Skip to main content

Equilibrium and Nonequilibrium Phase Transitions in a Continuum Model of an Anesthetized Cortex

  • Chapter
  • First Online:
Book cover Neural Fields

Abstract

In this chapter we investigate a range of dynamic behaviors accessible to a continuum model of the cerebral cortex placed close to the anesthetic phase transition. If the anesthetic transition from the high-firing (conscious) to the low-firing (comatose) state can be modeled as a jump between two equilibrium states of the cortex, then we can draw an analogy with the vapor-to-liquid phase transition of the van der Waals gas of classical thermodynamics. In this analogy, specific volume (inverse density) of the gas maps to cortical activity, with pressure and temperature being the analogs of anesthetic concentration and subcortical excitation. It is well known that at the thermodynamic critical point, large fluctuations in specific volume are observed; we find analogous critically-slowed fluctuations in cortical activity at its critical point. Unlike the van der Waals system, the cortical model can also exhibit nonequilibrium phase transitions in which the homogeneous equilibrium can destabilize in favor of slow global oscillations (Hopf temporal instability), stationary structures (Turing spatial instability), and chaotic spatiotemporal activity patterns (Hopf–Turing interactions). We comment on possible physiological and pathological interpretations for these dynamics. In particular, the turbulent state may correspond to the cortical slow oscillation between “up” and “down” states observed in nonREM sleep and clinical anesthesia.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Of course, the (specific volume) \(\equiv \) (firing rate) analogy is not perfect: the volume of a gas can increase without limit, but cortical firing rate is limited by biological constraints, implemented in the model by imposing a maximum firing rate \(Q_{e}^{\text{max}}\) (see Table 15.1).

  2. 2.

    See http://en.wikipedia.org/wiki/Maxwell_construction

References

  1. Alkire, M.T., Hudetz, A.G., Tononi, G.: Consciousness and anesthesia. Science 322(5903), 876–880 (2008)

    Google Scholar 

  2. Bazhenov, M., Timofeev, I., Steriade, M., Sejnowski, T.J.: Model of thalamocortical slow-wave sleep oscillations and transitions to activated states. J. Neurosci. 22(19), 8691–8704 (2002)

    Google Scholar 

  3. Bennett, M.V., Zukin, R.S.: Electrical coupling and neuronal synchronization in the mammalian brain. Neuron 41, 495–511 (2004)

    Article  Google Scholar 

  4. Cartwright, J.H.E.: Labyrinthine turing pattern formation in the cerebral cortex. J. Theor. Biol. 217(1), 97–103 (2002)

    Article  MathSciNet  Google Scholar 

  5. Casali, A.G., Gosseries, O., Rosanova, M., Boly, M., Sarasso, S., Casali, K.R., Casarotto, S., Bruno, M.A., Laureys, S., Tononi, G., Massimini, M.: A theoretically based index of consciousness independent of sensory processing and behavior. Sci. Transl. Med. 5(198), 198ra105 (2013)

    Google Scholar 

  6. Dadok, V., Szeri, A.J., Sleigh, J.W.: A probabilistic framework for a physiological representation of dynamically evolving sleep state. J. Comput. Neurosci. (2013). doi:10.1007/s10827-013-0489-x

    Google Scholar 

  7. Franks, N.P., Lieb, W.R.: Molecular and cellular mechanisms of general anaesthesia. Nature 367, 607–613 (1994)

    Article  Google Scholar 

  8. Friedman, E.B., Sun, Y., Moore, J.T., Hung, H.T., Meng, Q.C., Perera, P., Joiner, W.J., Thomas, S.A., Eckenhoff, R.G., Sehgal, A., Kelz, M.B.: A conserved behavioral state barrier impedes transitions between anesthetic-induced unconsciousness and wakefulness: evidence for neural inertia. PLoS One 5(7), e11,903 (2010)

    Google Scholar 

  9. Gajda, Z., Gyengesi, E., Hermesz, E., Ali, K.S., Szente, M.: Involvement of gap junctions in the manifestation and control of the duration of seizures in rats in vivo. Epilepsia 44(12), 1596–1600 (2003)

    Article  Google Scholar 

  10. Guldenmund, P., Demertzi, A., Boveroux, P., Boly, M., Vanhaudenhuyse, A., Bruno, M.A., Gosseries, O., Noirhomme, Q., Brichant, J.F., Bonhomme, V., Laureys, S., Soddu, A.: Thalamus, brainstem and salience network connectivity changes during propofol-induced sedation and unconsciousness. Brain Connect 3(3), 273–285 (2013)

    Article  Google Scholar 

  11. Hutt, A., Longtin, A.: Effects of the anesthetic agent propofol on neural populations. Cog. Neurodyn. 4(1), 37–59 (2010)

    Article  Google Scholar 

  12. Jacobson, G.M., Voss, L.J., Melin, S.M., Mason, J.P., Cursons, R.T., Steyn-Ross, D.A., Steyn-Ross, M.L., Sleigh, J.W.: Connexin36 knockout mice display increased sensitivity to pentylenetetrazol-induced seizure-like behaviors. Brain Res. 1360, 198–204 (2010)

    Article  Google Scholar 

  13. Jahromi, S.S., Wentlandt, K., Piran, S., Carlen, P.L.: Anticonvulsant actions of gap junctional blockers in an in vitro seizure model. J. Neurophysiol. 88(4), 1893–1902 (2002)

    Google Scholar 

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

    Google Scholar 

  15. Kramer, M.A., Kirsch, H.E., Szeri, A.J.: Pathological pattern formation and cortical propagation of epileptic seizures. J. R. Soc. Lond.: Interface 2, 113–207 (2005)

    Google Scholar 

  16. Kuizenga, K., Kalkman, C.J., Hennis, P.J.: Quantitative electroencephalographic analysis of the biphasic concentration–effect relationship of propofol in surgical patients during extradural analgesia. Br. J. Anaesth. 80, 725–732 (1998)

    Article  Google Scholar 

  17. Kuizenga, K., Wierda, J.M.K.H., Kalkman, C.J.: Biphasic EEG changes in relation to loss of consciousness during induction with thiopental, propofol, etomidate, midazolam or sevoflurane. Br. J. Anaesth. 86, 354–360 (2001)

    Article  Google Scholar 

  18. Lee, U., Ku, S., Noh, G., Baek, S., Choi, B., Mashour, G.A.: Disruption of frontal-parietal communication by ketamine, propofol, and sevoflurane. Anesthesiology 118(6), 1264–1275 (2013)

    Article  Google Scholar 

  19. Liley, D.T.J., Bojak, I.: Understanding the transition to seizure by modeling the epileptiform activity of general anesthetic agents. Clin. Neurophysiol. 22(5), 300–313 (2005)

    Google Scholar 

  20. 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.: Propofol anesthesia and sleep: a high-density EEG study. Sleep 34(3), 283–291A (2011)

    Google Scholar 

  21. Nilsen, K.E., Kelso, A.R., Cock, H.R.: Antiepileptic effect of gap-junction blockers in a rat model of refractory focal cortical epilepsy. Epilepsia 47(7), 1169–1175 (2006)

    Article  Google Scholar 

  22. Robinson, P.A., Rennie, C.J., Wright, J.J.: Propagation and stability of waves of electrical activity in the cerebral cortex. Phys. Rev. E 56, 826–840 (1997)

    Article  Google Scholar 

  23. Sears, F.W., Salinger, G.L.: Thermodynamics, Kinetic Theory, and Statistical Thermodynamics, 3rd edn. Addison-Wesley, Reading (1975)

    Google Scholar 

  24. Stanley, H.E.: Introduction to Phase Transitions and Critical Phenomena. Clarendon Press, Oxford (1971)

    Google Scholar 

  25. Steriade, M., Nuñez, A., Amzica, F.: A novel slow (\(<\) 1 Hz) oscillation of neocortical neurons in vivo: depolarizing and hyperpolarizing components. J. Neurosci. 13, 3252–3265 (1993)

    Google Scholar 

  26. Steyn-Ross, M.L., Steyn-Ross, D.A., Sleigh, J.W., Liley, D.T.J.: Theoretical electroencephalogram stationary spectrum for a white-noise-driven cortex: evidence for a general anesthetic-induced phase transition. Phys. Rev. E 60, 7299–7311 (1999)

    Article  Google Scholar 

  27. Steyn-Ross, M.L., Steyn-Ross, D.A., Sleigh, J.W., Wilcocks, L.C.: Toward a theory of the general anesthetic-induced phase transition of the cerebral cortex: I. A statistical mechanics analogy. Phys. Rev. E 64, 011,917 (2001)

    Google Scholar 

  28. Steyn-Ross, M.L., Steyn-Ross, D.A., Sleigh, J.W.: Modelling general anaesthesia as a first-order phase transition in the cortex. Prog. Biophys. Mol. Biol. 85(2–3), 369–385 (2004)

    Article  Google Scholar 

  29. Steyn-Ross, M.L., Steyn-Ross, D.A., Wilson, M.T., Sleigh, J.W.: Gap junctions mediate large-scale turing structures in a mean-field cortex driven by subcortical noise. Phys. Rev. E 76, 011,916 (2007)

    Google Scholar 

  30. Steyn-Ross, D.A., Steyn-Ross, M.L., Wilson, M.T., Sleigh, J.W.: Phase transitions in single neurons and neural populations: Critical slowing, anesthesia, and sleep cycles. In: Steyn-Ross, D.A., Steyn-Ross, M.L. (eds.) Modeling Phase Transitions in the Brain. Springer Series in Computational Neuroscience, vol. 4, chap. 1, pp. 1–26. Springer, New York (2010)

    Google Scholar 

  31. Steyn-Ross, D.A., Steyn-Ross, M.L., Sleigh, J.W., Wilson, M.T.: Progress in modeling EEG effects of general anesthesia: Biphasic response and hysteresis. In: Hutt, A. (ed.) Sleep and Anesthesia: Neural Correlates in Theory and Experiment. Springer Series in Computational Neuroscience, vol. 15, chap. 8, pp. 167–194. Springer, New York (2011)

    Chapter  Google Scholar 

  32. Voss, L.J., Jacobson, G., Sleigh, J.W., Steyn-Ross, D.A., Steyn-Ross, M.L.: Excitatory effects of gap junction blockers on cerebral cortex seizure-like activity in rats and mice. Epilepsia 50(8), 1971–1978 (2009)

    Article  Google Scholar 

  33. Wentlandt, K., Samoilova, M., Carlen, P.L., El Beheiry, H.: General anesthetics inhibit gap junction communication in cultured organotypic hippocampal slices. Anesth. Analg. 102(6), 1692–1698 (2006)

    Article  Google Scholar 

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

    Article  Google Scholar 

  35. Wozny, C., Williams, S.R.: Specificity of synaptic connectivity between layer 1 inhibitory interneurons and layer 2/3 pyramidal neurons in the rat neocortex. Cereb. Cortex 21(8), 1818–1826 (2011)

    Article  Google Scholar 

  36. Yang, L., Ling, D.S.F.: Carbenoxolone modifies spontaneous inhibitory and excitatory synaptic transmission in rat somatosensory cortex. Neurosci. Lett. 416, 221–226 (2007)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to D. Alistair Steyn-Ross .

Editor information

Editors and Affiliations

Appendix

Appendix

1.1 Model Equations

The cortex is modeled as a 2-D continuum of excitatory and inhibitory neurons, interconnected via resistive gap junctions and neurotransmitter -mediated chemical synapses . The spatially-averaged excitatory (inhibitory) soma potentials V e (V i ) obey partial differential equations,

$$\displaystyle\begin{array}{rcl} \tau _{e}\,\frac{\partial V _{e}} {\partial t} \: =\: V _{e}^{\text{rest}} + \Delta V _{ e}^{\text{rest}} - V _{ e} + \left [\rho _{e}\,\psi _{\mathit{ee}}\varPhi _{\mathit{ee}} +\rho _{i}\,\psi _{\mathit{ie}}\varPhi _{\mathit{ie}}\right ] + D_{1}\nabla ^{2}V _{ e}\,,& &{}\end{array}$$
(15.3)
$$\displaystyle\begin{array}{rcl} \tau _{i}\,\frac{\partial V _{i}} {\partial t} \: =\: V _{i}^{\text{rest}} - V _{ i} + \left [\rho _{e}\,\psi _{\mathit{ei}}\varPhi _{\mathit{ei}} +\rho _{i}\,\psi _{\mathit{ii}}\varPhi _{\mathit{ii}}\right ] + D_{2}\nabla ^{2}V _{ i}\,,& &{}\end{array}$$
(15.4)

with chemical-synaptic inputs [] enclosed in square brackets, and gap-junction inputs entering as diffusion terms \(D\nabla ^{2}V _{e,i}\). Here, τ b (b = e, i) is the soma time-constant; \(V _{b}^{\text{rest}}\) is the soma resting voltage; ρ b is the chemical synaptic strength with ρ e  > 0 (EPSP) and ρ i  < 0 (IPSP). These strengths are scaled by dimensionless reversal-potential functions ψ ab (a = e, i),

$$\displaystyle\begin{array}{rcl} \psi _{ab}(t)\: =\: \frac{V _{a}^{\text{rev}} - V _{b}(t)} {V _{a}^{\text{rev}} - V _{b}^{\text{rest}}}\,,& & {}\\ \end{array}$$

that are normalized to unity when the neuron is at its resting voltage, and are zero when the membrane voltage reaches the relevant reversal potential (see Table 15.1 for values). The Φ eb, ib functions in Eqs. (15.315.4) are chemical-synaptic input fluxes obeying second-order differential equations,

$$\displaystyle\begin{array}{rcl} \left ( \frac{d} {\mathit{dt}} +\gamma _{e}\right )^{2}\varPhi _{ \mathit{eb}}\: =\:\gamma _{ e}^{2}\big[N_{\mathit{ eb}}^{\alpha }\,\phi _{ \mathit{eb}}^{\alpha }(t) + N_{\mathit{ eb}}^{\beta }\,Q_{ e}(t) +\phi _{ \mathit{eb}}^{\text{sc}}(t)\big],& &{}\end{array}$$
(15.5)
$$\displaystyle\begin{array}{rcl} \left ( \frac{d} {\mathit{dt}} +\gamma _{i}\right )^{2}\varPhi _{ \mathit{ib}}\: =\:\gamma _{ i}^{2}N_{\mathit{ ib}}^{\beta }\,Q_{ i}(t),& &{}\end{array}$$
(15.6)

with dendritic rate constants γ e, i . The cortico-cortical and local connectivities N α and N β scale their respective incoming fluxes ϕ α, Q e, i respectively; these fluxes are supplemented by an unstructured subcortical stimulation \(\phi _{}^{\text{sc}}\) modeled as a small white-noise variation ξ(t) about a constant tone \(\langle \phi _{}^{\text{sc}}\rangle\),

$$\displaystyle\begin{array}{rcl} \phi _{\mathit{eb}}^{\text{sc}}(t)\: =\:\langle \phi _{ \mathit{ eb}}^{\text{sc}}\rangle \: +\:\alpha \sqrt{\langle \phi _{\mathit{ eb}}^{\text{sc}}\rangle }\,\xi _{\mathit{eb}}(t),& & {}\\ \end{array}$$

where α is a dimensionless noise-amplitude scale-factor. The local fluxes Q e, i in Eqs. (15.515.6) are defined by a sigmoidal mapping from soma voltage to firing rate,

$$\displaystyle\begin{array}{rcl} Q_{e,i}(t)\: =\: \frac{Q_{e,i}^{\max }} {1 +\exp \left [-C\left (V _{e,i}(t) -\theta _{e,i}\right )/\sigma _{e,i}\right ]}\,,& & {}\\ \end{array}$$

with \(C =\pi /\sqrt{3}\). Here, θ is the population-average threshold for firing, σ is its standard deviation, and Q max is the maximum firing rate.

The cortico-cortical flux ϕ α in Eq. (15.5) is generated by excitatory sources \(Q_{e}(\mathbf{r},t)\), and obeys a 2-D damped wave equation [22],

$$\displaystyle\begin{array}{rcl} \left [\left ( \frac{\partial } {\partial t} + v\varLambda _{\mathit{eb}}\right )^{2} -\, v^{2}\nabla ^{2}\right ]\phi _{\mathit{ eb}}^{\alpha }(\mathbf{r},t)\: =\: (v\,\varLambda _{\mathit{ eb}})^{2}\,Q_{ e}(\mathbf{r},t)\,,& &{}\end{array}$$
(15.7)

where Λ eb is the inverse-length scale for e → b axonal connections, and v is the axonal conduction speed.

The ∇2 diffusion terms in Eqs. (15.315.4) describe the voltage contribution arising from gap-junction currents between adjacent neurons. Gap-junction i-to-i coupling between inhibitory interneurons is substantially more abundant than e-to-e coupling between excitatory neurons [3], and in layer-1 of cortex, over 90 % of the neural density is inhibitory [35], suggesting the existence of a syncytium of interneuron-to-interneuron diffusive scaffolding that spans the cortex . In view of the relative dominance of i-to-i diffusion , we set the e-to-e diffusion strength D 1 to be small fraction of inhibitory diffusion D 2 (viz. \(D_{1} = D_{2}/100\)), with D 2 being an adjustable parameter (see [29] for detailed derivation and estimation of D 2 diffusive coupling).

1.2 Modeling Propofol Anesthesia

Inductive anesthetic agents, such as propofol , suppress neural activity by prolonging the opening of GABA (gamma−aminobutyric acid) channels on the postsynaptic neuron [7], allowing increased influx of chloride ions leading to hyperpolarization. We model propofol effect by simultaneously scaling both the inhibitory synaptic strength ρ i (in Eqs. (15.315.4)) and the dendritic rate-constant γ i (Eq. (15.6)) by a dimensionless scale-factor λ that is set to unity in the absence of propofol, and which grows proportionately to propofol concentration,

$$\displaystyle\begin{array}{rcl} \rho _{i} \rightarrow \lambda \rho _{i}^{0},\qquad \gamma _{ i} \rightarrow \gamma _{i}^{0}/\lambda & & {}\\ \end{array}$$

where γ i 0 and ρ i 0 are the anesthetic-free default values. This scaling prolongs the duration of the inhibitory postsynaptic potential (IPSP) without altering its peak amplitude [14] so that the area of the IPSP response (representing total charge transfer) increases linearly with drug concentration. We note that at very high propofol concentrations—well above the clinically relevant range—the charge-transfer versus drug-concentration curve shows saturation effects [14], but the assumption of linearity is accurate at low concentrations, and has been used by Hutt and Longtin [11] in their anesthesia modeling.

1.3 Linear Stability Analysis for Homogenous Stationary States

Equations (15.315.7) define the cortical model in terms of two first-order (V e, i soma voltages), and six second-order (\(\varPhi _{\mathit{ee},\mathit{ei}};\,\varPhi _{\mathit{ie},\mathit{ii}};\,\phi _{\mathit{ee},\mathit{ei}}\) firing-rate fluxes) partial differential equations (DEs). If we disable the subcortical noise and take note of the parameter symmetries evident in Table 15.1 (viz., \(N_{\mathit{ee}}^{\alpha } = N_{\mathit{ei}}^{\alpha }\); \(N_{\mathit{ee}}^{\beta } = N_{\mathit{ei}}^{\beta }\); \(N_{\mathit{ie}}^{\beta } = N_{\mathit{ii}}^{\beta }\)), the cortical system reduces to a set of two first-order and three second-order DEs, equivalent to eight first-order equations. We locate the homogenous equilibrium states by eliminating all space- the time-derivatives in differential equations (15.315.7) \((\nabla ^{2} = 0;\partial /\partial t = \partial ^{2}/\partial t^{2} = 0)\), then solving (numerically) the resulting set of nonlinear coupled algebraic equations for the steady-state firing rates (Q e , Q i ) of the excitatory and inhibitory neural populations as a function of anesthetic effect λ and resting potential offset Δ V e rest. The resulting distribution of homogeneous stationary states are displayed in Figs. 15.1a and 15.2a.

We define an eight-variable state vector \(\mathbf{X}\,=\,\left [V _{e},V _{i},\varPhi _{\mathit{eb}},\dot{\varPhi }_{\mathit{eb}},\varPhi _{\mathit{ib}},\dot{\varPhi }_{\mathit{ib}},\phi _{\mathit{eb}},\dot{\phi }_{\mathit{eb}}\right ]^{\text{T}}\) with homogeneous equilibrium value \(\mathbf{X}^{(0)}\). We examine the linear stability of this stationary state by imposing a small spatiotemporal disturbance \(\delta \mathbf{X}\) about \(\mathbf{X}^{(0)}\),

$$\displaystyle\begin{array}{rcl} \mathbf{X}(t,\mathbf{r}) =\mathbf{ X}^{(0)} +\delta \mathbf{ X}(t,\mathbf{r})& & {}\\ \end{array}$$

with \(\delta \mathbf{X}\) being a plane-wave perturbation

$$\displaystyle\begin{array}{rcl} \delta \mathbf{X}(t,\mathbf{r}) =\delta \mathbf{ X}(t)\,e^{i\mathbf{q}\cdot \mathbf{r}} =\delta \mathbf{ X}(0)\,e^{\Lambda t}\,e^{i\mathbf{q}\cdot \mathbf{r}}\,,& & {}\\ \end{array}$$

where \(\mathbf{q}\) is the wavevector with wavenumber \(\vert \mathbf{q}\vert = q\), and \(\Lambda \) is an eigenvalue whose real part gives the growth rate of the \(\delta \mathbf{X}(0)\) initial perturbation: if \(\text{Re}(\Lambda ) > 0\), an instability is predicted. Substituting \(\mathbf{X} =\mathbf{ X}^{(0)} +\delta \mathbf{ X}\) into Eqs. (15.315.7) and retaining only linear terms results in the matrix equation,

$$\displaystyle\begin{array}{rcl} \frac{d} {\mathit{dt}}\delta \mathbf{X} = \mathbf{J}(q)\,\delta \mathbf{X}\,,& & {}\\ \end{array}$$

where J is an 8 × 8 Jacobian matrix in which the ∇2 Laplacians for excitatory and inhibitory diffusion (Eqs. 15.315.4), and wave propagation (Eq. 15.7) appear as − q 2 terms. The eight eigenvalues owned by J describe the linearized dynamics of the homogeneous cortex. For each wavenumber q, we extract and plot the dominant eigenvalue—i.e., that eigenvalue whose real part is most positive (or least negative)—since this describes the most strongly growing (or most long-lived) mode at a given spatial frequency. The resulting \(\Lambda \) vs q dispersion curves are shown in Figs. 15.3 and 15.5.

Although the linear dispersion curve provides valuable guidance regarding the onset of instability (i.e., when the real part of the dominating eigenvalue will approach zero), it cannot predict accurately the new dynamics that will emerge once the homogeneous steady state has lost stability and the nonlinear terms can no longer be ignored. This mismatch between linear dispersion prediction and actual simulation outcome is nicely illustrated in Fig. 15.5b which suggests a zero-frequency instability at zero wavenumber will interact with a low frequency wave instability, while the simulation of Fig. 15.7 shows that the actual outcome is a 1.6-Hz Hopf oscillation at q = 0. Similarly, Fig. 15.5d indicates a zero-frequency instability at q = 0 competing with a stationary Turing, but the Fig. 15.9 simulation shows destabilization in favor of strongly turbulent unsteady interactions between Hopf and Turing instabilities.

Rights and permissions

Reprints and permissions

Copyright information

© 2014 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Steyn-Ross, D.A., Steyn-Ross, M.L., Sleigh, J.W. (2014). Equilibrium and Nonequilibrium Phase Transitions in a Continuum Model of an Anesthetized Cortex. In: Coombes, S., beim Graben, P., Potthast, R., Wright, J. (eds) Neural Fields. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-54593-1_15

Download citation

Publish with us

Policies and ethics