Skip to main content
Log in

Modeling multiple time scale firing rate adaptation in a neural network of local field potentials

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

Abstract

In response to stimulus changes, the firing rates of many neurons adapt, such that stimulus change is emphasized. Previous work has emphasized that rate adaptation can span a wide range of time scales and produce time scale invariant power law adaptation. However, neuronal rate adaptation is typically modeled using single time scale dynamics, and constructing a conductance-based model with arbitrary adaptation dynamics is nontrivial. Here, a modeling approach is developed in which firing rate adaptation, or spike frequency adaptation, can be understood as a filtering of slow stimulus statistics. Adaptation dynamics are modeled by a stimulus filter, and quantified by measuring the phase leads of the firing rate in response to varying input frequencies. Arbitrary adaptation dynamics are approximated by a set of weighted exponentials with parameters obtained by fitting to a desired filter. With this approach it is straightforward to assess the effect of multiple time scale adaptation dynamics on neural networks. To demonstrate this, single time scale and power law adaptation were added to a network model of local field potentials. Rate adaptation enhanced the slow oscillations of the network and flattened the output power spectrum, dampening intrinsic network frequencies. Thus, rate adaptation may play an important role in network dynamics.

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

Similar content being viewed by others

References

  • Abel, H. J., Lee, J. C., Callaway, J. C., & Foehring, R. C. (2004). Relationships between intracellular calcium and afterhyperpolarizations in neocortical pyramidal neurons. Journal of Neurophysiology, 91, 324–335.

    Article  CAS  PubMed  Google Scholar 

  • Adrian, E. D., & Zotterman, Y. (1926). The impulses produced by sensory nerve endings: part 2. The response of a single End-organ. The Journal of Physiology, 61, 151–171.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Arfken, G. B., & Weber, H.-J. (1995). Mathematical methods for physicists. San Diego: Academic.

    Google Scholar 

  • Barlow, HB. (1961) Possible principles underlying the transformation of sensory messages. In: Sensory communication. In W Rosenblith(ed.), MIT Press.

  • Barraza, D., Kita, H., & Wilson, C. J. (2009). Slow spike frequency adaptation in neurons of the rat subthalamic nucleus. Journal of Neurophysiology, 102, 3689–3697.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Benda, J., & Herz, A. V. (2003). A universal model for spike-frequency adaptation. Neural Computation, 15, 2523–2564.

    Article  PubMed  Google Scholar 

  • Brenner, N., de Ruyter, B. W., & van Steveninck, R. (2000). Adaptive rescaling maximizes information transmission. Neuron, 26, 695–702.

    Article  CAS  PubMed  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 

  • Connor, J. A., & Stevens, C. F. (1971). Prediction of repetitive firing behaviour from voltage clamp data on an isolated neurone soma. The Journal of Physiology, 213, 31–53.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • David, O., & Friston, K. J. (2003). A neural mass model for MEG/EEG: coupling and neuronal dynamics. NeuroImage, 20, 1743–1755.

    Article  PubMed  Google Scholar 

  • Dayan, P., & Abbott, L. F. (2001). Theoretical neuroscience: computational and mathematical modeling of neural systems. Cambridge: Massachusetts Institute of Technology Press.

    Google Scholar 

  • de Biase, S., Gigli, G. L., Valente, M., & Merlino, G. (2014). Lacosamide for the treatment of epilepsy. Expert Opinion on Drug Metabolism & Toxicology, 10, 459–468.

    Article  Google Scholar 

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

    Article  PubMed Central  PubMed  Google Scholar 

  • Demanuele, C., Broyd, S. J., Sonuga-Barke, E. J., & James, C. (2013). Neuronal oscillations in the EEG under varying cognitive load: a comparative study between slow waves and faster oscillations. Clinical Neurophysiology: Official Journal of the International Federation of Clinical Neurophysiology, 124, 247–262.

    Article  Google Scholar 

  • Descalzo, V. F., Nowak, L. G., Brumberg, J. C., McCormick, D. A., & Sanchez-Vives, M. V. (2005). Slow adaptation in fast-spiking neurons of visual cortex. Journal of Neurophysiology, 93, 1111–1118.

    Article  CAS  PubMed  Google Scholar 

  • Destexhe, A., Hughes, S. W., Rudolph, M., & Crunelli, V. (2007). Are corticothalamic ‘up’ states fragments of wakefulness? Trends in Neurosciences, 30, 334–342.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Drew, P. J., & Abbott, L. F. (2006). Models and properties of power-law adaptation in neural systems. Journal of Neurophysiology, 96, 826–833.

    Article  PubMed  Google Scholar 

  • Ermentrout, B. (1998). Linearization of F-I curves by adaptation. Neural Computation, 10, 1721–1729.

    Article  CAS  PubMed  Google Scholar 

  • Ermentrout, G. B., & Cowan, J. D. (1979). Temporal oscillations in neuronal nets. Journal of Mathematical Biology, 7, 265–280.

    Article  CAS  PubMed  Google Scholar 

  • Fairhall A, Bialek, W. (2002) Adaptive spike coding. In: The Handbook of Brain Theory and Neural Networks. In M. Arbib (ed.), MIT Press.

  • Fairhall, A. L., Lewen, G. D., de Ruyter, B. W., & van Steveninck, R. (2001a). Multiple timescales of adaptation in a neural code. In T. K. Leen (Ed.), Advances in Neural Information Processing Systems 13 (pp. 124–130). Cambridge: MIT Press.

    Google Scholar 

  • Fairhall, A. L., Lewen, G. D., Bialek, W., & de Ruyter Van Steveninck, R. R. (2001b). Efficiency and ambiguity in an adaptive neural code. Nature, 412, 787–792.

    Article  CAS  PubMed  Google Scholar 

  • Fellous, J. M., & Sejnowski, T. J. (2000). Cholinergic induction of oscillations in the hippocampal slice in the slow (0.5–2 Hz), theta (5–12 Hz), and gamma (35–70 Hz) bands. Hippocampus, 10, 187–197.

    Article  CAS  PubMed  Google Scholar 

  • Fleidervish, I. A., Friedman, A., & Gutnick, M. J. (1996). Slow inactivation of Na + current and slow cumulative spike adaptation in mouse and guinea-pig neocortical neurones in slices. The Journal of Physiology, 493(Pt 1), 83–97.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Gerstner, W., & Kistler, W. M. (2002). Spiking neuron models : single neurons, populations, plasticity. Cambridge: Cambridge University Press.

    Book  Google Scholar 

  • Harris, J., & Stöcker, H. (1998). Handbook of mathematics and computational science. New York: Springer.

    Book  Google Scholar 

  • Higgs, M. H., Slee, S. J., & Spain, W. J. (2006). Diversity of gain modulation by noise in neocortical neurons: regulation by the slow afterhyperpolarization conductance. The Journal of Neuroscience, 26, 8787–8799.

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Izhikevich, E. M. (2007). Dynamical systems in neuroscience: the geometry of excitability and bursting. Cambridge: MIT Press.

    Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  • Koch, C. (1999). Biophysics of computation: information processing in single neurons. New York: Oxford University Press.

    Google Scholar 

  • La Camera, G., Rauch, A., Luscher, H. R., Senn, W., & Fusi, S. (2004). Minimal models of adapted neuronal response to in vivo-like input currents. Neural Computation, 16, 2101–2124.

    Article  PubMed  Google Scholar 

  • La Camera, G., Rauch, A., Thurbon, D., Luscher, H. R., Senn, W., & Fusi, S. (2006). Multiple time scales of temporal response in pyramidal and fast spiking cortical neurons. Journal of Neurophysiology, 96(6), 3448–64.

    Article  PubMed  Google Scholar 

  • Liu, Y. H., & Wang, X. J. (2001). Spike-frequency adaptation of a generalized leaky integrate-and-fire model neuron. Journal of Computational Neuroscience, 10, 25–45.

    Article  CAS  PubMed  Google Scholar 

  • Lundqvist, M., Herman, P., Palva, M., Palva, S., Silverstein, D., & Lansner, A. (2013). Stimulus detection rate and latency, firing rates and 1-40Hz oscillatory power are modulated by infra-slow fluctuations in a bistable attractor network model. NeuroImage, 83, 458–471.

    Article  PubMed  Google Scholar 

  • Lundstrom, B. N., Higgs, M. H., Spain, W. J., & Fairhall, A. L. (2008). Fractional differentiation by neocortical pyramidal neurons. Nature Neuroscience, 11, 1335–1342.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Lundstrom. BN., Hong, S., Higgs, MH., Fairhall, AL. (2008b) Two Computational Regimes of a Single-Compartment Neuron Separated by a Planar Boundary in Conductance Space. Neural Comput

  • Lundstrom, B. N., Famulare, M., Sorensen, L. B., Spain, W. J., & Fairhall, A. L. (2009). Sensitivity of firing rate to input fluctuations depends on time scale separation between fast and slow variables in single neurons. Journal of Computational Neuroscience, 27, 277–290.

    Article  PubMed  Google Scholar 

  • Lundstrom, B. N., Fairhall, A. L., & Maravall, M. (2010). Multiple timescale encoding of slowly varying whisker stimulus envelope in cortical and thalamic neurons in vivo. The Journal of Neuroscience, 30, 5071–5077.

    Article  CAS  PubMed  Google Scholar 

  • Madison, D. V., & Nicoll, R. A. (1984). Control of the repetitive discharge of rat CA 1 pyramidal neurones in vitro. The Journal of Physiology, 354, 319–331.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Monto, S., Palva, S., Voipio, J., & Palva, J. M. (2008). Very slow EEG fluctuations predict the dynamics of stimulus detection and oscillation amplitudes in humans. The Journal of Neuroscience, 28, 8268–8272.

    Article  CAS  PubMed  Google Scholar 

  • Moran, R., Pinotsis, D. A., & Friston, K. (2013). Neural masses and fields in dynamic causal modeling. Frontiers in Computational Neuroscience, 7, 57.

    Article  PubMed Central  PubMed  Google Scholar 

  • Nunez, P. L., & Srinivasan, R. (2006). Electric fields of the brain : the neurophysics of EEG. Oxford: Oxford University Press.

    Book  Google Scholar 

  • Oppenheim, A. V., Schafer, R. W., & Buck, J. R. (1999). Discrete-time signal processing. Upper Saddle River: Prentice Hall.

    Google Scholar 

  • Pozzorini, C., Naud, R., Mensi, S., & Gerstner, W. (2013). Temporal whitening by power-law adaptation in neocortical neurons. Nature Neuroscience, 16, 942–948.

    Article  CAS  PubMed  Google Scholar 

  • Puccini, G. D., Sanchez-Vives, M. V., & Compte, A. (2007). Integrated mechanisms of anticipation and rate-of-change computations in cortical circuits. PLoS Computational Biology, 3, e82.

    Article  PubMed Central  PubMed  Google Scholar 

  • Sanchez-Vives, M. V., Nowak, L. G., & McCormick, D. A. (2000). Cellular mechanisms of long-lasting adaptation in visual cortical neurons in vitro. The Journal of Neuroscience, 20, 4286–4299.

    CAS  PubMed  Google Scholar 

  • Schwindt, P. C., Spain, W. J., Foehring, R. C., Chubb, M. C., & Crill, W. E. (1988). Slow conductances in neurons from cat sensorimotor cortex in vitro and their role in slow excitability changes. Journal of Neurophysiology, 59, 450–467.

    CAS  PubMed  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 

  • Steriade, M. (2006). Grouping of brain rhythms in corticothalamic systems. Neuroscience, 137, 1087–1106.

    Article  CAS  PubMed  Google Scholar 

  • Steriade, M., Contreras, D., Curro Dossi, R., & Nunez, A. (1993). The slow (<1 Hz) oscillation in reticular thalamic and thalamocortical neurons: scenario of sleep rhythm generation in interacting thalamic and neocortical networks. The Journal of Neuroscience, 13, 3284–3299.

    CAS  PubMed  Google Scholar 

  • Thorson, J., & Biederman-Thorson, M. (1974). Distributed relaxation processes in sensory adaptation. Science, 183, 161–172.

    Article  CAS  PubMed  Google Scholar 

  • Trappenberg, T. P. (2002). Fundamentals of Computational Neuroscience. USA: Oxford University Press.

    Google Scholar 

  • Tripp, B. P., & Eliasmith, C. (2010). Population models of temporal differentiation. Neural Computation, 22, 621–659.

    Article  PubMed  Google Scholar 

  • Ulrych, T., & Lasserre, M. (1966). Minimum-phase. Canadian Journal of Exploration Geophysics, 2, 22–32.

    Google Scholar 

  • Vanhatalo, S., Palva, J. M., Holmes, M. D., Miller, J. W., Voipio, J., & Kaila, K. (2004). Infraslow oscillations modulate excitability and interictal epileptic activity in the human cortex during sleep. Proceedings of the National Academy of Sciences of the United States of America, 101, 5053–5057.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Vinagre, B., Podlubny, I., Hernandez, A., & Feliu, V. (2000). Some approximations of fractional order operators used in control theory and applications. Fractional Calculus and Applied Analysis, 3, 231–248.

    Google Scholar 

  • Wang, X. J. (1998). Calcium coding and adaptive temporal computation in cortical pyramidal neurons. Journal of Neurophysiology, 79, 1549–1566.

    CAS  PubMed  Google Scholar 

  • Wark, B., Lundstrom, B. N., & Fairhall, A. (2007). Sensory adaptation. Current Opinion in Neurobiology, 17, 423–429.

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  • Wendling, F., Bellanger, J. J., Bartolomei, F., & Chauvel, P. (2000). Relevance of nonlinear lumped-parameter models in the analysis of depth-EEG epileptic signals. Biological Cybernetics, 83, 367–378.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

To Adrienne Fairhall, Matt Higgs, and John Oakley for insightful discussions and comments on the manuscript, and to the University of Washington Department of Neurology for support.

Conflict of interest

The author declares that he has no conflict of interest.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Brian Nils Lundstrom.

Additional information

Action Editor: A. Compte

Appendix

Appendix

1.1 Linear rate models with exponential rate adaptation

Rate models are often of the form r = f(x-a), where f is some function of the input x with a feedback current a that depends on the firing rate r via some equation a(r) describing adaptation dynamics (Benda and Herz 2003, La Camera et al. 2004). In one of the simplest such models the firing rate is a linear function of the stimulus and the negative feedback adaptation. The firing rate increases linearly with increasing stimulus, while the adaptation variable decays exponentially with time:

$$ \begin{array}{l}r=mx-ga\\ {}\frac{da}{dt}=\frac{-a}{\tau }+kr\end{array} $$
(13)

where r is the firing rate, x is the time-varying stimulus, a is the adaptation variable, τ is the relaxation time constant, and m, g, and k are gain constants. One way to see that this is a high pass system is to express the rate dependence in terms of only the stimulus, thus eliminating the adaptation variable a, and then examine the equation in the frequency domain. Expressing r(t) as only a function of x(t):

$$ \begin{array}{l}\frac{d}{dt}\left[\frac{m}{g}x-\frac{r}{g}\right]=\frac{-1}{\tau}\left[\frac{mx}{g}-\frac{r}{g}\right]+kr\\ {}\frac{dr}{dt}=-r\left(\frac{1}{\tau }+gk\right)+\frac{m}{\tau }x+m\frac{dx}{dt}\end{array} $$
(14)

With τ eff = (1/τ + gk) −1 this becomes:

$$ {\tau}_{eff}\frac{dr}{dt}=-r+m\left(1-gk{\tau}_{eff}\right)x+m{\tau}_{eff}\frac{dx}{dt} $$
(15)

which gives in the frequency domain:

$$ R=m\left(\frac{1-gk{\tau}_{eff}+i\omega {\tau}_{eff}}{1+i\omega {\tau}_{eff}}\right)X=m\left(1-\frac{gk{\tau}_{eff}}{1+i\omega {\tau}_{eff}}\right)X $$
(16)

and in the time domain:

$$ r(t)=m\left(x(t)-gkx(t)\ast {e}^{-t/{\tau}_{eff}}\right) $$
(17)

where 0 ≤ gkτ eff ≤1. When gkτ eff =1, it is clear that the system is that of adapting high pass filtering. In fact, as gkτ eff increases, the amount of adaptation increases and the response to a step increase reflects a decay time course that is increasingly exponential-like. Notice that τ eff ≤ τ, as has been previously shown (Ermentrout 1998, Wang 1998, Tripp and Eliasmith 2010). The transfer function of Eq. (16) is identical to that derived in Benda and Herz (see Eq 5.8, 2003), where the gain parameters were expressed as slopes of the firing rate-input curve. The filter is:

$$ H\left(\omega \right)=m\frac{m_{ss}/m+i\omega {\tau}_{eff}}{1+i\omega {\tau}_{eff}} $$
(18)

where m and m SS are the slopes of the firing rate-input curve at an initial unadapted time and an adapted steady state, respectively. Here, m SS = m(1-gkτ eff ) and since τ eff = (1/τ + gk) −1 one can see that τ eff = τ m SS /m. The degree of negative feedback provided by adaptation is expressed in the reduction of τ eff from τ, the more adaptation the smaller τ eff . Thus, the model of Eq. (16) is functionally equivalent to the universal exponential adaptation model derived in Benda and Herz (2003) from a standard spike-generating model of membrane potential and currents. Here, linear dynamics were assumed initially rather than as an approximation during the derivation. By claiming the adaptation variable a can be described as a function of the firing rate r, rather than of individual spikes, one is assuming that τ is much greater than 1/r. In order to separate the fast spike generator from slow adaptation, one assumes that fluctuations in a do not markedly affect the time course of the spike generator (Benda and Herz 2003).

1.2 Linear rate models with power law rate adaptation

One notable function that can be approximated by exponentials is the power law. This unique function has the special property of being scale invariant such that its shape is unchanged despite scaling the x-axis. In other words, power laws do not have a characteristic time scale. Adaptation in neocortical neurons has been found to perform a function that can be approximated by power law filters (Lundstrom et al. 2008a, Lundstrom et al. 2010). Power laws can be approximated by an infinite sum of exponentials as can be seen by using the definition of the gamma function (Thorson and Biederman-Thorson 1974, Fairhall and Bialek 2002):

$$ \varGamma (k)\equiv {\displaystyle {\int}_0^{\infty }{\lambda}^{k-1}{e}^{-\lambda }d\lambda }. $$

Further, the Fourier transform of a power law is obtained by using the gamma function and setting λ = iωt and dλ = iω dt:

$$ {\left(i\omega \right)}^{-k}=\frac{1}{\varGamma (k)}{\displaystyle {\int}_0^{\infty }{t}^{k-1}{e}^{-i\omega t}dt} $$

Thus, by using the definition of the Fourier transform, the Fourier transform of a power law is:

$$ {\left(i\omega \right)}^{\alpha}\overset{F}{\leftrightarrow }k{t}^{-\left(\alpha +1\right)} $$
(19)

where in the time domain t ≤0 = 0. Experimentally, (iω) α is a reasonable filter that describes the effect of adaptation on the stimulus (Lundstrom et al. 2008a, Lundstrom et al. 2010) with a model of adaptation as follows, shown below in the frequency domain:

$$ R\left(\omega \right)=k{\left(i\omega \right)}^{\alpha }X\left(\omega \right)+b $$
(20)

where k and b are constants and α describes degree of adaptation. In the time domain, the equivalent filter for H(ω) = iω α with a finite number of frequencies can be expressed as

$$ h(t)=\frac{1}{\varGamma \left(1-\alpha \right)}\frac{d}{dt}{\displaystyle {\int}_0^t{\left(t-\tau \right)}^{-\alpha }}\frac{ \sin nt}{\pi t}d\tau, \kern0.75em \left(0\le \alpha <1\right) $$

where the following approximation for the delta function with finite range of frequencies is used (Arfken and Weber 1995):

$$ {\delta}_n(t)=\frac{ \sin nt}{\pi t}=\frac{1}{2\pi }{\displaystyle {\int}_{-n}^n \exp (iwt)dw}. $$

However, this form of the filter is difficult to use. An easier-to-use filter can be found by using the approach described above, where resulting dynamics are such that either the adaptation variable or the response to a step impulse has the form of a power law. For example, assuming power law dynamics of the adaptation variable with h(t) = kt , the appropriate model would be:

$$ r(t)=mx(t)-mk{\displaystyle \int x\left(\tau \right){\left(t-\tau \right)}^{-\alpha }d\tau +b} $$
(21)

with rate reponse r to a time-varying stimulus x with constants m, b, and k, with α controlling the degree of adaptation. Alternately, if a model that has precisely a power law decay of power α to a step increase is desired, the derivative of the power law yields the result:

$$ r(t)=-m{\displaystyle \int x\left(\tau \right){\left(t-\tau \right)}^{-\alpha -1}d\tau }+b $$
(22)

with constants m and b and α controlling the degree of adaptation, as displayed in Fig. 7. Practically, discrete time power law filters as in Eqs. (21) and (22) are limited approximations for the frequency domain filter of (iω) α. They are not defined at time zero, and thus the initial response to a stimulus, which should not be dependent on any adaptation dynamics, may be discontinuous with the subsequent decay. In addition, a power law has an infinitely long tail, which can only be approximated by a finite length filter.

Fig. 7
figure 7

a A power law filter h with α =0.2, according to Eq (22). The magnitude of the filter was normalized such that its first point was equal to 1, which is not displayed for convenience. b The response of h convolved with a step function. Note that the initial peak response is determined by the arbitrary first point of the filter. The power law does not have a defined steady state, and thus eventually decays to zero given an infinite filter length, in contrast to an exponential filter which decays to some steady state

1.3 The relationship between magnitude and phase in minimum phase systems

Typically there is no precise relationship between the frequency-response magnitude, or gain, for a linear time invariant system and its phase. However, for systems characterized by a rational response function there is a relationship and for a subset of rational systems termed minimum phase systems, specifying the phase determines the magnitude to within a single scale factor, and vice versa (Oppenheim et al. 1999). A rational LTI system that is causal, stable, and that in the Laplace domain has all zeros on the left side of the s-plane (or inside the unit circle of the z-plane for the z-transform) is a minimum phase system (Ulrych and Lasserre 1966, Oppenheim et al. 1999). Causality assumes that the filter is right-sided, that is, it is equal to zero for negative time points, meaning that the output cannot precede the input. Stability implies a bounded output sequence for every bounded input sequence, which suggests that the discrete time impulse response is absolutely summable:

$$ {\displaystyle \sum_{n=-\infty}^{\infty}\left|h\left[n\right]\right|}<\infty $$
(23)

The last requirement for a minimum phase system, which amounts to assuming that the inverse system H is also causal and stable, is equivalent to requiring all τ n in the adaptation filter models to be nonnegative. Specifically, in the Laplace domain Eq. (10) has a numerator of the form:

$$ \left(1/{\tau}_1+s\right)\left(1/{\tau}_2+s\right)\ldots \left(1/{\tau}_n+s\right) $$
(24)

where zeros can be seen to be negative, that is, on the left side of the s-plane, as long as all τn are nonnegative. This implies that the magnitude and phase of the system are related through the Hilbert transfrom (Ulrych and Lasserre 1966, Oppenheim et al. 1999):

$$ \angle H\left(\omega \right)=-\frac{1}{\pi }PV{\displaystyle {\int}_{-\infty}^{\infty}\frac{ \log \left|H(x)\right|}{\omega -x}dx} $$
(25)

where ∠ H(ω) is the phase and PV signfies the principal value of the Cauchy integral of the Hilbert transform. The Hilbert transform can be understood as convolving a function with the filter 1/(π x), where the function in this case is the logarithm of the magnitude of H(ω). Approaching negative and positive infinity as values approach zero from the negative and positive side, respectively, 1/x is effectively a differentiating filter. Thus, from Eq. (25) the phase is related to the derivative of the magnitude. The phase is positive for high pass filters, meaning that rate adaptation gives rise to phase leads.

1.4 Implementing the Jansen and Rit model of EEG oscillations

The standard Jansen and Rit model (Jansen and Rit 1995) specified as a system differential equations was implemented with a fourth order Runge Kutta solver as:

  • $$ {y}_1^{\prime }={y}_4 $$
  • $$ {y}_2^{\prime }={y}_5 $$
  • $$ {y}_3^{\prime }={y}_6 $$
  • $$ {y}_4^{\prime }={k}_e/{\tau}_e\left(S+{c}_2Sgm\left({c}_1{y}_3\right)\right)-2/{\tau}_e{y}_4-1/{\tau}_e^2{y}_1 $$
  • $$ {y}_5^{\prime }={k}_i/{\tau}_i{c}_4Sgm\left({c}_3{y}_3\right)-2/{\tau}_i{y}_5-1/{\tau}_i^2{y}_2 $$
  • $$ {y}_6^{\prime }={k}_e/{\tau}_eSgm\left({y}_1-{y}_2\right)-2/{\tau}_e{y}_6-1/{\tau}_e^2{y}_3 $$

with sigmoid Sgm(v) = e 0 / [1 + exp(0.56(6 - v), external stimulus (S), constants c 1–4= [135 108 33.75 33.75], and parameters k e =3.25 mV, k i =22 mV, τ e =10 ms, τ i =20 ms, and e 0 = 5 Hz, unless otherwise noted. Primes indicate temporal first-order derivatives.ODEs were solved by a fourth-order fixed step Runge Kutta solver with dt =1–5 ms, with identical results obtained regardless. The stimulus was either constant or a sine wave. To stimulate rate adaptation with three time scales, the following system was implemented:

  • $$ {y}_1^{\prime }={y}_4 $$
  • $$ {y}_2^{\prime }={y}_5 $$
  • $$ {y}_3^{\prime }={y}_6 $$
  • $$ {y}_4^{\prime }={y}_{10} $$
  • $$ {y}_5^{\prime }={y}_{11} $$
  • $$ {y}_6^{\prime }={k}_e/{\tau}_eSgm\left({y}_7\right)-2/{\tau}_e{y}_6-1/{\tau}_e^2{y}_3 $$
  • $$ {y}_7^{\prime }={y}_8 $$
  • $$ {y}_8^{\prime }={y}_9 $$
  • $$ {y}_9^{\prime }=\left({y}_{10}^{\prime }-{y}_{11}^{\prime}\right)+{k}_1\left({y}_{10}-{y}_{11}\right)+{k}_2\left({y}_4-{y}_5\right)+{k}_3\left({y}_1-{y}_2\right)-{k}_4{y}_9-{k}_5{y}_8-{k}_6{y}_7 $$
  • $$ {y}_{10}^{\prime }={k}_e/{\tau}_e\left(S^{\prime }+{c}_1{c}_2{y}_6Sgm^{\prime}\left({c}_1{y}_3\right)-2/{\tau}_e{y}_{10}-/{\tau}_e^2{y}_4\right) $$
  • $$ {y}_{11}^{\prime }={k}_i/{\tau}_i{c}_3{c}_4{y}_6Sgm^{\prime}\left({c}_3{y}_3\right)-2/{\tau}_i{y}_{11}-1/{\tau}_i^2{y}_5 $$

with the following additional parameters:

  • $$ {k}_1=1/{\tau}_1+1/{\tau}_2+1/{\tau}_3 $$
  • $$ {k}_2=1/\left({\tau}_1+{\tau}_2\right)+1/\left({\tau}_1+{\tau}_3\right)+1/\left({\tau}_2+{\tau}_3\right) $$
  • $$ {k}_3=1/\left({\tau}_1{\tau}_2{\tau}_3\right) $$
  • $$ {k}_4={k}_1=k{g}_1+k{g}_2+k{g}_3 $$
  • $$ {k}_5={k}_2+k{g}_1/\left({\tau}_2+{\tau}_3\right)+k{g}_2\left({\tau}_1+{\tau}_3\right)+k{g}_3/\left({\tau}_1+{\tau}_2\right) $$
  • $$ {k}_6={k}_3+k{g}_1/\left({\tau}_2{\tau}_3\right)+k{g}_2\left({\tau}_1{\tau}_3\right)+k{g}_3/\left({\tau}_1{\tau}_2\right) $$

where kg n and τn as parameters governing the gain and time constant of each exponential filter. Similar smaller systems of ODEs were simulated for the cases with one or two exponential filters with identical results as those obtained by setting the appropriate gain parameters to zero above. This model can also be simulated with the corresponding filter or integral equations rather than the above differential equations. The overall output of the model is the difference between the post-synaptic excitatory and inhibitory membrane potentials to the pyramidal neurons (y 1 – y 2 ), consistent with what is thought to primarily underlie the signal of electroencephalography (Nunez and Srinivasan 2006).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lundstrom, B.N. Modeling multiple time scale firing rate adaptation in a neural network of local field potentials. J Comput Neurosci 38, 189–202 (2015). https://doi.org/10.1007/s10827-014-0536-2

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-014-0536-2

Keywords

Navigation