Skip to main content
Log in

Episodic activity in a heterogeneous excitatory network, from spiking neurons to mean field

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

Abstract

Many developing neural systems exhibit spontaneous activity (O’Donovan, Curr Opin Neurobiol 9:94–104, 1999; Feller, Neuron 22:653–656, 1999) characterized by episodes of discharge (active phases) when many cells are firing, separated by silent phases during which few cells fire. Various models exhibit features of episodic behavior by means of recurrent excitation for supporting an episode and slow activity-dependent depression for terminating one. The basic mechanism has been analyzed using mean-field, firing-rate models. Firing-rate models are typically formulated ad hoc, not derived from a spiking network description, and the effects of substantial heterogeneity amongst the units are not usually considered. Here we develop an excitatory network of spiking neurons (N-cell model) with slow synaptic depression to model episodic rhythmogenesis. This N-cell model displays episodic behavior over a range of heterogeneity in bias currents. Important features of the episodic behavior include orderly recruitment of individual cells during silent phases and existence of a dynamical process whereby a small critical subpopulation of intermediate excitability conveys synaptic drive from active to silent cells. We also derive a general self-consistency equation for synaptic drive that includes cell heterogeneity explicitly. We use this mean-field description to expose the dynamical bistability that underlies episodic behavior in the heterogeneous network. In a systematic numerical study we find that the robustness of the episodic behavior improves with increasing heterogeneity. Furthermore, the heterogeneity of depression variables (imparted by the heterogeneity in cellular firing thresholds) plays an important role in this improvement: it renders the network episodic behavior more robust to variations in excitability than if depression is uniformized. We also investigate the effects of noise vs. heterogeneity on the robustness of episodic behavior, especially important for the developing nervous system. We demonstrate that noise-induced episodes are very fragile, whereas heterogeneity-produced episodic rhythm is robust.

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
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

References

  • Beggs, J. M., & Plenz, D. (2003). Neuronal avalanches in neocortical circuits. Journal of Neuroscience, 23, 11167–11177.

    PubMed  CAS  Google Scholar 

  • Ben-Ari, Y. (2001). Developing networks play a similar melody. Trends in Neuroscience, 24, 353–360.

    Article  CAS  Google Scholar 

  • Borodinsky, L. N., Root, C. M., Cronin, J., Sann, S. B., Gu, X., & Spitzer, N. C. (2004). Activity-dependent homeostatic specification of transmitter expression in embryonic neurons. Nature, 429, 523–530.

    Article  PubMed  CAS  Google Scholar 

  • Burden, R. L., & Faires, J. D. (2001). Numerical analysis (7th ed.). Pacific Grove, CA: Brooks/Cole.

    Google Scholar 

  • Cherubini, E., Gaiarsa, J. L., & Ben-Ari, Y. (1991). GABA: an excitatory transmitter in early postnatal life. Trends in Neuroscience, 14, 515–519.

    Article  CAS  Google Scholar 

  • Chub, N., & O’Donovan, M. J. (1998). Blockade and recovery of spontaneous rhythmic activity after application of neurotransmitter antagonists to spinal networks of the chick embryo. Journal of Neuroscience, 18, 294–306.

    PubMed  CAS  Google Scholar 

  • Compte, A., Sanchez-Vives, M. V., McCormick, D. A., & Wang, X. J. (2003). Cellular and network mechanisms of slow oscillatory activity (<1 Hz) and wave propagations in a cortical network model. Journal of Neurophysiology, 89, 2707–2725.

    Article  PubMed  Google Scholar 

  • Ermentrout, B. (1994). Reduction of conductance based models with slow synapses to neural nets. Neural Computation, 6, 679–695.

    Article  Google Scholar 

  • Fedirchuk, B., Wenner, P., Whelan, P. J., Ho, S., Tabak, J., & O’Donovan, M. J. (1999). Spontaneous network activity transiently depresses synaptic transmission in the embryonic chick spinal cord. Journal of Neuroscience, 19, 2102–2112.

    PubMed  CAS  Google Scholar 

  • Feller, M. B. (1999). Spontaneous correlated activity in developing neural circuits. Neuron, 22, 653–656.

    Article  PubMed  CAS  Google Scholar 

  • Giugliano, M., Darbon, P., Arsiero, M., Lüscher, H.-R., & Streit, J. (2004). Single-neuron discharge properties and network activity in dissociated cultures of neocortex. Journal of Neurophysiology, 92, 977–996.

    Article  PubMed  CAS  Google Scholar 

  • Hansel, D., Mato, G., Meunier, C., & Neltner, L. (1998). On numerical simulations of integrate-and-fire neural networks. Neural Computation, 10, 467–483.

    Article  PubMed  CAS  Google Scholar 

  • Hanson, M. G., & Landmesser, L. T. (2004). Normal patterns of spontaneous activity are required for correct motor axon guidance and the expression of specific guidance molecules. Neuron, 43, 687–701.

    Article  PubMed  CAS  Google Scholar 

  • Katz, L. C., & Shatz, C. J. (1996). Synaptic activity and the construction of cortical circuits. Science, 274, 1133–1138.

    Article  PubMed  CAS  Google Scholar 

  • Latham, P. E., Richmond, B. J., Nelson, P. G., & Nirenberg, S. (2000). Intrinsic dynamics in neuronal networks. I. Theory. Journal of Neurophysiology, 83, 808–827.

    PubMed  CAS  Google Scholar 

  • Loebel, A., & Tsodyks, M. (2002). Computation by ensemble synchronization in recurrent networks with synaptic depression. Journal of Computational Neuroscience, 13, 111–124.

    Article  PubMed  Google Scholar 

  • Marchetti, C., Tabak, J., Chub, N., O’Donovan, M. J., & Rinzel, J. (2005). Modeling spontaneous activity in the developing spinal cord using activity-dependent variations of intracellular chloride. Journal of Neuroscience, 25, 3601–3612.

    Article  PubMed  CAS  Google Scholar 

  • O’Donovan, M. J. (1999). The origin of spontaneous activity in developing networks of the vertebrate nervous system. Current Opinions in Neurobiology, 9, 94–104.

    Article  CAS  Google Scholar 

  • Sernagor, E., Chub, N., Ritter, A., & O’Donovan, M. J. (1995). Pharmacological characterization of the rhythmic synaptic drive onto lumbosacral motoneurons in the chick embryo spinal cord. Journal of Neuroscience, 15, 7452–7464.

    PubMed  CAS  Google Scholar 

  • Shelley, M. J., & Tao, L. (2001). Efficient and accurate time-stepping schemes for integrate-and-fire neuronal networks. Journal of Computational Neuroscience, 11, 111–119.

    Article  PubMed  CAS  Google Scholar 

  • Shriki, O., Hansel, D., & Sompolinsky, H. (2003). Rate models for conductance-based cortical neuronal networks. Neural Computation, 15, 1809–1841.

    Article  PubMed  Google Scholar 

  • Stellwagen, D., & Shatz, C. J. (2002). An instructive role for retinal waves in the development of retinogeniculate connectivity. Neuron, 33, 357–367.

    Article  PubMed  CAS  Google Scholar 

  • Strogatz, S. H. (2000). Nonlinear dynamics and chaos: With applications to physics, biology, chemistry, and engineering. Cambridge, MA: Perseus.

    Google Scholar 

  • Tabak, J., O’Donovan, M. J., & Rinzel, J. (2006). Differential control of active and silent phases in relaxation models of neuronal rhythms. Journal of Computational Neuroscience, 21, 307–328.

    Article  PubMed  Google Scholar 

  • Tabak, J., Rinzel, J., & O’Donovan, M. J. (2001). The role of activity-dependent network depression in the expression and self-regulation of spontaneous activity in the developing spinal cord. Journal of Neuroscience, 21, 8966–8978.

    PubMed  CAS  Google Scholar 

  • Tabak, J., Senn, W., O’Donovan, M. J., & Rinzel, J. (2000). Modeling of spontaneous activity in developing spinal cord using activity-dependent depression in an excitatory network. Journal of Neuroscience, 20, 3041–3056.

    PubMed  CAS  Google Scholar 

  • Timofeev, I., Grenier, F., Bazhenov, M., Sejnowski, T. J., & Steriade, M. (2000). Origin of slow cortical oscillations in deafferented cortical slabs. Cerebral Cortex, 10, 1185–1199.

    Article  PubMed  CAS  Google Scholar 

  • Tsodyks, M., Uziel, A., & Markram, H. (2000). Synchrony generation in recurrent networks with frequency-dependent synapses. Journal of Neuroscience, 20(RC50), 1–5.

    Google Scholar 

  • van Vreeswijk, C., & Hansel, D. (2001). Patterns of synchrony in neural networks with spike adaptation. Neural Computation, 13, 959–992.

    Article  PubMed  Google Scholar 

  • Vladimirski, B. B. (2005). Modeling and analysis of spontaneous electrical episodic activity in the developing nervous system by means of a heterogeneous excitatory network of spiking neurons with slow synaptic depression and reduced mean-field models. Ph.D. diss., New York University, New York, NY.

  • Wenner, P., & O’Donovan, M. (2001). Mechanisms that initiate spontaneous network activity in the developing chick spinal cord. Journal of Neurophysiology, 86, 1481–1498.

    PubMed  CAS  Google Scholar 

  • Wiedemann, U. A., & Lüthi, A. (2003). Timing of network synchronization by refractory mechanisms. Journal of Neurophysiology, 90, 3902–3911.

    Article  PubMed  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  • Yvon, C., Czarnecki, A., & Streit, J. (2007). Riluzole-induced oscillations in spinal networks. Journal of Neurophysiology, 97, 3607–3620.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

J. Rinzel was funded in part by NIMH (MH62595-01). M. O’Donovan and B. Vladimirski were funded in part by the intramural program of NINDS, NIH.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Boris B. Vladimirski.

Additional information

Action Editor: Nicolas Brunel

Electronic Supplementary Material

Below is the link to the electronic supplementary material.

(PDF 375 kb)

Appendix

Appendix

1.1 Details of the mean-field description derivation

To obtain the expression for \(\hat{q}(g_{syn},\:I)\) given in (11), we perform the following steps.

  1. 1.

    We solve Eq. (6) analytically assuming that synapses are activated for a short period of time relative to the interspike interval, i.e., ε q  ≤ T(g syn , I). Using the integrating factor method, we obtain

    $$q_i(t)\!=\!\!\left\{ \begin{array}{ll} {\frac{\alpha_q}{\alpha_q+\beta_q}}\!+\!\left(\! q_i(0)\!-\!{\frac{\alpha_q}{\alpha_q+\beta_q}} \!\right)\! e^{-(\alpha_q+\beta_q)t} & \ 0 \!\leq \!t \!\leq \!\epsilon_q \\ e^{-\beta_q(t-\epsilon_q)}q_i(\epsilon_q) & \ \epsilon_q \!<\! t\! \leq\! T_i \!\equiv \!T(g_{syn},\:I_i) \end{array} \right.$$
    (19)

    Since T i is the period of neuron i, we must have

    $$\begin{array}{ll} q_i(T_i) & = q_i(0) \\ q_i(0) & = {\frac{\alpha_q}{\alpha_q+\beta_q}}\frac{e^{\beta_q\epsilon_q}-e^{-\alpha_q\epsilon_q}}{e^{\beta_q T_i}-e^{-\alpha_q\epsilon_q}} \end{array}$$
  2. 2.

    Find the temporal average of the analytical solution:

    $$\hat{q}(g_{syn},\:I_i)=\!\!\left\{ \begin{array}{ll} \!0, \mbox{ } T_i\!=\!+\infty \mbox{; otherwise:} & \ \nonumber\\ \!1/T_i \int_{0}^{T_i}q_i(t)dt \!=\!1/T_i \!\left( \int_{0}^{\epsilon_q} q_i(t)dt\!+\!\int_{\epsilon_q}^{T_i} \!q_i(t)dt \right) & \ \end{array} \right.$$

    For T i finite,

    $$\int_{0}^{\epsilon_q} q_i(t)dt = {\frac{\alpha_q}{\alpha_q+\beta_q}}\epsilon_q +\!\left(\! q_i(0)\!-\!{\frac{\alpha_q}{\alpha_q+\beta_q}} \right)\!\frac{1}{\alpha_q+\beta_q}\left(1\!-\!e^{-(\alpha_q+\beta_q)\epsilon_q}\right)$$
    (20)
    $$\begin{array}{*{20}ll}\int_{\epsilon_q}^{T_i} q_i(t)dt &= \left[{\frac{\alpha_q}{\alpha_q+\beta_q}}+\left( q_i(0)-{\frac{\alpha_q}{\alpha_q+\beta_q}} \right)e^{-(\alpha_q+\beta_q)\epsilon_q}\right] \frac{1}{\beta_q}\left(1-e^{-\beta_q(T_i-\epsilon_q)}\right) \\&= \frac{1}{\beta_q}\left({\frac{\alpha_q}{\alpha_q+\beta_q}}\left(1-e^{-(\alpha_q+\beta_q)\epsilon_q}\right) -\left(1-e^{-(\alpha_q+\beta_q)\epsilon_q}\right)q_i(0)\right)\end{array}$$
    (21)

    Hence, combining the similar terms in (20) and (21), we obtain:

    $$\hat{q}_i=r_i\left(c-\frac{d}{e^{\beta_q/r_i}-w}\right)$$
    (22)

    where

    $$\begin{array}{ll} c & = \frac{\alpha_q\epsilon_q}{\alpha_q+\beta_q}+\frac{\alpha_q^2}{(\alpha_q+\beta_q)^2\beta_q}\left(1-e^{-(\alpha_q+\beta_q)\epsilon_q}\right) \\ d & = \frac{\alpha_q^2}{(\alpha_q+\beta_q)^2\beta_q}\left(e^{\beta_q\epsilon_q}-e^{-\alpha_q\epsilon_q}\right)\left(1-e^{-(\alpha_q+\beta_q)\epsilon_q}\right) \\ w & = e^{-\alpha_q\epsilon_q} \end{array}$$

    In general,

    $$\hat{q}(g_{syn},\:I)=r(g_{syn},\: I) \left(c-\frac{d}{e^{\beta_q/r(g_{syn},\: I)}-w}\right)$$
    (23)

    Both (22) and (23) are valid for any firing rate. In particular, for low firing rates, we deduce from (11) that

    $$\hat{q}(g_{syn},\:I) \approx c r(g_{syn},\: I)$$
    (24)

    For high firing rates, by using Taylor’s Formula in powers of \(\beta_q/r(g_{syn},\: I)\) and keeping the linear part only, we obtain:

    $$\hat{q}(g_{syn},\:I) \approx \left(c\! - \!\frac{d}{1\!-\!w} \right)\! r(g_{syn},\: I)\! + \frac{b \beta_q}{(1\!-\!c)(1\!-\!c\!+\!\beta_q \tau_{ref})}$$
    (25)

    The β q τ ref term in the denominator is the minimum possible value of \(\beta_q/r(g_{syn},\: I)\). Thus, the temporally averaged synaptic activation is approximately linear in both cases, with the slope being shallower in the latter. The graph of \(\hat{q}(g_{syn},\:I)\) is shown in Fig. 11(b).

Fig. 11
figure 11

Panel (a): firing rate of the single leaky integrate-and-fire neuron without noise. Transition to repetitive firing occurs when Θ eff (g syn , I) = 1, i.e., along the straight line I = 1 + g syn (1 − V syn ). If either g syn or I is fixed at some level, the firing-rate as a function of the other variable possesses the characteristic logarithmic shape. Superimposed are also shown three firing-rate population profiles, corresponding to the three marked steady-state values of g syn on the lower, middle, and upper branches of the bifurcation diagram in Fig. 7(a). Panel (b): single-cell synaptic activation averaged over one interspike interval as a function of the cell’s firing rate. Note the linear parts for low and high firing rates, consistent with (24) and (25)

We also provide some additional details here. In this work, we use uniform distributions of bias currents on a finite interval (I min ; I max ). Then, (14) simplifies to

$$g_{\mathit{syn}}={\overline{g}}_{syn}\frac{1}{I_{max}-I_{min}}\int_{I_{min}}^{I_{max}} s(t, \: \ensuremath{I}) \hat{q}(g_{syn},\:I) dI$$
(26)

For general description purposes (e.g., this is how the bifurcation diagrams in Fig. 7 were generated), we can interpret (14) in terms of \({\overline{g}}_{syn}\) as a function of the self-consistent value of g syn . This allows us to avoid the numerical solution of (14) altogether, but is not applicable if the self-consistent value of g syn corresponding to a specific value of \({\overline{g}}_{syn}\) is required. The corresponding expression is

$${\overline{g}}_{syn}=\frac{g_{\mathit{syn}}}{\int_{-\infty}^{\infty}f(\ensuremath{I}) s(t, \: \ensuremath{I}) \hat{q}(g_{syn},\:I) dI}$$
(27)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Vladimirski, B.B., Tabak, J., O’Donovan, M.J. et al. Episodic activity in a heterogeneous excitatory network, from spiking neurons to mean field. J Comput Neurosci 25, 39–63 (2008). https://doi.org/10.1007/s10827-007-0064-4

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-007-0064-4

Keywords

Navigation