Skip to main content
Log in

Finite-size and correlation-induced effects in mean-field dynamics

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

Abstract

The brain’s activity is characterized by the interaction of a very large number of neurons that are strongly affected by noise. However, signals often arise at macroscopic scales integrating the effect of many neurons into a reliable pattern of activity. In order to study such large neuronal assemblies, one is often led to derive mean-field limits summarizing the effect of the interaction of a large number of neurons into an effective signal. Classical mean-field approaches consider the evolution of a deterministic variable, the mean activity, thus neglecting the stochastic nature of neural behavior. In this article, we build upon two recent approaches that include correlations and higher order moments in mean-field equations, and study how these stochastic effects influence the solutions of the mean-field equations, both in the limit of an infinite number of neurons and for large yet finite networks. We introduce a new model, the infinite model, which arises from both equations by a rescaling of the variables and, which is invertible for finite-size networks, and hence, provides equivalent equations to those previously derived models. The study of this model allows us to understand qualitative behavior of such large-scale networks. We show that, though the solutions of the deterministic mean-field equation constitute uncorrelated solutions of the new mean-field equations, the stability properties of limit cycles are modified by the presence of correlations, and additional non-trivial behaviors including periodic orbits appear when there were none in the mean field. The origin of all these behaviors is then explored in finite-size networks where interesting mesoscopic scale effects appear. This study leads us to show that the infinite-size system appears as a singular limit of the network equations, and for any finite network, the system will differ from the infinite system.

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
Fig. 11

Similar content being viewed by others

Notes

  1. Note that the activation functions are different in BCC and Bressloff models.

References

  • Abbott, L. F, & Van Vreeswijk, C. A. (1993). Asynchronous states in networks of pulse-coupled neuron. Physical Review, 48, 1483–1490.

    Article  PubMed  Google Scholar 

  • Amari, S. (1972). Characteristics of random nets of analog neuron-like elements. Syst. Man Cybernet., SMC-2.

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

    Article  PubMed  CAS  Google Scholar 

  • Amit, D. J., & Brunel, N. (1997). Model of global spontaneous activity and local structured delay activity during delay periods in the cerebral cortex. Cerebral Cortex, 7, 237–252.

    Article  PubMed  CAS  Google Scholar 

  • Arnold, V.I. (1981). Ordinary differential equations (Chap. 5). MIT Press.

  • Arnold, L. (1995). Random dynamical systems. Springer.

  • Beggs, J. M., & Plenz, D. (2004). Neuronal avalanches are diverse and precise activity patterns that are stable for many hours in cortical slice cultures. Journal of Neuroscience, 24, 5216–5229.

    Article  PubMed  CAS  Google Scholar 

  • Benayoun, M., Cowan, J. D., van Drongelen, W., & Wallace, E. (2010). Avalanches in a stochastic model of spiking neurons. PLoS Computational Biology, 6(7), e1000846. doi:10.1371/journal.pcbi.1000846.

    Article  PubMed  Google Scholar 

  • Boland, R. P., Galla, T., & McKane, A. J. (2008). How limit cycles and quasi-cycles are related in systems with intrinsic noise. Journal of Statistical Mechanics: Theory and Experiment, 9, P09001.

    Article  Google Scholar 

  • Bressfloff, P. (2010). Stochastic neural field theory and the system-size expansion. SIAM Journal on Applied Mathematics, 70, 1488–1521.

    Article  Google Scholar 

  • Bressfloff, P. (2010). Metastable states and quasicycles in a stochastic Wilson–Cowan model. Physical Review E, 82, 051903.

    Article  Google Scholar 

  • Brewer, J. (1978). Kronecker products and matrix calculus in system theory. IEEE Transactions on Circuits and Systems, CAS-25.

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

    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. (2003). Firing rate of noisy quadratic integrate-and-fire neurons. Neural Computation, 15, 2281–2306.

    Article  PubMed  Google Scholar 

  • Buice, M. A., & Cowan, J. D. (2007). Field theoretic approach to fluctuation effects for neural networks. Physical Review E, 75, 051919.

    Article  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • Cai, D., Tao, L., Shelley, M., & McLaughlin, D. W. (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, 101, 7757–7762.

    Article  CAS  Google Scholar 

  • Coombes, S., & Owen, M. R. (2005). Bumps, breathers, and waves in a neural network with spike frequency adaptation. Physical Review Letters 94, 148102-1–148102-4.

    Article  Google Scholar 

  • Doob, J. L. (1945). Markoff chains–denumerable case. Transactions of the American Mathematical Society, 58, 455–473.

    Google Scholar 

  • Dykman, M. I., Mori, E., Ross, J., & Hunt, P. M. (1994). Large fluctuations and optimal paths in chemical kinetics. Journal of Chemical Physics, 100, 5735–5750.

    Article  CAS  Google Scholar 

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

    Article  PubMed  Google Scholar 

  • El Boustani, S., & Destexhe, A. (2010). Brain dynamics at multiple scales: can one reconcile the apparent low-dimensional chaos of macroscopic variables with the seemingly stochastic behavior of single neurons? International Journal of Bifurcation and Chaos, 20, 1–16.

    Article  Google Scholar 

  • Ermentrout, B. (1998). Neural networks as spatio-temporal pattern-forming systems. Reports on Progress in Physics, 61, 353–430.

    Article  Google Scholar 

  • Ermentrout, B. (2002). Simulating, analyzing, and animating dynamical systems: A guide to XPPAUT for researchers and students. Society for Industrial Mathematics.

  • Ermentrout, G. B., & Cowan, J.D. (1979). A mathematical theory of visual hallucination patterns. Biological Cybernetics, 34, 137–150.

    Article  PubMed  CAS  Google Scholar 

  • Faugeras, O., Touboul, J., & Cessac, B. (2009). A constructive mean field analysis of multi population neural networks with random synaptic weights and stochastic inputs. Frontiers in Neuroscience, 3, 1. doi:10.3389/neuro.10.001.2009.

    Google Scholar 

  • Freidlin, M. I., & Wentzell, A. D. (1998). Random perturbations of dynamical systems. Springer Verlag.

  • Gaspard, P. (2002). Correlation time of mesoscopic chemical clocks. Journal of Chemical Physics, 117, 8905–8916.

    Article  CAS  Google Scholar 

  • Gillespie, D. T. (1976). A general method for numerically simulating the stochastic time evolution of coupled chemical reactions. Journal of Computational Physics, 22, 403–434.

    Article  CAS  Google Scholar 

  • Gillespie, D. T. (1977). Exact stochastic simulation of coupled chemical reactions. The Journal of Physical Chemistry, 81, 2340–2361.

    Article  CAS  Google Scholar 

  • Guckenheimer, J., & Holmes, P. J. (1983). Nonlinear oscillations, dynamical systems and bifurcations of vector fields. Applied mathematical sciences (Vol. 42). Springer.

  • Horsthemke, W., & Lefever, R. (2006). Noise-induced transitions. Springer.

  • Kurtz, T. G. (1976). Limit theorems and diffusion approximations for density dependent Markov chains. Mathematical Programming Studies, 5, 67.

    Google Scholar 

  • Kuznetsov, Y. A. (1998). Elements of applied bifurcation theory. Applied Mathematical Sciences (2nd ed.). Springer.

  • Laing, C. L., Troy, W. C., Gutkin, B., & Ermentrout, G. B. (2002). Multiple bumps in a neuronal model of working memory. SIAM Journal on Applied Mathematics 63, 62–97.

    Article  Google Scholar 

  • Levina, A., Herrmann, J. M., & Geisel, T. (2009). Phase transitions towards criticality in a neural system with adaptive interactions. Physical Review Letters, 102, 118110-1–118110-4.

    Article  Google Scholar 

  • Ly, C., & Tranchina, D. (2007). Critical analysis of dimension reduction by a moment closure method in a population density approach to neural network modeling. Neural Computation, 19, 2032–2092.

    Article  PubMed  Google Scholar 

  • McKane, A. J., Nagy, J. D., Newman, T. J., & Stefanini, M. O. (2007). Amplified biochemical oscillations in cellular systems. Journal of Statistical Physics, 71, 165.

    Article  Google Scholar 

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

    Article  Google Scholar 

  • Neudecker, H. (1969). Some theorems on matrix differentiation with special reference to kronecker matrix products. Journal of the American Statistical Association, 64, 953–963.

    Article  Google Scholar 

  • Ohira, T., & Cowan, J. D. (1993). Master-equation approach to stochastic neurodynamics. Physical Review E, 48(3), 2259–2266.

    Article  Google Scholar 

  • Plesser, H. E. (1999). Aspects of signal processing in noisy neurons. PhD thesis, Georg-August-Universität.

  • Rodriguez, R., & Tuckwell, H. C. (1996). A dynamical system for the approximate moments of nonlinear stochastic models of spiking neurons and networks. Mathematical and Computer Modeling, 31, 175–180.

    Article  Google Scholar 

  • Rodriguez, R., & Tuckwell, H. C. (1998). Noisy spiking neurons and networks: Useful approximations for firing probabilities and global behavior. Biosystems, 48, 187–194.

    Article  PubMed  CAS  Google Scholar 

  • Rolls, E. T., & Deco, G. (2010). The noisy brain: Stochastic dynamics as a principle of brain function. Oxford University Press.

  • Softky, W. R., & Koch, C. (1993). The highly irregular firing of cortical cells is inconsistent with temporal integration of random epsps. Journal of Neuroscience, 13, 334–350.

    PubMed  CAS  Google Scholar 

  • Teramae, J., Nakao, H., & Ermentrout, G. B. (2009). Stochastic phase reduction for a general class of noisy limit cycle oscillators. Physical Review Letters, 102, 194102.

    Article  PubMed  Google Scholar 

  • Touboul, J., & Destexhe, A. (2010). Can power-law scaling and neuronal avalanches arise from stochastic dynamics? PLoS ONE, 5, e8982.

    Article  Google Scholar 

  • Touboul, J., & Faugeras, O. (2007). The spikes trains probability distributions: a stochastic calculus approach. Journal of Physiology, 101(1–3), 78–98.

    PubMed  Google Scholar 

  • Touboul, J., & Faugeras, O. (2008) First hitting time of double integral processes to curved boundaries. Advances in Applied Probability, 40, 501–528.

    Article  Google Scholar 

  • Wainrib, G. (2010). Randomness in neurons: A multiscale probabilistic analysis. PhD thesis, Ecole Polytechnique.

  • 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 

  • Wilson, H. R., & Cowan, J. D. (1973). A mathematical theory of the functional dynamics of cortical and thalamic nervous tissue. Biological Cybernetics, 13, 55–80.

    CAS  Google Scholar 

Download references

Acknowledgement

This work was supported by NSF DMS 0817131.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jonathan D. Touboul.

Additional information

Action Editor: Nicolas Brunel

Appendices

A An alternative derivation of the moment equations

In this appendix we show that the moment equation corresponding to the Markov chain governed by Eq. (2) can be derived from a Rodriguez–Tuckwell moment expansion on the Langevin approximation of the Markov chain. Following Kurtz approach in Kurtz (1976), we know that the dynamics of the our rescaled variables p i  = n i /N i , which is a Markov chain in the initial setting, approaches as N goes to infinity a continuous diffusion process (X t ) t ≥ 0 satisfying the equation:

$$ \begin{array}{rll} dX_t^i&=\Big(-\alpha_i\,X^i_t + f_i(s_i(t,X_t))\Big)\;dt\\&{\kern6pt}\quad+ \displaystyle\frac 1 N \sqrt{\alpha_i\,X^i_t + f_i(s_i(t,X_t))}\;dW^i_t \end{array} $$
(13)

where \(S_i(t,X_t)=\sum_{j=1}^N w_{ij}X^j_t + I_i(t)\).

Following the works of Rodriguez and Tuckwell (Rodriguez and Tuckwell 1996, 1998), we derive from this equation the dynamical system governing the approximate moments of X. Denoting by m j the mean value of X j and by C ij the correlation between X i and X j , a direct application of Rodriguez and Tuckwell formula applied to our particular form of dynamical system yields:

$$ \begin{cases} \displaystyle\frac{\text{d}m_j}{\text{d}t}&=-\alpha_j m_j +f_j(s_j) + \displaystyle\frac 1 2 f_j^{\prime\prime}(s_j)\sum_{l=1}^N \sum_{p=1}^N w_{jp}w_{jl} C_{lp}\\ \displaystyle\frac{\text{d}C_{ij}}{\text{d}t}&=-(\alpha_i+\alpha_j)C_{ij}+ f_i^{\prime}(s_i)\sum_{l=1}^N w_{il} C_{lj} \\&{\kern5pt}+ f_j^{\prime}(s_j)\sum_{l=1}^N w_{jl} C_{li} + \delta_{ij} \Big[\alpha_i m_i +f_i(s_i) \\ &{\kern5pt}+ \displaystyle\frac 1 {N_i N_j} f_i^{\prime\prime}(s_i)\sum_{k,l}w_{ik}w_{il} C_{kl}\Big] \end{cases} $$

Where \(s_i=\sum_{j=1}^N w_{ij}m^j_t + I_i(t)\). These equations therefore appear as a perturbation of the BCC equations with an additional nonlinear term in the dynamics or order two. Truncation at order 1 in the small parameter 1/N yields exactly BCC equations. Moreover, it is interesting to note that these equations again correspond in the limit N→ ∞, to the infinite model described in Section 3.

B Kronecker algebra: some useful properties

In this appendix, we review and prove some useful properties of Kronecker products of matrixes. We recall the definition of the the function Vect transforming a M × N matrix into a M N-dimensional column vector, as defined in Neudecker (1969):

$$ \mathrm{Vect}: \begin{cases} \mathbb{R}^{M\times N} &\mapsto \mathbb{R}^{M\,N}\\ X & \mapsto [X_{11}, \ldots, X_{M1}, X_{12}(t), \ldots, X_{M2}(t), \\ & {\kern6pt}\ldots X_{1N}(t), \ldots, X_{MN}(t)]^T \end{cases} $$

Let us now denote by ⊗ the Kronecker product defined for A ∈ ℝm×n and B ∈ ℝr×s as the (m r) ×(n s) matrix:

$$ A\otimes B := \left( \begin{array}{cccc} a_{11} B & a_{12} B & \cdots & a_{1n}B\\ a_{21} B & a_{22} B & \cdots & a_{2n}B\\ \vdots & \vdots & \ddots & \vdots\\ a_{m1} B & a_{m2} B & \cdots & a_{mn}B\\ \end{array} \right) $$

For standard definitions and identities in the field of Kronecker products, the reader is referred to Brewer (1978). We recalled in the main text the following identities for A, B, D, G, X ∈ ℝM×M, I M be the M × M identity matrix and A · B or A B denote the standard matrix product:

$$ \begin{cases} \mathrm{Vect} (AXB) = (B^T \otimes A ) \mathrm{Vect}(X)\\ A\oplus A =A\otimes I_M + I_M \otimes A\\ (A \otimes B)\cdot (D \otimes G) = (A\cdot D) \times (B\cdot G) \end{cases}. $$
(14)

The relationship ⊕ is called Kronecker sum.

Proposition 2

Let A and B in M×M , and assume that A has the eigenvalues {λ i ; i = 1,..., M} and B the eigenvalues {μ i ; i = 1,..., M}. Then we have:

  • A ⊗ B has the eigenvalues \(\{\lambda_i\mu_j;\;(i,j)\in \{1,\ldots,M\}^2\}\) .

  • A ⊕ B has the eigenvalues \(\{\lambda_i+\mu_j;\;(i,j)\in \{1,\ldots,M\}^2\}\) .

Proof

Let λ i (resp.μ j ) be an eigenvalue of A (resp. B) with eigenvector u (resp. v), and define the matrix z = u·v T (i.e. z ij  = u i v j ). We have:

$$ \begin{array}{rll} \Big(A\oplus B \Big) \, \mathrm{Vect} (z) &= \mathrm{Vect}\Big(A \cdot u\cdot v^T + u\cdot v^T \cdot B^T\Big)\\ &= \mathrm{Vect}\Big(\lambda_i u \cdot v^T + u \cdot (B\cdot v)^T\Big) \\ &= \lambda_i \mathrm{Vect}(z) + \mathrm{Vect}\Big(u (\mu_j v^T)\Big)\\ &= (\lambda_i + \mu_j) \mathrm{Vect}(z). \end{array} $$

which entails that Vect(z) is an eigenvector of A(ν) ⊕ A(ν) associated with the eigenvalue λ i  + λ j . Similarly, we have for z = v · u T:

$$ \begin{array}{rll} \Big(A\otimes B \Big) \, \mathrm{Vect} (z) &= \mathrm{Vect}\Big(B \cdot v\cdot u^T A^T\Big)\\ &= \mathrm{Vect}\Big((\mu_j v) \cdot (\lambda_i \,u^T) \Big) = \lambda_i\mu_h \mathrm{Vect}(z) \end{array} $$

The dimension of A ⊕ B and A ⊗ B is M 2 and there are exactly M 2 linearly independent matrices z possible built in the proposed fashion, therefore we identified all possible eigenvalues. □

Theorem 3

Let Φ(t) be the solution of the matrix differential equation:

$$ \begin{cases} \frac{\text{d}\Phi(t)}{\text{d}t} &= A(t)\Phi(t)\\ \Phi(0) &= I_M \end{cases} $$

and Ψ(t) the solution of:

$$ \begin{cases} \frac{\text{d}\Psi(t)}{\text{d}t} &= (A(t)\oplus A(t))\Psi(t)\\ \Psi(0) &= I_{M^2} \end{cases} $$

We have Ψ(t) = Φ(t) ⊗ Φ(t).

Proof

Indeed, we have, using the basic properties of the Kronecker product recalled in Eq. (8)

$$ \begin{array}{lll} &&{\kern-6pt}\frac{\text{d}\Phi(t)\otimes \Phi(t)}{\text{d}t} \\&&= \frac{\text{d}\Phi(t)}{\text{d}t} \otimes \Phi(t) + \Phi(t) \otimes \frac{\text{d}\Phi(t)}{\text{d}t}\\ &&= \big(A(\nu(t))\cdot \Phi(t)\big)\otimes \Phi(t) + \Phi(t) \otimes \big(A(\nu(t))\cdot \Phi(t)\big) \\ &&= \big(A(\nu(t))\cdot \Phi(t)\big)\otimes \big(I_M\cdot\Phi(t)\big) + \big(I_M\cdot\Phi(t)\big) \otimes \big(A(\nu(t))\cdot \Phi(t)\big)\\ &&= (A(\nu(t))\otimes I_M)\cdot (\Phi(t)\otimes \Phi(t)) \\ &&{\kern6pt}+ (I_M\otimes A(\nu(t))) \cdot (\Phi(t)\otimes \Phi(t))\\ &&= \big(A(\nu(t))\otimes I_M + I_M\otimes A(\nu(t))\big) \cdot \big(\Phi(t)\otimes\Phi(t)\big)\\ &&= \big(A(\nu(t))\oplus A(\nu(t))\big) \cdot \big(\Phi(t)\otimes\Phi(t)\big) \end{array} $$

Therefore Φ(t) ⊗ Φ(t) satisfies the same differential equation as Ψ(t) and moreover, \(\Phi(0)\otimes \Phi(0)=I_M\otimes I_M=I_{M^2}\), and therefore by existence and uniqueness of the resolvent, Ψ(t) = Φ(t) ⊗ Φ(t). □

C Genericity of the Hopf bifurcation found

In this appendix we derive the expression of the first Lyapunov exponent of the bifurcation, which proves that the existence of the Hopf bifurcation exhibited in Section 4.1.2. In that section, we derived the expression of the Jacobian matrix at the considered fixed point:

$$ J=\left (\begin{array}{cc} -\alpha + w\, f^{\prime}(s) & \displaystyle\frac 1 2 \,f^{\prime\prime}(s)\,w^2\\ 2f^{\prime}(s)\displaystyle\frac w N & 2(-\alpha+w\, f^{\prime}(s)) \end{array}\right) $$

At the bifurcation point, we have − α * + w f (s) = 0, and therefore at this point the Jacobian matrix reads:

$$ J_0=\left (\begin{array}{cc} 0 & \displaystyle\frac 1 2 \,f^{\prime\prime}(I)\,w^2\\ 2f^{\prime}(I)\displaystyle\frac w N & 0 \end{array}\right) $$

The eigenvalues of this matrix under the assumptions of Section 4.1.2 are ±i ω 0 where \(\omega_0=\sqrt{-f^{\prime}(I)f^{\prime\prime}(I)w^3/N}\). We define q the right eigenvector of J 0 associated with i ω 0:

$$ q=\Big(-\displaystyle\frac{i}{\sqrt{2\,f^{\prime}(I)\,w/N}} \qquad 1\Big)^T $$

and p the right eigenvector of \(J_0^T\) associated with the eigenvalue 0:

$$ p=\Big(\displaystyle\frac{i}{\sqrt{-\displaystyle\frac 1 2 f^{\prime\prime}(I)\,w^2}} \qquad 1\Big)^T $$

For the sake of simplicity, we also name the components of the vector field of the system:

$$ \begin{cases} f_1(\binom{\nu}{C}, \alpha^*) &= w\,f^{\prime}(I)\,nu + f(w\nu+I)\\&{\kern6pt}+\displaystyle\frac 1 2 f^{\prime\prime}(w\nu+I)\,w^2\,C\\ f_2(\binom{\nu}{C}, \alpha^*) &= -2\,w\,f^{\prime}(I)\,C +\\&{\kern6pt} 2\,f^{\prime}(w\nu+I)\,w\,\big(C+\displaystyle\frac{\nu}{N}\big)\\ \end{cases}. $$

Following Kuznetsov (1998), we define \(B(\binom{x_1}{y_1}, \binom{x_2}{y_2})\) and \(C(\binom{x_1}{y_1}, \binom{x_2}{y_2},\binom{x_3}{y_3})\) the second and third derivatives of the vector field, which are bi- and tri-linear forms. We have the following expressions for these multilinear functions:

$$ \begin{cases} B_1(\binom{x_1}{y_1}, \binom{x_2}{y_2}) &= w^2\,f^{\prime\prime}(I)\,x_1\,x_2 + \displaystyle\frac 1 2\,f^{\prime\prime}(I)\,w^2\\&{\kern6pt}\times(x_1\,y_2 +y_1\,x_2)\\ B_2(\binom{x_1}{y_1}, \binom{x_2}{y_2}) &= 2\,f^{\prime\prime}(I)\,w^2\,(x_1\,y_2+y_1\,x_2) \\&{\kern6pt}+ \displaystyle\frac 4 N\,f^{\prime\prime}(I)\,w^2\,x_1 \, x_2\\ C_1(\binom{x_1}{y_1}, \binom{x_2}{y_2}, \binom{x_3}{y_3}) &= f^{(3)}(I)\,w^3\,x_1\,x_2\,x_3 + \displaystyle\frac 1 2 \, f^{(4)}(I)\,w^4\\&{\kern6pt}\times(x_1\,x_2\,y_3+x_1\,y_2\,x_3+y_1\,x_2\,x_3)\\ C_2(\binom{x_1}{y_1}, \binom{x_2}{y_2}, \binom{x_3}{y_3}) &= \displaystyle\frac 6 N\,f^{(3)}(I)\, w^3 \, x_1\,x_2\,x_3 + 2\,f^{(3)}(I)\,w^3 \\&{\kern6pt}\times(x_1\,x_2\,y_3+x_1\,y_2\,x_3+y_1\,x_2\,x_3) \end{cases} $$

We are now in position to compute the first Lyapunov exponent l 1(0) using the formula:

$$ \begin{array}{lll} &&{\kern-18pt}l_1(0)\\&{\kern3pt}=&\displaystyle\frac 1 {2\omega_0} \textrm{Re}\Big ( \langle p, C(q,q,\overline{q})\rangle - 2 \langle p, B(q, J_0^{-1} B(q, \overline{q}))\rangle \\ &&{\kern3pt}+ \langle p, B(\overline{q}, (2\,i\,\omega_0 Id - J_0)^{-1}B(q,q))\rangle \Big)\\ &{\kern3pt}=& \displaystyle\frac 1 {2\omega_0} (\mathcal{A}-2\mathcal{B}+\mathcal{C}) \end{array} $$

where \(\langle x, y \rangle\) denotes the complex inner product \(\overline{x}^T\cdot y\) and the sum of three terms denoted \(\mathcal{A}\), \(\mathcal{B}\) and \(\mathcal{C}\) are the real parts of the terms involved in the expression of the Lyapunov exponent. After straightforward but tedious calculations (that can be conveniently performed using a formal calculation tool such as Maple), we obtain:

$$ \begin{cases} \mathcal{A} &= \displaystyle\frac{w^2 \, N}{f^{\prime}(I)^2} f^{(3)}(I)\,f^{\prime}(I)\\ \mathcal{B} &= \displaystyle\frac{w^2 \, N}{f^{\prime}(I)^2} \left( -\displaystyle\frac{f^{\prime\prime}(I)^2}{\omega_0} - \displaystyle\frac 1 2 \displaystyle\frac{f^{(3)}(If^{\prime}(I))}{\omega_0} + 2 f^{\prime\prime}(I)^2\right)\\ \mathcal{C} &= -\displaystyle\frac{w^2\, N }{f^{\prime}(I)^2}\;\displaystyle\frac{2 f^{\prime\prime}(I)^2}{3} \end{cases} $$

which yields the expression for the Lyapunov exponent:

$$ \begin{array}{rll} l_1(0) &=& \displaystyle\frac{1}{2\,\omega_0} \left [ \mathcal{A} -2\,\mathcal{B}+\mathcal{C}\right ]\\ & =& \displaystyle\frac{w^2\,N}{2\,\omega_0\,f^{\prime}(I)^2} \left [ f^{(3)}(I)f^{\prime}(I) \left(1+\displaystyle\frac {1}{\omega_0}\right)\right.\\ &&\left.\vphantom{\left(1+\displaystyle\frac {1}{\omega_0}\right)}+ f^{\prime\prime}(I)^2\left(\displaystyle\frac{2}{\omega_0} - \displaystyle\frac{14}{3}\right)\right] \end{array} $$

D Finite-size effects in BCC two-populations model I

In this appendix, we study the two-populations BCC system corresponding to Model I and show that the finite-size effects are closely related to what is observed in Bressloff model as studied in Section 4.2. BCC finite-size equations read:

$$ \begin{cases} a_1^{\prime}&=-\alpha a_1+f(s_1)+\frac{1}{2}\,f^{\prime\prime}(s_1)\\&{\kern6pt}\times(w_{11}^2\,c_{11}+w_{12}^2\,c_{22}+2\,w_{12}\,w_{11}\,c_{12})\\ a_2^{\prime}&=-\alpha a_2+f(s_2)+\frac{1}{2}\,f^{\prime\prime}(s_2)\\&{\kern6pt}\times(w_{22}^2\,c_{22}+w_{21}^2\,c_{11}+2\,w_{22}\,w_{21}\,c_{12})\\ c_{11}^{\prime}&=-2\alpha c_{11}+2\,f^{\prime}(s_{1})\,(w_{11}\,c_{11}+w_{12}\,c_{12})\\&{\kern6pt}+2\,f^{\prime}(s_1)\,a_1\,w_{11}\,n_e\\ c_{22}^{\prime}&=-2\alpha c_{22}+2\,f^{\prime}(s_{2})\,(w_{21}\,c_{12}+w_{22}\,c_{22})\\&{\kern6pt}+2\,n_i\,f^{\prime}(s_2)\,w_{22}\,a_2\\ c_{12}^{\prime}&=-2\alpha c_{12}+f^{\prime}(s_{1})\,(w_{11}\,c_{12}+w_{12}\,c_{22}+a_2\,w_{12}\,n_i) \ldots \\ & \qquad + f^{\prime}(s_2)\,(w_{21}\,c_{11}+w_{22}\,c_{12}+w_{21}\,a_{1}\,n_e)\\ \end{cases} $$
(15)

where we denoted:

$$ \begin{cases} s_1 &= w_{11}\,a_1+w_{12}\,a_2+i_1\\ s_2 &= w_{21}\,a_2+w_{22}\,a_2+i_2 \end{cases} $$

Similarly to Bressloff case, BCC model features two families of limit cycles (see Fig. 12). One of these branches corresponds exactly to the branch of limit cycles of WC system starting from a Hopf bifurcation and disappearing through a homoclinic bifurcations. Two additional Hopf bifurcations appear related to the family of periodic orbit corresponding to the correlation-induced cycle evidence in the analysis of the infinite-size system. This branch of limit cycles exist whatever n, and loses stability through a Neimark-Sacker bifurcation as the number of neurons increases. This bifurcation generates chaos for large enough networks, which clearly does not exists in WC system.

Fig. 12
figure 12

Bifurcation diagram for the BCC system. Blue lines represent the equilibria, pink lines the extremal values of the cycles in the system. Bifurcations of equilibria are denoted with a red star, LP represents a saddle-node bifurcation (Limit Point), H a Hopf bifurcation. The four Hopf bifurcations share two families of limit cycles. The branch corresponding to the smaller values of i 1 undergoes two fold of limit cycles, and the other branch of limit cycle a Neimark Sacker (Torus) bifurcation

This family of limit cycles presents stability for networks containing between 100 and 17.000 neurons, corresponding to a finite size effect that appears only in the region of interest of the cortical columns. As the number of neurons increase, the system keep the same number and type of bifurcations, and the infinite model appears to be a singular case where different bifurcations meet (see Fig. 13), and very large networks lose the property of presenting a stable cycle, and present a chaotic behavior.

Fig. 13
figure 13

Codimension two bifurcations and chaotic behavior as the number of neuron increases

Rights and permissions

Reprints and permissions

About this article

Cite this article

Touboul, J.D., Ermentrout, G.B. Finite-size and correlation-induced effects in mean-field dynamics. J Comput Neurosci 31, 453–484 (2011). https://doi.org/10.1007/s10827-011-0320-5

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-011-0320-5

Keywords

Navigation