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:

Deep posteromedial cortical rhythm in dissociation

Abstract

Advanced imaging methods now allow cell-type-specific recording of neural activity across the mammalian brain, potentially enabling the exploration of how brain-wide dynamical patterns give rise to complex behavioural states1,2,3,4,5,6,7,8,9,10,11,12. Dissociation is an altered behavioural state in which the integrity of experience is disrupted, resulting in reproducible cognitive phenomena including the dissociation of stimulus detection from stimulus-related affective responses. Dissociation can occur as a result of trauma, epilepsy or dissociative drug use13,14, but despite its substantial basic and clinical importance, the underlying neurophysiology of this state is unknown. Here we establish such a dissociation-like state in mice, induced by precisely-dosed administration of ketamine or phencyclidine. Large-scale imaging of neural activity revealed that these dissociative agents elicited a 1–3-Hz rhythm in layer 5 neurons of the retrosplenial cortex. Electrophysiological recording with four simultaneously deployed high-density probes revealed rhythmic coupling of the retrosplenial cortex with anatomically connected components of thalamus circuitry, but uncoupling from most other brain regions was observed—including a notable inverse correlation with frontally projecting thalamic nuclei. In testing for causal significance, we found that rhythmic optogenetic activation of retrosplenial cortex layer 5 neurons recapitulated dissociation-like behavioural effects. Local retrosplenial hyperpolarization-activated cyclic-nucleotide-gated potassium channel 1 (HCN1) pacemakers were required for systemic ketamine to induce this rhythm and to elicit dissociation-like behavioural effects. In a patient with focal epilepsy, simultaneous intracranial stereoencephalography recordings from across the brain revealed a similarly localized rhythm in the homologous deep posteromedial cortex that was temporally correlated with pre-seizure self-reported dissociation, and local brief electrical stimulation of this region elicited dissociative experiences. These results identify the molecular, cellular and physiological properties of a conserved deep posteromedial cortical rhythm that underlies states of dissociation.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Multiregional widefield imaging of cortical activity reveals ketamine-induced retrosplenial rhythm.
Fig. 2: Retrosplenial in vivo electrophysiology.
Fig. 3: Ketamine- and optogenetically elicited dissociation-like behavioural phenotype.
Fig. 4: RSP HCN1 channels are required for ketamine to induce deep rhythm and dissociation-like behaviour.
Fig. 5: Human PMC rhythm and self-reported dissociation.

Similar content being viewed by others

Data availability

The datasets generated and analysed are available from the corresponding author upon reasonable request and at https://www.optogenetics.orgSource data are provided with this paper.

Code availability

Code used for data processing and analysis is available from the corresponding author upon reasonable request.

References

  1. Ferezou, I. et al. Spatiotemporal dynamics of cortical sensorimotor integration in behaving mice. Neuron 56, 907–923 (2007).

    Article  CAS  PubMed  Google Scholar 

  2. Mohajerani, M. H. et al. Spontaneous cortical activity alternates between motifs defined by regional axonal projections. Nat. Neurosci. 16, 1426 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Musall, S., Kaufman, M. T., Juavinett, A. L., Gluf, S. & Churchland, A. K. Single-trial neural dynamics are dominated by richly varied movements. Nat. Neurosci. 22, 1677–1686 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Kauvar, I. V. et al. Cortical observation by synchronous multifocal optical sampling reveals widespread population encoding of actions. Neuron 107, 351–367.e19 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Guo, Z. V. et al. Flow of cortical activity underlying a tactile decision in mice. Neuron 81, 179–194 (2014).

    Article  CAS  PubMed  Google Scholar 

  6. Wekselblatt, J. B., Flister, E. D., Piscopo, D. M. & Niell, C. M. Large-scale imaging of cortical dynamics during sensory perception and behavior. J. Neurophysiol. 115, 2852–2866 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Ma, Y. et al. Resting-state hemodynamics are spatiotemporally coupled to synchronized and symmetric neural activity in excitatory neurons. Proc. Natl Acad. Sci. USA 113, E8463–E8471 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Allen, W. E. et al. Global representations of goal-directed behavior in distinct cell types of mouse neocortex. Neuron 94, 891–907.e6 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Makino, H. et al. Transformation of cortex-wide emergent properties during motor learning. Neuron 94, 880–890.e8 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Chen, T.-W., Li, N., Daie, K. & Svoboda, K. A map of anticipatory activity in mouse motor cortex. Neuron 94, 866–879.e4 (2017).

    Article  CAS  PubMed  Google Scholar 

  11. Xiao, D. et al. Mapping cortical mesoscopic networks of single spiking cortical or sub-cortical neurons. eLife 6, e19976 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  12. Gilad, A., Gallero-Salas, Y., Groos, D. & Helmchen, F. Behavioral strategy determines frontal or posterior location of short-term memory in neocortex. Neuron 99, 814–828.e7 (2018).

    Article  CAS  PubMed  Google Scholar 

  13. American Psychiatric Association. Diagnostic and statistical manual of mental disorders 5th edn (American Psychiatric Association, 2013).

  14. 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  PubMed  Google Scholar 

  15. Guo, Z. V. et al. Procedures for behavioral experiments in head-fixed mice. PLoS ONE 9, e88678 (2014).

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  16. Gil-Sanz, C. et al. Lineage tracing using Cux2-Cre and Cux2-CreERT2 mice. Neuron 86, 1091–1099 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Gerfen, C. R., Paletzki, R. & Heintz, N. GENSAT BAC Cre-recombinase driver lines to study the functional organization of cerebral cortical and basal ganglia circuits. Neuron 80, 1368–1383 (2013).

    Article  CAS  PubMed  Google Scholar 

  18. Allen, W. E. et al. Thirst-associated preoptic neurons encode an aversive motivational drive. Science 357, 1149–1155 (2017).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  19. DeNardo, L. A. et al. Temporal evolution of cortical ensembles promoting remote memory retrieval. Nat. Neurosci. 22, 460–469 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Oh, S. W. et al. A mesoscale connectome of the mouse brain. Nature 508, 207–214 (2014).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  21. Hunnicutt, B. J. et al. A comprehensive thalamocortical projection map at the mesoscopic level. Nat. Neurosci. 17, 1276–1285 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Phillips, J. W. et al. A repeated molecular architecture across thalamic pathways. Nat. Neurosci. 22, 1925–1935 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. McCormick, D. A. & Pape, H. C. Properties of a hyperpolarization-activated cation current and its role in rhythmic oscillation in thalamic relay neurones. J. Physiol. (Lond.) 431, 291–318 (1990).

    Article  CAS  Google Scholar 

  24. Leresche, N., Lightowler, S., Soltesz, I., Jassik-Gerschenfeld, D. & Crunelli, V. Low-frequency oscillatory activities intrinsic to rat and cat thalamocortical cells. J. Physiol. (Lond.) 441, 155–174 (1991).

    Article  CAS  Google Scholar 

  25. Poulet, J. F. A., Fernandez, L. M. J., Crochet, S. & Petersen, C. C. H. Thalamic control of cortical states. Nat. Neurosci. 15, 370–372 (2012).

    Article  CAS  PubMed  Google Scholar 

  26. Fogerson, P. M. & Huguenard, J. R. Tapping the brakes: cellular and synaptic mechanisms that regulate thalamic oscillations. Neuron 92, P687–P704 (2016).

    Article  CAS  Google Scholar 

  27. 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).

    Article  CAS  PubMed  Google Scholar 

  28. Anis, N. A., Berry, S. C., Burton, N. R. & Lodge, D. The dissociative anaesthetics, ketamine and phencyclidine, selectively reduce excitation of central mammalian neurones by N-methyl-aspartate. Br. J. Pharmacol. 79, 565–575 (1983).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Ludwig, A., Zong, X., Jeglitsch, M., Hofmann, F. & Biel, M. A family of hyperpolarization-activated mammalian cation channels. Nature 393, 587–591 (1998).

    Article  CAS  PubMed  ADS  Google Scholar 

  30. Santoro, B. et al. Identification of a gene encoding a hyperpolarization-activated pacemaker channel of brain. Cell 93, P717–P729 (1998).

    Article  Google Scholar 

  31. Vogt, B. A. & Paxinos, G. Cytoarchitecture of mouse and rat cingulate cortex with human homologies. Brain Struct. Funct. 219, 185–192 (2014).

    Article  PubMed  Google Scholar 

  32. Foster, B. L. & Parvizi, J. Direct cortical stimulation of human posteromedial cortex. Neurology 88, 685–691 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  33. 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 

  34. Hua, T. et al. General anesthetics activate a potent central pain-suppression circuit in the amygdala. Nat. Neurosci. 23, 854–868 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  ADS  Google Scholar 

  36. Tomitaka, M., Tomitaka, S., Rajdev, S. & Sharp, F. R. Fluoxetine prevents PCP- and MK801-induced HSP70 expression in injured limbic cortical neurons of rats. Biol. Psychiatry 47, 836–841 (2000).

    Article  CAS  PubMed  Google Scholar 

  37. Olney, J. W., Labruyere, J. & Price, M. T. Pathological changes induced in cerebrocortical neurons by phencyclidine and related drugs. Science 244, 1360–1362 (1989).

    Article  CAS  PubMed  ADS  Google Scholar 

  38. Mason, M. F. et al. Wandering minds: the default network and stimulus-independent thought. Science 315, 393–395 (2007).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  39. Raichle, M. E. The brain’s default mode network. Annu. Rev. Neurosci. 38, 433–447 (2015).

    Article  CAS  PubMed  Google Scholar 

  40. Kohrs, R. & Durieux, M. E. Ketamine: teaching an old drug new tricks. Anesth. Analg. 87, 1186–1193 (1998).

    CAS  PubMed  Google Scholar 

  41. Green, S. M., Roback, M. G., Kennedy, R. M. & Krauss, B. Clinical practice guideline for emergency department ketamine dissociative sedation: 2011 update. Ann. Emerg. Med. 57, 449–461 (2011).

    Article  PubMed  Google Scholar 

  42. Schwenk, E. S. et al. Consensus guidelines on the use of intravenous ketamine infusions for acute pain management from the American Society of Regional Anesthesia and Pain Medicine, the American Academy of Pain Medicine, and the American Society of Anesthesiologists. Reg. Anesth. Pain Med. 43, 456–466 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  43. Jun, J. J. et al. Fully integrated silicon probes for high-density recording of neural activity. Nature 551, 232–236 (2017).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  44. Putzeys, J. et al. Neuropixels data-acquisition system: a scalable platform for parallel recording of 10 000+ electrophysiological signals. IEEE Trans. Biomed. Circuits Syst. 13, 1635–1644 (2019).

    Article  PubMed  Google Scholar 

  45. Stringer, C. et al. Spontaneous behaviors drive multidimensional, brainwide activity. Science 364, eaav7893 (2019).

    Article  CAS  Google Scholar 

  46. Allen, W. E. et al. Thirst regulates motivated behavior through modulation of brainwide neural population dynamics. Science 364, eaav3932 (2019).

    Article  CAS  Google Scholar 

  47. Zalocusky, K. A. et al. Nucleus accumbens D2R cells signal prior outcomes and control risky decision-making. Nature 531, 642–646 (2016).

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  48. Gunaydin, L. A. et al. Natural neural projection dynamics underlying social behavior. Cell 157, 1535–1551 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Lein, E. S. et al. Genome-wide atlas of gene expression in the adult mouse brain. Nature 445, 168–176 (2007).

    Article  CAS  PubMed  ADS  Google Scholar 

Download references

Acknowledgements

This work was supported by grants to K.D. from the National Institute on Drug Abuse (NIDA P50 Center), NIMH, DARPA, the Tarlton Foundation, the AE Foundation Borderline Research Fund, the NOMIS Foundation and the Else Kroner Fresenius Foundation. K.D. and L.L. were additionally supported by the NSF NeuroNex program. S.V., I.V.K. and E.R. are supported by a National Science Foundation Graduate Research Fellowship. F.G. is supported by a Walter V. and Idun Berry Postdoctoral Fellowship, a NARSAD Young Investigator Award, and a K99/R00 award from NIDA. J.H. and P.N. are supported by a Stanford Bio-X Interdisciplinary Initiatives Seeds Grants Program. P.N. received funding from the Wu Tsai Neurosciences Institute. J.P. is supported by R01MH109954 from the National Institutes of Mental Health. We thank J. H. Lui (Rbp4-Cre), S. Franco (Cux2-CreER), K. Masuda and L. Giocomo (HCN1), and W. E. Allen for contributing to widefield imaging and Neuropixels systems; C. Ramakrishnan for virus assistance; and S. Pak, A. Chibukhchyan, N. Pichamoorthy, C. Lee and C. Delacruz for administrative support and animal husbandry. We acknowledge L. Williams and L. Tozzi for detailed discussions regarding human imaging; and L. Giocomo, B. Heifets, B. Knutson, L. Fenno, E. Sylwestrak, Y. Chen, X. Sun, T. Machado and J. Kochalka for discussion. We thank members of the Stanford Comprehensive Epilepsy Center, including H. Kaur, T. Pham, L. Schumacher, D. Sebrell, A. Valderde, A. Joshi, M. Market, C. Halpern and B. Razavi. We also thank the Clinical Imaging and Stimulation subgroup in our laboratory for discussion and collaboration.

Author information

Authors and Affiliations

Authors

Contributions

S.V., I.V.K. and K.D. designed experiments and wrote the paper. S.V. and I.V.K. collaborated closely to perform all mouse experiments, analyse all datasets and make all figures. E.R. constructed the Neuropixels rig, designed Neuropixels experiments, performed recordings and processed data under the guidance of K.D. and L.L. S.V., I.V.K. and E.R. analysed Neuropixels data and made figures. F.G. performed electrophysiological recordings, analysis and made figures under the guidance of K.D. and R.C.M. J.M.H. performed the implantation of the epilepsy monitoring electrodes in the human research patient. J.P., J.M.H., P.N., T.O. and C.S.-S. collected human iEEG data. T.O. and P.N. performed human iEEG data curation and processing. J.P. performed the human electrical stimulation procedure. All authors made comments on the manuscript. K.D. supervised all aspects of the work.

Corresponding author

Correspondence to Karl Deisseroth.

Ethics declarations

Competing interests

All tools and protocols used during this study are freely available for non-profit use from the corresponding author upon request. Stanford University is in the process of submitting a patent application to further facilitate therapeutic translation of the findings reported in this study.

Additional information

Peer review information Nature thanks Thomas J. McHugh 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 Time course of ketamine-induced retrosplenial rhythm.

a, Prominent 1–3 Hz retrospenial activity after ketamine injection seen in a frequency response from Thy1-GCaMP6s (top) but not control (Thy1–GFP) mice (bottom). b, Expansion of the first 15 min of the recording shown in Fig. 1g. c, Power in 1–3 Hz band in motor (MOT), somatosensory (SS), visual (VIS), posterior parietal (PP) and retrosplenial (RSP) cortices (mean ± s.e.m., n = 6 mice). d, Example 410-nm channel-corrected trace of RSP and SS activity from 3 min before to 12 min after intraperitoneal injection of 50 mg kg−1 ketamine. Vertical lines indicate injection time. e, Magnifications of 30 s of data demonstrating oscillatory rhythm in the RSP but not the SS at ten minutes after injection, or before injection. f, PSD in the RSP during five consecutive days of 50 mg kg−1 ketamine injection (mean ± s.e.m., n = 6 mice). g, Summary of power in the 1–3 Hz band in the RSP during five consecutive days of 50 mg kg−1 ketamine injection (ns, corrected paired t-test comparison with first day, P = 0.10, 0.10, 0.80, 0.75, Hedge’s g = 0.73, 0.49, 0.12, 0.15). h, Direct comparison of oscillation in similarly aged male and female mice. PSD in the RSP and summary of power in the 1–3 Hz band. Both sexes have significant increase in power relative to pre-injection, but the magnitude is not significantly different between male and female mice (P < 0.01, corrected t-test versus 0, and P = 0.21, t-test between groups, Hedge’s g = 0.76). i, Video tracking of mouse facial movements. Facial movements did not show 1–3 Hz features before or after administration of ketamine (paired t-test P = 0.24, Hedge’s g = 0.80). j, Top, correlation of cortical areas imaged with widefield imaging, before and after saline injection (n = 5 mice, corrected paired t-test P = 0.85, 0.51, 0.37, 0.25, Hedge’s g = 0.08, 0.27, 0.36, 0.21). Bottom, correlation of cortical areas imaged with widefield imaging, before and after injection of 50 mg kg−1 ketamine (n = 5 mice, corrected paired t-test P = 0.009, 0.009, 0.009, 0.009, Hedge’s g = 3.88, 3.67, 3.75, 3.39, 3.36). k, Schematic of cranial window preparation and two-photon imaging of superficial neuropil. l, Example fullfield fluorescence trace from the RSP in a ketamine-injected Thy1-GCaMP6s mouse, imaged 30 μm below the cortical surface. m, Example fullfield fluorescence trace from the visual cortex, in the same mouse (same ketamine injection as in l). n, Quantification of oscillation frequency content from neuropil in cortical surface before (blue) and after (red) ketamine injection, n = 4 mice. o, Comparison of mean power in the 1–3 Hz band after ketamine injection; n = 4 mice (one-sided paired t-test, P = 0.035; Hedge’s g = 1.60). p, Quantification of oscillation frequency content before and after injection of 80 mg kg−1 ketamine, n = 5 mice; solid line represents the mean and shading denotes the s.e.m. q, After injection of 80 mg kg−1 ketamine, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 3.74, P = 0.025. Corrected two-sided paired t-tests P = 0.016, 0.034, 0.016, 0.034; Hedge’s g 1.92, 1.90, 1.66, 1.65. r, After injection of 80 mg kg−1 ketamine, example temporal video sequence of 410 nm-corrected fluorescence over 1 s in a Thy1-GCaMP6s mouse. A prominent rhythm is seen in the RSP.

Source data

Extended Data Fig. 2 PCP and MK801—but not memantine, saline or LSD—induce RSP-localized rhythms.

a, Quantification of oscillation frequency content before and after PCP injection, n = 5 mice. b, For PCP, example temporal video sequence of 410 nm-corrected fluorescence over 1 s in a Thy1-GCaMP6s mouse showing RSP rhythm. c, Quantification of oscillation frequency content before and after MK801 injection, n = 5 mice. d, For MK801, example temporal video sequence of 410 nm-corrected fluorescence over 1 s in a Thy1-GCaMP6s mouse, showing RSP rhythm. e, Quantification of oscillation frequency content before and after memantine injection, n = 5 mice. f, For memantine, example temporal video sequence of 410 nm-corrected fluorescence across 1 s in a Thy1-GCaMP6s mouse. Here, neither rhythms nor travelling waves are seen. g, Quantification of oscillation frequency content before and after saline injection, n = 5 mice. h, Quantification of mean power in 1–3 Hz band for saline (n = 5 mice), corrected two-sided paired t-tests (*P < 0.05, **P < 0.01, NS P > 0.05; Hedge’s g 0.36, 0.18, 0.14, 0.01). i, Quantification of oscillation frequency content before and after LSD injection, n = 5 mice. j, Quantification of mean power in 1–3 Hz band for LSD (n = 5 mice), corrected two-sided paired t-tests (*P < 0.05, **P < 0.01, NS P > 0.05; Hedge’s g LSD 2.05, 0.62, 1.07, 0.65). k, For LSD, example temporal video sequence of 410 nm-corrected fluorescence across 1 s in a Thy1-GCaMP6s mouse. Here, neither rhythms nor travelling waves are seen.

Source data

Extended Data Fig. 3 Other anaesthetic classes do not induce retrosplenial cortex-localized rhythms.

a, Quantification of oscillation frequency content before and after injection of 1.5 mg kg−1 dexmedetomidine, n = 5 mice. b, After injection of 1.5 mg kg−1 dexmedetomidine, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 2.66, P = 0.07. Corrected two-sided paired t-tests, P = 0.91, 0.27, 0.15, 0.91; Hedge’s g = 0.14, 0.77, 1.01, 0.06. c, Example temporal video sequence of 410 nm-corrected fluorescence after injection of 1.5 mg kg−1 dexmedetomidine over 1 s in a Thy1-GCaMP6s mouse. d, Quantification of oscillation frequency content before and after injection of 0.75 mg kg−1 dexmedetomidine, n = 5 mice. e, After injection of 0.75 mg kg−1 dexmedetomidine, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 2.340, P = 0.10. Corrected two-sided paired t-tests, P = 0.92, 0.92, 0.04, 0.86; Hedge’s g = 0.05, 0.11, 0.54, 0.13. f, Quantification of oscillation frequency content before and after injection of 0.35 mg kg−1 dexmedetomidine, n = 5 mice. g, After injection of 0.35 mg kg−1 dexmedetomidine, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 2.68, P = 0.07. Corrected two-sided paired t-tests, P = 0.29, 0.95, 0.013, 0.95; Hedge’s g = 0.68, 0.05, 0.32, 0.01. h, Quantification of oscillation frequency content before and after injection of 140 mg kg−1 propofol, n = 5 mice. i, After injection of 140 mg kg−1 propofol, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 6.46, P = 0.0027. Corrected two-sided paired t-tests, P = 0.0026, 0.0029, 0.0029, 0.28; Hedge’s g = 4.36, 3.07, 1.75, 0.61. j, Example temporal video sequence of 410 nm-corrected fluorescence after injection of 140 mg kg−1 propofol over 1 s in a Thy1-GCaMP6s mouse. k, Quantification of oscillation frequency content before and after injection of 70 mg kg−1 propofol, n = 5 mice. l, After injection of 70 mg kg−1 propofol, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 1.34, P = 0.30. Corrected two-sided paired t-tests, P = 0.11, 0.11, 0.11, 0.11; Hedge’s g = 0.79, 1.17, 0.37, 0.83. m, Quantification of oscillation frequency content before and after injection of 35 mg kg−1 propofol, n = 5 mice. n, For injection of 35 mg kg−1 propofol, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 2.69, P = 0.07. Corrected two-sided paired t-tests, P = 0.17, 0.07, 0.07, 0.17; Hedge’s g = 0.66, 1.56, 0.63, 0.82. o, Quantification of oscillation frequency content before and after injection of ketamine/xylazine, n = 5 mice. p, After injection of ketamine/xylazine, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 1.977, P = 0.15. Corrected two-sided paired t-tests, P = 0.08, 0.08, 0.82, 0.82; Hedge’s g = 1.44, 2.25, 0.08, 0.13. q, Example temporal video sequence of 410 nm-corrected fluorescence after injection of ketamine/xylazine over 1 s in a Thy1-GCaMP6s mouse.

Source data

Extended Data Fig. 4 A sedative and an analgesic do not induce the retrosplenial rhythm.

a, Quantification of oscillation frequency content before and after injection of 2 mg kg−1 diazepam, n = 5 mice. b, After injection of 2 mg kg−1 diazepam, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 39, P < 0.001. Corrected two-sided paired t-tests, P = 0.013, 0.028, 0.67, 0.013; Hedge’s g = 3.10, 2.07, 0.16, 1.27. c, Example temporal video sequence of 410 nm-corrected fluorescence after injection of 2 mg kg−1 diazepam over 1 s in a Thy1-GCaMP6s mouse. d, Quantification of oscillation frequency content before and 2 h after injection of buprenorphine SR, n = 5 mice. e, After injection of buprenorphine SR, comparison of mean power in the 1–3 Hz band, n = 5 mice. One-way ANOVA with repeated measures, F4,16 = 1.55, P = 0.23. Corrected two-sided paired t-tests, P = 0.51, 0.33, 0.33, 0.33; Hedge’s g = 0.17, 0.51, 0.33, 0.38. f, Example temporal video sequence of 410 nm-corrected fluorescence after injection of buprenorphine SR, over 1 s in a Thy1-GCaMP6s mouse.

Source data

Extended Data Fig. 5 Additional information for two photon imaging, TRAP and 32-contact electrophysiology experiments.

a, Schematic of two-photon, head-fixed imaging. GCaMP6m was expressed in layer-specific transgenic mice, and a cranial window was placed over the RSP. b, Epifluorescence images of coronal sections showing GCaMP6m expression restricted to either layer 2/3 or layer 5 neurons in the RSP. Scale bar, 100 μm. c, Single-cell activity traces of all recorded layer 2/3 retrosplenial neurons before ketamine injection, from one mouse. d, Quantification of oscillation frequency content in single layer 2/3 cells before and after ketamine injection, data show average and s.e.m. from n = 5 mice. e, Single-cell activity traces of all recorded layer 5 retrosplenial neurons before ketamine injection from one mouse. f, Quantification of oscillation frequency content in single layer 5 cells before and after ketamine injection, data show average and s.e.m. from n = 5 mice. g, Fluorescence images showing tdTomato expression after recombination of TRAP2;Ai14 mice, after saline injections. Scale bar, 1 mm. h, Magnified example image of the RSP after ketamine injection. Dense labelling (tdTomato+ neurons are black cells) in deep layers is observed. i, Quantification of tdTomato+ induction in the RSP in TRAP2;Ai14 mice after saline (blue, n = 6 coronal sections) or ketamine (red, n = 8 coronal sections) injections. Two-sided unpaired t-test, layer 2/3, P = 0.42; layer 5, P < 0.0004; Hedge’s g = 2.42. j, Left, histogram of spike rate during ON periods for mouse 3. Right, summary plot for three mice. k, Left, histogram of spike rate during OFF periods for mouse 3. Right, summary plot for three mice. l, Quantification of the percentage of recorded units that participated in each burst. Left, histogram of per cent units participating in a burst from mouse 3 (n = 249 bursts). Right, summary plot for three mice. m, Quantification of the number of spikes in each burst. Left, histogram from mouse 3. Right, summary plot for three mice. n, Quantification of the percentage of bursts that each neuron participates in. Left, histogram from mouse 3. Right, summary plot for three mice. o, For all 3 mice, correlation of rank sequence across all bursts over 2 min, compared to shuffling of unit IDs. Two-sided paired t-test, P < 0.0001 for all mice. The box denotes the mean and the bars the s.e.m. Hedge’s g = 2.30, 2.25, 3.93. p, Temporal organization of firing across bursts for each mouse. Left, rank of each neuron’s first spike in burst (averaged across bursts for each neuron). Right, rank of each neurons first spike with shuffling of unit IDs.

Source data

Extended Data Fig. 6 Brain-wide single-cell firing rates and spectral features.

a, Recording locations overlaid on coronal sections from the Allen Mouse Brain Atlas. b, Change in firing rate per brain region between 8–12 min before and 8–12 min after ketamine injection, each grey dot is an individual cell. Cells from all 4 sessions are included. c, Nucleus accumbens spike rasters before (top) and after (bottom) ketamine injection from a single session. d, PSD plots for cortical and subcortical regions in Fig. 2d–f; before (blue) and after (red) ketamine injection. e, PSD plots for additional subcortical regions; before (blue) and after (red) ketamine injection. f, Average regional power in the 1–3Hz band for cortical and subcortical regions from d. Each recording is represented as a single line. The total number of recorded units is indicated. g, Average regional power in the 1–3Hz band for additional subcortical regions from e. Each recording is represented as a single line. The total number of recorded units is indicated.

Source data

Extended Data Fig. 7 Brain-wide functional correlations with retrosplenial cortex and thalamic rhythms.

a, Before ketamine infusion, correlation with the average RSP trace for every recorded cell (grey dot) categorized by region. b, After ketamine infusion, correlation with the average RSP trace for every recorded cell (grey dot) categorized by region. c, For every neuron, change in correlation with the mean RSP trace. d, For every region, magnitude of correlation with the mean RSP trace. e, For every region, change in the correlation with mean RSP trace. Regional averages (bar) and s.e.m (line) are shown. f, For mouse 1, raster plot of spike trains from simultaneously recorded single units in somatosensory cortex (SS), anteromedial thalamus (AM), anteroventral thalamus (AV), lateral dorsal thalamus (LD) and retrosplenial cortex (RSP) before and after ketamine injection. g, For mouse 2, raster plot of spike trains from simultaneously recorded single units in somatosensory cortex (SS), anteromedial thalamus (AM), anteroventral thalamus (AV), lateral dorsal thalamus (LD) and retrosplenial cortex (RSP) before and after ketamine injection. h, Z-scored mean activity traces from three simultaneously recorded thalamic nuclei and retrosplenial cortex (green) corresponding to rasters in f. i, Histological assessment registered to the Allen Mouse Brain Atlas (2004)49. Electrode insertions revealed the location of mouse 2’s thalamic probe at the edge of the anteromedial thalamus, a possible explanation for this recording’s fewer AM neurons and weaker anti-correlation with RSP. Right, a coronal image from the Allen Mouse Brain Atlas shows the locations of thalamic nuclei.

Source data

Extended Data Fig. 8 Brain-wide single-cell functional connectivity matrices.

a, Correlation matrix between individual cells, grouped by brain region, before ketamine infusion for recording session 1. b, Correlation matrix between individual cells, grouped by brain region, after ketamine infusion for recording session 1. c, Change in regional correlation for recording session 1. d, Correlation matrix between individual cells, grouped by brain region, before ketamine infusion for recording session 2. e, Correlation matrix between individual cells, grouped by brain region, after ketamine infusion for recording session 2. f, Change in regional correlation for recording session 2. g, Correlation matrix between individual cells, grouped by brain region, before ketamine infusion for recording session 3. h, Correlation matrix between individual cells, grouped by brain region, after ketamine infusion for recording session 3. i, Change in regional correlation for recording session 3.

Source data

Extended Data Fig. 9 Additional characterization of dissociation-like phenotype elicited by ketamine and PCP.

a, Hot-plate test. Time to jump in ketamine-injected mice. The maximum allowable test time was 90 s. One-way ANOVA with repeated measures: jump, 1 × 10−12. Horizontal dotted line indicates the experimental time limit; mice that did not jump by 90 s were removed and the experiment ended. Corrected two-sided unpaired t-tests, *P < 0.05, **P < 0.01, ***P < 0.001. n = 5 mice in each group. For each dose in order, Glass’s ∆ effect size for jump = 0.55, 1.9, Inf, Inf. b, Rate of rears in hot-plate test. One-way ANOVA with repeated measures, P < 1 × 10−4. Two-sided unpaired t-tests. For each dose in order, Glass’s ∆ effect sizes = −0.82, −0.42, −24.6, −24.6. c, Latency to first flick, lick and rear in hot-plate test. The horizontal dotted line indicates the experimental time limit. Hedge’s g effect sizes for flick = 0.16, 0.34, −0.30, −0.17. Glass’s ∆ effect size for lick = 0.66, 0.17, 1.86, 7.62. Glass’s ∆ effect size for rear = 0.47, 0.85, Inf, 10.5. d, Tail suspension test. Percentage of time spent struggling. One-way ANOVA with repeated measures, F4,20 = 9.36, P = 0.0002, n = 5 mice for each group. Glass’s ∆ effect sizes = −0.30, 0.024, −2.49, −585. e, Time spent interacting with an intruder mouse in 2 min. One-way ANOVA with repeated measures, F4,20 = 15.8; P < 0.0001, n = 5 mice for each group. Glass’s ∆ effect sizes = −0.27, −0.67, −44.7, −79.5. f, Number of mice that corrected postural inversion within 5 s. Mice dosed with 50 mg kg−1 ketamine or less reflexively righted. Mice injected with 200 mg kg−1 ketamine did not right. n = 5 for each group. g, Hot-plate test, time to jump. The maximum allowable time is indicated by a dashed horizontal line (90 s); mice that did not jump by 90 s were removed and the experiment ended. For all statistical comparisons, each experimental group was compared with the saline (control) group via corrected Mann–Whitney U-test. *P < 0.05, **P < 0.01, ***P < 0.001; for each drug in order, Glass’s ∆ effect size for jump = Inf, 4.97, 0.35, −1.69, 0.60. h, Rearing rate was significantly decreased for the PCP group and increased for the LSD group. Glass’s ∆ = −10.9, −0.52, 0.34, 1.00, −0.10. i, Latencies to expression of affective (licking) and escape (rear) behaviours were significantly longer for the PCP mice. The maximum allowable time is indicated by dashed horizontal line (90 s). Hedge’s g effect sizes for flick = −0.78, 1.87, 1.07, −0.24, 1.05. Glass’s ∆ effect size for lick = 1.14, 0.75, 0.072, 0.96, 0.52. Glass’s ∆ effect size for rear = 12.4, 0.33, −0.63, −1.92, 0.39. j, Tail suspension test. Significantly decreased struggling time was observed in mice treated with PCP and diazepam compared with saline-injected mice (Mann–Whitney U-test with FDR correction, **P < 0.01). Glass’s ∆ effect sizes = −2.51, 0.074, −0.17, 0.58, −6.50. k, Social interaction. Significantly decreased social interaction time was observed in mice treated with PCP and buprenorphine (Mann–Whitney U-test with FDR,**P < 0.01). Hedge’s g effect sizes = −2.10, −2.13, −0.068, 1.88, −1.23. l, Righting reflex, showing the number of mice that right for each treatment group. For each group, all mice successfully corrected postural inversion. m, Open field test. Example traces of body position after treatment with each drug (2 example mice shown). Grey line represents full 5-min session, black line represents 20 s of tracking (from 2 min to 2 min 20 s).

Source data

Extended Data Fig. 10 Rhythmic optogenetic behaviour, alternative cortical stimulation locations, and behavioural controls.

a, Frequency content of rhythmic (magenta) and random (purple) illumination patterns. b, Hot-plate test: time to jump. The maximum allowable time is indicated by dashed horizontal line (90 s). Time-to-jump: YFP/rhythmic P = 0.020, YFP/random P = 0.87, rhythmic/random P = 0.054, corrected two-sided unpaired t-tests, n = 8 YFP, 7 rhythmic, 7 random mice. Hedge’s g effect sizes for jump = 1.58, −0.79, −1.2). c, Rate of rears, two-sided unpaired t-test, YFP/rhythmic P = 0.020, YFP/random P = 0.27, rhythmic/random P = 0.44. Hedge’s g effect sizes = −1.57, −0.68, −0.40. d, Latency to first withdrawal, lick and rear. The maximum allowable time is indicated by the dashed horizontal line (90 s). Two-sided unpaired t-tests, corrected P values for foot flick (>0.5 for all comparisons); foot lick: YFP/rhythmic P = 0.087, YFP/random P = 0.40, rhythmic/random P = 0.12; rear: P = 0.071, 0.42, 0.071; Hedge’s g effect sizes for flick = −0.0084, −0.43, −0.66; lick = 1.20, 0.42, −0.97; and rear = 1.07, −0.42, −1.2. e, Tail suspension test. Two-sided paired t-test for rhythmic/random, corrected P = 0.038; Two-sided unpaired t-tests: YFP/rhythmic corrected P = 0.31 and YFP/rand corrected P = 0.95. Hedge’s g effect sizes: YFP/rhythmic −0.61, rhythmic/random −0.58, YFP/random 0.032. f, Social interaction. Two-sided paired t-test for rhythmic/random, corrected P = 0.21; Two-sided unpaired t-tests: YFP/rhythmic corrected P = 0.21 and YFP/random corrected P = 0.39. Hedge’s g effect sizes: YFP/rhythmic = −0.85, rhythmic/random = 0.60, YFP/random = −0.42. g, Righting reflex, percentage of mice that right. In each group, all mice successfully corrected postural inversion. h, Average velocity in the open field test during optogenetic stimulation, comparing average light off and light on conditions within individual mice. Paired t-test, eNPAC P = 0.49, YFP P = 0.40. i, Time in centre in the open field test, comparing average light off and light on conditions within individual mice. Paired t-test, eNPAC P = 0.06, YFP P = 0.25. j, Velocity for each 3 min epoch. eNPAC (magenta) and YFP (grey). s.e.m. is shown. k, Time in centre for each 3 min epoch. eNPAC (magenta) and YFP (grey). s.e.m. is shown. l, Hot-plate test. Rate of jumps (P = 0.85), rearing (P = 0.03) and latencies. Two-sided paired t-test, n = 8 control mice and n = 6 S1 eNPAC mice. Hedge’s g effect size for jump = 0.09, rear = −1.27, latency to flick = −0.41, latency to lick = 0.78, latency to rear = 0.52. m, Tail suspension test. Percentage of time spent struggling for S1 eNPAC mice, with and without illumination, two-sided paired t-test, P = 0.73. Change in struggling time, P = 0.73. Hedge’s g = 0.11. n, Social interaction. Interaction time significantly decreased with illumination in S1 eNPAC mice, two-sided paired t-test, P = 0.015; percentage change with illumination. Hedge’s g = −0.29. o, Righting reflex for S1 YFP and ENPAC mice. p, Hot-plate test without ketamine. For rate of reflexive withdrawals (one-way ANOVA, F2,18 = 1.01, P = 0.38), paw-licks (one-way ANOVA, F2,18 = 2.2, P = 0.14), jump (one-way ANOVA, F2,18 = 0.55, P = 0.58), rearing (one-way ANOVA, F2,18 = 0.09, P = 0.91) and latencies (one-way ANOVA for each behaviour, P > 0.05). Controls are pooled from saline injected mice. q, Tail suspension test. Percentage of time spent struggling. One-way ANOVA, F2,20 = 0.31, P = 0.74. r, Social interaction. Interaction time was similar across groups; one-way ANOVA, F2,19 = 0.31, P = 0.74. s, Righting reflex for GRIN1-knockout and HCN1-knockout mice.

Source data

Extended Data Fig. 11 Additional information for long-range input and gene knockout experiments.

a, Percentage of ipsilateral inputs from non-retrosplenial regions across the whole brain (n = 2 mice). The bar denotes the mean. b, Top, example confocal image showing eNPHR3.0–YFP expression. Bottom, localizations of the fibre tip in opsin-expressing mice, each colour is an individual mouse. Scale bars, 500 μm. c, Frequency content of example recording in a dorsal thalamus NpHR3.0 mouse. Each time window (pre, stim, post) consists of 2 min. d, Density of GRIN1 (left) and HCN1 (right) cells in the RSP. GRIN1-expressing neurons were more dense in superficial than in deep layers (paired t-test, P = 0.023; Hedge’s g = −1.94), but HCN1-expressing neurons were mostly found in deep layers (paired t-test, P = 0.0066; Hedge’s g = 2.71). e, Confocal images showing expression of AAVdj-Ef1a-DIO-GCaMP6m and GRIN1 in control (left) and GRIN1-floxed (right) deep RSP. f, Quantification of gene disruption effectiveness in photometry experiments. Co-localization of GRIN1 and HCN1 with GCaMP6m was greatly reduced. g, Quantification of oscillation frequency content (PSD) before (blue) and after (red) ketamine injection, n = 7 mice in each group. h, Confocal images showing distribution of GRIN1+ neurons in wild-type and GRIN-knockout mice. i, Confocal images showing distribution of HCN1+ neurons in wild-type and HCN1-knockout mice. j, Quantification of gene disruption for behavioural experiments. Density of GRIN1+ cells (unpaired t-test, P = 0.0031; Hedge’s g = −2.22) and HCN1+ cells (unpaired t-test, P < 1 × 10−4, Hedge’s g = −4.64) were significantly reduced by injections of AAVdj-Ef1a-Cre (12 injections per mouse). k, Hot-plate test. No mice jumped before the maximum allowable time, indicated by the dashed horizontal line. Rate of rears was not significantly different (corrected Mann–Whitney U-test P > 0.1) for either group. Average latency to first lick was lower for HCN1-knockout mice. l, Tail suspension test. Significantly increased struggling time was observed in HCN1 and GRIN1 mice (corrected Mann–Whitney U-test, **P < 0.001. Glass’s ∆ effect sizes = 1.07, 1.13. m, Social interaction. Significantly increased social interaction time was observed in HCN1 and GRIN1 mice (corrected Mann–Whitney U-test,**P < 0.01). Glass’s ∆ effect sizes = 0.94, 1.39. n, Righting reflex, fraction of mice that right. For each group, all mice successfully corrected postural inversion.

Source data

Extended Data Fig. 12 Additional information on human iEEG recordings.

a, 3D rendering of electrode locations for right and left hemispheres. b, Screenshot of raw clinical iEEG recordings (no processing). The RC electrode is in the right PMC, with RC1 being the deepest contact. Vertical lines are 1 s. Note rhythm in contacts in right deep posteromedial cortex RC3–4 through RC8–9. One second is highlighted in blue showing an approximately 3 Hz rhythm. c, Frequency spectrum of iEEG traces from frontal lobe regions, parietal, temporal lobe cortical regions, and subcortical areas during aura (red) or non-seizure (blue) epochs. d, Quantification of power in the 3–4 Hz band across all contacts during the seizure aura-period (n = 2 seizures, 10 s for each aura). Deep posteromedial regions on both hemispheres exhibit the high power. Contacts that cross threshold (mean + 2.5 × s.d.) are coloured. e, Location of electrodes that elicited a dissociative experience (Fig. 5) that were not located in a spontaneously oscillating region of the cortex.

Source data

Extended Data Fig. 13 Locations of electrode contacts for human iEEG recordings.

MRI images show anatomical locations of electrode contacts displayed in Fig. 5.

Extended Data Fig. 14 Summary of behavioural effects in the rodent model.

Overview of effects from all rodent behaviour experiments (corresponding to Figs. 3, 4).

Supplementary information

Supplementary Information

This file contains Supplementary Notes 1 and 2.

Reporting Summary

Supplementary Video 1

Example widefield videos in Thy1-GcaMP6s mice Example widefield video of 410 nm-corrected fluorescence in a Thy1-GCaMP6s mouse, ten minutes after injection of either saline, ketamine 50 mg/kg, PCP 5 mg/kg, MK801 0.75 mg/kg, memantine 50mg/kg, LSD 0.3 mg/kg, dexmedetomidine 1.5 mg/kg, ketamine/xylazine 87.5/12.5 mg/kg, propofol 140 mg/kg, or diazepam 2 mg/kg. Color bar units are df/f.

Supplementary Video 2

Example two-photon video Example two-photon video of superficial neuropil in a Thy1-GCaMP6s mouse, ten minutes after ketamine injection. Field of view has side-length 550 µm.

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Vesuna, S., Kauvar, I.V., Richman, E. et al. Deep posteromedial cortical rhythm in dissociation. Nature 586, 87–94 (2020). https://doi.org/10.1038/s41586-020-2731-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-2731-9

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