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:

Structure-based design of a novel third-generation antipsychotic drug lead with potential antidepressant properties

Abstract

Partial agonist activity at the dopamine D2 receptor (DRD2) is a key feature of third-generation antipsychotics (TGAs). However, TGAs also act as antagonists or weak partial agonists to the serotonin (5-hydroxytryptamine; 5-HT) 2A receptor (5-HT2AR). Here we present the crystal structures of aripiprazole- and cariprazine-bound human 5-HT2AR. Both TGAs adopt an unexpected ‘upside-down’ pose in the 5-HT2AR binding pocket, with secondary pharmacophores inserted in a similar way to a ‘bolt’. This insight into the binding modes of TGAs offered a structural mechanism underlying their varied partial efficacies at 5-HT2AR and DRD2. These structures enabled the design of a partial agonist at DRD2/3 and 5-HT1AR with negligible 5-HT2AR binding that displayed potent antipsychotic-like activity without motor side effects in mice. This TGA lead also had antidepressant-like effects and improved cognitive performance in mouse models via 5-HT1AR. This work indicates that 5-HT2AR affinity is a dispensable contributor to the therapeutic actions of TGAs.

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 architecture of 5-HT2AR bound to aripiprazole.
Fig. 2: The exosite determines the efficacy of aripiprazole at DRD2 and 5-HT2AR.
Fig. 3: Comparison of exosite residues between 5-HT2AR and DRD2.
Fig. 4: Structure-guided design of selective DRD2 partial agonist.
Fig. 5: The antipsychotic-like efficacy of (–)-IHCH7041 in rodent animal models.
Fig. 6: (–)-IHCH7041 alleviates learning and memory impairments and cognitive disturbance induced by subchronic MK-801 treatment.

Similar content being viewed by others

Data availability

The structural files have been deposited on the RCSB PDB database with the PDB codes 7VOE and 7VOD. All related source data are publicly available.

References

  1. Perala, J. et al. Lifetime prevalence of psychotic and bipolar I disorders in a general population. Arch. Gen. Psychiatry 64, 19–28 (2007).

    Article  PubMed  Google Scholar 

  2. Creese, I., Burt, D. R. & Snyder, S. H. Dopamine receptor binding predicts clinical and pharmacological potencies of antischizophrenic drugs. Science 192, 481–483 (1976).

    Article  CAS  PubMed  Google Scholar 

  3. Roth, B. L., Sheffler, D. J. & Kroeze, W. K. Magic shotguns versus magic bullets: selectively non-selective drugs for mood disorders and schizophrenia. Nat. Rev. Drug Discov. 3, 353–359 (2004).

    Article  CAS  PubMed  Google Scholar 

  4. Meyer, J. M. et al. The Clinical Antipsychotic Trials Of Intervention Effectiveness (CATIE) Schizophrenia Trial: clinical comparison of subgroups with and without the metabolic syndrome. Schizophr. Res. 80, 9–18 (2005).

    Article  PubMed  Google Scholar 

  5. Keltner, N. L. & Johnson, V. Biological perspectives. Aripiprazole: a third generation of antipsychotics begins? Perspect. Psychiatr. Care 38, 157–159 (2002).

    Article  PubMed  Google Scholar 

  6. Leucht, S. et al. Comparative efficacy and tolerability of 15 antipsychotic drugs in schizophrenia: a multiple-treatments meta-analysis. Lancet 382, 951–962 (2013).

    Article  CAS  PubMed  Google Scholar 

  7. Kane, J. M. et al. Efficacy and safety of aripiprazole and haloperidol versus placebo in patients with schizophrenia and schizoaffective disorder. J. Clin. Psychiatry 63, 763–771 (2002).

    Article  CAS  PubMed  Google Scholar 

  8. Lieberman, J. A. Dopamine partial agonists: a new class of antipsychotic. CNS Drugs 18, 251–267 (2004).

    Article  CAS  PubMed  Google Scholar 

  9. Kern, R. S. et al. The neurocognitive effects of aripiprazole: an open-label comparison with olanzapine. Psychopharmacology (Berl.) 187, 312–320 (2006).

    Article  CAS  Google Scholar 

  10. Meltzer, H. Y. Clinical studies on the mechanism of action of clozapine: the dopamine-serotonin hypothesis of schizophrenia. Psychopharmacology (Berl.) 99 Suppl, S18–S27 (1989).

    Article  CAS  Google Scholar 

  11. Ebdrup, B. H., Rasmussen, H., Arnt, J. & Glenthoj, B. Serotonin 2A receptor antagonists for treatment of schizophrenia. Expert Opin. Investig. Drugs 20, 1211–1223 (2011).

    Article  CAS  PubMed  Google Scholar 

  12. Mamo, D. et al. Differential effects of aripiprazole on D2, 5-HT2, and 5-HT1A receptor occupancy in patients with schizophrenia: a triple tracer PET study. Am. J. Psychiatry 164, 1411–1417 (2007).

    Article  PubMed  Google Scholar 

  13. Natesan, S., Reckless, G. E., Barlow, K. B., Nobrega, J. N. & Kapur, S. Partial agonists in schizophrenia—why some work and others do not: insights from preclinical animal models. Int. J. Neuropsychopharmacol. 14, 1165–1178 (2011).

    Article  PubMed  Google Scholar 

  14. Kaar, S. J., Natesan, S., McCutcheon, R. & Howes, O. D. Antipsychotics: mechanisms underlying clinical response and side-effects and novel treatment approaches based on pathophysiology. Neuropharmacology 172, 107704 (2020).

    Article  CAS  PubMed  Google Scholar 

  15. Wang, S. et al. Structure of the D2 dopamine receptor bound to the atypical antipsychotic drug risperidone. Nature 555, 269–273 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Fan, L. et al. Haloperidol bound D2 dopamine receptor structure inspired the discovery of subtype selective ligands. Nat. Commun. 11, 1074 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Yin, J. et al. Structure of a D2 dopamine receptor–G-protein complex in a lipid membrane. Nature 584, 125–129 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Kimura, K. T. et al. Structures of the 5-HT2A receptor in complex with the antipsychotics risperidone and zotepine. Nat. Struct. Mol. Biol. 26, 121–128 (2019).

    Article  CAS  PubMed  Google Scholar 

  19. Im, D. et al. Structure of the dopamine D2 receptor in complex with the antipsychotic drug spiperone. Nat. Commun. 11, 6442 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Kim, K. et al. Structure of a hallucinogen-activated Gq-coupled 5-HT2A serotonin receptor. Cell 182, 1574–1588 e1519 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Shapiro, D. A. et al. Aripiprazole, a novel atypical antipsychotic drug with a unique and robust pharmacology. Neuropsychopharmacology 28, 1400–1411 (2003).

    Article  CAS  PubMed  Google Scholar 

  22. Kiss, B. et al. Cariprazine (RGH-188), a dopamine D3 receptor-preferring, D3/D2 dopamine receptor antagonist–partial agonist antipsychotic candidate: in vitro and neurochemical profile. J. Pharmacol. Exp. Ther. 333, 328–340 (2010).

    Article  CAS  PubMed  Google Scholar 

  23. Maeda, K. et al. Brexpiprazole I: in vitro and in vivo characterization of a novel serotonin-dopamine activity modulator. J. Pharmacol. Exp. Ther. 350, 589–604 (2014).

    Article  PubMed  CAS  Google Scholar 

  24. 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 

  25. Seeman, P. Clozapine, a fast-off-D2 antipsychotic. ACS Chem. Neurosci. 5, 24–29 (2014).

    Article  CAS  PubMed  Google Scholar 

  26. Venkatakrishnan, A. J. et al. Molecular signatures of G-protein-coupled receptors. Nature 494, 185–194 (2013).

    Article  CAS  PubMed  Google Scholar 

  27. Vass, M. et al. Aminergic GPCR–ligand interactions: a chemical and structural map of receptor mutation data. J. Med. Chem. 62, 3784–3839 (2019).

    Article  CAS  PubMed  Google Scholar 

  28. Klein Herenbrink, C. et al. Molecular determinants of the intrinsic efficacy of the antipsychotic aripiprazole. ACS Chem. Biol. 14, 1780–1792 (2019).

    Article  CAS  PubMed  Google Scholar 

  29. McCorvy, J. D. et al. Structure-inspired design of β-arrestin-biased ligands for aminergic GPCRs. Nat. Chem. Biol. 14, 126–134 (2018).

    Article  CAS  PubMed  Google Scholar 

  30. Frank, A., Kiss, D. J., Keseru, G. M. & Stark, H. Binding kinetics of cariprazine and aripiprazole at the dopamine D3 receptor. Sci. Rep. 8, 12509 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Chen, X. et al. Structure–functional selectivity relationship studies of β-arrestin-biased dopamine D2 receptor agonists. J. Med. Chem. 55, 7141–7153 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Chen, X. et al. Discovery of G protein-biased D2 dopamine receptor partial agonists. J. Med. Chem. 59, 10601–10618 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Imai, S. et al. Structural equilibrium underlying ligand-dependent activation of β2-adrenoreceptor. Nat. Chem. Biol. 16, 430–439 (2020).

    Article  CAS  PubMed  Google Scholar 

  35. Krogsgaard-Larsen, N., Jensen, A. A., Schroder, T. J., Christoffersen, C. T. & Kehler, J. Novel aza-analogous ergoline derived scaffolds as potent serotonin 5-HT6 and dopamine D2 receptor ligands. J. Med. Chem. 57, 5823–5828 (2014).

    Article  CAS  PubMed  Google Scholar 

  36. Besnard, J. et al. Automated design of ligands to polypharmacological profiles. Nature 492, 215–220 (2012).

    Article  CAS  PubMed  Google Scholar 

  37. Kroeze, W. K. et al. PRESTO-Tango as an open-source resource for interrogation of the druggable human GPCRome. Nat. Struct. Mol. Biol. 22, 362–369 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Martel, J. C. et al. WAY-100635 has high selectivity for serotonin 5-HT1A versus dopamine D4 receptors. Eur. J. Pharmacol. 574, 15–19 (2007).

    Article  CAS  PubMed  Google Scholar 

  39. Tanyeri, P. et al. Effects of ziprasidone, SCH23390 and SB277011 on spatial memory in the Morris water maze test in naive and MK-801 treated mice. Pharmacol. Biochem. Behav. 138, 142–147 (2015).

    Article  CAS  PubMed  Google Scholar 

  40. Thomas, E. H. X., Bozaoglu, K., Rossell, S. L. & Gurvich, C. The influence of the glutamatergic system on cognition in schizophrenia: a systematic review. Neurosci. Biobehav. Rev. 77, 369–387 (2017).

    Article  CAS  PubMed  Google Scholar 

  41. Kleven, M. S., Barret-Grevoz, C., Bruins Slot, L. & Newman-Tancredi, A. Novel antipsychotic agents with 5-HT1A agonist properties: role of 5-HT1A receptor activation in attenuation of catalepsy induction in rats. Neuropharmacology 49, 135–143 (2005).

    Article  CAS  PubMed  Google Scholar 

  42. Natesan, S., Reckless, G. E., Nobrega, J. N., Fletcher, P. J. & Kapur, S. Dissociation between in vivo occupancy and functional antagonism of dopamine D2 receptors: comparing aripiprazole to other antipsychotics in animal models. Neuropsychopharmacology 31, 1854–1863 (2006).

    Article  CAS  PubMed  Google Scholar 

  43. Bardin, L., Kleven, M. S., Barret-Grevoz, C., Depoortere, R. & Newman-Tancredi, A. Antipsychotic-like vs cataleptogenic actions in mice of novel antipsychotics having D2 antagonist and 5-HT1A agonist properties. Neuropsychopharmacology 31, 1869–1879 (2006).

    Article  CAS  PubMed  Google Scholar 

  44. Blier, P., Bergeron, R. & de Montigny, C. Selective activation of postsynaptic 5-HT1A receptors induces rapid antidepressant response. Neuropsychopharmacology 16, 333–338 (1997).

    Article  CAS  PubMed  Google Scholar 

  45. Blier, P. & Ward, N. M. Is there a role for 5-HT1A agonists in the treatment of depression? Biol. Psychiatry 53, 193–203 (2003).

    Article  CAS  PubMed  Google Scholar 

  46. Carhart-Harris, R. L. & Nutt, D. J. Serotonin and brain function: a tale of two receptors. J. Psychopharmacol. 31, 1091–1120 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Caffrey, M. & Cherezov, V. Crystallizing membrane proteins using lipidic mesophases. Nat. Protoc. 4, 706–731 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Yamashita, K., Hirata, K. & Yamamoto, M. KAMO: towards automated data processing for microcrystals. Acta Crystallogr. D Struct. Biol. 74, 441–449 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Foadi, J. et al. Clustering procedures for the optimal selection of data sets from multiple crystals in macromolecular crystallography. Acta Crystallogr. D Biol. Crystallogr. 69, 1617–1632 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Kabsch, W. Xds. Acta Crystallogr. D Biol. Crystallogr. 66, 125–132 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Kabsch, W. Integration, scaling, space-group assignment and post-refinement. Acta Crystallogr. D Biol. Crystallogr. 66, 133–144 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. McCoy, A. J. et al. Phaser crystallographic software. J. Appl. Crystallogr. 40, 658–674 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  57. Sherman, W., Day, T., Jacobson, M. P., Friesner, R. A. & Farid, R. Novel procedure for modeling ligand/receptor induced fit effects. J. Med. Chem. 49, 534–553 (2006).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  59. Wei, X. et al. Dopamine D1 or D2 receptor-expressing neurons in the central nervous system. Addict. Biol. 23, 569–584 (2018).

    Article  CAS  PubMed  Google Scholar 

  60. Neill, J. C. et al. Animal models of cognitive dysfunction and negative symptoms of schizophrenia: focus on NMDA receptor antagonism. Pharmacol. Ther. 128, 419–432 (2010).

    Article  CAS  PubMed  Google Scholar 

  61. Black, J. W. & Leff, P. Operational models of pharmacological agonism. Proc. R. Soc. Lond. B Biol. Sci. 220, 141–162 (1983).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by grants from the Ministry of Science and Technology of China (grant no. 2020YFA0509600), the National Natural Science Foundation of China (NSFC) (grant no. 32071197), the Strategic Priority Research Program of the Chinese Academy of Sciences (grant no. XDB19020111), Shanghai Science and Technology Committee (grant no. 19ZR1466200), the Thousand Talents Plan-Youth and Shanghai Rising-Star Program (grant no. 20QA1410600) to S.W.; and by the NSFC (grant no. 81703361), Shanghai Science and Technology Committee (grant no. 20S11901200), Shanghai Municipal Government and ShanghaiTech University to J.C. The diffraction data were collected at BL45XU and BL41XU of SPring-8 (JASRI proposal nos. 2020A2715 and 2020A2605). We thank K. Hirata, Y. Nakamura and K. Hasegawa for their help with data collection. We thank Z. Yang for her help with molecular docking. The Ki determinations of receptor binding profiles were generously provided by the National Institute of Mental Health’s Psychoactive Drug Screening Program, Contract no. HHSN-271-2018-00023-C (NIMH PDSP). The NIMH PDSP is directed by B. L. Roth at the University of North Carolina at Chapel Hill, and by Project Officer J. Driscoll at NIMH, Bethesda, MD, USA.

Author information

Authors and Affiliations

Authors

Contributions

Z.C. expressed the protein, purified the receptor, optimized crystallization conditions, grew crystals for data collection, collected diffraction data, performed ligand binding and functional assays, analyzed all data and assisted with preparing the manuscript. L.F. performed pharmacological assays, analyzed the data and assisted with preparing the manuscript. H.W. synthesized analogs and performed the analytical chemical analysis. J.Y. performed in vivo studies and analyzed the data. D.L. performed the in vitro slice electrophysiological recordings. F.N. performed the GPCRome assay and analyzed the data. J.Q. assisted with in vivo studies. Z. Luo provided independent structure quality control analysis and assisted with structure determination. Z. Liu supervised the slice electrophysiological recordings. J.C. supervised ligand design and synthesis, and edited the manuscript. S.W. was responsible for the overall project strategy, solved and analyzed the structure and management, and wrote the manuscript.

Corresponding authors

Correspondence to Jianjun Cheng or Sheng Wang.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Neuroscience thanks Arthur Christopoulos, Javier Gonzalez-Maeso, and Aashish Manglik for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Differential pharmacological behaviours of the TGAs at 5-HT2AR and DRD2.

a, Chemical structures of aripiprazole, brexpiprazole, and cariprazine. b-m, Concentration-response curves for: b-g, the Gαi-γ dissociation at DRD2; h-i, the β-arrestin1/2 recruitment at DRD2; j-k, the Gαq/11-γ9 dissociation at 5-HT2AR; and l-m, the β-arrestin1/2 recruitment at DRD2. EC50 and Emax values are shown in Supplementary Table 6. In all panels, n = 3 independent experiments, error bars represent SEM.

Source data

Extended Data Fig. 2 Unique binding pose of TGAs in 5-HT2AR and DRD2.

a, The introduced S3727.45N mutation substantially enhances purification yield. b, Electron density maps of aripiprazole and cariprazine. Fo-Fc omit maps were calculated by polder map. c, Overall view of 5-HT2AR/Cariprazine complex structure. Cariprazine is shown as sticks overlaid with transparent spheres. d, Conformational changes between 5-HT2AR/Aripiprazole and 5-HT2AR/25CN-NBOH (active, PDB code: 6WHA) are magnified at four key representative motifs. Conformations of these microswitches in the 5-HT2AR/Aripiprazole structure resemble the inactive states of 5-HT2AR/Risperidone (PDB code: 6A93). e, Alignment of OBPs across family A GPCRs. Aripiprazole and cariprazine stretch to a one-helix turn deeper position below W6.48. Ligands are represented as spheres.

Source data

Extended Data Fig. 3 The binding pose of cariprazine in 5-HT2AR and docking poses of TGAs in DRD2.

a, Top extracellular view of cariprazine and the OBP of 5-HT2AR. b, Schematic of cariprazine- binding residues at a 4.0 Å cut-off. The hydrogen bond between D1553.32 and cariprazine’s basic nitrogen is shown as grey dashed line. Primary and secondary pharmacophore and linker are indicated. c, 2Fo-Fc electron density map of 5-HT2AR exosite-related residues (blue mesh) contoured at 1σ. d, Overlay of DRD/Risperidone (PDB code: 6CM4) with two docked DRD2/Aripiprazole poses. The induced-fit docking with side chain configurations of F2435.47, F2445.48, and F3406.52 corrected recaptures aripiprazole’s binding pose in 5-HT2A/Aripiprazole crystal structure. e, Two different binding poses of cariprazine illustrated in rigid and induced-fit docking models of DRD2. In all panels, the Ballesteros-Weinstein numbering is shown as superscript.

Extended Data Fig. 4 Residue at 5.51 position plays a vital role in determining signalling transduction efficacies.

a, Sequence alignment of the OBP and exosite amino acid residues between DRD2 and 5-HT2AR. b-g, Testing pharmacology of the three TGAs at 5-HT2AR via Gαq9 dissociation and β-arrestin2 recruitment assay and consequent effects of 5.51 mutations when L5.51 mutated into alanine [A], valine [V], or phenylalanine [F]. h-k, assays were set as in b-g but for DRD2. In all panels, n = 3 independent experiments, error bars represent SEM. l, m, Cell surface expression levels of wild-type and 5-HT2AR / DRD2 mutants measured by NanoLuc system (see Methods). WT, wild-type receptor. n = 4 independent experiments performed in quadruplicate replicates.

Source data

Extended Data Fig. 5 The pharmacological profiles of compound 34 at wild-type and 5.51 mutants of DRD2.

a, Synthesis of compound 34 via modification of secondary pharmacophore (SP, highlighted in light blue). A smaller benzothiazole group in compound 34 means a longer interaction distance between SP and F5.51 in DRD2. b, c, Gαi19 dissociation and β-arrestin2 recruitment assay show a decreased Emax of compound 34 compared to aripiprazole. d, e, Mutating F5.51 into alanine [A], leucine [L], or tyrosine [Y] and consequent effects on compound 34’s pharmacological behaviour as evaluated by Gαi19 dissociation and β-arrestin2 recruitment assay. WT, wild-type receptor. In panels b-e, n = 3 independent experiments, error bars represent SEM.

Source data

Extended Data Fig. 6 (–)-IHCH7041 shows selective partial agonism at D2-like receptors.

a, Binding affinities of (–)-IHCH7041 to endogenous receptors, channels, and transporters. The values are shown in Supplementary Table 5. b, (–)-IHCH7041 at 1 μM was screened against 320 non-olfactory GPCRs for agonism in the arrestin recruitment TANGO assay. Each point shows luminescence normalized to basal level at a given GPCR. Data are mean ± SEM of normalized fold change. c-g, Receptors with over 2-fold response were further evaluated by concentration-response-curves in the arrestin recruitment TANGO assay. n = 3 independent experiments, error bars represent SEM.

Source data

Extended Data Fig. 7 The pharmacological profiles of (–)-IHCH7041 at DRD2, DRD3 and 5-HT1AR.

a, Heatmap illustration of transduction coefficients of three TGAs and (–)-IHCH7041 at DRD2, DRD3 and 5-HT1AR. b, Concentration–response curves demonstrating the abilities of three TGAs and (–)-IHCH7041 to activate DRD2, DRD3 and 5-HT1A receptors, EC50 and Emax values are shown in Supplementary Tables 68. All data were generated from BRET experiments, n = 3 independent experiments, error bars represent SEM.

Source data

Extended Data Fig. 8 In vivo characterization of (–)-IHCH7041’s pharmacokinetics and antipsychotic-like properties.

a, Metabolic rate of (–)-IHCH7041, n = 3 mice. b, Brain penetration of (–)-IHCH7041, n = 3 mice. c, d, Locomotor responses in mice are shown as 5-min binned intervals (c) or 0- to 45-min time interval (d) to the vehicle or different doses of aripiprazole followed 30 min later by 0.2 mg/kg MK-801. n = 8 mice per group. e, Bar graph of distance travelled after MK-801 administration (0- to 45-min time interval, Related Fig. 5a). f, (–)-IHCH7041(10 mg kg−1) and three TGAs (10 mg/kg) do not elicit recognizable catalepsy; haloperidol (2 mg kg−1) is used as the positive control to elicit catalepsy as assessed by latency to movement. Compounds were administered i.p. 60 min before the test. n = 8 mice per group. g, Evaluation of effects of (–)-IHCH7041 on learning and memory performance of mice subjected to subchronic MK-801 treatment (Related Fig. 6b). Latency to find the platform among (–)-IHCH7041 (0.00625, 0.025 or 0.1 mg kg−1), aripiprazole (3 mg kg−1), or vehicle groups was accessed by Morris water maze test in the acquisition phase. Compounds were injected i.p. 30 mins before each acquisition trial. n = 8 mice per group. In all panels, error bars, SEM. Veh, vehicle. *P < 0.05 and **P < 0.01 (Two-way ANOVA); specific statistical tests, information on reproducibility, and exact P values are reported in Methods and in Source data.

Source data

Extended Data Fig. 9 Antidepressant-like property of (–)-IHCH7041 in mouse models of depression.

a, Locomotor responses in male mice are shown as 5-min binned intervals to vehicle or different doses of (–)-IHCH7041 treatment 30 min prior to 6 mg/kg amphetamine i.p. injection. b, Bar graph of distance travelled after MK-801 administration (0- to 45-min time interval) and the calculation of ED50. c, Inhibition of hyperlocomotion by (–)-IHCH7041 in female mice. d, Antidepressant-like activity of (–)-IHCH7041 (0.4 mg kg−1) was accessed by forced swim test (FST) and its blockade by 5-HT1AR-selective antagonist WAY-100635 (10 mg kg−1), DRD2 antagonist haloperidol (0.2 mg kg−1) or DRD3-selective antagonist SB-277011 (10 mg kg−1) in mice exposed to acute restraint stress (ARS). e, Antidepressant activity of (–)-IHCH7041 was accessed by tail suspension test (TST), aripiprazole (3 mg kg−1) was tested in parallell in mice exposed to ARS. f, Antidepressant activity of (–)-IHCH7041 (0.4 mg kg−1) was accessed by TST and its blockade by 5-HT1AR-selective antagonist WAY-100635 (10 mg kg−1), DRD2 antagonist haloperidol (0.2 mg kg−1) or DRD3-selective antagonist SB-277011 (10 mg kg−1) in mice exposed to ARS. g, Antidepressant activity of (–)-IHCH7041 (0.4 mg kg−1) in corticosterone (Cort)-induced depression-like model and its blockade by 5-HT1AR-selective antagonist WAY-100635 (10 mg kg−1). Ketamine (10 mg kg−1) was used as an antidepressant control in (d), (f), and (g). n = 8 mice in panels a, c, e, and n = 10 mice in panels d, f, g, female mice are symbolled as filled grey circles in panels d, f, g, error bars represent SEM. Veh, vehicle. NS, not significant, *P < 0.05, **P < 0.01, ***P < 0.001 and ****P < 0.0001 (two-tailed unpaired t-tests); specific statistical tests, information on reproducibility, and exact P values are reported in Methods and in Source data.

Source data

Extended Data Fig. 10 Cognitive, learning, and memory impairments imposed by sub-chronic administration of MK-801 and its reversal by (–)-IHCH7041.

a, In Novel object recognition (NOR) test, injection of single dose of (–)-IHCH7041 (0.1 mg kg−1) reversed cognitive impairment. The cognition recovery effect of (–)-IHCH7041 was blocked by 5-HT1AR selective antagonist WAY-100635 (10 mg kg−1) but not by DRD2 or DRD3 antagonist, that is, haloperidol (0.2 mg kg−1) or SB-277011 (10 mg kg−1). b, c, In Morries water maze (MWZ) test, the learning and memory recovery effect of (–)-IHCH7041 was blocked by 5-HT1AR selective antagonist WAY-100635 (10 mg kg−1), but not by DRD2 or DRD3 antagonist, that is, haloperidol (0.2 mg kg−1) or SB-277011 (10 mg kg−1) both in acquisition phase (c) and probe phase (b). d, Sample trace showing evoked firing of mPFC layer 2/3 pyramidal neurons before and after bath application of (–)-IHCH7041 (10 μM). e, WAY-100635 (20 μM) blocked spiking-inhibitory effect of (–)-IHCH7041, n = 6 neurons from 3 mice. In panels a, b, and c, n = 8 mice. error bars in all panels represent SEM. Veh, vehicle. NS, not significant, *P < 0.05, **P < 0.01, ***P < 0.001 and ****P < 0.0001 (two-tailed unpaired t-tests in (a) and (b), two-way ANOVA in (c) and (e)); specific statistical tests, information on reproducibility, and exact P values are reported in Methods and in Source data.

Source data

Supplementary information

Supplementary Information

Supplementary Tables 1–8 and supplementary protocols of chemical synthesis

Reporting Summary

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 2

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 5

Statistical source data.

Source Data Extended Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, Z., Fan, L., Wang, H. et al. Structure-based design of a novel third-generation antipsychotic drug lead with potential antidepressant properties. Nat Neurosci 25, 39–49 (2022). https://doi.org/10.1038/s41593-021-00971-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-021-00971-w

This article is cited by

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