Skip to main content
Log in

The effect of dendritic voltage-gated conductances on the neuronal impedance: a quantitative model

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

Abstract

Neuronal impedance characterizes the magnitude and timing of the subthreshold response of a neuron to oscillatory input at a given frequency. It is known to be influenced by both the morphology of the neuron and the presence of voltage-gated conductances in the cell membrane. Most existing theoretical accounts of neuronal impedance considered the effects of voltage-gated conductances but neglected the spatial extent of the cell, while others examined spatially extended dendrites with a passive or spatially uniform quasi-active membrane. We derived an explicit mathematical expression for the somatic input impedance of a model neuron consisting of a somatic compartment coupled to an infinite dendritic cable which contained voltage-gated conductances, in the more general case of non-uniform dendritic membrane potential. The validity and generality of this model was verified through computer simulations of various model neurons. The analytical model was then applied to the analysis of experimental data from real CA1 pyramidal neurons. The model confirmed that the biophysical properties and predominantly dendritic localization of the hyperpolarization-activated cation current I h were important determinants of the impedance profile, but also predicted a significant contribution from a depolarization-activated fast inward current. Our calculations also implicated the interaction of I h with amplifying currents as the main factor governing the shape of the impedance-frequency profile in two types of hippocampal interneuron. Our results provide not only a theoretical advance in our understanding of the frequency-dependent behavior of nerve cells, but also a practical tool for the identification of candidate mechanisms that determine neuronal response properties.

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

References

  • Angelo, K., London, M., Christensen, S. R., & Häusser, M. (2007). Local and global effects of I(h) distribution in dendrites of mammalian neurons. Journal of Neuroscience, 27, 8643–8653.

    Article  PubMed  CAS  Google Scholar 

  • Biel, M., Wahl-Schott, C., Michalakis, S., & Zong, X. (2009). Hyperpolarization-activated cation channels: From genes to function. Physiological Reviews, 89, 847–885.

    Article  PubMed  CAS  Google Scholar 

  • Bishop, C. M. (2006). Pattern recognition and machine learning. Singapore: Springer.

    Google Scholar 

  • Borg-Graham, L. (1999). Interpretations of data and mechanisms for hippocampal pyramidal cell models. In P. S. Ulinski, E. G. Jones, & A. Peters (Eds.), Cortical models ed. (pp. 19–138). New York: Plenum Press.

    Google Scholar 

  • Bower, J. M., & Beeman, D. (1998). The book of GENESIS: Exploring realistic neural models with the GEneral NEural SImulation System (2nd ed.). New York, Inc.: Springer.

    Google Scholar 

  • Bressloff, P. C. (1999). Resonantlike synchronization and bursting in a model of pulse-coupled neurons with active dendrites. Journal of Computational Neuroscience, 6, 237–249.

    Article  PubMed  CAS  Google Scholar 

  • Cole, K. S. (1949). Some physical aspects of bioelectric phenomena. Proceedings of the National Academy of Sciences of the United States of America, 35, 558–566.

    Article  PubMed  CAS  Google Scholar 

  • Coombes, S., Timofeeva, Y., Svensson, C.-M., et al. (2007). Branching dendrites with resonant membrane: A “sum-over-trips” approach. Biological Cybernetics, 97, 137–149.

    Article  PubMed  CAS  Google Scholar 

  • Engel, T. A., Schimansky-Geier, L., Herz, A. V. M., et al. (2008). Subthreshold membrane-potential resonances shape spike-train patterns in the entorhinal cortex. Journal of Neurophysiology, 100, 1576–1589.

    Article  PubMed  CAS  Google Scholar 

  • Erchova, I., Kreck, G., Heinemann, U., & Herz, A. V. M. (2004). Dynamics of rat entorhinal cortex layer II and III cells: Characteristics of membrane potential resonance at rest predict oscillation properties near threshold. Journal of Physiology (London), 560, 89–110.

    Article  CAS  Google Scholar 

  • Golding, N. L., Mickus, T. J., Katz, Y., et al. (2005). Factors mediating powerful voltage attenuation along CA1 pyramidal neuron dendrites. Journal of Physiology (London), 568, 69–82.

    Article  CAS  Google Scholar 

  • Gutfreund, Y., Yarom, Y., & Segev, I. (1995). Subthreshold oscillations and resonant frequency in guinea-pig cortical neurons: Physiology and modelling. Journal of Physiology (London), 483(Pt 3), 621–640.

    CAS  Google Scholar 

  • Hines, M. (1984). Efficient computation of branched nerve equations. International Journal of Bio-Medical Computing, 15, 69–76.

    Article  PubMed  CAS  Google Scholar 

  • Hu, H., Vervaeke, K., Graham, L. J., & Storm, J. F. (2009). Complementary theta resonance filtering by two spatially segregated mechanisms in CA1 hippocampal pyramidal neurons. Journal of Neuroscience, 29, 14472–14483.

    Article  PubMed  CAS  Google Scholar 

  • Hu, H., Vervaeke, K., & Storm, J. F. (2002). Two forms of electrical resonance at theta frequencies, generated by M-current, h-current and persistent Na+ current in rat hippocampal pyramidal cells. Journal of Physiology (London), 545, 783–805.

    Article  CAS  Google Scholar 

  • Hutcheon, B., Miura, R. M., & Puil, E. (1996a). Models of subthreshold membrane resonance in neocortical neurons. Journal of Neurophysiology, 76, 698–714.

    PubMed  CAS  Google Scholar 

  • Hutcheon, B., Miura, R. M., & Puil, E. (1996b). Subthreshold membrane resonance in neocortical neurons. Journal of Neurophysiology, 76, 683–697.

    PubMed  CAS  Google Scholar 

  • Hutcheon, B., Miura, R. M., Yarom, Y., & Puil, E. (1994). Low-threshold calcium current and resonance in thalamic neurons: A model of frequency preference. Journal of Neurophysiology, 71, 583–594.

    PubMed  CAS  Google Scholar 

  • Hutcheon, B., & Yarom, Y. (2000). Resonance, oscillation and the intrinsic frequency preferences of neurons. Trends in Neurosciences, 23, 216–222.

    Article  PubMed  CAS  Google Scholar 

  • Jack, J. B., Noble, D., & Tsien, R. W. (1983). Electric current flow in excitable cells. USA: Oxford University Press.

    Google Scholar 

  • Johnston, D., Christie, B. R., Frick, A., et al. (2003). Active dendrites, potassium channels and synaptic plasticity. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 358, 667–674.

    Article  PubMed  CAS  Google Scholar 

  • Johnston, D., & Narayanan, R. (2008). Active dendrites: Colorful wings of the mysterious butterflies. Trends in Neurosciences, 31, 309–316.

    Article  PubMed  CAS  Google Scholar 

  • Káli, S., & Freund, T. F. (2005). Distinct properties of two major excitatory inputs to hippocampal pyramidal cells: A computational study. European Journal of Neuroscience, 22, 2027–2048.

    Article  PubMed  Google Scholar 

  • Koch, C. (1984). Cable theory in neurons with active, linearized membranes. Biological Cybernetics, 50, 15–33.

    Article  PubMed  CAS  Google Scholar 

  • Koch, C. (1998). Biophysics of computation: Information processing in single neurons. USA: Oxford University Press.

    Google Scholar 

  • Koch, C., & Poggio, T. (1985). A simple algorithm for solving the cable equation in dendritic trees of arbitrary geometry. Journal of Neuroscience Methods, 12, 303–315.

    Article  PubMed  CAS  Google Scholar 

  • Leung, L. S., & Yu, H. W. (1998). Theta-frequency resonance in hippocampal CA1 neurons in vitro demonstrated by sinusoidal current injection. Journal of Neurophysiology, 79, 1592–1596.

    PubMed  CAS  Google Scholar 

  • Llinás, R. R. (1988). The intrinsic electrophysiological properties of mammalian neurons: Insights into central nervous system function. Science, 242, 1654–1664.

    Article  PubMed  Google Scholar 

  • Lörincz, A., Notomi, T., Tamás, G., et al. (2002). Polarized and compartment-dependent distribution of HCN1 in pyramidal cell dendrites. Nature Neuroscience, 5, 1185–1193.

    Article  PubMed  Google Scholar 

  • Magee, J. C. (1998). Dendritic hyperpolarization-activated currents modify the integrative properties of hippocampal CA1 pyramidal neurons. Journal of Neuroscience, 18, 7613–7624.

    PubMed  CAS  Google Scholar 

  • Magee, J., Hoffman, D., Colbert, C., & Johnston, D. (1998). Electrical and calcium signaling in dendrites of hippocampal pyramidal neurons. Annual Review of Physiology, 60, 327–346.

    Article  PubMed  CAS  Google Scholar 

  • Mauro, A. (1961). Anomalous impedance, a phenomenological property of time-variant resistance. An analytic review. Biophysical Journal, 1, 353–372.

    Article  PubMed  CAS  Google Scholar 

  • Mauro, A., Conti, F., Dodge, F., & Schor, R. (1970). Subthreshold behavior and phenomenological impedance of the squid giant axon. Journal of General Physiology, 55, 497–523.

    Article  PubMed  CAS  Google Scholar 

  • Megías, 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  Google Scholar 

  • Narayanan, R., & Johnston, D. (2007). Long-term potentiation in rat hippocampal neurons is accompanied by spatially widespread changes in intrinsic oscillatory dynamics and excitability. Neuron, 56, 1061–1075.

    Article  PubMed  CAS  Google Scholar 

  • Narayanan, R., & Johnston, D. (2008). The h channel mediates location dependence and plasticity of intrinsic phase response in rat hippocampal neurons. Journal of Neuroscience, 28, 5846–5860.

    Article  PubMed  CAS  Google Scholar 

  • Pike, F. G., Goddard, R. S., Suckling, J. M., et al. (2000). Distinct frequency preferences of different types of rat hippocampal neurones in response to oscillatory input currents. Journal of Physiology (London), 529(Pt 1), 205–213.

    Article  CAS  Google Scholar 

  • Rall, W. (1959). Branching dendritic trees and motoneuron membrane resistivity. Experimental Neurology, 1, 491–527.

    Article  PubMed  CAS  Google Scholar 

  • Rall, W. (1962). Theory of physiological properties of dendrites. Annals of the New York Academy of Sciences, 96, 1071–1092.

    Article  PubMed  CAS  Google Scholar 

  • Rall, W., & Agmon-Snir, H. (1998). Cable theory for dendritic neurons. In C. Koch, & I. Segev (Eds.), Methods in neuronal modeling: From ions to networks (2 ed., pp. 27–92). Cambridge, MA: MIT Press.

    Google Scholar 

  • Reyes, A. (2001). Influence of dendritic conductances on the input-output properties of neurons. Annual Review of Neuroscience, 24, 653–675.

    Article  PubMed  CAS  Google Scholar 

  • Richardson, M. J. E., Brunel, N., & Hakim, V. (2003). From subthreshold to firing-rate resonance. Journal of Neurophysiology, 89, 2538–2554.

    Article  PubMed  Google Scholar 

  • Sjöström, P. J., Rancz, E. A., Roth, A., & Häusser, M. (2008). Dendritic excitability and synaptic plasticity. Physiological Reviews, 88, 769–840.

    Article  PubMed  Google Scholar 

  • Ströhmann, B., Schwarz, D. W., & Puil, E. (1994). Subthreshold frequency selectivity in avian auditory thalamus. Journal of Neurophysiology, 71, 1361–1372.

    PubMed  Google Scholar 

  • Stuart, G., & Spruston, N. (1998). Determinants of voltage attenuation in neocortical pyramidal neuron dendrites. Journal of Neuroscience, 18, 3501–3510.

    PubMed  CAS  Google Scholar 

  • Ulrich, D. (2002). Dendritic resonance in rat neocortical pyramidal cells. Journal of Neurophysiology, 87, 2753–2759.

    PubMed  Google Scholar 

  • Zemankovics, R., Káli, S., Paulsen, O., et al. (2010). Differences in subthreshold resonance of hippocampal pyramidal cells and interneurons: The role of h-current and passive membrane characteristics. Journal of Physiology (London), 588, 2109–2132.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We are grateful to Dr Norbert Hájos for initiating the project and for helpful discussions and comments regarding the manuscript. We thank Katalin Lengyel and Erzsébet Gregori for their excellent technical assistance. This work was supported by the Hungarian Scientific Research Fund (OTKA T049517, OTKA K60927, OTKA K83251).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Szabolcs Káli.

Additional information

Action Editor: Alain Destexhe

Appendices

Appendix A: Derivation of the impedance of a single-compartment model with multiple voltage-gated conductances

We start from Eqs. (6) and (7), and make the approximations \(m_{\infty }^{(j)}(V_{m})\approx m_{\infty }^{(j)}(V_{0})+b_{j}(V_{m}-V_{0})\) and \(\tau _{m}^{(j)}(V_{m})\approx \tau _{j}\). If we use these approximations in Eq. (7), we get

$$ \frac{\mathit{dm}_{j}}{\mathit{dt}}=\frac{m_{\infty }^{(j)}\big(V_{0}\big)+b_{j}\big(V_{m}-V_{0}\big)-m_{j}}{\tau _{j}} $$
(28)

In order to compute the impedance represented by the voltage-gated current I j , we introduce a formalism based on complex numbers: oscillatory variable X will be characterized by the complex number \(\bar{{X}}=A_{X}e^{i\phi _{X}}\) where A X is the amplitude and ϕ X is the phase (at time 0) of the oscillation of X; the actual value of X at time t is then given by \(X(t)=\Re (\bar{{X}}e^{i\omega t})=\Re (A_{X}e^{i(\omega t+\phi _{X})})=A_{X}\cos (\omega t+\phi _{X})\), where ω = 2πf, f is the oscillation frequency, and \(\Re (z)\) is the real part of complex number z. We will assume that our single-compartment model undergoes a small-amplitude sinusoidal voltage fluctuation given by \(V_{m}=V_{0}+A_{V}e^{i\omega t}\). If the dynamics of the current is approximately linear for small-amplitude voltage fluctuations (which holds in almost all cases), then the resulting current will also be sinusoidal at the same frequency, \(I_{j}=I_{j}^{0}+A_{j}e^{i(\omega t+\phi _{j})}\), and the complex impedance represented by this current is \(Z_{j}=\frac{V_{m}-V_{0}}{I_{j}-I_{j}^{0}}=\frac{A_{V}}{A_{j}}e^{-i\phi _{j}}\). As a first step in the calculation of Z j , we compute the response of gating variable m j to the voltage fluctuation, \(m_{j}=m_{j}^{0}+A_{m}e^{i(\omega t+\phi _{m})}\)(we have dropped the index ’j’ from A m and ϕ m to simplify notation). Substituting the appropriate complex expressions into Eq. (28), and noting that \(\frac{\mathit{dm}_{j}}{\mathit{dt}}=A_{m}i\omega e^{i(\omega t+\phi _{m})}\), we get \(\tau_{j}A_{m}i\omega e^{i(\omega t+\phi _{m})}=m_{\infty }^{(j)}(V_{0})+b_{j}A_{V}e^{i\omega t}-m_{j}^{0}-A_{m}e^{i(\omega t+\phi _{m})}\), from which, since the equation must hold for all t, we can infer both \(m_{j}^{0}=m_{\infty }^{(j)}(V_{0})\) and, more importantly, \(i\omega \tau _{j}A_{m}e^{i\phi _{m}}=b_{j}A_{V}-A_{m}e^{i\phi_{m}}\) or, rearranging terms, \(A_{m}e^{i\phi _{m}}=\frac{b_{j}A_{V}}{1+i\omega \tau _{j}}\). We are now in a position to calculate I j :

$$ \begin{array}{rll} I_{j}&=&\bar{{g}}_{j}m_{j}\big(V_{m}-E_{j}\big)\\ &=&\bar{{g}}_{j}\big(m_{j}^{0}+A_{m}e^{i\big(\omega t+\phi _{m}\big)}\big)\big(V_{0}+A_{V}e^{i\omega t}-E_{j}\big)\\ &\approx & \bar{{g}}_{j}m_{j}^{0}\big(V_{0}-E_{j}\big)+\bar{{g}}_{j}\big(V_{0}-E_{j}\big)A_{m}e^{i\big(\omega t+\phi _{m}\big)}\\ &&+\,\bar{{g}}_{j}m_{j}^{0}A_{V}e^{i\omega t}\\ &=&\bar{{g}}_{j}m_{j}^{0}\big(V_{0}-E_{j}\big)+\bar{{g}}_{j}\big(V_{0}-E_{j}\big)\frac{b_{j}A_{V}}{1+i\omega \tau _{j}}e^{i\omega t}\\ &&+\,\bar{{g}}_{j}m_{j}^{0}A_{V}e^{i\omega t} \end{array} $$
(29)

where, in the second step, we neglected the second-order term proportional to A V A m , and, in the last step, we made use of the expression we derived for \(A_{m}e^{i\phi _{m}}\) above. Since we also have \(I_{j}=I_{j}^{0}+A_{j}e^{i(\omega t+\phi _{j})}\), we see comparing terms (and defining \(g_{j}^{0}=\bar{{g}}_{j}m_{j}^{0}\)) that \(I_{j}^{0}=g_{j}^{0}(V_{0}-E_{j})\) and \(A_{j}e^{i\phi _{j}}=A_{V}\left[g_{j}^{0}+\frac{\bar{{g}}_{j}b_{j}(V_{0}-E_{j})}{1+i\omega \tau _{j}}\right]\). It follows that

$$ Z_{j}=\frac{A_{V}}{A_{j}}e^{-i\phi _{j}}=\frac{1}{g_{j}^{0}+\frac{\bar{{g}}_{j}b_{j}(V_{0}-E_{j})}{1+i\omega \tau _{j}}}. $$
(30)

Appendix B: Derivation of the impedance of a neuronal model with a semi-infinite active dendrite

1.1 B.1 Steady-state membrane potential distribution

In order to determine the steady-state voltage distribution in the dendrite, we start from Eqs. (11) and (12). Approximating the steady-state value of the gating variable as a linear expansion around V  ∞ , as embodied by Eq. (14), we replace Eq. (12) by

$${\frac{\partial m}{\partial t}=\frac{m_{{\infty }}\left(V_{{\infty }}\right)+b\big(V_{m}-V_{\infty }\big)-m}{\tau _{{m}}\left(V_{{m}}\right)}} $$
(31)

The steady-state solution of this equation is m = m  ∞ (V  ∞ ) + b(V m  − V  ∞ ). We substitute this expression into Eq. 11, and set its left-hand-side to 0, obtaining the steady-state equation

$$ \begin{array}{rll} \frac{d}{4R_{A}}\frac{d^{2}V_{0}}{\mathit{dx}^{2}}&=&g_{l}\big(V_{0}-E_{l}\big)\\&&+\,\bar{{g}}_{d} \big(V_{0}-E_{d}\big)\big[m_{\infty }\big(V_{\infty }\big)+b\big(V_{0}-V_{\infty }\big)\big] \end{array} $$
(32)

where V 0(x) is the steady-state voltage distribution along the dendrite. Making use of Eq. (13), which defined V  ∞ , Eq. (32) can be transformed into

$$ \begin{array}{rll} \frac{d}{4R_{A}}\frac{d^{2}V_{0}}{\mathit{dx}^{2}}&=&g_{l}\big(V_{0}-V_{\infty }\big)+\bar{{g}}_{d}\big(V_{0}-V_{\infty }\big)m_{\infty }\big(V_{\infty}\big)\\ &&+\bar{{g}}_{d}\big(V_{\infty }-E_{d}\big)b\big(V_{0}-V_{\infty}\big)\\ &&+\bar{{g}}_{d}\big(V_{0}-V_{\infty }\big)b\big(V_{0}-V_{\infty }\big) \end{array} $$
(33)

Of the four terms on the right-hand-side of Eq. (33), the first three depend linearly on V 0 − V  ∞ , while the fourth term depends on it quadratically. We will neglect this quadratic term in the following calculations based on our earlier assumption that V 0 − V  ∞  is relatively small; in the current context, it may be justified by noting that the fourth term in Eq. (33) relates to the second term as b(V 0 − V  ∞ ) relates to m  ∞ (V  ∞ ), a ratio which is small if V 0 − V  ∞  is small. In fact, we should neglect the fourth term also for the sake of the consistency of our approach since we have already neglected factors that would otherwise contribute second-order terms to Eq. (33) when we linearized the expression for m  ∞ (V  ∞ ).

Now let us define V′(x) = V 0(x) − V  ∞ , and substitute it into Eq. (33). We get

$$ \frac{d}{4R_{A}}\frac{d^{2}V'}{\mathit{dx}^{2}}=\big[g_{l}+\bar{{g}}_{d}m_{\infty }(V_{\infty })+\bar{{g}}_{d}b\big(V_{\infty }-E_{d}\big)\big]V' $$
(34)

a second-order linear differential equation, and the corresponding boundary conditions become \(\left(\frac{\mathit{dV}'}{\mathit{dx}}\right)_{x=\infty }=0\) and V′(0) = V 0(0) − V  ∞ , where V 0(0) is the holding potential at the soma. The solution of this equation is \(V'(x)=(V_{0}(0)-V_{\infty })e^{-x/\lambda }\), with λ given by Eq. (16).

Finally, returning to the original variables, we obtain the solution represented by Eq. (15).

1.2 B.2 Input impedance

In order to obtain the time-dependent solution, we start from Eqs. (11) and (12). Let us define V(x, t) = V m (x, t) − V 0(x) and m′(x, t) = m(x, t) − m  ∞ (V 0(x)). Plugging these expressions into Eq. (11) and noting that, following from the definition of V 0(x), \(\frac{d}{4R_{A}}\frac{d^{2}V_{0}}{\mathit{dx}^{2}}+g_{l}(E_{l}-V_{0})+m_{\infty }(V_{0})(E_{d}-V_{0})=0\), we get

$$ \begin{aligned}[b] C\frac{\partial V}{\partial t}={}&\frac{d}{4R_{A}}\frac{\partial ^{2}V}{\partial x^{2}}-g_{l}V-\bar{{g}}_{d}\big(V_{0}-E_{d}\big)m' \\ &-\bar{{g}}_{d}Vm_{\infty}\big(V_{0}\big)-\bar{{g}}_{d}Vm' \end{aligned} $$
(35)

Once again, we neglect the second-order last term containing Vm′ (as it should be much smaller in magnitude than either of the preceding two terms). The last significant approximation we make is that we assume that the time constant of the voltage-gated conductance can be considered uniform across the relevant membrane potential range, i.e., \({\tau _{{m}}\left(V_{{m}}\right)=\tau _{d}}\). With these simplifications and the approximation of Eq. (14), we arrive at (using the new variables)

$$ \begin{aligned}[b] C\frac{\partial V}{\partial t}={}&\frac{d}{4R_{A}}\frac{\partial ^{2}V}{\partial x^{2}}-g_{l}V\\ &-\bar{{g}}_{d}\big(V_{0}-E_{d}\big)m'-\bar{{g}}_{d}Vm_{\infty }\big(V_{0}\big) \end{aligned} $$
(36)

and

$$ \frac{\partial m'}{\partial t}=\frac{\mathit{bV}-m'}{\tau _{d}} $$
(37)

We will use Eqs. (36) and (37) to derive the impedance of this quasi-active semi-infinite dendrite.

Note that Eqs. (36) and (37) define a linear system of differential equations; therefore, in response to periodic sinusoidal perturbation (such as a sinusoidal current injection or sinusoidal variation in the membrane potential at the soma), all state variables (both V and m′) at all locations vary in a sinusoidal manner at the frequency of the input. Thus, for a given input frequency, the solution is completely characterized by providing the amplitude and phase of the oscillation of the state variables (relative to the amplitude and phase of the input) at each point along the dendrite. The calculations are again more straightforward if we employ the formalism based on complex numbers that we introduced in Appendix A. If we now substitute \(V=\bar{{V}}e^{i\omega t}\) and \(m'=\bar{{m}}e^{i\omega t}\) (where \(\bar{{V}}(x)\) and \(\bar{{m}}(x)\) are complex numbers which depend on x) into Eqs. (36) and (37), and note that the temporal differentiation on the left-hand-side is now straightforward, we get the following system of ordinary differential and algebraic equations:

$$ \begin{aligned}[b] i\omega C\bar{{V}}={}&\frac{d}{4R_{A}}\frac{d^{2}\bar{{V}}}{\mathit{dx}^{2}}-\big[g_{l}+\bar{{g}}_{d}m_{\infty }(V_{0})\big]\bar{{V}}\\ &-\bar{{g}}_{d}\big(V_{0}-E_{d}\big)\bar{{m}} \end{aligned} $$
(38)

and

$$ i\omega \tau _{d}\bar{{m}}=b\bar{{V}}-\bar{{m}} $$
(39)

From Eq. (39), we obtain \(\bar{{m}}=\frac{b}{1+i\omega \tau _{d}}\bar{{V}}\), which can be substituted into Eq. (38). We arrive at the second-order ordinary differential equation

$$ \begin{array}{rll} \frac{d}{4R_{A}}\frac{d^{2}\bar{{V}}}{\mathit{dx}^{2}}&=&i\omega C\bar{{V}}+\big[g_{l}+\bar{{g}}_{d}m_{\infty }(V_{0})\big]\bar{{V}}\\&& +\,\frac{\bar{{g}}_{d}b\big(V_{0}-E_{d}\big)}{1+i\omega \tau _{d}}\bar{{V}} \end{array} $$
(40)

The only remaining complexity hidden in the notation of Eq. (40) is that V 0 is actually a function of x. However, we have already calculated V 0(x), which is given by Eq. (15). In order to obtain \({m_{\infty }\left(V_{{0}}\right)}\), which is also a function of x, we combine Eq. (14) with Eq. (15), getting \(m_{\infty }(V_{0}(x))\approx m_{\infty }(V_{\infty })+b(V_{0}(x)-V_{\infty })=m_{\infty }(V_{\infty })+b(V_{0}(0)-V_{\infty })e^{-x/\lambda }\). If we now insert these expressions into Eq. (40) and rearrange terms, we get

$$ \begin{array}{rll} \frac{d}{4R_{A}}\frac{d^{2}\bar{{V}}}{\mathit{dx}^{2}}\!&=&\!\left\{g_{l}+i\omega C+\bar{{g}}_{d}m_{\infty }(V_{\infty }) \!+\!\frac{\bar{{g}}_{d}b(V_{\infty }-E_{d})}{1+i\omega \tau _{d}}\right.\\ &&+\left.\bar{{g}}_{d}b(V_{0}(0)\!-\!V_{\infty })\left[1\!+\!\frac{1}{1+i\omega \tau _{d}}\right]e^{-x/\lambda }\right\}\!\bar{{V}} \end{array} $$
(41)

We solved this second-order ordinary differential equation for \(\bar{{V}}(x)\), taking into account the boundary conditions \(\left(\frac{d\bar{{V}}}{\mathit{dx}}\right)_{x=\infty }=0\) and \(\bar{{V}}(0)=A_{0}\) (where A 0 is the amplitude of the somatic voltage fluctuation, which is assumed to be real for simplicity), using the MATLAB Symbolic Math Toolbox (version 3.2.2). Following algebraic manipulations involving Bessel function identities, we arrived at a numerically stable form of the solution:

$$ \bar{V}(x)=A_{0}\cdot {\frac{I_{2\lambda \sqrt{A}}\left(2\lambda \sqrt{B}e^{-x/2\lambda }\right)}{I_{2\lambda \sqrt{A}}\left(2\lambda \sqrt{B}\right)}} $$
(42)

where I ν (z) is the modified Bessel function (of order ν) of the first kind, and λ, A, and B are given by Eqs. (16), (18), and (19), respectively. The axial current at x = 0 (which is equal to the current flowing between the dendrite and the soma) can be computed as

$$ \begin{array}{rll} I&=&-{\frac{d^{2}\pi}{4R_{A}}}\left(\frac{d\bar{{V}}}{\mathit{dx}}\right)_{x=0} \\ &=&-{\frac{d^{2}\pi}{4R_{A}}}A_{0}\left[\sqrt{A}-\sqrt{B}\frac{I_{2\lambda \sqrt{A}-1}\left(2\lambda \sqrt{B}\right)}{I_{2\lambda \sqrt{A}}\left(2\lambda\sqrt{B}\right)}\right] \end{array} $$
(43)

The (complex) impedance of the dendrite as seen from the soma is then given by

$$ Z_{d}=\frac{\bar{V}(0)}{I}=-{\frac{4R_{A}}{d^{2}\pi }}A_{0}\left[\left(\frac{d\bar{{V}}}{\mathit{dx}}\right)_{x=0}\right]^{-1} $$
(44)

which, in this case, results in the expression shown in Eq. (17).

The soma-dendritic transfer impedance can be calculated as \(K^{(0)}_{\mathit{sx}}=\bar{V}(x)/I\) in the case of infinite somatic impedance (when all of the current injected at location 0 flows into the dendrite), resulting in the expression given by Eq. (24) in the main text. However, this result is straightforward to generalize to an arbitrary somatic impedance. We first note that the current I flowing into the dendrite from the soma is related to the external somatic current injection I e as \(I = \frac{Z_s}{Z_s+Z_d} I_e\), where Z s is the total somatic impedance, given by Eq. (9) divided by the somatic membrane area, and Z d is given by Eq. (17). In this general case, the transfer impedance can be calculated as

$$ K_{sx} = \bar{V}(x)/I_e = \frac{Z_s}{Z_s+Z_d} \bar{V}(x)/I = \frac{Z_s}{Z_s+Z_d} K^{(0)}_{\mathit{sx}} $$
(45)

Rights and permissions

Reprints and permissions

About this article

Cite this article

Káli, S., Zemankovics, R. The effect of dendritic voltage-gated conductances on the neuronal impedance: a quantitative model. J Comput Neurosci 33, 257–284 (2012). https://doi.org/10.1007/s10827-012-0385-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10827-012-0385-9

Keywords

Navigation