Skip to main content

Advertisement

Log in

Kinetic Models of Brain Activity

  • Published:
Brain Imaging and Behavior Aims and scope Submit manuscript

Abstract

Brain imaging sciences, like neurosciences in general, have predominantly been an empirical endeavour. This paper argues that the maturation of “kinetic models” of large-scale neuronal activity will provide a unifying theory to underpin brain imaging sciences. In particular, this framework will provide a means of unifying data from different imaging modalities, afford a direct link with cognitive theories of brain function, equip researchers with novel data analysis methodologies and underpin a dialogue in which theoretical formalisms are iteratively refined or refuted through empirical studies. Three steps are crucial to this endeavour: (1) The extension of models of spiking neural ensembles (where the states of all neurons are specified) to statistical models of neural “masses” (where only a few moments of the distribution of states are specified); (2) The refinement of “forward models”, such as neurovascular coupling, which map neuronal states to observables; and (3) A theory which links the distribution of neuronal states to cognitive operations, hence informing cognitive neuroscience experiments. We provide illustrative examples of kinetic models of neuronal dynamics at the mesoscopic scale, focusing on the manner by which sensory inputs modify the expression of ongoing “background” activity. The paper concludes with an overview of some of the cutting edge developments in kinetic models of brain activity.

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

Similar content being viewed by others

References

  • Achard, S., Salvador, R., Whitcher, B., Suckling, J., & Bullmore, E. (2006). A resilient, low-frequency, small-world human brain functional network with highly connected association cortical hubs. The Journal of Neuroscience, 26, 63–72. doi:10.1523/JNEUROSCI.3874-05.2006.

    Article  PubMed  CAS  Google Scholar 

  • Amari, S. (1975). Homogeneous nets of neuron-like elements. Biological Cybernetics, 17, 211–220. doi:10.1007/BF00339367.

    Article  PubMed  CAS  Google Scholar 

  • Amari, S. (1977). Dynamics of pattern formation in lateral-inhibition type neural fields. Biological Cybernetics, 27, 77–87. doi:10.1007/BF00337259.

    Article  PubMed  CAS  Google Scholar 

  • Ashwin, P., & Timme, M. (2005). When instability makes sense. Nature, 436, 36–37. doi:10.1038/436036b.

    Article  PubMed  CAS  Google Scholar 

  • Bassett, D. S., Meyer-Lindenberg, A., Achard, S., Duke, T., & Bullmore, E. (2006). Adaptive reconfiguration of fractal small-world human brain functional networks. Proceedings of the National Academy of Sciences of the United States of America, 103, 19219–19220. doi:10.1073/pnas.0606005103.

    Article  Google Scholar 

  • Breakspear, M., & Jirsa, V. (2007). Neuronal dynamics and brain connectivity. In A. R. McIntosh, & V. K. Jirsa (Eds.), Handbook of Brain Connectivity. Berlin: Springer.

    Google Scholar 

  • Breakspear, M., & Stam, K. J. (2005). Dynamics of a neural system with a multiscale architecture. Phil. Trans. R. Soc. B, 360, 1051–1074. doi:10.1098/rstb.2005.1643.

    Article  PubMed  Google Scholar 

  • Breakspear, M., Roberts, J. A., Terry, J. R., Rodrigues, S., & Robinson, P. A. (2005). A unifying explanation of generalized seizures via the bifurcation analysis of a dynamical brain model. Cerebral Cortex (New York, N.Y.), 16, 1296–1313. doi:10.1093/cercor/bhj072.

    Article  Google Scholar 

  • Breakspear, M., Terry, J., & Friston, K. (2003a). Modulation of excitatory synaptic coupling facilitates synchronization and complex dynamics in a nonlinear model of neuronal dynamics. Network: computation in Neural Systems, 14, 703–732. doi:10.1088/0954-898X/14/4/305.

    Article  Google Scholar 

  • Breakspear, M., Terry, J., Friston, K., Williams, L., Brown, K., Brennan, J., et al. (2003b). A disturbance of nonlinear interdependence in scalp EEG of subjects with first episode schizophrenia. NeuroImage, 20, 466–478. doi:10.1016/S1053-8119(03)00332-X.

    Article  PubMed  CAS  Google Scholar 

  • Breakspear, M., Williams, L., & Stam, K. (2004). Topographic analysis of phase dynamics in neural systems reveals formation and dissolution of ‘dynamic cell assemblies’. Journal of Computational Neuroscience, 16, 49–68. doi:10.1023/B:JCNS.0000004841.66897.7d.

    Article  PubMed  Google Scholar 

  • Bressler, S. L., & McIntosh, A. R. (2007). The role of neural context in large-scale neurocognitive network operations. In A. R. McIntosh, & V. K. Jirsa (Eds.), Handbook of brain connectivity. Berlin: Springer.

    Google Scholar 

  • Brunel, N., Chance, F., Fourcaud, N., & Abbott, L. (2001). Effects of synaptic noise and filtering on the frequency response of spiking neurons. Physical Review Letters, 86, 2186–2189. doi:10.1103/PhysRevLett.86.2186.

    Article  PubMed  CAS  Google Scholar 

  • Buchel, C., & Friston, K. J. (1997). Modulation of connectivity in visual pathways by attention: cortical interactions evaluated with structural equation modelling and fMRI. Cerebral Cortex (New York, N.Y.), 7, 768–778. doi:10.1093/cercor/7.8.768.

    Article  CAS  Google Scholar 

  • Coombes, S. (2005). Waves, bumps, and patterns in neural field theories. Biological Cybernetics, 93, 91–108. doi:10.1007/s00422-005-0574-y.

    Article  PubMed  CAS  Google Scholar 

  • Coombes, S. (2006). Neural fields. Scholarpedia, 1(6), 1373.

    Google Scholar 

  • Coombes, S., Lord, G. J., & Owen, M. R. (2003). Waves and bumps in neuronal networks with axo-dendritic synaptic interactions. Physica D Nonlinear Phenomena, 178, 219–241. doi:10.1016/S0167-2789(03)00002-2.

    Article  Google Scholar 

  • Coombes Venkov, N. A., Shiau, L., Bojak, I., Liley, D. T. J., & Laing, R. (2007). Modeling electrocortical activity through improved local approximations of integral neural field equations. Physical Review E, 76, 051901.

    Article  CAS  Google Scholar 

  • da Silva, F., Blanes, W., Kalitzin, S. N., Parra, J., Suffczynski, P., & Velis, D. N. (2003). Epilepsies as dynamical diseases of brain systems: basic models of the transition between normal and epileptic activity. Epilepsia, 44, 72–83. doi:10.1111/j.0013-9580.2003.12005.x.

    Article  Google Scholar 

  • Deco, G., & Rolls, E. T. (2003). Attention and working memory: a dynamical model of neuronal activity in the prefrontal cortex. The European Journal of Neuroscience, 18, 2374–2390. doi:10.1046/j.1460-9568.2003.02956.x.

    Article  PubMed  Google Scholar 

  • Deco, G., & Rolls, E. T. (2004). A neurodynamical cortical model of visual attention and invariant object recognition. Vision Research, 44, 621–644. doi:10.1016/j.visres.2003.09.037.

    Article  PubMed  Google Scholar 

  • Deco, G., & Rolls, E. T. (2005). Attention, short term memory, and action selection: a unifying theory. Progress in Neurobiology, 76, 236–256.

    PubMed  Google Scholar 

  • Deco, G., & Rolls, E. (2006). Decision-making and Weber’s law: a neurophysiological model. The European Journal of Neuroscience, 24, 901–916. doi:10.1111/j.1460-9568.2006.04940.x.

    Article  PubMed  Google Scholar 

  • Deco, G., Jirsa, V. K., Robinson, P. A., Breakspear, M., & Friston, K. J. (2008). The dynamic brain: from spiking neurons to neural masses and cortical fields. PLoS Computational Biology, 4, e1000092.

    Article  PubMed  Google Scholar 

  • De Schutter, , Ekeberg, O., Kotaleski, J. H., Achard, P., & Lansner, A. (2005). Biophysically detailed modelling of microcircuits and beyond. Trends in Neurosciences, 28, 562–569. doi:10.1016/j.tins.2005.08.002.

    Article  PubMed  CAS  Google Scholar 

  • Dhamala, M., Jirsa, V. K., & Ding, M. (2004). Enhancement of neural synchrony by time delay. Physical Review Letters, 92, 074104. doi:10.1103/PhysRevLett.92.074104.

    Article  PubMed  CAS  Google Scholar 

  • Ernst, M. O., & Banks, M. S. (2002). Humans integrate visual and haptic information in a statistically optimal fashion. Nature, 415, 429–433. doi:10.1038/415429a.

    Article  PubMed  CAS  Google Scholar 

  • Faisal, A. A., Selen, L. P. J., & Wolpert, D. M. (2008). Noise in the nervous system. Nature Neuroscience Reviews, 9, 292–303. doi:10.1038/nrn2258.

    Article  CAS  Google Scholar 

  • Fourcaud, N., & Brunel, N. (2002). Dynamics of the firing probability of noisy integrate-and-fire neurons. Neural Computation, 14, 2057–2110. doi:10.1162/089976602320264015.

    Article  PubMed  Google Scholar 

  • Freeman, W. J. (1975). Mass action in the nervous system: Examination of the neurophysiological basis of adaptive behaviour through the EEG. New York: Academic Press.

    Google Scholar 

  • Freeman, W. J. (1979). Nonlinear gain mediating cortical stimulus response relations. Biol Cybem, 33, 237–247. doi:10.1007/BF00337412.

    Article  CAS  Google Scholar 

  • Freeman, W. J. (1987). Simulation of chaotic EEG patterns with a dynamic model of the olfactory system. Biological Cybernetics, 56, 139–150. doi:10.1007/BF00317988.

    Article  PubMed  CAS  Google Scholar 

  • Freeman, W. J., & Breakspear, M. (2007). Scale-free neocortical dynamics. Scholarpedia, 2, 1357.

    Google Scholar 

  • Friston, K. J. (1997). Another neural code? NeuroImage, 5, 213–220. doi:10.1006/nimg.1997.0260.

    Article  PubMed  CAS  Google Scholar 

  • Friston, K. J. (2002). Bayesian estimation of dynamical systems: an application to fMRI. NeuroImage, 16, 513–530. doi:10.1006/nimg.2001.1044.

    Article  PubMed  CAS  Google Scholar 

  • Friston, K. J. (2008). DEM: a variational treatment of dynamic systems. NeuroImage, 41, 849–885. doi:10.1016/j.neuroimage.2008.02.054.

    Article  PubMed  CAS  Google Scholar 

  • Friston, K. J., & Büchel, C. (2000). Attentional modulation of effective connectivity from V2 to V5/MT in humans. Proceedings of the National Academy of Sciences of the United States of America, 97, 7591–7596. doi:10.1073/pnas.97.13.7591.

    Article  PubMed  CAS  Google Scholar 

  • Friston, K. J., Harrison, L., & Penny, W. (2003). Dynamic causal modelling. NeuroImage, 19, 1273–1302. doi:10.1016/S1053-8119(03)00202-7.

    Article  PubMed  CAS  Google Scholar 

  • Garrido, M. I., Kilner, J. M., Kiebel, S. J., & Friston, K. J. (2007). Evoked brain responses are generated by feedback loops. Proceedings of the National Academy of Sciences of the United States of America, 104, 20961–20966. doi:10.1073/pnas.0706274105.

    Article  PubMed  CAS  Google Scholar 

  • Gray, C., Konig, P., Engel, A., & Singer, W. (1989). Oscillatory responses in cat visual cortex exhibit inter-columnar synchronization which reflects global stimulus properties. Science, 338, 334–337.

    CAS  Google Scholar 

  • Greicius, M. D., Krasnow, D., Reiss, A. L., & Menon, V. (2003). Functional connectivity in te resting brain: a network analysis of the default mode hypothesis. Proceedings of the National Academy of Sciences of the United States of America, 100, 253–258. doi:10.1073/pnas.0135058100.

    Article  PubMed  CAS  Google Scholar 

  • Haken, H. (1977). Synergetics: An introduction. Nonequilibrium phase-transitions and self-organization in physics, chemistry and biology. Berlin: Springer.

    Google Scholar 

  • Haken, H. (1996). Principles of brain functioning. A synergetic approach to brain activity, behavior, and cognition. Berlin: Springer.

    Google Scholar 

  • Harrison, L. M., David, O., & Friston, K. J. (2005). Stochastic models of neuronal dynamics. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 360, 1075–1091. doi:10.1098/rstb.2005.1648.

    Article  PubMed  CAS  Google Scholar 

  • Hodgkin, A. L., & Huxley, A. F. (1952). A quantitative description of membrane current and its application to conduction and excitation in nerve. The Journal of Physiology, 117, 500–544.

    PubMed  CAS  Google Scholar 

  • Honey, C., Kötter, R., Breakspear, M., & Sporns, O. (2007). Network structure of cerebral cortex shapes functional connectivity on multiple time scales. Proceedings of the National Academy of Sciences of the United States of America, 104, 10240–10245. doi:10.1073/pnas.0701519104.

    Article  PubMed  CAS  Google Scholar 

  • Hopfield, J. J. (1982). Neural networks and physical systems with emergent collective computational abilities. Proceedings of the National Academy of Sciences of the United States of America, 79, 2554–2558. doi:10.1073/pnas.79.8.2554.

    Article  PubMed  CAS  Google Scholar 

  • Hopfield, J. J. (1984). Neurons with graded response have collective computational properties like those of two-state neurons. Proceedings of the National Academy of Sciences of the United States of America, 81, 3088–3092. doi:10.1073/pnas.81.10.3088.

    Article  PubMed  CAS  Google Scholar 

  • Izhikevich, E. M. (2005). Dynamical systems in neuroscience: The geometry of excitability and bursting. Cambridge, MA: MIT Press.

    Google Scholar 

  • Izhikevich, E. M., & Edelman, G. M. (2008). Large-scale model of mammalian thalamocortical systems. Proceedings of the National Academy of Sciences of the United States of America, 105, 3593–3598. doi:10.1073/pnas.0712231105.

    Article  PubMed  CAS  Google Scholar 

  • Jacobs, R. A. (1999). Optimal integration of texture and motion cues to depth. Vision Research, 39, 3621–3629. doi:10.1016/S0042-6989(99)00088-7.

    Article  PubMed  CAS  Google Scholar 

  • Jansen, B. H., & Rit, V. G. (1995). Electroencephalogram and visual evoked potential generation in a mathematical model of coupled cortical columns. Biological Cybernetics, 73, 357–366. doi:10.1007/BF00199471.

    Article  PubMed  CAS  Google Scholar 

  • Jirsa, V. K. (2004). Connectivity and dynamics of neural information processing. Neuroinformatics, 2, 183–204. doi:10.1385/NI:2:2:183.

    Article  PubMed  Google Scholar 

  • Jirsa, V. K., & Haken, H. (1996). Field theory of electromagnetic brain activity. Physical Review Letters, 77, 960–963. doi:10.1103/PhysRevLett.77.960.

    Article  PubMed  CAS  Google Scholar 

  • Jirsa, V. K., & Haken, H. (1997). A derivation of a macroscopic field theory of the brain from the quasi-microscopic neural dynamics. Physica D Nonlinear Phenomena, 99, 503–526. doi:10.1016/S0167-2789(96)00166-2.

    Article  Google Scholar 

  • Jirsa, V. K., & Kelso, J. A. S. (2000). Spatiotemporal pattern formation in continuous systems with heterogeneous connection topologies. Physical Review E: Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics, 62(6), 8462–8465. doi:10.1103/PhysRevE.62.8462.

    CAS  Google Scholar 

  • Jirsa, V. K., & Kelso, J. A. S. (2004). Coordination dynamics. Berlin: Springer.

    Google Scholar 

  • Jirsa, V. K., & McIntosh, A. R. (2007). Handbook of brain connectivity. Berlin: Springer.

    Book  Google Scholar 

  • Jones, S. R., Pritchett, D. L., Stufflebeam, S. M., Haamalanen, M. S., & Moore, C. I. (2007). Neural correlates of tactile detection: a combined magnetoencephalography and biophysically based computational modeling study. The Journal of Neuroscience, 27, 10751–10764. doi:10.1523/JNEUROSCI.0482-07.2007.

    Article  PubMed  CAS  Google Scholar 

  • Kiebel, S. J., Garrido, M. I., & Friston, K. J. (2007). Dynamic causal modelling of evoked responses: the role of intrinsic connections. NeuroImage, 36, 332–345. doi:10.1016/j.neuroimage.2007.02.046.

    Article  PubMed  Google Scholar 

  • Knight, B., Manin, D., & Sirovich, L. (1996). Dynamical models of interacting neuron populations. In E. C. Gerf (Ed.), Symposium on robotics and cybernetics: Computational engineering in systems applications. Lille, France: Cite Scienti_que.

    Google Scholar 

  • Kording, K. P., & Wolpert, D. M. (2004). Bayesian integration in sensorimotor learning. Nature, 427, 244–247. doi:10.1038/nature02169.

    Article  PubMed  CAS  Google Scholar 

  • Larter, R., & Speelman, B. (1999). A coupled ordinary differential equation lattice model for the simulation of epileptic seizures. Chaos (Woodbury, N.Y.), 9, 795–804. doi:10.1063/1.166453.

    Article  Google Scholar 

  • Liley, D. T. J., & Bojak, I. (2005). Understanding the transition to seizure by modeling the epileptiform Activity of general anesthetic agents. Journal of Clinical Neurophysiology, 22(5), 300–313.

    PubMed  CAS  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. doi:10.1088/0954-898X/13/1/303.

    Article  Google Scholar 

  • Logothetis, N. K. (2002). The neural basis of the BOLD fMRI signal. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 357, 1003–1037. doi:10.1098/rstb.2002.1114.

    Article  PubMed  Google Scholar 

  • Lopes da Silva, F. H., Hoeks, A., Smits, H., & Zetterberg, L. H. (1974). Model of brain rhythmic activity: the alpha-rhythm of the thalamus. Kybernetik, 15, 27–37. doi:10.1007/BF00270757.

    Article  PubMed  CAS  Google Scholar 

  • Makeig, S., Delorme, A., Westerfield, M., Jung, T. P., Townsend, J., et al. (2004). Electroencephalographic brain dynamics following manually responded visual targets. PLoS Biology, 2, e176.

    Article  PubMed  Google Scholar 

  • Marreiros, A., Daunizeau, J., Kiebel, S. J., & Friston, K. J. (2008a). Population dynamics: variance and the sigmoid activation function. NeuroImage, 42, 147–157. doi:10.1016/j.neuroimage.2008.04.239.

    Article  PubMed  Google Scholar 

  • Marreiros, A., Kiebel, S. J., & Friston, K. J. (2008b). Dynamic causal modelling for fMRI: a two-state model. NeuroImage, 39, 269–278. doi:10.1016/j.neuroimage.2007.08.019.

    Article  PubMed  CAS  Google Scholar 

  • McIntosh, A. R. M. (2004). Contexts and catalysts: a resolution of the localization and integration of function in the brain. Neuroinformatics, 2, 175–181. doi:10.1385/NI:2:2:175.

    Article  PubMed  Google Scholar 

  • Morris, C., & Lecar, H. (1981). Voltage oscillations in the barnacle giant muscle fiber. Biophysical Journal, 35, 193–213.

    PubMed  CAS  Google Scholar 

  • Nunez, P. L. (1974). The brain wave equation: a model for eeg. Mathematical Biosciences, 21, 279–297. doi:10.1016/0025–5564(74)90020–0.

    Article  Google Scholar 

  • Nunez, P. L. (1995). Neocortical dynamics and human brain rhythms. Oxford: Oxford University Press.

    Google Scholar 

  • Omurtag, A., Knight, B., & Sirovich, L. (2000). On the simulation of large populations of neurons. J. Comp. Neurosci., 8, 51.53.

    Google Scholar 

  • Penny, W. D., Stephan, K. E., Mechelli, A., & Friston, K. J. (2004). Comparing dynamic causal models. NeuroImage, 22, 1157–1172. doi:10.1016/j.neuroimage.2004.03.026.

    Article  PubMed  CAS  Google Scholar 

  • Qubbaj, M. R., & Jirsa, V. K. (2007). Neural field dynamics with heterogeneous connection topology. Physical Review Letters, 98, 238102. doi:10.1103/PhysRevLett.98.238102.

    Article  PubMed  CAS  Google Scholar 

  • Rabinovich, M. I., Huerta, R., & Laurent, G. (2008b). Transient dynamics for neural processing. Science, 321, 48–50. doi:10.1126/science.1155564.

    Article  PubMed  CAS  Google Scholar 

  • Rabinovich, M. I., Huerta, R., Varona, P., & Afraimovich, V. S. (2008a). Transient cognitive dynamics, metastability, and decision making. PLoS Computational Biology, 4, e1000072. doi:10.1371/journal.pcbi.1000072.

    Article  PubMed  CAS  Google Scholar 

  • Rennie, C., Robinson, P., & Wright, J. (2002). Unified neurophysical model of EEG spectra and evoked potentials. Biological Cybernetics, 86(6), 457–471. doi:10.1007/s00422-002-0310-9.

    Article  PubMed  CAS  Google Scholar 

  • Robinson, (2006). Patchy propagators, brain dynamics, and the generation of spatially structured gamma oscillations. Physical Review E, 73, 041904.

    Article  CAS  Google Scholar 

  • Robinson, P., Rennie, C., & Rowe, D. (2002). Dynamics of large-scale brain activity in normal arousal states and epileptic seizures. Physical Review E: Statistical, Nonlinear, and Soft Matter Physics, 65, 41924. doi:10.1103/PhysRevE.65.041924.

    CAS  Google Scholar 

  • Robinson, P. A., Rennie, C. A., & Wright, J. J. (1997). Propagation and stability of waves of electrical activity in the cerebral cortex. Physical Review E: Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics, 56, 826–840. doi:10.1103/PhysRevE.56.826.

    CAS  Google Scholar 

  • Robinson, P. A., Rennie, C. J., Wright, J. J., Bahramali, H., Gordon, E., & Rowe, D. L. (2001). Prediction of electroencephalographic spectra from neurophysiology. Physical Review E: Statistical, Nonlinear, and Soft Matter Physics, 63, 021903. doi:10.1103/PhysRevE.63.021903.

    CAS  Google Scholar 

  • Rodrigues, S., Terry, J., & Breakspear, M. (2006). On the genesis of spike-wave oscillations in a mean-field model of human thalamic and corticothalamic dynamics. Physics Letters. [Part A], 355, 352–357. doi:10.1016/j.physleta.2006.03.003.

    Article  CAS  Google Scholar 

  • Rubinov, M., Knock, S. A., Stam, C. J., Micheloyannis, S., Harris, A. W. F, Williams, L. M., et al. (2008) Small world properties of nonlinear brain activity in schizophrenia. Human Brain Mapping. doi:10.1002/hbm.20517

  • Salvador, R., Suckling, J., Coleman, M. R., Pickard, J. D., Menon, D., & Bullmore, E. (2004). Neurophysiological architecture of functional magnetic resonance images of human brain. Cerebral Cortex (New York, N.Y.), 15, 1332–1342. doi:10.1093/cercor/bhi016.

    Article  Google Scholar 

  • Shlesinger, M. F., Zaslavsky, G. M., & Klafter, J. (1993). Strange kinetics. Nature, 363, 31–37. doi:10.1038/363031a0.

    Article  CAS  Google Scholar 

  • Sporns, O., & Zwi, J. D. (2004). The small world of the cerebral cortex. Neuroinformatics, 2, 145–162. doi:10.1385/NI:2:2:145.

    Article  PubMed  Google Scholar 

  • Stam, C. J. (2004). Functional connectivity patterns of human magnetoencephalographic recordings: a ‘small-world’ network? Neuroscience Letters, 355, 25–28. doi:10.1016/j.neulet.2003.10.063.

    Article  PubMed  CAS  Google Scholar 

  • Stam, C. J., Jones, B. F., Nolte, G., Breakspear, M., & Scheltens, P. H. (2006). Small-world networks and functional connectivity in Alzheimer’s disease. Cerebral Cortex (New York, N.Y.), 17, 92–99. doi:10.1093/cercor/bhj127.

    Article  Google Scholar 

  • Stephan, K. E., Harrison, L. M., Kiebel, S. J., David, O., Penny, W. D., & Friston, K. J. (2007a). Dynamic causal models of neural system dynamics: current state and future extensions. Journal of Biosciences, 32, 129–144. doi:10.1007/s12038-007-0012-5.

    Article  PubMed  Google Scholar 

  • Stephan, K. E., Kasper, L., Harrison, L. M., Daunizeau, J., den Ouden, H. E. M., Breakspear, M., et al. (2008). Nonlinear dynamic causal models for fMRI. NeuroImage, 42, 649–662. doi:10.1016/j.neuroimage.2008.04.262.

    Article  PubMed  Google Scholar 

  • Stephan, K. E., Marshall, J. C., Penny, W. D., Friston, K. J., & Fink, G. R. (2007c). Inter-hemispheric integration of visual processing during task-driven lateralization. The Journal of Neuroscience, 27, 3512–3522. doi:10.1523/JNEUROSCI.4766-06.2007.

    Article  PubMed  CAS  Google Scholar 

  • Stephan, K. E., Weiskopf, N., Drysdale, P. M., Robinson, P. A., & Friston, K. J. (2007b). Comparing hemodynamic models with DCM. NeuroImage, 38, 387–401. doi:10.1016/j.neuroimage.2007.07.040.

    Article  PubMed  Google Scholar 

  • Valdes, P. A., Jimenez, J. C., Riera, J., Biscay, R., & Ozaki, T. (1999). Nonlinear EEG analysis based on a neural mass model. Biological Cybernetics, 81, 415–424. doi:10.1007/s004220050572.

    Article  PubMed  CAS  Google Scholar 

  • Valdes, P. A., Sanchez-Bornot, J. M., Sotero, R. C., Iturria-Medina, Y., Aleman-Gomez, Y., & Bosch-Bayard, J. (2008) Analyzing brain oscillations with EEG/fMRI fusion. Human Brain Mapping, in press.

  • Wang, X.-J. (2002). Probabilistic decision making by slow reverberation in cortical circuits. Neuron, 36, 955–968. doi:10.1016/S0896-6273(02)01092-9.

    Article  PubMed  CAS  Google Scholar 

  • Wilson, H. R., Blake, R., & Lee, S.-H. (2001). Dynamics of traveling waves in visual perception. Nature, 412, 907–911. doi:10.1038/35091066.

    Article  PubMed  CAS  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  • Wilson, H. R., & Cowan, J. D. (1973). A mathematical theory of the functional dynamics of cortical and thalamic nervous tissue. Kybernetik, 13(2), 55–80. doi:10.1007/BF00288786.

    Article  PubMed  CAS  Google Scholar 

  • Wong, K. F., & Wang, X. J. (2006). A recurrent network mechanism of time integration in perceptual decisions. The Journal of Neuroscience, 26, 1314–1328. doi:10.1523/JNEUROSCI.3733-05.2006.

    Article  PubMed  CAS  Google Scholar 

  • Wright, J. J. (2000). Modeling the whole brain in action. In E. Gordon (Ed.), Integrative neuroscience. London: Harwood.

    Google Scholar 

  • Zaslavsky, G. M. (2002). Chaos, fractional kinetics and anomalous transport. Physics Reports, 371, 461–580. doi:10.1016/S0370-1573(02)00331-9.

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank K. Friston, V. Jirsa, A.R. McIntosh, G. Deco, K. Stephan, P. Robinson, J.R. Terry, L. Harrison, P. Valdes-Soza, E. Bullmore, W. Freeman, C. Stam, O. Sporns and others for interesting discussions relating to noise, networks, dynamics and cortical function. This work was supported by Brain NRG JSMF22002082 and ARC grants DP0667065 and TS0669860.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael Breakspear.

Appendices

Appendix A: Biophysical models of spiking neurons

Biophysical models of spiking neurons derive from the Hodgkin Huxley framework. An elegant description of their dynamical principles is provided in Izhikevich (2005). A recent summary account is Breakspear and Jirsa (2007). The approach here is to provide plausible examples of dynamics within large networks of spiking neurons.

These models are based on ion currents through leaky and voltage-dependent transmembrane channels and have the form,

$$C\frac{{{\text{d}}V}}{{{\text{d}}t}} = I + {\sum\limits_{{\text{ion}}} {f_{{{\text{ion}}}} {\left( V \right)} \times {\left( {V - V_{{{\text{ion}}}} } \right)}} }$$
(3)

where V is the membrane potential, ion denotes various ion species (Na+, Ca2+, K+), f ion are the voltage-dependent conductances for each ion species and I is a current, either experimentally injected (as per in vivo preparations) or induced at the synapse by ligand-gated channels. The functions f ion represent the states of voltage-gated activation and deactivation channels, generating currents due to the gradient between the membrane potential V and the Nernst potential for that ion species V ion. For the present purposes, we incorporate only voltage-dependent Na+ and K+ activation channels,

$$f_{{\text{Na}}} \left( V \right) = g_{{\text{Na}}} .m\left( V \right),$$
(4)
$$f_{\text{K}} \left( V \right) = g_{\text{K}} .n\left( V \right).$$
(5)

The functions m and n are the voltage-dependent fraction of open channels of each respective ion species. These are then multiplied by the conductances g ion to give the ion currents. Sodium channels respond very quickly to changes in membrane potential, whereas potassium channels relax towards their steady state values relatively slowly. Hence,

$$m\left( V \right) = \frac{{m_{\max } }}{{1 + e^{\left( {{{V_m - V} \mathord{\left/ {\vphantom {{V_m - V} \sigma }} \right. \kern-\nulldelimiterspace} \sigma }} \right)} }}.$$
(6)

represents the instantaneous state of sodium channels, where m max is the maximum fraction of open sodium channels, V m the voltage at which half the channels are open, and σ controls the range of voltages over which the change occurs. On the other hand,

$$\frac{{{\text{d}}n\left( V \right)}}{{{\text{d}}t}} = \frac{{\left( {n_\infty \left( V \right) - n\left( V \right)} \right)}}{{\tau _n \left( V \right)}}$$
(7)

with,

$$n_\infty \left( V \right) = \frac{{n_{\max } }}{{1 + e^{\left( {{{V_m - V} \mathord{\left/ {\vphantom {{V_m - V} \sigma }} \right. \kern-\nulldelimiterspace} \sigma }} \right)} }}.$$
(8)

captures the relaxation of potassium conductance towards its voltage-driven state, with τ n the time constant of the potassium channel relaxation, and n max, V n , and σ have the same meaning as the sodium channels. As discussed in Izhikevich (2005), such a model is the minimum required in order to exhibit both type I and type II neuronal firing. Type I neurons (integrators) fire through a saddle node bifurcation and, because of this, act to integrate the average rate of synaptic input, firing at a proportional rate if a threshold is surpassed. Type II neurons (resonators) fire through a Hopf bifurcation and, hence, show a resonance to a preferred frequency. The influences of other neurons, nonspecific noise and sensory-evoked synaptic activity are introduced via the synaptic current term. Hence, if we take an ensemble of N neurons {1,2,…, N}, then the synaptic current at neuron j is due to three inputs,

$$I^j = I_{{\text{ext}}}^j + I_{{\text{noise}}}^j + \sum\limits_k {H_c \left( {g\left( {V^k } \right)} \right)} .$$
(9)

I ext represents a synaptic current induced by specific inputs from outside the local ensemble—such as due to a sensory stimulus relayed via the thalamus. I noise represents a non-specific stochastic input—representing all subcortical and cortico-cortical inputs not accounted for by I ext plus local stochastic influences such as amplified channel noise (Faisal et al. 2008). The final term represents inputs summed from all other neurons within the ensemble. The coupling function H is a connectivity map that introduces the presence and weighted strength of synaptic connections between neurons j and k. The function g incorporates two processes. The first is a thresholded step function that yields a brief burst of activity (a boxcar function) whenever neuron k issues an action potential. The second process is the conversion of this synaptic activity into a post-synaptic current at the index neuron j. For the present purpose, the post-synaptic current does not have an intrinsic time constant. Hence the frequency response of the neuron is determined by the membrane time constants. More interesting synaptic forms, which act as temporal filters, introduce additional time scales into such models (Fourcaud and Brunel 2002).

The parameters of the individual neurons were set in a physiologically realistic range and so that, in the absence of any synaptic current I = 0 each neuron is quiescent. In response to a sufficiently strong stochastic input, neurons fire stochastically due to intermittent subthreshold excursions. However, in the presence of a sufficiently strong external current I external each neuron undergoes a “bifurcation” from its quiescent state to spontaneous and periodic firing. This bifurcation is known as a “saddle-node” bifurcation (see Izhikevich 2005; Breakspear and Jirsa 2007).

The parameters are:

Capacitance:

C = 1;

Synaptic current (default):

I = 0;

Leaky channels:

V L = −80; g L = 8;

Sodium channels:

V Na = 60; g Na = 20; V m = −20; σ m = 15; m max = 1; τ m = 1;

Potassium channels:

V K = −90; g K = 10; V n  = −25; σ n  = 5; n max = 1; τ n  = 1;

Appendix B: Neural mass of cortical columns

Neural mass models aim to reduce the computational load of large networks of spiking neurons (Freeman 1975). The equations here begin with a model of the behavior of local ensembles of neurons, with dynamical variables V = [V W Z] representing ensemble averages over the extent of a local neural region (Eqs. 10, 11, 12, 13 and 14). The effect of the many thousands of inputs (i.e. the local mean field) into each local subsystem from other neurons is then introduced by coupling such nodes together through long-range excitatory projections (Eqs. 15 and 16) The model here was extended from a description by (Larter and Speelman 1999) in Breakspear et al. (2003a). An alternative neural mass model is the Jansen model (Jansen and Rit 1995) as recently further elaborated by Garrido et al. (2007). As with the spiking neural ensembles, the present goal of these neural mass models is to provide plausible examples of interesting and computationally relevant dynamics.

The dynamical variables represent the mean membrane potential of pyramidal cells, V, and inhibitory interneurons, Z, and the average number of ‘open’ potassium ion channels, W. The evolution equations are adapted from a study of epileptic seizures in hippocampal slices (Larter and Speelman 1999), which in turn are derived from the model of Morris and Lecar (1981) by introducing feedback and population effects. The mean cell membrane potential of the pyramidal cells is governed by the conductance of sodium, potassium and calcium ions through voltage- and ligand-gated membrane channels,

$$\frac{{{\text{d}}V}}{{{\text{d}}t}} = - \left( {g_{{\text{Ca}}} + r_{{\text{NMDA}}} a_{{\text{ee}}} Q_{\text{V}} } \right)m_{{\text{Ca}}} \left( {V - V_{{\text{Ca}}} } \right) - \left( {g_{{\text{na}}} m_{{\text{na}}} + a_{{\text{ee}}} Q_{\text{V}} } \right)\left( {V - V_{{\text{na}}} } \right) - g_{\text{K}} W\left( {V - V_{\text{K}} } \right) - g_{\text{L}} \left( {V - V_{\text{L}} } \right) + a_{{\text{ie}}} ZQ_{\text{Z}} + a_{{\text{ne}}} I_\delta ,$$
(10)
$$\frac{{{\text{d}}Z}}{{{\text{d}}t}} = b\left( {a_{{\text{ni}}} I_\delta + a_{{\text{ei}}} VQ_{\text{V}} } \right),$$
(11)

where g ion is the maximum conductance of each population of ion species if all channels are open, m ion is the fraction of channels open, V ion is the Nernst potential for that ion species, a ab are synaptic weights between respective populations and Q V(Z) is the firing rate of the excitatory (inhibitory) neurons. The fraction of open channels is determined by the sigmoid-shaped ‘neural activation function’,

$$m_{{\text{ion}}} = 0.5\left( {1 + {\text{tanh}}\left( {\frac{{V - V_{\text{T}} }}{{\delta _{{\text{ion}}} }}} \right)} \right),$$
(12)

where δ ion incorporates the variance of this distribution. The fraction of open potassium channels is slightly more complicated, being governed by W, with

$$\frac{{{\text{d}}W}}{{{\text{d}}t}} = \frac{{\phi \left( {m_k - W} \right)}}{\tau },$$
(13)

where ϕ is a temperature scaling factor and τ is a ‘relaxation’ time constant. Cell firing rate is also determined by sigmoid activation functions,

$$Q_{\text{V}} = 0.5xQ_{V_{\max } } \left( {1 + \tanh \left( {\frac{{V - V_{\text{T}} }}{{\delta _{\text{V}} }}} \right)} \right),$$
(14)

where the Q max is the maximum firing rate. An analogous term is also introduced for the inhibitory cells. The firing of these cell populations feeds back onto the ensemble through synaptic coupling to open ligand-gated channels and raise or lower the membrane potential accordingly. In the case of excitatory-to-inhibitory and inhibitory-to-excitatory connections, this is modelled as additional inputs to the flow of ions across the membrane channel, weighted by functional synaptic factors, a ei and a ie. In the case of excitatory to excitatory connections, the rate of firing Q v is assumed to lead to a proportional release of glutamate neurotransmitter across the synapse, onto two classes of ligand-gated ion channels: (1) AMPA channels, which open an additional population of sodium channels, and (2) NMDA receptors, which open an additional population of voltage-gated sodium and calcium channels. r NMDA incorporates the ratio of NMDA to AMPA receptors. Finally, a stochastic synaptic current term, \(I_\delta \) is also present and weighted by the synaptic efficacy term a ne.

Each of the set of Eqs. 10, 11, 12, 13 and 14 govern the dynamics within each local cell assembly, which hence require inputs from other cortical sources. Coupling between N nodes is introduced as competitive agonist excitatory action at the same populations of NMDA and AMPA receptors. Locating the i-th node at position x i this is represented as,

$$\begin{array}{*{20}l} {{\frac{{dV{\left( {x_{i} ,t} \right)}}}{{dt}} = F_{{\text{a}}} {\left( {V{\left( {x_{i} ,t} \right)} + c{\sum\limits_j {H_{{ij}} V{\left( {x_{j} ,t} \right)}} }} \right)},} \hfill} & {{i = 1,2, \ldots ,N} \hfill} \\ \end{array}$$
(15)

for i,j = 1,…,N. The function F represents the within-node dynamics as given by B1 and H ij is the connection strength between nodes i and j. The parameter c represents the strength of excitatory coupling between cortical columns. If c = 0 the systems evolve independently. c > 0 introduces interdependences between consecutive columns. c = 1 corresponds to maximum coupling, with excitatory input from outside each column surpassing excitatory input from within each column.

The spatial kernel H ij can hence be viewed as a ‘synaptic footprint’—the dependence of coupling strength on distance. In the present simulations we employ,

$$\begin{array}{*{20}c} {H_{{ij}} = \frac{1}{{{\left| {x_{i} - x_{j} } \right|}^{{\mathbf{K}}} }},}{j \ne i,} \\ { = - 1,}{j = i,} \\ \end{array} $$
(16)

K > 0 is a constant vector which parameterises the shape of the synaptic footprint. When K is close to zero, H ij is almost constant, so that ‘mean field’ effects dominate. Increasing K decreases the effective support of H ij towards a neighborhood of x i . Note that K is of the same dimension of the modelled array, in order to permit anisotropic coupling K 1 K 2.

All physiologically measurable parameters (conductances, threshold potentials and Nernst potentials) are set to their accepted values (Larter and Speelman 1999). The behavior of this model is more fully described in Breakspear et al. (2003a).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Breakspear, M., Knock, S. Kinetic Models of Brain Activity. Brain Imaging and Behavior 2, 270–288 (2008). https://doi.org/10.1007/s11682-008-9033-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11682-008-9033-4

Keywords

Navigation