Skip to main content
Log in

A finite volume method for stochastic integrate-and-fire models

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

Abstract

The stochastic integrate and fire neuron is one of the most commonly used stochastic models in neuroscience. Although some cases are analytically tractable, a full analysis typically calls for numerical simulations. We present a fast and accurate finite volume method to approximate the solution of the associated Fokker-Planck equation. The discretization of the boundary conditions offers a particular challenge, as standard operator splitting approaches cannot be applied without modification. We demonstrate the method using stationary and time dependent inputs, and compare them with Monte Carlo simulations. Such simulations are relatively easy to implement, but can suffer from convergence difficulties and long run times. In comparison, our method offers improved accuracy, and decreases computation times by several orders of magnitude. The method can easily be extended to two and three dimensional Fokker-Planck equations.

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
Fig. 11

Similar content being viewed by others

Notes

  1. I modified the proof.

References

  • Apfaltrer, F., Ly, C., & Tranchina, D. (2006). Population density methods for stochastic neurons with realistic synaptic kinetics: Firing rate dynamics and fast computational methods. Network: Computation in Neural Systems, 17(4), 373–418.

    Article  Google Scholar 

  • Barth, T., & Ohlberger, M. (2004). Finite volume methods: Foundation and analysis. Encyclopedia of computational mechanics. New York: Wiley.

    Google Scholar 

  • Bruneau, C. H., Marpeau, F., & Saad, M. (2005). Numerical simulation of the miscible displacement of radionuclides in a heterogeneous porous medium. International Journal for Numerical Methods in Fluids, 49(10), 1053–1085.

    Article  CAS  Google Scholar 

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

    Article  PubMed  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  PubMed  CAS  Google Scholar 

  • Brunel, N., & Latham, P. E. (2003). Firing rate of the noisy quadratic integrate-and-fire neuron. Neural Computation, 15(10), 2281–2306.

    Article  PubMed  Google Scholar 

  • Burkitt, A. N. (2006). A review of the integrate-and-fire neuron model: I. homogeneous synaptic input. Biological Cybernetics, 95(1), 1–19.

    Article  PubMed  CAS  Google Scholar 

  • Courant, R., Friedrichs, K., & Lewy, H. (1928). Uber die partiellen differenzengleichungen der mathematischen Physik. Mathematische Annalen, 100(1), 32–74.

    Article  Google Scholar 

  • Doiron, B., Chacron, M. J., Maler, L., Longtin, A., & Bastian, J. (2003). Inhibitory feedback required for network oscillatory responses to communication but not prey stimuli. Nature, 421(6922), 539–43.

    Article  PubMed  CAS  Google Scholar 

  • Ermentrout, G. B., & Kopell, N. (1986). Parabolic bursting in an excitable system coupled with a slow oscillation. SIAM Journal on Applied Mathematics, 46(2), 233–253.

    Article  Google Scholar 

  • Galán, R. F., Ermentrout, G. B., & Urban, N. N. (2007). Stochastic dynamics of uncoupled neural oscillators: Fokker-Planck studies with the finite element method. Physical Review E, 76(5), 56110.

    Article  Google Scholar 

  • Gardiner, C. W. (1985). Handbook of stochastic methods. New York: Springer.

    Google Scholar 

  • Godlewski, E., & Raviart, P. A. (1990). Hyperbolic systems of conservation laws. Mathématiques et applications (Vol. 3/4). Paris: Ellipses.

    Google Scholar 

  • Harten, A. (1983). High resolution schemes for hyperbolic conservation laws. Journal of Computational Physics, 49, 357–392.

    Article  Google Scholar 

  • Kloeden, P. E., & Platen, E. (1992). Numerical solution of stochastic differential equations. New York: Springer.

    Google Scholar 

  • Lapicque, L. (1907). Recherches quantitatives sur l’excitation electrique des nerfs traitee comme une polarisation. Journal de Physiologie et de Pathologie Générale (Paris), 9, 622–635.

    Google Scholar 

  • Lascaux, P., & Théodor, R. (1986–1987). Analyse numérique matricielle appliquée à l’Art de l’Ingénieur (Vol. 1,2). Paris: Masson.

    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(2), 808–827.

    PubMed  CAS  Google Scholar 

  • Lindner, B. (2001). Coherence and stochastic resonance in nonlinear dynamical systems. Ph.D. thesis, Humboldt University.

  • Lindner, B., & Longtin, A. (2006). Comment on “Characterization of subthreshold voltage fluctuations in neuronal membranes” by M. Rudolph and A. Destexhe. Neural Computation, 18(8), 1896.

    Article  PubMed  Google Scholar 

  • Lindner, B., & Schimansky-Geier, L. (2001). Transmission of noise coded versus additive signals through a neronal ensemble. Physical Review Letters, 86, 2934–2937.

    Article  PubMed  CAS  Google Scholar 

  • Mattia, M., & Del Giudice, P. (2002). Population dynamics of interacting spiking neurons. Physical Review E, 66(5), 51917.

    Article  Google Scholar 

  • Mattia, M., & Del Giudice, P. (2004). Finite-size dynamics of inhibitory and excitatory interacting spiking neurons. Physical Review E, 70(5), 52903.

    Article  Google Scholar 

  • Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P. (2007). Numerical recipes: The art of scientific computing (3rd ed.) Cambridge: Cambridge University Press.

    Google Scholar 

  • Rall, W. (1995). The theoretical foundation of dendritic function: Selected papers of Wilfrid Rall with commentaries. Boston: MIT.

    Google Scholar 

  • Renart, A., Brunel, N., & Wang, X. J. (2004). Mean-field theory of irregularly spiking neuronal populations and working memory in recurrent cortical networks. In Computational neuroscience: A comprehensive approach (pp. 431–490). Boca Raton: Chapman & Hall.

    Google Scholar 

  • Risken, H. (1989). The Fokker-Planck equation: Methods of solution and applications. Berlin: Springer.

    Google Scholar 

  • Roe, P. L. (1984). Generalized formulation of TVD Lax-Wendroff schemes. ICASE Rep. 53–84.

  • Rudolph, M., & Destexhe, A. (2005). An extended analytic expression for the membrane potential distribution of conductance-based synaptic noise. Neural Computation, 17(11), 2301–2315.

    Article  PubMed  CAS  Google Scholar 

  • Strang, G. (1968). On the construction and comparison of difference schemes. SIAM Journal on Numerical Analysis, 5, 506–517.

    Article  Google Scholar 

  • Sweby, P. K. (1984). High resolution schemes using flux limiters for hyperbolic conservation laws. SIAM Journal on Numerical Analysis, 21 995–1011.

    Article  Google Scholar 

  • Toro, E. F. (2001). Godunov methods: Theory and applications. Dordrecht: Kluwer/Plenum.

    Google Scholar 

  • Tuckwell, H. C. (1988). Introduction to theoretical neurobiology. Cambridge: Cambridge University Press.

    Google Scholar 

  • Tuckwell, H. C. (1989). Stochastic processes in the neurosciences. Philadelphia: SIAM.

    Google Scholar 

Download references

Acknowledgements

We thank Cheng Ly, Roberto Fernández Galán and Brent Doiron for helpful discussions. This research was supported by NSF grants DMS-0604429 and DMS-0817649, and a Texas ARP/ATP award.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Fabien Marpeau.

Additional information

Action Editor: G. Bard Ermentrout

Appendices

Appendix A: Construction of numerical scheme (1112) and Proof of Proposition 1

Footnote 1First, consider the second order Taylor expansion

(26)

Then, integrating over one cell using \(\partial_{t}P=-\partial_{V}(Pf)\), \(\partial_{tt}P=\partial_{V }\big( f\partial_{V }(Pf)\big) =f\partial_{V V }(Pf)\!+\!\partial_{V }f\partial_{V }(Pf)\) and the integration by parts formula, one has

(27)

For the time being, let us neglect the terms containing Δt 2, so that (26) reduces to the first order Taylor expansion. Then, one of the simplest stable schemes can be derived by using the first order upwind approximation \(P(t_{n},V_{i+{\frac{1}{2}} })f\big(V_{i+{\frac{1}{2}} }\big)\thickapprox P_{i}^{n}f_{i+{\frac{1}{2}} }^{+}+P_{i+1}^{n}f_{i+{\frac{1}{2}} }^{-}\) in the first term of the right-hand side of (27). We obtain the numerical scheme (1112) with φ ≡ 0, which is first order accurate in space and time. Such accuracy does not properly capture some possible drift effects, such as sharp fronts (see for instance Bruneau et al. 2005; Godlewski and Raviart 1990). It has been observed that using a second order approximation of the space derivatives improves drastically the accuracy of numerical solutions to hyperbolic equations, even if the accuracy in time is still of first order. For this reason, we use instead the centered discretization \(P\big(t_{n},V_{i+{\frac{1}{2}} }\big)f\big(V_{i+{\frac{1}{2}} }\big)\thickapprox f_{i+{\frac{1}{2}} }\big( \gamma_{i+{\frac{1}{2}} }P_{i}^{n}+\big(1-\gamma_{i+{\frac{1}{2}} }\big)P_{i+1}^{n}\big) \) in the first term of the right-hand side of (27), where the coefficient \(\gamma _{i+{\frac{1}{2}} }=\frac{1}{2}\frac{\Delta V_{i+1}}{\Delta V_{i+{\frac{1}{2}} }}\) interpolates the numerical solution at the interface \(V_{i+{\frac{1}{2}} }\), thus allowing us to handle variable size mesh elements. Finally, we further improve the accuracy of our scheme by including the term \(\Big[ f^{2}\partial_{V } P(t_{n},.) \Big]_{V_{i-{\frac{1}{2}} }}^{V_{i+{\frac{1}{2}}}}\) and using a centered approximation, \(f^{2}\big(V_{i+{\frac{1}{2}} }\big)\partial_{V } P\big(t_{n},V_{i+{\frac{1}{2}} }\big) \thickapprox f_{i+{\frac{1}{2}} }^{2}\frac{P_{i+1}^{n}-P_{i}^{n}}{\Delta V }\) for all i. This results in the numerical scheme (1112) with φ ≡ 1, which is more accurate in space and time than the upwind scheme, but unstable. Also notice that it becomes second order accurate in time if the drift term is constant, hence cancelling out the terms containing \(\partial_{V}f\) in Eq. (27).

Indeed, the flux (12) is a perturbation of the upwind scheme, so a compromise between accuracy and stability is offered by selecting φ satisfying (13).

In order to prove Proposition 1, the resulting scheme (1112) is written as \(P_{i}^{n+{\frac{1}{2}} }=P_{i}^{n}\Big( 1- \Delta t\frac{f_{i+{\frac{1}{2}} }-f_{i-{\frac{1}{2}} }}{\Delta V_{i}}\Big) + A_{i+{\frac{1}{2}} }^{n}\Delta P_{i+{\frac{1}{2}} }^{n} - B_{i-{\frac{1}{2}} }^{n}\Delta P_{i-{\frac{1}{2}} }^{n}\), with

$$\begin{array}{lll} A_{i+{\frac{1}{2}} }^{n}&=&-\frac{\Delta tf_{i+{\frac{1}{2}} }^{-}}{\Delta V_{i}}\left( 1-\frac{1}{2}\left( 1+\frac{\Delta tf_{i+{\frac{1}{2}} }^{-}}{\Delta V_{i+1}}\right) \frac{\Delta V_{i+1}}{\Delta V_{i+{\frac{1}{2}} }}\phi _{i+{\frac{1}{2}} }^{m} \right. \\ &&{\kern45pt}\left.+ \frac{1}{2}\left( 1+\frac{\Delta tf_{i-{\frac{1}{2}} }^{-}}{\Delta V_{i}}\right) \frac{\Delta V_{i}}{\Delta V_{i-{\frac{1}{2}} }}\frac{\phi _{i-{\frac{1}{2}} }^{m}}{r_{i-{\frac{1}{2}} }^{m}}\right) , \\ B_{i-{\frac{1}{2}} }^{n}&=&\frac{\Delta tf_{i-{\frac{1}{2}} }^{+}}{\Delta V_{i}}\left( 1-\frac{1}{2}\left( 1-\frac{\Delta tf_{i-{\frac{1}{2}} }^{+}}{\Delta V_{i-1}}\right) \frac{\Delta V_{i-1}}{\Delta V_{i-{\frac{1}{2}} }}\phi _{i-{\frac{1}{2}} }^{p}\right.\\ &&{\kern40pt} \left.+\frac{1}{2}\left( 1-\frac{\Delta tf_{i+{\frac{1}{2}} }^{+}}{\Delta V_{i}}\right) \frac{\Delta V_{i}}{\Delta V_{i+{\frac{1}{2}} }}\frac{\phi_{i+{\frac{1}{2}} }^{p}}{r_{i+{\frac{1}{2}} }^{p}}\right) . \end{array}$$

Using property (13) and re-phrasing

$$\begin{array}{lll} P_{i}^{n+{\frac{1}{2}} }\! &=& \! P_{i}^{n}\left(\! 1 \!-\! \frac{\Delta t}{\Delta V_{i}} \left(\! f_{i+{\frac{1}{2}} } \!-\! f_{i-{\frac{1}{2}} }\right)^{+} \!-\! A_{i+{\frac{1}{2}} }^{n} \!-\! B_{i-{\frac{1}{2}} }^{n} \right) \\ &&- \frac{\Delta t}{\Delta V_{i}} \left(\! f_{i+{\frac{1}{2}} } \!-\! f_{i-{\frac{1}{2}} }\right)^{-} \! P_{i}^{n} \!+\! A_{i+{\frac{1}{2}} }^{n}P_{i+1}^{n} \!+\! B_{i-{\frac{1}{2}} }^{n}P_{i-1}^{n} , \end{array}$$

one has \(A_{i+{\frac{1}{2}} }^{n}\!\ge\! 0\), \(B_{i-{\frac{1}{2}} }^{n}\!\ge\! 0\) and \(1 \!-\! \frac{\Delta t}{\Delta V_{i}} ( f_{i+{\frac{1}{2}} }-f_{i-{\frac{1}{2}} })^{+} - A_{i+{\frac{1}{2}} }^{n} - B_{i-{\frac{1}{2}} }^{n} \ge 0\) under conditions (15) and (16), implying that the scheme (1112) is nonnegativity preserving and l  ∞ –stable.

Appendix B: Proof of Proposition 2

Recursively assuming that \(P_{i}^{n}\ge 0\) for a given index n and all i, which is true for n = 0, Proposition 1 and boundary condition imply that \(P_{i}^{n+{\frac{1}{2}} }\ge 0\) for all i under conditions (1516). Then, multiplying (11) by ΔV i and summing from i = 1 to i = N, one has \(\| P^{n+{\frac{1}{2}} }\|_{l^{1}}=\| P^{n}\|_{l^{1}}\).

Next, P n + 1 is obtained by inverting the linear system MP n + 1 = S whose matrix M is given in Fig. 3 and the right-hand side S is defined by \(S_{i}=\Delta V_{i}P_{i}^{n+{\frac{1}{2}} }\) if i ≠ i R , and \(S_{i_{R}}=\Delta V_{i_{R}}P_{i_{R}}^{n+{\frac{1}{2}} }-\int_{t_{n}-\tau_{R}}^{ \min (t_{n+1}-\tau_{R},t_{n}) }F(t)\, dt\).

Indeed, the matrix M is strongly column diagonally dominant with positive diagonal coefficients and nonpositive off-diagonal coefficients. So it is invertible and M  − 1 is a positive matrix. Since the right-hand side S has nonnegative entries, the solution of the system, P n + 1, is component-wise nonnegative.

Finally, sum (24) and (19) from i = 1 to i = N to obtain

$$\| P^{n+1}\|_{l^{1}}=\| P^{n+{\frac{1}{2}} }\|_{l^{1}}+ \int_{t_{n}}^{t_{n+1}}F(t)\, dt - \int_{t_{n}-\tau_{R}}^{t_{n+1}-\tau_{R}}F(t)\, dt\ ,$$

thus recursively implying (25).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Marpeau, F., Barua, A. & Josić, K. A finite volume method for stochastic integrate-and-fire models. J Comput Neurosci 26, 445–457 (2009). https://doi.org/10.1007/s10827-008-0121-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-008-0121-7

Keywords

Navigation