Skip to main content

Advertisement

Log in

A coarse-graining framework for spiking neuronal networks: from strongly-coupled conductance-based integrate-and-fire neurons to augmented systems of ODEs

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

Abstract

Homogeneously structured, fluctuation-driven networks of spiking neurons can exhibit a wide variety of dynamical behaviors, ranging from homogeneity to synchrony. We extend our partitioned-ensemble average (PEA) formalism proposed in Zhang et al. (Journal of Computational Neuroscience, 37(1), 81–104, 2014a) to systematically coarse grain the heterogeneous dynamics of strongly coupled, conductance-based integrate-and-fire neuronal networks. The population dynamics models derived here successfully capture the so-called multiple-firing events (MFEs), which emerge naturally in fluctuation-driven networks of strongly coupled neurons. Although these MFEs likely play a crucial role in the generation of the neuronal avalanches observed in vitro and in vivo, the mechanisms underlying these MFEs cannot easily be understood using standard population dynamic models. Using our PEA formalism, we systematically generate a sequence of model reductions, going from Master equations, to Fokker-Planck equations, and finally, to an augmented system of ordinary differential equations. Furthermore, we show that these reductions can faithfully describe the heterogeneous dynamic regimes underlying the generation of MFEs in strongly coupled conductance-based integrate-and-fire neuronal networks.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

References

  • Abbott, L.F., & van Vreeswijk, C.A. (1993). Asynchronous states in networks of pulse-coupled neurons. Physical Review E, 48, 1483–1488.

    Article  CAS  Google Scholar 

  • Anderson, J., Lampl, I., Reichova, I., Carandini, M., Ferster, D. (2000). Stimulus dependence of two-state fluctuations of membrane potential in cat visual cortex. Nature Neuroscience, 3(6), 617–621.

    Article  CAS  PubMed  Google Scholar 

  • Bak, P., Tang, C., Wiesenfeld, K. (1987). Self-organized criticality: an explanation of 1/f noise. Physical Review Letters, 59(4), 381–384.

    Article  CAS  PubMed  Google Scholar 

  • Battaglia, D., & Hansel, D. (2011). Synchronous chaos and broad band gamma rhythm in a minimal multi-layer model of primary visual cortex. PLoS Computational Biology, 7.

  • Buzsaki, G., & Wang, X.J. (2012). Mechanisms of gamma oscillations. Annual Reviews in the Neurosciences, 35, 203–225.

    Article  CAS  Google Scholar 

  • Bornholdt, S., & Rohl, T. (2003). Self-organized critical neural networks. Physical Review E, 67, 066118.

    Article  CAS  Google Scholar 

  • Bressloff, P.C. (2015). Path-integral methods for analyzing the effects of fluctuations in stochastic hybrid neural networks. Journal of Mathematical Neuroscience, 5, 4.

    Article  PubMed  PubMed Central  Google Scholar 

  • Brunel, N. (2000). Dynamics of sparsely connected networks of excitatory and inhibitory spiking neurons. The Journal of Comparative Neurology, 8, 183–208.

    CAS  Google Scholar 

  • Brunel, N., & Hakim, V. (1999). Fast global oscillations in networks of integrate-and-fire neurons with low firing rates. Neural Computation, 11, 1621–1671.

    Article  CAS  PubMed  Google Scholar 

  • Bruzsaki, G., & Draguhn, A. (2004). Neuronal oscillations in cortical networks. Science, 304, 1926–1929.

    Article  CAS  Google Scholar 

  • Buice, M.A., & Chow, C.C. (2007). Correlations, fluctuations, and stability of a finite-size network of coupled oscillators. Physical Review E, 76, 031118.1-031118.25.

    Article  CAS  Google Scholar 

  • Buice, M.A., Cowan, J.D., Chow, C.C. (2010). Systematic fluctuation expansion for neural network activity equations. Neural Computation, 22(2), 377–426.

    Article  PubMed  PubMed Central  Google Scholar 

  • Cai, D., Tao, L., Shelley, M., McLaughlin, D. (2004). An effective kinetic representation of fluctuation-driven neuronal networks with application to simple and complex cells in visual cortex. Proceedings of the National Academy of Sciences of the United States of America, 101(20), 7757–7762.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Cai, D., Tao, L., Rangan, A., McLaughlin, D. (2006). Kinetic theory for neuronal network dynamics. Communication in Mathematical Sciences, 4, 97–127.

    Article  Google Scholar 

  • Cardanobile, S., & Rotter, S. (2010). Multiplicatively interacting point processes and applications to neural modeling. Journal of Computational Neuroscience, 28, 267–284.

    Article  PubMed  Google Scholar 

  • Churchland, M.M., & et al. (2010). Stimulus onset quenches neural variability: a widespread cortical phenomenon. Nature Neuroscience, 13, 3:369–378.

    Article  PubMed  CAS  Google Scholar 

  • Csicsvari, J., Hirase, H., Mamiya, A., Buzsaki, G. (2000). Ensemble patterns of hippocampal ca3-ca1 neurons during sharp wave-associated population events. Neuron, 28, 585–594.

    Article  CAS  PubMed  Google Scholar 

  • Destexhe, A., & Pare, D. (1999). Impact of network activity on the integrative properties of neocortical pyramidal neurons in vivo. Journal of Neurophysiology, 81, 1531–1547.

    Article  CAS  PubMed  Google Scholar 

  • Destexhe, A., Rudolph, M., Pare, D. (2003). The high-conductance state of neocortical neurons in vivo. Nature Reviews. Neuroscience, 4, 739–751.

    Article  CAS  PubMed  Google Scholar 

  • Dehghani, N., Hatsopoulos, N.G., Haga, N.G., Parker, R.A., Greger, B., Halgren, E., Cash, S.S., Destexhe, A. (2012). Avalanche analysis from multi-electrode ensemble recordings in cat, monkey and human cerebral cortex during wakefulness and sleep. Frontiers in Physiology, 3.

  • DeVille, L., & Zheng, Y. (2014). Synchrony and periodicity in excitable neural networks with multiple subpopulations. SIAM Journal on Applied Dynamical Systems, 13(3), 1060–1081.

    Article  Google Scholar 

  • El Boustani, S., & Destexhe, A. (2009). A master equation formalism for macroscopic modeling of asynchronous irregular activity states. Neural Computation, 21(1), 46–100.

    Article  PubMed  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Fries, P. (2009). Neuronal gamma-band synchronization as a fundamental process in cortical computation. Annual Review of Neuroscience, 32, 209–24.

    Article  CAS  PubMed  Google Scholar 

  • Grill-Spector, K., & Weiner, K. (2014). The functional architecture of the ventral temporal cortex and its role in categorization. Nature Reviews in the Neurosciences, 15, 536–548.

    Article  CAS  Google Scholar 

  • Hahn, G., Petermann, T., Havenith, M.N., Yu, S., Singer, W., Plenz, D., Nikolic, D. (2010). Neuronal avalanches in spontaneous activity in vivo. Journal of Neurophysiology, 104, 3313–3322.

    Article  Google Scholar 

  • Hansel, D., & Sompolinsky, H. (1996). Chaos and synchrony in a model of a hypercolumn in visual cortex. Journal of Computational Neuroscience, 3, 7–34.

    Article  CAS  PubMed  Google Scholar 

  • Hatsopoulos, N.G., Ojakangas, C.L., Paniniski, L., Donoghue, J.P. (1998). Information about movement direction obtained from synchronous activity of motor cortical neurons. Proceedings of the National Academy of Sciences, 95, 15706–15711.

    Article  CAS  Google Scholar 

  • Helias, M., Deger, M., Rotter, S., Diesmann, M. (2010). Instantaneous nonlinear processing by pulse-coupled threshold units. PLoS Computational Biology, 6(9), e1000929.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  • Hertz, A.V.M., & Hopfield, J.J. (1995). Earthquake cycles and neural reverberations: collective oscillations in systems with pulse-coupled threshold elements. Physical Review Letters, 75(6), 1222–1225.

    Article  Google Scholar 

  • Hu, Y., Trousdale, J., Josic, K., Shea-Brown, E. (2013). Motif statistics and spike correlations in neuronal networks. Journal of Statistical Mechanics, P03012, 1–51.

    Google Scholar 

  • Kenet, T., Bibitchkov, D., Tsodyks, M., Grinvald, A., Arieli, A. (2003). Spontaneously emerging cortical representations of visual attributes. Nature, 425, 954–956.

    Article  CAS  PubMed  Google Scholar 

  • Knight, B. (1972). The relationship between the firing rate of a single neuron and the level of activity in a population of neurons. The Journal of General Physiology, 59, 734.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Koch, C. (1999). Biophysics of computation. Oxford: Oxford University Press.

    Google Scholar 

  • Kohn, A., & Smith, M.A. (2005). Stimulus dependence of neuronal correlation in primary visual cortex of the macaque. Journal of Neuroscience, 25, 3661–73.

    Article  CAS  PubMed  Google Scholar 

  • Kriener, B., Tetzlaff, T., Aertsen, A., Diesmann, M., Rotter, S. (2008). Correlatilons and population dynamics in cortical networks. Neural Computation, 20, 2185–2226.

    Article  PubMed  Google Scholar 

  • Ledoux, E., & Brunel, N. (2011). Dynamics of networks of excitatory and inhibitory neurons in response to time-dependent inputs. Frontiers in Computational Neuroscience, 5, 25,1–17.

    Article  Google Scholar 

  • Leinekugel, X., Khazipov, R., Cannon, R., Hirase, H., Ben-Ari, Y., Buzsaki, G. (2002). Correlated bursts of activity in the neonatal hippocampus in vivo. Science, 296, 2049–2052.

    Article  CAS  PubMed  Google Scholar 

  • Litwin-Kumar, A., & Doiron, B. (2012). Slow dynamics and high variability in balanced cortical networks with clustered connections. Nature Neuroscience, 15(11), 1498–1505.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Mazzoni, A., Broccard, F.D., Garcia-Perez, E., Bonifazi, P., Ruaro, M.E., Torre, V. (2007). On the dynamics of the spontaneous activity in neuronal networks. PloS One, 5, e439.

    Article  CAS  Google Scholar 

  • Nykamp, D. (2000). A population density approach that facilitates large scale modeling of neural networks: analysis and application to orientation tuning. Journal of Computational Neuroscience, 8, 19–50.

    Article  CAS  PubMed  Google Scholar 

  • Newhall, K. A., Kovac̆ic̆, G., Kramer, P.R., et al. (2010). Cascade-induced synchrony in stochastically driven neuronal networks. Physical Review E, 82(1), 041903.

    Article  CAS  Google Scholar 

  • Ohira, T., & Cowan, J.D. (1993). Master-equation approach to stochastic neurodynamics. Physical Review. E, Statistical Physics, Plasmas, Fluids, and Related Interdisciplinary Topics, 48(3), 2259–2266.

    CAS  PubMed  Google Scholar 

  • Omurtag, A., Kaplan, E., Knight, B., Sirovich, L. (2000). A population approach to cortical dynamics with an application to orientation tuning. Network: Computation in Neural Systems, 11(4), 247–260.

    Article  CAS  Google Scholar 

  • Ostojic, S., & Brunel, N. (2011). From spiking neuron models to linear-nonlinear models. PLoS Computational Biology, 7, 1:e1001056.

    Article  PubMed  CAS  Google Scholar 

  • Percival, D.B., & Walden, A.T. (1993). Spectral analysis for physical applications. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Petermann, T., Thiagarajan, T.C., Lebedev, M.A., Nicolelis, M.A.L., Chailvo, D.R., Plenz, D. (2009). Spontaneous cortical activity in awake monkeys composed of neuronal avalanches. Proceedings of the National Academy of Sciences, 106, 37:15921–15926.

    Article  Google Scholar 

  • Plenz, D., Stewart, C.V., Shew, W., Yang, H., Klaus, A., Bellay, T. (2011). Multi-electrode array recordings of neuronal avalanches in organotypic cultures. Journal of Visualized Experiments, 54, 2949.

    Google Scholar 

  • Poil, S.S., Hardstone, R., Mansvelder, H.D., Linkenkaer-Hansen, K. (2012). Critical-state dynamics of avalanches and oscillations jointly emerge from balanced excitation/inhibition in neuronal networks. The Journal of Neuroscience, 33, 9817–9823.

    Article  CAS  Google Scholar 

  • Rangan, A.V. (2009). Diagrammatic expansion of pulse-coupled network dynamics. Physical Reviews Letters, 102, 158101.

    Article  CAS  Google Scholar 

  • Rangan, A.V., & Cai, D. (2006). Maximum-entropy closures for kinetic theories of neuronal network dynamics. Physical Review Letters, 96, 178101.

    Article  PubMed  CAS  Google Scholar 

  • Rangan, A.V., & Young, L.S. (2013a). Dynamics of spiking neurons: between homogeneity and synchrony. Journal of Computational Neuroscience, 34(3), 433–460.

    Article  PubMed  Google Scholar 

  • Rangan, A.V., & Young, L.S. (2013b). Emergent dynamics in a model of visual cortex. Journal of Computational Neuroscience, 35(2), 155–167.

    Article  PubMed  PubMed Central  Google Scholar 

  • Richardson, M.J. (1918). Effects of synaptic conductance on the voltage distribution and firing rate of spiking neurons. Physical Review E, 69(05), 2004.

    Google Scholar 

  • Robert, P., & Touboul, J. (2016). On the dynamics of random networks. Journal of Statistical Physics, 165, 545–584.

    Article  Google Scholar 

  • Roopum, A.K., Kramer, M.A., Carracedo, L.M., Kaiser, M., Davies, C.H., Traub, R.D., Kopell, N.J., Whittington, M.A. (2008). Temporal interactions between cortical rhythms. Frontiers in Neuroscience, 2, 145–154.

    Article  Google Scholar 

  • Roxin, A., Brunel, N., Hansel, D., Mongillo, G., Vreeswijk, C.V. (2011). On the distribution of firing rates in networks of cortical neurons. The Journal of Neuroscience, 31(45), 16217–16226.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Sakata, S., & Harris, K.D. (2009). Laminar structure of spontaneous and sensory-evoked population activity in auditory cortex. Neuron, 12(3), 404–418.

    Article  CAS  Google Scholar 

  • Samonds, J.M., Zhou, Z., Bernard, M.R., Bonds, A.B. (2005). Synchronous activity in cat visual cortex encodes collinear and cocircular contours. Journal of Neurophysiology, 95, 4:2602–2616.

    PubMed  Google Scholar 

  • Seejnowski, T.J., & Paulsen, O. (2006). Network oscillations: emerging computational principles. The Journal of Neuroscience, 26, 1673–1676.

    Article  CAS  Google Scholar 

  • Singer, W. (1999). Neuronal synchrony: a versatile code for the definition of relations? Neuron, 24, 49–65.

    Article  CAS  PubMed  Google Scholar 

  • Sirovich, L., Omurtag, A., Knight, B. (2000). Dynamics of neuronal populations; the equilibrium solution. SIAM Journal on Applied Mathematics, 60, 2009–2028.

    Article  Google Scholar 

  • Shelley, M., McLaughlin, D., Shapley, R., Wielaard, J. (2002). States of high conductance in a large-scale model of the visual cortex. Journal of Computational Neuroscience, 13, 93–109.

    Article  PubMed  Google Scholar 

  • Shew, S., Yang, H., Yu, S., Roy, R., Plenz, D. (2011). Information capacity and transmission are maximized in balanced cortical networks with neuronal avalanches. The Journal of Neuroscience, 31, 55–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Stern, E.A., Kincaid, A.E., Wilson, C.J. (1997). Spontaneous subthreshold membrane potential fluctuations and action potential variability of rat corticostriatal and striatal neurons in vivo. Journal of Neurophysiology, 77, 1697–1715.

    Article  CAS  PubMed  Google Scholar 

  • Storch, H., & Zwiers, F.W. (2001). Statistical analysis in climate research. Cambridge University Press.

  • Touboul, J. (2014). Propagation of chaos in neural fields. Annals of Applied Probability, 24, 1298–1328.

    Article  Google Scholar 

  • Traub, R.D., Jeffreys, J., Whittington, M. (1999). Fast oscillations in cortical circuits. Cambridge: MIT Press.

    Book  Google Scholar 

  • Vogels, T.P., & Abbott, L.F. (2005). Signal propagation and logic gating in networks of integrate-and-fire neurons. The Journal of Neuroscience, 25, 10786–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Werner, G. (2007). Metastability, criticality and phase transitions in brain and its models. BioSystems, 90, 496–508.

    Article  PubMed  Google Scholar 

  • Xiao, Z.C., Zhang, J.W., Sornborger, A.T., Tao, L. (2308). Cusps enable line attractors for neural computation. Physical Review E, 96(05), 2017.

    Google Scholar 

  • Yu, Y., & Ferster, D. (2010). Membrane potential synchrony in primary visual cortex during sensory stimulation. Neuron, 68, 1187–1201.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Yu, S., Yang, H., Nakahara, H., Santos, G.S., Nikolic, D., Plenz, D. (2011). Higher-order interactions characterized in cortical activity. The Journal of Neuroscience, 31, 17514–17526.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  • Zerlaut, Y., Chemla, S., Chavane, F., Destexhe, A. (2018). Modeling mesoscopic cortical dynamics using a mean-field model of conductance-based networks of adaptive exponential integrate-and-fire neurons. Journal of Computational Neuroscience, 44, 45–61.

    Article  PubMed  Google Scholar 

  • Zhang, J., & Rangan, A.V. (2015). A reduction for spiking integrate-and-fire network dynamics ranging from homogeneity to synchrony. Journal of Computational Neuroscience, 38(2), 355–404.

    Article  CAS  PubMed  Google Scholar 

  • Zhang, J.W., Zhou, D., Cai, D., Rangan, A.V. (2014a). A coarse-grained framework for spiking neuronal networks: between homogeneity and synchrony. Journal of Computational Neuroscience, 37(1), 81–104.

    Article  PubMed  Google Scholar 

  • Zhang, J.W., Newhall, K., Zhou, D., Rangan, A.V. (2014b). Distribution of correlated spiking events in a population-based approach for integrate-and-fire networks. Journal of Computational Neuroscience, 36(2), 279–295.

    Article  PubMed  Google Scholar 

  • Zhao, L.Q., Beverlin, B., Netoff, T., Nykamp, D.Q. (2011). Synchronization from second order network connectivity statistics. Frontiers in Computational Neuroscience, 5(28).

Download references

Acknowledgments

This work was partially supported by the Natural Science Foundation of China through grants 11771035 (J.Z.), 91430216 (J.Z.), U1530401 (J.Z.), 31771147 (Y.S., L.T.) and 91232715 (L.T.), by the Open Research Fund of the State Key Laboratory of Cognitive Neuroscience and Learning grant CNLZD1404 (Y.S., L.T.), and by the Beijing Municipal Science andTechnology Commission under contract Z151100000915070 (Y.S., L.T.).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Louis Tao.

Ethics declarations

Conflict of interests

The authors declare that they have no conflict of interest.

Additional information

Action Editor: Alain Destexhe

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: spike resolution

An important feature of the neuronal network we considered in this paper is that the synapses are instantaneous. In other words, the postsynaptic impulse response functions are taken to be δ-functions with no synaptic delay. This δ-function impulse response allows us to determine whether or not one neuron ‘caused’ another neuron to fire, i.e., the instantaneous synapses ensure that cascade-induced firing-events transpire instantaneously. Thus the feedforward inputs will not be confused with the network spikes. To clarify the dynamics that occurs whenever multiple neurons cross threshold simultaneously at time tspk, we consider our system to be a specific limit of the following fast conductance system

$$\begin{array}{@{}rcl@{}} \frac{d{V}^{{Q}}_{i}(t)}{dt}&=&- g^{E} \left( V^{{Q}}_{i}-{V}_{E}\right) -g^{I} \left( {V}^{{Q}}_{i}-{V}_{{I}}\right),\\ \tau_{{E}} \frac{d g^{{E}}(t)}{dt} &=& - g^{{E}} + \sum\limits_{k}{S}^{{QE}} \delta \left( t-t_{jk}^{{E}}\right), \\ \tau_{I}\frac{dg^{{I}}(t)}{dt} &=& - g^{{I}} +\sum\limits_{k}{S}^{{QI}} \delta \left( t-t_{jk}^{{I}}\right), \end{array} $$
(34)

where the gQ represent synaptic conductances, and the decay times τEτI → 0. Now at the time tspk, we freeze the macroscopic time t of the original system and create an “infinitesimal-time” system which evolves according to an infinitesimal time τQ. Since the MFE is initiated by one excitatory neuron firing before others, at tspk the state-variables of the original system are given by

$$ {V}_{j}^{{Q}}(t^{-}_{\text{{spk}}}) \text{ for } j = 1,{\cdot}s, {N}_{{Q}} \text{ with } {V}_{1}^{{E}}(t_{\text{{spk}}}^{-}) \geq {V}_{{T}}. $$
(35)

This infinitesimal-time system will use the infinitesimal time τ to describe the effects of the excitatory conductances in Eq. (36) by taking τEτI → 0. This is, at this macro-time tspk, we solve the dynamics of the infinitesimal system

$$\begin{array}{@{}rcl@{}} \frac{d{V}^{{Q}}_{i}(\tau)}{d\tau}&=&- g^{{E}} \left( V^{{Q}}_{i}-{V}_{{E}}\right) -g^{{I}} \left( V^{{Q}}_{i}-{V}_{{I}}\right),\\ \frac{d g^{{E}}(\tau)}{d\tau} &=& - g^{{E}} + \sum\limits_{j\neq i}{S}^{{QE}} \delta \left( \tau-\tau_{j}^{{E}}\right) \\ \frac{dg^{{I}}(\tau)}{d\tau} &=& - g^{{I}} +\sum\limits_{j\neq i}{S}^{{QI}} \delta \left( \tau-\tau_{jk}^{{I}}\right), \end{array} $$
(36)

by fixing τE = τI = 1, letting the micro-time τ, and using the same conditions as Eq. (35). The spike-times \(\tau _{j}^{{Q}}\) record the infinitesimal time at which neurons j fires. By denoting r = gI/gE and g(τ) = gEeτ, without ambiguity, we delete the subscript for a while, and rewrite the first equation in (36) as

$$\frac{dv}{d\tau} = -g(\tau) (1+r)(v-{V}_{s}), $$

where Vs = (VE + rVI)/(1 + r). Noteing that \(\frac {dv}{d \tau } = \frac {dv}{dg}\frac {dg}{d\tau }\), we arrive at

$$\frac{dv}{dg} = (1+r)(v -{V}_{s}) $$

with the solution v(τ) = Vs + gE(V0Vs)e1+r(1 − eτ). The voltage will evolve to a steady state of

$$\max(v) = \lim\limits_{\tau\rightarrow \infty}v(\tau) = {V}_{s} + g^{{E}}({V}_{0}-{V}_{s}){e}^{1+r}. $$

If max(v) ≥ VT, the neuron will fire at time

$$\tau_{j}^{{Q}} = \tau^{\prime} - \log \left[1 + \frac{1}{(1+r)g^{{E}}} \log \frac {{V}_{{T}}-{V}_{s}}{{V}_{0}-{V}_{s}}\right], $$

where τ is the time that the last previous neuron fires. By using this analytical formula, we can resolve the trajectory to machine precision during this infinitesimal time until the time \(\tau = \min \limits _{j,{ Q}} \tau _{j}^{{Q}}\). Once a single neuron fires, it is clamped at VR for a refractory period, and never fires again during this infinitesimal time. On the other hand, we can update the voltages of the other neurons that do not fire at this macro-time tspk. After this infinitesimal time, we go back to the macro-time t to evolve the system. Thus we obtain a well-posed system of ODEs. One can see other similar implementation in Appendix B of Ref. Rangan and Young (2013a) and Appendix 12 of Ref. Zhang and Rangan (2015).

Appendix B: derivation of the Master equation f B, single

In this section, the Master equation fB, single is used to evolve the single-neuron distributions \(\boldsymbol {{\rho }}_{{{ single}}}\left (v,t\right ) = ({\rho }^{{E}}_{{{single}}}, {\rho }^{{I}}_{{{single}}})\). Over the small voltage interval [V1, V2], the total probability changes as a function of time. The difference from time t1 to t1 + Δt is determined by the probability flux at the boundaries V1 and V2

$${\int}_{v={V}_{1}}^{v={V}_{2}}\left[{\rho}^{{Q}}\left( v,t\right)\right]_{t=t_{1}}^{t=t_{1}+{\Delta} t}dv=f_{{{B, single}}}^{{Q}}\left( {\rho}^{{Q}},v,t\right) {,} $$

where fB, single = (fB, singleE, fB, singleI) is given by

$$\begin{array}{@{}rcl@{}} && f^{{Q}}_{{B},{{single}}}\left( {\rho}^{{Q}},v,t\right) ={\int}_{t_{1}}^{t_{1}+{\Delta} t}\left[{g}_{{L}}\left( v-{V}_{{L}}\right) {\rho}^{{Q}}\left( v,t\right) \right]_{v={V}_{1}}^{v={V}_{2}}dt \\ &&\quad\quad+\sum\limits_{{R}\in \left\{ E,I,Y\right\}}{\int}_{t_{1}}^{t_{1}+{\Delta} t}{\int}_{T({V}_{1},{S}^{{QR}})}^{T({V}_{2},{S}^{{QR}})}m_{{{ net}}}^{{QR}}\left( t\right) {\cdot} {\rho}^{{Q}}\left( v,t\right) dvdt \\ &&\quad\quad-\sum\limits_{{R}\in \left\{ { E,I,Y}\right\}}{\int}_{t_{1}}^{t_{1}+{\Delta} t}{\int}_{v={V}_{1}}^{v={V}_{2}}m_{{{net}}}^{{QR}}\left( t\right) {\cdot} \mathbf{1}_{\left[-\infty ,T({V}_{T},{S}^{{QR}})\right] }\left( v\right) {\cdot} {\rho}^{{Q}}\left( v,t\right) dvdt \\ &&\quad\quad-{\int}_{t_{1}}^{t_{1}+{\Delta} t}{\int}_{v={V}_{1}}^{v={V}_{2}}m_{{{net}}}^{{QY}}\left( t\right) {\cdot} \mathbf{{\gamma}}^{{Q}}\left( t\right) {\cdot} \mathbf{1}_{\left[T({V}_{{T}},{S}^{{QY}}),{V}_{{T}}\right] }\left( v\right) {\cdot} {\rho}^{{Q}}\left( v,t\right) dvdt \\ &&\quad\quad+\mathbf{1}_{\left[{V}_{1},{V}_{2}\right] }\left( {V}_{{L}}\right) {\int}_{t_{1}}^{t_{1}+{\Delta} t}{\int}_{T({V}_{{T}},{S}^{{EY}})}^{{V}_{{T}}}m_{{{net}}}^{{QY}}\left( t\right) {\cdot} \mathbf{{\gamma}}^{{Q}}\left( t\right) {\cdot} {\rho}^{{Q}}\left( v,t\right) dvdt{,} \end{array} $$
(37)

where the fraction γI, and γE are used to condition fB, single not to produce MFEs. The motivation to introduce γI and γE can be found in Zhang et al. (2014a). Generally, γI ≡ 1 and γE is given by the fraction of time that:

$$\begin{array}{@{}rcl@{}} {\gamma}^{E}\left( t^{+}\right) &= & \left[{\int}_{- \infty}^{T({V}_{{T}},{S}^{{EE}})}{\rho}^{{E}}_{{{single}}}\left( v,t^{-}\right) dv\right]^{{N}_{{ E}}-1}\\&&\times\left[{\int}_{-\infty}^{T({V}_{{T}},{S}^{{IE}})}{\rho}^{{I}}_{{{single}}}\left( v,t^{-}\right) dv\right]^{{N}_{{I}}}{.} \end{array} $$

The network firing rates are given by:

$$\begin{array}{@{}rcl@{}} m_{{net}}^{{QE}}\left( t^{+}\right) &=&\left( {N}_{{Q}}-\delta_{{QE}}\right) m_{{{single}}}^{{E}}\left( t^{-}\right),\\ \text{and } m_{{{ net}}}^{{QI}}\left( t^{+}\right) &=&\left( {N}_{{Q}}-\delta_{{QI}}\right) m_{{{single}}}^{{I}}\left( t^{-}\right) {,} \end{array} $$
(38)

where the single-neuron firing rates msingleQ (t) represents the instantaneous firing rate of a single neuron in an ensemble driven by the various mnetQR (t), and is calculated by the flux of ρQ (v, t) over the threshold point VT at time t in form of

$$\begin{array}{@{}rcl@{}} m_{{{single}}}^{{Q}}\left( t^{-}\right) = {\int}_{t_{1}}^{t_{1}+{\Delta} t}{\int}_{T({V}_{{T}},{S}^{{EY}})}^{{V}_{{T}}}m_{{{net}}}^{{QY}}\left( t\right) {\cdot} \mathbf{{\gamma}}^{{Q}}\left( t\right) {\cdot} {\rho}^{{Q}}\left( v,t\right). \end{array} $$

Appendix C: derivation of the standard Master equation

Here we derive the Master equation from system of conductance-based IF neuronal networks (1). We start by considering what can happen during a single time-interval of length Δt? The coarsest approximation involves choosing from amongst 5 possibilities: (i) the neuron starts out held in refractory and remains at VL, (ii) the neuron starts out in (VI, VT) and decays naturally, (iii) the neuron starts out in (VI, VT) and gets a excitatory kick by the external input, or (iv,v) the neuron starts out in (VI, VT) and gets a kick of type-Q with synaptic coupling. The simplest possibility is (i), which can be treated by a straightforward computation. In possibility (ii), the pure decay case, v gets mapped to

$$\begin{array}{@{}rcl@{}} { R}_{{L}}^{-1}\left( v,{g}_{{L}}{\Delta} t\right) ={V}_{{L}}+\left( v-{V}_{{L}}\right){\cdot} {e}^{{g}_{{L}}{\Delta} t}={T}_{{L}}\left( v,{g}_{{ L}}{\Delta} t\right). \end{array} $$

Thus, for a small dv, the distribution ρ (v, t)dv within the interval [v, v + dv] is replaced by

$$\begin{array}{@{}rcl@{}} {\rho} \left( {T}_{{L}}\left( v,{g}_{{L}}{\Delta} t\right),t\right)\!{\cdot}\! {T}_{{L}}^{\prime }\left( v,{g}_{{L}}{\Delta} t\right) dv = {\rho}\left( {V}_{{L}} + \left( v - {V}_{{L}}\right){\cdot} {e}^{{g}_{{L}}{\Delta} t},t\right)\!{\cdot}\! {e}^{{g}_{{L}}{\Delta} t}dv. \end{array} $$
(39)

Now let us discuss possibilities (iii,iv,v) which correspond to getting a kick from the external input and other excitatory or inhibitory neurons in the network. For example, given a single kick of the external Poisson input, the distribution concentrated in the interval [V1, V2]gets mapped to the interval [R (V1, SQY) , R (V2, SQY)], and is mapped into by the interval [T (V1, SQY) , T (V2, SQY)]. Thus, if V1 = v, and V2 = v + dv for dv small, then ρ (v)dv is replaced by ρ (T (v, SQY) , t)⋅T(v, SQY) dv, implying that, for each voltage v, the distribution ρ (v) is replaced by the distribution \({\rho } \left ({T}\left (v,{S}^{{QY}}\right ),t\right ) {\cdot } {T}^{\prime }\left (v,{S}^{{QY}}\right ) ={\rho } \left ({V}_{{E}}+\left ({V}-{V}_{{E}}\right ){\cdot } {e}^{{S}^{{QY}}},t\right ) {\cdot } {e}^{{S}^{{QY}}}\). Similarly, we have corresponding maps for excitatory and inhibitory network spikes.

Thus, combining these 5 possibilities, ρ (v, t) evolves according to

$$\begin{array}{@{}rcl@{}} {\rho}^{{Q}} \left( v,t+{\Delta} t\right) &=&\left[1-({m}^{{QY}}_{{net}} + {N}_{{E}}{m}^{{E}} + {N}_{{I}}{m}^{I}){\Delta} t\right] {\cdot} {\rho}^{{Q}} \left( {{V}}_{{L}}+\left( {V}-{V}_{{L}}\right) {\cdot} {e}^{{g}_{{L}}{\Delta} t},t\right) {\cdot} {e}^{{g}_{{L}}{\Delta} t} \\ &&+{m}^{{QY}}_{{net}}(t){\Delta} t{\cdot} {\rho}^{{Q}} \left( {V}_{{E}}+\left( v-{V}_{{E}}\right){\cdot} {e}^{{S}^{{QY}}},t\right) {\cdot} {e}^{{S}^{{QY}}} \\ &&+{N}_{{E}}{m}^{{E}}(t){\Delta} t{\cdot} {\rho}^{{Q}} \left( {V}_{{E}}+\left( v-{V}_{{E}}\right){\cdot} {e}^{{S}^{{QE}}},t\right) {\cdot} {e}^{{S}^{{QE}}} \\ &&+{N}_{{I}}{m}^{{I}}(t){\Delta} t{\cdot} {\rho}^{{Q}} \left( {V}_{{I}}+\left( v-{V}_{{I}}\right) {\cdot} {e}^{{S}^{{QI}}},t\right) {\cdot} {e}^{{S}^{{QI}}}+\mathcal{{S}}^{{Q}}\left( t\right) {\cdot} \delta\left( {V}-{V}_{{R}}\right) \end{array} $$
(40)

with Q ∈{E, I}. Taylor expanding,

$$\begin{array}{@{}rcl@{}} {\rho}^{{Q}} \left( {V}_{{L}}+\left( {V}-{V}_{{L}}\right) {\cdot} {e}^{{g}_{{L}}{\Delta} t},t\right){\cdot} {e}^{{g}_{{L}}{\Delta} t,t} = {\rho}^{{Q}}(v,t) + {g}_{{L}}{\Delta} t {\partial}_{v}[(v-{V}_{{L}}){\rho}^{{Q}}(v,t)] + o({\Delta} t). \end{array} $$

Substituting the above into Eq. (40) and ignoring the terms of order ot) and higher, we arrive at

$$\begin{array}{@{}rcl@{}} {\rho}^{{Q}} \left( v,t+{\Delta} t\right) - {\rho}^{{Q}}(v,t) &=&{g}_{L}{\Delta} t {\partial}_{v} \left[(v-{V}_{{L}}){\rho}^{{Q}}(v,t)\right] +{m}^{{QY}}_{{net}}(t){\Delta} t {\cdot} \left[{\rho}^{{Q }}\left( T(v,{S}^{{QY}}),t\right) {\cdot} {T}^{\prime}(v,{S}^{{QY}}) - {\rho}^{{ Q}}(v,t)\right] \\ &&+{N}_{{E}}{m}^{{E}}(t){\Delta} t{\cdot} \left[{\rho}^{{Q }}\left( T(v,{S}^{{QE}}),t\right){\cdot} T^{\prime}(v,{S}^{{QE}}) - {\rho}^{{Q}}(v,t)\right] \\ &&+{N}_{{I}}{m}^{{I}}(t){\Delta} t {\cdot} \left[{\rho}^{{Q}} \left( T(v,{S}^{{QI}}),t\right) {\cdot} {T}^{\prime}(v,{S}^{{QI}}) - {\rho}^{{Q}}(v,t)\right] +\mathcal{{S}}^{{Q}}\left( t\right) {\cdot} \delta\left( {V}-{V}_{{R}}\right), \end{array} $$
(41)

where \(\mathcal {S}^{Q}(t)\) is the flux across the threshold and is given by

$$\begin{array}{@{}rcl@{}} \mathcal{S}^{Q}\left( t\right) &=&{m}^{QY}_{net}(t) {\int}_{T({V}_{T},{S}^{QY})}^{{V}_{T}}{\rho}^{Q}(v,t)dv +{N}_{E}{m}^{E}(t){\int}_{T({V}_{T},{S}^{QE})}^{{V}_{T}} {\rho}^{Q}(v,t)dv. \end{array} $$
(42)

Let Δt → 0, we have the Master equation in the following form

$$\begin{array}{@{}rcl@{}} \frac{d{\rho}^{Q} \left( v,t\right)}{dt} &=&{g}_{L}{\partial}_{v} \left[(v-{V}_{L}){\rho}^{Q}(v,t)\right] +{m}^{QY}_{net}(t) {\cdot} \left[{\rho}^{Q} \left( T(v,{S}^{QY}),t\right) {\cdot} T^{\prime}(v,{S}^{QY}) - {\rho}^{Q}(v,t)\right] \\ &&+{N}_{E}{m}^{E}(t){\cdot} \left[{\rho}^{Q} \left( T(v,{S}^{QE}),t\right){\cdot} T^{\prime}(v,{S}^{QE}) - {\rho}^{Q}(v,t)\right] \\ &&+{N}_{I}{m}^{I}(t) {\cdot} \left[{\rho}^{Q} \left( T(v,{S}^{QI}),t\right) {\cdot} T^{\prime}(v,{S}^{QI}) - {\rho}^{Q}(v,t)\right]+\mathcal{S}^{Q}\left( t\right) {\cdot} \delta\left( V-{V}_{R}\right). \end{array} $$
(43)

For a similar derivation of Eq. (43), see Nykamp (2000) and Cai et al. (2006).

Appendix D: derivation of Fokker-Planck equation

Similar to the reductions performed in Rangan and Cai (2006) and Helias et al. (2010), we take a single neuron driven only by the external Poisson input for example. More complicated case can be computed similarly. Let \(s ={e}^{{S}^{{QY}}} - 1\), we can rewrite

$$T\left( v,{S}^{{QY}}\right) = v+s(v-{V}_{{E}}), \text{and} T^{\prime}(v,{S}^{{QY}})= 1+s. $$

Ignoring the network synaptic connection for now, and considering only the external Poisson input, thus the Master Eq. (8) is given by

$$\begin{array}{@{}rcl@{}} {\partial}_{t} {\rho}^{{Q}}(v,t) & = & {g}_{{L}}{\partial}_{v} \left[(v - {{V}}_{{L}}){\rho}^{{Q}}(v,t)\right] + {m}^{{QY}}_{{net}}(t) \\&&{\cdot} \left[{\rho}^{{Q}} \left( v + (v - {V}_{{E}})s,t\right)(1 + s) - {\rho}^{{Q}}(v,t)\right]. \end{array} $$
(44)

Taylor expanding and keeping the first order in s, we get

$$ {\rho}^{{Q}} (v+(v-{{V}}_{{E}})s)\approx {\rho}^{{Q}}(v) + (v-{V}_{{E}})s{\partial}_{v}{\rho}^{{Q}} + \frac{(v-{V}_{{ E}})^{2}{S}^{2}}{2}{\partial}_{vv}{\rho}^{{Q}}, $$
(45)

thus we can approximate (44) by

$$\begin{array}{@{}rcl@{}} {\partial}_{t} {\rho}^{{Q}}(v,t) &=& {g}_{{L}}{\partial}_{v} \left[(v-{{V}}_{{L}}){\rho}^{{Q}}(v,t)\right] \\&&+{m}^{{QY}}_{{net}}(t){\cdot} \left[(v-{V}_{{E}})s(1+s){\partial}_{v}{\rho}^{{Q}}+ \frac{(v-{V}_{{E}})^{2}{S}^{2}(1+s)}{2}{\partial}_{vv}{\rho}^{{Q}} + s{\rho}^{{Q}}\right]. \end{array} $$
(46)

Noting the relationships

$$\begin{array}{@{}rcl@{}} (v-{V}_{E})s{\partial}_{v}{\rho}^{{Q}} &= &{\partial}_{v} \left[s(v-{V}_{{E}}){\rho}^{{Q}}\right] -s{\rho}^{{Q}}, \\ \frac{(v-{V}_{{E}})^{2}{S}^{2}}{2}{\partial}_{vv}{\rho}^{{Q}} &=&{\partial}_{v}\left[\frac{(v-{V}_{{E}})^{2}{S}^{2}}{2}{\partial}_{v}{\rho}^{{Q}}\right] \\&&- {\partial}_{v}\left[{S}^{2}(v-{{V}}_{{E}}){\rho}^{{Q}}\right] + {S}^{2}{\rho}^{{Q}}, \end{array} $$

we have

$$\begin{array}{@{}rcl@{}} &&(v-{V}_{{E}})s{\partial}_{v}{\rho}^{{Q}} + \frac{(v-{V}_{{E}})^{2}{S}^{2}}{2}{\partial}_{vv}{\rho}^{{Q}} \\&=&{\partial}_{v}\left[(s-{S}^{2})(v-{V}_{{E}}){\rho}^{{Q}} + \frac{(v-{V}_{{E}})^{2}{S}^{2}}{2}{\partial}_{v}{\rho}^{{Q}}\right] - (s-{S}^{2}){\rho}^{{Q}}. \end{array} $$
(47)

Substituting (47) into (46), we arrive at the Fokker-Planck equation

$$\begin{array}{@{}rcl@{}} {\partial}_{t} {\rho}^{{Q}}(v,t) &= &{\partial}_{v} \{\left[{g}_{{L}}(v-{V}_{{L}}) + {m}^{{QY}}_{{net}}(t) (s-{S}^{3})(v-{V}_{{E}})\right]{\rho}^{{Q}}(v,t) \} \\ && +{\partial}_{v}\{\frac{{m}^{{QY}}_{{net}}(t)(1+s){S}^{2}(v-{V}_{{E}})^{2}}{2}{\partial}_{v}{\rho}^{{Q}} \} + {m}^{{QY}}_{{net}}(t){S}^{3}{\rho}^{{Q}}. \end{array} $$
(48)

We remark that the last term \({m}^{{QY}}_{net}(t){S}^{3}{\rho }^{{Q}}\) in Eq. (48) has a small influence on the distribution when the parameter s is small. Hence, if we only keep the second order of s in the Taylor’s expansion of Eq. (45), we can ignore this term in most practical analysis. Similarly, for the full IF network, we let \(s = {e}^{{S}^{{QY}}} - 1, {s}_{1} = {e}^{{S}^{{QE}}} - 1, {s}_{2} = {e}^{{S}^{{QI}}} - 1\), and can approximate the Master Eq. (8) by a FP-type Eq. (9).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, J., Shao, Y., Rangan, A.V. et al. A coarse-graining framework for spiking neuronal networks: from strongly-coupled conductance-based integrate-and-fire neurons to augmented systems of ODEs. J Comput Neurosci 46, 211–232 (2019). https://doi.org/10.1007/s10827-019-00712-w

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-019-00712-w

Keywords

Navigation