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:

Ketamine triggers a switch in excitatory neuronal activity across neocortex

Abstract

The brain can become transiently disconnected from the environment while maintaining vivid, internally generated experiences. This so-called ‘dissociated state’ can occur in pathological conditions and under the influence of psychedelics or the anesthetic ketamine (KET). The cellular and circuit mechanisms producing the dissociative state remain poorly understood. We show in mice that KET causes spontaneously active neurons to become suppressed while previously silent neurons become spontaneously activated. This switch occurs in all cortical layers and different cortical regions, is induced by both systemic and cortical application of KET and is mediated by suppression of parvalbumin and somatostatin interneuron activity and inhibition of NMDA receptors and HCN channels. Combined, our results reveal two largely non-overlapping cortical neuronal populations—one engaged in wakefulness, the other contributing to the KET-induced brain state—and may lay the foundation for understanding how the brain might become disconnected from the surrounding environment while maintaining internal subjective experiences.

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: KET induces a switch in the active population of L2/3 neurons in S1.
Fig. 2: KET induces a neuronal switch in the input and output cortical layers of S1.
Fig. 3: KET induces widespread reorganization of L2/3 activity across the neocortex.
Fig. 4: Local KET delivery in S1 induces a neuronal switch in the active population of L2/3 neurons.
Fig. 5: Systemic and local KET suppresses activity in S1 interneuron subtypes.
Fig. 6: KET-induced suppression of SST or PV interneurons is necessary for the activity switch in L2/3 pyramidal neurons.
Fig. 7: Inhibition of NMDARs or HCN channels drives a neuronal activity switch.
Fig. 8: Combined inhibition of NMDARs and HCN channels recapitulates the effects of KET administration.

Similar content being viewed by others

Data availability

Figures and associated data will be available to the scientific community via figshare.

References

  1. Sleigh, J., Harvey, M., Voss, L. & Denny, B. Ketamine—more mechanisms of action than just NMDA blockade. Trends Anaesth. Crit. Care 4, 76–81 (2014).

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

  3. Krystal, J. H. et al. Subanesthetic effects of the noncompetitive NMDA antagonist, ketamine, in humans. Psychotomimetic, perceptual, cognitive, and neuroendocrine responses. Arch. Gen. Psychiatry 51, 199–214 (1994).

    Article  CAS  Google Scholar 

  4. Li, D. & Mashour, G. A. Cortical dynamics during psychedelic and anesthetized states induced by ketamine. Neuroimage 196, 32–40 (2019).

    Article  CAS  Google Scholar 

  5. Akeju, O. et al. Electroencephalogram signatures of ketamine anesthesia-induced unconsciousness. Clin. Neurophysiol. 127, 2414–2422 (2016).

    Article  Google Scholar 

  6. Schwartz, M. S., Virden, S. & Scott, D. F. Effects of ketamine on the electroencephalograph. Anaesthesia 29, 135–140 (1974).

    Article  CAS  Google Scholar 

  7. Corssen, G. & Domino, E. F. Dissociative anesthesia: further pharmacologic studies and first clinical experience with the phencyclidine derivative CI-581. Anesth. Analg. 45, 29–40 (1966).

    Article  CAS  Google Scholar 

  8. Domino, E. F. Taming the ketamine tiger. 1965. Anesthesiology 113, 678–684 (2010).

    Article  Google Scholar 

  9. Domino, E. F., Chodoff, P. & Corssen, G. Pharmacologic effects of Ci-581, a new dissociative anesthetic, in man. Clin. Pharmacol. Ther. 6, 279–291 (1965).

    Article  CAS  Google Scholar 

  10. Vesuna, S. et al. Deep posteromedial cortical rhythm in dissociation. Nature 586, 87–94 (2020).

    Article  CAS  Google Scholar 

  11. Abdallah, C. G. et al. The effects of ketamine on prefrontal glutamate neurotransmission in healthy and depressed subjects. Neuropsychopharmacology 43, 2154–2160 (2018).

    Article  CAS  Google Scholar 

  12. Moghaddam, B., Adams, B., Verma, A. & Daly, D. Activation of glutamatergic neurotransmission by ketamine: a novel step in the pathway from NMDA receptor blockade to dopaminergic and cognitive disruptions associated with the prefrontal cortex. J. Neurosci. 17, 2921–2927 (1997).

    Article  CAS  Google Scholar 

  13. Ali, F. et al. Ketamine disinhibits dendrites and enhances calcium signals in prefrontal dendritic spines. Nat. Commun. 11, 72 (2020).

    Article  CAS  Google Scholar 

  14. 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  Google Scholar 

  15. Patel, I. M. & Chapin, J. K. Ketamine effects on somatosensory cortical single neurons and on behavior in rats. Anesth. Analg. 70, 635–644 (1990).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  17. 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 

  18. MacDonald, J. F., Miljkovic, Z. & Pennefather, P. Use-dependent block of excitatory amino acid currents in cultured neurons by ketamine. J. Neurophysiol. 58, 251–266 (1987).

  19. Chen, X., Shu, S. & Bayliss, D. A. HCN1 channel subunits are a molecular substrate for hypnotic actions of ketamine. J. Neurosci. 29, 600–609 (2009).

    Article  CAS  Google Scholar 

  20. 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 

  21. Freund, T. F. & Katona, I. Perisomatic inhibition. Neuron 56, 33–42 (2007).

    Article  CAS  Google Scholar 

  22. Chiu, C. Q. et al. Compartmentalization of GABAergic inhibition by dendritic spines. Science 340, 759–762 (2013).

    Article  CAS  Google Scholar 

  23. Pi, H.-J. et al. Cortical interneurons that specialize in disinhibitory control. Nature 503, 521–524 (2013).

    Article  CAS  Google Scholar 

  24. Cryan, J. F., Mombereau, C. & Vassout, A. The tail suspension test as a model for assessing antidepressant activity: review of pharmacological and genetic studies in mice. Neurosci. Biobehav. Rev. 29, 571–625 (2005).

    Article  CAS  Google Scholar 

  25. Corne, S. J. & Pickering, R. W. A possible correlation between drug-induced hallucinations in man and a behavioural response in mice. Psychopharmacologia 11, 65–78 (1967).

    Article  CAS  Google Scholar 

  26. Broekkamp, C. L., Rijk, H. W., Joly-Gelouin, D. & Lloyd, K. L. Major tranquillizers can be distinguished from minor tranquillizers on the basis of effects on marble burying and swim-induced grooming in mice. Eur. J. Pharmacol. 126, 223–229 (1986).

    Article  CAS  Google Scholar 

  27. Bouet, V. et al. The adhesive removal test: a sensitive method to assess sensorimotor deficits in mice. Nat. Protoc. 4, 1560–1564 (2009).

    Article  CAS  Google Scholar 

  28. Blain-Moraes, S., Lee, U., Ku, S., Noh, G. & Mashour, G. A. Electroencephalographic effects of ketamine on power, cross-frequency coupling, and connectivity in the alpha bandwidth. Front. Syst. Neurosci. 8, 114 (2014).

    Article  Google Scholar 

  29. Sachidhanandam, S., Sreenivasan, V., Kyriakatos, A., Kremer, Y. & Petersen, C. C. H. Membrane potential correlates of sensory perception in mouse barrel cortex. Nat. Neurosci. 16, 1671–1677 (2013).

    Article  CAS  Google Scholar 

  30. Dana, H. et al. Thy1-GCaMP6 transgenic mice for neuronal population imaging in vivo. PLoS ONE 9, e108697 (2014).

    Article  Google Scholar 

  31. Lu, J. et al. Role of endogenous sleep-wake and analgesic systems in anesthesia. J. Comp. Neurol. 508, 648–662 (2008).

    Article  Google Scholar 

  32. Nelson, C. L., Burk, J. A., Bruno, J. P. & Sarter, M. Effects of acute and repeated systemic administration of ketamine on prefrontal acetylcholine release and sustained attention. Psychopharmacology (Berl.) 161, 168–179 (2002).

    Article  CAS  Google Scholar 

  33. Långsjö, J. W. et al. S-ketamine anesthesia increases cerebral blood flow in excess of the metabolic needs in humans. Anesthesiology 103, 258–268 (2005).

    Article  Google Scholar 

  34. Ouelhazi, A. et al. Effects of ketamine on orientation selectivity and variability of neuronal responses in primary visual cortex. Brain Res. 1725, 146462 (2019).

    Article  CAS  Google Scholar 

  35. 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  Google Scholar 

  36. Urban, D. J. & Roth, B. L. DREADDs (designer receptors exclusively activated by designer drugs): chemogenetic tools with therapeutic utility. Annu. Rev. Pharmacol. Toxicol. 55, 399–417 (2015).

    Article  CAS  Google Scholar 

  37. Miller, R. D. et al. Miller’s Anesthesia (Elsevier Health Sciences, 2014).

  38. Mashour, G. A. Top-down mechanisms of anesthetic-induced unconsciousness. Front. Syst. Neurosci. 8, 115 (2014).

    Article  Google Scholar 

  39. Constantinidis, C. & Goldman-Rakic, P. S. Correlated discharges among putative pyramidal neurons and interneurons in the primate prefrontal cortex. J. Neurophysiol. 88, 3487–3497 (2002).

    Article  Google Scholar 

  40. Rao, S. G., Williams, G. V. & Goldman-Rakic, P. S. Isodirectional tuning of adjacent interneurons and pyramidal cells during working memory: evidence for microcolumnar organization in PFC. J. Neurophysiol. 81, 1903–1916 (1999).

    Article  CAS  Google Scholar 

  41. Alexander, L. et al. Fractionating blunted reward processing characteristic of anhedonia by over-activating primate subgenual anterior cingulate cortex. Neuron 101, 307–320 (2019).

    Article  CAS  Google Scholar 

  42. Zhou, H. et al. Ketamine reduces aversion in rodent pain models by suppressing hyperactivity of the anterior cingulate cortex. Nat. Commun. 9, 3751 (2018).

    Article  Google Scholar 

  43. Tokay, T. et al. HCN1 channels constrain DHPG-induced LTD at hippocampal Schaffer collateral-CA1 synapses. Learn. Mem. 16, 769–776 (2009).

    Article  CAS  Google Scholar 

  44. Zhang, K. Essential roles of AMPA receptor GluA1 phosphorylation and presynaptic HCN channels in fast-acting antidepressant responses of ketamine. Sci. Signal. 9, ra123 (2016).

    Article  Google Scholar 

  45. Cichon, J. & Gan, W.-B. Branch-specific dendritic Ca2+ spikes cause persistent synaptic plasticity. Nature 520, 180–185 (2015).

    Article  CAS  Google Scholar 

  46. Li, W., Ma, L., Yang, G. & Gan, W.-B. REM sleep selectively prunes and maintains new synapses in development and learning. Nat. Neurosci. 20, 427–437 (2017).

    Article  CAS  Google Scholar 

  47. Zhou, Y. REM sleep promotes experience-dependent dendritic spine elimination in the mouse cortex. Nat. Commun. 11, 4819 (2020).

    Article  CAS  Google Scholar 

  48. Suzuki, M. & Larkum, M. E. Dendritic calcium spikes are clearly detectable at the cortical surface. Nat. Commun. 8, 276 (2017).

    Article  Google Scholar 

  49. David, F. et al. Suppression of hyperpolarization-activated cyclic nucleotide-gated channel function in thalamocortical neurons prevents genetically determined and pharmacologically induced absence seizures. J. Neurosci. 38, 6615–6627 (2018).

    Article  CAS  Google Scholar 

  50. Labarrera, C. Adrenergic modulation regulates the dendritic excitability of layer 5 pyramidal neurons in vivo. Cell Rep. 23, 1034–1044 (2018).

    Article  CAS  Google Scholar 

  51. Wasilczuk, A. Z., Proekt, A., Kelz, M. B. & McKinstry-Wu, A. R. High-density electroencephalographic acquisition in a rodent model using low-cost and open-source resources. J. Vis. Exp. 117, 54908 (2016).

  52. Siegle, J. H. Open Ephys: an open-source, plugin-based platform for multichannel electrophysiology. J. Neural Eng. 14, 045003 (2017).

    Article  Google Scholar 

  53. Hudson, A. E., Calderon, D. P., Pfaff, D. W. & Proekt, A. Recovery of consciousness is mediated by a network of discrete metastable activity states. Proc. Natl Acad. Sci. USA 111, 9283–9288 (2014).

    Article  Google Scholar 

  54. Cichon, J. et al. Imaging neuronal activity in the central and peripheral nervous systems using new Thy1.2-GCaMP6 transgenic mouse lines. J. Neurosci. Methods 334, 108535 (2020).

    Article  CAS  Google Scholar 

  55. Chen, Q. et al. Imaging neural activity using Thy1-GCaMP transgenic mice. Neuron 76, 297–308 (2012).

    Article  CAS  Google Scholar 

  56. Cichon, J., Blanck, T. J. J., Gan, W.-B. & Yang, G. Activation of cortical somatostatin interneurons prevents the development of neuropathic pain. Nat. Neurosci. 20, 1122–1132 (2017).

    Article  CAS  Google Scholar 

  57. Yang, G. et al. Sleep promotes branch-specific formation of dendritic spines after learning. Science 344, 1173–1178 (2014).

    Article  CAS  Google Scholar 

  58. Adler, A., Zhao, R., Shin, M. E., Yasuda, R. & Gan, W.-B. Somatostatin-expressing interneurons enable and maintain learning-dependent sequential activation of pyramidal neurons. Neuron 102, 202–216 (2019).

    Article  CAS  Google Scholar 

  59. Zhao, R. et al. Neuropathic pain causes pyramidal neuronal hyperactivity in the anterior cingulate cortex. Front. Cell. Neurosci. 12, 107 (2018).

    Article  Google Scholar 

Download references

Acknowledgements

We thank W. Gan for insightful discussions, M. Fina for animal management and genotyping and S. Nikonov of the University of Pennsylvania Vision Research Center for technical and imaging support. This work was supported by the University of Pennsylvania Department of Anesthesia Dripps scholarship and FAER MRTG to J.C.; the Training Program in Neuroengineering and Medicine T32NS091006 to A.W.; National Institutes of Health (NIH) R01GM124023-01A1 and 5R01NS113366 to A.P.; NIH R01GM088156 to M.B.K.; and NIH R01EY020765 to D.C.

Author information

Authors and Affiliations

Authors

Contributions

J.C. initiated the project. J.C. performed imaging and behavioral experiments. J.C. processed and analyzed the data, with input from A.P., M.B.K. and D.C. A.W. processed and analyzed EEG recordings. J.C., A.P., L.L.L. and M.B.K wrote the paper, with input from D.C.

Corresponding authors

Correspondence to Joseph Cichon or Alex Proekt.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Neuroscience thanks Jordan Hamm, Ken Solt 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 Subhypnotic ketamine impairs sensorimotor processing but not gross animal movement.

(a, b) Duration effect of single dose of ketamine (KET; 50 mg/kg: 20 mice, 100 mg/kg: 11 mice) or saline (11 mice) on time for complete removal of adhesive from snout (TAR) (a) and forelimb withdrawal response to air puff (b). (c) Example images of marble burial test under different conditions. Note that saline-injected mice bury a fraction of marbles over 30 min (middle image). (d) Summary of results of marble burial test for 16 mice in each treatment group. Mice injected with KET at either 50 or 100 mg/kg fail to bury a single marble (example image in c). Number of marbles buried within 30 min was 0 in KET (50 or 100 mg/kg) vs. 4.1 ± 1 in saline-injected mice. One-sided Kruskal-Wallis with Dunn’s multiple comparisons test, P < 0.001. (e) Tracking of total distance and average speed for 14 mice over 5 min period before and after KET at 50 mg/kg in a standard mouse cage. Two-sided Wilcoxon rank test, P > 0.05 for both parameters. Exact P values in figure. Error bars show s.e.m.

Extended Data Fig. 2 Various ketamine-induced layer 2/3 responses in the forelimb region of primary somatosensory cortex.

(a) Line scatterplots of L2/3 neurons before and after KET at different doses, 50 (49 cells from 4 regions), 100 (37 cells from 3 regions), 150 mg/kg (34 cells from 3 regions) from single animals. Scatterplot of entire L2/3 population for 50 and 100 mg/kg in Fig. 1g. (b) Representative fluorescent traces (left) of individual L2/3 axonal boutons imaged in layer 1 under wakefulness and KET 100 mg/kg. Line scatterplots (right) of individual boutons before and after KET 100 mg/kg reveals emergence of a switch. (c) Difference in calcium activity induced by KET at 10 mg/kg (left, 85 neurons from 3 mice) or 150 mg/kg IP (right, 64 neurons from 2 mice) from baseline wakefulness vs. wakefulness activity. KET at 150 mg/kg (one-sided Pearson correlation: r = −0.62, P = 7.8 × 10−8), but not at 10 mg/kg (r = −0.06, P = 0.56), induced a neuronal switch in L2/3 activity.

Extended Data Fig. 3 GABAergic hypnotic agents induce net suppression of L2/3 pyramidal cell activity.

(a, b) Representative GCaMP6 traces of individual L2/3 neurons under wakefulness and different hypnotic agents (a, sevoflurane (sevo), yellow shaded region; b, midazolam (midaz), green shaded region). (c) (Left) Line scatterplots and (Right) box plots of average spontaneous rate of L2/3 activity before and after volatile (sevo or isoflurane/iso, purple) and midaz hypnosis in S1. Spontaneous activity: Sevo group (71 neurons from 4 mice), awake: 4.0 ± 0.8 vs. Sevo: 1.0 ± 0.4, P = 2.8 × 10−8; Iso group (78 neurons from 7 mice), awake: 2.4 ± 0.5 vs. Iso: 0.1 ± 0.1, P < 1 × 10−15; Midaz group (85 neurons from 3 mice), awake: 2.8 ± 0.6 vs. Midaz: 0.4 ± 0.1, P = 4 × 10−15). (d) Peak ΔF/F0 before and after hypnosis (Sevo: awake: 238 ± 30% vs. Sevo: 47 ± 5%, P = 2.9 × 10−10; Iso: awake: 102 ± 12% vs. Iso: 14 ± 0.5%, P < 1 × 10−15; Midaz: awake: 83 ± 7% vs. Midaz: 30 ± 2%, P = 5.6 × 10−12). Same cells from c. (e) Relative change in L2/3 activity for different hypnotic agents. Unlike KET, GABAergic hypnotics induced a uniform suppression of L2/3 neuronal activity (fraction of cells with reduced activity under GABAergic agents 65–78% vs. KET ~40%). While KET induced substantial activation of the network (~40–45% of cells), GABAergic agents activated fewer than 10% of cells. Box and whisker plot show min to max, centre (median), 25th and 75th percentile box bounds. One-sided Kruskal-Wallis with Dunn’s multiple comparisons test in c, d, ***P < 0.001.

Extended Data Fig. 4 Ketamine-induced activity switch not explained by random changes in activity.

(a–d) Shows joint distribution of activity of L2/3 neurons under normal wakefulness (x-axis) and after ketamine (KET)/saline administration (y-axis). The left column shows the experimentally observed joint probability distribution. Middle column shows the joint distribution expected if activity after KET/saline administration did not depend on baseline activity (Methods). (a) Saline control, (b–d) KET at 10, 50, and 100 mg/kg administered systemically. The right column shows statistical deviations of the observed joint distribution (left column) from one expected if activity during normal wakefulness and after drug administration was independent (middle column). Statistical significance was computed using a two-sided permutation test consisting of shuffling neuron IDs randomly 1000 times. Deviations were statistically significant at P < 0.01 (bootstrap estimates). Red points indicate locations where experimentally observed probability was higher than chance, blue points indicate locations where experimentally observed probability was lower than chance, white points show areas where no statistically significant differences were detected. The regions of statistical deviations directly abutting the x and y axes under KET at 50 and 100 mg/kg indicate that KET preferentially suppresses active neurons and activates a subset of silent neurons.

Extended Data Fig. 5 Layer 4 and 5 calcium responses under various conditions.

(a, b) Representative GCaMP6 traces of individual neurons (a, top for L4, b, bottom for L5) under wakefulness and following saline injection (left, purple shaded region) and KET at 100 mg/kg (right, red shaded region). (c) Line scatterplots (left) and average spontaneous activity (right) of individual L4 and L5 activity before and after saline. Saline did not significantly increase calcium activity compared to wakefulness. L4 (105 neurons from 9 mice), awake: 3.8 ± 0.6 vs. saline: 3.9 ± 0.5, P > 0.99. L5 (97 neurons from 8 mice), awake: 3.4 ± 0.7 vs. saline: 2.9 ± 0.7, P = 0.63. (d) Peak ΔF/F0 in saline-treated mice (same cells from c). L4, awake: 98 ± 7% vs. saline: 90 ± 6%, P = 0.30. L5, awake: 112 ± 12% vs. saline: 95 ± 7%, P = 0.30. Box and whisker plot show min to max, centre (median), 25th and 75th percentile box bounds. One-sided Kruskal-Wallis with Dunn’s multiple comparisons test in c, d.

Extended Data Fig. 6 Ketamine at 100 mg/kg induces a neuronal switch in L2/3 activity across the neocortex.

(a) Two-photon images of L2/3 neurons from various cortical regions in Thy1-GCaMP6f mice. Scale bar: 20 µm. (b) Cartoon of mouse brain in center of panel demarcates imaging locations (teal boxes) across the neocortex. Surrounding traces are representative L2/3 neurons from these regions under wakefulness and KET 100 mg/kg. (c) The effect of KET was negatively correlated with the activity in wakefulness across all imaged cortical regions. M2 (65 from 4 mice; r = −0.64, P = 1.1 × 10−8), M1 (140 from 4 mice; r = −0.56, P = 3.6 × 10−13), vibrissal (v) S1 (65 from 2 mice; r = −0.50, P = 2.5 × 10−5), and V1 (120 from 4 mice; r = −0.55, P = 2.2 × 10−10). (d) L2/3 activity switch index across saline and KET-injected mice for various regions. M2 (P = 0.001), M1 (P = 2.7 × 10−10), vS1 (P = 0.003), and V1 (P = 1.6 × 10−8). Error bars show s.e.m. One-sided Pearson correlations in c, One-sided Kruskal-Wallis with Dunn’s multiple comparisons test in d, ***P < 0.001.

Extended Data Fig. 7 Labelling different inhibitory interneurons subtypes in the somatosensory cortex with GCaMP6 and their responses to ketamine at 100 mg/kg.

(a) Example two-photon imaging fields of interneuron subtypes (SST, pink; PV cells, blue; VIP cells, green) expressing GCaMP6f in the S1. Interneurons outlined with white circles. Scale bar: 20 µm. Representative images carried out on at least 3 animals per group. (b) Pie graphs show the portion of cells increasing (green)/decreasing (red)/no change (black) following saline administration for each interneuron subtypes. (c, d) Comparison between spontaneous rate of activity (c) and peak signals (d) across interneuron subtypes following KET at 100 mg/kg. SST: 82 cells from 6 mice, PV: 84 cells from 4 mice, VIP: 85 mice from 5 mice (cell numbers per subtype italicized in c). Note the strong reduction of spontaneous interneuron activity and peak signals (One-sided Kruskal-Wallis with Dunn’s multiple comparisons for c: P < 0.05 for SST and P < 0.001 for all other comparisons; d: P < 0.001 for all comparisons). Exact P values in figure. Error bars show s.e.m.

Extended Data Fig. 8 Systemic administration of CNO does not impair ketamine’s neuronal activity switch in L2/3 neurons of S1.

(a) Coronal slices (left) of S1 region from interneuron-specific Cre-positive mice expressing DREADD variant hM3D(Gq)-mCherry. Scale bar: 100 µm. (Right) Number of hM3D(Gq)-mCherry positive cells in a 1 mm2 region of forelimb S1 (slice thickness 30 µm). 15 (SST), 16 (PV), and 16 (VIP) S1 slices from 4 mice in each group. Box and whisker plot show min to max, centre (median), 25th and 75th percentile box bounds. (b) Coronal slice of S1 region from mouse expressing AAV-hSynapsin-1-tdTomato. Scale bar: 100 µm. (c) Representative traces of L2/3 pyramidal cells under wakefulness, CNO, then KET at 50 mg/kg in mice expressing AAV-hSynapsin-1-tdTomato. (d) L2/3 cell activity in the setting of CNO vs. its response to KET at 50 mg/kg in mice expressing AAV-hSyn1-tdTomato (97 neurons from 3 mice). KET at 50 mg/kg induced a strong neuronal switch (Pearson correlation r = −0.68, P = 2 × 10−14) and a significant elevation in activity switch index (KET 4.9 ± 0.5) whereas CNO injection did not (1.8 ± 0.2, P > 0.05). (e) Spontaneous rate of activity of L2/3 pyramidal neurons before/baseline (BL), after interneuron subtype activation (CNO activation of hM3D(Gq)) and following KET treatment at 50 mg/kg. Same cells from Fig. 6d. One-sided Kruskal-Wallis, P < 0.001 with Dunn’s multiple comparisons and P values listed in graph. Representative images and traces carried out on at least 3 animals per group. Error bars show s.e.m. *P < 0.05, **P < 0.01, ***P < 0.001.

Extended Data Fig. 9 Ketamine-induced neuronal switch attenuated in the presence of local midazolam to S1.

(a) Representative GCaMP6 traces of L2/3 pyramidal neurons under wakefulness and local (L-) midazolam at 100 (79 neurons from 3 mice) and 500 (125 neurons from 3 mice) µM. (b) L-midazolam reduced the normalized calcium response as compared to wakefulness indicating the likely facilitation of spontaneous GABAA receptor activity. Same cells from a. Error bars show s.e.m. (c) Scatterplot of L2/3 activity in the presence of L-midazolam vs. KET at 50 mg/kg. KET-induced neuronal switch was attenuated in the presence of L-midazolam (100 µM: r = 0.06, P = 0.58; 500 µM: r = −0.14, P = 0.13). Two-sided Wilcoxon matched-pairs signed-rank test in b, Pearson correlations in c. ***P < 0.001.

Extended Data Fig. 10 Layer 2/3 calcium responses under various drug conditions.

(a) Representative GCaMP6 traces of individual L2/3 neurons from S1 under wakefulness and following local (nifedipine at 100 µM and mecamylamine at 100 µM) or systemic (fentanyl at 0.3 mg/kg IP and ephedrine at 20 mg/kg IP) drug injection (color shaded regions). These potential KET molecular targets were ineffective inducers of an activity switch in L2/3 neurons (see Fig. 7). (b) Experimental design (top) for sequential local (L-) blockade of NMDA-R and AMPA-R in vivo. (Bottom) Representative GCaMP6 traces of L2/3 pyramidal neurons under wakefulness followed over 40 min followed by sequential delivery of L-MK801 at 100 µM and then DNQX at 100 µM (73 cells from 2 mice). (c) Compared to baseline recording in wakefulness, L-MK801 significantly reduced (61%) the L2/3 activity. The remaining cohort of active cells in the presence of L-MK801 could be completely silenced in the presence of L-DNQX. Same cells from b. One-sided Kruskal-Wallis with Dunn’s multiple comparisons test in c. (d, e) Scatter plots of L2/3 activity modulated by either L-MK801 (73 neurons from 2 mice, one-sided Pearson correlation r = −0.57, P = 1.6 × 10−7) or systemic KET at 50 mg/kg (94 neurons from 3 mice, r = −0.86, P = < 1 × 10−15) vs. L-DNQX effect. L-DNQX blocks newly activated cells induced by KET. Error bars show s.e.m. ***P < 0.001.

Supplementary information

Supplementary Information

Supplementary Figs. 1–3

Reporting Summary

Supplementary Video

Mice held in tail suspension exhibit escape-like behaviors (pre-KET). Post-KET (50 mg kg−1), mice no longer attempt to escape, and truncal movement is minimal. Meanwhile, KET induces a sustained vertical head twitch response.

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

Cichon, J., Wasilczuk, A.Z., Looger, L.L. et al. Ketamine triggers a switch in excitatory neuronal activity across neocortex. Nat Neurosci 26, 39–52 (2023). https://doi.org/10.1038/s41593-022-01203-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-022-01203-5

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