Skip to main content
Erschienen in: The Cerebellum 3/2012

Open Access 01.09.2012

Controlling Ca2+-Activated K+ Channels with Models of Ca2+ Buffering in Purkinje Cells

verfasst von: Haroon Anwar, Sungho Hong, Erik De Schutter

Erschienen in: The Cerebellum | Ausgabe 3/2012

Abstract

Intracellular Ca2+ concentrations play a crucial role in the physiological interaction between Ca2+ channels and Ca2+-activated K+ channels. The commonly used model, a Ca2+ pool with a short relaxation time, fails to simulate interactions occurring at multiple time scales. On the other hand, detailed computational models including various Ca2+ buffers and pumps can result in large computational cost due to radial diffusion in large compartments, which may be undesirable when simulating morphologically detailed Purkinje cell models. We present a method using a compensating mechanism to replace radial diffusion and compared the dynamics of different Ca2+ buffering models during generation of a dendritic Ca2+ spike in a single compartment model of a PC dendritic segment with Ca2+ channels of P- and T-type and Ca2+-activated K+ channels of BK- and SK-type. The Ca2+ dynamics models used are (1) a single Ca2+ pool; (2) two Ca2+ pools, respectively, for the fast and slow transients; (3) detailed Ca2+ dynamics with buffers, pump, and diffusion; and (4) detailed Ca2+ dynamics with buffers, pump, and diffusion compensation. Our results show that detailed Ca2+ dynamics models have significantly better control over Ca2+-activated K+ channels and lead to physiologically more realistic simulations of Ca2+ spikes and bursting. Furthermore, the compensating mechanism largely eliminates the effect of removing diffusion from the model on Ca2+ dynamics over multiple time scales.
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1007/​s12311-010-0224-3) contains supplementary material, which is available to authorized users.
Model scripts All the scripts used to run simulations in this manuscript are available at ModelDB (http://​senselab.​med.​yale.​edu/​ModelDB).

Introduction

Purkinje cells (PCs) are known to express large conductance (BK) [110] and small conductance (SK) [6, 1114] Ca2+-activated K+ (KCa) channels on their dendrites. KCa channels together with voltage-gated Ca2+ channels significantly control the dendritic excitability [6, 15]. Ca2+ entering through the voltage-gated Ca2+ channels causes the free cytosolic Ca2+ concentration to rise, which in turn controls the activation of KCa channels. A significant difference between those channels is their spatial separation. It has been shown that BK channels are in closer vicinity of Ca2+ sources compared to SK channels, suggesting that BK channels require a brief large amount (~10–100 μM) of Ca2+, while SK channels are activated by lower concentrations (~0–2 μM) [1618]. Intracellular Ca2+ mechanisms like diffusion, endogenous buffers, internal stores, and pumps significantly control the cytosolic spread of Ca2+ and shape the Ca2+ signal that activates KCa channels [19]. Therefore, it is important to model intracellular Ca2+ dynamics carefully in a biophysically detailed PC model.
Existing PC models use a Ca2+ pool with a single relaxation time [2022], which can reasonably approximate the fast transient [23] and can activate a BK channel well [17, 18]. On the other hand, it provides inappropriate Ca2+ signals to activate SK channels to support their role in excitability modulation. A simple and effective solution is to use two Ca2+ pools [24], respectively, for the fast and slow transient. A more comprehensive and biophysically realistic solution is to use a detailed model [25] with Ca2+ buffers, diffusion of those buffers, and diffusion of Ca2+ and Ca2+ extrusion pumps. Large compartments also require modeling of radial diffusion of Ca2+ ions and buffers [26]. The presence of diffusion in these detailed models can result in a large computational cost [27], which may need to be avoided in simulating morphologically detailed PC models.
In this study, we compare the effectiveness of different Ca2+ buffering models in controlling Ca2+-activated K+ channels. We build a single compartment model of a PC dendritic segment, including P-type and T-type voltage-gated Ca2+ channels and BK-type and SK-type Ca2+-activated K+ channels, and tune this to generate Ca2+ spikes. Intracellular Ca2+ dynamics are modeled using a single Ca2+ pool, two Ca2+ pools, or detailed Ca2+ dynamics with calbindin (CB), parvalbumin (PV), a pump, and diffusion. Our results show that a detailed Ca2+ dynamics model with buffers, pumps, and diffusion has better control over Ca2+-activated K+ channels and can generate physiologically more realistic Ca2+ spikes. We also introduce a method to specify a diffusion compensating mechanism (DCM) that can replace the radial diffusion in the model. Our result shows that the consequences of removing diffusion from the model on the simulated Ca2+ dynamics can be largely eliminated by these compensating mechanisms for both short and long time scales. This allows the generation of physiologically realistic Ca2+ spikes that can be simulated with significantly less computational cost.

Materials and Methods

PC Model for Ca2+ Spikes

A dendritic compartment with a diameter of 4 μm and a length of 20 μm was used as a PC dendrite model for simulating dendritic Ca2+ spikes. All the simulations were run in the NEURON simulation environment [28]. The details about the currents used in our model are given below.

P-Type Ca2+ Channel

The P-type calcium current was based on data from Swensen and Bean [29] and included three activation (m) gates. The current density was based on the Goldman–Hodgkin–Katz equation [30]. The equations describing its kinetics are summarized below:
$$ \begin{array}{*{20}{c}} {{I_{{{\rm{C}}{{\rm{a}}^{{2 + }}}}}} = \overline {{P_{{{\rm{C}}{{\rm{a}}^{{2 + }}}}}}} \times {m^3} \times {g_{\rm{GHK}}}} \hfill \\{{m_{\infty }} = \frac{1}{{1 + {e^{{ - \frac{{V + 24.758}}{{8.429}}}}}}}} \hfill \\{{\tau_m} = \left\{ {\begin{array}{*{20}{c}} {0.2702 + 1.1622{e^{{ - \frac{{{{\left( {V + 22.098} \right)}^2}}}{{164.19}}}}}} \hfill & {{\hbox{if }}V \geqslant - {40}\,{\hbox{mV,}}} \hfill \\{0.6923{e^{{\frac{{V - 4.7}}{{1,089.372}}}}}} \hfill & {\hbox{otherwise}} \hfill \\\end{array} } \right\}} \hfill \\\end{array} $$
The factor g GHK is a current per unit permeability.

T-Type Ca2+ Channel

The low threshold voltage-activated, T-type calcium current was modeled using Cav 3.1 data from Iftinca et al. [31]. It included two activation (m) gates and an inactivation (h) gate. The current density is based on the Goldman–Hodgkin–Katz equation [30]. The equations describing its kinetics are summarized below:
$$ \begin{array}{*{20}{c}} {{I_{{{\rm{C}}{{\rm{a}}^{{2 + }}}}}} = \overline {{P_{{{\rm{C}}{{\rm{a}}^{{2 + }}}}}}} \times {m^2} \times h \times {g_{\rm{GHK}}}} \hfill \\{\begin{array}{*{20}{c}} {{m_{\infty }} = \frac{1}{{1 + {e^{{ - \frac{{V + 52}}{5}}}}}},} \hfill & {{\tau_m} = \left\{ {\begin{array}{*{20}{c}} 1 \hfill & {i{\hbox{f}}\,V\, \leqslant - 90\,{\hbox{mV}},} \hfill \\{1 + \frac{1}{{{e^{{\frac{{V + 40}}{9}}}} + {e^{{ - \frac{{V + 102}}{{18}}}}}}}} \hfill & {\hbox{otherwise}} \hfill \\\end{array} } \right\}} \hfill \\\end{array} } \hfill \\{\begin{array}{*{20}{c}} {{h_{\infty }} = \frac{1}{{1 + {e^{{\frac{{V + 72}}{7}}}}}},} \hfill & {{\tau_h} = 15 + \frac{1}{{{e^{{\frac{{V + 32}}{7}}}}}}} \hfill \\\end{array} } \hfill \\\end{array} $$
The factor g GHK is a current per unit permeability.

BK-Type KCa Channel

The BK-type KCa channel was based on a kinetic scheme proposed by Cox et al. [32]. The channel model has a single voltage-dependent gate and four binding sites for Ca2+ ions, and Ca2+ binding at each site modulates the opening/closing rate coefficients. The kinetic scheme and rate constants used in the BK model were obtained from scheme II and patch 3 data in Table III in reference [32].

SK-Type KCa Channel

We used the SK-type channel from a Golgi cell model [33], which was based on data from Hirschberg et al. [34].

Leak Current

The leak current was modeled as a linear voltage-independent conductance following Ohm's law:
$$ {I_{\rm{leak}}} = {G_{\rm{leak}}}\left( {V - {E_{\rm{leak}}}} \right), $$
where \( {G_{\rm{leak}}} = {10^{{ - 6}}}\,{\hbox{S/c}}{{\hbox{m}}^2} \) and \( {E_{\rm{leak}}} = - 61\,{\hbox{mV}} \).

Intracellular and Extracellular Ca2+

Intracellular Ca2+ was modeled using different buffering mechanisms described in the following section. Extracellular Ca2+ was maintained at 2 mM during all the simulations.

Ca2+ Buffering Models

Intracellular Ca2+ was modeled using the following Ca2+ buffering mechanisms.

Single Pool

The exponential decaying Ca2+ pool was modeled as
$$ \frac{{d\left[ {{\text{Ca}}_i^{{2 + }}} \right]}}{{dt}} = - \frac{{{I_{{{\rm{C}}{{\rm{a}}^{{2 + }}}}}}(t)}}{{2Fd}} - \beta \left( {\left[ {{\hbox{Ca}}_i^{{2 + }}} \right] - \left[ {{\hbox{Ca}}_o^{{2 + }}} \right]} \right), $$
where \( \left[ {{\hbox{Ca}}_i^{{2 + }}} \right] \) is intracellular Ca2+, \( \left[ {{\hbox{Ca}}_o^{{2 + }}} \right] \) is Ca2+ at rest and is 45 nM, \( {I_{{{\rm{C}}{{\rm{a}}^{{2 + }}}}}}(t) \) is total current through P-type and T-type Ca2+ channels, F is Faraday's constant, d is depth of a submembrane shell to define the volume for effective Ca2+ concentration, and β is the decay time constant. Instead of using commonly used values for β and d, we tuned these parameters to approximate the Ca2+ transients given by detailed Ca2+ dynamics, as described in the “Tuning of Single Pool Models” section.

Two Pools

Fast and slow Ca2+ transients, \( {\left[ {{\hbox{Ca}}_i^{{2 + }}} \right]_{\rm{f}}} \) and \( {\left[ {{\hbox{Ca}}_i^{{2 + }}} \right]_{\rm{s}}} \), were modeled using two pools with different decay time constants, β f and β s (β f > β s), and depths, d f and d s (d f < d s). The total intracellular Ca2+ \( \left[ {{\hbox{Ca}}_i^{{2 + }}} \right] \) was modeled as the weighted sum of fast and slow Ca2+ transients.
$$ \left[ {Ca_i^{{2 + }}} \right] = {f_{\rm{f}}}{\left[ {Ca_i^{{2 + }}} \right]_{\rm{f}}} + {f_{\rm{s}}}{\left[ {Ca_i^{{2 + }}} \right]_{\rm{s}}} $$
The parameters β f, β s, d f, d s, f f, and f s were tuned to approximate the Ca2+ transients given by detailed Ca2+ dynamics, as described in the “Tuning of Double Pool” section.

Detailed Ca2+ Dynamics

The detailed Ca2+ dynamics model included calbindin (CB) and parvalbumin (PV) as buffers. In addition to Ca2+, both PV and 80% of CB were diffusible. A single surface-based Ca2+ pump was modeled using Michaelis–Menten kinetics [35]. The kinetics of CB and PV was obtained from a pre-existing Ca2+ dynamics model [25] and is described in Table 1. The kinetics of the pump was tuned to approximately match the Ca2+ decay measured at very high intracellular Ca2+ concentration [36], where the buffers are almost saturated and the decay is solely due to pump and diffusion. The outer radial shell had a depth (d) of 0.1 μm, whereas other radial shells had a depth of 0.2 μm each.
Table 1
Detailed Ca2+ dynamics model parameters
Parameter
Value
Reference
[Ca2+] at rest
45 nM
[43]
[Mg2+] at rest
590 μM
[25]
Diffusion rates
  
DCa
0.233 μm2/ms
[44]
DCB
0.028 μm2/ms
[25]
DPV
0.043 μm2/ms
[25]
Calbindin
  
Concentration
0.16 mM
a
k on,fast
43.5 ms−1 mM−1
[45]
k off,fast
3.58 × 10−2 ms−1
[45]
k on,slow
5.5 ms−1 mM−1
[45]
k off,slow
0.26 × 10−2 ms−1
[45]
Parvalbumin
  
Concentration
0.08 mM
a
k on,Ca
107 ms−1 mM−1
[25]
k off,Ca
9.5 × 10−4 ms−1
[46]
k on,Mg
0.8 ms−1 mM−1
[25]
k off,Mg
2.5 × 10−2 ms−1
[46]
Pump
  
Density
1 × 10−9 mol cm−2
Estimated
k on
3 × 10−3 ms−1 mM−1
Estimated
k off
1.75 × 10−5 ms−1
Estimated
k ext
7.255 × 10−5 ms−1 mM−1
Estimated
Fraction of immobile calbindin
0.2
[25]
aData provided by Dr. Klaus M. Stiefel, personal communication

Detailed Ca2+ Dynamics with Diffusion Compensation

The detailed Ca2+ dynamics model was modified by removing the diffusion of Ca2+ and buffers. A diffusion compensating mechanism (DCM) was included to compensate for the reduced removal of Ca2+ from the submembrane region due to the lack of diffusion towards the center of the compartment. The DCM follows a standard buffering scheme:
$$ \left[ {{\hbox{C}}{{\hbox{a}}^{{2 + }}}} \right] + \left[ {\hbox{DCM}} \right]\mathop {{{k_{\rm{off}}}}}{\overset{{{k_{\rm{on}}}}}{\longleftrightarrow}}\left[ {{\hbox{C}}{{\hbox{a}}^{{2 + }}} - {\hbox{DCM}}} \right] $$
The total concentration of DCM and rate constants (k on and k off) were tuned to compensate for the diffusion of Ca2+ and mobile buffers in the absence of radial diffusion. Moreover, the depth of the single submembrane shell, which was used to compute the effective Ca2+ concentration for KCa channels, was also a tunable parameter.

Tuning of Ca2+ Buffering Models

Generating Target Traces

A single compartment was modeled with a diameter of 4 μm and a length of 20 μm. The P-type channel was included in the model for Ca2+ influx, and the detailed Ca2+ dynamics were used as a buffering model. A voltage step protocol, shown in Fig. 1a, was used to depolarize the compartment to the voltage at which physiological dendritic Ca2+ spikes are generated (an example of an experimental recording is shown for comparison in Fig. 1b). Then, different conductance values for P-type Ca2+ channel were used to generate peak amplitude of Ca2+ concentrations of 0.5, 1, 2, 4, and 8 μM in the volume defined by the submembrane shell. These simulated calcium transients formed the target traces for the automated fitting.
Similarly, the experimental voltage trace of Ca2+ spikes, shown in Fig. 1b, was used as the voltage clamp control signal to generate calcium transients with peak amplitudes of Ca2+ concentrations of ~1.25, 2.5, 5, and 10 μM. These simulated calcium transients were later used to tune parameters for pool-based Ca2+ buffering models.

Tuning of Single Pool Models

The same model as described above was used, but now with a single pool. We used Neurofitter [37], an automated parameter search method to find optimal values of d and β to match the single pool simulations with the target traces. We used the random search mode, and model traces were compared to the target using a standard root mean square (RMS) error measure. While using the step voltage protocol (shown in Fig. 1a), since each of the traces used in fitting had two distinct features: (1) a fast rise and decay and (2) a slow decay (Fig. 3), we separated those features by specifying two time periods, 500 to 550 ms and 550 to 5,000 ms, over which separate RMS values were computed and added together. Similarly, while using the experimental voltage protocol (shown in Fig. 1b), we separated each of the traces used in fitting by specifying two time periods, 10 to 25 ms and 25 to 50 ms, over which separate RMS values were computed and added together. The best values for d and β obtained to fit the data from Fig. 1b were used with the single pool model to simulate dendritic Ca2+ spikes.

Tuning of Double Pool

Similar procedures were used to search for the values of β f, β s, d f, d s, f f, and f s values.

Tuning of DCM

We replaced diffusion of Ca2+, of free mobile buffers, and Ca2+-bound mobile buffers (keeping the total buffer concentrations constant) from the detailed Ca2+ dynamics model with a single equation for DCM. Similar procedures were used as in the case of tuning the single pool, except that we only used target traces generated using the step voltage protocol (shown in Fig. 1a), to find values of [DCM], k on, k off, and d. In order to find a robust solution, Neurofitter was run with five different seeds for random number generation. The five best solutions found using each seed were selected. We assume that the parameters for the diffusion compensating mechanism should be robust for different Ca2+ concentration peak amplitudes, but not for different compartment diameters. Therefore, the fitting procedure was repeated for diameters of 0.8, 2, 4, 6, 8, 10, 12, 16, and 20 μm.

DCM Parameter Interpolation

In actual compartmental models of neurons [38], a compartment can have an arbitrary diameter. Therefore, the parameters fitted to specific diameters were fitted to continuous functions. The 25 best solutions for each of the four parameters were plotted against corresponding diameter values. Matlab routines were used to fit the data with a single exponential function, double exponential function, and fifth order polynomial function based on an RMS measure. Later, those functions were used to predict the diffusion compensating parameter values for any compartment diameter.

Tuning Channel Densities for Dendritic Ca2+ Spikes

A dendritic Ca2+ spike, with three spikelets, was generated using the detailed Ca2+ dynamics model by hand tuning the channel density values. Later, Neurofitter was used to search proper channel densities for models with single pool, double pool, or DCM to generate similar dendritic Ca2+ spikes. The search was based on the RMS error measurement. We used five features for the fitness function. Three features comprised the three spikelets; other features were the pre-spike depolarization and post-spike hyperpolarization.

Results

Detailed Ca2+ Dynamics Model

A single compartment model with 4 μm diameter and 20 μm length was simulated with a P-type Ca2+ channel and the detailed Ca2+ dynamics model. The detailed Ca2+ dynamics model included calbindin (CB) and parvalbumin (PV) as buffers. The four binding sites of CB were simulated by assuming a pair of fast and a pair of slow binding sites as, respectively, a single fast and a single slow binding site (described as CB_f_s). PV was simulated together with Mg2+. Due to medium affinity of PV for Mg2+, its binding kinetics to Ca2+ is significantly affected by Mg2+. Diffusion of Ca2+, PV, and 80% of CB was included in the model. We simulated the voltage step of Fig. 1a in our model using different conductance values for P-type Ca2+ channel to generate peak amplitudes of 0.5 to 8 μM Ca2+ concentration. As the intracellular Ca2+ concentration started rising due to the influx through the P-type channel, the buffers and the pump became active. The changes in concentration of Ca2+ and Ca2+-bound buffers and outflux due to the pump over time are shown in Fig. 2.

Dynamics of Single Pool and Double Pool Models

The pool parameters were tuned with Neurofitter [37], as described in “Materials and Methods” section. The best solution for the single pool, using the step voltage protocol (Fig. 1a), was β = 1.35/ms and d = 0.891 μm. The best parameter values for the double pool were β f = 3.77/ms, d f = 0.351 μm, β s = 0.00306/ms, d s = 0.928 μm, f f = 0.994, and f s = 0.006.
Figure 3 compares the intracellular Ca2+ profiles obtained using the tuned single pool and double pool models with those in the detailed Ca2+ dynamics model. The comparison demonstrates that the single pool approximates the Ca2+ transient only during the active Ca2+ influx and the initial fast decay. Moreover, the approximation is valid only for a small range of Ca2+ influx. Therefore, the single pool cannot approximate the effects of physiological buffering. The double pool can approximate Ca2+ transient during the active Ca2+ influx and following fast and slow Ca2+ decays, but its validity is also limited to a small range of Ca2+ influx values. Although the double pool model could not approximate the full range of detailed Ca2+ concentrations, it approximated the major components of the detailed model reasonably well, i.e., the fast and slow transients. Therefore, a double pool model is expected to provide a better control over activation of KCa channels than a single pool model.
Since these estimated parameters belong to phenomenological models, the parameter values could vary based on the applied protocol. To estimate the ability of pool-based models to capture more complex Ca2+ kinetics, we applied also the experimental waveform (Fig. 1b) and estimated new model parameters. The best solution for the single pool was very different from the model used in Fig. 3: β = 6.86/ms and d = 0.169 μm. The same applied to the best parameter values for the double pool, though the ratio between fast and slow pool did not change: β f = 7.33/ms, d f = 0.167 μm, β s = 0.00795/ms, d s = 0.683 μm, f f = 0.995, and f s = 0.005. The behavior of these models (Supplementary Fig. 1) was similar to that observed in Fig. 3: their optimal concentration range is limited, and a single pool model cannot capture the slow transients.

Diffusion Compensation Model

Modeling buffers and pumps without diffusion of Ca2+ and buffers resulted in an accumulation of Ca2+ in the submembrane volume defined by a shell of depth 0.1 μm in compartments with a diameter of 1 μm or larger (Fig. 4). Therefore, we developed a compensating mechanism, DCM, to reduce the steep rise in peak amplitudes of intracellular Ca2+. Because the detailed Ca2+ dynamics model had multiple buffers with different buffer affinities, the effects of diffusion depended on the amount of Ca2+ entry. We developed DCM to robustly reproduce model Ca2+ transients in compartments with a wide range of diameters and for physiological peak amplitudes in the submembrane shell of between 0.5 and 8 μM.
The parameters for DCM were estimated using Neurofitter [37]. Examples for different levels of Ca2+ influx using the voltage step protocol demonstrate the effect of removing diffusion, and the good compensation by the DCM over the entire range is shown in Fig. 4.

Derivation of the Parameter Predictors

We showed in Fig. 4 that DCM can effectively compensate for excluding diffusion, but the parameter fitting produced values only for specific compartment diameters. To use DCM in a morphologically detailed PC model, we need a way to predict the DCM parameter values for any diameter present in the PC. Figure 5 shows the results of automated parameter fitting (see “Materials and Methods” section), for nine diameters. Using Matlab, the four parameters (concentration of DCM, [DCM]; forward rate constant, k on; the backward rate constant, k off; and the depth of submembrane shell, d) were fitted to functions (Fig. 5). The prediction functions for the parameters are
$$ \begin{array}{*{20}{c}} {\left[ {\hbox{DCM}} \right] = 64.2 - 57.3{e^{{ - \frac{\rm{diam}}{{1.4}}}}}} \hfill \\{{k_{\rm{on}}} = 0.162 - 0.106{e^{{ - \frac{\rm{diam}}{{2.29}}}}}} \hfill \\{{k_{\rm{off}}} = \left\{ {\begin{array}{*{20}{c}} {0.000267 + 0.0167{e^{{ - \frac{\rm{diam}}{{0.722}}}}} + 0.0028{e^{{ - \frac{\rm{diam}}{4}}}}} \hfill & {i{\hbox{f}}\,{\hbox{diam}} \geqslant {2,}} \hfill \\{0.003} \hfill & {\hbox{otherwise}} \hfill \\\end{array} } \right\}} \hfill \\{d = \frac{\text{diam}}{{4 \times \left( { - 0.674 + 1.94\left( {\hbox{diam}} \right) + 0.289{{\left( {\hbox{diam}} \right)}^2} - 3.33 \times {{10}^{{ - 2}}}{{\left( {\hbox{diam}} \right)}^3} + 1.55 \times {{10}^{{ - 3}}}{{\left( {\hbox{diam}} \right)}^4} - 2.55 \times {{10}^{{ - 5}}}{{\left( {\hbox{diam}} \right)}^5}} \right)}}} \hfill \\\end{array} $$
The fitted functions are very close to the mean values of the parameter distributions and reasonably capture the relationship between the DCM parameters and the diameter of the compartment. Therefore, we expect these functions to predict DCM parameters for any diameter in the PC. This is demonstrated for two unfitted diameters in Fig. 6. It shows that the fitted functions indeed produce good compensation for diffusion in the detailed model. Therefore, we can use the fitted functions with morphologically detailed PC models.

Dendritic Ca2+ Spikes

Dendritic Ca2+ spikes with multiple spikelets comparable to those observed in vitro (Fig. 1b) were simulated using a single pool, a double pool, detailed Ca2+ dynamics, and DCM. Since the buffering models used in the simulations were different approximations of detailed Ca2+ dynamics, we expected the intracellular Ca2+ transients generated by these models to show considerable differences (Supplementary Fig. 3). Therefore, we first optimized values for the maximal conductance (G max and P max) of ion channels (current activation profiles are shown in Supplementary Fig. 4) in the model to generate as comparable Ca2+ spikes as possible. This is identical to the procedure that would be used to generate a neuron model with only one type of Ca2+ dynamics model [38]. The resulting values are listed in Table 2. The resultant voltage traces of the Ca2+ spikes are shown in Fig. 7 (see also Supplementary Fig. 2).
Table 2
Maximal conductance values used to generate dendritic Ca2+ spikes
Buffering model
P max (CaP) ×10−4 cm/s
P max (CaT) ×10−6 cm/s
G max (BK) ×10−2 S/cm2
G max (SK) ×10−4 S/cm2
Single pool
1.87
7.32
5.34
6.07
Double pool
1.95
7.43
5.65
4.68
Detailed Ca2+ dynamics model
2.00
8.00
7.00
3.10
Diffusion compensated model
2.2
8.2
7.01
3.02
Comparing the dendritic Ca2+ spikes generated using different buffering models (Fig. 7a), we clearly see that the dendritic spikes generated using the pool-based models are different from the spikes generated using the other buffering models and strongly depend on the pool models used (compare Supplementary Fig. 2 with Fig. 7a). The amplitude as well as the width of second and third spikelets are different from the spikelets generated using the detailed Ca2+ model. Another noticeable difference exists in the after hyperpolarization of these bursts. These differences are due to the different activation of KCa channels, in particular of the SK channel (Supplementary Fig. 4). On the other hand, the dendritic Ca2+ spikes simulated using DCM approximated the Ca2+ spikes generated using the detailed Ca2+ dynamics model quite well.
A comparison of dendritic Ca2+ spikes over a time period of 50 ms may not reveal the effect of diffusion at longer time scales. Therefore, we ran all simulations for 200 s and compared the ability of different buffering mechanisms to capture the bursting behavior over that time. Fig. 7c, d shows the spontaneous dendritic Ca2+ bursting using the detailed Ca2+ dynamics model, DCM, single pool, and double pool models. Each of the bursts contains three Ca2+ spikelets (examples shown in Fig. 7a, b). The frequency of bursts in case of pool-based models is approximately five to six times higher than the frequency of bursts in the detailed Ca2+ dynamics model. However, the frequency of bursts for the DCM model is similar to that of the detailed model.
Comparing the burst of Ca2+ spikes around 57 s (marked with red asterisk in Fig. 7c, d) in Fig. 7b, we see that the shape has changed compared to the first burst (Fig. 7a), which is due to the buildup of Ca2+ over longer time. Pool-based models failed to capture this burst shape adaptation, while DCM reproduced it fairly well.
Next, we investigated the ability of DCM to approximate characteristics of diffusion over the longer time period when the bursting period was modified by current injection. The mean inter-burst interval plotted against the injected current is shown in Fig. 7e. The behavior of detailed Ca2+ dynamics model and DCM is similar at most current injection levels, except for bumps around 0.002 pA. These bumps represent the transition of the firing pattern from bursts with three spikelets to bursts with two spikelets (shown in Fig. 7f), which occurs at lower current levels in the detailed model. As the pool models do not show burst shape adaptation, they also entirely fail to capture this transition to two spikelet bursts. The spontaneous bursting period range is limited. Therefore, to compare the different buffering models at faster bursting rates, we injected current pulses at 1 Hz (Supplementary Fig. 5). Under these conditions, the Ca2+ buildup is more pronounced leading to a SK current mediated after hyperpolarization and a progressive decrease in the number of spikes in each burst, but the relative accuracy of the different buffering models is similar to that shown in Fig. 7.
Based on the data shown in Fig. 7, we conclude that DCM is quite effective over a wide range of conditions. While the approximation is not perfect, it is clearly much superior compared to the pool-based models. Finally, we briefly explored the robustness of DCM to changes in the Ca2+ buffers. Supplementary Fig. 6 shows a comparison of detailed and DCM models when the concentration of the CB and PV buffers is either halved or doubled (DCM parameters unchanged but ion channel densities retuned). The DCM approximation is very good for the half buffer concentration model, but less adequate for the double buffer concentration one. Therefore, we recommend to retune the DCM parameters whenever the detailed Ca2+ dynamics model is changed.

Discussion

We compared the effectiveness of different Ca2+ buffering mechanisms in controlling KCa channels expressed in PCs. The PC is known to express high amount of mobile endogenous buffers like calbindin and parvalbumin [39]. Ca2+ entering through voltage-gated channels is quickly removed by these buffers [19]. The internal Ca2+ stores and pumps together with calbindin and parvalbumin control the intracellular Ca2+ available to KCa channels. Therefore, careful modeling of intracellular Ca2+ dynamics is essential to simulate the physiological control of KCa channels.
The detailed Ca2+ dynamics model introduced by Schmidt et al. [25] includes calbindin, parvalbumin, pumps, and diffusion of Ca2+, parvalbumin, and calbindin. The model was tuned for Ca2+ dynamics in a single spine coupled to a dendrite. Since the spine is a small compartment, only longitudinal diffusion was modeled from the spine through the spine neck. In this study, we modeled dendrites, which are relatively large compartments; therefore, we modified the Schmidt et al. [25] model to incorporate radial diffusion towards the center of the compartment using the formalization implemented in NEURON [40]. Recent Ca2+ dynamics models [25, 41] used different pump rates to fit their data. We tuned the pump rate parameters to match the Ca2+ profiles recorded from a PC at high concentrations of Ca2+, at which most of the buffers were saturated, and the decay could be attributed to diffusion and pump only [36].
The single Ca2+ pool has been the choice for modeling intracellular Ca2+ in almost all compartmental models. The volume of sub-compartment to compute the effective Ca2+ concentration for activation of KCa channels is often defined by a submembrane shell with a depth of 0.2 μm [20]. However, a wide range of decay rates (β values) is used from 10/ms [20] to 0.02/ms [42]. These parameter values used in single pool models can reasonably approximate the changes in microdomains [23] and can be used to control a BK-type KCa channel [17, 18]. SK-type KCa channels, which are relatively far from Ca2+ channels [18], require smaller amounts of Ca2+ (0–2 μM) and activate at relatively long time scales. Therefore, it is impossible to control both BK and SK channels correctly using a single pool. We confirmed this prediction by comparing the Ca2+ profiles (Supplemental Fig. 3) and dendritic Ca2+ spikes (Fig. 7 and Supplementary Figs. 2 and 5) generated by single pool and detailed models. We also considered a Ca2+ pool model with two time constants as a phenomenological solution to the problem. While such models fitted simple Ca2+ spikes better (Fig. 3), their concentration range is limited, and they fail to simulate the period of repetitive bursting and burst shape adaptation (Fig. 7).
The limitation of using detailed Ca2+ dynamics is the computational cost of underlying diffusion [27]. We computed run times for simulations of 10 s with detailed Ca2+ dynamics and different compartment diameters (Fig. 8). It took 45 s to run the simulation for a 2-μm diameter compartment. The run time increase was supra-linear with increase in diameter; for 20 μm, it was 1,076 s. This suggests that the computation cost will increase tremendously for simulating a morphologically detailed PC model, which we want to avoid. In this study, we proposed a method to approximate the effect of diffusion. Since only Ca2+ in the submembrane region is required to control KCa channels, it is not essential to compute the Ca2+ concentration in and around the center of compartments. We replaced diffusion with a compensating mechanism, DCM, that made up for the Ca2+ diffusing towards the center. We confirmed that the DCM approximated the effect of diffusion robustly in the range of 0.5–8 μM Ca2+ for dendritic compartments with diameters of 1–20 μm and over time ranges from tens of milliseconds to hundreds of seconds. Though BK channels are reported to sense approximately 100 μM of Ca2+, those high concentrations are only available in nanodomains [18]. For volumes defined by ~0.1 μm shells, 8–10 μM of Ca2+ concentration should represent the limit of the physiological range [17]. To achieve this level of accuracy, it is important to fit the DCM parameters to the detailed Ca2+ dynamics model used (Figs. 4 and 5). For dendritic compartments with diameters less than 1 μm, the effect of diffusion is not significant and can be compensated by tuning only the depth (d) of the submembrane shell (results not shown in this work).
The method proposed in this paper is not limited to modeling PCs. The DCM parameters can be tuned to replace diffusion in large compartmental models of other neuronal types. To formulate robust DCM parameters, it is necessary to have a faithful detailed Ca2+ dynamics model of the specific neuronal type. Then, for the desired physiological range of Ca2+ influx and range of diameters in the morphology, DCM parameter equations can be derived to replace diffusion.
The use of the DCM resulted in a decrease of the run time to 8 s with a loss of dependence on diameter (Fig. 8). We conclude that using a detailed Ca2+ buffering model combined with DCM is a clear improvement compared to the Ca2+ pool for modeling Ca2+ dynamics in a large neuron model.

Acknowledgements

We acknowledge Dr. Klaus M. Stiefel for providing the detailed Ca2+ dynamics model scripts which were extended for use in this work. This work was supported by OISTPC.

Conflicts of interest

There is no conflict of interest in the work presented in this manuscript.

Open Access

This article is distributed under the terms of the Creative Commons Attribution Noncommercial License which permits any noncommercial use, distribution, and reproduction in any medium, provided the original author(s) and source are credited.
Open AccessThis is an open access article distributed under the terms of the Creative Commons Attribution Noncommercial License (https://​creativecommons.​org/​licenses/​by-nc/​2.​0), which permits any noncommercial use, distribution, and reproduction in any medium, provided the original author(s) and source are credited.
Model scripts All the scripts used to run simulations in this manuscript are available at ModelDB (http://​senselab.​med.​yale.​edu/​ModelDB).

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Neurologie & Psychiatrie

Kombi-Abonnement

Mit e.Med Neurologie & Psychiatrie erhalten Sie Zugang zu CME-Fortbildungen der Fachgebiete, den Premium-Inhalten der dazugehörigen Fachzeitschriften, inklusive einer gedruckten Zeitschrift Ihrer Wahl.

e.Med Neurologie

Kombi-Abonnement

Mit e.Med Neurologie erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes, den Premium-Inhalten der neurologischen Fachzeitschriften, inklusive einer gedruckten Neurologie-Zeitschrift Ihrer Wahl.

Anhänge

Electronic supplementary material

Literatur
1.
Zurück zum Zitat Gähwiler BH, Llano I. Sodium and potassium conductances in somatic membranes of rat Purkinje cells from organotypic cerebellar cultures. J Physiol. 1989;417:105–22.PubMed Gähwiler BH, Llano I. Sodium and potassium conductances in somatic membranes of rat Purkinje cells from organotypic cerebellar cultures. J Physiol. 1989;417:105–22.PubMed
2.
Zurück zum Zitat Gruol DL. Chronic exposure to alcohol during development alters the membrane properties of cerebellar Purkinje neurons in culture. Brain Res. 1991;558(1):1–12.PubMedCrossRef Gruol DL. Chronic exposure to alcohol during development alters the membrane properties of cerebellar Purkinje neurons in culture. Brain Res. 1991;558(1):1–12.PubMedCrossRef
3.
Zurück zum Zitat Knaus HG, Schwarzer C, Koch RO, Eberhart A, Kaczorowski GJ, Glossmann H, et al. Distribution of high-conductance Ca(2+)-activated K+ channels in rat brain: targeting to axons and nerve terminals. J Neurosci. 1996;16(3):955–63.PubMed Knaus HG, Schwarzer C, Koch RO, Eberhart A, Kaczorowski GJ, Glossmann H, et al. Distribution of high-conductance Ca(2+)-activated K+ channels in rat brain: targeting to axons and nerve terminals. J Neurosci. 1996;16(3):955–63.PubMed
4.
Zurück zum Zitat Jacquin TD, Gruol DL. Ca2+ regulation of a large conductance K+ channel in cultured rat cerebellar Purkinje neurons. Eur J Neurosci. 1999;11(2):735–9.PubMedCrossRef Jacquin TD, Gruol DL. Ca2+ regulation of a large conductance K+ channel in cultured rat cerebellar Purkinje neurons. Eur J Neurosci. 1999;11(2):735–9.PubMedCrossRef
5.
Zurück zum Zitat Womack MD, Khodakhah K. Characterization of large conductance Ca2+-activated K+ channels in cerebellar Purkinje neurons. Eur J Neurosci. 2002;16(7):1214–22.PubMedCrossRef Womack MD, Khodakhah K. Characterization of large conductance Ca2+-activated K+ channels in cerebellar Purkinje neurons. Eur J Neurosci. 2002;16(7):1214–22.PubMedCrossRef
6.
Zurück zum Zitat Edgerton JR, Reinhart PH. Distinct contributions of small and large conductance Ca2+-activated K+ channels to rat Purkinje neuron function. J Physiol. 2003;548(1):53–69.PubMedCrossRef Edgerton JR, Reinhart PH. Distinct contributions of small and large conductance Ca2+-activated K+ channels to rat Purkinje neuron function. J Physiol. 2003;548(1):53–69.PubMedCrossRef
7.
Zurück zum Zitat Haghdoost-Yazdi H, Janahmadi M, Behzadi G. Iberiotoxin-sensitive large conductance Ca2+-dependent K+ (BK) channels regulate the spike configuration in the burst firing of cerebellar Purkinje neurons. Brain Res. 2008;1212:1–8.PubMedCrossRef Haghdoost-Yazdi H, Janahmadi M, Behzadi G. Iberiotoxin-sensitive large conductance Ca2+-dependent K+ (BK) channels regulate the spike configuration in the burst firing of cerebellar Purkinje neurons. Brain Res. 2008;1212:1–8.PubMedCrossRef
8.
Zurück zum Zitat Cheron G, Sausbier M, Sausbier U, Neuhuber W, Ruth P, Dan B, et al. BK channels control cerebellar Purkinje and golgi cell rhythmicity in vivo. PLoS ONE. 2009;4(11):e7991.PubMedCrossRef Cheron G, Sausbier M, Sausbier U, Neuhuber W, Ruth P, Dan B, et al. BK channels control cerebellar Purkinje and golgi cell rhythmicity in vivo. PLoS ONE. 2009;4(11):e7991.PubMedCrossRef
9.
Zurück zum Zitat Kaufmann WA, Ferraguti F, Fukazawa Y, Kasugai Y, Shigemoto R, Laake P, et al. Large-conductance calcium-activated potassium channels in Purkinje cell plasma membranes are clustered at sites of hypolemmal microdomains. J Comp Neurol. 2009;515(2):215–30.PubMedCrossRef Kaufmann WA, Ferraguti F, Fukazawa Y, Kasugai Y, Shigemoto R, Laake P, et al. Large-conductance calcium-activated potassium channels in Purkinje cell plasma membranes are clustered at sites of hypolemmal microdomains. J Comp Neurol. 2009;515(2):215–30.PubMedCrossRef
10.
Zurück zum Zitat Womack MD, Hoang C, Khodakhah K. Large conductance calcium-activated potassium channels affect both spontaneous firing and intracellular calcium concentration in cerebellar Purkinje neurons. Neuroscience. 2009;162(4):989–1000.PubMedCrossRef Womack MD, Hoang C, Khodakhah K. Large conductance calcium-activated potassium channels affect both spontaneous firing and intracellular calcium concentration in cerebellar Purkinje neurons. Neuroscience. 2009;162(4):989–1000.PubMedCrossRef
11.
Zurück zum Zitat Stocker M, Pedarzani P. Differential distribution of three Ca(2+)-activated K(+) channel subunits, SK1, SK2, and SK3, in the adult rat central nervous system. Mol Cell Neurosci. 2000;15(5):476–93.PubMedCrossRef Stocker M, Pedarzani P. Differential distribution of three Ca(2+)-activated K(+) channel subunits, SK1, SK2, and SK3, in the adult rat central nervous system. Mol Cell Neurosci. 2000;15(5):476–93.PubMedCrossRef
12.
Zurück zum Zitat Cingolani LA, Gymnopoulos M, Boccaccio A, Stocker M, Pedarzani P. Developmental regulation of small-conductance Ca2+-activated K+ channel expression and function in rat Purkinje neurons. J Neurosci. 2002;22(11):4456–67.PubMed Cingolani LA, Gymnopoulos M, Boccaccio A, Stocker M, Pedarzani P. Developmental regulation of small-conductance Ca2+-activated K+ channel expression and function in rat Purkinje neurons. J Neurosci. 2002;22(11):4456–67.PubMed
13.
Zurück zum Zitat Womack MD, Khodakhah K. Somatic and dendritic small-conductance calcium-activated potassium channels regulate the output of cerebellar Purkinje neurons. J Neurosci. 2003;23(7):2600–7.PubMed Womack MD, Khodakhah K. Somatic and dendritic small-conductance calcium-activated potassium channels regulate the output of cerebellar Purkinje neurons. J Neurosci. 2003;23(7):2600–7.PubMed
14.
Zurück zum Zitat Yazdi HH, Janahmadi M, Behzadi G. The role of small-conductance Ca2+-activated K+ channels in the modulation of 4-aminopyridine-induced burst firing in rat cerebellar Purkinje cells. Brain Res. 2007;1156:59–66.PubMedCrossRef Yazdi HH, Janahmadi M, Behzadi G. The role of small-conductance Ca2+-activated K+ channels in the modulation of 4-aminopyridine-induced burst firing in rat cerebellar Purkinje cells. Brain Res. 2007;1156:59–66.PubMedCrossRef
15.
Zurück zum Zitat Womack MD, Khodakhah K. Dendritic control of spontaneous bursting in cerebellar Purkinje cells. J Neurosci. 2004;24(14):3511–21.PubMedCrossRef Womack MD, Khodakhah K. Dendritic control of spontaneous bursting in cerebellar Purkinje cells. J Neurosci. 2004;24(14):3511–21.PubMedCrossRef
16.
Zurück zum Zitat Goldberg JA, Wilson CJ. Control of spontaneous firing patterns by the selective coupling of calcium currents to calcium-activated potassium currents in striatal cholinergic interneurons. J Neurosci. 2005;25(44):10230–8.PubMedCrossRef Goldberg JA, Wilson CJ. Control of spontaneous firing patterns by the selective coupling of calcium currents to calcium-activated potassium currents in striatal cholinergic interneurons. J Neurosci. 2005;25(44):10230–8.PubMedCrossRef
17.
Zurück zum Zitat Zhuge R, Fogarty KE, Tuft RA, Walsh JV. Spontaneous transient outward currents arise from microdomains where BK channels are exposed to a mean Ca(2+) concentration on the order of 10 microm during a Ca(2+) spark. J Gen Physiol. 2002;120(1):15–27.PubMedCrossRef Zhuge R, Fogarty KE, Tuft RA, Walsh JV. Spontaneous transient outward currents arise from microdomains where BK channels are exposed to a mean Ca(2+) concentration on the order of 10 microm during a Ca(2+) spark. J Gen Physiol. 2002;120(1):15–27.PubMedCrossRef
18.
Zurück zum Zitat Fakler B, Adelman JP. Control of K(ca) channels by calcium nano/microdomains. Neuron. 2008;59(6):873–81.PubMedCrossRef Fakler B, Adelman JP. Control of K(ca) channels by calcium nano/microdomains. Neuron. 2008;59(6):873–81.PubMedCrossRef
19.
Zurück zum Zitat Hartmann J, Konnerth A. Determinants of postsynaptic Ca2+ signaling in Purkinje neurons. Cell Calcium. 2005;37(5):459–66.PubMedCrossRef Hartmann J, Konnerth A. Determinants of postsynaptic Ca2+ signaling in Purkinje neurons. Cell Calcium. 2005;37(5):459–66.PubMedCrossRef
20.
Zurück zum Zitat De Schutter E, Bower JM. An active membrane model of the cerebellar Purkinje cell. I. Simulation of current clamps in slice. J Neurophysiol. 1994;71(1):375–400.PubMed De Schutter E, Bower JM. An active membrane model of the cerebellar Purkinje cell. I. Simulation of current clamps in slice. J Neurophysiol. 1994;71(1):375–400.PubMed
21.
Zurück zum Zitat Khaliq ZM, Gouwens NW, Raman IM. The contribution of resurgent sodium current to high-frequency firing in Purkinje neurons: an experimental and modeling study. J Neurosci. 2003;23(12):4899–912.PubMed Khaliq ZM, Gouwens NW, Raman IM. The contribution of resurgent sodium current to high-frequency firing in Purkinje neurons: an experimental and modeling study. J Neurosci. 2003;23(12):4899–912.PubMed
22.
Zurück zum Zitat Akemann W, Knöpfel T. Interaction of Kv3 potassium channels and resurgent sodium current influences the rate of spontaneous firing of Purkinje neurons. J Neurosci. 2006;26(17):4602–12.PubMedCrossRef Akemann W, Knöpfel T. Interaction of Kv3 potassium channels and resurgent sodium current influences the rate of spontaneous firing of Purkinje neurons. J Neurosci. 2006;26(17):4602–12.PubMedCrossRef
23.
Zurück zum Zitat Sherman A, Keizer J, Rinzel J. Domain model for ca2(+)-inactivation of Ca2+ channels at low channel density. Biophys. 1990;58(4):985–95.CrossRef Sherman A, Keizer J, Rinzel J. Domain model for ca2(+)-inactivation of Ca2+ channels at low channel density. Biophys. 1990;58(4):985–95.CrossRef
24.
Zurück zum Zitat Tegnér J, Grillner S. Interactive effects of the gababergic modulation of calcium channels and calcium-dependent potassium channels in lamprey. J Neurophysiol. 1999;81(3):1318–29.PubMed Tegnér J, Grillner S. Interactive effects of the gababergic modulation of calcium channels and calcium-dependent potassium channels in lamprey. J Neurophysiol. 1999;81(3):1318–29.PubMed
25.
Zurück zum Zitat Schmidt H, Stiefel KM, Racay P, Schwaller B, Eilers J. Mutational analysis of dendritic Ca2+ kinetics in rodent Purkinje cells: role of parvalbumin and calbindin d28k. J Physiol. 2003;551(1):13–32.PubMedCrossRef Schmidt H, Stiefel KM, Racay P, Schwaller B, Eilers J. Mutational analysis of dendritic Ca2+ kinetics in rodent Purkinje cells: role of parvalbumin and calbindin d28k. J Physiol. 2003;551(1):13–32.PubMedCrossRef
26.
Zurück zum Zitat Yamada WM, Koch C, Adams PR. Multiple channels and calcium dynamics. In: Methods in neuronal modeling: from ions to networks. Cambridge: MIT; 1998. p. 137–70. Yamada WM, Koch C, Adams PR. Multiple channels and calcium dynamics. In: Methods in neuronal modeling: from ions to networks. Cambridge: MIT; 1998. p. 137–70.
27.
Zurück zum Zitat De Schutter E. Modeling intracellular calcium dynamics. In: Computational modeling methods for neuroscientists. Cambridge: MIT; 2009. p. 93–106. De Schutter E. Modeling intracellular calcium dynamics. In: Computational modeling methods for neuroscientists. Cambridge: MIT; 2009. p. 93–106.
28.
Zurück zum Zitat Hines ML, Carnevale NT. NEURON: a tool for neuroscientists. Neuroscientist. 2001;7(2):123–35.PubMedCrossRef Hines ML, Carnevale NT. NEURON: a tool for neuroscientists. Neuroscientist. 2001;7(2):123–35.PubMedCrossRef
29.
Zurück zum Zitat Swensen AM, Bean BP. Robustness of burst firing in dissociated Purkinje neurons with acute or long-term reductions in sodium conductance. J Neurosci. 2005;25(14):3509–20.PubMedCrossRef Swensen AM, Bean BP. Robustness of burst firing in dissociated Purkinje neurons with acute or long-term reductions in sodium conductance. J Neurosci. 2005;25(14):3509–20.PubMedCrossRef
30.
Zurück zum Zitat Hille B. Selective permeability: independence. In: Ion channels and excitable membranes. Sunderland: Sinauer Associates; 2001. p. 441–70. Hille B. Selective permeability: independence. In: Ion channels and excitable membranes. Sunderland: Sinauer Associates; 2001. p. 441–70.
31.
Zurück zum Zitat Iftinca M, McKay BE, Snutch TP, McRory JE, Turner RW, Zamponi GW. Temperature dependence of t-type calcium channel gating. Neuroscience. 2006;142(4):1031–42.PubMedCrossRef Iftinca M, McKay BE, Snutch TP, McRory JE, Turner RW, Zamponi GW. Temperature dependence of t-type calcium channel gating. Neuroscience. 2006;142(4):1031–42.PubMedCrossRef
32.
Zurück zum Zitat Cox DH, Cui J, Aldrich RW. Allosteric gating of a large conductance ca-activated K+ channel. J Gen Physiol. 1997;110(3):257–81.PubMedCrossRef Cox DH, Cui J, Aldrich RW. Allosteric gating of a large conductance ca-activated K+ channel. J Gen Physiol. 1997;110(3):257–81.PubMedCrossRef
33.
Zurück zum Zitat Solinas S, Forti L, Cesana E, Mapelli J, De Schutter E, D’Angelo E. Computational reconstruction of pacemaking and intrinsic electroresponsiveness in cerebellar golgi cells. Front Cell Neurosci. 2007;1:2.PubMedCrossRef Solinas S, Forti L, Cesana E, Mapelli J, De Schutter E, D’Angelo E. Computational reconstruction of pacemaking and intrinsic electroresponsiveness in cerebellar golgi cells. Front Cell Neurosci. 2007;1:2.PubMedCrossRef
34.
Zurück zum Zitat Hirschberg B, Maylie J, Adelman JP, Marrion NV. Gating of recombinant small-conductance ca-activated K+ channels by calcium. J Gen Physiol. 1998;111(4):565–81.PubMedCrossRef Hirschberg B, Maylie J, Adelman JP, Marrion NV. Gating of recombinant small-conductance ca-activated K+ channels by calcium. J Gen Physiol. 1998;111(4):565–81.PubMedCrossRef
35.
Zurück zum Zitat Sala F, Hernández-Cruz A. Calcium diffusion modeling in a spherical neuron. Relevance of buffering properties. Biophys J. 1990;57(2):313–24.PubMedCrossRef Sala F, Hernández-Cruz A. Calcium diffusion modeling in a spherical neuron. Relevance of buffering properties. Biophys J. 1990;57(2):313–24.PubMedCrossRef
36.
Zurück zum Zitat Maeda H, Ellis-Davies GC, Ito K, Miyashita Y, Kasai H. Supralinear Ca2+ signaling by cooperative and mobile Ca2+ buffering in Purkinje neurons. Neuron. 1999;24(4):989–1002.PubMedCrossRef Maeda H, Ellis-Davies GC, Ito K, Miyashita Y, Kasai H. Supralinear Ca2+ signaling by cooperative and mobile Ca2+ buffering in Purkinje neurons. Neuron. 1999;24(4):989–1002.PubMedCrossRef
37.
Zurück zum Zitat Van Geit W, Achard P, De Schutter E. Neurofitter: a parameter tuning package for a wide range of electrophysiological neuron models. Front Neuroinformatics. 2007;1:1. Van Geit W, Achard P, De Schutter E. Neurofitter: a parameter tuning package for a wide range of electrophysiological neuron models. Front Neuroinformatics. 2007;1:1.
38.
Zurück zum Zitat De Schutter E, van Geit W. Modeling complex neurons. In: Computational modeling methods for neuroscientists. Cambridge: MIT; 2009. p. 259–84. De Schutter E, van Geit W. Modeling complex neurons. In: Computational modeling methods for neuroscientists. Cambridge: MIT; 2009. p. 259–84.
39.
Zurück zum Zitat Baimbridge KG, Celio MR, Rogers JH. Calcium-binding proteins in the nervous system. Trends Neurosci. 1992;15(8):303–8.PubMedCrossRef Baimbridge KG, Celio MR, Rogers JH. Calcium-binding proteins in the nervous system. Trends Neurosci. 1992;15(8):303–8.PubMedCrossRef
40.
Zurück zum Zitat Carnevale NT, Hines M. How to expand NEURON s library of mechanisms. In: The neuron book. Cambridge: Cambridge University Press; 2006. p. 207–64. Carnevale NT, Hines M. How to expand NEURON s library of mechanisms. In: The neuron book. Cambridge: Cambridge University Press; 2006. p. 207–64.
41.
Zurück zum Zitat Doi T, Kuroda S, Michikawa T, Kawato M. Inositol 1,4,5-trisphosphate-dependent Ca2+ threshold dynamics detect spike timing in cerebellar Purkinje cells. J Neurosci. 2005;25(4):950–61.PubMedCrossRef Doi T, Kuroda S, Michikawa T, Kawato M. Inositol 1,4,5-trisphosphate-dependent Ca2+ threshold dynamics detect spike timing in cerebellar Purkinje cells. J Neurosci. 2005;25(4):950–61.PubMedCrossRef
42.
Zurück zum Zitat Traub RD, Llinas R. The spatial distribution of ionic conductances in normal and axotomized motorneurons. Neuroscience. 1977;2(6):829–49.CrossRef Traub RD, Llinas R. The spatial distribution of ionic conductances in normal and axotomized motorneurons. Neuroscience. 1977;2(6):829–49.CrossRef
43.
Zurück zum Zitat Airaksinen MS, Eilers J, Garaschuk O, Thoenen H, Konnerth A, Meyer M. Ataxia and altered dendritic calcium signaling in mice carrying a targeted null mutation of the calbindin d28k gene. Proc Natl Acad Sci U S A. 1997;94(4):1488–93.PubMedCrossRef Airaksinen MS, Eilers J, Garaschuk O, Thoenen H, Konnerth A, Meyer M. Ataxia and altered dendritic calcium signaling in mice carrying a targeted null mutation of the calbindin d28k gene. Proc Natl Acad Sci U S A. 1997;94(4):1488–93.PubMedCrossRef
44.
Zurück zum Zitat Allbritton NL, Meyer T, Stryer L. Range of messenger action of calcium ion and inositol 1,4,5-trisphosphate. Science. 1992;258:1812–5.PubMedCrossRef Allbritton NL, Meyer T, Stryer L. Range of messenger action of calcium ion and inositol 1,4,5-trisphosphate. Science. 1992;258:1812–5.PubMedCrossRef
45.
Zurück zum Zitat Nägerl UV, Novo D, Mody I, Vergara JL. Binding kinetics of calbindin-d (28 k) determined by flash photolysis of caged Ca(2+). Biophys J. 2000;79(6):3009–18.PubMedCrossRef Nägerl UV, Novo D, Mody I, Vergara JL. Binding kinetics of calbindin-d (28 k) determined by flash photolysis of caged Ca(2+). Biophys J. 2000;79(6):3009–18.PubMedCrossRef
46.
Zurück zum Zitat Lee SH, Schwaller B, Neher E. Kinetics of Ca2+ binding to parvalbumin in bovine chromaffin cells: Implications for [Ca2+] transients of neuronal dendrites. J Physiol. 2000;525(2):419–32.PubMedCrossRef Lee SH, Schwaller B, Neher E. Kinetics of Ca2+ binding to parvalbumin in bovine chromaffin cells: Implications for [Ca2+] transients of neuronal dendrites. J Physiol. 2000;525(2):419–32.PubMedCrossRef
Metadaten
Titel
Controlling Ca2+-Activated K+ Channels with Models of Ca2+ Buffering in Purkinje Cells
verfasst von
Haroon Anwar
Sungho Hong
Erik De Schutter
Publikationsdatum
01.09.2012
Verlag
Springer-Verlag
Erschienen in
The Cerebellum / Ausgabe 3/2012
Print ISSN: 1473-4222
Elektronische ISSN: 1473-4230
DOI
https://doi.org/10.1007/s12311-010-0224-3

Weitere Artikel der Ausgabe 3/2012

The Cerebellum 3/2012 Zur Ausgabe

Leitlinien kompakt für die Neurologie

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Akuter Schwindel: Wann lohnt sich eine MRT?

28.04.2024 Schwindel Nachrichten

Akuter Schwindel stellt oft eine diagnostische Herausforderung dar. Wie nützlich dabei eine MRT ist, hat eine Studie aus Finnland untersucht. Immerhin einer von sechs Patienten wurde mit akutem ischämischem Schlaganfall diagnostiziert.

Niedriger diastolischer Blutdruck erhöht Risiko für schwere kardiovaskuläre Komplikationen

25.04.2024 Hypotonie Nachrichten

Wenn unter einer medikamentösen Hochdrucktherapie der diastolische Blutdruck in den Keller geht, steigt das Risiko für schwere kardiovaskuläre Ereignisse: Darauf deutet eine Sekundäranalyse der SPRINT-Studie hin.

Frühe Alzheimertherapie lohnt sich

25.04.2024 AAN-Jahrestagung 2024 Nachrichten

Ist die Tau-Last noch gering, scheint der Vorteil von Lecanemab besonders groß zu sein. Und beginnen Erkrankte verzögert mit der Behandlung, erreichen sie nicht mehr die kognitive Leistung wie bei einem früheren Start. Darauf deuten neue Analysen der Phase-3-Studie Clarity AD.

Viel Bewegung in der Parkinsonforschung

25.04.2024 Parkinson-Krankheit Nachrichten

Neue arznei- und zellbasierte Ansätze, Frühdiagnose mit Bewegungssensoren, Rückenmarkstimulation gegen Gehblockaden – in der Parkinsonforschung tut sich einiges. Auf dem Deutschen Parkinsonkongress ging es auch viel um technische Innovationen.

Update Neurologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.