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:

GluN2A mediates ketamine-induced rapid antidepressant-like responses

Abstract

Ketamine was thought to induce rapid antidepressant responses by inhibiting GluN2B-containing N-methyl-d-aspartic acid (NMDA) receptors (NMDARs), which presents a promising opportunity to develop better antidepressants. However, adverse side effects limit the broader application of ketamine and GluN2B inhibitors are yet to be approved for clinical use. It is unclear whether ketamine acts solely through GluN2B-dependent mechanisms. The present study reports that the loss of another major NMDAR subunit, GluN2A, in adult mouse brains elicits robust antidepressant-like responses with limited impact on the behaviors that mimic the psychomimetic effects of ketamine. The antidepressant-like behavioral effects of broad NMDAR channel blockers, such as ketamine and MK-801 (dizocilpine), were mediated by the suppression of GluN2A, but not by the inhibition of GluN2B. Moreover, treatment with ketamine or MK-801 rapidly increased the intrinsic excitability of hippocampal principal neurons through GluN2A, but not GluN2B. Together, these findings indicate that GluN2A mediates ketamine-triggered rapid antidepressant-like responses.

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: MK-801-induced fast antidepressant and anxiolytic effects, not hyperlocomotion, were occluded in GluN2A−/− mice.
Fig. 2: GluN2A removal in adult mouse brains induced antidepressant-like behaviors but not anxiolytic or hyperlocomotive behaviors.
Fig. 3: KO of GluN2A in excitatory neurons, not in inhibitory neurons, led to antidepressant, anxiolytic-like and hyperlocomotive behaviors in mice.
Fig. 4: KD of hippocampal GluN2A promoted antidepressant-like behaviors.
Fig. 5: GluN2A did not regulate the excitatory and inhibitory synaptic transmission of CA1 pyramidal neurons.
Fig. 6: Suppression of GluN2A elevated intrinsic excitability of CA1 pyramidal neurons in a cell autonomous manner.
Fig. 7: Acute treatment with MK-801 or ketamine increased excitability of CA1 pyramidal neurons in WT, but not in GluN2A−/−, mice.

Similar content being viewed by others

Data availability

The datasets generated during and analyzed during the present study are available from the corresponding author on reasonable request. Source data are provided with this paper.

Code availability

The present study did not use any customized code or mathematical algorithm.

References

  1. Nierenberg, A. A. et al. Timing of onset of antidepressant response with fluoxetine treatment. Am. J. Psychiatry 157, 1423–1428 (2000).

    Article  CAS  PubMed  Google Scholar 

  2. Al-Harbi, K. S. Treatment-resistant depression: therapeutic trends, challenges, and future directions. Patient Prefer. Adherence 6, 369–388 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  3. Duman, R. S. Ketamine and rapid-acting antidepressants: a new era in the battle against depression and suicide. F1000Res https://doi.org/10.12688/f1000research.14344.1 (2018).

  4. Gibbons, R. D. et al. Early evidence on the effects of regulators’ suicidality warnings on SSRI prescriptions and suicide in children and adolescents. Am. J. Psychiatry 164, 1356–1363 (2007).

    Article  PubMed  Google Scholar 

  5. Kim, J., Farchione, T., Potter, A., Chen, Q. & Temple, R. Esketamine for treatment-resistant depression−first FDA-approved antidepressant in a new class. N. Engl. J. Med. 381, 1–4 (2019).

    Article  PubMed  Google Scholar 

  6. Ng, S. H., Tse, M. L., Ng, H. W. & Lau, F. L. Emergency department presentation of ketamine abusers in Hong Kong: a review of 233 cases. Hong Kong Med J. 16, 6–11 (2010).

    CAS  PubMed  Google Scholar 

  7. Trullas, R. & Skolnick, P. Functional antagonists at the NMDA receptor complex exhibit antidepressant actions. Eur. J. Pharmacol. 185, 1–10 (1990).

    Article  CAS  PubMed  Google Scholar 

  8. Iosifescu, D. V. et al. Efficacy and safety of AXS-05 (dextromethorphan-bupropion) in patients with major depressive disorder: a phase 3 randomized clinical trial (GEMINI). J. Clin. Psychiatry https://doi.org/10.4088/JCP.21m14345 (2022).

  9. Nowak, G., Ordway, G. A. & Paul, I. A. Alterations in the N-methyl-d-aspartate (NMDA) receptor complex in the frontal cortex of suicide victims. Brain Res. 675, 157–164 (1995).

    Article  CAS  PubMed  Google Scholar 

  10. Liu, R. et al. Correlation of functional GRIN2A gene promoter polymorphisms with schizophrenia and serum d-serine levels. Gene 568, 25–30 (2015).

    Article  CAS  PubMed  Google Scholar 

  11. Paoletti, P., Bellone, C. & Zhou, Q. NMDA receptor subunit diversity: impact on receptor properties, synaptic plasticity and disease. Nat. Rev. Neurosci. 14, 383–400 (2013).

    Article  CAS  PubMed  Google Scholar 

  12. Miller, O. H. et al. GluN2B-containing NMDA receptors regulate depression-like behavior and are critical for the rapid antidepressant actions of ketamine. eLife 3, e03581 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Jimenez-Sanchez, L., Campa, L., Auberson, Y. P. & Adell, A. The role of GluN2A and GluN2B subunits on the effects of NMDA receptor antagonists in modeling schizophrenia and treating refractory depression. Neuropsychopharmacology 39, 2673–2680 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Gerhard, D. M. et al. GABA interneurons are the cellular trigger for ketamine’s rapid antidepressant actions. J. Clin. Invest. 130, 1336–1349 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Henter, I. D., Park, L. T. & Zarate, C. A. Jr. Novel glutamatergic modulators for the treatment of mood disorders: current status. CNS Drugs 35, 527–543 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Zanos, P. et al. Ketamine and ketamine metabolite pharmacology: insights into therapeutic mechanisms. Pharm. Rev. 70, 621–660 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Aguilar-Valles, A. et al. Antidepressant actions of ketamine engage cell-specific translation via eIF4E. Nature 590, 315–319 (2021).

    Article  CAS  PubMed  Google Scholar 

  18. Cui, Y. et al. Astroglial Kir4.1 in the lateral habenula drives neuronal bursts in depression. Nature 554, 323–327 (2018).

    Article  CAS  PubMed  Google Scholar 

  19. Autry, A. E. et al. NMDA receptor blockade at rest triggers rapid behavioural antidepressant responses. Nature 475, 91–95 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Li, N. et al. mTOR-dependent synapse formation underlies the rapid antidepressant effects of NMDA antagonists. Science 329, 959–964 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Zanos, P. et al. NMDAR inhibition-independent antidepressant actions of ketamine metabolites. Nature 533, 481–486 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Wilkinson, S. T. & Sanacora, G. A new generation of antidepressants: an update on the pharmaceutical pipeline for novel and rapid-acting therapeutics in mood disorders based on glutamate/GABA neurotransmitter systems. Drug Discov. Today 24, 606–615 (2019).

    Article  CAS  PubMed  Google Scholar 

  23. Berberich, S. et al. Lack of NMDA receptor subtype selectivity for hippocampal long-term potentiation. J. Neurosci. 25, 6907–6910 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Francija, E. et al. GluN2A-ERK-mTOR pathway confers a vulnerability to LPS-induced depressive-like behaviour. Behav. Brain Res. 417, 113625 (2022).

    Article  CAS  PubMed  Google Scholar 

  25. Boyce-Rustay, J. M. & Holmes, A. Genetic inactivation of the NMDA receptor NR2A subunit has anxiolytic- and antidepressant-like effects in mice. Neuropsychopharmacology 31, 2405–2414 (2006).

    Article  CAS  PubMed  Google Scholar 

  26. Lim, A. L., Taylor, D. A. & Malone, D. T. Consequences of early life MK-801 administration: long-term behavioural effects and relevance to schizophrenia research. Behav. Brain Res. 227, 276–286 (2012).

    Article  CAS  PubMed  Google Scholar 

  27. Kiselycznyk, C. et al. NMDA receptor subunits and associated signaling molecules mediating antidepressant-related effects of NMDA-GluN2B antagonism. Behav. Brain Res. 287, 89–95 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Koike, H., Fukumoto, K., Iijima, M. & Chaki, S. Role of BDNF/TrkB signaling in antidepressant-like effects of a group II metabotropic glutamate receptor antagonist in animal models of depression. Behav. Brain Res. 238, 48–52 (2013).

    Article  CAS  PubMed  Google Scholar 

  29. Lang, E. et al. Molecular and cellular dissection of NMDA receptor subtypes as antidepressant targets. Neurosci. Biobehav. Rev. 84, 352–358 (2018).

    Article  CAS  PubMed  Google Scholar 

  30. Pothula, S. et al. Cell-type specific modulation of NMDA receptors triggers antidepressant actions. Mol. Psychiatry https://doi.org/10.1038/s41380-020-0796-3 (2020).

  31. Campbell, S. & Macqueen, G. The role of the hippocampus in the pathophysiology of major depression. J. Psychiatry Neurosci. 29, 417–426 (2004).

    PubMed  PubMed Central  Google Scholar 

  32. Gu, X., Zhou, L. & Lu, W. An NMDA receptor-rependent mechanism underlies inhibitory synapse development. Cell Rep. 14, 471–478 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Tatard-Leitman, V. M. et al. Pyramidal cell selective ablation of N-methyl-d-aspartate receptor 1 causes increase in cellular and network excitability. Biol. Psychiatry 77, 556–568 (2015).

    Article  CAS  PubMed  Google Scholar 

  34. Kim, C. S., Chang, P. Y. & Johnston, D. Enhancement of dorsal hippocampal activity by knockdown of HCN1 channels leads to anxiolytic- and antidepressant-like behaviors. Neuron 75, 503–516 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Han, Y. et al. HCN-channel dendritic targeting requires bipartite interaction with TRIP8b and regulates antidepressant-like behavioral effects. Mol. Psychiatry 22, 458–465 (2017).

    Article  CAS  PubMed  Google Scholar 

  36. Kernell, D. Input resistance, electrical excitability, and size of ventral horn cells in cat spinal cord. Science 152, 1637–1640 (1966).

    Article  CAS  PubMed  Google Scholar 

  37. Hou, G. & Zhang, Z. W. NMDA receptors regulate the development of neuronal intrinsic excitability through cell-autonomous mechanisms. Front. Cell Neurosci. 11, 353 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  38. Zanos, P. et al. NMDA receptor activation-dependent antidepressant-relevant behavioral and synaptic actions of ketamine. J. Neurosci. 43, 1038–1050 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Frizelle, P. A., Chen, P. E. & Wyllie, D. J. Equilibrium constants for (R)-[(S)-1-(4-bromo-phenyl)-ethylamino]-(2,3-dioxo-1,2,3,4-tetrahydroquinoxalin-5-yl)-methyl]-phosphonic acid (NVP-AAM077) acting at recombinant NR1/NR2A and NR1/NR2B N-methyl-d-aspartate receptors: implications for studies of synaptic transmission. Mol. Pharmacol. 70, 1022–1032 (2006).

    Article  CAS  PubMed  Google Scholar 

  40. Russo, S. J. & Nestler, E. J. The brain reward circuitry in mood disorders. Nat. Rev. Neurosci. 14, 609–625 (2013).

    Article  CAS  PubMed  Google Scholar 

  41. Abdallah, C. G. et al. The nucleus accumbens and ketamine treatment in major depressive disorder. Neuropsychopharmacology 42, 1739–1746 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Khan, A. R. et al. Neurite atrophy in dorsal hippocampus of rat indicates incomplete recovery of chronic mild stress induced depression. NMR Biomed. 32, e4057 (2019).

    Article  PubMed  Google Scholar 

  43. Felix-Ortiz, A. C. & Tye, K. M. Amygdala inputs to the ventral hippocampus bidirectionally modulate social behavior. J. Neurosci. 34, 586–595 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Bagot, R. C. et al. Ventral hippocampal afferents to the nucleus accumbens regulate susceptibility to depression. Nat. Commun. 6, 7062 (2015).

    Article  CAS  PubMed  Google Scholar 

  45. Kim, C. S. & Johnston, D. Antidepressant effects of (S)-ketamine through a reduction of hyperpolarization-activated current Ih. iScience 23, 101239 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Homayoun, H. & Moghaddam, B. NMDA receptor hypofunction produces opposite effects on prefrontal cortex interneurons and pyramidal neurons. J. Neurosci. 27, 11496–11500 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Quirk, M. C., Sosulski, D. L., Feierstein, C. E., Uchida, N. & Mainen, Z. F. A defined network of fast-spiking interneurons in orbitofrontal cortex: responses to behavioral contingencies and ketamine administration. Front. Syst. Neurosci. 3, 13 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Picard, N., Takesian, A. E., Fagiolini, M. & Hensch, T. K. NMDA 2A receptors in parvalbumin cells mediate sex-specific rapid ketamine response on cortical activity. Mol. Psychiatry 24, 828–838 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Widman, A. J. & McMahon, L. L. Disinhibition of CA1 pyramidal cells by low-dose ketamine and other antagonists with rapid antidepressant efficacy. Proc. Natl Acad. Sci. USA 115, E3007–E3016 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Lu, W., Bushong, E. A., Shih, T. P., Ellisman, M. H. & Nicoll, R. A. The cell-autonomous role of excitatory synaptic transmission in the regulation of neuronal structure and function. Neuron 78, 433–439 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Kase, D. & Imoto, K. The role of HCN channels on membrane excitability in the nervous system. J. Signal Transduct. 2012, 619747 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  52. Burrone, J., O’Byrne, M. & Murthy, V. N. Multiple forms of synaptic plasticity triggered by selective suppression of activity in individual neurons. Nature 420, 414–418 (2002).

    Article  CAS  PubMed  Google Scholar 

  53. Kadotani, H. et al. Motor discoordination results from combined gene disruption of the NMDA receptor NR2A and NR2C subunits, but not from single disruption of the NR2A or NR2C subunit. J. Neurosci. 16, 7859–7867 (1996).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Zhou, J. et al. NMDA receptor-dependent prostaglandin-endoperoxide synthase 2 induction in neurons promotes glial proliferation during brain development and injury. Cell Rep. 38, 110557 (2022).

    Article  CAS  PubMed  Google Scholar 

  55. Ruzankina, Y. et al. Deletion of the developmentally essential gene ATR in adult mice leads to age-related phenotypes and stem cell loss. Cell Stem Cell 1, 113–126 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Goebbels, S. et al. Genetic targeting of principal neurons in neocortex and hippocampus of NEX-Cre mice. Genesis 44, 611–621 (2006).

    Article  CAS  PubMed  Google Scholar 

  57. Vong, L. et al. Leptin action on GABAergic neurons prevents obesity and reduces inhibitory tone to POMC neurons. Neuron 71, 142–154 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Kaneko, R. et al. Inhibitory neuron-specific Cre-dependent red fluorescent labeling using VGAT BAC-based transgenic mouse lines with identified transgene integration sites. J. Comp. Neurol. 526, 373–396 (2018).

    Article  CAS  PubMed  Google Scholar 

  59. Madisen, L. et al. A robust and high-throughput Cre reporting and characterization system for the whole mouse brain. Nat. Neurosci. 13, 133–140 (2010).

    Article  CAS  PubMed  Google Scholar 

  60. Yang, Y. et al. Ketamine blocks bursting in the lateral habenula to rapidly relieve depression. Nature 554, 317–322 (2018).

    Article  CAS  PubMed  Google Scholar 

  61. Ryan, T. J., Roy, D. S., Pignatelli, M., Arons, A. & Tonegawa, S. Memory. Engram cells retain memory under retrograde amnesia. Science 348, 1007–1013 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Paxinos, G. & Franklin, K. B. J. The Mouse Brain In Stereotaxic Coordinates, 4th edn (Elsevier Science & Technology, 2013).

  63. Zhang, Q., He, Q., Wang, J., Fu, C. & Hu, H. Use of TAI-FISH to visualize neural ensembles activated by multiple stimuli. Nat. Protoc. 13, 118–133 (2018).

    Article  PubMed  Google Scholar 

  64. Chen, Y. et al. Activity-induced Nr4a1 regulates spine density and distribution pattern of excitatory synapses in pyramidal neurons. Neuron 83, 431–443 (2014).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank S. Nakanishi (Osaka Bioscience Institute and Kyoto University, Japan) for providing the Grin2a−/− mice (by the RIKEN BRC through the National Bio-Resource Project of MEXT, Japan). We thank K. He (Interdisciplinary Research Center on Biology and Chemistry, Chinese Academy of Sciences, China), J. Hu (ShanghaiTech University, China), Z. Qiu (Institute of Neuroscience, Chinese Academy of Sciences, China) and X. Xu (Institute of Neuroscience, Chinese Academy of Sciences, China) for helping us to acquire the genetically engineered tool mice. We thank the staff members of the Animal Facility at the National Facility for Protein Science in Shanghai, Shanghai Advanced Research Institute and the Chinese Academy of Sciences for providing assistance in mouse breeding and maintenance. We thank M. Sheng (Broad Institute and Massachusetts Institute of Technology, USA) and J. Yuan (Interdisciplinary Research Center on Biology and Chemistry, Chinese Academy of Sciences, China) for helping with the manuscript preparation. The present study was supported by the following agencies: Shanghai Municipal Science and Technology Major Project to Y.C. (grant no. 2019SHZDZX02) and Natural Science Foundation of Shanghai to Y.G. (grant nos. 19ZR1468600 and 201409003800).

Author information

Authors and Affiliations

Authors

Contributions

Y.C. and Y.G. conceived the idea and designed the study. T.S., Y.L. and C.F. performed the experiments and statistical analysis, and wrote the manuscript under the supervision of Y.C. and Y.G.

Corresponding authors

Correspondence to Yang Geng or Yelin Chen.

Ethics declarations

Competing interests

Y.C. is a visiting professor of Shanghai Jiao Tong University and a founder of Synphatec (Shanghai) Biopharmaceutical Technology Co., Ltd. Y.C., Y.G., T.S., Y.L. and C.F. are inventors on China patent application (no. 202110901806.8) held by Shanghai Institute of Organic Chemistry, Chinese Academy of Sciences.

Peer review

Peer review information

Nature Neuroscience thanks Gerard Sanacora, Panos Zanos 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

Extended Data Fig. 1 MK-801-induced fast antidepressant and anxiolytic effects, not hyperlocomotion, were occluded in GluN2A−/− mice, related to Fig. 1.

a and b, GluN2A expression was evaluated with FISH (a) and western blot (b) in cortex or hippocampus from WT and GluN2A−/− mice. Scale bar showed 200 μm. Samples of (b) derive from the same experiment and that gels/blots were processed in parallel. cf, Measurements of GluN2A−/− mice and their WT littermates with OFT (c, p = 0.00020, d, p = 0.0040, WT n = 16, GluN2A−/− n = 14) and NSFT (e, p = 0.0031, f, WT n = 10, GluN2A−/− n = 11). g, Escape latency of WT and GluN2A−/− mice in learned helplessness test (LH) (WT-Sham n = 7, WT-IES n = 20, p = 0.0013; GluN2A−/−-Sham n = 10, GluN2A−/−-IES n = 19, p < 0.0001). hj, Effects of MK-801 treatment on WT and GluN2A−/− mice were evaluated with OFT (WT-Sal n = 6, WT-MK-801 n = 8, GluN2A−/−-Sal n = 8, GluN2A−/−-MK-801 n = 9). h, WT-Sal v.s. WT-MK-801 p = 0.00040; WT-Sal v.s. GluN2A−/−-Sal p = 0.0080; GluN2A−/−-Sal v.s. GluN2A−/−- MK-801 p = 0.0010; i, WT-Sal v.s. WT-MK-801 p < 0.0001; WT-Sal v.s. GluN2A−/−-Sal p = 0.0080; GluN2A−/−-Sal v.s. GluN2A−/−-MK-801 p = 0.0019; j, WT-Sal v.s. WT-MK-801 p = 0.00010; WT-Sal v.s. GluN2A−/−-Sal p < 0.0001. k, Inhibition of GluN2A-containing NMDAR-mediated currents by ketamine, (2R, 6R)-HNK and LY341495. IComp/IGlu represents the normalized remaining currents after co-perfusion with each compound. Veh v.s. Ket p < 0.0001, one-way ANOVA. l, Schematic diagram showing the experimental timeline. m, Quantitation of the immobility time from WT and GluN2A−/− in TST after 7 days fluoxetine (Flx) treatment (WT-Sal n = 10, WT-Flx n = 12, p = 0.0046; GluN2A−/−-Sal n = 9, GluN2A−/−-Flx n = 12, p = 0.030; WT-Sal v.s. GluN2A−/−-Sal p = 0.0039). Error bars show SEM. Two-way ANOVA (g, h, i, j, m) and Student’s t test (two-tailed) (cf). * p < 0.05, ** p < 0.01, *** p < 0.001, **** p < 0.0001.

Source data

Extended Data Fig. 2 Validation of adult age-specific induced GluN2A knockout in mice, related to Fig. 2.

a, Body weight of UBC-CreERT/Grin2aflox/flox mice treated with TAM (n = 22) or Veh (n = 20). bd, GluN2A expression level was measured from hippocampus of UBC-CreERT/Grin2aflox/flox mice after treated with TAM or Veh (b, c, western blot, Veh n = 7, TAM n = 6) (d, qRT-PCR, Veh n = 5, TAM n = 5). c, p < 0.0001; d, p < 0.0001; ei, Evaluation of Grin2aflox/flox (2A flox) and UBC-CreERT/Grin2aflox/flox (UBC-2A flox) mice with OFT (e), EPM (f, g), TST (h, p = 0.011) or FST (i, p = 0.025). The behavioral analysis was done 1 month after treatment with TAM (2A flox n = 8, UBC-2A flox n = 9). j and k, Verification of GluN2A knockout in UBC-CreERT/Grin2aflox/flox mice treated with TAM after inescapable footshocks training in LH test presented in Fig. 2i–n (Veh n = 5, TAM n = 6). k, p < 0.0001. Samples of (b)/(j) derive from the same experiment and that gels/blots were processed in parallel. Error bars show SEM. One-sample t test (c, k), two-way ANOVA (eg) and Student’s t test (two-tailed) (d, h, i). * p < 0.05, **** p < 0.0001.

Source data

Extended Data Fig. 3 Validation of excitatory-neuron specific GluN2A conditional knockout mice (Nex-2A cKO), related to Fig. 3.

a, Greyscale images showing tdTomato signal in sagittal brain sections of Nex-Cre/Ai9 mice. C-cortex, H-hippocampus, S-striatum, T-thalamus, A-amygdala, M-midbrain. Scale bars showed 500 μm (top) and 200 μm (bottom) respectively. b and c, Representative greyscale images of Grin2a FISH in hippocampus (b) or cortex (c) of WT and Nex-2A cKO mice. Scale bar showed 50 μm in b and 100 μm in c. d and e, Western blot analysis of GluN2A in hippocampus (d) and cortex (e) of Nex-2A WT and cKO mice. Samples of (d)/(e) derive from the same experiment and that gels/blots were processed in parallel. f and g, Quantification of GluN2A expression level by western blot (f) and RT-PCR (g) in hippocampus or cortex of Nex-2A WT and cKO mice (f, Hipp: Nex-2A WT n = 4, cKO n = 5, p < 0.0001; Ctx: Nex-2A WT n = 4, cKO n = 3, p = 0.010), (g, Hipp: Nex-2A WT n = 4, cKO n = 3, p = 0.0045; Ctx: Nex-2A WT n = 4, cKO n = 3, p = 0.00020). One-sample t test (f) and Student’s t test (two-tailed) (g). Error bars show SEM. ** p < 0.01, *** p < 0.001, **** p < 0.0001.

Source data

Extended Data Fig. 4 Validation of inhibitory-neuron specific GluN2A conditional knockout mice (vGAT-2A cKO), related to Fig. 3.

a, Greyscale images showing tdTomato signal in sagittal brain sections from vGAT-Cre/Ai9 mice. C-cortex, H-hippocampus, S-striatum, T-thalamus, A-amygdala, M-midbrain. Scale bars showed 500 μm (top) and 200 μm (bottom), respectively. b and c, Western blot analysis of GluN2A expression level in hippocampus of vGAT-2A cKO or WT mice, quantified in b (vGAT-2A WT n = 3, vGAT-2A cKO n = 3). Samples of (b) derive from the same experiment and that gels/blots were processed in parallel. d, Quantification of GluN2A mRNA level in hippocampus of vGAT-2A cKO or WT mice (vGAT-2A WT n = 4, vGAT-2A cKO n = 6). eg, Grin2a FISH and fluorescent staining of GAD1 in hippocampus of vGAT-2A cKO or WT mice (e; yellow arrows indicated the GAD1 positive cells that co-expressed with GluN2A; cyan arrowhead indicated GAD1 positive cells without GluN2A signal). f, Area of stratum radiatum (SR) and stratum lacunosum & moleculare (SLM). g, Density of GluN2A positive cells in SR and SLM (vGAT-2A WT n = 4, vGAT-2A cKO n = 4; p = 0.0024,). Scale bar showed 200 μm. h, Representative traces of evoked NMDAR-mediated EPSCs from CA1 pyramidal neurons or interneurons in WT or vGAT-2A cKO mice after incubated with Veh (black) or GluN2B selective inhibitor Ro-25 6981 (blue). i and j, Quantitation of the decay time (i) or +Ro25/-Ro25 ratio of the amplitudes of evoked NMDAR-EPSCs (j) (Pyr: pyramidal neurons, vGAT-2A WT n = 3, vGAT-2A cKO n = 5; INs: interneurons, vGAT-2A WT n = 5, vGAT-2A cKO n = 5;). i, INs: p = 0.018; j, INs: p = 0.012. Error bars show SEM. One-sample t test (c, d), two-way ANOVA (f, i, j) and Student’s t test (two-tailed) (g).* p < 0.05, ** p < 0.01.

Source data

Extended Data Fig. 5 Hippocampal interneurons of GluN2A−/− mice had normal excitability, related to Fig. 6.

aj, Representative current-clamp recording traces after injection of a 330 pA depolarizing current (a, f), quantitation of AP number (b, g), RMP (c, h), sample traces after step hyperpolarizing current injections (d, i), current-voltage curves and RIn (e, j and inserts) from pyramidal neurons of mPFC (ae) or visual cortex (fj) of WT (grey) or GluN2A−/− (red) mice. (mPFC: WT n = 9, GluN2A−/− n = 16; visual cortex: WT n = 13, GluN2A−/− n = 9). b, p(130 pA) = 0.012, p(150 pA) = 0.018; e, p = 0.031, p(RIn) = 0.031. kp, Representative current-clamp recording traces after injection of a 330 (k) or 210 (n) pA depolarizing current (k, n), quantitation of AP number after depolarizing current injections (l, o), RMP (m, p) from non-fast spiking (km) or fast-spiking (np) interneurons from CA1 of WT (grey) or GluN2A−/− (red) mice. (Non-fast, WT n = 12, GluN2A−/− n = 7; Fast, WT n = 13, GluN2A−/− n = 14). 6–8 weeks old mice were used for electrophysiological recordings. Error bars show SEM. Two-way ANOVA (b, e, g, j, l, o) or Student’s t test (two-tailed) (c, e (insert), h, j (insert), m, p). * p < 0.05.

Extended Data Fig. 6 Cell type specific removal of GluN2A did not alter excitatory or inhibitory synaptic transmission of CA1 pyramidal neurons, related to Fig. 6.

af, Representative traces (a, d), average individual events (b, e), quantitation of cumulative probability and average amplitude of events (c, f) of mIPSCs and sIPSCs recorded from 4–6 weeks old Nex-2A WT (black) or Nex-2A cKO (olive) mice (mIPSCs: Nex-2A WT n = 12, Nex-2A cKO n = 14; sIPSCs: Nex-2A WT n = 16, Nex-2A cKO n = 14). gr, Representative traces, average individual events, quantitation of cumulative probability and average amplitude of events of mIPSCs (gi), sIPSCs (jl), mEPSCs (mo) and sEPSCs (pr) recorded from vGAT-2A WT (black) or vGAT-2A cKO (purple) mice. (mIPSCs: vGAT-2A WT n = 18, vGAT-2A cKO n = 21; sIPSCs: vGAT-2A WT n = 15, vGAT-2A cKO n = 13; mEPSCs: vGAT-2A WT n = 13, vGAT-2A cKO n = 10; sEPSCs: vGAT-2A WT n = 12, vGAT-2A cKO n = 16) 6–8 weeks mcie were used for electrophysiological recordings. Error bars showed SEM. Cumulative frequency was analyzed with Kolmogorov-Smirnov test. Peak amplitudes were analyzed with Student’s t test (two-tailed).

Extended Data Fig. 7 GluN2B was not responsible to the effects of MK-801 in intrinsic excitability and antidepressant-like behaviors, related to Fig. 6.

a, Representative recording traces after injection of 330 pA depolarizing current in CA1 pyramidal neurons treated with Veh or 3 μM Ro-25 6981 for 1–2 hours. b and c, Quantitation of the number of action potentials (b) after injection of depolarizing currents and RMP (c) recorded from neurons after Veh or Ro-25 6981 treatments (Veh n = 18, Ro 25-6981 n = 19). di, number of AP (e, p = 0.0074), RMP (f, p = 0.041), current-voltage curve (h, p = 0.00070), input resistance (RIn) (i, p = 0.00080) and sample trace recorded from CA1 pyramidal neurons after incubated with ‘Ro 25-6981 + Veh’ (n = 10) or ‘Ro 25-6981 + 10 μM MK-801’ (n = 10) for 1–2 hours from WT mice (g). j and k, Effects of MK-801 on WT mice-treated with GluN2B inhibitor Ro 25-6981 (10 mpk) were evaluated with TST (j; Ro 25-6981 + Sal n = 10; Ro 25-6981 + MK-801 n = 10, p < 0.0001) and FST (k; Ro 25-6981 + Sal n = 10; Ro 25-6981 + MK-801 n = 9, p < 0.0001). All the recordings were measured from CA1 pyramidal neurons acutely prepared from 6-8 weeks old WT mice. Error bars show SEM. Two-way ANOVA (b, e, h) or Student’s t test (two-tailed) (c, f, i, j, k). * p < 0.05, ** p < 0.01, *** p < 0.001, **** p < 0.0001.

Extended Data Fig. 8 Sag potentials were not altered in CA1 pyramidal neurons from GluN2A−/− mice.

a, Representative voltage sag were generated by hyperpolarized current of −100 pA. b, Normalized voltage sag of WT and GluN2A−/−, which were measured at the last 50-ms of 1-s stimulation (WT n = 14, GluN2A−/− n = 22). Error bars show SEM. Two-way ANOVA.

Supplementary information

Source data

Source Data Extended Data Fig. 1

Unprocessed western blots.

Source Data Extended Data Fig. 1

Unprocessed western blots.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 2

Unprocessed western blots.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 3

Unprocessed western blots.

Source Data Extended Data Fig. 4

Unprocessed western blots.

Source Data Extended Data Fig. 4

Unprocessed western blots.

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

Su, T., Lu, Y., Fu, C. et al. GluN2A mediates ketamine-induced rapid antidepressant-like responses. Nat Neurosci 26, 1751–1761 (2023). https://doi.org/10.1038/s41593-023-01436-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-023-01436-y

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