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:

Conformational transitions of a neurotensin receptor 1–Gi1 complex

Abstract

Neurotensin receptor 1 (NTSR1) is a G-protein-coupled receptor (GPCR) that engages multiple subtypes of G protein, and is involved in the regulation of blood pressure, body temperature, weight and the response to pain. Here we present structures of human NTSR1 in complex with the agonist JMV449 and the heterotrimeric Gi1 protein, at a resolution of 3 Å. We identify two conformations: a canonical-state complex that is similar to recently reported GPCR–Gi/o complexes (in which the nucleotide-binding pocket adopts more flexible conformations that may facilitate nucleotide exchange), and a non-canonical state in which the G protein is rotated by about 45 degrees relative to the receptor and exhibits a more rigid nucleotide-binding pocket. In the non-canonical state, NTSR1 exhibits features of both active and inactive conformations, which suggests that the structure may represent an intermediate form along the activation pathway of G proteins. This structural information, complemented by molecular dynamics simulations and functional studies, provides insights into the complex process of G-protein activation.

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: Cryo-EM structures of hNTSR1–Gi1 complex.
Fig. 2: Structural comparison of hNTSR1 in C and NC states.
Fig. 3: Structural comparison of Gαi1 in C and NC states.
Fig. 4: Comparison of receptor–G-protein interfaces in C and NCstates.
Fig. 5: Interactions specifically observed in NC-state hNTSR1–Gi1 complex.
Fig. 6: Proposed model of hNTSR1 activation.

Similar content being viewed by others

Data availability

All data generated or analysed during this study are included in the published Article and Supplementary Information. The cryo-EM density maps for the hNTSR1–Gi1 complex in C and NC states have been deposited in the Electron Microscopy Data Bank (EMDB) under accession codes EMD-20180 and EMD-20181, respectively. The coordinates for the models of hNTSR1–Gi1 in both states have been deposited in the PDB under accession numbers 6OS9 and 6OSA, respectively. All other data are available upon request to the corresponding authors.

References

  1. Carraway, R. & Leeman, S. E. The isolation of a new hypotensive peptide, neurotensin, from bovine hypothalami. J. Biol. Chem. 248, 6854–6861 (1973).

    CAS  PubMed  Google Scholar 

  2. Vincent, J. P., Mazella, J. & Kitabgi, P. Neurotensin and neurotensin receptors. Trends Pharmacol. Sci. 20, 302–309 (1999).

    Article  CAS  Google Scholar 

  3. Boules, M., Li, Z., Smith, K., Fredrickson, P. & Richelson, E. Diverse roles of neurotensin agonists in the central nervous system. Front. Endocrinol. 4, 36 (2013).

    Article  CAS  Google Scholar 

  4. Wu, Z., Martinez-Fong, D., Trédaniel, J. & Forgez, P. Neurotensin and its high affinity receptor 1 as a potential pharmacological target in cancer therapy. Front. Endocrinol. 3, 184 (2013).

    Article  Google Scholar 

  5. Mustain, W. C., Rychahou, P. G. & Evers, B. M. The role of neurotensin in physiologic and pathologic processes. Curr. Opin. Endocrinol. Diabetes Obes. 18, 75–82 (2011).

    Article  CAS  Google Scholar 

  6. Schroeder, L. E. & Leinninger, G. M. Role of central neurotensin in regulating feeding: implications for the development and treatment of body weight disorders. Biochim. Biophys. Acta 1864, 900–916 (2018).

    Article  CAS  Google Scholar 

  7. Tanaka, K., Masu, M. & Nakanishi, S. Structure and functional expression of the cloned rat neurotensin receptor. Neuron 4, 847–854 (1990).

    Article  CAS  Google Scholar 

  8. Chalon, P. et al. Molecular cloning of a levocabastine-sensitive neurotensin binding site. FEBS Lett. 386, 91–94 (1996).

    Article  CAS  Google Scholar 

  9. Mazella, J. et al. Structure, functional expression, and cerebral localization of the levocabastine-sensitive neurotensin/neuromedin N receptor from mouse brain. J. Neurosci. 16, 5613–5620 (1996).

    Article  CAS  Google Scholar 

  10. Mazella, J. et al. The 100-kDa neurotensin receptor is gp95/sortilin, a non-G-protein-coupled receptor. J. Biol. Chem. 273, 26273–26276 (1998).

    Article  CAS  Google Scholar 

  11. Kitabgi, P. Targeting neurotensin receptors with agonists and antagonists for therapeutic purposes. Curr. Opin. Drug Discov. Devel. 5, 764–776 (2002).

    CAS  PubMed  Google Scholar 

  12. Besserer-Offroy, É. et al. The signaling signature of the neurotensin type 1 receptor with endogenous ligands. Eur. J. Pharmacol. 805, 1–13 (2017).

    Article  CAS  Google Scholar 

  13. White, J. F. et al. Structure of the agonist-bound neurotensin receptor. Nature 490, 508–513 (2012).

    Article  ADS  CAS  Google Scholar 

  14. Egloff, P. et al. Structure of signaling-competent neurotensin receptor 1 obtained by directed evolution in Escherichia coli. Proc. Natl Acad. Sci. USA 111, E655–E662 (2014).

    Article  CAS  Google Scholar 

  15. Krumm, B. E., White, J. F., Shah, P. & Grisshammer, R. Structural prerequisites for G-protein activation by the neurotensin receptor. Nat. Commun. 6, 7895 (2015).

    Article  ADS  CAS  Google Scholar 

  16. Krumm, B. E. et al. Structure and dynamics of a constitutively active neurotensin receptor. Sci. Rep. 6, 38564 (2016).

    Article  ADS  CAS  Google Scholar 

  17. Ballesteros, J. A. & Weinstein, H. in Receptor Molecular Biology, vol. 25 (ed. Sealfon, S. C.) Ch. 19 (Elsevier, 1995).

  18. Dubuc, I. et al. JMV 449: a pseudopeptide analogue of neurotensin-(8–13) with highly potent and long-lasting hypothermic and analgesic effects in the mouse. Eur. J. Pharmacol. 219, 327–329 (1992).

    Article  CAS  Google Scholar 

  19. Koehl, A. et al. Structure of the μ-opioid receptor–Gi protein complex. Nature 558, 547–552 (2018).

    Article  ADS  CAS  Google Scholar 

  20. Noble, A. J. et al. Reducing effects of particle adsorption to the air–water interface in cryo-EM. Nat. Methods 15, 793–795 (2018).

    Article  CAS  Google Scholar 

  21. Rasmussen, S. G. et al. Structure of a nanobody-stabilized active state of the β2 adrenoceptor. Nature 469, 175–180 (2011).

    Article  ADS  CAS  Google Scholar 

  22. Huang, W. et al. Structural insights into μ-opioid receptor activation. Nature 524, 315–321 (2015).

    Article  ADS  CAS  Google Scholar 

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

    Article  ADS  CAS  Google Scholar 

  24. Capper, M. J. & Wacker, D. How the ubiquitous GPCR receptor family selectively activates signalling pathways. Nature 558, 529–530 (2018).

    Article  ADS  CAS  Google Scholar 

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

    Article  ADS  CAS  Google Scholar 

  26. Latorraca, N. R., Venkatakrishnan, A. J. & Dror, R. O. GPCR dynamics: structures in motion. Chem. Rev. 117, 139–155 (2017).

    Article  CAS  Google Scholar 

  27. Draper-Joyce, C. J. et al. Structure of the adenosine-bound human adenosine A1 receptor–Gi complex. Nature 558, 559–563 (2018).

    Article  ADS  CAS  Google Scholar 

  28. García-Nafría, J., Nehmé, R., Edwards, P. C. & Tate, C. G. Cryo-EM structure of the serotonin 5-HT1B receptor coupled to heterotrimeric Go. Nature 558, 620–623 (2018).

    Article  ADS  Google Scholar 

  29. Wall, M. A. et al. The structure of the G protein heterotrimer Giα1β1γ2. Cell 83, 1047–1058 (1995).

    Article  CAS  Google Scholar 

  30. Sun, D. et al. Probing Gαi1 protein activation at single-amino acid resolution. Nat. Struct. Mol. Biol. 22, 686–694 (2015).

    Article  CAS  Google Scholar 

  31. Thomas, T. C., Schmidt, C. J. & Neer, E. J. G-protein alpha o subunit: mutation of conserved cysteines identifies a subunit contact surface and alters GDP affinity. Proc. Natl Acad. Sci. USA 90, 10295–10299 (1993).

    Article  ADS  CAS  Google Scholar 

  32. Iiri, T., Herzmark, P., Nakamoto, J. M., van Dop, C. & Bourne, H. R. Rapid GDP release from G in patients with gain and loss of endocrine function. Nature 371, 164–168 (1994).

    Article  ADS  CAS  Google Scholar 

  33. Posner, B. A., Mixon, M. B., Wall, M. A., Sprang, S. R. & Gilman, A. G. The A326S mutant of Giα1 as an approximation of the receptor-bound state. J. Biol. Chem. 273, 21752–21758 (1998).

    Article  CAS  Google Scholar 

  34. Grishina, G. & Berlot, C. H. A surface-exposed region of G in which substitutions decrease receptor-mediated activation and increase receptor affinity. Mol. Pharmacol. 57, 1081–1092 (2000).

    CAS  PubMed  Google Scholar 

  35. Hu, J. et al. Structural basis of G protein-coupled receptor–G protein interactions. Nat. Chem. Biol. 6, 541–548 (2010).

    Article  CAS  Google Scholar 

  36. Hillenbrand, M., Schori, C., Schöppe, J. & Plückthun, A. Comprehensive analysis of heterotrimeric G-protein complex diversity and their interactions with GPCRs in solution. Proc. Natl Acad. Sci. USA 112, E1181–E1190 (2015).

    Article  ADS  CAS  Google Scholar 

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

    Article  Google Scholar 

  38. Sounier, R. et al. Propagation of conformational changes during μ-opioid receptor activation. Nature 524, 375–378 (2015).

    Article  ADS  CAS  Google Scholar 

  39. Gregorio, G. G. et al. Single-molecule analysis of ligand efficacy in β2AR–G-protein activation. Nature 547, 68–73 (2017).

    Article  ADS  CAS  Google Scholar 

  40. Van Eps, N. et al. Gi- and Gs-coupled GPCRs show different modes of G-protein binding. Proc. Natl Acad. Sci. USA 115, 2383–2388 (2018).

    Article  Google Scholar 

  41. Dror, R. O. et al. Structural basis for nucleotide exchange in heterotrimeric G proteins. Science 348, 1361–1365 (2015).

    Article  ADS  CAS  Google Scholar 

  42. 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  Google Scholar 

  43. Zhang, K. Gctf: real-time CTF determination and correction. J. Struct. Biol. 193, 1–12 (2016).

    Article  ADS  CAS  Google Scholar 

  44. Scheres, S. H. Processing of structurally heterogeneous cryo-EM data in RELION. Methods Enzymol. 579, 125–157 (2016).

    Article  CAS  Google Scholar 

  45. Heymann, J. B. Single particle reconstruction and validation using Bsoft for the map challenge. J. Struct. Biol. 204, 90–95 (2018).

    Article  Google Scholar 

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

    Article  Google Scholar 

  47. 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 

  48. Williams, C. J. et al. MolProbity: more and better reference data for improved all-atom structure validation. Protein Sci. 27, 293–315 (2018).

    Article  CAS  Google Scholar 

  49. Pettersen, E. F. et al. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004).

    Article  CAS  Google Scholar 

  50. Goddard, T. D. et al. UCSF ChimeraX: meeting modern challenges in visualization and analysis. Protein Sci. 27, 14–25 (2018).

    Article  CAS  Google Scholar 

  51. Jacobson, M. P. et al. A hierarchical approach to all-atom protein loop prediction. Proteins 55, 351–367 (2004).

    Article  CAS  Google Scholar 

  52. Jacobson, M. P., Friesner, R. A., Xiang, Z. & Honig, B. On the role of the crystal environment in determining protein side-chain conformations. J. Mol. Biol. 320, 597–608 (2002).

    Article  CAS  Google Scholar 

  53. Eswar, N. et al. Comparative protein structure modeling using MODELLER. Curr. Protoc. Protein Sci. 50, 2.9.1–2.9.31 (2007).

    Article  Google Scholar 

  54. Ghanouni, P. et al. The effect of pH on β2 adrenoceptor function. evidence for protonation-dependent activation. J. Biol. Chem. 275, 3121–3127 (2000).

    Article  CAS  Google Scholar 

  55. Ranganathan, A., Dror, R. O. & Carlsson, J. Insights into the role of Asp792.50 in β2 adrenergic receptor activation from molecular dynamics simulations. Biochemistry 53, 7283–7296 (2014).

    Article  CAS  Google Scholar 

  56. Lomize, M. A., Lomize, A. L., Pogozheva, I. D. & Mosberg, H. I. OPM: orientations of proteins in membranes database. Bioinformatics 22, 623–625 (2006).

    Article  CAS  Google Scholar 

  57. Betz, R.M. Dabble. https://doi.org/10.5281/zenodo.836914 (2018).

  58. Vilardaga, J. P., Bünemann, M., Krasel, C., Castro, M. & Lohse, M. J. Measurement of the millisecond activation switch of G protein-coupled receptors in living cells. Nat. Biotechnol. 21, 807–812 (2003).

    Article  CAS  Google Scholar 

  59. Lewis, G. N. A new principle of equilibrium. Proc. Natl Acad. Sci. USA 11, 179–183 (1925).

    Article  ADS  CAS  Google Scholar 

  60. Astumian, R. D. Microscopic reversibility as the organizing principle of molecular machines. Nat. Nanotechnol. 7, 684–688 (2012).

    Article  ADS  CAS  Google Scholar 

  61. Hopkins, C. W., Le Grand, S., Walker, R. C. & Roitberg, A. E. Long-time-step molecular dynamics through hydrogen mass repartitioning. J. Chem. Theory Comput. 11, 1864–1874 (2015).

    Article  CAS  Google Scholar 

  62. Ryckaert, J.-P., Ciccotti, G. & Berendsen, H. J. Numerical integration of the Cartesian equations of motion of a system with constraints: molecular dynamics of n-alkanes. J. Comput. Phys. 23, 327–341 (1977).

    Article  ADS  CAS  Google Scholar 

  63. Huang, J. & MacKerell, A. D., Jr. CHARMM36 all-atom additive protein force field: validation based on comparison to NMR data. J. Comput. Chem. 34, 2135–2145 (2013).

    Article  CAS  Google Scholar 

  64. Klauda, J. B. et al. Update of the CHARMM all-atom additive force field for lipids: validation on six lipid types. J. Phys. Chem. B 114, 7830–7843 (2010).

    Article  CAS  Google Scholar 

  65. MacKerell, A. D. et al. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 102, 3586–3616 (1998).

    Article  CAS  Google Scholar 

  66. Best, R. B., Mittal, J., Feig, M. & MacKerell, A. D. Jr. Inclusion of many-body effects in the additive CHARMM protein CMAP potential results in enhanced cooperativity of α-helix and β-hairpin formation. Biophys. J. 103, 1045–1051 (2012).

    Article  ADS  CAS  Google Scholar 

  67. Best, R. B. et al. Optimization of the additive CHARMM all-atom protein force field targeting improved sampling of the backbone φ, ψ and side-chain χ1 and χ2 dihedral angles. J. Chem. Theory Comput. 8, 3257–3273 (2012).

    Article  CAS  Google Scholar 

  68. Salomon-Ferrer, R., Götz, A. W., Poole, D., Le Grand, S. & Walker, R. C. Routine microsecond molecular dynamics simulations with AMBER on GPUs. 2. Explicit solvent particle mesh ewald. J. Chem. Theory Comput. 9, 3878–3888 (2013).

    Article  CAS  Google Scholar 

  69. Pearlman, D. A. et al. AMBER, a package of computer programs for applying molecular mechanics, normal mode analysis, molecular dynamics and free energy calculations to simulate the structural and energetic properties of molecules. Comput. Phys. Commun. 91, 1–41 (1995).

    Article  ADS  CAS  Google Scholar 

  70. Humphrey, W., Dalke, A. & Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph14, 33–38 (1996).

    Article  CAS  Google Scholar 

  71. Coleman, D. E. & Sprang, S. R. Structure of Giα1•GppNHp, autoinhibition in a Gα protein–substrate complex. J. Biol. Chem. 274, 16669–16672 (1999).

    Article  CAS  Google Scholar 

  72. Grundmann, M. et al. Lack of beta-arrestin signaling in the absence of active G proteins. Nat. Commun. 9, 341 (2018).

    Article  ADS  Google Scholar 

  73. Robert, X. & Gouet, P. Deciphering key features in protein structures with the new ENDscript server. Nucleic Acids Res. 42, W320–W324 (2014)

    Article  CAS  Google Scholar 

  74. Hattori, M., Hibbs, R. E. & Gouaux, E. A fluorescence-detection size-exclusion chromatography-based thermostability assay for membrane protein precrystallization screening. Structure 20, 1293–1299 (2012).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank Y. S. Kim for assistance with HEK cell maintenance and transfection; B. White for assistance with Sf9 insect cell maintenance and mini-preparation of plasmids; S. Maeda for the P1 virus of scFv16; K. Geiselhart and M. Lima for administrative support of the project; and M. Masureel and S. Lavington for useful discussions on the manuscript. C.-M.S. acknowledges the Sigrid Jusélius Foundation and the International Human Frontier Science Program (Long-Term Fellowship LT000916-2018-L). R.F. was funded by grant NNF15OC0015268 from the Novo Nordisk Foundation and the Stanford Bio-X Program. This work was supported by National Institutes of Health (NIH) grants R01GM127359 (R.O.D.), R01GM083118 (B.K.K. and G.S.) and R01NS028471 (B.K.K.), the PRIME JP17gm5910013 (A.I.) and the LEAP JP17gm0010004 (A.I. and J.A.) from the Japan Agency for Medical Research and Development, JSPS KAKENHI 19H03163 (H.E.K.) and 17K08264 (A.I.), and the Mathers Foundation (G.S. and B.K.K.). B.K.K. is a Chan–Zuckerberg Biohub Investigator.

Peer review information

Nature thanks Shangyu Dang, Daniel Wacker and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

H.E.K. started the project, performed the molecular cloning, expressed and purified proteins, prepared the NTSR1–Gi1 complexes, refined the structure from cryo-EM density maps, performed FSEC-TS and GTPase-Glo assays, and analysed GPCR–G-protein interactions, with R.F. Y.Z. obtained cryo-EM images with the help of H.H. and processed cryo-EM data to generate 3D maps. C.-M.S. and N.R.L. performed and analysed the molecular dynamics simulations under the supervision of R.O.D. R.F. performed in silico analysis of GPCR–G-protein complexes. A.I. and F.M.N.K. performed nano-BiT G-protein dissociation assay under supervision of J.A. K.K.K. helped with G protein and scFv16 purification. D.H. provided critical input on structural analysis. W.H. contributed to the early stages of the project, including HEK cell transfection. H.E.K. prepared the initial manuscript and H.E.K., G.S. and B.K.K. wrote the paper with input from all the authors.

Corresponding authors

Correspondence to Brian K. Kobilka or Georgios Skiniotis.

Ethics declarations

Competing interests

B.K.K. is a founder of and consultant for ConfometRx, Inc.

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 Structure-based sequence alignment of class-A GPCRs.

The sequences shown are those are for hNTSR1, rNTSR1, mouse μOR, human cannabinoid receptor 1 (CB1), human rhodopsin, human 5-hydroxytryptamine receptor 1B (5HT1B), human A1 adenosine receptor (A1AR), human β2AR and human A2 adenosine receptor (A2AR). The sequence alignment was created using GPCRdb (http://www.gpcrdb.org) and ESPript 373 servers. Secondary structure elements for hNTSR1 are shown as coils and arrows. PIF, PAF, PLF or LVF, DRY or ERY, and NPXXY motifs are highlighted in green, red and blue, respectively. The truncated sequences of hNTSR1(∆ICL3) are highlighted in grey.

Extended Data Fig. 2 Preparation and cryo-EM of the full-length hNTSR1–Gi1–scFv16 and the hNTSR1(∆ICL3)–Gi1–scFv16 complexes.

a, Representative elution profile (out of more than three independent runs) of full-length hNTSR1 (hNTSR1(FL), comprising residues 20–418 and A85L mutation) in complex with Gi1 and scFv16, on a Superdex 200 Increase 10/300 GL. b, Representative 3D classifications of the hNTSR1(FL)–Gi1–scFv16 complex. The C-state and NC-state complex maps are coloured in cyan and red, respectively. The black arrow indicates the partially disordered α-helical domain. c, Representative cryo-EM micrograph of the hNTSR1(∆ICL3) in complex with Gi1 and scFv16. d, Representative 2D averages showing different views of the hNTSR1(∆ICL3)–Gi1–scFv16 complex. e, Flow chart of cryo-EM data processing. f, Local resolutions of C1 state (left) and NC1 state (right). Full view of the RELION local-resolution-filtered map coloured by local resolution. g, Representative 3D classifications of the hNTSR1(∆ICL3)–Gi1–scFv16 complex. The black arrow indicates the partially disordered α-helical domain. h, Gold-standard Fourier shell correlation plots. i, Fourier shell correlation curves for the final model versus the final map and the half maps.

Extended Data Fig. 3 Functional comparison between hNTSR1(FL) and hNTSR1(∆ICL3).

a, FSEC-TS74 for hNTSR1(FL)–Gi1 (left) and hNTSR1(∆ICL3)–Gi1 (right). Each profile is a representative of two independent experiments. Only about 50% of the hNTSR1(FL)–Gi1 complex survives after a 45 °C incubation for 10 min, whereas over 90% of the hNTSR1(FL)–Gi1 complex survives after the same heat stress. b, GTPase-Glo assay39 of hNTSR1(FL) and hNTSR1(∆ICL3). Sample sizes for both hNTSR1(FL) and hNTSR1(∆ICL3) are 3. The intrinsic GTP hydrolysis activities of Gi1 heterotrimer and Gq heterotrimer are enhanced by hNTSR1. The guanine-nucleotide exchange factor activities of hNTSR1(FL) and hNTSR1(∆ICL3) proteins are equally potent. Symbols and bars represent individual value and mean of a single experiment performed in triplicate. c, Cell-surface expression level. HEK293 cells transiently expressing a Flag-epitope-tagged NTSR1 construct were analysed by flow cytometry. Sample sizes are shown in parentheses. Centre lines and error bars represent mean and s.e.m. of the indicated experiments. One-way analysis of variance (ANOVA) with Dunnett’s post hoc test was used to assess statistical analyses (ANOVA P value = 0.90, not significantly different (NS) among the four samples). dg, NanoBiT G-protein dissociation assay. Concentration–response curves for G-protein dissociation signals (d, top) and their summary (d, bottom), for hNTSR1(FL) and hNTSR1(∆ICL3). Symbols and error bars represent mean and s.e.m. of indicated independent numbers of experiments, each performed in duplicate. e, Heat map of NanoBiT G-protein dissociation signals for hNTSR1(∆ICL3) (10 μM JMV449), β2AR (10 μM isoproterenol) and μOR (10 μM DAMGO). Mean values of test GPCR-specific signal-changes (differences in NanoBiT-G protein dissociation signal between test GPCR-transfected cells and mock-transfected cells) are shown. Sample sizes for Gs, Gi1, Go, Gq and G13 are as follows: 5, 5, 5, 5 and 5 (hNTSR1), 6, 5, 3, 3 and 3 (β2AR) and 7, 7, 6, 5 and 5 (μOR). Unlike β2AR and μOR, the NTSR1 agonist (JMV449) causes a signal decrease for all G-proteins (e), which suggest that all G proteins can be recognized and activated by hNTSR1, and dissociated into Gα and Gβγ subunits. f, The summary of NanoBiT G-protein dissociation assay of wild-type hNTSR1 and the hNTSR1(S93A/L94A/R294A/H373A) mutant for full-length constructs. Concentration–response curves are shown in Fig. 5b. We used an unpaired t-test with correction for multiple comparisons using the Holm–Sidak method. NS, not significantly different from wild type; **P < 0.01. g, NanoBiT G-protein dissociation assay of wild-type hNTSR1 and the hNTSR1(S93A/L94A/R294A/H373A) mutant for ∆ICL3 constructs. Concentration–response curves of Gs, Gi1, Go, Gq and G13 signalling (top), and the summary of the assay result (bottom). Symbols and error bars (top) represent mean and s.e.m. of indicated independent numbers of experiments (bottom), each performed in duplicate. We used an unpaired t-test with correction for multiple comparisons using the Holm–Sidak method. NS, not significantly different from wild type; *P < 0.05, **P < 0.01, ***P < 0.001.

Extended Data Fig. 4 Structural comparisons of micro-conformers observed in NC- and C-state hNTSR1(∆ICL3)–Gi1 complexes.

a, Side (top) and extracellular (bottom) views of the superposed structures of three conformers in the C state. b, Side (top) and extracellular (bottom view (bottom) of the superimposed structures of two conformers in the NC state. In each micro-conformer, the G protein is 4–5° rotated relative to the receptor.

Extended Data Fig. 5 Cryo-EM map quality.

a, b, Density and model for transmembrane helices of hNTSR1, α5 helix of Gαi1 and JMV449 in C-state (a) and NC-state (b) complexes. c, Putative cholesterol observed near TM6 and TM7, and the neighbouring side chains in the putative binding site of C-state (left) and NC-state (right) complexes.

Extended Data Fig. 6 Agonist-peptide binding to NTSR1.

ac, Agonist peptide and the neighbouring side chains in the ligand-binding site of active rNTSR1 (rNTSR1(ELF)) (a), C-state complex (b) and NC-state complex (c). Black dashed lines represent hydrogen bonds.

Extended Data Fig. 7 Structural comparison of NTSR1, β2AR and μOR.

a, b, Comparison of TM6, DRY motif and NPXXY motif between C-state hNTSR1 (blue), NC-state hNTSR1 (red), active β2AR (orange) (PDB 3SN6) and active μOR (purple) (PDB 6DDF). Black double-headed arrow represents the conformational difference of TM6 between the receptor in the Gs complex and the receptor in the Gi complex (a). Y7.53 in the NPXXY motif packs against R3.50 in C-state hNTSR1, active β2AR and active μOR, but there is no direct interaction between Y7.53 and R3.50 in NC-state hNTSR1 (b). c, Comparison of the cytoplasmic half of TM7 between C-state hNTSR1 (blue), NC-state hNTSR1 (red), inactive-state rNTSR1 (rNTSR1-inact, grey), and active-state rNTSR1 (rNTSR1-act, green). NC-state hNTSR1 is superimposes well onto rNTSR1(TM86V/ΔIC3A), which suggests that TM7 adopts an inactive-like conformation in NC-state hNTSR1.

Extended Data Fig. 8 Structural comparison of G proteins and GPCR–G-protein complexes.

a, Overall structures of Gαi from C-state hNTSR1–Gi1 (yellow), NC-state hNTSR1–Gi1 (grey), μOR–Gi1 (green), rhodopsin–Gi1 (purple) and A1AR–Gi2 (pink) complexes. The α-helical domain of the rhodopsin–Gi1 complex is removed for clarity. b, The π–π stacking interaction between the α5 helix and β6 strand, which is specifically observed in Gi complexes. c, d, Side view and extracellular view of the superimposed structures of C-state hNTSR1–Gi1 complex (blue, hNTSR1; yellow, Gi1) and β2AR-Gs complex (grey, β2AR; orange, Gs) (c), and C-state hNTSR1–Gi1 complex (blue, hNTSR1; yellow, Gi1) and μOR–Gi1 complex (green, μOR; grey, Gi1) (d).

Extended Data Fig. 9 Dynamics of the nucleotide-binding pocket in NC and C states.

a, Cryo-EM density, shown in two different contour levels, corresponding to the α5–β6 loop of Gαi1 from C-state (top) and NC-state (bottom) hNTSR1–Gi1 complex. b, Summary of molecular dynamics simulation conditions. c, Dynamics of the P-loop and switch II regions during molecular dynamics simulations of the C-state (left) and NC-state (right) complexes. The figures show superposed frames sampled every 50 ns from 5 independent simulations for each state. In these simulations, the P-loop and switch-II regions show similar flexibility in both the C-state and NC-state complexes. d, Representative molecular dynamics simulations initiated from the C-state hNTSR1–Gi1 complex (left) and the NC-state hNTSR1–Gi1 complex (right). The r.m.s.d. of the NPXXY motif relative to the NC-state structure (top) and the distance between TM3 and TM6 (bottom) are plotted for each simulation. In both C-state and NC-state hNTSR1–Gi1 complexes, the NPXXY region and TM6 consistently retain the conformations that are observed in the cryo-EM structures. e, The r.m.s.d. of α5 to the cryo-EM C-state Gi1 for each simulation of the NC-state and C-state complexes. The trajectories were aligned on TM1–TM4 of the receptor, and the r.m.s.d. was calculated for the backbone atoms of residues 329 to 354 of α5. f, Dynamics of the α5–β6 loop for each independent simulation of the C- and NC-state complexes. Frames are sampled every 20 ns from each individual simulation. In these simulations, the α5–β6 loop shows enhanced conformational variability in the C-state complex compared to the NC-state complex. g, The calculated solvent-accessible surface area (SASA) of the nucleotide-binding pocket in the NC and C states. The solvent accessibility of the pocket is consistently larger for the C state (P = 0.00027976943074583155, using Welch’s two-sided t-test and treating each simulation as an independent data point).

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

Supplementary information

Supplementary Information

Supplementary Discussion about a putative cholesterol binding site in hNTSR1, and information about the amino acid sequences of the constructs used in the NanoBiT G-protein dissociation assay.

Reporting Summary

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kato, H.E., Zhang, Y., Hu, H. et al. Conformational transitions of a neurotensin receptor 1–Gi1 complex. Nature 572, 80–85 (2019). https://doi.org/10.1038/s41586-019-1337-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-1337-6

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