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:

A µ-opioid receptor modulator that works cooperatively with naloxone

Abstract

The µ-opioid receptor (µOR) is a well-established target for analgesia1, yet conventional opioid receptor agonists cause serious adverse effects, notably addiction and respiratory depression. These factors have contributed to the current opioid overdose epidemic driven by fentanyl2, a highly potent synthetic opioid. µOR negative allosteric modulators (NAMs) may serve as useful tools in preventing opioid overdose deaths, but promising chemical scaffolds remain elusive. Here we screened a large DNA-encoded chemical library against inactive µOR, counter-screening with active, G-protein and agonist-bound receptor to ‘steer’ hits towards conformationally selective modulators. We discovered a NAM compound with high and selective enrichment to inactive µOR that enhances the affinity of the key opioid overdose reversal molecule, naloxone. The NAM works cooperatively with naloxone to potently block opioid agonist signalling. Using cryogenic electron microscopy, we demonstrate that the NAM accomplishes this effect by binding a site on the extracellular vestibule in direct contact with naloxone while stabilizing a distinct inactive conformation of the extracellular portions of the second and seventh transmembrane helices. The NAM alters orthosteric ligand kinetics in therapeutically desirable ways and works cooperatively with low doses of naloxone to effectively inhibit various morphine-induced and fentanyl-induced behavioural effects in vivo while minimizing withdrawal behaviours. Our results provide detailed structural insights into the mechanism of negative allosteric modulation of the µOR and demonstrate how this can be exploited in vivo.

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: DEL screen for new µOR allosteric modulators.
Fig. 2: 368 inhibits turnover together with naloxone.
Fig. 3: Structural mechanism of 368 NAM activity.
Fig. 4: 368 potentiates naloxone activity in vivo.

Similar content being viewed by others

Data availability

All data supporting the findings of this study are available within the article, extended data, Supplementary Information or the associated data tables. Original raw data will be provided upon request for all experiments, including supporting information. The cryo-EM density map has been deposited into the Electron Microscopy Data Bank under accession code EMD-44635. Model coordinates have been deposited into the PDB under accession number 9BJKSource data are provided with this paper.

Code availability

This manuscript does not report any new code.

References

  1. Jamison, R. N. & Mao, J. Opioid analgesics. Mayo Clin. Proc. 90, 957–968 (2015).

    Article  CAS  PubMed  Google Scholar 

  2. O’Donnell, J. K., Halpin, J., Mattson, C. L., Goldberger, B. A. & Gladden, R. M. Deaths involving fentanyl, fentanyl analogs, and U-47700—10 states, July–December 2016. MMWR Morb. Mortal. Wkly Rep. 66, 1197–1202 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Centers for Disease Control and Prevention. Vital signs: overdoses of prescription opioid pain relievers—United States, 1999–2008. MMWR Morb. Mortal. Wkly Rep. 60, 1487–1492 (2011).

    Google Scholar 

  4. Rudd, R. A. et al. Increases in heroin overdose deaths—28 states, 2010 to 2012. MMWR Morb. Mortal. Wkly Rep. 63, 849–854 (2014).

    PubMed  PubMed Central  Google Scholar 

  5. National Academies of Sciences, Engineering, and Medicine. Pain Management and the Opioid Epidemic: Balancing Societal and Individual Benefits and Risks of Prescription Opioid Use (National Academies Press, 2017).

  6. Abdelal, R. et al. Real-world study of multiple naloxone administration for opioid overdose reversal among bystanders. Harm. Reduct. J. 19, 49 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Thompson, G. L. et al. Biased agonism of endogenous opioid peptides at the μ-opioid receptor. Mol. Pharmacol. 88, 335–346 (2015).

    Article  CAS  PubMed  Google Scholar 

  8. Benarroch, E. E. Endogenous opioid systems: current concepts and clinical correlations. Neurology 79, 807–814 (2012).

    Article  PubMed  Google Scholar 

  9. De Neve, J. et al. Comprehensive overview of biased pharmacology at the opioid receptors: biased ligands and bias factors. RSC Med. Chem. 12, 828–870 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  10. Wang, Y. et al. Structures of the entire human opioid receptor family. Cell https://doi.org/10.1016/j.cell.2022.12.026 (2023).

  11. Rasmussen, K., White, D. A. & Acri, J. B. NIDA’s medication development priorities in response to the opioid crisis: ten most wanted. Neuropsychopharmacology 44, 657–659 (2019).

    Article  PubMed  Google Scholar 

  12. Faouzi, A. et al. Structure-based design of bitopic ligands for the µ-opioid receptor. Nature 613, 767–774 (2023).

    Article  ADS  CAS  PubMed  Google Scholar 

  13. Bueno, A. B. et al. Structural insights into probe-dependent positive allosterism of the GLP-1 receptor. Nat. Chem. Biol. 16, 1105–1110 (2020).

    Article  CAS  PubMed  Google Scholar 

  14. Vaysse, P. J.-J., Gardner, E. L. & Zukin, S. Modulation of rat brain opioid receptors by cannabinoids. J. Pharmacol. Exp. Ther. 241, 534–539 (1987).

    CAS  PubMed  Google Scholar 

  15. Rothman, R. B. et al. Salvinorin A: allosteric interactions at the μ-opioid receptor. J. Pharmacol. Exp. Ther. 320, 801–810 (2007).

    Article  CAS  PubMed  Google Scholar 

  16. Franzini, R. M. & Randolph, C. Chemical space of DNA-encoded libraries: miniperspective. J. Med. Chem. 59, 6629–6644 (2016).

    Article  CAS  PubMed  Google Scholar 

  17. Halford, B. How DNA-encoded libraries are revolutionizing drug discovery. Chemical & Engineering News (19 June 2017).

  18. Gironda-Martínez, A., Donckele, E. J., Samain, F. & Neri, D. DNA-encoded chemical libraries: a comprehensive review with succesful stories and future challenges. ACS Pharmacol. Transl. Sci. 4, 1265–1279 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  19. Ahn, S. et al. Small-molecule positive allosteric modulators of the β2-adrenoceptor isolated from DNA-encoded libraries. Mol. Pharmacol. 94, 850–861 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  21. Liu, X. et al. Mechanism of intracellular allosteric β2AR antagonist revealed by X-ray crystal structure. Nature 548, 480–484 (2017).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  22. Liu, X. et al. Mechanism of β2 AR regulation by an intracellular positive allosteric modulator. Science 364, 1283–1287 (2019).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  24. Hilger, D. et al. Structural insights into differences in G protein activation by family A and family B GPCRs. Science 369, eaba3373 (2020).

    Article  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. Livingston, K. E. & Traynor, J. R. Allostery at opioid receptors: modulation with small molecule ligands. Br. J. Pharmacol. 175, 2846–2856 (2018).

    Article  CAS  PubMed  Google Scholar 

  27. Che, T. et al. Structure of the nanobody-stabilized active state of the kappa opioid receptor. Cell 172, 55–67.e15 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Che, T. et al. Nanobody-enabled monitoring of kappa opioid receptor states. Nat. Commun. 11, 1145 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Robertson, M. J. et al. Structure determination of inactive-state GPCRs with a universal nanobody. Nat. Struct. Mol. Biol. https://doi.org/10.1038/s41594-022-00859-8 (2022).

  30. Manglik, A. et al. Crystal structure of the μ-opioid receptor bound to a morphinan antagonist. Nature 485, 321–326 (2012).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  33. Váradi, A. et al. Mitragynine/corynantheidine pseudoindoxyls as opioid analgesics with mu agonism and delta antagonism, which do not recruit β-arrestin-2. J. Med. Chem. 59, 8381–8397 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  34. Smits, S. E. Antagonism by naloxone of morphine-induced single-dose dependence and antinociception in mice. Res. Commun. Chem. Pathol. Pharmacol. 15, 689–696 (1976).

    CAS  PubMed  Google Scholar 

  35. Bloom, A. S., Dewey, W. L., Harris, L. S. & Brosius, K. K. The correlation between antinociceptive activity of narcotics and their antagonists as measured in the mouse tail-flick test and increased synthesis of brain catecholamines. J. Pharmacol. Exp. Ther. 198, 33–41 (1976).

    CAS  PubMed  Google Scholar 

  36. Székely, J. I., Dunai-Kovács, Z., Miglécz, E., Rónai, A. Z. & Bajusz, S. In vivo antagonism by naloxone of morphine, beta-endorphin and a synthetic enkephalin analog. J. Pharmacol. Exp. Ther. 207, 878–883 (1978).

    PubMed  Google Scholar 

  37. Chakraborty, S. et al. Oxidative metabolism as a modulator of kratom’s biological actions. J. Med. Chem. 64, 16553–16572 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Wilson, L. L. et al. Lyophilized kratom tea as a therapeutic option for opioid dependence. Drug Alcohol Depend. 216, 108310–108318 (2020).

    Article  CAS  PubMed  Google Scholar 

  39. Wilson, L. L. et al. Kratom alkaloids, natural and semi-synthetic, show less physical dependence and ameliorate opioid withdrawal. Cell. Mol. Neurobiol. 41, 1131–1143 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Kandasamy, R. et al. Positive allosteric modulation of the mu-opioid receptor produces analgesia with reduced side effects. Proc. Natl Acad. Sci. USA 118, e2000017118 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Draper-Joyce, C. J. et al. Positive allosteric mechanisms of adenosine A1 receptor-mediated analgesia. Nature 597, 571–576 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  42. Slosky, L. M. et al. β-arrestin-biased allosteric modulator of NTSR1 selectively attenuates addictive behaviors. Cell 181, 1364–1379.e14 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Persechino, M., Hedderich, J. B., Kolb, P. & Hilger, D. Allosteric modulation of GPCRs: from structural insights to in silico drug discovery. Pharmacol. Ther. 237, 108242 (2022).

    Article  CAS  PubMed  Google Scholar 

  44. 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 

  45. Cong, Z. et al. Molecular insights into ago-allosteric modulation of the human glucagon-like peptide-1 receptor. Nat. Commun. 12, 3763 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  46. Xu, J. et al. Structural and dynamic insights into supra-physiological activation and allosteric modulation of a muscarinic acetylcholine receptor. Nat. Commun. 14, 376 (2023).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  47. Hedderich, J. B. et al. The pocketome of G-protein-coupled receptors reveals previously untargeted allosteric sites. Nat. Commun. 13, 2567 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kumar, K. K. et al. Structure of a signaling cannabinoid receptor 1–G protein complex. Cell https://doi.org/10.1016/j.cell.2018.11.040 (2019).

  49. Wang, H. et al. Structure-based evolution of G protein-biased μ-opioid receptor agonists. Angew. Chem. Int. Ed. https://doi.org/10.1002/anie.202200269 (2022).

  50. Zivanov, J. et al. New tools for automated high-resolution cryo-EM structure determination in RELION-3. eLife 7, e42166 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  51. 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 

  52. Rohou, A. & Grigorieff, N. CTFFIND4: fast and accurate defocus estimation from electron micrographs. J. Struct. Biol. 192, 216–221 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  53. 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 

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

    Article  ADS  PubMed  Google Scholar 

  55. Moriarty, N. W., Grosse-Kunstleve, R. W. & Adams, P. D. Electronic ligand builder and optimization workbench (eLBOW): a tool for ligand coordinate and restraint generation. Acta Crystallogr. D Biol. Crystallogr. 65, 1074–1080 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  56. Adams, P. D. et al. The Phenix software for automated determination of macromolecular structures. Methods 55, 94–106 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. 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  PubMed  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).

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

  60. Jo, S., Lim, J. B., Klauda, J. B. & Im, W. CHARMM-GUI Membrane Builder for mixed bilayers and its application to yeast membranes. Biophys. J. 97, 50–58 (2009).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  61. Jo, S., Kim, T. & Im, W. Automated builder and database of protein/membrane complexes for molecular dynamics simulations. PLoS ONE 2, e880 (2007).

  62. Lee, J. et al. CHARMM-GUI Membrane Builder for complex biological membrane simulations with glycolipids and lipoglycans. J. Chem. Theory Comput. 15, 775–786 (2019).

  63. Yu, W., He, X., Vanommeslaeghe, K. & Mackerell, A. D. Jr. Extension of the CHARMM general force field to sulfonyl-containing compounds and its utility in biomolecular simulations. J. Comput. Chem. 33, 2451–2468 (2012).

  64. Vanommeslaeghe, K. et al. CHARMM general force field: a force field for drug-like molecules compatible with the CHARMM all-atom additive biological force fields. J. Comput. Chem. 31, 671–690 (2009).

  65. 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 

  66. Jorgensen, W. L., Chandrasekhar, J., Madura, J. D., Impey, R. W. & Klein, M. L. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys. 79, 926–935 (1983).

    Article  ADS  CAS  Google Scholar 

  67. Eastman, P. et al. OpenMM 7: rapid development of high performance algorithms for molecular dynamics. PLoS Comput. Biol. 13, e1005659 (2017).

  68. 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  PubMed  Google Scholar 

  69. Ryckaert, J.-P., Ciccotti, G. & Berendsen, H. J. C. 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 

  70. 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).

  71. Gowers, R. J. et al. MDAnalysis: a Python package for the rapid analysis of molecular dynamics simulations. In Proc. 15th Python in Science Conference (eds Benthall, S. & Rostrup, S.) 98–105 (SciPy, 2016).

  72. Kilkenny, C., Browne, W. J., Cuthill, I. C., Emerson, M. & Altman, D. G. Improving bioscience research reporting: the ARRIVE guidelines for reporting animal research. Animals 4, 35–44 (2014).

    Article  PubMed Central  Google Scholar 

  73. Chakraborty, S. et al. A novel mitragynine analog with low-efficacy mu opioid receptor agonism displays antinociception with attenuated adverse effects. J. Med. Chem. 64, 13873–13892 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Wilson, L. L. et al. Characterization of CM-398, a novel selective sigma-2 receptor ligand, as a potential therapeutic for neuropathic pain. Molecules 27, 3617 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Uprety, R. et al. Controlling opioid receptor functional selectivity by targeting distinct subpockets of the orthosteric site. eLife 10, e56519 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Reilley, K. J. et al. Identification of two novel, potent, low-liability antinociceptive compounds from the direct in vivo screening of a large mixture-based combinatorial library. AAPS J. 12, 318–329 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Brice-Tutt, A. C. et al. Multifunctional opioid receptor agonism and antagonism by a novel macrocyclic tetrapeptide prevents reinstatement of morphine-seeking behaviour. Br. J. Pharmacol. 177, 4209–4222 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. 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 

  79. 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 

Download references

Acknowledgements

We thank staff at WuXi Apptec for providing the DEL and off-DNA synthesis of hits; J. Su for providing structure–activity relationship information for chemical families; and M. Robertson for providing the µORκ and Nb6 vectors. Cryo-EM data were collected at S2C2. K.K.K. was supported by the American Diabetes Association (ADA) Postdoctoral Fellowship. E.S.O. was supported by the American Heart Association (AHA) Postdoctoral Fellowship. S.M. and J.P.M. are supported by RO1DA057790. B.K.K. was supported by the Chan Zuckerberg Biohub and by R01DA036246.

Author information

Authors and Affiliations

Authors

Contributions

E.S.O., V.A.R., S.M., J.P.M. and B.K.K. wrote the manuscript with input from all authors. E.S.O., W.H., K.K. and B.K.K. designed the DEL screening strategy. E.S.O. performed DEL selections. E.W. and E.S.O. designed, optimized and performed the radioligand-binding experiments. E.S.O., W.H. and Y.S. optimized and performed the GTP turnover assays. J.M.M. performed the in-cell cAMP experiments. E.S.O. formed the complexes for the cryo-EM studies, collected and processed cryo-EM data with assistance from H.W. and C.Z., built the structural models with assistance from K.K., and performed and analysed the molecular dynamics simulations. V.A.R. synthesized (±)-368, its pure enantiomers, developed a scale-up synthesis of (±)-368 and chemical characterization under S.M.’s supervision. K.A. carried out pharmacokinetics analysis of (±)-368 under S.M.’s supervision. A.E. and Q.J. performed TRUPATH experiments with supervision from T.C. S.O.E., H.R.H. and J.P.M. performed and analysed all behavioural pharmacology experiments. B.J.K. developed and performed the chiral analysis.

Corresponding authors

Correspondence to Jay P. McLaughlin, Susruta Majumdar or Brian K. Kobilka.

Ethics declarations

Competing interests

B.K.K. is a founder and consultant for ConfometRx. S.M. is a founder of Sparian Biosciences. E.S.O., K.K.K., V.A.R., S.M. and B.K.K. have filed a patent around the new NAM compound acting through µOR. B.J.K. is an employee of Lotus Separations. The remaining authors declare no other competing interests.

Peer review

Peer review information

Nature thanks Catherine Cahill 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 Initial biochemical characterization of µOR allosteric modulators from DEL screen.

(a) Chemical properties of 368. MW; molecular weight (Da). logP; predicted octanol/water partition coefficient. PSA; polar surface area (Å2). ROTN; rotatable bonds. HBD; hydrogen bond donors. HBA; hydrogen bond acceptors. RCount; number of rings. ARCount; number of aromatics. CNSMPO; central nervous system multiparameter optimization. LogP was predicted using QikProp in Schrödinger, and other properties were calculated using ChemDraw. (b) Excess concentrations of 368 have no impact (P = 0.084) on Gi1 intrinsic turnover in the absence of receptor. Data are displayed as the average ± s.d. with n = 4 individual experiments. (c) We show using a direct 3H-naloxone binding experiment that increasing concentrations of 368 result in an increased antagonist affinity for µOR-containing membranes. Fitted affinity values are shown along with 95% confidence intervals in parentheses. Data are displayed as average values with error bars corresponding to the standard deviations of n = 4 measurements (d) The GTP turnover assay was used to show that 20 µM 368 inhibits turnover for a wide variety of orthosteric site conditions, ranging from slight inhibition of basal signaling (P = 0.0115) to no detectable effect on naloxone turnover (P = 0.314) to substantial inhibition of moderate partial agonist turnover (mitragynine pseudoindoxyl, MP, P = 0.005) to peptide (DAMGO, P < 0.0001) or small molecule (BU72, P < 0.0001) full agonists (all orthosteric molecules also present at 20 µM, data are displayed as the average ± s.d. with n = 4 individual experiments). P values for all of the above were calculated using an unpaired t-test (two tailed) and are denoted as follows: ns (P > 0.05), * (P ≤ 0.05), ** (P ≤ 0.01), *** (P ≤ 0.001), and **** (P ≤ 0.0001). Titrations of (e) morphine, (f) fentanyl, and (g) met-enkephalin result in activation of an assortment of G-proteins (Gi1, pink; Gi2, orange; Gi3, pale green; GoA, green; GoB, blue; and Gz, purple) as observed in the TRUPATH assay by a change in BRET signal as the Gα and Gβγ subunits separate. Activation by all 3 agonists was also calculated in the presence of 90 µM 368 (bottom panels). The data are displayed as the average the average ± s.d. with n = 6 individual experiments. The average log(EC50) for all 6 G-protein activation curves within each panel is shown as a black line (with dashed grey lines representing the standard deviation among different G protein subtypes). The average fold change in EC50 upon addition of 368 for morphine (14.5), fentanyl (16.2) and met-enkephalin (15.8) is shown. (h) The presence of excess (90 µM) 368 results in decreased agonist potencies for a variety of orthosteric agonists (fentanyl, circles; met-enkephalin, squares; morphine, triangles) across a series of Gi/o family G-protein effectors as observed by the TRUPATH assay for G-protein activation. The calculated log(EC50) for all agonist/G-protein combinations are right-shifted by ~1-1.5 units in the presence of 368. Data are displayed as average changes in log(EC50 values) for each condition with error bars corresponding to the additive fitted error 95% confidence intervals for EC50 values with and without 368. (i) This Gi/o family inhibition results in dampened cAMP inhibition in cells with morphine (dark blue vs. pink) and met-enkephalin (green vs. purple). Titration of 368 at EC80 concentrations of orthosteric agonists results in the reversal of cAMP inhibition, though with weaker potencies than those observed biochemically. Data are displayed as the average the average ± s.d. with n = 5 individual experiments.

Source Data

Extended Data Fig. 2 CryoEM structure determination of µORκ-Nb6 bound to naloxone and 368.

(a) Schematic of µORκ purification and complex formation with Nb6. (b) Cryo-EM data collection and processing pipeline, showing representative micrographs of the µORκ-Nb6 complex, reference-free 2D cryo-EM class averages, and processing flow chart. This includes motion and CTF correction in Relion, followed by particle selection, 2D and 3D classifications, density map reconstructions, “gold standard” FSC curves in Cryosparc, and the final density map colored by local resolution. Also included is the cryo-EM density map and model for the seven transmembrane helices of the µORκ.

Extended Data Fig. 3 Stereochemistry & MD analysis of 368.

(a) Comparison of the chemical structures and raw sequencing counts in the DEL selections for a series of molecules in the enriched 368 family. S-stereoisomers are substantially more enriched than their R-counterparts. (b) Chiral chromatography trace demonstrating separation of racemic 368, conditions which were then used to calculate enantiomeric excess in (R)-368 (c) and (S)-368 (d). Calculated peak parameters for (b-d) are shown below the respective traces. (e) Comparison of the individually synthesized stereoisomers of the racemic 368 “hit”, demonstrating that the R-368 is >100 fold weaker than the S-368 isomer, consistent with the DEL enrichment data. Data are displayed as the average value ± s.d. with n = 4 individual experiments. (f) GTP turnover assay again demonstrating that S-368 retains the ability to potently inhibit fentanyl (5 µM)-induced GTP turnover like the racemic 368 “hit”, while the R-368 isomer does not display full inhibition even at 20 µM. Data are displayed as the average ± s.d. with n = 6 individual experiments. (g) Accordingly, the S-isomer of 368 was modeled and placed into the cryo-EM density map (blue), along with an alternate, sub-optimal pose (green) that fits into overlapping but distinct areas in cryoEM density. (h) Both poses were then subjected to MD simulation for three independent 200 ns simulations (without Nb6) (green). An alternate pose from the “opposite” orientation of the NAM in the binding site that (sub-optimally) fits into the cryoEM density was also subjected to three independent 200 ns simulations (blue). The overall root mean square deviations (r.m.s.d.) throughout the trajectory were calculated for protein Cα and all atoms in 368 and naloxone for comparison. The time-dependent r.m.s.d. of 368 throughout all trajectories is displayed in (h) and the average of each of the runs with error bars representing the s.d. of three independent simulation averages is displayed in (i). (j) Example simulation snapshots were overlaid by Cα alignment for both poses at 4 ns increments. For our chosen pose, the conformation of 368 remains nearly constant throughout all three simulations, with r.m.s.d. values near that for the protein Cα (h-j). While Cα and naloxone remain stable in both poses (i), the alternate 368 pose is unstable (h-j), resulting in very high r.m.s.d. values and extensive conformational sampling in the orthosteric vestibule region (j).

Source Data

Extended Data Fig. 4 Mutational analysis of 368 binding site.

(a) Titration of (S)-368 against either wild type µOR (red), A323L µOR (purple) or I71W µOR (pink) expressing membranes demonstrates that both mutations substantially inhibit the ability of the NAM to enhance 3H-naloxone affinity. Data are displayed as the average ± s.d. with n = 4 individual experiments. (b) Expression levels of wild type human µOR (normalized to 100%) compared to mutations used in TRUPATH studies. Data are displayed as the average ± s.e.m. with n = 9 individual experiments. (c) Morphine dose-response curves observed by Gi1 recruitment to human µOR using the TRUPATH assay, comparing wild type receptor with a series of point mutants (A323L, I144E, H319L). The residue numbers correspond to those in the mouse µOR sequence used for the structural studies. Data are displayed as the average ± s.e.m. with n = 9 individual experiments. (d) Fentanyl dose-response curves observed by Gi1 recruitment to human µOR using the TRUPATH assay, again comparing wild type with point mutants. Data are displayed as the average ± s.e.m. with n = 9 individual experiments. Dose-response curves in c and d were fit using a three-parameter model for bottom, top, and logEC50 values. Log EC50 values for each curve, along with the 95% confidence interval range, are displayed for all curves.

Source Data

Extended Data Fig. 5 Probe dependence and opioid receptor selectivity of 368.

(a) Comparison of the extracellular vestibule regions of the previous β-FNA-bound µOR and the current naloxone-NAM-bound µOR. The presence of 368 sterically restricts the ability of small molecule antagonists to enter/exit the orthosteric site. (b) Alignment of the receptor regions of various ligand-bound µOR structures demonstrate that small molecule orthosteric compounds (top; naloxone [present work], lofentanil [PDB: 7T2H], MP [PDB:7T2G]) have little steric clash with 368 (calculated as the number of atoms in the orthosteric ligand within 1.5 Å of 368), while both peptide agonists (bottom; β-endorphin [PDB: 8F7Q], DAMGO [PDB: 6DDE]) have clear and substantial steric clash when overlaid with the 368 binding site. (c-e) TRUPATH assays for Gi1 activation for three different opioid receptor/ligand pairs; (c) µOR/DAMGO, (d) δOR/DPDPE, (e) κOR/U50488. Data are displayed as the average ± s.d. with n = 6 individual experiments. Fitted agonist EC50 values for each curve are shown, along with the error values which correspond to the 95% confidence interval for the fitted values. 368 has the largest impact on DAMGO activation of µOR and shows some activity against δOR activation by DPDPE, but has no effects on κOR activation by U50488. (f) Sequence alignment of human µOR, δOR and κOR at structural elements with interaction with 368 (residues within 6 Å denoted with an *). Numbering refers to positions within the human µOR. Red asterisks denote residues with important interactions with 368 but have side chains that are predicted to clash or no longer form productive interactions with 368.

Source Data

Extended Data Fig. 6 In vivo behavior of 368.

(a) Pharmacokinetics measurements of 368 at 10 mg/kg administered intravenously. 368 enters the brain with a maximum concentration of 1.66 uM, well above the observed affinity of the compound from radioligand binding (133 nM), but significantly lower than the concentration needed to inhibit agonist activity in cell assays. Data are displayed as the averages ± s.e.m. with n = 4 individuals. (b) In the absence of naloxone, 368 does not have significant effects on morphine-induced antinociception in the 55 °C warm-water tail-withdrawal assay (F(14,84) = 0.80, p = 0.67; Two-way RM ANOVA; n = 8 mice per group). This contrasts to the observed potentiated antagonism when in the presence of a low-dose (0.1 mg/kg) of naloxone (Fig. 4b). (c,d) The CLAMS assay with n = 8 mice demonstrates that in the absence of orthosteric compounds, 368 alone (10 mg/kg) has no significant impact on ambulation (F(6,108) = 0.36, p = 0.90; Two-way RM ANOVA) (c) and respiration (F(6,108) = 1.19, p = 0.32; Two-way RM ANOVA) (d). (e,f) CLAMS assay with n = 12 mice/group (n = 8 for morphine alone) demonstrating a dose- and time-dependent ability of 368 to potentiate low-dose (0.1 mg/kg) naloxone in reversing morphine-induced hyperlocomotion (F(24,306) = 4.27, p < 0.0001; Two-way RM ANOVA; e) and respiratory depression (F(24,294) = 5.00, p < 0.0001; Two-way RM ANOVA; f), with a maximal effect observed at 100 mg/kg 368 for both. (g) Antinociceptive time course experiment with n = 8 mice/group demonstrating that a 0.3 mg/kg dose of naloxone is able to partially reverse fentanyl (0.1 mg/kg)-induced antinociception on its own (factor: dose x time; Two-way RM ANOVA, F(7,98) = 3.883, p = 0.001; purple vs. orange curves), addition of 368 results in a significant enhancement of this naloxone-induced reversal (Two-way RM ANOVA,F(7,98) = 7.069, p = <0.0001; orange vs. dark red curves, n = 12 mice). Data for are displayed as the average percent antinociception ± s.e.m. of n = 8 or 12 mice as detailed above. (h) Antinociceptive time course experiment demonstrating that a 1.0 mg/kg dose of naloxone are not significantly able to reverse fentanyl (0.1 mg/kg)-induced antinociception (F(7,98) = 1.947, p = 0.070; purple vs. orange curves), addition of 368 results in a significant enhancement of this naloxone-induced reversal (F(7,98) = 7.077, p = <0.0001; orange vs. dark red curves). Data for are displayed as the average percent antinociception ± s.e.m. of n = 8 mice.

Source Data

Extended Data Table 1 Enrichment properties of selected 368 family members
Extended Data Table 2 Summary of TRUPATH potency values for Gi/o/z activation
Extended Data Table 3 Cryo-EM data collection, refinement and validation statistics
Extended Data Table 4 Summary of behavioral endpoints of naloxone-precipitated withdrawal in saline-treated or morphine-dependent mice following administration of either vehicle or 368 and naloxone (NAL)

Supplementary information

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

O’Brien, E.S., Rangari, V.A., El Daibani, A. et al. A µ-opioid receptor modulator that works cooperatively with naloxone. Nature (2024). https://doi.org/10.1038/s41586-024-07587-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41586-024-07587-7

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