On the mechanisms underlying the depolarization block in the spiking dynamics of CA1 pyramidal neurons | Journal of Computational Neuroscience Skip to main content
Log in

On the mechanisms underlying the depolarization block in the spiking dynamics of CA1 pyramidal neurons

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

Abstract

Under sustained input current of increasing strength neurons eventually stop firing, entering a depolarization block. This is a robust effect that is not usually explored in experiments or explicitly implemented or tested in models. However, the range of current strength needed for a depolarization block could be easily reached with a random background activity of only a few hundred excitatory synapses. Depolarization block may thus be an important property of neurons that should be better characterized in experiments and explicitly taken into account in models at all implementation scales. Here we analyze the spiking dynamics of CA1 pyramidal neuron models using the same set of ionic currents on both an accurate morphological reconstruction and on its reduction to a single-compartment. The results show the specific ion channel properties and kinetics that are needed to reproduce the experimental findings, and how their interplay can drastically modulate the neuronal dynamics and the input current range leading to a depolarization block. We suggest that this can be one of the rate-limiting mechanisms protecting a CA1 neuron from excessive spiking activity.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
¥17,985 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Japan)

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
Fig. 12

Similar content being viewed by others

References

  • Andrasfalvy, B. K., & Magee, J.C. (2001). Distance-dependent increase in AMPA receptor number in the dendrites of adult hippocampal CA1 pyramidal neurons. Journal of Neuroscience, 21, 9151–9159.

    PubMed  CAS  Google Scholar 

  • Ayala, G. F., Dichter, M., Gumnit, R. J., Matsumoto, H., & Spencer, W. A. (1973). Genesis of epileptic interictal spikes. New knowledge of cortical feedback systems suggests a neurophysiological explanation of brief paroxysms. Brain Research, 52, 1–17.

    Article  PubMed  CAS  Google Scholar 

  • Colbert, C. M., & Pan, E. (2002). Ion channel properties underlying axonal action potential initiation in pyramidal neurons. Nature Neuroscience, 5, 533–538.

    Article  PubMed  CAS  Google Scholar 

  • Dichter, M. A., & Ayala, G. F. (1987). Cellular mechanisms of epilepsy: A status report. Science, 237, 157–164.

    Article  PubMed  CAS  Google Scholar 

  • Gasparini, S., & Magee, J. C. (2002). Phosphorylation-dependent differences in the activation properties of distal and proximal dendritic Na  +  channels in rat CA1 hippocampal neurons. Journal of Physiology, 541.3, 665–672.

    Article  Google Scholar 

  • Golomb, D., Yue, C., & Yaari, Y. (2006). Contribution of persistent Na  +  current and M-type K  +  current to somatic bursting in CA1 pyramidal cells: Combined experimental and modeling study. Journal of Neurophysiology, 96, 1912–1926.

    Article  PubMed  CAS  Google Scholar 

  • Hassard, B. (1978). Bifurcation of periodic solutions of the Hodgkin–Huxley model for the squid giant axon. Journal of Theoretical Biology, 71, 401–420.

    Article  PubMed  CAS  Google Scholar 

  • Hemond, P., Epstein, D., Boley, A., Migliore, M., Ascoli, G. A., & Jaffe, D. B. (2008). Distinct classes of pyramidal cells exhibit mutually exclusive firing patterns in hippocampal area CA3b. Hippocampus, 18, 411–424.

    Article  PubMed  Google Scholar 

  • Hines, M. L., & Carnevale, N. T. (2003). The NEURON simulation environment. In The handbook of brain theory and neural networks (2nd ed., pp. 769–773). Cambridge: MIT Press.

    Google Scholar 

  • Hoffman, D. A., & Johnston, D. (1998). Downregulation of transient K+ channels in dendrites of hippocampal CA1 pyramidal neurons by activation of PKA and PKC. Journal of Neuroscience, 18, 3521–3528.

    PubMed  CAS  Google Scholar 

  • Hoffman, D. A., Magee, J. C., Colbert, C. M., & Johnston, D. (1997). K+ channel regulation of signal propagation in dendrites of hippocampal pyramidal neurons. Nature, 387, 869–875.

    Article  PubMed  CAS  Google Scholar 

  • Ito, H., & Schuman, E. M. (2009). Distance-dependent homeostatic synaptic scaling mediated by A-type potassium channels. Frontiers in Cellular Neuroscience, 3, 1–15.

    Article  Google Scholar 

  • Jiang, L., Sun, S., Nedergaard, M., & Kang, J. (2000). Paired pulse modulation at individual GABAergic synapses in rat hippocampus. Journal of Physiology, 523.2, 425–439.

    Article  Google Scholar 

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

    Google Scholar 

  • Magee, J. C. (1998). Dendritic hyperpolarization-activated currents modify the integrative properties of hippocampal CA1 pyramidal neurons. Journal of Neuroscience, 18, 7613–7824.

    PubMed  CAS  Google Scholar 

  • Marasco, A., & Romano, A. (2001). Scientific computing with mathematica: Mathematical problems for ordinary differential equations. Boston: Birkhauser. ISBN 0-8176-4205-6.

    Google Scholar 

  • Marie, H., Morishita, W., Yu, X., Calakos, N., & Malenka, R. C. (2005). Generation of silent synapses by acute in vivo expression of CaMKIV and CREB. Neuron, 45, 741–752.

    Article  PubMed  CAS  Google Scholar 

  • McCormick, D. A., & Contreras, D. (2001). On the cellular and network bases of epileptic seizures. Annual Review of Physiology, 63, 815–846.

    Article  PubMed  CAS  Google Scholar 

  • Megias, M., Emri, Z., Freund, T. F., & Gulyás, A. I. (2001). Total number and distribution of inhibitory and excitatory synapses on hippocampal CA1 pyramidal cells. Neuroscience, 102, 527–540.

    Article  PubMed  CAS  Google Scholar 

  • Meng, X., Lu, Q., & Rinzel, J. (2011). Control of firing patterns by two transient potassium currents: Leading spike, latency, bistability. Journal of Computational Neuroscience, 31, 117–136.

    Article  PubMed  Google Scholar 

  • Migliore, M. (1996). Modeling the attenuation and failure of action potentials in the dendrites of hippocampal neurons. Biophysical Journal, 71, 2394–2403.

    Article  PubMed  CAS  Google Scholar 

  • Migliore, M., Hoffman, D., Magee, J., & Johnston D. (1999). Role of an A-type K  +  conductance in the back-propagation of action potentials in the dendrites of hippocampal pyramidal neurons. Journal of Computational Neuroscience, 7, 5–15.

    Article  PubMed  CAS  Google Scholar 

  • Migliore, M., & Shepherd, G. M. (2002). Emerging rules for the distributions of active dendritic conductances. Nature Reviews. Neuroscience, 3, 362–370.

    Article  PubMed  CAS  Google Scholar 

  • Migliore, M., & Shepherd, G. M. (2005). Opinion: An integrated approach to classifying neuronal phenotypes. Nature Reviews. Neuroscience, 6, 810–818.

    Article  PubMed  CAS  Google Scholar 

  • Moczydlowski, E., & Latorre, R. (1983). Gating kinetics of 2 + -activated K  +  channels from rat muscle incorporated into planar lipid bilayers. Journal of General Physiology, 82, 511–542.

    Article  PubMed  CAS  Google Scholar 

  • Nowacki, J., Osinga, H. M., Browna, J. T., Randall, A. D., & Tsaneva-Atanasova, K. (2011). A unified model of CA1/3 pyramidal cells: An investigation into excitability. Progress in Biophysics and Molecular Biology, 105, 34–48.

    Article  PubMed  Google Scholar 

  • Pinsky, P. F., & Rinzel, J. (1994). Intrinsic and network rhythmogenesis in a reduced Traub model for CA3 neurons. Journal of Computational Neuroscience, 1, 39.

    Article  PubMed  CAS  Google Scholar 

  • Poirazi, P., Brannon, T., & Mel, B. W. (2003). Arithmetic of subthreshold synaptic summation in a model CA1 pyramidal cell. Neuron, 37, 977–987.

    Article  PubMed  CAS  Google Scholar 

  • Poirazi, P., Brannon, T., & Mel, B. W. (2003). Pyramidal neuron as 2-layer neural network. Neuron, 37, 989–999.

    Article  PubMed  CAS  Google Scholar 

  • Pospischil, M., Toledo-Rodriguez, M., Monier, C., Piwkowska, Z., Bal, T., Frégnac, Y., et al. (2008). Minimal Hodgkin–Huxley type models for different classes of cortical and thalamic neurons. Biological Cybernetics, 99, 427–441.

    Article  PubMed  Google Scholar 

  • Remy, S., Beck, H., & Yaari, Y. (2010). Plasticity of voltage-gated ion channels in pyramidal cell dendrites. Current Opinion in Neurobiology, 20, 503–509.

    Article  PubMed  CAS  Google Scholar 

  • Rüdiger, S. (2010). Practical bifurcation and stability analysis practical bifurcation and stability analysis. In Springer series: Interdisciplinary applied mathematics (Vol. 5, 3rd ed.). New York: Springer.

    Google Scholar 

  • Samsonovich, A. V., & Ascoli, G. A. (2005). Statistical determinants of dendritic morphology in hippocampal pyramidal neurons: A hidden Markov model. Hippocampus, 15, 166–183.

    Article  PubMed  Google Scholar 

  • Scorza, C. A., Araujo, B. H., Leite, L. A., Torres, L. B., Otalora, L. F., Oliveira, M. S., et al. (2011). Morphological and electrophysiological properties of pyramidal-like neurons in the stratum oriens of Cornu ammonis 1 and Cornu ammonis 2 area of Proechimys. Neuroscience, 177, 252–268.

    Article  PubMed  CAS  Google Scholar 

  • Shah, M. M., Migliore, M., Valencia, I., Cooper, E. C., & Brown, D. A. (2008). Functional significance of axonal Kv7 channels in hippocampal pyramidal neurons. Proceedings of the National Academy of Sciences of the United States of America, 105, 7869–7874.

    Article  PubMed  CAS  Google Scholar 

  • Spruston, N., Schiller, Y., Stuart, G., & Sakmann, B. (1995). Activity-dependent action potential invasion and calcium influx into hippocampal CA1 dendrites. Science, 268, 297–300.

    Article  PubMed  CAS  Google Scholar 

  • Traub, R. D., & Wong, R. K. (1982). Cellular mechanism of neuronal synchronization in epilepsy. Science, 216, 745–747.

    Article  PubMed  CAS  Google Scholar 

  • Traub, R. D., Wong, R. K., Miles, R., & Michelson, H. (1991). A model of a CA3 hippocampal pyramidal neuron incorporating voltage-clamp data on intrinsic conductances. Journal of Neurophysiology, 66, 635–650.

    PubMed  CAS  Google Scholar 

  • Troy, W. C. (1974). Oscillatory phenomena in nerve conduction equations. Ph.D. dissertation, SUNY at Buffalo.

  • Troy, W. C. (1978). The bifurcation of periodic solutions in the Hodgkin–Huxley equations. Quarterly of Applied Mathematics, 36, 73–83.

    Google Scholar 

  • Xiong, W., & Chen, W. R. (2002). Dynamic gating of spike propagation in the mitral cell lateral dendrites. Neuron, 34, 115–126.

    Article  PubMed  CAS  Google Scholar 

  • Zemankovics, R., Káli, S., Paulsen, O., Freund, T. F., & Hájos, N. (2010). Differences in subthreshold resonance of hippocampal pyramidal cells and interneurons: The role of h-current and passive membrane characteristics. Journal of Physiology, 588(12), 2109–2132.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgements

Financial support from “Compagnia di San Paolo” is gratefully acknowledged. We thank Drs. S. Cuomo and P. De Michele (Department of Mathematics and Applications “Renato Caccioppoli”, University of Naples Federico II) for assistance in running the parallel version of our morphological model and for the use of the S.Co.P.E. Grid infrastructure of University of Naples Federico II.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daniela Bianchi.

Additional information

Action Editor: Frances K. Skinner

D. Bianchi and A. Marasco contributed equally to this work.

Appendix

Appendix

1.1 Ionic currents

The full morphological model includes the following 17 types of ionic channels, most of them distributed non-uniformly along the somatodendritic region. More precisely, the K M and K A potassium currents are those described by Shah et al. (2008), whereas the other types of currents have been distributed as in model of Poirazi et al. (2003b).

  • Soma: a leak current, a transient sodium (NaT) current, a delay-rectifier potassium (K DR) current, an A-type potassium current (K A), a M type potassium current (K M), a mixed conductance hyper-polarization activated h-current, three types of voltage dependent calcium currents (namely LVA T-type current, a HVA R-type current, a HVA L-type current), two types of calcium dependent potassium currents (a slow AHP current and a medium fast AHP current);

  • Axon: a NaT current, a K DR current, a leak current, a K A current and K M current;

  • Basal dendrites: a NaT current, a K DR current, a leak current, a K A current and a h-current;

  • Apical trunk: all currents included in the soma with the exception of K M current and the insertion of a K A current (the HVA R and L type are different from those in soma);

  • Apical dendrites: all currents included in the apical trunk with an insertion of a persistent sodium current.

Ionic currents were modeled following a Hodgkin-Huxley-like formalism as follows:

$$ I_{j}=\bar{g}_{j}m_{j}^{a_{j}}h_{j}^{b_{j}}(V-E_{j}), $$
(2)

where \(\bar{g}_{j}\) represents the maximal ionic conductance, (m j, h j) and (a j, b j) are the gating variables for activation and inactivation and their exponents, respectively, and E j is the reversal potential associated with the particular ion or ions that make up the current. The dynamics of the gating variables (m j , h j ) is governed by an ODE of the form

$$ \frac{d\chi}{dt}= \frac{\chi_{\infty}-\chi}{\tau_{\chi}} $$
(3)

where the activation and inactivation steady-state functions χ  ∞ , and their time constant τ χ are given by

$$ \chi _{\infty }=\frac{\alpha _{\chi }(V)}{\alpha _{\chi }(V)+\beta _{\chi }(V)},\quad \chi _{\infty }=\frac{1}{1+\exp \left[ \left( V-V_{\chi }^{1/2}\right) /k_{\chi }\right] }, $$
(4)
$$ \tau_{\chi } =\frac{Q}{\alpha _{\chi }(V)+\beta _{\chi }(V)}, $$
(5)
$$ \tau _{\chi } =\tau _{\chi }^{0}+\frac{\bar{\tau}_{\chi }}{G_{\chi }\left(\exp \left[ \gamma _{\chi }\left( V-V_{\chi }^{1/2}\right) /\widetilde{k} _{\chi }\right] +\exp \left[ \left( \gamma _{\chi }-1\right) \left( V-V_{\chi }^{1/2}\right) /\widetilde{k}_{\chi }\right] \right)}, $$
(6)

In the following subsections we report only the kinetic details of the currents directly involved with the results discussed in this paper. The parameters of all the other currents were not modified from their original values.

1.1.1 Transient NaT current

The NaT current was implemented according to Shah et al. (2008), so Eq. (2) becomes:

$$I_{\rm NaT}= \overline{g}_{\rm NaT} \; m_{\rm NaT}^3 \;h_{\rm NaT}(V-50). $$
(7)

The dynamics of the gating variable m NaT is described by Eqs. (3), (4)1 and (5) as follows

$$ \begin{array}{rll} \alpha _{\rm m_{\rm NaT}} &=&\left\{ \begin{array}{ll} \dfrac{0.4\;(V-V_{\rm m_{\rm NaT}}^{1/2})}{1-\exp[-(V-V_{\rm m_{\rm NaT}}^{1/2})/k_{\rm m_{\rm NaT}}]},\\\\ |V-V_{\rm m_{\rm NaT}}^{1/2}|>10^{-6}\;\rm mV \\\\ 0.4 \;k_{\rm m_{\rm NaT}}, & \mathrm{otherwise} \end{array} \right. , \\ \beta _{\rm m_{\rm NaT}} &=&\left\{ \begin{array}{ll} \dfrac{-0.124\;(V-V_{\rm m_{\rm NaT}}^{1/2})}{1-\exp[(V-V_{\rm m_{\rm NaT}}^{1/2})/k_{\rm m_{\rm NaT}}]},\\\\ |V_{\rm m_{\rm NaT}}^{1/2}-V|>10^{-6}\;mV \\\\ 0.124\; k_{\rm m_{\rm NaT}}, & \mathrm{otherwise} \end{array} \right. , \\ \tau_{\rm m_{\rm NaT}} &=&\left\{ \begin{array}{ll} \dfrac{3}{\alpha _{\rm m_{\rm NaT} }(V)+\beta _{\rm m_{\rm NaT} }(V)},\\\\ \dfrac{3}{\alpha _{\rm m_{\rm NaT} }(V)+\beta _{\rm m_{\rm NaT} }(V)}> 0.02 \rm \;ms\\\\ 0.02 \rm \;ms, & \mathrm{otherwise} \end{array} \right. , \end{array} $$

The dynamics of the gating variable h NaT is described by Eqs. (3), (4)2, and (5), i.e.,

$$ \begin{array}{rll} (h_{\rm NaT})_{\infty}&=&\frac{1}{1+\exp[(V-V_{\rm h_{\rm NaT}}^{1/2})/k_{\rm h_{\rm NaT}}]}, \\[6pt] \alpha _{\rm h_{\rm NaT}} &=&\left\{ \begin{array}{ll} \dfrac{ 0.03 \;(V+45)}{1- \exp[-(V+45)/1.5]},\\\\ |V-V_{\rm h_{\rm NaT}}^{1/2}|>10^{-6}\rm \;mV \\\\ 0.045 , & \mathrm{otherwise} \end{array} \right. , \\ \beta _{\rm h_{\rm NaT}} &=&\left\{ \begin{array}{ll} \dfrac{- 0.01 \; (V+45)}{1- \exp[(V+45)/1.5]},\\\\ |V_{\rm h_{\rm NaT}}^{1/2}-V|>10^{-6} \rm \;mV \\\\ 0.015 , & \mathrm{otherwise} \end{array} \right. \end{array} $$

All the parameters written in this section are listed in Table 3.

1.1.2 The delayed rectifier K  +  current

The delayed rectifier K  +  current (\(I_{\rm K_{\rm DR}}\)) taken from Shah et al. (2008) is given by:

$$ I_{K_{DR}}= \overline{g}_{K_{DR}} \; m_{K_{DR}}\; (V+77) $$
(8)

The dynamics of \(m_{\rm K_{\rm DR}}\) is described by Eqs. (3), (4)2 and (6), with \(V_{\rm m_{\rm K_{\rm DR}}}^{1/2}=13\;\rm mV\), \(k_{\rm K_{\rm DR}}=-8.824\rm \;mV\), \(\tau_{\rm m_{\rm K_{\rm DR}}}^{0}=0\rm \;ms\), \(\gamma_{\rm m_{\rm K_{\rm DR}}}=0.7\), \(G_{\rm K_{\rm DR}}=1\), \(\bar{\tau}_{\rm m_{\rm K_{\rm DR}}}=\frac{1}{0.02}\rm \; ms\), \(\widetilde{k}_{\rm s_{\rm NaT}}=8.824\rm \;mV\), and a minimum value for \(\tau_{\rm m_{\rm K_{\rm DR}}}\) of 2 ms.

1.1.3 The potassium M-type current

The M current (\(I_{K_{M}}\)) was modeled according to Shah et al. (2008), namely

$$ I_{K_{M}}= \overline{g}_{K_{M}} \; m_{K_{M}}\; (V+77) $$
(9)

The dynamics of the gating variable \(m_{K_{M}}\) is described by Eqs. (3), (6) and (4)2 with the following formula:

$$ (m_{K_{M}})_{\infty}=\frac{1}{1+\exp[(V-V_{m_{K_{M}}}^{1/2})/(k_{m_{K_{M}}}]} $$
(10)

Some parameters are listed in Table 3 and the others are set as follows: \(\tau_{m_{K_{M}}}^{0}=60\rm \; ms\), \(G_{m_{K_{M}}}=1\).

1.1.4 The m-type Ca-dependent potassium current, mAHP

The medium AHP current I mAHP, from Moczydlowski and Latorre (1983), is given by

$$ I_{\rm mAHP}= \overline{g}_{\rm mAHP} \; m_{\rm mAHP}\; (V+77) $$
(11)

The dynamics of the gating variable m mAHP is described by Eqs. (3), (4)1 and (5). The α m and β m are given by:

$$ \begin{array}{rll}\alpha_{\rm mAHP}&=&\frac{0.48} {1+ \frac{0.18}{[Ca^{2+}]_{i}}\exp[-2 d_{1} V \cdot 37.775]}, \\ \beta_{\rm mAHP}&=&\frac{0.28}{1+\frac{[Ca^{2+}]_{i}}{ 0.011\;\exp[-2 d_{2} V \cdot 37.775]}} \end{array} $$
(12)

\([Ca^{2+}]_{i}\) is the internal calcium concentration, and the other parameters are listed in Table 3.

1.2 Mathematical description of the somatic model

We recall that the single-compartment model takes into account only the ten somatic ionic currents listed in Section “Ionic currents”, i.e.,

  • one transient Na  +  current I NaT;

  • three K  +  currents: one delayed rectifier \(I_{\rm K_{\rm DR}}\), one muscarinic-sensitive \(I_{K_{M}}\), and one A-type \(I_{K_{A}}\);

  • three Ca 2 +  currents: one LVA T-type current I CaT, one HVA R-type current I CaR, and one HVA L-type current I CaL;

  • one h current I h ;

  • two Ca 2 +  −activated K  +  currents: one slow I sAHP, and one medium I mAHP.

Then, the current balance equation for the somatic membrane potential V becomes

$$ \begin{aligned}[b] C_{m}\displaystyle \frac{dV}{dt}={}&-I_{\rm NaT}-I_{K_{\rm DR}}-I_{K_{M}}-I_{K_{A}}\\ &-I_{\rm CaT}-I_{\rm CaR}-I_{\rm CaL}-I_{h}\\ &-I_{\rm sAHP}-I_{\rm mAHP}-I_{\rm leak}+I_{\rm inj}. \end{aligned} $$
(13)

Thus, closing the membrane equation (13) with the ionic currents, the dynamics of gating variables, and the intracellular Ca 2 +  concentration, \(\left[ Ca^{2+}\right]_{i}\), the model consists of a system of seventeen nonlinear ODEs in the unknown functions

$$ \begin{array} [t]{c} V,m_{\rm Na_{\rm T}},h_{\rm Na_{\rm T}},s_{\rm Na_{\rm T}},m_{\rm K_{\rm DR}},m_{\rm K_{\rm M}},m_{\rm K_{\rm A}},h_{\rm K_{\rm A}},m_{\rm CaT},h_{\rm CaT}, \\ m_{\rm CaR},h_{\rm CaR},m_{\rm CaL},m_{h},m_{\rm sAHP},m_{\rm mAHP},\left[ Ca^{2+}\right]_{i}. \end{array} $$
(14)

The model is analytically intractable, and solutions may be obtained only by numerical integration. The Cauchy problem for our model is:

$$ \left\{ \begin{array}{l} \displaystyle\frac{d\mathbf{X}}{dt}=\mathbf{f}\left( \mathbf{X},I_{\rm ext}\right) , \\\\ \mathbf{X}\left( 0,I_{\rm ext}\right) =\mathbf{X}_{0}\left( I_{\rm ext}\right), \end{array} \right. $$
(15)

where the components of the vector X are listed in Eq. (14), and \(\mathbf{f}\left( \mathbf{X},I_{\rm ext}\right)\) is the vector whose components are the functions in the right-hand side of the ODEs. In addition, we set

$$ f_{1}\left(\mathbf{X},I_{\rm ext}\right)= I_{\rm ext}-G\left(\mathbf{X}\right). $$
(16)

In order to compare our dynamical model with experiments, we have to fix the values of all parameters and the external current. We set the potential V to the resting value V rest and all the gating variables and intracellular Ca 2 +  concentration to the corresponding steady-state values, i.e.,

$$ X_{j,0}=\left\{ \begin{array}{ll} V_{\rm rest} & \quad j=1, \\\\ X_{j}^{\infty }\left( V_{\rm rest}\right) & \quad j=2,\ldots ,14, \\\\ m_{\rm sAHP}^{\infty }\left( \left[ Ca^{2+}\right]_{i} ^{\infty}\right) & \quad j=15, \\\\ m_{\rm mAHP}^{\infty }\left(V_{\rm rest}, \left[ Ca^{2+}\right]_{i} ^{\infty}\right) & \quad j=16, \\\\ \left[ Ca^{2+}\right]_{i} ^{\infty } & \quad j=17, \end{array} \right. $$
(17)

where \(\left[ Ca^{2+}\right]_{i} ^{\infty }\) is a solution of the nonlinear equation

$$ f_{17}\left( V_{\rm rest},m_{\rm CaL}^{\infty },m_{\rm CaR}^{\infty },h_{\rm CaR}^{\infty },\left[ Ca^{2+} \right]_{i} \right) =0. $$
(18)

For the all simulations in Section 3.1.3 we have set the initial data as in Eq. (17) in which V rest = − 70 mV.

A fundamental step in the qualitative analysis of our dynamical system is finding the equilibria. To start with, we calculate and characterize the stationary states X * which are defined by the condition \(\mathbf{f}\left(\mathbf{X^*},I_{\rm ext}\right)=0\). As in the Hodgkin–Huxley-type models, a suitable numerical procedure was applied to reduce the vectorial equation \(\mathbf{f}\left(\mathbf{X^*},I_{\rm ext}\right)=0\) to the nonlinear scalar equation

$$ f_{1}\left(\mathbf{X^*},I_{\rm ext}\right)= I_{\rm ext}-g\left(V^{*}\right)=0, $$
(19)

where \(g\left(V\right)=G\left(V,X_{2}^{\infty }\left( V\right),...,X_{17}^{\infty }\left( V\right)\right)\). Owing to the presence of the calcium currents, the last three equations for the equilibria become

$$ \begin{aligned}[b] f_{15}\left(m_{\rm sAHP}^{*}, \left[ Ca^{2+}\right]_{i} ^{*}\right)&=0, \\ f_{16}\left(m_{\rm mAHP}^{*},V^{*}, \left[ Ca^{2+}\right]_{i} ^{*}\right)&=0, \\ f_{17}\left( V^{*},m_{\rm CaL}^{* },m_{\rm CaR}^{*},h_{\rm CaR}^{*},\left[ Ca^{2+} \right]_{i}^{*} \right) &=0, \end{aligned} $$
(20)

consequently a nonstandard numerical procedure was applied to derive Eq. (19). This last equation, which supplies the stationary point X * as a function of I ext, can be solved only numerically. Linearizing the \(\mathbf{f}\left(\mathbf{X^*},I_{\rm ext}\right)\) around the stationary states, we investigated their stability properties by varying the external current I ext. The stability of the equilibria is characterized by the eigenvalues λ 1, ..., λ 17 of the usual Jacobian matrix

$$ J\left( I_{\rm ext}\right) =\frac{\partial f_{j}}{\partial X_{k}}\left( \mathbf{ X}^{\ast},I_{\rm ext}\right), $$
(21)

and all bifurcations diagrams are obtained using the procedure originally introduced by Troy for the Hodgkin–Huxley model in Troy (1974) (see also Hassard 1978; Troy 1978).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Bianchi, D., Marasco, A., Limongiello, A. et al. On the mechanisms underlying the depolarization block in the spiking dynamics of CA1 pyramidal neurons. J Comput Neurosci 33, 207–225 (2012). https://doi.org/10.1007/s10827-012-0383-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-012-0383-y

Keywords

Navigation