Skip to main content

Advertisement

Log in

How the cortico-thalamic feedback affects the EEG power spectrum over frontal and occipital regions during propofol-induced sedation

  • Published:
Journal of Computational Neuroscience Aims and scope Submit manuscript

Abstract

Increasing concentrations of the anaesthetic agent propofol initially induces sedation before achieving full general anaesthesia. During this state of anaesthesia, the observed specific changes in electroencephalographic (EEG) rhythms comprise increased activity in the δ− (0.5−4 Hz) and α− (8−13 Hz) frequency bands over the frontal region, but increased δ− and decreased α−activity over the occipital region. It is known that the cortex, the thalamus, and the thalamo-cortical feedback loop contribute to some degree to the propofol-induced changes in the EEG power spectrum. However the precise role of each structure to the dynamics of the EEG is unknown. In this paper we apply a thalamo-cortical neuronal population model to reproduce the power spectrum changes in EEG during propofol-induced anaesthesia sedation. The model reproduces the power spectrum features observed experimentally both in frontal and occipital electrodes. Moreover, a detailed analysis of the model indicates the importance of multiple resting states in brain activity. The work suggests that the α−activity originates from the cortico-thalamic relay interaction, whereas the emergence of δ−activity results from the full cortico-reticular-relay-cortical feedback loop with a prominent enforced thalamic reticular-relay interaction. This model suggests an important role for synaptic GABAergic receptors at relay neurons and, more generally, for the thalamus in the generation of both the δ− and the α− EEG patterns that are seen during propofol anaesthesia sedation.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

References

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

    Article  CAS  PubMed  Google Scholar 

  • Alkire, M.T., Hudetz, A.G., & Tononi, G. (2008). Consciousness and anesthesia. Science, 322, 876–880. doi:10.1126/science.1149213.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Amari, S. (1977). Dynamics of pattern formation in lateral-inhibition type neural fields. Biological Cybernetics, 27, 77–87.

    Article  CAS  PubMed  Google Scholar 

  • Antkowiak, B. (2002). In vitro networks: cortical mechanisms of anaesthetic action. British Journal of Anaesthesia, 89(1), 102–111.

    Article  CAS  PubMed  Google Scholar 

  • Belelli, D., Harrison, N.L., Maguire, J., Macdonald, R.L., Walker, M.C., & Cope, D.W. (2009). Extra-synaptic GABA A receptors: form, pharmacology, and function. The Journal of Neuroscience, 29(41), 12757–12763.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Bojak, I., & Liley, D.T.J. (2005). Modeling the effects of anesthesia on the electroencephalogram. Physical Review E, 71, 041902.

    Article  CAS  Google Scholar 

  • Boly, M., Moran, R., Murphy, M., Boveroux, P., Bruno, M.A., Noirhomme, Q., Ledoux, D., Bonhomme, V., Brichant, J.F., Tononi, G., Laureys, S., & Friston, K.I. (2012). Connectivity changes underlying spectral EEG changes during propofol-induced loss of consciousness. The Journal of Neuroscience, 32(20), 7082–7090.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Breakspear, M., Roberts, J., Terry, J.R., Rodrigues, S., Mahant, N., & Robinson, P.A. (2006). A unifying explanation of primary generalized seizures through nonlinear brain modeling and bifurcation analysis. Cerebral Cortex, 16, 1296–1313.

    Article  CAS  PubMed  Google Scholar 

  • Bressloff, P. (2012). Spatiotemporal dynamics of continuum neural fields. Journal of Physics A, 45, 033001.

    Article  Google Scholar 

  • Byoung-Kyong, M. (2010). A thalamic reticular networking model of consciousness. Theoretical Biology and Medical Modelling, 7.

  • Chau, P-L. (2010). New insights into the molecular mechanisms of general anaesthetics. British Journal of Pharmacology, 161, 288–307.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Chiang, A.K., Rennie, C.J., Robinson, P.A., Roberts, J.A., Rigozzi, M.K., Whitehouse, R.W., Hamilton, R.J., & Gordon, E. (2008). Automated characterization of multiple alpha peaks in multi-site electroencephalograms. Journal of Neuroscience Methods, 168, 396–411.

    Article  CAS  PubMed  Google Scholar 

  • Ching, S., Cimenser, A., Purdon, P.L., Brown, E.N., & Kopell, N.J. (2010). Thalamocortical model for a propofol-induced-rhythm associated with loss of consciousness. Proceedings of the National Academy of Sciences of the United States of America, 107(52), 22665–22670.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Ching, S., Purdon, P.L., Vijayand, S., Kopell, N.J., & Brown, E.N. (2012). A neurophysiological metabolic model for burst suppression. Proceedings of the National Academy of Sciences of the United States of America, 109 (8), 3095–3100.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Cimenser, A., Purdon, P.L., Pierce, E.T., Walsh, J.L., Salazar-Gomez, A.F., Harrell, P.G., Tavares-Stoeckel, C., Habeeb, K., & Brown, E.N. (2011). Tracking brain states under general anesthesia by using global coherence analysis. Proceedings of the National Academy of Sciences of the United States of America, 108(21), 8832–8837.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Dang-Vu, T.T., Schabus, M., Desseilles, M., Albouy, G., Boly, M., Darsaud, A., Gais, S., Rauchs, G., Sterpenich, V., Vandewalle, G., Carrier, J., Moonen, G., Balteau, E., Degueldre, C., Luxen, A., Phillips, C., & Maquet, P. (2008). Spontaneous neural activity during human slow wave sleep. Proceedings of the National Academy of Sciences of the United States of America, 105(39), 15160–15165.

    Article  PubMed Central  PubMed  Google Scholar 

  • David, O., & Friston, K.J. (2003). A neural mass model for meg/eeg: coupling and neuronal dynamics. NeuroImage, 20, 1743– 1755.

    Article  PubMed  Google Scholar 

  • David, O., Kiebel, S.J., Harrison, L.M., Mattout, J., Kilner, J.M., & Friston, K.J. (2006). Dynamic causal modeling of evoked responses in eeg and meg. NeuroImage, 30, 1255–1272.

    Article  PubMed  Google Scholar 

  • Deco, G., Jirsa, V.K., Robinson, P.A., Breakspear, M., & Friston, K. (2008). The dynamic brain: from spiking neurons to neural masses and cortical fields. PLoS Computational Biology, 4(8).

  • Drover, J.D., Schiff, N.D., & Victor, J.D. (2010). Dynamics of coupled thalamocortical modules. Journal of Computational Neuroscience, 28, 605–616.

    Article  PubMed  Google Scholar 

  • Farrant, M., & Nusser, Z. (2005). Variations on an inhibitory theme: phasic and tonic activation of GABA A receptors. Nature Reviews Neuroscience, 6, 215–229.

    Article  CAS  PubMed  Google Scholar 

  • Feshchenko, V.A., Veselis, R.A., & Reinsel, R.A. (2004). Propofol-induced alpha rhythm. Neuropsychobiology, 50(3), 257–266.

    Article  CAS  PubMed  Google Scholar 

  • Foster, B.L., Bojak, I., & Liley, D.T.J. (2008). Population based models of cortical drug response: insights from anaesthesia. Cognitive Neurodynamics, 2, 283–296.

    Article  PubMed Central  PubMed  Google Scholar 

  • Franks, N.P. (2008). General anesthesia: from molecular targets to neuronal pathways of sleep and arousal. Nature Reviews Neuroscience, 9, 370–386. doi:10.1038/nrn2372.

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  • Freeman, W.J. (1979). Nonlinear gain mediating cortical stimulus-response relations. Biological Cybernetics, 33, 237–247.

    Article  CAS  PubMed  Google Scholar 

  • Freestone, D.R., Aram, P., Dewar, M., Scerri, K., Grayden, D.B., & Kadirkamanathan, V. (2011). A data-driven framework for neural field modeling. NeuroImage, 56(3), 1043–1058.

    Article  CAS  PubMed  Google Scholar 

  • Friston, K.J., Harrison, L., & Penny, W. (2003). Dynamic causal modelling. NeuroImage, 19, 273–1302.

    Google Scholar 

  • Garcia, P.S., Kolesky, S.E., & Jenkins, A. (2010). General anesthetic actions on GABA A receptors. Current Neuropharmacology, 8(1), 2–9.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Grasshoff, C., Drexler, B., Rudolph, U., & Antkowiak, B. (2006). Anaesthetic drugs: linking molecular actions to clinical effects. Current Pharmaceutical Design, 12(28), 3665–3679.

    Article  CAS  PubMed  Google Scholar 

  • Gugino, L.D., Chabot, R.J., Prichep, L.S., John, E.R., Formanek, V., & Aglio, L.S. (2001). Quantitative EEG changes associated with loss and return of conscious- ness in healthy adult volunteers anaesthetized with propofol or sevoflurane. British Journal of Anaesthesia, 87, 421–428.

    Article  CAS  PubMed  Google Scholar 

  • Hashemi, M., Hutt, A., & Sleigh, J. (2014). Anesthetic action on extra-synaptic receptors: effects in neural population models of EEG activity. Journals of Frontiers in Systems Neuroscience, 8(232).

  • Hazeaux, C., Tisserant, D., Vespignani, H., Hummer-Sigiel, L., Kwan-Ning, V., & Laxenaire, M.C. (1987). Electroencephalographic changes produced by propofol. Annales Françaises d’Anesthèsie et de Rèanimation, 6, 261–266.

    Article  CAS  PubMed  Google Scholar 

  • Hindriks, R., & van Putten, M.J.A.M. (2012). Meanfield modeling of propofol-induced changes in spontaneous EEG rhythms. Neuroimage, 60, 2323–2344.

    Article  CAS  PubMed  Google Scholar 

  • Hindriks, R., & van Putten, M.J.A.M. (2013). Thalamo-cortical mechanisms underlying changes in amplitude and frequency of human alpha oscillations. Neuroimage, 70, 150–163.

    Article  PubMed  Google Scholar 

  • Hutt, A. (2012). The population firing rate in the presence of GABAergic tonic inhibition in single neurons and application to general anaesthesia. Cognitive Neurodynamics, 6, 227–237.

    Article  PubMed Central  PubMed  Google Scholar 

  • Hutt, A. (2013). The anaesthetic propofol shifts the frequency of maximum spectral power in EEG during general anaesthesia: analytical insights from a linear model. Frontiers in Computational Neuroscience, 7, 2.

    Article  PubMed Central  PubMed  Google Scholar 

  • Hutt, A., & Buhry, L. (2014). Study of GABAergic extra-synaptic tonic inhibition in single neurons and neural populations by traversing neural scales: application to propofol-induced anaesthesia. Journal of Computational Neuroscience. in press.

  • Hutt, A., & Longtin, A. (2009). Effects of the anesthetic agent propofol on neural populations. Cognitive Neurodynamics, 4(1), 37–59.

  • Hutt, A., Bestehorn, M., & Wennekers, T. (2003). Pattern formation in intracortical neuronal fields. Network: Computation in Neural Systems, 14, 351–368.

    Article  Google Scholar 

  • Jirsa, V.K., & Haken, H. (1996). Field theory of electromagnetic brain activity. Physical Review Letters, 77 (5), 960–963.

    Article  CAS  PubMed  Google Scholar 

  • Johnson, B.W., Sleigh, J.W., Kirk, I.J., & Williams, M.L. (2003). High-density EEG mapping during general anaesthesia with xenon and propofol: a pilot study. Anaesthesia and Intensive Care, 31(2), 155–163.

    CAS  PubMed  Google Scholar 

  • Kitamura, A., Marszalec, W., Yeh, J.Z., & Narahashi, T. (2002). Effects of halothane and propofol on excitatory and inhibitory synaptic transmission in rat cortical neurons. Journal de Pharmacologie, 304(1), 162–171.

    Google Scholar 

  • Kullmann, D.M., Ruiz, A., Rusakov, D.M., Scott, R., Semyanov, A., & Walker, M.C. (2005). Presynaptic, extrasynaptic and axonal GABA A receptors in the cns: where and why?. Progress in Biophysics and Molecular Biology, 87, 33–46.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Lee, U., Oh, G., Kim, S., Noh, G., Choi, B., & Mashour, G.A. (2010). Brain networks maintain a scale-free organization across consciousness, anesthesia, and recovery: Evidence for adaptive reconfiguration. Anesthesiology, 113(5), 1081–1091.

    Article  PubMed Central  PubMed  Google Scholar 

  • Lewis, L.D., Weiner, V.S., Mukamel, E.A., Donoghue, J.A., Eskandar, E.N., Madsen, J.R., Anderson, W.S., Hochberg, L.R., Cash, S.S., Brown, E.N., & Purdon, P.L. (2012). Rapid fragmentation of neuronal networks at the onset of propofol-induced unconsciousness. Proceedings of the National Academy of Sciences of the United States of America, 109(21), 3377–3386.

    Article  Google Scholar 

  • Liley, D.T.J., & Bojak, I. (2005). Understanding the transition to seizure by modeling the epileptiform activity of general anaesthetic agents. Journal of Clinical Neurophysiology, 22, 300–313.

    CAS  PubMed  Google Scholar 

  • Liley, D.T.J., & Walsh, M. (2013). The mesoscopic modeling of burst suppression during anesthesia. Frontiers in Computational Neuroscience, 7, 46.

    Article  PubMed Central  PubMed  Google Scholar 

  • Liley, D.T.J., & Wright, J.J. (1994). Intracortical connectivity of pyramidal and stellate cells: estimates of synaptic densities and coupling symmetry. Network: Computation in Neural Systems, 5, 175–189.

    Article  Google Scholar 

  • Liley, D.T.J., Cadusch, P.J., & Dafilis, M.P. (2002). A spatially continuous mean field theory of electrocortical activity. Network: Computation in Neural Systems, 13, 67–113.

    Article  Google Scholar 

  • Maquet, P., Degueldre, C., Delfiore, G., Aerts, J., Peters, J., Luxen, A., & Franck, G. (1997). Functional neuroanatomy of human slow wave sleep. The Journal of Neuroscience, 17(8), 2807–2812.

    CAS  PubMed  Google Scholar 

  • Massimini, M., Ferrarelli, F., Huber, R., Esser, S.K., Singh, H., & Tononi, G. (2005). Breakdown of cortical effective connectivity during sleep. Science, 309, 2228–2232.

    Article  CAS  PubMed  Google Scholar 

  • McCarthy, M.M., Brown, E.N., & Kopell, N. (2008). Potential network mechanisms mediating electroencephalographic beta rhythm changes during propofol-induced paradoxical excitation. The Journal of Neuroscience, 28(50), 13488–13504.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Molaee-Ardekani, B., Senhadji, L., Shamsollahi, M.B., Vosoughi-Vahdat, B., & Wodey, E. (2007). Brain activity modeling in general anesthesia: Enhancing local mean-field models using a slow adaptive firing rate. Physical Review E, 76, 041911.

    Article  CAS  Google Scholar 

  • Mukamel, E.A., Pirondini, E., Babadi, B., Wong, K.F.k., Pierce, E.T., Harrell, P.G., Walsh, J.L., Salazar-Gomez, A.F., Cash, S.S., Eskandar, E.N., Weiner, V.S., Brown, E.N., & Purdon, P.L. (2014). A transition in brain state during propofol-induced unconsciousness. Journal of Neuroscience, 34(3), 839–845.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Murphy, M., Bruno, M.A., Riedner, B.A., Boveroux, P., Noirhomme, Q., Landsness, E.C., Brichant, J.F., Phillips, C., Massimini, M., Laureys, S., Tononi, G., & Boly, M. (2011). Propofol anesthesia and sleep: A high-density EEG study. Sleep, 34(3), 283–291.

    PubMed Central  PubMed  Google Scholar 

  • Nunez, P.L. (1974). The brain wave equation: A model for the EEG. Mathematical Biosciences, 21, 279–291.

    Article  Google Scholar 

  • Nunez, P.L. (1981). Electrical fields of the brain. Oxford: Oxford University Press.

    Google Scholar 

  • Nunez, P.L., & Srinivasan, R. (2006). Electric fields of the brain: The neurophysics of EEG. New York - Oxford: Oxford University Press.

    Book  Google Scholar 

  • Orser, B.A. (2006). Extrasynaptic GABA A receptors are critical targets for sedative-hypnotic drugs. Journal of Clinical Sleep Medicine, 2, 12–8.

    Google Scholar 

  • Pinotsis, D.A., Moran, R.J., & Friston, K.J. (2012). Dynamic causal modeling with neural fields. NeuroImage, 59(2), 1261–1274.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Purdon, P.L., Pierce, E.T., Mukamel, E.A., Prerau, M.J., Walsh, J.L., Wong, K.F., Salazar-Gomez, A.F., Harrell, P.G., Sampson, A.L., Cimenser, A., Ching, S., Kopell, N.J., Tavares-Stoeckel, C., Habeeb, K., Merhar, R., & Brown, E.N. (2013). Electroencephalogram signatures of loss and recovery of consciousness from propofol. Proceedings of the National Academy of Sciences of the United States of America, 110 (12), 1142–1151.

    Article  Google Scholar 

  • Reed, M.D., Yamashita, T.S., Marx, C.M., Myers, C.M., & Blumer, J.L. (1996). A pharmacokinetically based propofol dosing strategy for sedation of the critically ill, mechanically ventilated pediatric patient. Critical Care Medicine, 24(9), 1473–1481.

    Article  CAS  PubMed  Google Scholar 

  • Rennie, C.J., Robinson, P.A., & Wright, J.J. (2002). Unified neurophysical model of EEG spectra and evoked potentials. Biological Cybernetics, 86, 457–471.

    Article  CAS  PubMed  Google Scholar 

  • Robinson, P.A. (2003). Neurophysical theory of coherence and correlations of electroencephalographic signals. Journal of Theoretical Biology, 222, 163–175.

    Article  CAS  PubMed  Google Scholar 

  • Robinson, P.A., Loxley, P.N., O’Connor, S.C., & Rennie, C.J. (2001b). Modal analysis of corticothalamic dynamics, electroencephalographic spectra, and evoked potentials. Physical Review E, 63, 041909.

    Article  CAS  Google Scholar 

  • Robinson, P.A., Rennie, C.J., & Rowe, D.L. (2002). Dynamics of large-scale brain activity in normal arousal states and eplieptic seizures. Physical Review E, 65(4), 041924.

    Article  CAS  Google Scholar 

  • Robinson, P.A., Rennie, C.J., & Wright, J.J. (1997). Propagation and stability of waves of electrical activity in the cerebral cortex. Physical Review E, 56, 826–840.

    Article  CAS  Google Scholar 

  • Robinson, P.A., Wright, J.J., & Rennie, C.J. (1998). Synchronous oscillations in the cerebral cortex. Physical Review E, 57, 4578–4588.

    Article  CAS  Google Scholar 

  • Robinson, P.A., Rennie, C.J., Wright, J.J., Bahramali, H., Gordon, E., & Rowe, D. (2001a). Prediction of electroencephalographic spectra from neurophysiology. Physical Review E, 63, 201903.

    Google Scholar 

  • Robinson, P.A., Rennie, C.J., Rowe, D.L., & O’Connor, S.C. (2004). Estimation of multiscale neurophysiologic parameters by electroencephalographic means. Human Brain Mapping, 23, 53–72.

    Article  CAS  PubMed  Google Scholar 

  • Rowe, D.L, Robinson, P.A., & Rennie, C.J. (2004). Estimation of neurophysiological parameters from the waking EEG using a biophysical model of brain dynamics. Journal of Theoretical Biology, 231(3), 413–433.

    Article  PubMed  Google Scholar 

  • Rudolph, U., & Antkowiak, B. (2004). Molecular and neuronal substrates for general anaesthetics. Nature Reviews Neuroscience, 5, 709–720.

    Article  CAS  PubMed  Google Scholar 

  • San-Juan, D., Chiappa, K.H., & Cole, A.J. (2010). Propofol and the electroencephalogram. Clinical Neurophysiology, 121(7), 998–1006.

    Article  CAS  PubMed  Google Scholar 

  • Sellers, K.K., Bennett, D.V., Hutt, A., & Frohlich, F. (2013). Anesthesia differentially modulates spontaneous network dynamics by cortical area and layer. Journal of Neurophysiology. in press.

  • Semyanov, A., Walker, M.C., Kullmann, D.M., & Silver, R.A. (2004). Tonically active GABAA receptors: modulating gain and maintaining the tone. Trends in Neurosciences, 27(5), 262–269.

  • Siegel, J.M. (2009). Sleep viewed as a state of adaptive inactivity. Nature. Reviews in the Neurosciences, 10, 747–753.

    CAS  Google Scholar 

  • Spiegler, A., Kiebel, S.J., Atay, F.M., & Knosche, T.R. (2010). Bifurcation analysis of neural mass models: Impact of extrinsic inputs and dendritic time constants. NeuroImage, 52(3), 1041–1058.

    Article  PubMed  Google Scholar 

  • Spiegler, A., Knosche, T.R., Schwab, K., Haueisen, J., & Atay, F.M. (2011). Modeling brain resonance phenomena using a neural mass model. PLoS Computational Biology, 7(12), 1002298.

    Article  Google Scholar 

  • Steriade, M., Contreras, D., Curro Dossi, R., & Nunez, A. (1993). The slow (< 1 hz) oscillation in reticular thalamic and thalamocortical neurons: scenario of sleep rhythm generation in interacting thalamic and neocortical networks. The Journal of Neuroscience, 13(8), 3284–3299.

    CAS  PubMed  Google Scholar 

  • Steyn-Ross, M.L., Steyn-Ross, D.A., & Sleigh, J.W. (2004). Modelling general anaesthesia as a first-order phase transition in the cortex. Progress in Biophysics and Molecular Biology, 85(2-3), 369– 385.

    Article  CAS  PubMed  Google Scholar 

  • Steyn-Ross, M.L., Steyn-Ross, D.A., Sleigh, J.W., & Liley, D.T.J. (1999). Theoretical electroencephalogram stationary spectrum for a white-noise-driven cortex: Evidence for a general anesthetic-induced phase transition. Physical Review E, 60(6), 7299–7311.

    Article  CAS  Google Scholar 

  • Steyn-Ross, M.L., Steyn-Ross, D.A., Sleigh, J.W., & Wilcocks, L.C. (2001a). Toward a theory of the general-anesthetic-induced phase transition of the cerebral cortex: I. a thermodynamic analogy. Physical Review E, 64, 011917.

    Article  CAS  Google Scholar 

  • Steyn-Ross, M.L., Steyn-Ross, D.A., Sleigh, J.W., & Wilcocks, L.C. (2001b). Toward a theory of the general-anesthetic-induced phase transition of the cerebral cortex: Ii. numerical simulations, spectra entropy, and correlation times. Physical Review E, 64, 011918.

    Article  CAS  Google Scholar 

  • Vanini, G., & Baghdoyan, H.A. (2013). Extrasynaptic GABA A receptors in rat pontine reticular formation increase wakefulness. Sleep, 36(3), 337–343.

    PubMed Central  PubMed  Google Scholar 

  • Victor, J.D., Drover, J.D., Conte, M.M., & Schiff, N.D. (2011). Mean-field modeling of thalamocortical dynamics and a model-driven approach to EEG analysis. Proceedings of the National Academy of Sciences of the United States of America, 108, 15631– 15638.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Vijayan, S., Ching, S., Cimenser, A., Purdon, P.L., Brown, E.N., & Kopell, N.J. (2013). Thalamocortical mechanisms for the anteriorization of alpha rhythms during propofol-induced unconsciousness. The Journal of Neuroscience, 33(27), 11070–11075.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Wieloch, T., & Nikolich, K. (2006). Mechanisms of neural plasticity following brain injury. Current Opinion in Neurobiology, 16(3), 258–264.

    Article  CAS  PubMed  Google Scholar 

  • Wilson, H.R., & Cowan, J.D. (1972). Excitatory and inhibitory interactions in localized populations of model neurons. Biophysical Journal, 12, 1–24.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  • Wilson, M.T., Sleigh, J.W., Steyn-Ross, A., & Steyn-Ross, M.L. (2006). General anesthetic-induced seizures can be explained by a mean-field model of cortical dynamics. Anesthesiology, 104(3), 588–593.

    Article  PubMed  Google Scholar 

  • Wright, J.J., & Liley, D.T.J. (1996). Dynamics of the brain at global and microscopic scales: Neural networks and the EEG. Behavioral and Brain Sciences, 19, 285–320.

    Article  Google Scholar 

  • Ying, S., & Goldstein, P. (2001). Propofol effects on the thalamus: Modulation of GABAergic synaptic inhibition and suppression of neuronal excitability. Abstract Viewer/Itinerary Planner Washington. DC: Society for Neuroscience, 89(411).

  • Ying, S.W., & Goldstein, P.A. (2005a). Propofol-block of SK channels in reticular thalamic neurons enhances GABAergic inhibition in relay neurons. Journal of Neurophysiology, 93, 1935–1948.

  • Ying, S.W., & Goldstein, P.A. (2005b). Propofol suppresses synaptic responsiveness of somatosensory relay neurons to excitatory input by potentiating GABA A receptor chloride channels. Molecular Pain, 1, 2.

    Article  PubMed Central  PubMed  Google Scholar 

  • Zhang, M., Wei, G.W., & Wei, C.H. (2002). Transition from intermittency to periodicity in lag synchronization in coupled roessler oscillators. Physical Review E, 65, 036202.

  • Zhou, C., Liu, J., & Chen, X.D. (2012). General anesthesia mediated by effects on ion channels. World Journal of Critical Care Medicine, 1, 80–93.

    Article  PubMed Central  PubMed  Google Scholar 

Download references

Acknowledgments

The authors acknowledge funding from the European Research Council for support under the European Union’s Seventh Framework Programme (FP7/2007-2013)/ERC grant agreement no. 257253.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Meysam Hashemi.

Additional information

Conflict of interests

The authors declare that they have no conflict of interest.

Appendices

Appendix A: Stationary states

Under the assumption of the spatial homogeneity, the mean excitatory and inhibitory postsynaptic potentials in the cortical pyramidal neurons (E), cortical inhibitory neurons (I), thalamo-cortical relay neurons (S) and thalamic reticular neurons (R) shown in Fig. 1 obey

$$\begin{array}{@{}rcl@{}} \hat {L}_{e} {V^{e}_{E}}(t)&=& K_{EE}S_{C}\left[{V^{e}_{E}}(t)-{V^{i}_{E}}(t)\right]\\ &&+ K_{ES}S_{T}\left[{V^{e}_{S}}(t-\tau)-{V^{i}_{S}}(t-\tau)\right],\\ \hat{L}_{i} {V^{i}_{E}}(t)&=& f_{C}(p) K_{EI} S_{C}\left[{V^{e}_{I}}(t)-{V^{i}_{I}}(t)\right] ,\\ \hat {L}_{e} {V^{e}_{I}}(t)&=& K_{IE}S_{C}\left[{V^{e}_{E}}(t)-{V^{i}_{E}}(t)\right] ,\\ \hat {L}_{i} {V^{i}_{I}}(t)&=& K_{II}S_{C}\left[{V^{e}_{I}}(t)-{V^{i}_{I}}(t)\right] ,\\ \hat{L}_{e} {V^{e}_{S}}(t)&=& K_{SE}S_{C}\left[{V^{e}_{E}}(t-\tau)-{V^{i}_{E}}(t-\tau)\right]+I(t) ,\\ \hat{L}_{i} {V^{i}_{S}}(t)&=& f_{T}(p) K_{SR} S_{T}\left[{V^{e}_{R}}(t)\right] ,\\ \hat{L}_{e} {V^{e}_{R}}(t)&=& K_{RE} S_{C}\left[{V^{e}_{E}}(t-\tau)-{V^{i}_{E}}(t-\tau)\right]\\ &&+K_{RS} S_{T}\left[{V^{e}_{S}}(t)-{V^{i}_{S}}(t)\right], \end{array} $$
(19)

In this model, the long-range propagation of the signals has been considered by the connection between cortex and thalamus associated with a constant time delay and the model follows closely Victor et al. (2011) and Drover et al. (2010) showing that long connections between cortico-cortical populations through white matter are not necessary to describe experimental EEG dynamics. In the case of constant external input I(t)=I 0, the spatially-homogeneous resting states of Eqs. (19) can be obtained by \(dV^{e,i}_{a}/dt=0\), for a∈{E,I,S,R} leading to

$$\begin{array}{@{}rcl@{}} V^{*e}_{E}&=& K_{EE} S_{C}\left[V^{*e}_{E}-V^{*i}_{E}\right]+K_{ES} S_{T}\left[V^{*e}_{S}-V^{*i}_{S}\right],\\ V^{*i}_{E}&=& f_{C}(p)K_{EI}S_{C}\left[V^{*e}_{I}-V^{*i}_{I}\right],\\ V^{*e}_{I}&=& K_{IE} S_{C}\left[V^{*e}_{E}-V^{*i}_{E}\right],\\ V^{*i}_{I}&=& K_{II}S_{C}\left[V^{*e}_{I}-V^{*i}_{I}\right],\\ V^{*e}_{S}&=& K_{SE} S_{C}\left[V^{*e}_{E}-V^{*i}_{E}\right] +I_{0},\\ V^{*i}_{S}&=& f_{T}(p) K_{SR}S_{T}\left[V^{*e}_{R}\right],\\ V^{*e}_{R}&=& {K_{RE} S_{C}\left[V^{*e}_{E}-V^{*i}_{E}\right]+K_{RS} S_{T}\left[V^{*e}_{S}-V^{*i}_{S}\right]}. \end{array} $$
(20)

The dynamic of the reduced model are described by

$$\begin{array}{@{}rcl@{}} \hat{L}_{e}{V^{e}_{E}}(t)&=& K_{ES} S_{T}\left[{V^{e}_{S}}(t-\tau)-{V^{i}_{S}}(t-\tau)\right] ,\\ \hat{L}_{e}{V^{e}_{S}}(t)&= &K_{SE} S_{C}\left[{V^{e}_{E}}(t-\tau)\right]+I(t) ,\\ \hat{L}_{i}{V^{i}_{S}}(t)&=& f_{T}(p) K_{SR} S_{T}\left[{V^{e}_{R}}(t)\right] ,\\ \hat{L}_{e}{V^{e}_{R}}(t)&=& K_{RE} S_{C}\left[{V^{e}_{E}}(t-\tau)\right]\\ &&+K_{RS} S_{T}\left[{V^{e}_{S}}(t)-{V^{i}_{S}}(t)\right], \end{array} $$
(21)

where the resting states of the system obey

$$\begin{array}{@{}rcl@{}} V^{*e}_{E}&=& K_{ES} S_{T}\left[V^{*e}_{S}-V^{*i}_{S}\right],\\ V^{*e}_{S}&=& K_{SE} S_{C}\left[V^{*e}_{E}\right] +I_{0},\\ V^{*i}_{S}&=& f_{T}(p)K_{SR} S_{T}\left[V^{*e}_{R}\right],\\ V^{*e}_{R}&=& {K_{RE} S_{C}\left[V^{*e}_{E}\right]+K_{RS} S_{T}\left[V^{*e}_{S}-V^{*i}_{S}\right]}. \end{array} $$
(22)

By inserting these equations into each other

$$\begin{array}{@{}rcl@{}} V^{*e}_{E}&=&K_{SE} S_{T} \left[ K_{SE} S_{C}\left[V^{*e}_{E}\right]+I_{0} \right.\\ &&\qquad\qquad\left.- f_{T}(p) K_{SR} S_{T}\left[H\left[V^{*e}_{E}\right]\right]\right], \end{array} $$
(23)

where \(H[V^{*e}_{E}] \equiv {K_{RE} S_{C}[V^{*e}_{E}]+\frac {K_{RS}}{K_{ES}} V^{*e}_{E}} \). Since all the resting states \(V^{*e,i}_{a}\) for a∈{E,R,S} can be written as an implicit function of \(V^{*e}_{E}\), the number of solutions of \(V^{*e}_{E}\), i.e., the number of roots of Eq. (23), is identical to the number of resting states (Robinson et al. 1997; Robinson et al. 1998; Robinson et al. 2004).

Appendix B: Theoretical power spectrum

The solution of Eq. (14) for t is

$$ \boldsymbol{Y}(t)={\int}_{-\infty}^{\infty} \boldsymbol{G}(t-t^{\prime}) \boldsymbol{\xi}(t^{\prime})dt^{\prime}, $$
(24)

with the matrix Greens function \(\boldsymbol {G}\in {\mathbb R}^{N\times N}\), that has dimension N = 7. Substituting Eq. (24) into Eq. (14) leads to

$$ \hat{\boldsymbol{L}} (\partial/\partial t,p) \boldsymbol{G}(t) = \boldsymbol {A }(p) \boldsymbol{G}(t)+\boldsymbol {B }(p)\boldsymbol{G}(t-\tau) + \boldsymbol{1} \delta(t) , $$
(25)

with the unitary matrix \(\boldsymbol {1} \in \mathbb {R}^{N \times N}\). Then the Fourier transform of the matrix Greens function

$$ \boldsymbol{\tilde{G}}(\nu,p)=\frac{1}{\sqrt{2 \pi}} \left[{\boldsymbol{L}}(\nu,p) -\boldsymbol{A}(p)-\boldsymbol{B}(p)e^{-2 \pi i\nu \tau} \right]^{-1}, $$
(26)

and the Wiener-Khinchine theorem defines the power spectral density matrix to

$$ \boldsymbol P(\nu)= 2 \kappa \sqrt{2\pi}\boldsymbol{\tilde{G}}(\nu,p) \boldsymbol{\tilde{G}}^{\top} (-\nu,p), $$

where the high index ⊤ denotes the transposed vector or matrix. Essentially we assume that the EEG is generated by the activity of pyramidal cortical cells. By virtue of the specific choice of external input to relay neurons, the power spectrum of the EEG just depends on one matrix component of the Greens function by

$$\begin{array}{@{}rcl@{}} P_{E}(\nu)&=&{2 \kappa} \sqrt{2\pi}{\tilde{G}}_{1,5}(\nu,p) {\tilde{G}}_{1,5}(-\nu,p)\\ &=&{2 \kappa}\sqrt{2\pi} \left\vert {\tilde{G}}_{1,5}(\nu,p)\right\vert^{2}. \end{array} $$
(27)

In detail, we find

$$\begin{array}{@{}rcl@{}} \boldsymbol{Y}(t)\! &=&\! \left( {V^{e}_{E}}(t)-V^{*e}_{E}, {V^{i}_{E}}(t)-V^{*i}_{E},{V^{e}_{I}}(t)-V^{*e}_{I},{V^{i}_{I}}(t)\right.\\ &&\left.\!-V^{*i}_{I},{V^{e}_{S}}(t)-V^{*e}_{S}, {V^{i}_{S}} (t)\! -\! V^{*i}_{S}, {V^{e}_{R}}(t)-V^{*e}_{R}\right)^{\top}, \end{array} $$

and the diagonal matrix \(\hat {\boldsymbol {L}} (\partial /\partial t,p)\) with the entries \(\hat {L}_{1,1} = \hat {L}_{3,3} = \hat {L}_{5,5} = \hat {L}_{7,7}=\hat {L}_{e}(\nu )\), \(\hat {L}_{2,2} =\hat {L}_{4,4}=\hat {L}_{6,6} = \hat {L}_{i}(\nu ,p)\) and

$$\begin{array}{@{}rcl@{}} \boldsymbol{ A}(p)&=& \left( {\begin{array}{ccccccc} K_{1}&-K_{1}&0&0&0&0&0 \\ 0&0&f_{C}(p)K_{3}&-f_{C}(p)K_{3}&0&0&0 \\ K_{4}&-K_{4}&0&0&0&0&0 \\ 0&0&K_{5}&-K_{5}&0&0 &0\\ 0&0&0&0&0&0&0 \\ 0&0&K_{5}&-K_{5}&0&0 & f_{T}(p)K_{7}\\ 0&0&0&0&K_{9}&-K_{9}&0 \end{array}} \right),\\ \boldsymbol{B}(p) &=& \left( {\begin{array}{ccccccc} 0&0&0&0&K_{2}&-K_{2}&0 \\ 0&0&0&0&0&0&0 \\ 0&0&0&0&0&0 &0\\ 0&0&0&0&0&0&0 \\ K_{6}&-K_{6}&0&0&0&0&0 \\ 0&0&0&0&0&0&0 \\ K_{8}&-K_{8}&0&0&0&0 &0 \end{array}} \right) \end{array} $$

with

$$\begin{array}{@{}rcl@{}} \hat {L}_{e}(\nu,p)&=&\left( 1+\frac{2 \pi i \nu}{\alpha_{e}} \right)\left( 1+\frac{2 \pi i \nu}{\beta_{e}} \right),\\ \hat{L}_{i}(\nu,p)&=&\left( 1+\frac{2 \pi i \nu}{\alpha_{i}} \right)\left( 1+\frac{2 \pi i \nu}{\beta_{i}} \right),\\ \end{array} $$
$$\begin{array}{@{}rcl@{}} K_{1}&=&K_{EE}\frac{d S_{C}[V]}{d V} \mid_{V=\left( V^{*e}_{E}-V^{*i}_{E}\right)},\\ K_{2}&=&K_{ES}\frac{d S_{T}[ V]}{d V} \mid_{V=\left( V^{*e}_{S}-V^{*i}_{S}\right)},\\ K_{3}&=&K_{EI}\frac{d S_{C}[V]}{d V} \mid_{V=\left( V^{*e}_{I}-V^{*i}_{I}\right)},\\ K_{4}&=&K_{IE}\frac{d S_{C}[V]}{d V}\mid_{V=\left( V^{*e}_{E}-V^{*i}_{E}\right)},\\ K_{5}&=&K_{II}\frac{d S_{C}[V]}{d V}\mid_{V=\left( V^{*e}_{I}-V^{*i}_{I}\right)},\\ K_{6}&=&K_{SE}\frac{d S_{C}[V]}{d V}\mid_{V=\left( V^{*e}_{E}-V^{*i}_{E}\right)},\\ K_{7}&=&K_{SR} \frac{d S_{T}[V]}{d V} \mid_{V=V^{*e}_{R}},\\ K_{8}&=&K_{RE} \frac{d S_{C}[ V]}{d V}\mid_{V=\left( V^{*e}_{E}-V^{*i}_{E}\right)},\\ K_{9}&=&K_{RS}\frac{d S_{T}[V]}{d V} \mid_{V=\left( V^{*e}_{S}-V^{*i}_{S}\right)}. \end{array} $$

The constants K i , i=1,…,9 depend on the anaesthetic factor p since they are evaluated at the resting state depending on p itself. Hence

$$\begin{array}{@{}rcl@{}} \boldsymbol{\tilde{G}}(\nu,p)=\frac{1}{\sqrt{2 \pi}} \left[\begin{array}{ccccccc} \hat {L}_{e}(\nu)-K_{1}&K_{1}&0&0&-K_{2} e^{-2 \pi i\nu\tau}&K_{2} e^{-2 \pi i\nu\tau}&0 \\ 0&\hat {L}_{i}(\nu)&-f_{C}(p)K_{3}&f_{C}(p)K_{3}&0&0&0 \\ -K_{4}&K_{4}&\hat {L}_{e}(\nu)&0&0&0&0 \\ 0&0&-K_{5}&\hat {L}_{i}+K_{5}(\nu)&0&0&0\\ -K_{6}e^{-2 \pi i\nu\tau}&K_{6}e^{-2 \pi i\nu\tau}&0&0 &\hat {L}_{e}(\nu)&0&0\\ 0& 0&0& 0&0&\hat {L}_{i} (\nu,p)&-f_{T}(p)K_{7}\\ -K_{8}e^{-2 \pi i\nu\tau}&K_{8}e^{-2 \pi i\nu\tau}&0&0&-K_{9} &K_{9}&\hat {L}_{e}(\nu) \end{array}\right]^{-1} \end{array} $$
(28)

and finally

$$ P_{E}(\nu)={2 \kappa}\sqrt{2\pi} \left\vert {\tilde{G}}_{1,5}(\nu,p)\right\vert^{2}. $$
(29)

Appendix C: Contribution to power spectrum

Here we consider the reduced model and parametrize the contribution of PSPs to the power spectrum in the δ− and α− frequency bands. Substituting the ansatz Y(t)=e λt u with eigenfunction \(\mathbf {u}=[ {u^{e}_{E}}, {u^{e}_{S}}, {u^{i}_{S}}, {u^{e}_{R}}]^{\top } \) into Eq. (14) yields

$$\begin{array}{@{}rcl@{}} \left (\frac{\lambda}{\alpha_{e} }\! +\! 1 \right )\! \left (\frac{\lambda}{\beta_{e} } +1 \right )\! {u^{e}_{E}}\! &=&\! K_{2} ({u^{e}_{S}}-{u^{i}_{S}}) e^{-\lambda \tau}, \\ \left (\frac{\lambda}{\alpha_{e} }\! +\! 1 \right )\! \left (\frac{\lambda}{\beta_{e} } +1 \right )\! {u^{e}_{S}}\! &=&\! K_{6} {u^{e}_{E}} e^{-\lambda \tau} , \\ \left (\frac{\lambda}{\alpha_{i} }\! +\! 1 \right )\! \left (\frac{\lambda}{\beta_{i}} +1 \right )\! {u^{i}_{S}}\! &=&\! f_{T}(p) K_{7} {u^{e}_{R}}, \\ \left (\frac{\lambda}{\alpha_{e} }\! +\! 1 \right )\! \left (\frac{\lambda}{\beta_{e} }\! +\! 1 \right )\! {u^{e}_{R}}\! &=&\! K_{8} {u^{e}_{E}} e^{-\lambda \tau}\! +\! K_{9} \left( {u^{e}_{S}}\! -\! {u^{i}_{S}}\right). \end{array} $$
(30)

Now we can write all the elements of eigenfunction u in terms of the first element \({u^{e}_{E}}\) as follows

$$\begin{array}{@{}rcl@{}} {u^{e}_{E}} &=&u_{1}(\lambda) {u^{e}_{E}}, \\ {u^{e}_{S}} &=& \left( \frac {K_{6} e^{- \lambda \tau}} {\left( 1+ \frac{\lambda}{\alpha_{e}}\right)\left( 1+ \frac{\lambda}{\beta_{e}}\right)} \right) {u^{e}_{E}} \equiv u_{2}(\lambda) {u^{e}_{E}}, \\ {u^{i}_{S}} &=& \left( \frac{K_{6} e^{- \lambda \tau}} {\left( 1+ \frac{\lambda}{\alpha_{e}}\right)\left( 1+ \frac{\lambda}{\beta_{e}}\right)} - \frac{\left( 1+ \frac{\lambda}{\alpha_{e}}\right)\left( 1+ \frac{\lambda}{\beta_{e}}\right)} {K_{2} e^{- \lambda \tau}} \right)\\ {u^{e}_{E}} &\equiv& u_{3}(\lambda) {u^{e}_{E}}, \\ {u^{e}_{R}} &=& \frac{\left( 1+ \frac{\lambda}{\alpha_{i}}\right)\left( 1+ \frac{\lambda}{\beta_{i}}\right)} {f_{T}(p) K_{7} } \left( \frac{K_{6} e^{- \lambda \tau}} {\left( 1+ \frac{\lambda}{\alpha_{e}}\right)\left( 1+ \frac{\lambda}{\beta_{e}}\right)}\right.\\ &&\left. - \frac{\left( 1+\frac{\lambda}{\alpha_{e}}\right)\left( 1+ \frac{\lambda}{\beta_{e}}\right)} {K_{2} e^{- \lambda \tau}} \right) {u^{e}_{E}} \equiv u_{4}(\lambda) {u^{e}_{E}}. \end{array} $$
(31)

Then the associated normalized eigenfunction with known eigenvalue λ is \(\hat {\mathbf {u}} = C\left (1,u_{2}(\lambda ),u_{3}(\lambda ),u_{4}(\lambda )\right )^{\top }\), where \(C=\frac {1}{\sqrt {1+{u^{2}_{2}}+{u^{2}_{3}}+{u^{2}_{4}}}}\). All elements of vector Y(t)=[y 1(t),y 2(t),y 3(t),y 4(t)] can be written as \(y_{n}(t)=\hat u_{n} e^{\lambda t}+ \hat {u}_{n}^{\ast } e^{\lambda ^{\ast } t}\) for n=1,...,4. Here the superscript ∗ denotes the complex conjugate. Let λ=γ+2π i ν and \(\hat u_{n} =R_{n}+i I_{n}\), then y n (t) becomes

$$\begin{array}{@{}rcl@{}} y_{n}(t) &=& (R_{n}+iI_{n}) e^{(\gamma+2\pi i \nu)t}+(R_{n}-iI_{n}) e^{(\gamma-2\pi i \nu)t}\\ &=&2 e^{\gamma t}(R_{n}\cos(2\pi \nu t) -I_{n} \sin(2\pi \nu t)), \end{array} $$
(32)

and the contribution of excitatory and inhibitory currents to power in a certain frequency ν, in different populations can be defined by Eq. (18).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hashemi, M., Hutt, A. & Sleigh, J. How the cortico-thalamic feedback affects the EEG power spectrum over frontal and occipital regions during propofol-induced sedation. J Comput Neurosci 39, 155–179 (2015). https://doi.org/10.1007/s10827-015-0569-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-015-0569-1

Keywords

Navigation