Skip to main content Accessibility help
×
Hostname: page-component-5c6d5d7d68-7tdvq Total loading time: 0 Render date: 2024-08-07T22:58:39.906Z Has data issue: false hasContentIssue false

Chapter 19 - The Thalamus and Sleep

from Section 8: - Arousal

Published online by Cambridge University Press:  12 August 2022

Michael M. Halassa
Affiliation:
Massachusetts Institute of Technology
Get access

Summary

The architecture of the thalamus and its reciprocal connections with multiple cortical and subcortical structures are essential to the generation of the thalamo-cortical network oscillations associated with attention, sleep, and consciousness. This chapter provides an overview of the cellular mechanisms underlying thalamo-cortical network oscillations occurring during sleep and their contribution to the architecture of the sleep–wake cycle, including the onset and stability of non–rapid eye movement (NREM) and rapid eye movement (REM) sleep. It further summarizes the influence of the brainstem neuromodulatory system on thalamo-cortical network activity during wakefulness and sleep. Finally, the association between these mechanisms and synaptic plasticity in thalamo-cortical networks is described in the context of sleep-dependent consolidation, or weakening, of previously acquired information in health and disease.

Type
Chapter
Information
The Thalamus , pp. 361 - 381
Publisher: Cambridge University Press
Print publication year: 2022

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Abrams, J. K., Johnson, P. L., Hollis, J. H., & Lowry, C. A. (2004). Anatomic and functional topography of the dorsal raphe nucleus. Annals of the New York Academy of Sciences, 1018, 4657. https://doi.org/10.1196/annals.1296.005Google Scholar
Adamantidis, A. R., Gutierrez Herrera, C., & Gent, T. C. (2019). Oscillating circuitries in the sleeping brain. Nature Reviews Neuroscience, 20(12), 746762. https://doi.org/10.1038/s41583-019-0223-4Google Scholar
Alkire, M., Hudetz, A., & Tononi, G. (2008). Consciousness and anesthesia. Science, 322(November), 139152. https://doi.org/10.1016/B978-0-12-800948-2.00009-1CrossRefGoogle ScholarPubMed
Alkire, M. T., Asher, C. D., Franciscus, A. M., & Hahn, E. L. (2009). Thalamic microinfusion of antibody to a voltage-gated potassium channel restores consciousness during anesthesia. Anesthesiology, 110(4), 766773. https://doi.org/10.1097/ALN.0b013e31819c461cGoogle Scholar
Alkire, M. T., McReynolds, J. R., Hahn, E. L., & Trivedi, A. N. (2007). Thalamic microinjection of nicotine reverses sevoflurane-induced loss of righting reflex in the rat. Anesthesiology, 107(2), 264272. https://doi.org/10.1097/01.anes.0000270741.33766.24Google Scholar
Anaclet, C., Griffith, K., & Fuller, P. M. (2018). Activation of the GABAergic parafacial zone maintains sleep and counteracts the wake-promoting action of the psychostimulants armodafinil and caffeine. Neuropsychopharmacology, 43(2), 415425. https://doi.org/10.1038/npp.2017.152Google Scholar
Anaclet, C., Pedersen, N. P., Ferrari, L. L., Venner, A., Bass, C. E., Arrigoni, E., & Fuller, P. M. (2015). Basal forebrain control of wakefulness and cortical rhythms. Nature Communications, 6, 114. https://doi.org/10.1038/ncomms9744CrossRefGoogle ScholarPubMed
Anafi, R. C., Kayser, M. S., & Raizen, D. M. (2019). Exploring phylogeny to find the function of sleep. Nature Reviews Neuroscience, 20(2), 109116. https://doi.org/10.1038/s41583-018-0098-9Google Scholar
Anderson, M. P., Mochizuki, T., Xie, J., Fischler, W., Manger, J. P., Talley, E. M., Scammell, T. E., & Tonegawa, S. (2005). Thalamic Cav3.1 T-type Ca2+ channel plays a crucial role in stabilizing sleep. Proceedings of the National Academy of Sciences of the United States of America, 102(5), 17431748.Google Scholar
Aston-Jones, G., & Cohen, J. D. (2005). An integrative theory of locus coeruleus-norepinephrine function: adaptive gain and optimal performance. Annual Review of Neuroscience, 28(1), 403450. https://doi.org/10.1146/annurev.neuro.28.061604.135709CrossRefGoogle ScholarPubMed
Astori, S., & Lüthi, A. (2013). Synaptic plasticity at intrathalamic connections via CaV3.3 T-type Ca2+ channels and GluN2B-containing NMDA receptors. Journal of Neuroscience, 33(2), 624630. https://doi.org/10.1523/JNEUROSCI.3185-12.2013Google Scholar
Astori, S., Wimmer, R. D., Prosser, H. M., Corti, C., Corsi, M., & Liaudet, N. (2011). The CaV3.3 calcium channel is the major sleep spindle pacemaker in thalamus. Proceedings of the National Academy of Sciences of the United States of America, 108(33), 1382313828. https://doi.org/10.1073/pnas.1105115108Google Scholar
Aton, S. J., Suresh, A., Broussard, C., & Frank, M. G. (2014). Sleep promotes cortical response potentiation following visual experience. Sleep, 37(7), 11631170. https://doi.org/10.5665/sleep.3830CrossRefGoogle ScholarPubMed
Baker, A. P., Brookes, M. J., Rezek, I. A., Smith, S. M., Behrens, T., Smith, P. J. P., & Woolrich, M. (2014). Fast transient networks in spontaneous human brain activity. eLife, 2014(3), 118. https://doi.org/10.7554/eLife.01867Google Scholar
Bal, T., von Krosigk, M., & McCormick, D. A. (1995). Role of the ferret perigeniculate nucleus in the generation of synchronized oscillations in vitro. Journal of Physiology, 483(3), 665685. https://doi.org/10.1113/jphysiol.1995.sp020613Google Scholar
Bandarabadi, M., Boyce, R., Herrera, C. G., Bassetti, C. L., Williams, S., Schindler, K., & Adamantidis, A. (2019). Dynamic modulation of theta-gamma coupling during rapid eye movement sleep. Sleep, 42(12), 111. https://doi.org/10.1093/sleep/zsz182Google Scholar
Bandarabadi, M., Herrera, C. G., Gent, T. C., Bassetti, C., Schindler, K., & Adamantidis, A. R. (2020). A role for spindles in the onset of rapid eye movement sleep. Nature Communications, 11(5247). https://doi.org/10.1038/s41467-020-19076-2Google Scholar
Bassetti, C. L. (2005). Sleep and stroke. Seminars in Neurology, 25. https://doi.org/10.1016/B978-0-12-804074-4.00006-6CrossRefGoogle ScholarPubMed
Bassetti, C., Mathis, J., Gugger, M., Lovblad, K., & Hess, C. W. (1996). Hypersomnia following paramedian thalamic stroke: a report of 12 patients. Annals of Neurology, 39, 471480.CrossRefGoogle ScholarPubMed
Benson, K. L. (2015). Sleep in schizophrenia: pathology and treatment. Sleep Medicine Clinics, 10(1), 4955. https://doi.org/10.1016/j.jsmc.2014.11.001CrossRefGoogle ScholarPubMed
Bergel, A., Deffieux, T., Demené, C., Tanter, M., & Cohen, I. (2018). Local hippocampal fast gamma rhythms precede brain-wide hyperemic patterns during spontaneous rodent REM sleep. Nature Communications, 9(1). https://doi.org/10.1038/s41467-018-07752-3CrossRefGoogle ScholarPubMed
Blethyn, K. L., Hughes, S. W., Tóth, T. I., Cope, D. W., & Crunelli, V. (2006). Neuronal basis of the slow (< Hz) oscillation in neurons of the nucleus reticularis thalami in vitro. Journal of Neuroscience, 26(9), 24742486. https://doi.org/10.1523/JNEUROSCI.3607-05.2006Google Scholar
Bonjean, M., Baker, T., Lemieux, M., Timofeev, I., Sejnowski, T., & Bazhenov, M. (2011). Corticothalamic feedback controls sleep spindle duration in vivo. Journal of Neuroscience, 31(25), 91249134. https://doi.org/10.1523/JNEUROSCI.0077-11.2011Google Scholar
Boucetta, S., Cisse, Y., Mainville, L., Morales, M., & Jones, B. E. (2014). Discharge profiles across the sleep–waking cycle of identified cholinergic, GABAergic, and glutamatergic neurons in the pontomesencephalic tegmentum of the rat. Journal of Neuroscience, 34(13), 47084727. https://doi.org/10.1523/JNEUROSCI.2617-13.2014Google Scholar
Broicher, T., Wettschureck, N., Munsch, T., Coulon, P., Meuth, S. G., Kanyshkova, T., Seidenbecher, T., Offermanns, S., Pape, H. C., & Budde, T. (2008). Muscarinic ACh receptor-mediated control of thalamic activity via G q/G11-family G-proteins. Pflugers Archive European Journal of Physiology, 456(6), 10491060. https://doi.org/10.1007/s00424-008-0473-xGoogle Scholar
Brown, E. N., Lydic, R., & Schiff, N. D. (2018). General anesthesia, sleep, and coma. New England Journal of Medicine, 363(27), 26382650. https://doi.org/10.1001/jama.306.20.2283Google Scholar
Campbell, S. S., & Tobler, I. (1984). Animal sleep: a review of sleep duration across phylogeny. Neuroscience and Biobehavioral Reviews, 8(3), 269300. https://doi.org/10.1016/0149-7634(84)90054-XGoogle Scholar
Chapin, E. M., & Andrade, R. (2001). A 5-HT7 receptor-mediated depolarization in the anterodorsal thalamus. II. Involvement of the hyperpolarization-activated current Ih. Journal of Pharmacology and Experimental Therapeutics, 297(1), 403409.Google Scholar
Chauvette, S., Seigneur, J., & Timofeev, I. (2012). Sleep Oscillations in the thalamocortical system induce long-term neuronal plasticity. Neuron, 75(6), 11051113. https://doi.org/10.1016/j.neuron.2012.08.034Google Scholar
Chen, K. S., Xu, M., Zhang, Z., Chang, W. C., Gaj, T., Schaffer, D. V., & Dan, Y. (2018). A hypothalamic switch for REM and non-REM sleep. Neuron, 97(5), 1168–1176.e4. https://doi.org/10.1016/j.neuron.2018.02.005Google Scholar
Chung, S., Weber, F., Zhong, P., Tan, C. L., Nguyen, T. N., Beier, K. T., Hörmann, N., Chang, W. C., Zhang, Z., Do, J. P., Yao, S., Krashes, M. J., Tasic, B., Cetin, A., Zeng, H., Knight, Z. A., Luo, L., & Dan, Y. (2017). Identification of preoptic sleep neurons using retrograde labelling and gene profiling. Nature, 545(7655), 477481. https://doi.org/10.1038/nature22350CrossRefGoogle ScholarPubMed
Churgin, M. A., Szuperak, M., Davis, K. C., Raizen, D. M., Fang-Yen, C., & Kayser, M. S. (2019). Quantitative imaging of sleep behavior in Caenorhabditis elegans and larval Drosophila melanogaster. Nature Protocols, 14(5), 15551588. https://doi.org/10.1038/s41596-019-0146-6Google Scholar
Clark, M., McDevitt, R., & Neumaier, J. (2006). Quantitative mapping of tryptophan hydroxylase-2, 5-HT1A, 5-HT1B, and serotonin transporter expression across the anteroposterior axis of the rat dorsal and median raphe nuclei. Journal of Comparative Neurology, 498(5), 611623. https://doi.org/10.1002/cneGoogle Scholar
Colangelo, C., Shichkova, P., Keller, D., & Markram, H. (2019). Cellular, synaptic and network effects of acetylcholine in the neocortex. Frontiers in Neural Circuits, 13, 24. https://doi.org/10.3389/fncir.2019.00024Google Scholar
Contreras, D., Dossi, R. C., & Steriade, M. (1992). Bursting and tonic discharges in two classes of reticular thalamic neurons. Journal of Neurophysiology, 68(3), 973977. https://doi.org/10.1152/jn.1992.68.3.973Google Scholar
Crandall, S. R., Govindaiah, G., & Cox, C. L. (2010). Low-threshold Ca2+ current amplifies distal dendritic signaling in thalamic reticular neurons. Journal of Neuroscience, 30(46), 1541915429. https://doi.org/10.1523/JNEUROSCI.3636-10.2010Google Scholar
Crunelli, V., & Hughes, S. W. (2010). The slow (1 Hz) rhythm of non-REM sleep: a dialogue between three cardinal oscillators. Nature Neuroscience, 13(1), 917. https://doi.org/10.1038/nn.2445CrossRefGoogle ScholarPubMed
Crunelli, V., Larincz, M. L., Connelly, W. M., David, F., Hughes, S. W., Lambert, R. C., Leresche, N., & Errington, A. C. (2018). Dual function of thalamic low-vigilance state oscillations: rhythm-regulation and plasticity. Nature Reviews Neuroscience, 19(2), 107118. https://doi.org/10.1038/nrn.2017.151Google Scholar
Cueni, L., Canepari, M., Lujan, R., Emmenegger, Y., Watanabe, M., Bond, C. T., Franken, P., Adelman, J. P., & Luthi, A. (2008). T-type Ca2+ channels, SK2 channels and SERCAs gate sleep-related oscillations in thalamic dendrites. Nature Neuroscience, 11(6), 683692. https://doi.org/10.1038/nn.2124Google Scholar
Curró Dossi, R., Pare, D., & Steriade, M. (1991). Short-lasting nicotinic and long-lasting muscarinic depolarizing responses of thalamocortical neurons to stimulation of mesopontine cholinergic nuclei. Journal of Neurophysiology, 65(3), 393406. https://doi.org/10.1152/jn.1991.65.3.393Google Scholar
Dahan, L., Astier, B., Vautrelle, N., Urbain, N., Kocsis, B., & Chouvet, G. (2007). Prominent burst firing of dopaminergic neurons in the ventral tegmental area during paradoxical sleep. Neuropsychopharmacology, 32, 12321241. https://doi.org/10.1038/sj.npp.1301251Google Scholar
Dang-vu, T. T., Buxton, O. M., & Solet, J. M. (2010). Spontaneous brain rhythms predict sleep stability in the face of noise. Current Biology, 20(15), 626627. https://doi.org/10.1016/j.cub.2010.06.032Google Scholar
David, F., Schmiedt, J. T., Taylor, H. L., Orban, G., Di Giovanni, G., Uebele, V. N., Renger, J. J., Lambert, R. C., Leresche, N., & Crunelli, V. (2013). Essential thalamic contribution to slow waves of natural sleep. Journal of Neuroscience, 33(50), 1959919610. https://doi.org/10.1523/JNEUROSCI.3169-13.2013Google Scholar
De-miguel, F. F., Leon-pinzon, C., Noguez, P., & Mendez, B. (2015). Serotonin release from the neuronal cell body and its long-lasting effects on the nervous system. Philosophical Transactions of the Royal Society B, 370(1672).Google Scholar
Deschfines, M., Paradis, M., Roy, J. P., & Steriade, M. (1984). Electrophysiology of neurons of lateral thalamic nuclei in cat: resting properties and burst discharges. Journal of Neurophysiology, 51(6), 11961219.Google Scholar
Destexhe, A., Hughes, S. W., Rudolph, M., & Crunelli, V. (2007). Are corticothalamic “up” states fragments of wakefulness? Trends in Neurosciences, 30(7), 334342. https://doi.org/10.1016/j.tins.2007.04.006Google Scholar
Devilbiss, D. M., & Waterhouse, B. D. (2011). Phasic and tonic patterns of locus coeruleus output differentially modulate sensory network function in the awake rat. Journal of Neurophysiology, 105(1), 6987. https://doi.org/10.1152/jn.00445.2010Google Scholar
Durkin, J., Suresh, A. K., Colbath, J., Broussard, C., Wu, J., Zochowski, M., & Aton, S. J. (2017). Cortically coordinated NREM thalamocortical oscillations play an essential, instructive role in visual system plasticity. Proceedings of the National Academy of Sciences of the United States of America, 114(39), 1048510490. https://doi.org/10.1073/pnas.1710613114Google Scholar
Eban-Rothschild, A., Rothschild, G., Giardino, W. J., Jones, J. R., & De Lecea, L. (2016). VTA dopaminergic neurons regulate ethologically relevant sleep-wake behaviors. Nature Neuroscience, 19(10), 13561366. https://doi.org/10.1038/nn.4377Google Scholar
Economo, M. N., Viswanathan, S., Tasic, B., Bas, E., Winnubst, J., Menon, V., Graybuck, L. T., Nguyen, T. N., Smith, K. A., Yao, Z., Wang, L., Gerfen, C. R., Chandrashekar, J., Zeng, H., Looger, L. L., & Svoboda, K. (2018). Distinct descending motor cortex pathways and their roles in movement. Nature, 563(7729), 7984. https://doi.org/10.1038/s41586-018-0642-9Google Scholar
El Mansari, M., Sakai, K., & Jouvet, M. (1989). Unitary characteristics of presumptive cholinergic tegmental neurons during the sleep-waking cycle in freely moving cats. Experimental Brain Research, 76(3), 519529. https://doi.org/10.1007/BF00248908Google Scholar
Errington, A. C., Renger, J. J., Uebele, V. N., & Crunelli, V. (2010). State-dependent firing determines intrinsic dendritic Ca2+ signaling in thalamocortical neurons. Journal of Neuroscience, 30(44), 1484314853. https://doi.org/10.1523/JNEUROSCI.2968-10.2010Google Scholar
Eschenko, O., Magri, C., Panzeri, S., & Sara, S. J. (2012). Noradrenergic neurons of the locus coeruleus are phase locked to cortical up-down states during sleep. Cerebral Cortex, 22, 426435. https://doi.org/10.1093/cercor/bhr121CrossRefGoogle ScholarPubMed
Favero, M., Varghese, G., & Castro-Alamancos, M. A. (2012). The state of somatosensory cortex during neuromodulation. Journal of Neurophysiology, 108(4), 10101024. https://doi.org/10.1152/jn.00256.2012Google Scholar
Fernandez, L. M. J., Vantomme, G., Osorio-Forero, A., Cardis, R., Béard, E., & Lüthi, A. (2018). Thalamic reticular control of local sleep in mouse sensory cortex. eLife, 7, 125. https://doi.org/10.7554/eLife.39111Google Scholar
Ferrarelli, F., & Tononi, G. (2017). Reduced sleep spindle activity point to a TRN-MD thalamus-PFC circuit dysfunction in schizophrenia. Schizophrenia Research, 180, 3643. https://doi.org/10.1016/j.schres.2016.05.023Google Scholar
Fort, P., Bassetti, C. L., & Luppi, P. H. (2009). Alternating vigilance states: new insights regarding neuronal networks and mechanisms. European Journal of Neuroscience, 29(9), 17411753. https://doi.org/10.1111/j.1460-9568.2009.06722.xGoogle Scholar
Francesconi, W., Muller, C. M., & Singer, W. (1988). Cholinergic mechanisms in the reticular control of transmission in the cat lateral geniculate nucleus. Journal of Neurophysiology, 59(6), 16901718. https://doi.org/10.1152/jn.1988.59.6.1690Google Scholar
Fuller, P., Sherman, D., Pedersen, N. P., Saper, C. B., & Lu, J. (2011). Reassessment of the structural basis of the ascending arousal system. Journal of Comparative Neurology, 519(5), 933956. https://doi.org/10.1002/cne.22559Google Scholar
Gazea, M., Furdan, S., Sere, P., Oesch, L., Molnár, B., Giovanni, G. Di, Fenno, L. E., Ramakrishnan, C., Mattis, J., Deisseroth, K., Dymecki, S. M., Adamantidis, A. R., & Lőrincz, M. L. (2021). Reciprocal lateral hypothalamic and raphe GABAergic projections promote wakefulness. Journal of Neuroscience, 41(22), 48404849. https://doi.org/10.1523/JNEUROSCI.2850-20.2021Google Scholar
Gent, T. C., Bandarabadi, M., Herrera, C. G., & Adamantidis, A. R. (2018). Thalamic dual control of sleep and wakefulness. Nature Neuroscience, 21(7), 974984. https://doi.org/10.1038/s41593-018-0164-7Google Scholar
Gent, T. C., Bassetti, C. L. A., & Adamantidis, A. R. (2018). Sleep-wake control and the thalamus. Current Opinion in Neurobiology, 52, 188197. https://doi.org/10.1016/j.conb.2018.08.002Google Scholar
Glin, L., Arnaud, C., Berracochea, D., Galey, D., Jaffard, R., & Gottesmann, C. (1991). The intermediate stage of sleep in mice. Physiology and Behavior, 50(5), 951953. https://doi.org/10.1016/0031–9384(91)90420-SGoogle Scholar
Guilleminault, C., Quera-salva, M. A., & Goldberg, M. P. (1993). Pseudo-hypersomnia and pre-sleep behaviour with bilateral paramedian thalamic lesions. Brain, 116(6), 15491563. https://doi.org/10.1093/brain/116.6.1549Google Scholar
Halassa, M. M., & Kastner, S. (2017). Thalamic functions in distributed cognitive control. Nature Neuroscience, 20(12), 16691679. https://doi.org/10.1038/s41593-017-0020-1Google Scholar
Halassa, M. M., Siegle, J. H., Ritt, J. T., Ting, J. T., Feng, G., & Moore, C. I. (2011). Selective optical drive of thalamic reticular nucleus generates thalamic bursts and cortical spindles. Nature Neuroscience, 14(9), 11181120. https://doi.org/10.1038/nn.2880Google Scholar
Hassani, O. K., Henny, P., Lee, M. G., & Jones, B. E. (2010). GABAergic neurons intermingled with orexin and MCH neurons in the lateral hypothalamus discharge maximally during sleep. European Journal of Neuroscience, 32(3), 448457. https://doi.org/10.1111/j.1460-9568.2010.07295.xGoogle Scholar
Hassani, O. K., Lee, M. G., & Jones, B. E. (2009). Melanin-concentrating hormone neurons discharge in a reciprocal manner to orexin neurons across the sleep-wake cycle. Proceedings of the National Academy of Sciences of the United States of America, 106(7), 24182422. https://doi.org/10.1073/pnas.0811400106Google Scholar
Hayashi, Y., Kashiwagi, M., Yasuda, K., Ando, R., Kanuka, M., Sakai, K., & Itohara, S. (2015). Cells of a common developmental origin regulate REM/non-REM sleep and wakefulness in mice. Science, 350(6263), 957962.Google Scholar
Hebb, D. O. (1949). The organization of behavior: a neuropsychological theory. New York: Wiley.Google Scholar
Helfrich, R. F., Lendner, J. D., Mander, B. A., Guillen, H., Paff, M., Mnatsakanyan, L., Vadera, S., Walker, M. P., Lin, J. J., & Knight, R. T. (2019). Bidirectional prefrontal-hippocampal dynamics organize information transfer during sleep in humans. Nature Communications, 10(1), 116. https://doi.org/10.1038/s41467-019-11444-xGoogle Scholar
Hermann, D. M., Siccoli, M., Brugger, P., Wachter, K., Mathis, J., Achermann, P., & Bassetti, C. L. (2008). Evolution of neurological, neuropsychological and sleep-wake disturbances after paramedian thalamic stroke. Stroke, 39(1), 6268. https://doi.org/10.1161/STROKEAHA.107.494955CrossRefGoogle ScholarPubMed
Herrera, C. G., Cadavieco, M. C., Jego, S., Ponomarenko, A., Korotkova, T., & Adamantidis, A. (2016). Hypothalamic feedforward inhibition of thalamocortical network controls arousal and consciousness. Nature Neuroscience, 19(2), 290298. https://doi.org/10.1038/nn.4209Google Scholar
Hong, J., Ha, G. E., Kwak, H., Lee, Y., Jeong, H., Suh, P. G., & Cheong, E. (2020). Destabilization of light NREM sleep by thalamic PLCβ4 deletion impairs sleep-dependent memory consolidation. Scientific Reports, 10(1), 114. https://doi.org/10.1038/s41598-020-64377-7Google Scholar
Honjoh, S., Sasai, S., Schiereck, S. S., Nagai, H., Tononi, G., & Cirelli, C. (2018). Regulation of cortical activity and arousal by the matrix cells of the ventromedial thalamic nucleus. Nature Communications, 9(1). https://doi.org/10.1038/s41467-018-04497-xGoogle Scholar
Hopkins, W., & Johnston, D. (1984). Frequency-dependent noradrenergic modulation of long-term potentiation in the hippocampus abstract. Science, 226(4672), 350352.Google Scholar
Hubbard, J., Gent, T. C., Hoekstra, M. M. B., Emmenegger, Y., Mongrain, V., Landolt, H. P., Adamantidis, A. R., & Franken, P. (2020). Rapid fast-delta decay following prolonged wakefulness marks a phase of wake-inertia in NREM sleep. Nature Communications, 11(1), 116. https://doi.org/10.1038/s41467-020–16915-0Google Scholar
Huber, R., Ghilardi, M. F., Massimini, M., & Tononi, G. (2004). Local sleep and learning. Nature, 430(6995), 7881. https://doi.org/10.1038/nature02663Google Scholar
Hughes, S. W., Cope, D. W., Blethyn, K. L., & Crunelli, V. (2002). Cellular mechanisms of the slow (< Hz) oscillation in thalamocortical neurons in vitro. Neuron, 33(6), 947958. https://doi.org/10.1016/S0896-6273(02)00623-2Google Scholar
Hughes, S. W., Lo, M., Cope, D. W., Blethyn, K. L., Ke, K. A., & Parri, H. R. (2004). Synchronized oscillations at α and θ frequencies in the lateral geniculate nucleus. Neuron, 42, 253268.Google Scholar
Jing, M., Li, Y., Zeng, J., Huang, P., Skirzewski, M., Kljakic, O., Peng, W., Qian, T., Tan, K., Zou, J., Trinh, S., Wu, R., Zhang, S., Pan, S., Hires, S. A., Xu, M., Li, H., Saksida, L. M., Prado, V. F., … Li, Y. (2020). An optimized acetylcholine sensor for monitoring in vivo cholinergic activity. Nature Methods, 17(11), 11391146. https://doi.org/10.1038/s41592-020-0953-2Google Scholar
Jones, B. E. (2005). From waking to sleeping: neuronal and chemical substrates. Trends in Pharmacological Sciences, 26(11), 578586. https://doi.org/10.1016/j.tips.2005.09.009Google Scholar
Jones, B. E., & Yang, T. ‐Z. (1985). The efferent projections from the reticular formation and the locus coeruleus studied by anterograde and retrograde axonal transport in the rat. Journal of Comparative Neurology, 242(1), 5692. https://doi.org/10.1002/cne.902420105Google Scholar
Jouvet, M., Michel, F., & Courjon, J. (1959). Sur un stade d’activité éléctrique cérébrale rapide au cours du sommeil physiologique. Comptes Rendus des Seances de la Societe de Biologie et de Ses Filiales, 153, 10241028.Google Scholar
Kayama, Y., Ohta, M., & Jodo, E. (1992). Firing of “possibly” cholinergic neurons in the rat laterodorsal tegmental nucleus during sleep and wakefulness. Brain Research, 569(2), 210220. https://doi.org/10.1016/0006-8993(92)90632-JGoogle Scholar
Kayser, M. S., Yue, Z., & Sehgal, A. (2014). A critical period of sleep for development of courtship circuitry and behavior in Drosophila. Science, 344(6181), 269274. https://doi.org/10.1126/science.1250553Google Scholar
Kim, A., Latchoumane, C., Lee, S., Kim, G. B., Cheong, E., Augustine, G. J., & Shin, H. S. (2012). Optogenetically induced sleep spindle rhythms alter sleep architectures in mice. Proceedings of the National Academy of Sciences of the United States of America, 109(50), 2067320678. https://doi.org/10.1073/pnas.1217897109Google Scholar
Kirszenblat, L., & van Swinderen, B. (2015). The yin and yang of sleep and attention. Trends in Neurosciences, 38(12), 776786. https://doi.org/10.1016/j.tins.2015.10.001Google Scholar
Kocsis, B., Varga, V., Dahan, L., & Sik, A. (2006). Serotonergic neuron diversity: identification of raphe neurons with discharge time-locked to the hippocampal theta rhythm. Proceedings of the National Academy of Sciences of the United States of America, 103(4), 10591064. https://doi.org/10.1073/pnas.0508360103Google Scholar
Kolmac, C., & Mitrofanis, J. (1999). Organization of the basal forebrain projection to the thalamus in rats. Neuroscience Letters, 272(3), 151154. https://doi.org/10.1016/S0304-3940(99)00614-XGoogle Scholar
Kroeger, D., Ferrari, L. L., Petit, G., Mahoney, C. E., Fuller, P. M., Arrigoni, E., & Scammell, T. E. (2017). Cholinergic, glutamatergic, and GABAergic neurons of the pedunculopontine tegmental nucleus have distinct effects on sleep/wake behavior in mice. Journal of Neuroscience, 37(5), 13521366. https://doi.org/10.1523/JNEUROSCI.1405-16.2016Google Scholar
Krone, L. B., Yamagata, T., Blanco-duque, C., Guillaumin, M. C. C., Kahn, M. C., Vinne, V. Van Der, Mckillop, Tam, L. E., Peirson, S. K. E., Akerman, S. N., Hoerder-suabedissen, C. J., Molnár, A., Z., & Vyazovskiy, V. V. (2021). A role for the cortex in sleep–wake regulation. Nature Neuroscience, 24(September). https://doi.org/10.1038/s41593-021–00894-6Google Scholar
Krueger, J. M., Nguyen, J. T., Dykstra-Aiello, C. J., & Taishi, P. (2019). Local sleep. Sleep Medicine Reviews, 43, 1421. https://doi.org/10.1016/j.smrv.2018.10.001Google Scholar
Krueger, J. M., & Tononi, G. (2012). Local use-dependent sleep; synthesis of the new paradigm. Current Topics in Medicinal Chemistry, 11(19), 24902492. https://doi.org/10.2174/156802611797470330Google Scholar
Larkum, M. (2013). A cellular mechanism for cortical associations: an organizing principle for the cerebral cortex. Trends in Neurosciences, 36(3), 141151. https://doi.org/10.1016/j.tins.2012.11.006Google Scholar
Latchoumane, C. F. V., Ngo, H. V. V., Born, J., & Shin, H. S. (2017). Thalamic spindles promote memory formation during sleep through triple phase-locking of cortical, thalamic, and hippocampal rhythms. Neuron, 95(2), 424435.e6. https://doi.org/10.1016/j.neuron.2017.06.025Google Scholar
Lecci, S., Fernandez, L. M. J., Weber, F. D., Cardis, R., Chatton, J.-Y., Born, J., & Luthi, A. (2017). Coordinated infraslow neural and cardiac oscillations mark fragility and offline periods in mammalian sleep. Science Advances, 3. https://doi.org/10.3389/fphys.2017.00847Google Scholar
Lee, J., Kim, D., & Shin, H. (2004). Lack of delta waves and sleep disturbances during non-rapid eye movement sleep in mice lacking α1 G-subunit of T-type calcium channels. Proceedings of the National Academy of Sciences of the United States of America, 101(52), 1819518199.Google Scholar
Lemieux, M., Chen, J. Y., Lonjers, P., Bazhenov, M., & Timofeev, I. (2014). The impact of cortical deafferentation on the neocortical slow oscillation. Journal of Neuroscience, 34(16), 56895703. https://doi.org/10.1523/JNEUROSCI.1156-13.2014Google Scholar
Leung, L. C., Wang, G. X., Madelaine, R., Skariah, G., Kawakami, K., Deisseroth, K., Urban, A. E., & Mourrain, P. (2019). Neural signatures of sleep in zebrafish. Nature, 571(7764), 198204. https://doi.org/10.1038/s41586-019-1336-7Google Scholar
Lioudyno, M. I., Birch, A. M., Tanaka, B. S., Sokolov, Y., Goldin, A. L., Chandy, K. G., Hall, J. E., & Alkire, M. T. (2013). Shaker-related potassium channels in the central medial nucleus of the thalamus are important molecular targets for arousal suppression by volatile general anesthetics. Journal of Neuroscience, 33(41), 1631016322. https://doi.org/10.1523/JNEUROSCI.0344-13.2013Google Scholar
Liu, D., Li, W., Ma, C., Zheng, W., Yao, Y., Tso, C. F., Zhong, P., Chen, X., Song, J. H., Choi, W., Paik, S. B., Han, H., & Dan, Y. (2020). A common hub for sleep and motor control in the substantia nigra. Science, 367(6476), 440445. https://doi.org/10.1126/science.aaz0956Google Scholar
Liu, K., Kim, J., Kim, D. W., Zhang, Y. S., Bao, H., Denaxa, M., Lim, S. A., Kim, E., Liu, C., Wickersham, I. R., Pachinis, V., Hattar, S., Song, J., Brown, S. P., & Blackshaw, S. (2017). Lhx6-positive GABA-releasing neurons of the zona incerta promote sleep. Nature, 548(7669), 582587. https://doi.org/10.1038/nature23663Google Scholar
Llinás, R. R., & Steriade, M. (2006). Bursting of thalamic neurons and states of vigilance. Journal of Neurophysiology, 95(6), 32973308. https://doi.org/10.1152/jn.00166.2006Google Scholar
Loomis, L. A., Newton, E. H., & Hobart, A. G. (1938). Distribution of disturbance-patterns in the human electroencephalogram, with special reference to sleep. Journal of Neurophysiology, 1, 413430.Google Scholar
Lőrincz, M. L., Gunner, D., Bao, Y., Connelly, W. M., Isaac, J. T. R., Hughes, S. W., & Crunelli, V. (2015). A distinct class of slow (~0.2–2 Hz) intrinsically bursting layer 5 pyramidal neurons determines UP/DOWN state dynamics in the neocortex. Journal of Neuroscience, 35(14), 54425458. https://doi.org/10.1523/JNEUROSCI.3603-14.2015Google Scholar
Lugaresi, E., Medori, R., Montagna, P., Baruzzi, A., Cortelli, P., Lugaresi, A., Tinuper, P., Zucconi, M., & Gambetti, P. (1986). Fatal familial insomnia and dysautonomia with selective degeneration of thalamic nuclei. New England Journal of Medicine, 315(16), 9971003. https://doi.org/10.1056/NEJM198610163151605Google Scholar
Luo, Y. J., Li, Y. D., Wang, L., Yang, S. R., Yuan, X. S., Wang, J., Cherasse, Y., Lazarus, M., Chen, J. F., Qu, W. M., & Huang, Z. L. (2018). Nucleus accumbens controls wakefulness by a subpopulation of neurons expressing dopamine D1 receptors. Nature Communications, 9(1). https://doi.org/10.1038/s41467-018-03889-3Google Scholar
Luppi, P. H., Peyron, C., & Fort, P. (2017). Not a single but multiple populations of GABAergic neurons control sleep. Sleep Medicine Reviews, 32, 8594. https://doi.org/10.1016/j.smrv.2016.03.002Google Scholar
Lüthi, A. (2014). Sleep spindles: where they come from, what they do. Neuroscientist, 20 (3), 243256. https://doi.org/10.1177/1073858413500854Google Scholar
Magnin, M., Bastuji, H., Garcia-Larrea, L., & Mauguière, F. (2004). Human thalamic medial pulvinar nucleus is not activated during paradoxical sleep. Cerebral Cortex, 14(8), 858862. https://doi.org/10.1093/cercor/bhh044Google Scholar
Mahon, S., Vautrelle, N., Pezard, L., Slaght, S. J., Deniau, J. M., Chouvet, G., & Charpier, S. (2006). Distinct patterns of striatal medium spiny neuron activity during the natural sleep-wake cycle. Journal of Neuroscience, 26(48), 1258712595. https://doi.org/10.1523/JNEUROSCI.3987-06.2006Google Scholar
Mai, J. K., & Majtanik, M. (2019). Toward a common terminology for the thalamus. Frontiers in Neuroanatomy, 12(January), 123. https://doi.org/10.3389/fnana.2018.00114Google Scholar
Mak-Mccully, R. A., Rolland, M., Sargsyan, A., Gonzalez, C., Magnin, M., Chauvel, P., Rey, M., Bastuji, H., & Halgren, E. (2017). Coordination of cortical and thalamic activity during non-REM sleep in humans. Nature Communications, 8(May). https://doi.org/10.1038/ncomms15499Google Scholar
Manoach, D. S., Pan, J. Q., Purcell, S. M., & Stickgold, R. (2016). Reduced sleep spindles in schizophrenia: a treatable endophenotype that links risk genes to impaired cognition? Biological Psychiatry, 80(8), 599608. https://doi.org/10.1016/j.biopsych.2015.10.003Google Scholar
Manoach, D. S., Thakkar, K. N., Stroynowski, E., Ely, A., McKinley, S. K., Wamsley, E., Djonlagic, I., Vangel, M. G., Goff, D. C., & Stickgold, R. (2010). Reduced overnight consolidation of procedural learning in chronic medicated schizophrenia is related to specific sleep stages. Journal of Psychiatric Research, 44(2), 112120. https://doi.org/10.1016/j.jpsychires.2009.06.011Google Scholar
Maquet, P., Laureys, S., Peigneux, P., Fuchs, S., Petiau, C., Phillips, C., Aerts, J., Del Fiore, G., Degueldre, C., Meulemans, T., Luxen, A., Franck, G., Van Der Linden, M., Smith, C., & Cleeremans, A. (2000). Experience-dependent changes in changes in cerebral activation during human REM sleep. Nature Neuroscience, 3(8), 831836. https://doi.org/10.1038/77744Google Scholar
Maquet, P., Peters, J., Aerts, J., Delfiore, G., Degueldre, C., Luxen, A., & Franck, G. (1996). Functional neuroanatomy of human rapid-eye-movement sleep and dreaming. Nature, 383, 163166.CrossRefGoogle ScholarPubMed
Massimini, M., Ferrarelli, F., Esser, S. K., Riedner, B. A., Huber, R., Murphy, M., Peterson, M. J., & Tononi, G. (2007). Triggering sleep slow waves by transcranial magnetic stimulation. Proceedings of the National Academy of Sciences of the United States of America, 104(20), 84968501.Google Scholar
Massimini, M., Huber, R., Ferrarelli, F., Hill, S., & Tononi, G. (2004). The sleep slow oscillation as a traveling wave. Journal of Neuroscience, 24(31), 68626870. https://doi.org/10.1523/JNEUROSCI.1318-04.2004Google Scholar
McBride, R. L., & Sutin, J. (1976). Projections of the locus coeruleus and adjacent pontine tegmentum in the cat. Journal of Comparative Neurology, 165(3), 265284. https://doi.org/10.1002/cne.901650302Google Scholar
McCormick, D. A., McGinley, M. J., & Salkoff, D. B. (2015). Brain state dependent activity in the cortex and thalamus. Current Opinion in Neurobiology, 31, 133140. https://doi.org/10.1016/j.conb.2014.10.003Google Scholar
McCormick, D. A., & Pape, H. C. (1988). Acetylcholine inhibits identified interneurons in the cat lateral geniculate nucleus. Nature, 334(6179), 246248. https://doi.org/10.1038/334246a0Google Scholar
McCormick, D. A., & Pape, H. C. (1990). Noradrenergic and serotoninergic modulation of a hyperpolarization-activated cation current in thalamic relay neurons. Journal of Physiology, 431, 319342.Google Scholar
McCormick, D. A., Pape, H. C., & Williamson, A. (1991). Actions of norepinephrine in the cerebral cortex and thalamus: implications for function of the central noradrenergic system. Progress in Brain Research, 88(C), 293305. https://doi.org/10.1016/S0079-6123(08)63817–0Google Scholar
McCormick, D. A., & Prince, D. A. (1986). Acetylcholine induces burst firing in thalamic reticular neurones by activating a potassium conductance. Nature, 319, 402405.Google Scholar
McKenna, J. T., & Vertes, R. P. (2004). Afferent projections to nucleus reuniens of the thalamus. Journal of Comparative Neurology, 480(2), 115142. https://doi.org/10.1002/cne.20342Google Scholar
Mesbah-Oskui, L., Horner, R. L., Orser, B. A., & Horner, R. L. (2014). Thalamic δ-subunit containing GABAA receptors promote electrocortical signatures of deep non-REM sleep but do not mediate the effects of etomidate at the thalamus in vivo. Journal of Neuroscience, 34(37), 1225312266. https://doi.org/10.1523/JNEUROSCI.0618-14.2014Google Scholar
Mölle, M., Bergmann, T. O., Marshall, L., & Born, J. (2011). Fast and slow spindles during the sleep slow oscillation: disparate coalescence and engagement in memory processing. Sleep, 34(10), 14111421. https://doi.org/10.5665/SLEEP.1290Google Scholar
Morairty, S. R., Dittrich, L., Pasumarthi, R. K., Valladao, D., Heiss, J. E., Gerashchenko, D., & Kilduff, T. S. (2013). A role for cortical nNOS/NK1 neurons in coupling homeostatic sleep drive to EEG slow wave activity. Proceedings of the National Academy of Sciences of the United States of America, 110(50), 2027220277. https://doi.org/10.1073/pnas.1314762110Google Scholar
Moruzzi, G., & Magoun, H. W. (1949). Brain stem reticular formation and activation of the EEG. Electroencephalography and Clinical Neurophysiology, 1(1–4), 455473.Google Scholar
Moxon, K. A., Devilbiss, D. M., Chapin, J. K., & Waterhouse, B. D. (2007). Influence of norepinephrine on somatosensory neuronal responses in the rat thalamus: a combined modeling and in vivo multi-channel, multi-neuron recording study. Brain Research, 1147(1), 105123. https://doi.org/10.1016/j.brainres.2007.02.006Google Scholar
Muehlroth, B. E., Sander, M. C., Fandakova, Y., Grandy, T. H., Rasch, B., Shing, Y. L., & Werkle-Bergner, M. (2019). Precise slow oscillation–spindle coupling promotes memory consolidation in younger and older adults. Scientific Reports, 9(1), 115. https://doi.org/10.1038/s41598-018-36557-zGoogle Scholar
Nath, R. D., Bedbrook, C. N., Abrams, M. J., Basinger, T., Bois, J. S., Prober, D. A., Sternberg, P. W., Gradinaru, V., & Goentoro, L. (2017). The jellyfish Cassiopea exhibits a sleep-like state. Current Biology, 27(19), 2984–2990.e3. https://doi.org/10.1016/j.cub.2017.08.014Google Scholar
Newman, E. L., Gupta, K., Climer, J. R., Monaghan, C. K., & Hasselmo, M. E. (2012). Cholinergic modulation of cognitive processing: insights drawn from computational models. Frontiers in Behavioral Neuroscience. https://doi.org/10.3389/fnbeh.2012.00024Google Scholar
Nichols, A. L. A., Eichler, T., Latham, R., & Zimmer, M. (2017). A global brain state underlies C. elegans sleep behavior. Science, 356(6344), 12471256. https://doi.org/10.1126/science.aam6851Google Scholar
Oishi, Y., Xu, Q., Wang, L., Zhang, B. J., Takahashi, K., Takata, Y., Luo, Y. J., Cherasse, Y., Schiffmann, S. N., De Kerchove D’Exaerde, A., Urade, Y., Qu, W. M., Huang, Z. L., & Lazarus, M. (2017). Slow-wave sleep is controlled by a subset of nucleus accumbens core neurons in mice. Nature Communications, 8(1), 112. https://doi.org/10.1038/s41467-017–00781-4Google Scholar
Pace-Schott, E. F., & Hobson, J. A. (2002). The neurobiology of sleep: genetics, cellular physiology and subcortical networks. Nature Reviews Neuroscience, 3(8), 591605. https://doi.org/10.1038/nrn895Google Scholar
Pape, H. C., & McCormick, D. A. (1989). Noradrenaline and serotonin selectively modulate thalamic burst firing by enhancing a hyperpolarization-activated cation current. Nature, 340, 715718. https://doi.org/10.1038/246170a0Google Scholar
Parent, A., Paré, D., Smith, Y., & Steriade, M. (1988). Basal forebrain cholinergic and noncholinergic projections to the thalamus and brainstem in cats and monkeys. Journal of Comparative Neurology, 277(2), 281301. https://doi.org/10.1002/cne.902770209Google Scholar
Pedersen, N. P., Ferrari, L., Venner, A., Wang, J. L., Abbott, S. B. G., Vujovic, N., Arrigoni, E., Saper, C. B., & Fuller, P. M. (2017). Supramammillary glutamate neurons are a key node of the arousal system. Nature Communications, 8(1). https://doi.org/10.1038/s41467-017-01004-6Google Scholar
Pellegrini, C., Lecci, S., Lüthi, A., & Astori, S. (2016). Suppression of sleep spindle rhythmogenesis in mice with deletion of CaV3.2 and CaV3.3 T-type Ca2+ channels. Sleep, 39(4), 875885.Google Scholar
Phillips, J. W., Schulmann, A., Hara, E., Winnubst, J., Liu, C., Valakh, V., Wang, L., Shields, B. C., Korff, W., Chandrashekar, J., Lemire, A. L., Mensh, B., Dudman, J. T., Nelson, S. B., & Hantman, A. W. (2019). A repeated molecular architecture across thalamic pathways. Nature Neuroscience, 22(11), 19251935. https://doi.org/10.1038/s41593-019-0483-3Google Scholar
Piantoni, G., Halgren, E., & Cash, S. S. (2017).Spatiotemporal characteristics of sleep spindles depend on cortical location. NeuroImage, 146(June), 236245. https://doi.org/10.1016/j.neuroimage.2016.11.010Google Scholar
Picciotto, M. R., Higley, M. J., & Mineur, Y. S. (2012). Acetylcholine as a neuromodulator: cholinergic signaling shapes nervous system function and behavior. Neuron, 76(1), 116129. https://doi.org/10.1016/j.neuron.2012.08.036Google Scholar
Pita-Almenar, J. D., Yu, D., Lu, H. C., & Beierlein, M. (2014). Mechanisms underlying desynchronization of cholinergic-evoked thalamic network activity. Journal of Neuroscience, 34(43), 1446314474. https://doi.org/10.1523/JNEUROSCI.2321-14.2014Google Scholar
Portas, C. M., Bjorvatn, B., & Ursin, R. (2000). Serotonin and the sleep/wake cycle: special emphasis on microdialysis studies. Progress in Neurobiology, 60, 1335.Google Scholar
Poulet, J. F. A., Fernandez, L. M. J., Crochet, S., & Petersen, C. C. H. (2012). Thalamic control of cortical states. Nature Neuroscience, 15(3), 370372. https://doi.org/10.1038/nn.3035Google Scholar
Poulet, J. F. A., & Petersen, C. C. H. (2008). Internal brain state regulates membrane potential synchrony in barrel cortex of behaving mice. Nature, 454(7206), 881885. https://doi.org/10.1038/nature07150Google Scholar
Puentes-Mestril, C., & Aton, S. J. (2017). Linking network activity to synaptic plasticity during sleep: Hypotheses and recent data. Frontiers in Neural Circuits, 11(September), 118. https://doi.org/10.3389/fncir.2017.00061Google Scholar
Qiu, M. H., Chen, M. C., Fuller, P. M., & Lu, J. (2016). Stimulation of the Pontine parabrachial nucleus promotes wakefulness via extra-thalamic forebrain circuit nodes. Current Biology, 26(17), 23012312. https://doi.org/10.1016/j.cub.2016.07.054Google Scholar
Raizen, D. M., Zimmerman, J. E., Maycock, M. H., Ta, U. D., You, Y. J., Sundaram, M. V., & Pack, A. I. (2008). Lethargus is a Caenorhabditis elegans sleep-like state. Nature, 451(7178), 569572. https://doi.org/10.1038/nature06535Google Scholar
Rikhye, R. V., Wimmer, R. D., & Halassa, M. M. (2018). Toward an integrative theory of thalamic function. Annual Review of Neuroscience, 41(March), 163183. https://doi.org/10.1146/annurev-neuro-080317-062144Google Scholar
Rodenkirch, C., Liu, Y., Schriver, B. J., & Wang, Q. (2019). Locus coeruleus activation enhances thalamic feature selectivity via norepinephrine regulation of intrathalamic circuit dynamics. Nature Neuroscience, 22(1), 120133. https://doi.org/10.1038/s41593-018–0283-1Google Scholar
Sanchez-Vives, M. V., Descalzo, V. F., Reig, R., Figueroa, N. A., Compte, A., & Gallego, R. (2008). Rhythmic spontaneous activity in the piriform cortex. Cerebral Cortex, 18(5), 11791192. https://doi.org/10.1093/cercor/bhm152Google Scholar
Sanchez-Vives, M. V., & McCormick, D. A. (2000). Cellular and network mechanisms of rhythmic recurrent activity in neocortex. Nature Neuroscience, 3(10), 10271034. https://doi.org/10.1038/79848Google Scholar
Santamaria, J., Pujol, M., Orteu, N., Solanas, A., Cardenal, C., Santacruz, P., Chimeno, E., & Moon, P. (2000). Unilateral thalamic stroke does not decrease ipsilateral sleep spindles. Sleep, 23(3), 333339. https://doi.org/10.1093/sleep/23.3.1Google Scholar
Saper, C. B., Fuller, P. M., Pedersen, N. P., Lu, J., & Scammell, T. E. (2010). Sleep state switching. Neuron, 68(6), 10231042. https://doi.org/doi:10.1016/j.neuron.2010.11.032.Google Scholar
Sarasso, S., D’Ambrosio, S., Fecchio, M., Casarotto, S., Viganò, A., Landi, C., Mattavelli, G., Gosseries, O., Quarenghi, M., Laureys, S., Devalle, G., Rosanova, M., & Massimini, M. (2020). Local sleep-like cortical reactivity in the awake brain after focal injury. Brain, 143(12), 36723684. https://doi.org/10.1093/brain/awaa338Google Scholar
Schabus, M., Dang-Vu, T. T., Albouy, G., Balteau, E., Boly, M., Carrier, J., Darsaud, A., Degueldre, C., Desseilles, M., Gais, S., Phillips, C., Rauchs, G., Schnakers, C., Sterpenich, V., Vandewalle, G., Luxen, A., & Maquet, P. (2007). Hemodynamic cerebral correlates of sleep spindles during human non-rapid eye movement sleep. Proceedings of the National Academy of Sciences of the United States of America, 104(32), 1316413169. https://doi.org/10.1073/pnas.0703084104Google Scholar
Schabus, M., Dang-vu, T. T., Philip, D., Heib, J., Boly, M., Vandewalle, G., Schmidt, C., Albouy, G., Darsaud, A., & Gais, S. (2012). The fate of incoming stimuli during NREM sleep is determined by spindles and the phase of the slow oscillation. Frontiers in Neurology, 3(April), 111. https://doi.org/10.3389/fneur.2012.00040Google Scholar
Seibt, J., Richard, C. J., Sigl-Glöckner, J., Takahashi, N., Kaplan, D. I., Doron, G., De Limoges, D., Bocklisch, C., & Larkum, M. E. (2017). Cortical dendritic activity correlates with spindle-rich oscillations during sleep in rodents. Nature Communications, 8(1). https://doi.org/10.1038/s41467-017-00735-wGoogle Scholar
Seugnet, L., Suzuki, Y., Donlea, J. M., Gottschalk, L., & Shaw, P. J. (2011). Sleep deprivation during early-adult development results in long-lasting learning deficits in adult Drosophila. Sleep, 34(2), 137146. https://doi.org/10.1093/sleep/34.2.137Google Scholar
Sherman, S. M. (2016). Thalamus plays a central role in ongoing cortical functioning. Nature Neuroscience, 19(4), 533541. https://doi.org/10.1038/nn.4269Google Scholar
Siclari, F., Baird, B., Perogamvros, L., Bernardi, G., LaRocque, J. J., Riedner, B., Boly, M., Postle, B. R., & Tononi, G. (2017). The neural correlates of dreaming. Nature Neuroscience, 20(6), 872878. https://doi.org/10.1038/nn.4545Google Scholar
Siclari, F., Bernardi, G., Cataldi, J., & Tononi, G. (2018). Dreaming in NREM sleep: a high-density EEG study of slow waves and spindles. Journal of Neuroscience, 38(43), 91759185. https://doi.org/10.1523/JNEUROSCI.0855–18.2018Google Scholar
Siclari, , F., Bernardi, F., Riedner, B. A., LaRocque, J. J., Benca, R. M., & Tononi, G. (2014). Two distinct synchronization processes in the transition to sleep. Sleep, 37.Google Scholar
Siclari, F., & Tononi, G. (2017). Local aspects of sleep and wakefulness. Current Opinion in Neurobiology, 44, 222227. https://doi.org/10.1016/j.conb.2017.05.008Google Scholar
Sieber, A. R., Min, R., & Nevian, T. (2013). Non-Hebbian long-term potentiation of inhibitory synapses in the thalamus. Journal of Neuroscience, 33(40), 1567515685. https://doi.org/10.1523/JNEUROSCI.0247-13.2013Google Scholar
Sillito, A. M., Kemp, J. A., & Berardi, N. (1983). The cholinergic influence on the function of the cat dorsal lateral geniculate nucleus (dLGN). Brain Research, 280(2), 299307. https://doi.org/10.1016/0006–8993(83)90059–8CrossRefGoogle ScholarPubMed
Sriji, S., Akhtar, N., & Mallick, H. N. (2020). Mediodorsal thalamus lesion increases paradoxical sleep in rats. Sleep Science, 16. https://doi.org/10.5935/1984-0063.20190155Google Scholar
Stanton, P. K., & Sarvey, J. M. (1985). Depletion of norepinephrine, but not serotonin, reduces long-term potentiation in the dentate gyrus of rat hippocampal slices. Journal of Neuroscience, 5(8), 21692176. https://doi.org/10.1523/jneurosci.05-08-02169.1985Google Scholar
Steriade, M. (2006). Grouping of brain rhythms in corticothalamic systems. Neuroscience, 137(4), 10871106. https://doi.org/10.1016/j.neuroscience.2005.10.029Google Scholar
Steriade, M., Curro Dossi, R., Pare, D., & Oakson, G. (1991). Fast oscillations (20–40 Hz) in thalamocortical systems and their potentiation by mesopontine cholinergic nuclei in the cat. Proceedings of the National Academy of Sciences of the United States of America, 88(10), 43964400. https://doi.org/10.1073/pnas.88.10.4396Google Scholar
Steriade, M., Datta, S., Paré, D., Oakson, G., & Curró Dossi, R. (1990). Neuronal activities in brain-stem cholinergic nuclei related to tonic activation processes in thalamocortical systems. Journal of Neuroscience, 10(8), 25412559. https://doi.org/10.1523/jneurosci.10-08-02541.1990Google Scholar
Steriade, M., Nunez, A., & Amzica, F. (1993a). A novel slow (< 1 Hz) oscillation of neocortical neurons in vivo: depolarizing and hyperpolarizing components. Journal of Neuroscience, 13(8), 32523265. https://doi.org/10.1523/jneurosci.13-08-03252.1993Google Scholar
Steriade, M., Nunez, A., & Amzica, F. (1993b). Intracellular analysis of relations between the slow (< Hz) neocortical oscillation and other sleep rhythms of the electroencephalogram. Journal of Neuroscience, 13(8), 32663283. https://doi.org/10.1016/S0040-4039(97)01107–6Google Scholar
Steriade, M., Timofeev, I., & Grenier, F. (2001). Natural waking and sleep states: a view from inside neocortical neurons. Journal of Neurophysiology, 85(5), 19691985. https://doi.org/10.1152/jn.2001.85.5.1969Google Scholar
Stucynski, J. A., Schott, A. L., Baik, J., Chung, S., & Weber, F. (2021). Regulation of REM sleep by inhibitory neurons in the dorsomedial medulla. Current Biology, 32, 114. https://doi.org/10.1016/j.cub.2021.10.030Google Scholar
Sun, Y. G., Pita-Almenar, J. D., Wu, C. S., Renger, J. J., Uebele, V. N., Lu, H. C., & Beierlein, M. (2013). Biphasic cholinergic synaptic transmission controls action potential activity in thalamic reticular nucleus neurons. Journal of Neuroscience, 33(5), 20482059. https://doi.org/10.1523/JNEUROSCI.3177-12.2013Google Scholar
Takahashi, T. M., Sunagawa, G. A., Soya, S., Abe, M., Sakurai, K., Ishikawa, K., Yanagisawa, M., Hama, H., Hasegawa, E., Miyawaki, A., Sakimura, K., Takahashi, M., & Sakurai, T. (2020). A discrete neuronal circuit induces a hibernation-like state in rodents. Nature, 583(7814), 109114. https://doi.org/10.1038/s41586-020-2163-6Google Scholar
Takata, Y., Oishi, Y., Zhou, X. Z., Hasegawa, E., Takahashi, K., Cherasse, Y., Sakurai, T., & Lazarus, M. (2018). Sleep and wakefulness are controlled by ventral medial midbrain/pons GABAergic neurons in mice. Journal of Neuroscience, 38(47), 1008010092. https://doi.org/10.1523/JNEUROSCI.0598-18.2018Google Scholar
Timofeev, I., & Chauvette, S. (2017). Sleep slow oscillation and plasticity. Current Opinion in Neurobiology, 44, 116126. https://doi.org/10.1016/j.conb.2017.03.019Google Scholar
Timofeev, I., Grenier, F., Bazhenov, M., Sejnowski, T. J., & Steriade, M. (2000). Origin of slow cortical oscillations in deafferented cortical slabs. Cerebral Cortex, 10(12), 11851199. https://doi.org/10.1093/cercor/10.12.1185Google Scholar
Timofeev, I., & Steriade, M. (1996). Low-frequency rhythms in the thalamus of intact-cortex and decorticated cats. Journal of Neurophysiology, 76(6), 41524168. https://doi.org/10.1152/jn.1996.76.6.4152Google Scholar
Tononi, G., Boly, M., Massimini, M., & Koch, C. (2016). Integrated information theory: From consciousness to its physical substrate. Nature Reviews Neuroscience, 17(7), 450461. https://doi.org/10.1038/nrn.2016.44Google Scholar
Tononi, G., & Cirelli, C. (2014). Sleep and the price of plasticity: from synaptic and cellular homeostasis to memory consolidation and integration. Neuron, 81(1), 1234. https://doi.org/10.1016/j.neuron.2013.12.025Google Scholar
Totah, N. K. B., Logothetis, N. K., & Eschenko, O. (2019). Noradrenergic ensemble-based modulation of cognition over multiple timescales. Brain Research, 1709(December), 5066. https://doi.org/10.1016/j.brainres.2018.12.031Google Scholar
Tsai, C. J., Nagata, T., Liu, C. Y., Suganuma, T., Kanda, T., Miyazaki, T., Liu, K., Saitoh, T., Nagase, H., Lazarus, M., Vogt, K. E., Yanagisawa, M., & Hayashi, Y. (2021). Cerebral capillary blood flow upsurge during REM sleep is mediated by A2a receptors. Cell Reports, 36(7), 109558. https://doi.org/10.1016/j.celrep.2021.109558Google Scholar
Vantomme, G., Osorio-Forero, A., Lüthi, A., & Fernandez, L. M. J. (2019). Regulation of local sleep by the thalamic reticular nucleus. Frontiers in Neuroscience, 13(June), 18. https://doi.org/10.3389/fnins.2019.00576Google Scholar
Varela, C. (2013). The gating of neocortical information by modulators. Journal of Neurophysiology, 109(5), 12291232. https://doi.org/10.1152/jn.00701.2012Google Scholar
Varela, C. (2014). Thalamic neuromodulation and its implications for executive networks. Neural Circuits, 8(June), 122. https://doi.org/10.3389/fncir.2014.00069Google Scholar
Varela, C., & Sherman, S. M. (2007). Differences in response to muscarinic activation between first and higher order thalamic relays. Journal of Neurophysiology, 98(6), 35383547. https://doi.org/10.1152/jn.00578.2007Google Scholar
Varela, C., & Sherman, S. M. (2009). Differences in response to serotonergic activation between first and higher order thalamic nuclei. Cerebral Cortex, 19(8), 17761786. https://doi.org/10.1093/cercor/bhn208Google Scholar
Verret, L., Goutagny, R., Fort, P., Cagnon, L., Salvert, D., Léger, L., Boissard, R., Salin, P., Peyron, C., & Luppi, P. H. (2003). A role of melanin-concentrating hormone producing neurons in the central regulation of paradoxical sleep. BMC Neuroscience, 4, 110. https://doi.org/10.1186/1471-2202-4-19Google Scholar
Vertes, R. P., Linley, S. B., & Hoover, W. B. (2015). Limbic circuitry of the midline thalamus. Neuroscience and Biobehavioral Reviews, 54, 89107. https://doi.org/10.1016/j.neubiorev.2015.01.014Google Scholar
Vyazovskiy, V. V., Olcese, U., Hanlon, E. C., Nir, Y., Cirelli, C., & Tononi, G. (2011). Local sleep in awake rats. Nature, 472(7344), 443447. https://doi.org/10.1038/nature10009Google Scholar
Vyazovskiy, V. V., Olcese, U., Lazimy, Y. M., Faraguna, U., Esser, S. K., Williams, J. C., Cirelli, C., & Tononi, G. (2009). Cortical firing and sleep homeostasis. Neuron, 63(6), 865878. https://doi.org/10.1016/j.neuron.2009.08.024Google Scholar
Wamsley, E. J., Tucker, M. A., Shinn, A. K., Ono, K. E., McKinley, S. K., Ely, A. V., Goff, D. C., Stickgold, R., & Manoach, D. S. (2012). Reduced sleep spindles and spindle coherence in schizophrenia: mechanisms of impaired memory consolidation? Biological Psychiatry, 71(2), 154161. https://doi.org/10.1016/j.biopsych.2011.08.008Google Scholar
Warby, S. C., Wendt, S. L., Welinder, P., Munk, E. G. S., Carrillo, O., Sorensen, H. B. D., Jennum, P., Peppard, P. E., Perona, P., & Mignot, E. (2014). Sleep-spindle detection: crowdsourcing and evaluating performance of experts, non-experts and automated methods. Nature Methods, 11(4), 385392. https://doi.org/10.1038/nmeth.2855Google Scholar
Watson, B. O., Levenstein, D., Greene, J. P., Gelinas, J. N., & Buzsáki, G. (2016). Network homeostasis and state dynamics of neocortical sleep. Neuron, 90(4), 839852. https://doi.org/10.1016/j.neuron.2016.03.036Google Scholar
Watts, A., Gritton, H. J., Sweigart, J., & Poe, G. R. (2012). Antidepressant suppression of non-REM sleep spindles and REM sleep impairs hippocampus-dependent learning while augmenting striatum-dependent learning. Journal of Neuroscience, 32(39), 1341113420. https://doi.org/10.1523/JNEUROSCI.0170-12.2012Google Scholar
Weber, F., & Dan, Y. (2016). Circuit-based interrogation of sleep control. Nature, 538(7623), 5159. https://doi.org/10.1038/nature19773Google Scholar
Wells, M. F., Wimmer, R. D., Schmitt, L. I., Feng, G., & Halassa, M. M. (2016). Thalamic reticular impairment underlies attention deficit in Ptchd1 Y’mice. Nature, 532(7597), 5863. https://doi.org/10.1038/nature17427Google Scholar
Weyand, T. G., Boudreaux, M., Guido, W., Theodore, G., Boudreaux, M., & Guido, W. (2000). Burst and tonic response modes in thalamic neurons during sleep and wakefulness. Journal of Physiology, 85(3), 11071118.Google Scholar
White, N. S., & Alkire, M. T. (2003). Impaired thalamocortical connectivity in humans during general-anesthetic- induced unconsciousness. NeuroImage, 19(2), 402411. https://doi.org/10.1016/S1053-8119(03)00103-4Google Scholar
Wu, W., Cui, L., Fu, Y., Tian, Q., Liu, L., Zhang, X., Du, N., Chen, Y., Qiu, Z., Song, Y., Shi, F. D., & Xue, R. (2016). Sleep and cognitive abnormalities in acute minor thalamic infarction. Neuroscience Bulletin, 32(4), 341348. https://doi.org/10.1007/s12264-016-0036-7Google Scholar
Xu, M., Chung, S., Zhang, S., Zhong, P., Ma, C., Chang, W. C., Weissbourd, B., Sakai, N., Luo, L., Nishino, S., & Dan, Y. (2015). Basal forebrain circuit for sleep-wake control. Nature Neuroscience, 18(11), 16411647. https://doi.org/10.1038/nn.4143Google Scholar
Yamazaki, R., Toda, H., Libourel, P. A., Hayashi, Y., Vogt, K. E., & Sakurai, T. (2020). Evolutionary origin of distinct NREM and REM sleep. Frontiers in Psychology, 11(December), 18. https://doi.org/10.3389/fpsyg.2020.567618Google Scholar
Yang, S. R., Hu, Z. Z., Luo, Y. J., Zhao, Y. N., Sun, H. X., Yin, D., Wang, C. Y., Yan, Y. D., Wang, D. R., Yuan, X. S., Ye, C. B., Guo, W., Qu, W. M., Cherasse, Y., Lazarus, M., Ding, Y. Q., & Huang, Z. L. (2018). The rostromedial tegmental nucleus is essential for non-rapid eye movement sleep. PLoS Biology, 16(4), 129. https://doi.org/10.1371/journal.pbio.2002909Google Scholar
Yuan, X.-S., Wang, L., Dong, H., Qu, W.-M., Yang, S.-R., Cherasse, Y., Lazarus, M., Schiffmann, S. N., d’Exaerde, A. de K., Li, R.-X., & Huang, Z.-L. (2017). Striatal adenosine A2A receptor neurons control active-period sleep via parvalbumin neurons in external globus pallidus. eLife, 6, 124. https://doi.org/10.7554/elife.29055Google Scholar
Zhang, Ze, Liu, W. Y., Diao, Y. P., Xu, W., Zhong, Y. H., Zhang, J. Y., Lazarus, M., Liu, Y. Y., Qu, W. M., & Huang, Z. L. (2019). Superior colliculus GABAergic neurons are essential for acute dark induction of wakefulness in mice. Current Biology, 29(4), 637–644.e3. https://doi.org/10.1016/j.cub.2018.12.031Google Scholar
Zhang, Zhe, Ferretti, V., Güntan, I., Moro, A., Steinberg, E. A., Ye, Z., Zecharia, A. Y., Yu, X., Vyssotski, A. L., Brickley, S. G., Yustos, R., Pillidge, Z. E., Harding, E. C., Wisden, W., & Franks, N. P. (2015). Neuronal ensembles sufficient for recovery sleep and the sedative actions of α 2 adrenergic agonists. Nature Neuroscience, 18(4), 553561. https://doi.org/10.1038/nn.3957Google Scholar
Zhang, Zhe, Zhong, P., Hu, F., Barger, Z., Ren, Y., Ding, X., Li, S., Weber, F., Chung, S., Palmiter, R. D., & Dan, Y. (2019). An excitatory circuit in the perioculomotor midbrain for non-REM sleep control. Cell, 177(5), 1293–1307.e16. https://doi.org/10.1016/j.cell.2019.03.041Google Scholar
Zhong, P., Zhang, Z., Barger, Z., Ma, C., Liu, D., Ding, X., & Dan, Y. (2019). Control of non-REM sleep by midbrain neurotensinergic neurons. Neuron, 104(4), 795–809.e6. https://doi.org/10.1016/j.neuron.2019.08.026Google Scholar
Zhu, J., & Heggelund, P. (2001). Muscarinic regulation of dendritic and axonal outputs of rat thalamic interneurons: a new cellular mechanism for uncoupling distal dendrites. Journal of Neuroscience, 21(4), 11481159. https://doi.org/10.1523/jneurosci.21–04-01148.2001Google Scholar
Zhu, L., Blethyn, K. L., Cope, D. W., Tsomaia, V., Crunelli, V., & Hughes, S. W. (2006). Nucleus- and species-specific properties of the slow (< 1 Hz) sleep oscillation in thalamocortical neurons. Neuroscience, 141, 621636. https://doi.org/10.1016/j.neuroscience.2006.04.069Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×