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.

General anaesthesia: from molecular targets to neuronal pathways of sleep and arousal

Key Points

  • It is now accepted that general anaesthetics act at specific molecular targets, and a relatively small number of likely ion channels and receptors have been identified — the γ-aminobutyric acid type A receptor, two-pore-domain K+ channels and N-methyl-D-aspartate receptors top the list. Anaesthetics probably affect these channels by selectively binding to specific conformational states.

  • Many anaesthetic determinants on these targets have been investigated, partly in the hope that they could be introduced as 'silent' mutations in animals to test the in vivo significance of the putative target; this approach is beginning to yield exciting results.

  • Focus in this field of research is gradually shifting to the question of how effects on these targets might translate into effects in the intact animal, and one possibility that has emerged is that anaesthetic-induced loss of consciousness and deep sleep might have common neuronal mechanisms.

  • Support for this idea comes from human imaging studies, which have revealed striking similarities between the two states. This similarity is due, in large part, to the thalamic deactivation that occurs during deep sleep and during anaesthetic-induced loss of consciousness.

  • Further evidence comes from in vitro studies on the neuronal pathways that underlie behavioural arousal and those that are responsible for the initiation and maintenance of sleep.

  • Current evidence suggests that anaesthetics might induce reversible loss of consciousness by acting on arousal and sleep pathways that involve both the thalamus and the hypothalamus.

Abstract

The mechanisms through which general anaesthetics, an extremely diverse group of drugs, cause reversible loss of consciousness have been a long-standing mystery. Gradually, a relatively small number of important molecular targets have emerged, and how these drugs act at the molecular level is becoming clearer. Finding the link between these molecular studies and anaesthetic-induced loss of consciousness presents an enormous challenge, but comparisons with the features of natural sleep are helping us to understand how these drugs work and the neuronal pathways that they affect. Recent work suggests that the thalamus and the neuronal networks that regulate its activity are the key to understanding how anaesthetics cause loss of consciousness.

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

Figure 1: Loss of consciousness in humans occurs over a very narrow range of anaesthetic concentrations and correlates with loss of the righting reflex in rodents.
Figure 2: Three key molecular targets and some known determinants of their anaesthetic sensitivities.
Figure 3: Functional brain imaging reveals similarities between anaesthetic-induced loss of consciousness and deep natural sleep.
Figure 4: Thalamic oscillations.

References

  1. Franks, N. P. & Lieb, W. R. Molecular and cellular mechanisms of general anaesthesia. Nature 367, 607–614 (1994).

    CAS  PubMed  Google Scholar 

  2. Campagna, J. A., Miller, K. W. & Forman, S. A. Mechanisms of actions of inhaled anesthetics. N. Engl. J. Med. 348, 2110–2124 (2003).

    CAS  PubMed  Google Scholar 

  3. Sonner, J. M. et al. Inhaled anesthetics and immobility: mechanisms, mysteries, and minimum alveolar anesthetic concentration. Anesth. Analg. 97, 718–740 (2003).

    CAS  PubMed  Google Scholar 

  4. Rudolph, U. & Antkowiak, B. Molecular and neuronal substrates for general anaesthetics. Nature Rev. Neurosci. 5, 709–720 (2004).

    CAS  Google Scholar 

  5. Macdonald, R. L. & Barker, J. L. Different actions of anticonvulsant and anesthetic barbiturates revealed by use of cultured mammalian neurons. Science 200, 775–777 (1978).

    CAS  PubMed  Google Scholar 

  6. Nicoll, R. A. Pentobarbital: differential postsynaptic actions on sympathetic ganglion cells. Science 199, 451–452 (1978).

    CAS  PubMed  Google Scholar 

  7. Leeb-Lundberg, F., Snowman, A. & Olsen, R. W. Barbiturate receptor sites are coupled to benzodiazepine receptors. Proc. Natl Acad. Sci. USA 77, 7468–7472 (1980).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Farrant, M. & Nusser, Z. Variations on an inhibitory theme: phasic and tonic activation of GABAA receptors. Nature Rev. Neurosci. 6, 215–229 (2005).

    CAS  Google Scholar 

  9. de Sousa, S. L., Dickinson, R., Lieb, W. R. & Franks, N. P. Contrasting synaptic actions of the inhalational general anesthetics isoflurane and xenon. Anesthesiology 92, 1055–1066 (2000).

    CAS  PubMed  Google Scholar 

  10. Raines, D. E., Claycomb, R. J., Scheller, M. & Forman, S. A. Nonhalogenated alkane anesthetics fail to potentiate agonist actions on two ligand-gated ion channels. Anesthesiology 95, 470–477 (2001).

    CAS  PubMed  Google Scholar 

  11. Hill-Venning, C., Belelli, D., Peters, J. A. & Lambert, J. J. Subunit-dependent interaction of the general anaesthetic etomidate with the GABAA receptor. Br. J. Pharmacol. 120, 749–756 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Bai, D. et al. Distinct functional and pharmacological properties of tonic and quantal inhibitory postsynaptic currents mediated by γ-aminobutyric acidA receptors in hippocampal neurons. Mol. Pharmacol. 59, 814–824 (2001). This paper provided a comprehensive demonstration of the distinct pharmacological properties of extrasynaptic GABA A receptors and, in particular, their sensitivity to propofol.

    CAS  PubMed  Google Scholar 

  13. Bieda, M. C. & MacIver, M. B. Major role for tonic GABAA conductances in anesthetic suppression of intrinsic neuronal excitability. J. Neurophysiol. 92, 1658–1667 (2004).

    CAS  PubMed  Google Scholar 

  14. Caraiscos, V. B. et al. Selective enhancement of tonic GABAergic inhibition in murine hippocampal neurons by low concentrations of the volatile anesthetic isoflurane. J. Neurosci. 24, 8454–8458 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Belelli, D., Peden, D. R., Rosahl, T. W., Wafford, K. A. & Lambert, J. J. Extrasynaptic GABAA receptors of thalamocortical neurons: a molecular target for hypnotics. J. Neurosci. 25, 11513–11520 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Belelli, D., Lambert, J. J., Peters, J. A., Wafford, K. & Whiting, P. J. The interaction of the general anesthetic etomidate with the GABAA receptor is influenced by a single amino acid. Proc. Natl Acad. Sci. USA 94, 11031–11036 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Mihic, S. J. et al. Sites of alcohol and volatile anaesthetic action on GABAA and glycine receptors. Nature 389, 385–389 (1997). This important paper, together with reference 16, showed that individual mutations in GABA A receptors could remove their in vitro anaesthetic sensitivity.

    CAS  PubMed  Google Scholar 

  18. Jurd, R. et al. General anesthetic actions in vivo strongly attenuated by a point mutation in the GABAA receptor β3 subunit. FASEB J. 17, 250–252 (2003). This landmark paper, together with reference 19, showed that point mutations in GABA A receptors could greatly affect their in vivo anaesthetic potencies.

    CAS  PubMed  Google Scholar 

  19. Reynolds, D. S. et al. Sedation and anesthesia mediated by distinct GABAA receptor isoforms. J. Neurosci. 23, 8608–8617 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Sonner, J. M. et al. Effect of isoflurane and other potent inhaled anesthetics on minimum alveolar concentration, learning, and the righting reflex in mice engineered to express α1 γ-aminobutyric acid type A receptors unresponsive to isoflurane. Anesthesiology 106, 107–113 (2007).

    CAS  PubMed  Google Scholar 

  21. Krasowski, M. D. & Harrison, N. L. The actions of ether, alcohol and alkane general anaesthetics on GABAA and glycine receptors and the effects of TM2 and TM3 mutations. Br. J. Pharmacol. 129, 731–743 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Nishikawa, K., Jenkins, A., Paraskevakis, I. & Harrison, N. L. Volatile anesthetic actions on the GABAA receptors: contrasting effects of α1(S270) and β2(N265) point mutations. Neuropharmacology 42, 337–345 (2002).

    CAS  PubMed  Google Scholar 

  23. Mascia, M. P., Trudell, J. R. & Harris, R. A. Specific binding sites for alcohols and anesthetics on ligand-gated ion channels. Proc. Natl Acad. Sci. USA 97, 9305–9310 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Bali, M. & Akabas, M. H. Defining the propofol binding site location on the GABAA receptor. Mol. Pharmacol. 65, 68–76 (2004).

    CAS  PubMed  Google Scholar 

  25. Eckenhoff, R. G., Knoll, F. J., Greenblatt, E. P. & Dailey, W. P. Halogenated diazirines as photolabel mimics of the inhaled haloalkane anesthetics. J. Med. Chem. 45, 1879–1886 (2002).

    CAS  PubMed  Google Scholar 

  26. Darbandi-Tonkabon, R. et al. Photoaffinity labeling with a neuroactive steroid analogue. 6-azi-pregnanolone labels voltage-dependent anion channel-1 in rat brain. J. Biol. Chem. 278, 13196–13206 (2003).

    CAS  PubMed  Google Scholar 

  27. Husain, S. S. et al. 2-(3-Methyl-3H-diaziren-3-yl)ethyl 1-(1-phenylethyl)-1H-imidazole-5-carboxylate: a derivative of the stereoselective general anesthetic etomidate for photolabeling ligand-gated ion channels. J. Med. Chem. 46, 1257–1265 (2003).

    CAS  PubMed  Google Scholar 

  28. Li, G. D. et al. Identification of a GABAA receptor anesthetic binding site at subunit interfaces by photolabeling with an etomidate analog. J. Neurosci. 26, 11599–11605 (2006). This impressive piece of work identified a binding site for etomidate on the GABA A receptor using photoaffinity labelling.

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Krasowski, M. D. et al. Propofol and other intravenous anesthetics have sites of action on the γ-aminobutyric acid type A receptor distinct from that for isoflurane. Mol. Pharmacol. 53, 530–538 (1998).

    CAS  PubMed  Google Scholar 

  30. Koltchine, V. V. et al. Agonist gating and isoflurane potentiation in the human γ-aminobutyric acid type A receptor determined by the volume of a second transmembrane domain residue. Mol. Pharmacol. 56, 1087–1093 (1999).

    CAS  PubMed  Google Scholar 

  31. Jenkins, A. et al. Evidence for a common binding cavity for three general anesthetics within the GABAA receptor. J. Neurosci. 21, RC136 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Jenkins, A., Andreasen, A., Trudell, J. R. & Harrison, N. L. Tryptophan scanning mutagenesis in TM4 of the GABAA receptor α1 subunit: implications for modulation by inhaled anesthetics and ion channel structure. Neuropharmacology 43, 669–678 (2002).

    CAS  PubMed  Google Scholar 

  33. Schofield, C. M. & Harrison, N. L. Transmembrane residues define the action of isoflurane at the GABAA receptor alpha-3 subunit. Brain Res. 1032, 30–35 (2005).

    CAS  PubMed  Google Scholar 

  34. Carlson, B. X., Engblom, A. C., Kristiansen, U., Schousboe, A. & Olsen, R. W. A single glycine residue at the entrance to the first membrane-spanning domain of the γ-aminobutyric acid type A receptor β2 subunit affects allosteric sensitivity to GABA and anesthetics. Mol. Pharmacol. 57, 474–484 (2000).

    CAS  PubMed  Google Scholar 

  35. Krasowski, M. D., Nishikawa, K., Nikolaeva, N., Lin, A. & Harrison, N. L. Methionine 286 in transmembrane domain 3 of the GABAA receptor β subunit controls a binding cavity for propofol and other alkylphenol general anesthetics. Neuropharmacology 41, 952–964 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Siegwart, R., Jurd, R. & Rudolph, U. Molecular determinants for the action of general anesthetics at recombinant α2β3γ2 γ-aminobutyric acidA receptors. J. Neurochem. 80, 140–148 (2002).

    CAS  PubMed  Google Scholar 

  37. Chang, C. S., Olcese, R. & Olsen, R. W. A single M1 residue in the β2 subunit alters channel gating of GABAA receptor in anesthetic modulation and direct activation. J. Biol. Chem. 278, 42821–42828 (2003).

    CAS  PubMed  Google Scholar 

  38. Zeller, A., Arras, M., Jurd, R. & Rudolph, U. Identification of a molecular target mediating the general anesthetic actions of pentobarbital. Mol. Pharmacol. 71, 852–859 (2007).

    CAS  PubMed  Google Scholar 

  39. Zeller, A., Arras, M., Lazaris, A., Jurd, R. & Rudolph, U. Distinct molecular targets for the central respiratory and cardiac actions of the general anesthetics etomidate and propofol. FASEB J. 19, 1677–1679 (2005).

    CAS  PubMed  Google Scholar 

  40. Cirone, J. et al. γ-aminobutyric acid type A receptor β2 subunit mediates the hypothermic effect of etomidate in mice. Anesthesiology 100, 1438–1445 (2004).

    CAS  PubMed  Google Scholar 

  41. Chandra, D. et al. GABAA receptor α4 subunits mediate extrasynaptic inhibition in thalamus and dentate gyrus and the action of gaboxadol. Proc. Natl Acad. Sci. USA 103, 15230–15235 (2006). This paper provided a definitive demonstration that the key pharmacological effects of the sedative agent gaboxadol are mediated by α4-subunit-containing GABA A receptors, which are thought to be predominantly expressed extrajunctionally.

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Cheng, V. Y. et al. α5GABAA receptors mediate the amnestic but not sedative-hypnotic effects of the general anesthetic etomidate. J. Neurosci. 26, 3713–3720 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Hall, A. C., Lieb, W. R. & Franks, N. P. Stereoselective and non-stereoselective actions of isoflurane on the GABAA receptor. Br. J. Pharmacol. 112, 906–910 (1994).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Tomlin, S. L., Jenkins, A., Lieb, W. R. & Franks, N. P. Stereoselective effects of etomidate optical isomers on GABAA receptors and animals. Anesthesiology 88, 708–717 (1998).

    CAS  PubMed  Google Scholar 

  45. Tomlin, S. L., Jenkins, A., Lieb, W. R. & Franks, N. P. Preparation of barbiturate optical isomers and their effects on GABAA receptors. Anesthesiology 90, 1714–1722 (1999).

    CAS  PubMed  Google Scholar 

  46. Dickinson, R., White, I., Lieb, W. R. & Franks, N. P. Stereoselective loss of righting reflex in rats by isoflurane. Anesthesiology 93, 837–843 (2000).

    CAS  PubMed  Google Scholar 

  47. Nicoll, R. A. & Madison, D. V. General anesthetics hyperpolarize neurons in the vertebrate central nervous system. Science 217, 1055–1057 (1982).

    CAS  PubMed  Google Scholar 

  48. Franks, N. P. & Lieb, W. R. Volatile general anaesthetics activate a novel neuronal K+ current. Nature 333, 662–664 (1988). This paper provided the first characterization of an anaesthetic-activated K+ current. This current was later shown (see reference 54) to be mediated by a 2PK channel.

    CAS  PubMed  Google Scholar 

  49. Goldstein, S. A., Bockenhauer, D., O'Kelly, I. & Zilberberg, N. Potassium leak channels and the KCNK family of two-P-domain subunits. Nature Rev. Neurosci. 2, 175–184 (2001).

    CAS  Google Scholar 

  50. Patel, A. J. & Honore, E. Properties and modulation of mammalian 2P domain K+ channels. Trends Neurosci. 24, 339–346 (2001).

    CAS  PubMed  Google Scholar 

  51. Patel, A. J. et al. Inhalational anesthetics activate two-pore-domain background K+ channels. Nature Neurosci. 2, 422–426 (1999). This important paper provided the first demonstration that mammalian baseline K+ channels can be activated by various volatile general anaesthetics, establishing 2PK channels as likely targets for general anaesthetics.

    CAS  PubMed  Google Scholar 

  52. Liu, C., Au, J. D., Zou, H. L., Cotten, J. F. & Yost, C. S. Potent activation of the human tandem pore domain K channel TRESK with clinical concentrations of volatile anesthetics. Anesth. Analg. 99, 1715–1722 (2004).

    CAS  PubMed  Google Scholar 

  53. Berg, A. P., Talley, E. M., Manger, J. P. & Bayliss, D. A. Motoneurons express heteromeric TWIK-related acid-sensitive K+ (TASK) channels containing TASK-1 (KCNK3) and TASK-3 (KCNK9) subunits. J. Neurosci. 24, 6693–6702 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Andres-Enguix, I. et al. Determinants of the anesthetic sensitivity of two-pore domain acid-sensitive potassium channels: molecular cloning of an anesthetic-activated potassium channel from Lymnaea stagnalis. J. Biol. Chem. 282, 20977–20990 (2007).

    CAS  PubMed  Google Scholar 

  55. Gruss, M. et al. Two-pore-domain K+ channels are a novel target for the anesthetic gases xenon, nitrous oxide, and cyclopropane. Mol. Pharmacol. 65, 443–452 (2004).

    CAS  PubMed  Google Scholar 

  56. Ries, C. R. & Puil, E. Mechanism of anesthesia revealed by shunting actions of isoflurane on thalamocortical neurons. J. Neurophysiol. 81, 1795–1801 (1999). This study showed that isoflurane could affect the firing properties of thalamocortical neurons by enhancing their baseline K+ conductance.

    CAS  PubMed  Google Scholar 

  57. Sirois, J. E., Lei, Q., Talley, E. M., Lynch, C. & Bayliss, D. A. The TASK-1 two-pore domain K+ channel is a molecular substrate for neuronal effects of inhalation anesthetics. J. Neurosci. 20, 6347–6354 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Washburn, C. P., Sirois, J. E., Talley, E. M., Guyenet, P. G. & Bayliss, D. A. Serotonergic raphe neurons express TASK channel transcripts and a TASK-like pH- and halothane-sensitive K+ conductance. J. Neurosci. 22, 1256–1265 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Westphalen, R. I., Krivitski, M., Amarosa, A., Guy, N. & Hemmings, H. C. Jr. Reduced inhibition of cortical glutamate and GABA release by halothane in mice lacking the K+ channel, TREK-1. Br. J. Pharmacol. 152, 939–945 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Talley, E. M. & Bayliss, D. A. Modulation of TASK-1 (Kcnk3) and TASK-3 (Kcnk9) potassium channels: volatile anesthetics and neurotransmitters share a molecular site of action. J. Biol. Chem. 277, 17733–17742 (2002).

    CAS  PubMed  Google Scholar 

  61. Honoré, E., Maingret, F., Lazdunski, M. & Patel, A. J. An intracellular proton sensor commands lipid- and mechano-gating of the K+ channel TREK-1. EMBO J. 21, 2968–2976 (2002).

    PubMed  PubMed Central  Google Scholar 

  62. Franks, N. P. & Lieb, W. R. Stereospecific effects of inhalational general anesthetic optical isomers on nerve ion channels. Science 254, 427–430 (1991).

    CAS  PubMed  Google Scholar 

  63. Heurteaux, C. et al. TREK-1, a K+ channel involved in neuroprotection and general anesthesia. EMBO J. 23, 2684–2695 (2004). This paper provided an important demonstration that TREK1-knockout mice had reduced sensitivities to volatile anaesthetics but unchanged sensitivities to a barbiturate.

    CAS  PubMed  PubMed Central  Google Scholar 

  64. Linden, A. M. et al. The in vivo contributions of TASK-1-containing channels to the actions of inhalation anesthetics, the a2 adrenergic sedative dexmedetomidine, and cannabinoid agonists. J. Pharmacol. Exp. Ther. 317, 615–626 (2006).

    CAS  PubMed  Google Scholar 

  65. Cull-Candy, S., Brickley, S. & Farrant, M. NMDA receptor subunits: diversity, development and disease. Curr. Opin. Neurobiol. 11, 327–335 (2001).

    CAS  PubMed  Google Scholar 

  66. Perouansky, M., Baranov, D., Salman, M. & Yaari, Y. Effects of halothane on glutamate receptor-mediated excitatory postsynaptic currents. A patch-clamp study in adult mouse hippocampal slices. Anesthesiology 83, 109–119 (1995).

    CAS  PubMed  Google Scholar 

  67. Plested, A. J., Wildman, S. S., Lieb, W. R. & Franks, N. P. Determinants of the sensitivity of AMPA receptors to xenon. Anesthesiology 100, 347–358 (2004).

    CAS  PubMed  Google Scholar 

  68. Flohr, H., Glade, U. & Motzko, D. The role of the NMDA synapse in general anesthesia. Toxicol. Lett. 100–101, 23–29 (1998).

    PubMed  Google Scholar 

  69. Yamakura, T. & Harris, R. A. Effects of gaseous anesthetics nitrous oxide and xenon on ligand-gated ion channels. Comparison with isoflurane and ethanol. Anesthesiology 93, 1095–1101 (2000).

    CAS  PubMed  Google Scholar 

  70. Hollmann, M. W., Liu, H. T., Hoenemann, C. W., Liu, W. H. & Durieux, M. E. Modulation of NMDA receptor function by ketamine and magnesium. Part II: interactions with volatile anesthetics. Anesth. Analg. 92, 1182–1191 (2001).

    CAS  PubMed  Google Scholar 

  71. Ogata, J. et al. Effects of anesthetics on mutant N-methyl-D-aspartate receptors expressed in Xenopus oocytes. J. Pharmacol. Exp. Ther. 318, 434–443 (2006).

    CAS  PubMed  Google Scholar 

  72. Solt, K., Eger, E. I. & Raines, D. E. Differential modulation of human N-methyl-D-aspartate receptors by structurally diverse general anesthetics. Anesth. Analg. 102, 1407–1411 (2006).

    CAS  PubMed  Google Scholar 

  73. Franks, N. P., Dickinson, R., de Sousa, S. L., Hall, A. C. & Lieb, W. R. How does xenon produce anaesthesia? Nature 396, 324 (1998).

    CAS  PubMed  Google Scholar 

  74. Jevtovic-Todorovic, V. et al. Nitrous oxide (laughing gas) is an NMDA antagonist, neuroprotectant and neurotoxin. Nature Med. 4, 460–463 (1998). This was the first study to identify the NMDA receptor as a plausible target for nitrous oxide. The work was complemented by reference 73, which proposed that the same target was important for xenon.

    CAS  PubMed  Google Scholar 

  75. Watanabe, I., Takenoshita, M., Sawada, T., Uchida, I. & Mashimo, T. Xenon suppresses nociceptive reflex in newborn rat spinal cord in vitro; comparison with nitrous oxide. Eur. J. Pharmacol. 496, 71–76 (2004).

    CAS  PubMed  Google Scholar 

  76. Wilhelm, S., Ma, D., Maze, M. & Franks, N. P. Effects of xenon on in vitro and in vivo models of neuronal injury. Anesthesiology 96, 1485–1491 (2002).

    CAS  PubMed  Google Scholar 

  77. Dickinson, R. et al. Competitive inhibition at the glycine site of the N-methyl-D-aspartate receptor by the anesthetics xenon and isoflurane: evidence from molecular modeling and electrophysiology. Anesthesiology 107, 756–767 (2007).

    CAS  PubMed  Google Scholar 

  78. Carter, A. J. Antagonists of the NMDA receptor-channel complex and motor coordination. Life Sci. 57, 917–929 (1995).

    CAS  PubMed  Google Scholar 

  79. Lingenhohl, K. & Pozza, M. F. Reevaluation of ACEA 1021 as an antagonist at the strychnine-insensitive glycine site of the N-methyl-D-aspartate receptor. Neuropharmacology 37, 729–737 (1998).

    CAS  PubMed  Google Scholar 

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

    PubMed  Google Scholar 

  81. Harrison, N. L., Kugler, J. L., Jones, M. V., Greenblatt, E. P. & Pritchett, D. B. Positive modulation of human GABAA and glycine receptors by the inhalation anesthetic isoflurane. Mol. Pharmacol. 44, 628–632 (1993).

    CAS  PubMed  Google Scholar 

  82. Downie, D. L., Hall, A. C., Lieb, W. R. & Franks, N. P. Effects of inhalational general anaesthetics on native glycine receptors in rat medullary neurones and recombinant glycine receptors in Xenopus oocytes. Br. J. Pharmacol. 118, 493–502 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Mascia, M. P., Machu, T. K. & Harris, R. A. Enhancement of homomeric glycine receptor function by long-chain alcohols and anaesthetics. Br. J. Pharmacol. 119, 1331–1336 (1996).

    CAS  PubMed  PubMed Central  Google Scholar 

  84. Sirois, J. E., Lynch, C. & Bayliss, D. A. Convergent and reciprocal modulation of a leak K+ current and Ih by an inhalational anaesthetic and neurotransmitters in rat brainstem motoneurones. J. Physiol. 541, 717–729 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  85. Chen, X. et al. HCN subunit-specific and cAMP-modulated effects of anesthetics on neuronal pacemaker currents. J. Neurosci. 25, 5803–5814 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  86. Ying, S. W., Abbas, S. Y., Harrison, N. L. & Goldstein, P. A. Propofol block of Ih contributes to the suppression of neuronal excitability and rhythmic burst firing in thalamocortical neurons. Eur. J. Neurosci. 23, 465–480 (2006).

    PubMed  Google Scholar 

  87. Funahashi, M., Mitoh, Y. & Matsuo, R. The sensitivity of hyperpolarization-activated cation current (Ih) to propofol in rat area postrema neurons. Brain Res. 1015, 198–201 (2004).

    CAS  PubMed  Google Scholar 

  88. Ouyang, W., Wang, G. & Hemmings, H. C. Jr. Isoflurane and propofol inhibit voltage-gated sodium channels in isolated rat neurohypophysial nerve terminals. Mol. Pharmacol. 64, 373–381 (2003).

    CAS  PubMed  Google Scholar 

  89. Shiraishi, M. & Harris, R. A. Effects of alcohols and anesthetics on recombinant voltage-gated Na+ channels. J. Pharmacol. Exp. Ther. 309, 987–994 (2004).

    CAS  PubMed  Google Scholar 

  90. OuYang, W. & Hemmings, H. C. Jr. Isoform-selective effects of isoflurane on voltage-gated Na+ channels. Anesthesiology 107, 91–98 (2007).

    CAS  PubMed  Google Scholar 

  91. Fiset, P. et al. Brain mechanisms of propofol-induced loss of consciousness in humans: a positron emission tomographic study. J. Neurosci. 19, 5506–5513 (1999). This paper, together with reference 95, highlights the relatively selective deactivation of the thalamus and other specific brain regions that occurs at, or during, anaesthetic-induced loss of consciousness.

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Bonhomme, V. et al. Propofol anesthesia and cerebral blood flow changes elicited by vibrotactile stimulation: a positron emission tomography study. J. Neurophysiol. 85, 1299–1308 (2001).

    CAS  PubMed  Google Scholar 

  93. Kaisti, K. K. et al. Effects of sevoflurane, propofol, and adjunct nitrous oxide on regional cerebral blood flow, oxygen consumption, and blood volume in humans. Anesthesiology 99, 603–613 (2003). This was a comprehensive study into the effects of anaesthetics on regional cerebral blood flow in humans.

    CAS  PubMed  Google Scholar 

  94. Laitio, R. M. et al. Effects of xenon anesthesia on cerebral blood flow in humans: a positron emission tomography study. Anesthesiology 106, 1128–1133 (2007).

    PubMed  Google Scholar 

  95. Alkire, M. T., Haier, R. J. & Fallon, J. H. Toward a unified theory of narcosis: brain imaging evidence for a thalamocortical switch as the neurophysiologic basis of anesthetic-induced unconsciousness. Conscious. Cogn. 9, 370–386 (2000).

    CAS  PubMed  Google Scholar 

  96. White, N. S. & Alkire, M. T. Impaired thalamocortical connectivity in humans during general-anesthetic-induced unconsciousness. Neuroimage 19, 402–411 (2003).

    PubMed  Google Scholar 

  97. Braun, A. R. et al. Regional cerebral blood flow throughout the sleep-wake cycle. An H2(15)O PET study. Brain 120, 1173–1197 (1997). This paper provided an impressive and thorough analysis of the changes in cerebral blood flow that occur during the sleep–wake cycle.

    PubMed  Google Scholar 

  98. Kajimura, N. et al. Activity of midbrain reticular formation and neocortex during the progression of human non-rapid eye movement sleep. J. Neurosci. 19, 10065–10073 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  99. Maquet, P. Functional neuroimaging of normal human sleep by positron emission tomography. J. Sleep Res. 9, 207–231 (2000).

    CAS  PubMed  Google Scholar 

  100. Laureys, S., Owen, A. M. & Schiff, N. D. Brain function in coma, vegetative state, and related disorders. Lancet Neurol. 3, 537–546 (2004).

    PubMed  Google Scholar 

  101. Cavanna, A. E. & Trimble, M. R. The precuneus: a review of its functional anatomy and behavioural correlates. Brain 129, 564–583 (2006).

    PubMed  Google Scholar 

  102. Steriade, M., McCormick, D. A. & Sejnowski, T. J. Thalamocortical oscillations in the sleeping and aroused brain. Science 262, 679–685 (1993). This paper provided a beautifully clear description of how thalamocortical oscillations are generated during the initiation and maintenance of sleep.

    CAS  PubMed  Google Scholar 

  103. McCormick, D. A. & Bal, T. Sleep and arousal: thalamocortical mechanisms. Annu. Rev. Neurosci. 20, 185–215 (1997).

    CAS  PubMed  Google Scholar 

  104. Huguenard, J. R. & McCormick, D. A. Thalamic synchrony and dynamic regulation of global forebrain oscillations. Trends Neurosci. 30, 350–356 (2007).

    CAS  PubMed  Google Scholar 

  105. Angel, A. The G. L. Brown lecture. Adventures in anaesthesia. Exp. Physiol. 76, 1–38 (1991). This paper reviewed a series of comprehensive studies designed to investigate at the anatomical level how anaesthetics affect the transmission of sensory information from the periphery to the cortex. Although it did not address anaesthetic-induced LOC, the article drew particular attention to the role of the thalamus and thalamocortical connectivity in anaesthetic action.

    CAS  PubMed  Google Scholar 

  106. Bal, T., von Krosigk, M. & McCormick, D. A. Synaptic and membrane mechanisms underlying synchronized oscillations in the ferret lateral geniculate nucleus in vitro. J. Physiol. 483, 641–663 (1995).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Le Masson, G., Renaud-Le Masson, S., Debay, D. & Bal, T. Feedback inhibition controls spike transfer in hybrid thalamic circuits. Nature 417, 854–858 (2002).

    CAS  PubMed  Google Scholar 

  108. Hofle, N. et al. Regional cerebral blood flow changes as a function of delta and spindle activity during slow wave sleep in humans. J. Neurosci. 17, 4800–4808 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Clark, D. L. & Rosner, B. S. Neurophysiologic effects of general anesthetics. I. The electroencephalogram and sensory evoked responses in man. Anesthesiology 38, 564–582 (1973).

    CAS  PubMed  Google Scholar 

  110. Keifer, J. C., Baghdoyan, H. A. & Lydic, R. Pontine cholinergic mechanisms modulate the cortical electroencephalographic spindles of halothane anesthesia. Anesthesiology 84, 945–954 (1996). This paper provided evidence that the spindles in the EEG that are generated by halothane might be regulated by pontine cholinergic mechanisms.

    CAS  PubMed  Google Scholar 

  111. Sloan, T. B. Anesthetic effects on electrophysiologic recordings. J. Clin. Neurophysiol. 15, 217–226 (1998).

    CAS  PubMed  Google Scholar 

  112. Gugino, L. D. et al. Quantitative EEG changes associated with loss and return of consciousness in healthy adult volunteers anaesthetized with propofol or sevoflurane. Br. J. Anaesth. 87, 421–428 (2001).

    CAS  PubMed  Google Scholar 

  113. Voss, L. & Sleigh, J. Monitoring consciousness: the current status of EEG-based depth of anaesthesia monitors. Best Pract. Res. Clin. Anaesthesiol. 21, 313–325 (2007).

    PubMed  Google Scholar 

  114. Detsch, O., Vahle-Hinz, C., Kochs, E., Siemers, M. & Bromm, B. Isoflurane induces dose-dependent changes of thalamic somatosensory information transfer. Brain Res. 829, 77–89 (1999).

    CAS  PubMed  Google Scholar 

  115. Vahle-Hinz, C. & Detsch, O. What can in vivo electrophysiology in animal models tell us about mechanisms of anaesthesia? Br. J. Anaesth. 89, 123–142 (2002).

    CAS  PubMed  Google Scholar 

  116. Sugiyama, K., Muteki, T. & Shimoji, K. Halothane-induced hyperpolarization and depression of postsynaptic potentials of guinea pig thalamic neurons in vitro. Brain Res. 576, 97–103 (1992).

    CAS  PubMed  Google Scholar 

  117. Meuth, S. G. et al. Contribution of TWIK-related acid-sensitive K+ channel 1 (TASK1) and TASK3 channels to the control of activity modes in thalamocortical neurons. J. Neurosci. 23, 6460–6469 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  118. Jia, F. et al. An extrasynaptic GABAA receptor mediates tonic inhibition in thalamic VB neurons. J. Neurophysiol. 94, 4491–4501 (2005).

    CAS  PubMed  Google Scholar 

  119. Joksovic, P. M., Brimelow, B. C., Murbartian, J., Perez-Reyes, E. & Todorovic, S. M. Contrasting anesthetic sensitivities of T-type Ca2+ channels of reticular thalamic neurons and recombinant Cav3.3 channels. Br. J. Pharmacol. 144, 59–70 (2005).

    CAS  PubMed  Google Scholar 

  120. Cudeiro, J. & Sillito, A. M. Looking back: corticothalamic feedback and early visual processing. Trends Neurosci. 29, 298–306 (2006).

    CAS  PubMed  Google Scholar 

  121. Lukatch, H. S. & MacIver, M. B. Synaptic mechanisms of thiopental-induced alterations in synchronized cortical activity. Anesthesiology 84, 1425–1434 (1996).

    CAS  PubMed  Google Scholar 

  122. Hentschke, H., Schwarz, C. & Antkowiak, B. Neocortex is the major target of sedative concentrations of volatile anaesthetics: strong depression of firing rates and increase of GABAA receptor-mediated inhibition. Eur. J. Neurosci. 21, 93–102 (2005).

    PubMed  Google Scholar 

  123. Velly, L. J. et al. Differential dynamic of action on cortical and subcortical structures of anesthetic agents during induction of anesthesia. Anesthesiology 107, 202–212 (2007).

    CAS  PubMed  Google Scholar 

  124. Sherin, J. E., Shiromani, P. J., McCarley, R. W. & Saper, C. B. Activation of ventrolateral preoptic neurons during sleep. Science 271, 216–219 (1996). This paper, together with reference 126, showed the importance of the activation of VLPO neurons during sleep.

    CAS  PubMed  Google Scholar 

  125. Sherin, J. E., Elmquist, J. K., Torrealba, F. & Saper, C. B. Innervation of histaminergic tuberomammillary neurons by GABAergic and galaninergic neurons in the ventrolateral preoptic nucleus of the rat. J. Neurosci. 18, 4705–4721 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Szymusiak, R., Alam, N., Steininger, T. L. & McGinty, D. Sleep-waking discharge patterns of ventrolateral preoptic/anterior hypothalamic neurons in rats. Brain Res. 803, 178–188 (1998).

    CAS  PubMed  Google Scholar 

  127. McGinty, D. & Szymusiak, R. Brain structures and mechanisms involved in the generation of NREM sleep: focus on the preoptic hypothalamus. Sleep Med. Rev. 5, 323–342 (2001).

    PubMed  Google Scholar 

  128. McGinty, D. et al. Sleep-promoting functions of the hypothalamic median preoptic nucleus: inhibition of arousal systems. Arch. Ital. Biol. 142, 501–509 (2004).

    CAS  PubMed  Google Scholar 

  129. Suntsova, N., Szymusiak, R., Alam, M. N., Guzman-Marin, R. & McGinty, D. Sleep-waking discharge patterns of median preoptic nucleus neurons in rats. J. Physiol. 543, 665–677 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Modirrousta, M., Mainville, L. & Jones, B. E. Gabaergic neurons with α2-adrenergic receptors in basal forebrain and preoptic area express c-Fos during sleep. Neuroscience 129, 803–810 (2004).

    CAS  PubMed  Google Scholar 

  131. Szymusiak, R. & McGinty, D. Sleep-related neuronal discharge in the basal forebrain of cats. Brain Res. 370, 82–92 (1986).

    CAS  PubMed  Google Scholar 

  132. Szymusiak, R. & McGinty, D. Sleep-waking discharge of basal forebrain projection neurons in cats. Brain Res. Bull. 22, 423–430 (1989).

    CAS  PubMed  Google Scholar 

  133. Bremer, F. Cerebral hypnogenic centers. Ann. Neurol. 2, 1–6 (1977).

    CAS  PubMed  Google Scholar 

  134. Saper, C. B., Chou, T. C. & Scammell, T. E. The sleep switch: hypothalamic control of sleep and wakefulness. Trends Neurosci. 24, 726–731 (2001). This short review provides an excellent overview of the importance of the hypothalamus in the sleep–wake cycle. It includes a persuasive argument regarding the possible importance of reciprocal inhibitory connections between sleep- and wake-active neurons and the resulting flip-flop bi-stability.

    CAS  PubMed  Google Scholar 

  135. Chou, T. C. et al. Afferents to the ventrolateral preoptic nucleus. J. Neurosci. 22, 977–990 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  136. Lu, J., Sherman, D., Devor, M. & Saper, C. B. A putative flip-flop switch for control of REM sleep. Nature 441, 589–594 (2006).

    CAS  PubMed  Google Scholar 

  137. Tung, A., Szafran, M. J., Bluhm, B. & Mendelson, W. B. Sleep deprivation potentiates the onset and duration of loss of righting reflex induced by propofol and isoflurane. Anesthesiology 97, 906–911 (2002).

    CAS  PubMed  Google Scholar 

  138. Tung, A., Lynch, J. P. & Mendelson, W. B. Prolonged sedation with propofol in the rat does not result in sleep deprivation. Anesth. Analg. 92, 1232–1236 (2001).

    CAS  PubMed  Google Scholar 

  139. Lydic, R. & Baghdoyan, H. A. Sleep, anesthesiology, and the neurobiology of arousal state control. Anesthesiology 103, 1268–1295 (2005).

    PubMed  Google Scholar 

  140. Steriade, M., Pare, D., Parent, A. & Smith, Y. Projections of cholinergic and non-cholinergic neurons of the brainstem core to relay and associational thalamic nuclei in the cat and macaque monkey. Neuroscience 25, 47–67 (1988).

    CAS  PubMed  Google Scholar 

  141. Datta, S. Neuronal activity in the peribrachial area: relationship to behavioral state control. Neurosci. Biobehav. Rev. 19, 67–84 (1995).

    CAS  PubMed  Google Scholar 

  142. Jones, B. E. From waking to sleeping: neuronal and chemical substrates. Trends Pharmacol. Sci. 26, 578–586 (2005).

    CAS  PubMed  Google Scholar 

  143. Alkire, M. T., McReynolds, J. R., Hahn, E. L. & Trivedi, A. N. Thalamic microinjection of nicotine reverses sevoflurane-induced loss of righting reflex in the rat. Anesthesiology 107, 264–272 (2007).

    CAS  PubMed  Google Scholar 

  144. Meuret, P., Backman, S. B., Bonhomme, V., Plourde, G. & Fiset, P. Physostigmine reverses propofol-induced unconsciousness and attenuation of the auditory steady state response and bispectral index in human volunteers. Anesthesiology 93, 708–717 (2000).

    CAS  PubMed  Google Scholar 

  145. Hirota, K., Hashimoto, Y. & Lambert, D. G. Interaction of intravenous anesthetics with recombinant human M1-M3 muscarinic receptors expressed in chinese hamster ovary cells. Anesth. Analg. 95, 1607–1610 (2002).

    CAS  PubMed  Google Scholar 

  146. Flood, P., Ramirez-Latorre, J. & Role, L. α4 β2 neuronal nicotinic acetylcholine receptors in the central nervous system are inhibited by isoflurane and propofol, but α7-type nicotinic acetylcholine receptors are unaffected. Anesthesiology 86, 859–865 (1997).

    CAS  PubMed  Google Scholar 

  147. Violet, J. M., Downie, D. L., Nakisa, R. C., Lieb, W. R. & Franks, N. P. Differential sensitivities of mammalian neuronal and muscle nicotinic acetylcholine receptors to general anesthetics. Anesthesiology 86, 866–874 (1997).

    CAS  PubMed  Google Scholar 

  148. Flood, P., Sonner, J. M., Gong, D. & Coates, K. M. Heteromeric nicotinic inhibition by isoflurane does not mediate MAC or loss of righting reflex. Anesthesiology 97, 902–905 (2002).

    CAS  PubMed  Google Scholar 

  149. Devor, M. & Zalkind, V. Reversible analgesia, atonia, and loss of consciousness on bilateral intracerebral microinjection of pentobarbital. Pain 94, 101–112 (2001). This paper, together with reference 150, provided a potentially important demonstration that microinjections of barbiturates into the brain stem can induce reversible LORR and delta oscillations in the EEG that mirror those seen with systemic application.

    CAS  PubMed  Google Scholar 

  150. Sukhotinsky, I. et al. Neural pathways associated with loss of consciousness caused by intracerebral microinjection of GABAA-active anesthetics. Eur. J. Neurosci. 25, 1417–1436 (2007).

    CAS  PubMed  Google Scholar 

  151. Parvizi, J. & Damasio, A. R. Neuroanatomical correlates of brainstem coma. Brain 126, 1524–1536 (2003).

    PubMed  Google Scholar 

  152. Aston-Jones, G. & Bloom, F. E. Activity of norepinephrine-containing locus coeruleus neurons in behaving rats anticipates fluctuations in the sleep-waking cycle. J. Neurosci. 1, 876–886 (1981).

    CAS  PubMed  PubMed Central  Google Scholar 

  153. Berridge, C. W. & Foote, S. L. Effects of locus coeruleus activation on electroencephalographic activity in neocortex and hippocampus. J. Neurosci. 11, 3135–3145 (1991).

    CAS  PubMed  PubMed Central  Google Scholar 

  154. Berridge, C. W. & Waterhouse, B. D. The locus coeruleus-noradrenergic system: modulation of behavioral state and state-dependent cognitive processes. Brain Res. Brain Res. Rev. 42, 33–84 (2003).

    PubMed  Google Scholar 

  155. Ma, J., Shen, B., Stewart, L. S., Herrick, I. A. & Leung, L. S. The septohippocampal system participates in general anesthesia. J. Neurosci. 22, RC200 (2002).

    PubMed  PubMed Central  Google Scholar 

  156. Correa-Sales, C., Rabin, B. C. & Maze, M. A hypnotic response to dexmedetomidine, an α2 agonist, is mediated in the locus coeruleus in rats. Anesthesiology 76, 948–952 (1992). This key study showed that the adrenergic agonist dexmedetomidine is likely to exert a substantial part of its sedative/hypnotic effects through direct actions on the locus coeruleus.

    CAS  PubMed  Google Scholar 

  157. Nelson, L. E. et al. The α2-adrenoceptor agonist dexmedetomidine converges on an endogenous sleep-promoting pathway to exert its sedative effects. Anesthesiology 98, 428–436 (2003).

    CAS  PubMed  Google Scholar 

  158. Ursin, R. Serotonin and sleep. Sleep Med. Rev. 6, 55–69 (2002).

    PubMed  Google Scholar 

  159. Kanbayashi, T., Honda, K., Kodama, T., Mignot, E. & Nishino, S. Implication of dopaminergic mechanisms in the wake-promoting effects of amphetamine: a study of D- and L-derivatives in canine narcolepsy. Neuroscience 99, 651–659 (2000).

    CAS  PubMed  Google Scholar 

  160. Isaac, S. O. & Berridge, C. W. Wake-promoting actions of dopamine D1 and D2 receptor stimulation. J. Pharmacol. Exp. Ther. 307, 386–394 (2003).

    CAS  PubMed  Google Scholar 

  161. Kelland, M. D., Chiodo, L. A. & Freeman, A. S. Anesthetic influences on the basal activity and pharmacological responsiveness of nigrostriatal dopamine neurons. Synapse 6, 207–209 (1990).

    CAS  PubMed  Google Scholar 

  162. Lu, J., Jhou, T. C. & Saper, C. B. Identification of wake-active dopaminergic neurons in the ventral periaqueductal gray matter. J. Neurosci. 26, 193–202 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  163. Jasper, H. H. & Tessier, J. Acetylcholine liberation from cerebral cortex during paradoxical (REM) sleep. Science 172, 601–602 (1971).

    CAS  PubMed  Google Scholar 

  164. Metherate, R., Cox, C. L. & Ashe, J. H. Cellular bases of neocortical activation: modulation of neural oscillations by the nucleus basalis and endogenous acetylcholine. J. Neurosci. 12, 4701–4711 (1992).

    CAS  PubMed  PubMed Central  Google Scholar 

  165. Rasmusson, D. D., Clow, K. & Szerb, J. C. Modification of neocortical acetylcholine release and electroencephalogram desynchronization due to brainstem stimulation by drugs applied to the basal forebrain. Neuroscience 60, 665–677 (1994).

    CAS  PubMed  Google Scholar 

  166. Szymusiak, R. Magnocellular nuclei of the basal forebrain: substrates of sleep and arousal regulation. Sleep 18, 478–500 (1995).

    CAS  PubMed  Google Scholar 

  167. Mendelson, W. B. Sleep induction by microinjection of pentobarbital into the medial preoptic area in rats. Life Sci. 59, 1821–1828 (1996).

    CAS  PubMed  Google Scholar 

  168. Tung, A., Bluhm, B. & Mendelson, W. B. The hypnotic effect of propofol in the medial preoptic area of the rat. Life Sci. 69, 855–862 (2001).

    CAS  PubMed  Google Scholar 

  169. Haas, H. & Panula, P. The role of histamine and the tuberomamillary nucleus in the nervous system. Nature Rev. Neurosci. 4, 121–130 (2003).

    CAS  Google Scholar 

  170. Panula, P., Yang, H. Y. & Costa, E. Histamine-containing neurons in the rat hypothalamus. Proc. Natl Acad. Sci. USA 81, 2572–2576 (1984).

    CAS  PubMed  PubMed Central  Google Scholar 

  171. Strecker, R. E. et al. Extracellular histamine levels in the feline preoptic/anterior hypothalamic area during natural sleep-wakefulness and prolonged wakefulness: an in vivo microdialysis study. Neuroscience 113, 663–670 (2002).

    CAS  PubMed  Google Scholar 

  172. Ko, E. M., Estabrooke, I. V., McCarthy, M. & Scammell, T. E. Wake-related activity of tuberomammillary neurons in rats. Brain Res. 992, 220–226 (2003).

    CAS  PubMed  Google Scholar 

  173. Takahashi, K., Lin, J. S. & Sakai, K. Neuronal activity of histaminergic tuberomammillary neurons during wake-sleep states in the mouse. J. Neurosci. 26, 10292–10298 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  174. Nelson, L. E. et al. The sedative component of anesthesia is mediated by GABAA receptors in an endogenous sleep pathway. Nature Neurosci. 5, 979–984 (2002). This paper, together with reference 157, showed how anaesthetics might mediate their effects in the brain by recruiting natural sleep pathways in the hypothalamus.

    CAS  PubMed  Google Scholar 

  175. Mammoto, T. et al. Interactions between neuronal histamine and halothane anesthesia in rats. J. Neurochem. 69, 406–411 (1997).

    CAS  PubMed  Google Scholar 

  176. Kushikata, T. et al. Orexinergic neurons and barbiturate anesthesia. Neuroscience 121, 855–863 (2003).

    CAS  PubMed  Google Scholar 

  177. Kelz, M. B. et al. An essential role for orexins in emergence from general anesthesia. Proc. Natl Acad. Sci. USA 105, 1309–1314 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  178. Sakurai, T. The neural circuit of orexin (hypocretin): maintaining sleep and wakefulness. Nature Rev. Neurosci. 8, 171–181 (2007).

    CAS  Google Scholar 

  179. Estabrooke, I. V. et al. Fos expression in orexin neurons varies with behavioral state. J. Neurosci. 21, 1656–1662 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  180. Lee, M. G., Hassani, O. K. & Jones, B. E. Discharge of identified orexin/hypocretin neurons across the sleep-waking cycle. J. Neurosci. 25, 6716–6720 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  181. Hagan, J. J. et al. Orexin A activates locus coeruleus cell firing and increases arousal in the rat. Proc. Natl Acad. Sci. USA 96, 10911–10916 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  182. Huang, Z. L. et al. Arousal effect of orexin A depends on activation of the histaminergic system. Proc. Natl Acad. Sci. USA 98, 9965–9970 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  183. Abraham, M. H., Lieb, W. R. & Franks, N. P. Role of hydrogen bonding in general anesthesia. J. Pharm. Sci. 80, 719–724 (1991).

    CAS  PubMed  Google Scholar 

  184. Krasowski, M. D. Contradicting a unitary theory of general anesthetic action: a history of three compounds from 1901 to 2001. Bull. Anesth. Hist. 21, 1, 4–8 (2003).

    Google Scholar 

  185. Seeman, P. The membrane actions of anesthetics and tranquilizers. Pharmacol. Rev. 24, 583–655 (1972).

    CAS  PubMed  Google Scholar 

  186. Miller, K. W., Paton, W. D., Smith, R. A. & Smith, E. B. The pressure reversal of general anesthesia and the critical volume hypothesis. Mol. Pharmacol. 9, 131–143 (1973).

    CAS  PubMed  Google Scholar 

  187. Trudell, J. R. A unitary theory of anesthesia based on lateral phase separations in nerve membranes. Anesthesiology 46, 5–10 (1977).

    CAS  PubMed  Google Scholar 

  188. Franks, N. P. & Lieb, W. R. Molecular mechanisms of general anaesthesia. Nature 300, 487–493 (1982).

    CAS  PubMed  Google Scholar 

  189. Franks, N. P. & Lieb, W. R. Partitioning of long-chain alcohols into lipid bilayers: implications for mechanisms of general anesthesia. Proc. Natl Acad. Sci. USA 83, 5116–5120 (1986).

    CAS  PubMed  PubMed Central  Google Scholar 

  190. Franks, N. P. & Lieb, W. R. Do general anaesthetics act by competitive binding to specific receptors? Nature 310, 599–601 (1984). This influential publication directed the field towards identifying specific protein targets rather than exploring non-specific lipid mechanisms of anaesthetic action.

    CAS  PubMed  Google Scholar 

  191. Franks, N. P. & Lieb, W. R. Mapping of general anaesthetic target sites provides a molecular basis for cutoff effects. Nature 316, 349–351 (1985).

    CAS  PubMed  Google Scholar 

  192. Curry, S., Lieb, W. R. & Franks, N. P. Effects of general anesthetics on the bacterial luciferase enzyme from Vibrio harveyi: an anesthetic target site with differential sensitivity. Biochemistry 29, 4641–4652 (1990).

    CAS  PubMed  Google Scholar 

  193. Dickinson, R., Franks, N. P. & Lieb, W. R. Thermodynamics of anesthetic/protein interactions. Temperature studies on firefly luciferase. Biophys. J. 64, 1264–1271 (1993).

    CAS  PubMed  PubMed Central  Google Scholar 

  194. Franks, N. P. & Lieb, W. R. Temperature dependence of the potency of volatile general anesthetics: implications for in vitro experiments. Anesthesiology 84, 716–720 (1996).

    CAS  PubMed  Google Scholar 

  195. Bhattacharya, A. A., Curry, S. & Franks, N. P. Binding of the general anesthetics propofol and halothane to human serum albumin. High resolution crystal structures. J. Biol. Chem. 275, 38731–38738 (2000).

    CAS  PubMed  Google Scholar 

  196. Liu, R., Loll, P. J. & Eckenhoff, R. G. Structural basis for high-affinity volatile anesthetic binding in a natural 4-helix bundle protein. FASEB J. 19, 567–576 (2005).

    PubMed  Google Scholar 

  197. Colloc'h, N. et al. Protein crystallography under xenon and nitrous oxide pressure: comparison with in vivo pharmacology studies and implications for the mechanism of inhaled anesthetic action. Biophys. J. 92, 217–224 (2007).

    CAS  PubMed  Google Scholar 

  198. Bertaccini, E. J., Trudell, J. R. & Franks, N. P. The common chemical motifs within anesthetic binding sites. Anesth. Analg. 104, 318–324 (2007).

    CAS  PubMed  Google Scholar 

  199. Bai, D., Pennefather, P. S., MacDonald, J. F. & Orser, B. A. The general anesthetic propofol slows deactivation and desensitization of GABAA receptors. J. Neurosci. 19, 10635–10646 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  200. Li, X. & Pearce, R. A. Effects of halothane on GABAA receptor kinetics: evidence for slowed agonist unbinding. J. Neurosci. 20, 899–907 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  201. Rusch, D., Zhong, H. & Forman, S. A. Gating allosterism at a single class of etomidate sites on α1β2γ2L GABAA receptors accounts for both direct activation and agonist modulation. J. Biol. Chem. 279, 20982–20992 (2004).

    PubMed  Google Scholar 

  202. Sofroniew, M. V., Priestley, J. V., Consolazione, A., Eckenstein, F. & Cuello, A. C. Cholinergic projections from the midbrain and pons to the thalamus in the rat, identified by combined retrograde tracing and choline acetyltransferase immunohistochemistry. Brain Res. 329, 213–223 (1985).

    CAS  PubMed  Google Scholar 

  203. Pare, D., Smith, Y., Parent, A. & Steriade, M. Projections of brainstem core cholinergic and non-cholinergic neurons of cat to intralaminar and reticular thalamic nuclei. Neuroscience 25, 69–86 (1988).

    CAS  PubMed  Google Scholar 

  204. Satoh, K. & Fibiger, H. C. Cholinergic neurons of the laterodorsal tegmental nucleus: efferent and afferent connections. J. Comp. Neurol. 253, 277–302 (1986).

    CAS  PubMed  Google Scholar 

  205. Gallopin, T. et al. Identification of sleep-promoting neurons in vitro. Nature 404, 992–995 (2000).

    CAS  PubMed  Google Scholar 

  206. Lindvall, O., Bjorklund, A., Nobin, A. & Stenevi, U. The adrenergic innervation of the rat thalamus as revealed by the glyoxylic acid fluorescence method. J. Comp. Neurol. 154, 317–347 (1974).

    CAS  PubMed  Google Scholar 

  207. Jones, B. E. & Yang, T. Z. The efferent projections from the reticular formation and the locus coeruleus studied by anterograde and retrograde axonal transport in the rat. J. Comp. Neurol. 242, 56–92 (1985).

    CAS  PubMed  Google Scholar 

  208. Krout, K. E., Belzer, R. E. & Loewy, A. D. Brainstem projections to midline and intralaminar thalamic nuclei of the rat. J. Comp. Neurol. 448, 53–101 (2002).

    PubMed  Google Scholar 

  209. Vertes, R. P. A PHA-L analysis of ascending projections of the dorsal raphe nucleus in the rat. J. Comp. Neurol. 313, 643–668 (1991).

    CAS  PubMed  Google Scholar 

  210. Peyron, C. et al. Neurons containing hypocretin (orexin) project to multiple neuronal systems. J. Neurosci. 18, 9996–10015 (1998).

    CAS  PubMed  PubMed Central  Google Scholar 

  211. Bayer, L. et al. Selective action of orexin (hypocretin) on nonspecific thalamocortical projection neurons. J. Neurosci. 22, 7835–7839 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  212. Inagaki, N. et al. Organization of histaminergic fibers in the rat brain. J. Comp. Neurol. 273, 283–300 (1988).

    CAS  PubMed  Google Scholar 

  213. Panula, P., Pirvola, U., Auvinen, S. & Airaksinen, M. S. Histamine-immunoreactive nerve fibers in the rat brain. Neuroscience 28, 585–610 (1989).

    CAS  PubMed  Google Scholar 

  214. Steininger, T. L., Gong, H., McGinty, D. & Szymusiak, R. Subregional organization of preoptic area/anterior hypothalamic projections to arousal-related monoaminergic cell groups. J. Comp. Neurol. 429, 638–653 (2001).

    CAS  PubMed  Google Scholar 

  215. Methippara, M. M., Alam, M. N., Szymusiak, R. & McGinty, D. Preoptic area warming inhibits wake-active neurons in the perifornical lateral hypothalamus. Brain Res. 960, 165–173 (2003).

    CAS  PubMed  Google Scholar 

  216. Uschakov, A., Gong, H., McGinty, D. & Szymusiak, R. Efferent projections from the median preoptic nucleus to sleep- and arousal-regulatory nuclei in the rat brain. Neuroscience 150, 104–120 (2007).

    CAS  PubMed  Google Scholar 

  217. Henny, P. & Jones, B. E. Innervation of orexin/hypocretin neurons by GABAergic, glutamatergic or cholinergic basal forebrain terminals evidenced by immunostaining for presynaptic vesicular transporter and postsynaptic scaffolding proteins. J. Comp. Neurol. 499, 645–661 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  218. Semba, K., Reiner, P. B., McGeer, E. G. & Fibiger, H. C. Brainstem afferents to the magnocellular basal forebrain studied by axonal transport, immunohistochemistry, and electrophysiology in the rat. J. Comp. Neurol. 267, 433–453 (1988).

    CAS  PubMed  Google Scholar 

  219. Katoh, T., Bito, H. & Sato, S. Influence of age on hypnotic requirement, bispectral index, and 95% spectral edge frequency associated with sedation induced by sevoflurane. Anesthesiology 92, 55–61 (2000).

    CAS  PubMed  Google Scholar 

  220. Smith, C. et al. The interaction of fentanyl on the Cp50 of propofol for loss of consciousness and skin incision. Anesthesiology 81, 820–828; discussion 826A (1994).

    CAS  PubMed  Google Scholar 

  221. Rashin, A. A., Rashin, B. H., Rashin, A. & Abagyan, R. Evaluating the energetics of empty cavities and internal mutations in proteins. Protein Sci. 6, 2143–2158 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  222. Richardson, J. E. et al. A conserved tyrosine in the β2 subunit M4 segment is a determinant of γ-aminobutyric acid type A receptor sensitivity to propofol. Anesthesiology 107, 412–418 (2007).

    CAS  PubMed  Google Scholar 

  223. Ronald, K. M., Mirshahi, T. & Woodward, J. J. Ethanol inhibition of N-methyl-D-aspartate receptors is reduced by site-directed mutagenesis of a transmembrane domain phenylalanine residue. J. Biol. Chem. 276, 44729–44735 (2001).

    CAS  PubMed  Google Scholar 

  224. Honse, Y., Ren, H., Lipsky, R. H. & Peoples, R. W. Sites in the fourth membrane-associated domain regulate alcohol sensitivity of the NMDA receptor. Neuropharmacology 46, 647–654 (2004).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

I thank S. Brickley and A. Zecharia for many helpful and stimulating discussions during the writing of this Review and their assistance in the preparation of the figures, P. Brick and S. Curry for help with the crystallographic images and C. Robledo for providing the EEG data in figure 4. I also thank the Medical Research Council UK for support.

Author information

Authors and Affiliations

Authors

Supplementary information

Supplementary information S1 (table)

Free aqueous concentrations for human loss of consciousness and rodent loss of righting reflex* (PDF 178 kb)

Related links

Related links

FURTHER INFORMATION

Nicholas Franks' homepage

Glossary

Tonic inhibition

Background inhibition resulting from persistent low-level activation of GABAA receptors. The main component of this current is thought to be mediated by extrasynaptic receptors with relatively high affinities for GABA.

Phasic inhibition

Inhibition resulting from the transient release of a high concentration of GABA from presynaptic terminals.

Sulphydryl reagents

Various chemical reagents that selectively react with SH groups and which are hence used to label the sulphur-containing amino acids (cysteine and methionine).

Photoreactive anaesthetics

Modified anaesthetics that contain groups (such as diazirines) that are converted into reactive species following stimulation by light; they are used to selectively and irreversibly label an anaesthetic-binding site.

Righting reflex

The postural response of an animal when placed on its back or side to reorient itself such that its paws or feet are oriented towards the ground.

Blood–brain barrier

A layer of tightly apposed endothelial cells that lines blood capillaries and forms a barrier to the diffusion of substances from the blood into the brain. Apolar molecules, such as general anaesthetics, however, readily permeate this barrier.

Positron-emission tomography

(PET). A technique that uses positron emitters, such as 15O, to measure local changes in cerebral blood flow, which can be interpreted as reflecting local neuronal activity.

Somatosensory-evoked potentials

Electrical signals generated by the nervous system following a mechanical or, more usually, an electrical stimulation in the periphery that reflect sequential activation of neural structures along the somatosensory pathways.

Inhibitory shunt

Inhibition of excitatory inputs owing to an increase in neuronal membrane conductance.

Encephalitis lethargica

A condition that affected millions of people in the 1920s and which was first described by Constantin von Economo. It was characterized by a high fever, profound lethargy and many other neurological changes that, in extreme cases, led to irreversible coma. It was sometimes described as 'sleeping sickness'.

c-fos

An immediate-early-gene product that can be upregulated by various cellular stimuli, including neuronal excitability. It is sometimes used as a surrogate measure of neuronal activity.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Franks, N. General anaesthesia: from molecular targets to neuronal pathways of sleep and arousal. Nat Rev Neurosci 9, 370–386 (2008). https://doi.org/10.1038/nrn2372

Download citation

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrn2372

This article is cited by

Search

Quick links

Nature Briefing

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

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