Mouse

Deep posteromedial cortical rhythm in dissociation

AbstractAdvanced 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.

Data availabilityThe datasets generated and analysed are available from the corresponding author upon reasonable request and at https://www.optogenetics.org. Source data are provided with this paper.Code availabilityCode used for data processing and analysis is available from the corresponding author upon reasonable request.References1.Ferezou, I. et al. Spatiotemporal dynamics of cortical sensorimotor integration in behaving mice. Neuron 56, 907–923 (2007).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).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).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).PubMed 

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

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).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).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).PubMed 

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

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

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

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

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

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

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

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

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

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

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

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).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).ADS 
PubMed 

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

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

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

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).PubMed 

Google Scholar 
45.Stringer, C. et al. Spontaneous behaviors drive multidimensional, brainwide activity. Science 364, eaav7893 (2019).
Google Scholar 
46.Allen, W. E. et al. Thirst regulates motivated behavior through modulation of brainwide neural population dynamics. Science 364, eaav3932 (2019).
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).ADS 
PubMed 
PubMed Central 

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

Google Scholar 
Download referencesAcknowledgementsThis 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 informationAuthor notesThese authors contributed equally: Sam Vesuna, Isaac V. KauvarAffiliationsDepartment of Bioengineering, Stanford University, Stanford, CA, USASam Vesuna, Isaac V. Kauvar, Ethan Richman, Felicity Gore, Tomiko Oskotsky, Paul Nuyujukian & Karl DeisserothDepartment of Electrical Engineering, Stanford University, Stanford, CA, USAIsaac V. Kauvar & Paul NuyujukianDepartment of Neurosurgery, Stanford University, Stanford, CA, USATomiko Oskotsky, Jaimie M. Henderson & Paul NuyujukianDepartment of Neurology and Neurological Sciences, Stanford University, Stanford, CA, USAClara Sava-Segal & Josef ParviziDepartment of Biology, Stanford University, Stanford, CA, USALiqun LuoHoward Hughes Medical Institute, Stanford University, Stanford, CA, USALiqun Luo & Karl DeisserothDepartment of Psychiatry and Behavioral Sciences, Stanford University, Stanford, CA, USAFelicity Gore, Robert C. Malenka & Karl DeisserothContributionsS.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 authorCorrespondence 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 informationPeer 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 tablesExtended 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 
Read More

Show More

Related Articles

Leave a Reply

Your email address will not be published.

Back to top button
Close
Close