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:

Antidepressant actions of ketamine engage cell-specific translation via eIF4E

Abstract

Effective pharmacotherapy for major depressive disorder remains a major challenge, as more than 30% of patients are resistant to the first line of treatment (selective serotonin reuptake inhibitors)1. Sub-anaesthetic doses of ketamine, a non-competitive N-methyl-d-aspartate receptor antagonist2,3, provide rapid and long-lasting antidepressant effects in these patients4,5,6, but the molecular mechanism of these effects remains unclear7,8. Ketamine has been proposed to exert its antidepressant effects through its metabolite (2R,6R)-hydroxynorketamine ((2R,6R)-HNK)9. The antidepressant effects of ketamine and (2R,6R)-HNK in rodents require activation of the mTORC1 kinase10,11. mTORC1 controls various neuronal functions12, particularly through cap-dependent initiation of mRNA translation via the phosphorylation and inactivation of eukaryotic initiation factor 4E-binding proteins (4E-BPs)13. Here we show that 4E-BP1 and 4E-BP2 are key effectors of the antidepressant activity of ketamine and (2R,6R)-HNK, and that ketamine-induced hippocampal synaptic plasticity depends on 4E-BP2 and, to a lesser extent, 4E-BP1. It has been hypothesized that ketamine activates mTORC1–4E-BP signalling in pyramidal excitatory cells of the cortex8,14. To test this hypothesis, we studied the behavioural response to ketamine and (2R,6R)-HNK in mice lacking 4E-BPs in either excitatory or inhibitory neurons. The antidepressant activity of the drugs is mediated by 4E-BP2 in excitatory neurons, and 4E-BP1 and 4E-BP2 in inhibitory neurons. Notably, genetic deletion of 4E-BP2 in inhibitory neurons induced a reduction in baseline immobility in the forced swim test, mimicking an antidepressant effect. Deletion of 4E-BP2 specifically in inhibitory neurons also prevented the ketamine-induced increase in hippocampal excitatory neurotransmission, and this effect concurred with the inability of ketamine to induce a long-lasting decrease in inhibitory neurotransmission. Overall, our data show that 4E-BPs are central to the antidepressant activity of ketamine.

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: 4E-BP1 and 4E-BP2 are required for the antidepressant effect of ketamine.
Fig. 2: Differential effects of 4E-BP1 and 4E-BP2 in neuronal populations in response to ketamine and (2R,6R)-HNK.
Fig. 3: Ketamine-induced increase in excitatory synaptic transmission in pyramidal cells depends on the expression of 4E-BP2 in inhibitory neurons.
Fig. 4: Inhibitory synaptic transmission in pyramidal cells and its depression by ketamine depend on the expression of 4E-BP2 in inhibitory interneurons.

Similar content being viewed by others

Data availability

All data generated and analysed during this study are included in this published Article and its Supplementary Information filesSource data are provided with this paper.

References

  1. Rush, A. J. et al. Acute and longer-term outcomes in depressed outpatients requiring one or several treatment steps: a STAR*D report. Am. J. Psychiatry 163, 1905–1917 (2006).

    Article  Google Scholar 

  2. Hirota, K. & Lambert, D. G. Ketamine: its mechanism(s) of action and unusual clinical uses. Br. J. Anaesth. 77, 441–444 (1996).

    Article  CAS  Google Scholar 

  3. Kashiwagi, K. et al. Channel blockers acting at N-methyl-d-aspartate receptors: differential effects of mutations in the vestibule and ion channel pore. Mol. Pharmacol. 61, 533–545 (2002).

    Article  CAS  Google Scholar 

  4. Berman, R. M. et al. Antidepressant effects of ketamine in depressed patients. Biol. Psychiatry 47, 351–354 (2000).

    Article  CAS  Google Scholar 

  5. Zarate, C. A., Jr et al. A randomized trial of an N-methyl-d-aspartate antagonist in treatment-resistant major depression. Arch. Gen. Psychiatry 63, 856–864 (2006).

    Article  CAS  Google Scholar 

  6. Lapidus, K. A. et al. A randomized controlled trial of intranasal ketamine in major depressive disorder. Biol. Psychiatry 76, 970–976 (2014).

    Article  CAS  Google Scholar 

  7. Zarate, C. A., Jr & Machado-Vieira, R. Ketamine: translating mechanistic discoveries into the next generation of glutamate modulators for mood disorders. Mol. Psychiatry 22, 324–327 (2017).

    Article  CAS  Google Scholar 

  8. Workman, E. R., Niere, F. & Raab-Graham, K. F. Engaging homeostatic plasticity to treat depression. Mol. Psychiatry 23, 26–35 (2018).

    Article  CAS  Google Scholar 

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

    Article  CAS  ADS  Google Scholar 

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

    Article  CAS  ADS  Google Scholar 

  11. Fukumoto, K. et al. Activity-dependent brain-derived neurotrophic factor signaling is required for the antidepressant actions of (2R,6R)-hydroxynorketamine. Proc. Natl Acad. Sci. USA 116, 297–302 (2019).

    Article  CAS  Google Scholar 

  12. Saxton, R. A. & Sabatini, D. M. mTOR signaling in growth, metabolism, and disease. Cell 169, 361–371 (2017).

    Article  CAS  Google Scholar 

  13. Sonenberg, N. & Hinnebusch, A. G. Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell 136, 731–745 (2009).

    Article  CAS  Google Scholar 

  14. Miller, O. H., Moran, J. T. & Hall, B. J. Two cellular hypotheses explaining the initiation of ketamine’s antidepressant actions: direct inhibition and disinhibition. Neuropharmacology 100, 17–26 (2016).

    Article  CAS  Google Scholar 

  15. Kavalali, E. T. & Monteggia, L. M. The ketamine metabolite 2R,6R-hydroxynorketamine blocks NMDA receptors and impacts downstream signaling linked to antidepressant effects. Neuropsychopharmacology 43, 221–222 (2018).

    Article  CAS  Google Scholar 

  16. Suzuki, K., Nosyreva, E., Hunt, K. W., Kavalali, E. T. & Monteggia, L. M. Effects of a ketamine metabolite on synaptic NMDAR function. Nature 546, E1–E3 (2017).

    Article  CAS  Google Scholar 

  17. Paul, R. K. et al. (R,S)-Ketamine metabolites (R,S)-norketamine and (2S,6S)-hydroxynorketamine increase the mammalian target of rapamycin function. Anesthesiology 121, 149–159 (2014).

    Article  CAS  Google Scholar 

  18. Adaikkan, C., Taha, E., Barrera, I., David, O. & Rosenblum, K. Calcium/calmodulin-dependent protein kinase II and eukaryotic elongation factor 2 kinase pathways mediate the antidepressant action of ketamine. Biol. Psychiatry 84, 65–75 (2018).

    Article  CAS  Google Scholar 

  19. Miller, O. H., Grabole, N., Wells, I. & Hall, B. Genome-wide translating mRNA analysis following ketamine reveals novel targets for antidepressant treatment. Preprint at https://doi.org/10.1101/254904 (2018).

  20. Murrough, J. W. et al. Rapid and longer-term antidepressant effects of repeated ketamine infusions in treatment-resistant major depression. Biol. Psychiatry 74, 250–256 (2013).

    Article  CAS  Google Scholar 

  21. Aguilar-Valles, A., Matta-Camacho, E. & Sonenberg, N. in The Oxford Handbook of Neuronal Protein Synthesis (ed. Wayne Sossin) (Oxford Univ. Press, 2018).

  22. Banko, J. L. et al. The translation repressor 4E-BP2 is critical for eIF4F complex formation, synaptic plasticity, and memory in the hippocampus. J. Neurosci. 25, 9581–9590 (2005).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  24. 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  Google Scholar 

  25. Duman, R. S., Sanacora, G. & Krystal, J. H. Altered connectivity in depression: GABA and glutamate neurotransmitter deficits and reversal by novel treatments. Neuron 102, 75–90 (2019).

    Article  CAS  Google Scholar 

  26. Ng, L. H. L. et al. Ketamine and selective activation of parvalbumin interneurons inhibit stress-induced dendritic spine elimination. Transl. Psychiatry 8, 272 (2018).

    Article  CAS  Google Scholar 

  27. 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  Google Scholar 

  28. Moda-Sava, R. N. et al. Sustained rescue of prefrontal circuit dysfunction by antidepressant-induced spine formation. Science 364, eaat8078 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  ADS  Google Scholar 

  30. Bidinosti, M. et al. Postnatal deamidation of 4E-BP2 in brain enhances its association with raptor and alters kinetics of excitatory synaptic transmission. Mol. Cell 37, 797–808 (2010).

    Article  CAS  Google Scholar 

  31. Gkogkas, C. G. et al. Autism-related deficits via dysregulated eIF4E-dependent translational control. Nature 493, 371–377 (2013).

    Article  CAS  ADS  Google Scholar 

  32. Jernigan, C. S. et al. The mTOR signaling pathway in the prefrontal cortex is compromised in major depressive disorder. Prog. Neuropsychopharmacol. Biol. Psychiatry 35, 1774–1779 (2011).

    Article  CAS  Google Scholar 

  33. Abdallah, C. et al. Rapamycin, an immunosuppressant and mTORC1 inhibitor, triples the antidepressant response rate of ketamine at 2 weeks following treatment: a double-blind, placebo-controlled, cross-over, randomized clinical trial. Preprint at https://doi.org/10.1101/500959 (2018).

  34. Tsukiyama-Kohara, K. et al. Adipose tissue reduction in mice lacking the translational inhibitor 4E-BP1. Nat. Med. 7, 1128–1132 (2001).

    Article  CAS  Google Scholar 

  35. Bodnoff, S. R., Suranyi-Cadotte, B., Aitken, D. H., Quirion, R. & Meaney, M. J. The effects of chronic antidepressant treatment in an animal model of anxiety. Psychopharmacology (Berl.) 95, 298–302 (1988).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank A. Sylvestre, A. Lafrance, I. Harvey and T. Degenhard for technical assistance. A.A.-V. was a recipient of FRQS (Fond Recherche Québec-Santé) and CIHR (Canadian Institutes for Health Research) postdoctoral fellowships. D.D.G. is a recipient of FRQS and CIHR postdoctoral fellowships. J.-C.L. is the recipient of the Canada Research Chair in Cellular and Molecular Neurophysiology. This work was supported by a CIHR Foundation grant to N. Sonenberg (FND-148423), and a CIHR project grant to J.-C.L. (PJT-153311).

Author information

Authors and Affiliations

Authors

Contributions

A.A.-V. designed the study, performed experiments, supervised students, analysed data and wrote the manuscript. D.D.G. performed behavioural experiments and data analyses. E.M.-C. performed western blot experiments, supervised students and edited the manuscript. M.J.E. performed electrophysiological experiments. A.K. performed electrophysiological experiments and data analyses. A.S. performed immunohistochemistry experiments and contributed to the writing of the manuscript. M.L.-C. and A.T.-B. contributed to behavioural tests. S.B. performed data analyses. G.M.R., S.S. and N. Salmaso contributed to experimental design, data analysis and manuscript editing. G.G., J.-C.L. and N. Sonenberg supervised students, designed experiments and contributed to writing the manuscript.

Corresponding authors

Correspondence to Argel Aguilar-Valles or Nahum Sonenberg.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature thanks Kimberly Raab-Graham and the other, anonymous, reviewer(s) 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 figures and tables

Extended Data Fig. 1 Expression of 4E-BP1 and 4E-BP2 in the prefrontal cortex and hippocampus.

a, Representative western blot analysis (from 3 independent experiments) of 4E-BP1 and 4E-BP2 in the prefrontal cortex and hippocampus of wild-type (Eif4ebp1/2+/+ +/+), 4E-BP1 knock out (Eif4ebp1−/−), 4E-BP2 knock out (Eif4ebp2−/−), and 4E-BP1/2 double knock out mice (Eif4ebp1/2−/−) (uncropped blots are presented in Supplementary Information). b, c, Representative confocal images of immunostaining for 4E-BP2 (green), CAMK2α (red)/GAD67 (magenta) in the hippocampus. The arrows indicate area showed in the insert, which is a higher magnification of the images. Insert scale bar: 10 μm. Brain sections were additionally counterstained with DAPI (blue). d, Representative confocal images of immunostaining for 4E-BP1 (green), GFAP (red) in the hippocampus. Brain sections were additionally counterstained with DAPI (blue). The white box indicates area shown in e. f, Representative confocal image of immunostaining for 4E-BP1 (green) and CD11b (magenta). Brain sections were additionally counterstained with DAPI (blue). All immunofluorescence experiments were repeated in 3 independent experiments.

Extended Data Fig. 2 4E-BP1 and 4E-BP2 are required for the antidepressant effect of ketamine.

a, A group of Eif4ebp1/2+/+, Eif4ebp1−/−, and Eif4ebp2−/− mice were treated with saline or ketamine (IP, 10 mg kg−1) and tested 24 h after injection in the FST (2-way ANOVA and Fisher’s LSD, n = 11, 13, 11, 12, 4, 7 from left to right). b, Wild-type (Eif4ebp1/2+/+), Eif4ebp1−/− and Eif4ebp2−/− mice were treated with saline or ketamine (IP, 10 or 30 mg kg−1) and tested in the FST 6 days after treatment (2-way ANOVA and Fisher’s LSD, n = 13, 19, 7, 11, 10, 9, 16, 14, 6 from left to right). c, A different group of mice treated as in a was tested in the open field, and general activity was measured as distance travelled (2-way ANOVA, n = 8, 8, 8, 6, 6, 5, 6, 6, 7 from left to right). d, Latency to feed in the home cage was measured immediately after the NSF task in Eif4ebp1/2+/+ +/+, Eif4ebp1−/− and Eif4ebp2−/− mice, treated with saline or Ket (corresponding to Fig. 1b; 2-way ANOVA). e, Latency to feed in the home cage was also measured after the NSF task in the three mouse strains, following saline or (2R,6R)-HNK treatment (corresponding to Fig. 1c; 2-way ANOVA). f, In the same group of mice as in e (and Fig. 1c) average distance travelled during the NSF test was measured (expressed as an average per minute to correct for differential time spent in the arena; 2-way ANOVA). g, Mice of the three genotypes were also tested for immobility in the tail suspension test, 1 h after saline or Ket treatments (2-way ANOVA and Fisher’s LSD, n = 8, 9, 9, 10, 10, 7 from left to right). h, Eif4ebp1/2+/+ +/+, Eif4ebp1−/− and Eif4ebp2−/− mice were treated with fluoxetine (Fluox, IP, 3 mg kg−1, 0.5 h) and tested in the FST (2-way ANOVA and Fisher’s LSD, n = 8, 10, 9, 8, 6, 6 from left to right). Mean ± s.e.m.; * P < 0.05; ** P < 0.01 vs. Sal-treated mice of the same genotype.

Source data

Extended Data Fig. 3 Generation of Eif4ebp1 and Eif4ebp2 floxed genes in mice.

Schematic representation of the targeting construct used for the generation of Eif4ebp1 and Eif4ebp2 knockout mice. FRT (FLPase recognition target) and loxP sequences are indicated. Positive (PGK-neo) selection markers are indicated in the schemes on the Neomycin cassette (Neo, neomycin). Numbered grey boxes represent exons in the Eif4ebp1 and Eif4ebp2 genes. The Eif4ebp1 and Eif4ebp2 neomycin alleles were produced by homologous recombination. FLPase was used to generate the Eif4ebp1 and Eif4ebp2 floxed alleles. The genotyping PCR primers are indicated for each construct, as well as the length of the PCR product.

Extended Data Fig. 4 Effects of cell-specific mutation of Eif4ebp1 and Eif4ebp2 on home cage feeding, locomotion and forced swim test.

a, Latency to feed in the home cage was measured 1 h after saline or Ket (IP, 10 mg kg−1) treatments in control, mice lacking Eif4ebp1 in CAMK2α+ (Eif4ebp1Ex) or GAD2+ neurons (Eif4ebp1In), or mice mutant for Eif4ebp2 in excitatory (Eif4ebp2Ex) or inhibitory (Eif4ebp2In) neurons (corresponding to mice in Fig. 2i; 2-way ANOVA and Dunnett’s). b, In these same mice, distance travelled during the NSF task was measured. Data were expressed as average distance travelled per minute to account for differences in latency to feed (2-way ANOVA and Dunnett’s). c, Control, Eif4ebp1Ex, Eif4ebp1In, Eif4ebp2Ex and Eif4ebpIn mice were tested 6 days (144 h) after saline or Ket treatment (IP, 10 mg kg−1) in the FST (2-way ANOVA and Fisher’s LSD, n = 17, 15, 7, 10, 12, 17, 8, 9, 5, 7 from left to right). Mean ± s.e.m.; * P < 0.05 vs. control (main effect of genotype, in a) or vs. Sal-treated mice of the same genotype.

Source data

Extended Data Fig. 5 Effects of 4E-BP2 mutation in inhibitory neurons on amplitude of sEPSC and mIPSC.

a, Cumulative distribution of sEPSC amplitude in CA1 pyramidal neurons of control and Eif4ebp2In mice treated with saline or Ket (mice recorded in Fig. 3; Kolmogorov–Smirnov test). b, Amplitude of sEPSCs recorded from pyramidal neurons in CA1 of control and Eif4ebp2In mice (mice recorded in Fig. 3; 2-way ANOVA). c, Cumulative distribution of mIPSC amplitude in CA1 pyramidal neurons from mice treated as in a (mice recorded in Fig. 4; Kolmogorov–Smirnov test). d, Amplitude of mIPSC in CA1 pyramidal neurons in mice treated as in a (mice recorded in Fig. 4; 2-way ANOVA). Mean ± s.e.m.; * P < 0.05 vs. control-Sal.

Source data

Extended Data Fig. 6 The absence of 4E-BP2 in inhibitory neurons does not affect mEPSCs in pyramidal neurons.

a, Representative whole-cell recordings of mEPSCs in CA1 pyramidal neurons of mice treated with saline or Ket 24 h before (IP, 10 mg kg−1). Control and Eif4ebp2In mice were used. b, Cumulative inter-event interval (Kolmogorov–Smirnov test, n = 7, 11, 6, 17 from left from right) and c, frequency of mEPSCs measured in a (2-way ANOVA). d, Cumulative (Kolmogorov–Smirnov test) and e, average distribution of mEPSC amplitude measured in a (2-way ANOVA). Mean ± s.e.m.; * P < 0.05 vs. Eif4ebp2In-Sal.

Source data

Supplementary information

Supplementary Information

Uncropped Western blots from Extended Data Figure 1a. Uncropped images of the films for the Western blot analyses of 4E-BP1 and 4E-BP2 in the prefrontal cortex and hippocampus of wildtype (Eif4ebp1/2+/+ +/+), 4E-BP1 knock out (Eif4ebp1-/-), 4E-BP2 knock out (Eif4ebp2-/-), and 4E-BP1/2 double knock out mice (Eif4ebp1/2-/-). Images for 4E-BP1, 4E-BP2 and GAPDH are presented.

Reporting Summary

Supplementary Table 1

Details of statistical analyses.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Aguilar-Valles, A., De Gregorio, D., Matta-Camacho, E. et al. Antidepressant actions of ketamine engage cell-specific translation via eIF4E. Nature 590, 315–319 (2021). https://doi.org/10.1038/s41586-020-03047-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-03047-0

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing