Skip to main content
Log in

The role of synaptic facilitation in spike coincidence detection

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

Abstract

Using a realistic model of activity dependent dynamical synapse, which includes both depressing and facilitating mechanisms, we study the conditions in which a postsynaptic neuron efficiently detects temporal coincidences of spikes which arrive from N different presynaptic neurons at certain frequency f. A numerical and analytical treatment of that system shows that: (1) facilitation enhances the detection of correlated signals arriving from a subset of presynaptic excitatory neurons, and (2) the presence of facilitation yields to a better detection of firing rate changes in the presynaptic activity. We also observed that facilitation determines the existence of an optimal input frequency which allows the best performance for a wide (maximum) range of the neuron firing threshold. This optimal frequency can be controlled by means of facilitation parameters. Finally, we show that these results are robust even for very noisy signals and in the presence of synaptic fluctuations produced by the stochastic release of neurotransmitters.

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

Similar content being viewed by others

Notes

  1. Note that it is the synaptic conductance, rather than the synaptic current, which depends on A SE ·y(t). Our assumption for the current, however, is a good approximation when the membrane potential V(t) is below the firing threshold V th and τ m  ≫ τ in , so that V(t) remains constant during the temporal variation of the synaptic conductance.

  2. The specific value of Δ is not too critical for the results found if moderate values are used. In particular, it is convenient to have Δ~τ ref since this is a natural window for spike detection.

References

  • Abbott, L. F., & Regehr, W. G. (2004). Synaptic computation. Nature, 431, 796–803.

    Article  PubMed  CAS  Google Scholar 

  • Abbott, L. F., Varela, J. A., Sen, K., & Nelson, S. B. (1997). Synaptic depression and cortical gain control. Science, 275(5297), 220–224.

    Article  PubMed  CAS  Google Scholar 

  • Azouz, R., & Gray, C. M. (2000). Dynamic spike threshold reveals a mechanism for synaptic coincidence detection in cortical neurons in vivo. Proceedings of the National Academy of Sciences of the United States of America, 97(14), 8110–8115.

    Article  PubMed  CAS  Google Scholar 

  • Bertram, R., Sherman, A., & Stanley, E. F. (1996). Single-domain/bound calcium hypothesis of transmitter release and facilitation. Journal of Neurophysiology, 75, 1919–1931.

    PubMed  CAS  Google Scholar 

  • Buia, C. I., & Tiesinga, P. H. E. (2005). Rapid temporal modulation of synchrony in cortical interneuron networks with synaptic plasticity. Neurocomputing, 65–66, 809–815.

    Article  Google Scholar 

  • Cortes, J. M., Torres, J. J., Marro, J., Garrido, P. L., & Kappen, H. J. (2006). Effects of fast presynaptic noise in attractor neural networks. Neural Computation, 18, 614–633.

    Article  PubMed  CAS  Google Scholar 

  • de la Rocha, J., & Parga, N. (2005). Short-term synaptic depression causes a non-monotonic response to correlated stimuli. Journal of Neuroscience, 25, 8416–8431.

    Article  PubMed  Google Scholar 

  • Destexhe, A., & Marder, E. (2004). Plasticity in single neuron and circuit computations. Nature, 431, 789–795.

    Article  PubMed  CAS  Google Scholar 

  • Dobrunz, L. E., & Stevens, C. F. (1997). Heterogeneity of release probability, facilitation, and depletion at central synapses. Neuron, 18, 995–1008.

    Article  PubMed  CAS  Google Scholar 

  • Holcman, D., & Tsodyks, M. (2006). The emergence of up and down states in cortical networks. PLoS Computational Biology, 2(3), 174–181.

    Article  Google Scholar 

  • Hopfield, J. J. (1982). Neural networks and physical systems with emergent collective computational abilities. Proceedings of the National Academy of Sciences of the United States of America, 79, 2554–2558.

    Article  PubMed  CAS  Google Scholar 

  • Kistler, W., & van Hemmen, J. (1999). Short-term synaptic plasticity and network behavior. Neural Computation, 11, 1579–1594.

    Article  PubMed  CAS  Google Scholar 

  • Koch, C. (1999). Biophysics of computation: Information processing in single neurons. London: Oxford University Press.

    Google Scholar 

  • Markram, H., Wang, Y., & Tsodyks, M. (1998). Differential signaling via the same axon of neocortical pyramidal neurons. Proceedings of the National Academy of Sciences of the United States of America, 95, 5323–5328.

    Article  PubMed  CAS  Google Scholar 

  • Matveev, V., & Wang, X. J. (2000). Differential short-term synaptic plasticity and transmission of complex spike trains: To depress or to facilitate? Cerebral Cortex, 10, 1143–1153.

    Article  PubMed  CAS  Google Scholar 

  • McAdams, C., & Maunsell, J. (1999). Effects of attention on orientation-tuning functions of single neurons in macaque cortical area v4. Journal of Neuroscience, 19, 431–441.

    PubMed  CAS  Google Scholar 

  • Pantic, L., Torres, J. J., & Kappen, H. J. (2003). Coincidence detection with dynamic synapses. Network: Computation in Neural Systems, 14, 17–33.

    Article  Google Scholar 

  • Pantic, L., Torres, J. J., Kappen, H. J., & Gielen, S. C. A. M. (2002). Associative memmory with dynamic synapses. Neural Computation, 14, 2903–2923.

    Article  PubMed  Google Scholar 

  • Plesser, H. E., & Gerstner, W. (2000). Noise in integrate-and-fire neurons: From stochastic input to escape rates. Neural Computation 12, 367–384.

    Article  PubMed  CAS  Google Scholar 

  • Torres, J. J., Cortes, J., Marro, J., & Kappen, H. (2007). Competition between synaptic depression and facilitation in attractor neural networks. Neural Computation, 19(10), q-bio.NC/0604019.

  • Tsodyks, M. V., & Markram, H. (1997). The neural code between neocortical pyramidal neurons depends on neurotransmitter release probability. Proceedings of the National Academy of Sciences of the United States of America, 94, 719–723.

    Article  PubMed  CAS  Google Scholar 

  • Tsodyks, M. V., Pawelzik, K., & Markram, H. (1998). Neural networks with dynamic synapses. Neural Computation, 10, 821–835.

    Article  PubMed  CAS  Google Scholar 

  • Wang, Y., Markram, H., Goodman, P. H., Berger, T. K., Ma, J., & Goldman-Rakic, P. S. (2006). Heterogeneity in the pyramidal network of the medial prefrontal cortex. Nature Neuroscience, 9, 534–542.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the MEyC–FEDER project FIS2005-00791 and the Junta de Andalucía project FQM–165. We thank useful discussions with J. Marro.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jorge F. Mejías.

Additional information

Action Editor: Xiao-Jing Wang

Appendix

Appendix

1.1 Analytical derivation of the error function E(f,V th )

In this section we derived analytical expressions for the functions appearing in the definition of the error function (5) used to obtain theoretically the regions for good spike CD in the (f,V th) parameter space.

First, we assume that the total presynaptic current can be divided in two terms: a signal term containing the correlated embedded signal and a noise term formed by the background of uncorrelated spikes.

1.1.1 Noise contribution

To take into account the noise generated by N − M uncorrelated spikes trains, we assume that the current at time t = t * + τ generated by a single spike arriving to the synapse i at time t * is given by

$$ I_{i}(\tau,t^*)=I_{peak}\exp(-\tau/\tau_{in}) $$
(20)

where I peak represents the averaged stationary EPSC amplitude obtained after stimulation with a periodic spike train, assumption that we also suppose valid for a Poisson distributed spike train. After this consideration, one easily obtains from Eqs. (13) that

$$ I_{peak}=A_{SE}\frac{U_\infty(1-\exp(-1/f\tau_{rec}))}{1-(1-U_\infty)\exp(-1/f\tau_{rec})} \label{peak} $$
(21)

with U  ∞  = u  ∞ (1 − U SE ) + U SE , where u  ∞  is the value of u(t) in the stationary state (t→ ∞). For a periodic spike train, u  ∞  is given by

$$ u_\infty=U_{SE}\frac{\exp(-1/f\tau_{fac})}{1-(1-U_{SE})\exp(-1/f\tau_{fac})}. $$
(22)

We can compute the mean noise contribution of the current and fluctuations using the standard expressions

$$\begin{array}{*{20}l} {{I_{{noise}} \equiv {\left\langle I \right\rangle },} \hfill} \\ {{\sigma ^{2}_{{I_{{noise}} }} \equiv {\left\langle {I^{2} } \right\rangle } - {\left\langle I \right\rangle }^{2} } \hfill} \\ \end{array} $$
(23)

From these definitions and using the central limit theorem we obtain

$$ I_{noise}\!=\!(N-M)A_{SE}f\tau_{in}U_{\infty}\frac{1-\exp(-1/f\tau_{rec})}{1\! -\! (1-U_\infty)\exp(\! -1/f\tau_{rec})} $$
(24)

where we assumed that τ in  ≪ τ rec .

If we neglect fluctuations (\(\sigma_{I_{noise}}=0\)), we can write V noise  = R in I noise . Using this expression one can compute N falses taking into account that false firing occurs when V noise  > V th so by a direct integration of Eq. (4) in a period of time T gives N falses  ≈ T/{τ ref  − τ m ln (1 − V th /V noise )} (Koch 1999). Now using that f = N inputs /T, we finally obtain as in Pantic et al. (2003)

$$ N_{falses}=\frac{\theta(V_{noise}-V_{th})N_{inputs}}{f(\tau_{ref}-\tau_{m}\ln(1-V_{th}/V_{noise}))} \label{falses} $$
(25)

where θ(x) is the Heaviside step function, which takes into account that for V noise  < V th N falses  = 0.

To take into account fluctuations of I noise one can use the so called hazard function approximation (Plesser and Gerstner 2000) but it has been reported that it gives the same results as those obtained using the formula (25) for high frequencies and, on the contrary to the expression (25), it does not work properly for small frequencies (Pantic et al. 2003). Therefore, hereafter we will neglect fluctuations in I noise and use Eq. (25) as an approximatively valid expression to analytically compute N falses .

1.1.2 Signal contribution

To analyse the signal contribution (arising from M coincident spikes) we used the same method developed in Pantic et al. (2003) for the case of only depressing synapses. That is, assuming that V(0;t *) is the membrane potential at t = t * when M coincident spikes arrive, by direct integration of the Eq. (4) the membrane potential at time t = t * + τ is

$$ V(\tau;t^{*})=e^{\tau/\tau_{m}}\left \{ V(0;t^*) + \frac{R_{in}M I_{peak}}{\tau_{m}\alpha}[e^{\alpha\tau}-1]\right \} \label{voltage} $$
(26)

where \(\alpha= \frac{\tau_{in}-\tau_{m}}{\tau_{in}\tau_{m}}\) and I peak is given by Eq. (21) including all the effects due to synaptic depression and facilitation. If the next signal event (M coincident spikes) occurs at t = t′ one can obtain the following recurrence relation:

$$ V(0;t')=e^{\Delta t/\tau_{m}}\left \{V(0;t^*) + \frac{R_{in}M I_{peak}}{\tau_{m}\alpha}[e^{\alpha\Delta t}-1]\right \} $$
(27)

with Δt = t′ − t *, which allows for computing the stationary value for the membrane potential at the exact time of the signal event arrival (see also Kistler and van Hemmen 1999), that is:

$$ V_{st}=e^{-\Delta t/\tau_{m}}\frac{R_{in}M I_{peak}}{\tau_{m}\alpha}\frac{e^{\alpha \Delta t} -1}{(1-e^{-\Delta t/\tau_{m}})}. $$
(28)

We define V signal as the maximum of the membrane potential reached between the arrival of two consecutive signal events separated by a time Δt. This can be easily computed from Eq. (26) with V(0,t *) replaced by V st :

$$ V_{signal}=\Biggl[\frac{\tau_{m}(1-\exp(-1/f\tau_{m}))}{\tau_{in}(1-\exp(-1/f\tau_{in}))}\Biggr]^{\frac{\tau_{m}}{\tau_{in}-\tau_{m}}}R_{in}MI_{peak} $$
(29)

where we consider \(\tau=\Delta t\backsimeq 1/f.\)

The expression of V signal allows for an evaluation of the number of failures assuming that N failures  = N inputs  − N hits . Then, one obtains by direct integration of Eq. (4) and using the same reasoning as for N falses case that

$$N_{{failures}} = N_{{inputs}} {\left[ {1 - \frac{{\theta {\left( {V_{{noise}} + V_{{signal}} - V_{{th}} } \right)}}}{{f{\left[ {\tau _{{ref}} - \tau _{m} \ln {\left( {1 - {{\left( {V_{{th}} - V_{{signal}} } \right)}} \mathord{\left/ {\vphantom {{{\left( {V_{{th}} - V_{{signal}} } \right)}} {V_{{noise}} }}} \right. \kern-\nulldelimiterspace} {V_{{noise}} }} \right)}} \right]}}}} \right]}$$
(30)

where we have considered a hit event every time V noise  + V signal reach V th . Note that from Eq. (30) if V noise  + V signal  < V th we will have N failures  = N inputs .

Expression for N falses , N failures allows to theoretically compute the number of errors in the CD maps.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Mejías, J.F., Torres, J.J. The role of synaptic facilitation in spike coincidence detection. J Comput Neurosci 24, 222–234 (2008). https://doi.org/10.1007/s10827-007-0052-8

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-007-0052-8

Keywords

Navigation