayna, D. & Behrens, M. Activation profile of TAS2R2, the 26th human bitter taste receptor. Mol. Nutr. Food Res. 67, e2200775 (2023).

    Article  PubMed  Google Scholar 

  7. Chandrashekar, J. et al. T2Rs function as bitter taste receptors. Cell 100, 703–711 (2000).

    Article  CAS  PubMed  Google Scholar 

  8. Chandrashekar, J., Hoon, M. A., Ryba, N. J. & Zuker, C. S. The receptors and cells for mammalian taste. Nature 444, 288–294 (2006).

    Article  ADS  CAS  PubMed  Google Scholar 

  9. Zhang, Y. et al. Coding of sweet, bitter, and umami tastes: different receptor cells sharing similar signaling pathways. Cell 112, 293–301 (2003).

    Article  CAS  PubMed  Google Scholar 

  10. Deshpande, D. A. et al. Bitter taste receptors on airway smooth muscle bronchodilate by localized calcium signaling and reverse obstruction. Nat. Med. 16, 1299–1304 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Kim, D., Woo, J. A., Geffken, E., An, S. S. & Liggett, S. B. Coupling of airway smooth muscle bitter taste receptors to intracellular signaling and relaxation is via Gαi1,2,3. Am. J. Respir. Cell Mol. Biol. 56, 762–771 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Medapati, M. R. et al. Bitter taste receptor T2R14 detects quorum sensing molecules from cariogenic Streptococcus mutans and mediates innate immune responses in gingival epithelial cells. FASEB J. 35, e21375 (2021).

    Article  CAS  PubMed  Google Scholar 

  13. Shaik, F. A., Medapati, M. R. & Chelikani, P. Cholesterol modulates the signaling of chemosensory bitter taste receptor T2R14 in human airway cells. Am. J. Physiol. 316, L45–L57 (2019).

    CAS  Google Scholar 

  14. Ziegler, F., Steuer, A., Di Pizio, A. & Behrens, M. Physiological activation of human and mouse bitter taste receptors by bile acids. Commun. Biol. 6, 612 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Waterloo, L. et al. Discovery of 2-aminopyrimidines as potent agonists for the bitter taste receptor TAS2R14. J. Med. Chem. 66, 3499–3521 (2023).

    Article  CAS  PubMed  Google Scholar 

  16. Tabula Sapiens, C. et al. The Tabula Sapiens: a multiple-organ, single-cell transcriptomic atlas of humans. Science 376, eabl4896 (2022).

    Article  Google Scholar 

  17. Inoue, A. et al. Illuminating G-protein-coupling selectivity of GPCRs. Cell 177, 1933–1947.e1925 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Marheineke, K., Grunewald, S., Christie, W. & Reilander, H. Lipid composition of Spodoptera frugiperda (Sf9) and Trichoplusia ni (Tn) insect cells used for baculovirus infection. FEBS Lett. 441, 49–52 (1998).

    Article  CAS  PubMed  Google Scholar 

  19. Deshpande, I. et al. Smoothened stimulation by membrane sterols drives Hedgehog pathway activity. Nature 571, 284–288 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Huang, P. et al. Structural basis of smoothened activation in Hedgehog signaling. Cell 174, 312–324.e316 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Xu, W. et al. Structural basis for strychnine activation of human bitter taste receptor TAS2R46. Science 377, 1298–1304 (2022).

    Article  ADS  CAS  PubMed  Google Scholar 

  22. Rasmussen, S. G. 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 

  23. Di Pizio, A. et al. Comparing class A GPCRs to bitter taste receptors: structural motifs, ligand interactions and agonist-to-antagonist ratios. Methods Cell. Biol. 132, 401–427 (2016).

    Article  PubMed  Google Scholar 

  24. Muenks, A., Zepeda, S., Zhou, G., Veesler, D. & DiMaio, F. Automatic and accurate ligand structure determination guided by cryo-electron microscopy maps. Nat. Commun. 14, 1164 (2023).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  25. Olsen, R. H. J. et al. TRUPATH, an open-source biosensor platform for interrogating the GPCR transducerome. Nat. Chem. Biol. 16, 841–849 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Di Pizio, A. et al. Rational design of agonists for bitter taste receptor TAS2R14: from modeling to bench and back. Cell. Mol. Life Sci. 77, 531–542 (2020).

    Article  ADS  PubMed  Google Scholar 

  27. Dagan-Wiener, A. et al. BitterDB: taste ligands and receptors database in 2019. Nucleic Acids Res. 47, D1179–D1185 (2019).

    Article  PubMed  Google Scholar 

  28. Krumm, B. E. et al. Neurotensin receptor allosterism revealed in complex with a biased allosteric modulator. Biochemistry 62, 1233–1248 (2023).

    Article  CAS  PubMed  Google Scholar 

  29. Zhao, L. H. et al. Conserved class B GPCR activation by a biased intracellular agonist. Nature 621, 635–641 (2023).

    Article  ADS  CAS  PubMed  Google Scholar 

  30. Yang, F. et al. Structural basis of GPBAR activation and bile acid recognition. Nature 587, 499–504 (2020).

    Article  ADS  CAS  PubMed  Google Scholar 

  31. Slosky, L. M. et al. β-Arrestin-biased allosteric modulator of NTSR1 selectively attenuates addictive behaviors. Cell 181, 1364–1379.e1314 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Tamura, T. et al. Identification of an orally active small-molecule PTHR1 agonist for the treatment of hypoparathyroidism. Nat. Commun. 7, 13384 (2016).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. Audet, M. & Stevens, R. C. Emerging structural biology of lipid G protein-coupled receptors. Protein Sci. 28, 292–304 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Chrencik, J. E. et al. Crystal structure of antagonist bound human lysophosphatidic acid receptor 1. Cell 161, 1633–1643 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Srivastava, A. et al. High-resolution structure of the human GPR40 receptor bound to allosteric agonist TAK-875. Nature 513, 124–127 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  36. Hua, T. et al. Crystal structure of the human cannabinoid receptor CB1. Cell 167, 750–762.e714 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Hori, T. et al. Na+-mimicking ligands stabilize the inactive state of leukotriene B4 receptor BLT1. Nat. Chem. Biol. 14, 262–269 (2018).

    Article  CAS  PubMed  Google Scholar 

  38. Cao, C. et al. Structural basis for signal recognition and transduction by platelet-activating-factor receptor. Nat. Struct. Mol. Biol. 25, 488–495 (2018).

    Article  CAS  PubMed  Google Scholar 

  39. Taniguchi, R. et al. Structural insights into ligand recognition by the lysophosphatidic acid receptor LPA6. Nature 548, 356–360 (2017).

    Article  ADS  CAS  PubMed  Google Scholar 

  40. Bansal, P. D., Dutta, S. & Shukla, D. Activation mechanism of the human Smoothened receptor. Biophys. J. 122, 1400–1413 (2023).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  41. Radhakrishnan, A., Rohatgi, R. & Siebold, C. Cholesterol access in cellular membranes controls Hedgehog signaling. Nat. Chem. Biol. 16, 1303–1313 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Bufe, B., Hofmann, T., Krautwurst, D., Raguse, J. D. & Meyerhof, W. The human TAS2R16 receptor mediates bitter taste in response to beta-glucopyranosides. Nat. Genet. 32, 397–401 (2002).

    Article  CAS  PubMed  Google Scholar 

  43. Tan, S. M. & Seetoh, W. G. Construction of a bioluminescence-based assay for bitter taste receptors (TAS2Rs). Sci Rep. 12, 17658 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  44. Cao, C. et al. Structure, function and pharmacology of human itch GPCRs. Nature 600, 170–175 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  Google Scholar 

  46. 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 

  47. Han, J. et al. Ligand and G-protein selectivity in the κ-opioid receptor. Nature 617, 417–425 (2023).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  48. 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 

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

    Article  ADS  CAS  PubMed  PubMed Central  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  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  52. Chen, V. B. et al. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr. D 66, 12–21 (2010).

    Article  ADS  CAS  PubMed  Google Scholar 

  53. 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 

  54. Klein, U., Gimpl, G. & Fahrenholz, F. Alteration of the myometrial plasma membrane cholesterol content with beta-cyclodextrin modulates the binding affinity of the oxytocin receptor. Biochemistry 34, 13784–13793 (1995).

    Article  CAS  PubMed  Google Scholar 

  55. Gumpper, R. H., Fay, J. F. & Roth, B. L. Molecular insights into the regulation of constitutive activity by RNA editing of 5HT2C serotonin receptors. Cell Rep. 40, 111211 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. 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 

  57. Brooks, B. R. et al. CHARMM: the biomolecular simulation program. J. Comput. Chem. 30, 1545–1614 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. 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 

  59. Lee, J. et al. CHARMM-GUI input generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM simulations using the CHARMM36 additive force field. J. Chem. Theory Comput. 12, 405–413 (2016).

    Article  CAS  PubMed  Google Scholar 

  60. Wu, E. L. et al. CHARMM-GUI Membrane Builder toward realistic biological membrane simulations. J. Comput. Chem. 35, 1997–2004 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Roe, D. R. & Cheatham, T. E. 3rd PTRAJ and CPPTRAJ: software for processing and analysis of molecular dynamics trajectory data. J. Chem. Theory Comput. 9, 3084–3095 (2013).

    Article  CAS  PubMed  Google Scholar 

  62. Michaud-Agrawal, N., Denning, E. J., Woolf, T. B. & Beckstein, O. MDAnalysis: a toolkit for the analysis of molecular dynamics simulations. J. Comput. Chem. 32, 2319–2327 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Chovancova, E. et al. CAVER 3.0: a tool for the analysis of transport pathways in dynamic protein structures. PLoS Comput. Biol. 8, e1002708 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by NIH grant U24DK116195, the NIMH Psychoactive Drug Screening Program, and the Michael Hooker Distinguished Professorship to B.L.R. The authors acknowledge E. Robinson and J. Strauss for technical assistance in this project. This work utilized the NMR Spectrometer Systems at Mount Sinai acquired with funding from NIH SIG grants 1S10OD025132 and 1S10OD028504. This work was also supported by NIH grants T32HD040127 and K99GM149931 to D.D.K.

Author information

Authors and Affiliations

Authors

Contributions

Y.K. designed the experiments, performed cloning of mutagenesis constructs, performed expression and purification of the TAS2R14–Gi1–scFv16 and TAS2R14–Ggust–scFv16 complexes, prepared the cryo-EM grids, processed cryo-EM data, built and refined the models, analysed structural data, performed the BRET2 and cAMP inhibition assays, and prepared the figures, tables and manuscript. R.H.G. helped with cryo-EM data processing, performed the molecular docking and molecular dynamics simulations, edited the manuscript, and prepared the figures and manuscript for the molecular docking and molecular dynamics simulations. Y.L. performed the cloning of the mutagenesis constructs and conducted BRET2 assays. D.D.K. conducted RT–qPCR for TAS2R14 RNA abundance and prepared the figures and manuscript for the RT–qPCR and mRNA analyses. Y.X. and Z.D. synthesized cmpd28.1. C.C. helped with the purification and provided the expression plasmid for TAS2R14 and Gi1. B.E.K. provided the expression plasmid for Ggust. M.K.J. helped with the BRET2 assays. S.Z. helped with the purification. J.J. supervised the synthesis of cmpd28.1. B.L.R. supervised the entire project and edited the manuscript.

Corresponding author

Correspondence to Bryan L. Roth.

Ethics declarations

Competing interests

J.J. is a co-founder and equity shareholder in Cullgen Inc., a scientific co-founder and scientific advisory board member of Onsero Therapeutics Inc., and a consultant for Cullgen Inc., EpiCypher Inc., Accent Therapeutics Inc. and TavotekBiotherapeutics Inc. The Jin laboratory received research funds from Celgene Corporation, Levo Therapeutics Inc., Cullgen Inc. and Cullinan Oncology Inc. B.L.R. is a co-founder of Onsero and Epiodyne and a scientific advisory board member of Escient Pharmaceuticals and Lassogen. University of North Carolina has licensed TRUPATH technology to many pharmaceutical and biotechnology companies.

Peer review

Peer review information

Nature thanks Antonella di Pizio, Danielle Reed and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 Cryo-EM data processing for the TAS2R14–Gi1–scFv16 complex.

a, Workflow of cryo-EM data processing for TAS2R14–Gi1–scFv16 complex. Masks used for focused classification and local refinement were colored pink and magenta, respectively. The locally refined map is rotated by 90° based on the y-axis from the globally refined map for clarity. b, A representative motion-corrected micrograph imaged at a nominal 81,000 × magnification. The experiment was repeated three times with similar results. c, Representative 2D class averages without alignment. d, Map-to-model FSC curve calculated by mtriage in Phenix software. e-j, Representative local resolution (e), Gold-standard Fourier shell correlation (GSFSC) curves adopted from cryoSPARCv4 (f), and particle angular distributions (g) for globally refined map (e-g) and locally refined map (h-j).

Extended Data Fig. 2 Cryo-EM data processing for the TAS2R14–Ggust–scFv16 complex.

a, Workflow of cryo-EM data processing for TAS2R14–Ggust–scFv16 complex. Masks used for focused classification and local refinement were colored pink and magenta, respectively. The locally refined map is rotated by 90° based on the y-axis from the globally refined map for clarity. b, A representative motion-corrected micrograph imaged at a nominal 81,000 × magnification. The experiment was repeated three times with similar results. c, Representative 2D class averages without alignment. d, Map-to-model FSC curve calculated by mtriage in Phenix software. e-j, Representative local resolution (e), Gold-standard Fourier shell correlation (GSFSC) curves adopted from cryoSPARCv4 (f), and particle angular distributions (g) for globally refined map (e-g) and locally refined map (h-j).

Extended Data Fig. 3 Structural features of TAS2R14 coupled with Gi1 and Ggust.

a-j, Representative atomic model-fitted EM map for the structure of TAS2R14–Gi1–scFv16 (a-e), and TAS2R14–Ggust–scFv16 (f-j), respectively. Cholesterol and cmpd28.1 which are shown as model-fitted EM maps were observed in the orthosteric pocket and allosteric site, respectively (b,g). Cmpd28.1 in the allosteric site interacts with Hα5 of Gαi1 (c) and Gαgust (h). No interactions were observed between ICL1 and Gβ1 in both complexes (d,i). e, W124 of ICL2 forms a hydrogen bond with R32 of αN of Gα in TAS2R14–Gi1–scFv16. The distances between R125 of ICL2 and E28 of αN of Gα in TAS2R14–Ggust–scFv16 is 9 Å indicated by a red arrow. j, E28 of αN of Gαgust forms a hydrogen bond and an ionic interaction with W124 and R125 of ICL2, respectively. k, Two orthogonal views aligned onto the receptors between TAS2R14–Ggust–scFv16 and TAS2R14–Gi1–scFv16. Black dashed lines indicate two HαN helices distant by 8.5 Å. Models are presented as cylinders for clarity. l-n, Structural comparison of TAS2R14–Ggust with reported structure TAS2R46-miniGs/gust (PDB 7XP6). This comparison highlights TMs 4, 5, and 7 (l), as well as 6 (m). n, Conserved residue W3.32 (red circle) of TAS2R14 in the orthosteric pocket is located in the same position as TAS2R46. Cholesterol and strychnine are shown as yellow and purple sticks, respectively. o-r, Structural comparison between TAS2R14–Gi1 and class A GPCR β2AR–Gs (PDB 3SN6) to highlight the conserved motifs in class A GPCRs which are CWxP (o), PIF (p), DRY (q), and NPxxY (r) motif. A large movement of TMs of TAS2R14–Gi1 is shown in the black arrows. A yellow dashed line indicates an interaction between H2767.49 of TAS2R14–Gi1 and cmpd28.1 (r). s,t Interactions formed between the residues N241.50 and S2777.50 (s) as well as W983.41 and P1905.50 (t).

Extended Data Fig. 4 Sequence alignment of TAS2R family and Gα proteins.

a, The residues of TAS2R14 mentioned throughout the text are labeled as generic residue numbering. The black squares, stars, and circles indicate the residues involved in the interactions with cholesterol, cmpd28.1, and the α5 helices of Gαi1 and Gαgust. b, Sequence alignment of α5 helix of Gα subunits. Labeled residues interact with TAS2R14. The sequence was aligned using GPCR generic residue numbers based on GPCRdb (https://gpcrdb.org) (a,b).

Extended Data Fig. 5 Computational analyses for TAS2R14.

a,b, EMERALD docking based on our cryo-EM map and model. a, The top-scoring pose for the cholesterol can recapitulate the modeled binding pose off of the map. b, The cmpd28.1 binding pose can also be recapitulated with EMERALD and the top scoring pose. c,d, Docking of bile acids to our TAS2R14 structure. The following bile acids were docked using Glide (Maestro v13.3): Cholate, Chendeoxycholate, Deoxycholate, Glycocholate, Lithocholate, Taurocholate, Taurolithocholate, and Ursodeoxycholate to orthosteric pocket occupied by cholesterol. The top score was taken from each docking run and visualized in ChimeraX. e-g, PCA of Simulations Reveal Conformational Change. The first two principal components for each simulation of the receptor alone state with CLR and cmpd28.1 (e), just CLR (f), or just cmpd28.1 (g). Circled is the conformational change seen in TM5/TM6 when traversing PC1. TM6 swings much further out only in the presence of CLR and cmpd28.1, not in the other simulations.

Extended Data Fig. 6 Activation of TAS2R14 by cholesterol and cmpd28.1.

a,b, Basal activities of TAS2R14 coupled with Ggust (a) and Gi1 (b) measured by BRET2 assays. Data represent mean ± s.e.m. of n = 3 biological replicates. The activation levels according to the increase of relative amounts of transfection of TAS2R14 were normalized by BRET ratio from the pcDNA3.1-transfected conditions. c-g, Split luciferase biosensor cAMP inhibition assay upon cholesterol depletion and replenishment. The values represent the mean ± s.e.m. of n = 3–6 biologically independent experiments. Significant differences were determined by ordinary one-way ANOVA with Dunnett’s test for multiple comparison analysis with reference to the pcDNA3.1 (c,d) and WT (e-g). h-j, BRET2 assays for TAS2R14 mutations at the orthosteric binding pocket in the absence/presence of cmpd28.1. Data represent mean ± s.e.m. of n = 3 biological replicates. h,i, All values are subtracted by net BRET of pcDNA3.1 in each of the six cellular conditions, separately. j, The net BRET values were subtracted by the net BRET of WT in the absence of cmpd28.1. Statistical differences were determined by ordinary two-way ANOVA with Dunnett’s test for multiple comparison analysis with reference to each of WT with either 0 μM or 10 μM cmpd28.1, separately. k-n, Agonist-stimulated TAS2R14 (k) and TAS2R46 (m) activation upon conditions of cholesterol depletion and replenishment. Representative column graphs for TAS2R14 (l) and TAS2R46 (n) adopted from the values at 100 μM agonists in a dose-response curve for TAS2R14 (k) and TAS2R46 (m), respectively. Data represent mean ± s.e.m. of n = 6 (k, l) and 3 (m, n). Statistical differences were determined by ordinary one-way ANOVA with Dunnett’s test for multiple comparison analysis with reference to the no-treatment condition (l, n). o-q, BRET2 assays with a dose-response of cmpd28.1 under the conditions of no treatment (o), depletion (p), and replenishment (q). All the significant differences were indicated by ****P < 0.0001, ***P < 0.001, **P < 0.01, and *P < 0.05, and ns (not significantly different) (c-g, j,l,n).

Source Data

Extended Data Fig. 7 TAS2R14 activation by intracellular cmpd28.1.

a, Close-up view of structures of cmpd28.1 in the allosteric site by aligning between TAS2R14–Gi1 and TAS2R14–Ggust. The ligands and interacting residues are shown as sticks. Hydrogen bonds are indicated as yellow dashed lines. b, Close-up view of the intracellular allosteric site of structures aligned between TAS2R14-Gi1 and TAS2R46-mGs/gust (PDB 7XP6). The labeling of the conserved residues of TAS2R46 compared to TAS2R14 was omitted for clarity. c, d, Structural superposition of TAS2R14-Gi1 with either NTSR1-Go (c) or PTH1R-Gs (d). The ligands bound to the intracellular allosteric sites were shown as sticks. e-g, Representative agonist-stimulated BRET2 assays upon no treatment (e), cholesterol depletion (f), and replenishment (g) using mutations at the allosteric site. h-j, Split luciferase biosensor cAMP inhibition assays for mutations of TAS2R14 at the allosteric site with a dose-response of cmpd28.1 (h), flufenamic acid (i), and aristolochic acid I (j). k-n, Substitution of L353 of Gα which interacts with cmpd28.1 with Alanine affects the coupling of TAS2R14 to Gαi1 (k) and Gαgust (l) as well as NTSR1 to Gαi1 (m) and Gαgust (n). Data represent mean ± s.e.m. of n = 3 biological replicates (e-n).

Source Data

Extended Data Fig. 8 MD Simulations confirm cholesterol and cmpd28.1 pockets.

a, Overall simulation results for cmpd28.1 and cholesterol (CLR). Representative average RMSD with the top and bottom part of the box plot showing the 75th percentile and 25th percentile, respectively. The simulations were independently repeated three times. b-f, Average RMSD plot by frame shown in solid blue (cmpd28.1) and teal (CLR). The grey shaded area is the calculated standard deviation across each frame across each simulation. Cartoon representations are shown on the right-hand side of each panel of selected frames from the MD simulation. Each one has either circled cmpd28.1 (blue) or CLR in (teal). Particularly focus on panels (d,e) which show cmpd28.1 occupying other sites than the intracellular pocket of the receptor. TAS2R14 are labeled as red stars.

Extended Data Fig. 9 Models for representative cavity between the orthosteric pocket and allosteric site in TAS2R14.

a, Representative orthosteric and allosteric site of TAS2R14. The pockets are occupied by cholesterol and cmpd28.1 shown as spheres. b, A model for electrostatic potential surface showing the orthosteric binding pocket engaged by cholesterol of TAS2R14. Red, negative (−5 kT e−1); blue, positive (+5 kT e−1). c, Extensive aromatic residues which are abundant in TM5 and TM6 form a hydrophobic network along with the cavity. The network interacts with cholesterol and cmpd28.1. d, A representative model for TAS2R14-Gi1 with the deep and elongated cavity which was predicted using Caver 3.0.3 software63. The cavity was colored orange. Four predicted tunnels for the entrance of cholesterol are indicated by black squares (e-h). e-h, Close-up views of the predicted entrance from d. Surrounding residues are shown as sticks. i-k, Cavities predicted by Caver 3.0.363 in SMO receptor (PDB 6O3C) (i), GPBAR (PDB 7CFN) (j), and TAS2R46 (PDB 7XP6) (k). The calculated cavities were colored green (i-k).

Extended Data Table 1 Cryo-EM data collection, refinement, and validation statistics

Supplementary information

Supplementary Information

Supplementary Figs 1 and 2 and Supplementary Methods.

Reporting Summary

Source data

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

Kim, Y., Gumpper, R.H., Liu, Y. et al. Bitter taste receptor activation by cholesterol and an intracellular tastant. Nature 628, 664–671 (2024). https://doi.org/10.1038/s41586-024-07253-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-024-07253-y

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