Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2006 Sep 7;577(Pt 1):205–220. doi: 10.1113/jphysiol.2006.113043

Contribution of morphology and membrane resistance to integration of fast synaptic signals in two thalamic cell types

Marie-Claude Perreault 1, Morten Raastad 1
PMCID: PMC2000667  PMID: 16959860

Abstract

Thalamocortical cells (TCs) and interneurons (INs) in the lateral geniculate nucleus process visual information from the retina. The TCs have many short dendrites, whereas the INs have fewer and longer dendrites. Because of these morphological differences, it has been suggested that transmission of synaptic signals from dendritic synapses to soma is more efficient in TCs than in INs. However, a higher membrane resistance (Rm) for the INs could, in theory, compensate for the attenuating effect of their long dendrites and allow distal synaptic inputs to significantly depolarize the soma. Compartmental models were made from biocytin filled TCs (n = 15) and INs (n = 3) and adjusted to fit the current- and voltage-clamp recordings from the individual cells. The confidence limits for the passive electrical parameters were explored by simulating the influence of noise, morphometric errors and non-uniform and active conductances. One of the useful findings was that Rm was accurately estimated despite realistic levels of active conductance. Simulations to explore the somatic influence of dendritic synapses showed that a small (0.5 nS) excitatory synapse placed at different dendritic positions gave similar somatic potentials in the individual TCs, within the TC population and also between TCs and INs. A linear increase in the conductance of the synapse gave increases in somatic potentials that were more sublinear in INs than TCs. However, when the total synaptic conductance was increased by simultaneously activating many small, spatially distributed synapses, the INs converted the synaptic signals to soma potentials almost as efficiently as the TCs. Thus, INs can transfer fast synaptic signals to soma as efficiently as TCs except when the focal conductance is large.


The lateral geniculate nucleus (LGN) of the thalamus processes visual information using two main cell types; the excitatory thalamocortical cells (TCs) and the local inhibitory interneurons (INs). Interactions between morphology and membrane properties (passive and active) determine how these neurons integrate synaptic information.

As TCs and INs have strikingly different morphology and because the TCs have much lower somatic input resistance (Rin) than INs, at least during low network activity (Pape et al. 1994; Williams et al. 1996; Zhu & Uhlrich, 1997; Zhu et al. 1999a; Zhu & Heggelund, 2001; Briska et al. 2003; Perreault et al. 2003), their integration properties may be rather different. For instance, it has been suggested that synaptic signals originating in the dendrites are transmitted much more efficiently to soma in TCs than in INs (Bloomfield et al. 1987; Bloomfield & Sherman, 1989; Briska et al. 2003; Sherman, 2004).

The study of the conversion of synaptic conductances to somatic voltages relies on accurate neuronal models because direct voltage measurements at individual dendritic synapses are not yet available. The models that are needed must include a description of the cell morphology and the parameters describing the passive electrical behaviour of the cells (i.e. the membrane resistance (Rm), the cytoplasmic resistivity (Ri) and the membrane capacitance (Cm) (Barrett & Crill, 1974; Clements & Redman, 1989; Rall et al. 1992)).

In most types of neurons, the exact values of the passive parameters have been difficult to determine. Whereas Cm is thought to be uniform (Gentet et al. 2000), Rm and perhaps also Ri (Golding et al. 2005) may vary substantially within and between cell populations. The estimation of Rm in particular has been problematic for several reasons. First, Rm has often been assumed to be uniform over the cell membrane but this may not be the case. A non-uniform Rm may introduce a bias in the estimation of this parameter (London et al. 1999).

The presence of time- and voltage-dependent currents may also make it difficult to estimate Rm. Whereas this problem can be reduced by blocking the active currents (Major et al. 1994), this procedure may give an artificially high Rm. In the present study, we pursued a new trend whereby only a few (Stuart & Spruston, 1998; Roth & Häusser, 2001; Golding et al. 2005) or no active currents are blocked (this study and Briska et al. 2003). We use simulations to quantify the errors in Rm introduced by active currents as well as four additional types of models errors: inaccuracies due to the experimental noise, inaccuracies in the length and diameter of the dendrites (Roth & Häusser, 2001; Steuber et al. 2004; Szilagyi & De Schutter, 2004; Ambros-Ingerson & Holmes, 2005; Holmes et al. 2006), and violations in the assumption of a uniform Rm (London et al. 1999).

We found that Rm was larger in INs than in TCs and was accurately estimated at realistic levels of active conductance. Furthermore, the higher Rm of the INs compensated for the attenuating effect of their long dendrites. Because of this, distal synaptic inputs could significantly depolarize the soma. Parts of this work have been published previously in abstract form (Perreault et al. 2000).

Methods

Preparation

The experiments were performed using brain slices prepared from 21-day-old rats. Some of the present material (half of the cells) has been used in a previous study (Perreault et al. 2003). All procedures were approved by the National Animal Research Authority in Norway. The animals were deeply anaesthetized with halothane and decapitated. Blocks of brain tissue were quickly removed, submerged in ice-cold, oxygenated (95% O2–5% CO2) solution containing (mm): NaCl 125, KCl 2.5, NaH2PO4 1.25, NaHCO3 25, MgCl2 1, d-glucose 25 and CaCl2 2 at pH 7.4 and placed on a cold plate of an HR2-microslicer (Sigmann Elektronik, Germany) where coronal slices (300 μm thick) containing the LGN, but not the reticular nucleus, were cut. The slices were then incubated at 37.0 ± 0.5°C for a recovery period of > 30 min before they were transferred to an immersion recording chamber. The oxygenated physiological solution in the recording chamber was kept at room temperature (23.5 ± 0.5°C). Neurons were visualized using a 40 × water immersion objective and infrared differential interference contrast (IR-DIC) video microscopy (Zeiss Axioplan-FS2, Germany).

Electrophysiology

Patch electrodes were made from borosilicate capillaries (Harvard Apparatus, UK) and had resistance of 5–7 MΩ when filled with a solution containing (mm): potassium gluconate 115, Hepes 10, MgATP 2, MgCl2 2, Na2ATP 2, GTP 0.3 and KCl 20 at pH 7.3. Liquid junction potential was approximately −13 mV and was not included in the measurement of the resting membrane potential. Following gigaohm seals and electrode capacitance compensation, whole-cell recordings from the somata of INs and TCs were made in current-clamp (CC) and voltage-clamp (VC) modes using a HEKA 9 amplifier (Heka Elektronik, Germany). Voltage- and current-clamp recordings were made without serial resistance compensation and bridge balance. Instead the access resistance was included as a free variable in the simulations (see below). Stimulus delivery and data acquisition was performed using the Pulse software (Heka Elektronik). For CC experiments, we used 300 and 600 ms current steps that hyperpolarized the membrane less than −10 mV. For VC experiments, we used 600 ms long, −10 mV steps. Both CC and VC responses consisted of averages of three to six traces.

Histology and morphological reconstructions

All cells recorded were filled with biocytin (0.25%) added to the intracellular pipette solution. After recording, the slices were fixed in 0.1 m phosphate buffer containing 4% paraformaldehyde for up to 2 weeks. The slices were subsequently processed for biocytin labelling (Horikawa & Armstrong, 1988) using avidin–biotin peroxidase complex (ABC-Elite kit; Vector Laboratories, USA) and diaminobenzidine (DAB) reaction. Slices were mounted in the aqueous medium Mowiol (Hoechst AG, Frankfurt AM, Germany). Morphological reconstruction of the labelled cells were done using 40 × or 100 × water immersion objective and the computerized tracing program Neurolucida (MicroBright Field, USA).

Compartmental models

For each cell, the Neurolucida file containing the coordinates, diameters and branching of the dendritic and axonal processes was translated into NEURON program format (Hines, 1989; Hines & Carnevale, 1997) using custom-built software. These data were used to build a compartmental model for each neuron. The soma was modelled as a cylinder with diameter and height equal to the average between the maximal width and maximal length of the soma contour. The average number of compartments was 190 and the average compartment length 0.05 electrotonic units using Rm = 10 kΩ cm2 and Ri = 100 Ω cm.

The distortion/shrinkage error associated with the present histological procedure (no re-slicing, air-drying or dehydrating) has been estimated to be < 5% (Roth & Häusser, 2001). In agreement with this estimate, we found a small and not significant increase of 3.3 ± 0.1% (P = 0.7, n = 12) in the diameters of TC primary dendrites (measured at a distance of 6.3 ± 0.9 μm from soma) when we compared the measurements taken from the living slices (2.9 ± 0.1 μm) with those taken after mounting the slices (3.0 ± 0.3 μm). The models were not corrected for this small distortion. Similarly, we did not correct for the possibility of an increased area due to dendritic spines because we found very few spines present on our cells. Similar observations have been made in the LGN of ferrets (82 spines per cell and 0.04 spines per μm, Rocha & Sur, 1995).

Simulations and fitting procedures

The responses produced by our models (simulated responses) were fitted to the experimental CC and VC traces. As with the experimental traces, the baseline of the simulated traces was adjusted to zero by subtracting the trace level before the steps (average over a period of 50 ms). The reversal potential for the passive leak conductance was also set to the actual value of the resting potential recorded during the experiments. The time steps used in the simulations were 0.3 ms and 0.1 ms for CC and VC, respectively. In the simulation with active membrane properties (see below) we also tested time steps 1/10 of these values. This gave a difference in the resulting parameter estimates which was < 2% in all individual cases.

We routinely fitted 300 and 50 ms of the CC and VC traces, respectively. For the CC traces, fitting started at the onset of the applied current pulse. For VC traces, fitting started 3 ms after the beginning of the voltage step to avoid electrode capacitance artifacts. To find the best fits, a simplex algorithm (Press et al. 1986) was used to minimize the squared error between the simulated and the experimental traces. The free variables in our models were Rm, Cm, Ri as well as the access resistance (Racc). The latter was modelled either as an additional voltage step of amplitude equal to electrode resistance × injected current (CC) or as the electrode resistance (VC). We chose to include Racc as a free variable because simulations (explained in the Results) showed that this strategy gave the most accurate estimates for Ri, with only a modest reduction of the accuracy of the other parameters. Moreover, the commonly used estimate based on the initial peak current in VC requires an accurate measurement of the somatic voltage (Jackson, 1992), which is not available with single-electrode voltage clamp.

Simulations of model errors

To explore the errors that may result from factors not accounted for in the original models, we performed a series of simulations where we modified the original models in the following ways (cf. ‘modified models’ in Fig. 3). We multiplied all diameters and lengths in the original models with a factor drawn from a uniform distribution between 0.8 and 1.2. We introduced active and non-uniform conductances (described separately below) also by drawing values for maximal conductances from the ranges given in the table in Fig. 3. We also added noise. This noise was a scaled version of noise recorded during the real experiments and was added arithmetically to the synthetic traces. The scaling was done so that the standard deviation of the noise covered the full range of standard deviations observed experimentally. Altogether, a data set of five factors describing the errors that distinguished these ‘modified models’ from the original models was obtained. Each factor was assigned 10 different values resulting in 10 modified models from each original model.

Figure 3. Simulating model errors.

Figure 3

A, illustration of the error simulation procedure. The table gives the ranges from which error values were picked to modify the original models so they contained a set of errors that distinguished them from the original models (details given in Methods). These modified models were then given a set of predefined values for Rm, Cm, Ri and Racc, voltage and current responses were simulated, and the noise added (two topmost curves). The original models were then used to produce charging curves (two bottom curves) to fit the simulated responses by adjusting Rm, Cm, Ri and Racc until the sum of squared differences were at the minimum (similar to Fig. 1). B, the inverse of the Rm error (quantified as 1/Rm,fit/1/Rm,predef) and the Cm error (Cm,fit/Cm,predef) correlated well with the surface error. C, we normalized the 1/Rm and Cm errors by dividing by the surface errors and were then left with the error not accounted for by the surface error. The remaining errors (distance from the correct value of 1.0) increased with increasing noise both in CC and VC. The dotted line indicates the average noise level recorded during the real experiments.

The modified models were given predefined values, Rm,predef, Cm,predef, Ri,predef, Racc,predef, chosen to be compatible with our recordings and commonly used values in the literature: Ri = 100 Ω cm, Cm = 1.0 μF cm−2, Rm = 35 and 100 kΩ cm2 (TCs and INs, respectively) and Racc = individual cell value. We could then produce VC and CC charging curves, which we call ‘synthetic curves’.

Finally, we used our original models to fit these synthetic curves and obtained parameters (Rm,fit, Cm,fit, Ri,fit, and Racc,fit). These parameters, of course, were slightly different from the predefined values and these errors in the estimates were quantified by taking the ratio between the fit and predef values.

Non-uniform Rm

To simulate non-uniform Rm the passive membrane conductance per cm2 (Gx) at a given compartment x was computed as:

graphic file with name tjp0577-0205-m1.jpg

where Gsoma is the conductance at soma, dx is distance between soma and compartment x, dmax is distance between soma and most remote dendritic compartment, and k is a constant giving the conductance gradient.

The average Rm (Rm, avg, used to compare with Rm,predef) was then calculated as:

graphic file with name tjp0577-0205-m2.jpg

Active membrane properties

We made Rm voltage sensitive by including, in soma, Hodgkin–Huxley-type models of fast Na+ and K+ currents and, in all compartments, hyperpolarization-activated cationic (Ih) and low-threshold calcium (It) currents and a Ca2+ extrusion process, as described by Destexhe et al. (1996). These types of conductances were chosen because they have been characterized in TCs (Jahnsen & Llinas, 1984a,b; Suzuki & Rogawski, 1989; Coulter et al. 1989; Crunelli et al. 1989; Hernandez-Cruz & Pape, 1989; Pape & McCormick, 1989; McCormick & Pape, 1990; Huguenard & Prince, 1992) and INs (McCormick & Pape, 1988; Pape et al. 1994; Pape & McCormick, 1995; Williams et al. 1996; Zhu et al. 1999a,b).

The model description files for the above conductances were downloaded from http://cns.iaf.cnrsgif.fr/abstracts/Dlgn96.html (HH2.mod, It.mod, Ih.mod (iar), and cadecay.mod (cad)). Temperature was set to 24°C.

Before voltage or current steps were applied to the modified models, a period of 6 s was used to let the membrane potential stabilize. The depolarizing effect of Ih was compensated so that the somatic membrane potential was similar to that recorded in the real experiments. This was done by adjusting (with a manual, interactive procedure) the reversal potential of the passive leak conductance. Rm,avg at the end of the steps was then estimated as described above.

Simulations of synaptic conductance

We simulated the activation of one excitatory synapse with small peak conductance (0.5 nS, reversal potential set to 0 mV). The synaptic conductance was modelled as an alpha function:

graphic file with name tjp0577-0205-m3.jpg

where Gmax is the peak conductance and τ is the time constant of the conductance change. We monitored the voltage produced both locally and at soma. To study how synaptic position affected the somatic potential, the synapse was placed at 50 different dendritic locations with a uniform probability over the dendritic surface. Using uniform probability relative to dendritic length (rather than surface) did not influence the conclusions herein. In this series of simulations, we used Ri = 100 Ω cm, Cm = 1.0 μF cm−2 and the Rm estimates from the individual cells. The Rm values were taken from the VC fits because simulations described in the results showed that these were more reliable than the values obtained from the CC fits.

Statistics

Differences between the means were tested with a Student's t test unless stated otherwise. The correlations between the best-fit parameters (Fig. 2) were examined using Pearson's correlation coefficient (r).

Figure 2. Best-fitting parameters for all recorded neurons.

Figure 2

A, the Rin values for the 15 TCs (median with 10%–90% confidence limits, 170–370 MΩ) were clearly smaller than the Rin values of the three INs (0.5–1.1 GΩ)(P < 0.001, Mann–Whitney U test), both for CC and VC recordings (measurements at steady state). B, Rm values for TCs (20–54 kΩ cm2) were outside the range of the Rm values for INs (83–124 kΩ cm2), both in CC and VC modes (median 33 and 106 kΩ cm2 for TCs and INs, respectively, P < 0.001, Mann–Whitney U test). C, correlation between Rin and Rm in the combined TC and IN population (Pearson's r = 0.78, P < 0.001) and in the TC population alone (r = 0.55, P = 0.04), without correlation between Rin and area (r = −0.2, P = 0.51). D, TCs and INs had similar surface areas. E, both cell types had similar Cm both in CC and VC modes (median 1.2 μF cm−2 in both cases). F, negative correlation between Rm and Cm (r = −0.75, P = 0.001). G and H, the Ri was similar in the two cell populations with a median of 79 Ω cm with 10%–90% confidence limits of 22–424 Ω cm for the TCs and a median of 113 Ω cm for the INs (VC mode). Ri probably contributed very little to Rin which was almost completely accounted for by the ratio between Rm and area.

To quantify differences in linearity in the current–voltage (I–V) plots (Fig. 4), we calculated an R2 value from I–V data for all models. The R2 values were not used the conventional way (i.e. to test the hypothesis that there is no correlation between two variables). Instead, we used R2 only to quantify the deviation from a straight line and made our own estimates for how R2 would be distributed if the models were passive. This was done by estimating R2 for the I–V curves produced by the passive neuronal models with only experimentally recorded noise added. The resulting population of R2 values represented the expectation for linearity. Then we compared this expectation with the R2 values from non-linear models (and original data) to test whether the hypothesis of linearity could be rejected.

Figure 4. Rm and Cm errors due to active conductances.

Figure 4

A, I–V curves for the experimental data (seven TCs and one IN) were linear (visual assessment). B, the responses of a TC modified model with 0.2 mS cm−2 of Gt and Gh showing obvious active membrane behaviours (sag and overshoot). Yet, the corresponding I–V curves (C) were surprisingly linear, both in CC and VC (measurements at steady state). D, to quantify how much the data deviated from a straight line we compared R2 distributions (see Methods). The mean of the R2 distribution for the purely passive models (Passive) was compared to the means for the models with 0.2 mS cm−2 of Gt and Gh (0.4), models with 1 mS cm−2 of Gt and Gh (2.0) and the real experiments (Exp.). The real experiments were not significantly different from purely passive models (Passive) or models with 0.4 mS cm−2 total active conductance. E, the Rm and Cm errors increased with increasing Gt and Gh (the sum of them given at the x-axis), except for Cm in VC which had low error over the entire simulated range of active conductances. F, the sum of squared differences (SSq) between the best fitting charging curve and the charging curve of real or simulated experiments was a sensitive measure to detect the amount of Gt and Gh. The sum of Gt and Gh in 180 models (10 from each reconstructed neuron) is plotted along the x-axis. For comparison we plotted the SSq value from the real experiments (Exp.) and positioned it along the x-axis to fit in the rank of increasing SSq values.

Results

We recorded from 15 TCs and three INs (Fig. 1Aa and Bb, respectively), labelled them and used their morphology to construct compartmental models (see Methods). Somatic recordings were used to record their responses to current and voltage command steps in CC and VC modes, respectively. The steps were delivered when the somatic membrane potential was at rest (mean of −60.0 ± 0.9 mV for TCs and −59.8 ± 0.7 mV for INs).

Figure 1. Fitting model responses to experimental records.

Figure 1

Aa and Ba, biocytin-filled TC and IN with surface areas 19627 μm2 and 20266 μm2, respectively, were used to construct compartmental models. Axons are represented by dotted lines. Ab, the response recorded from the TC cell (continuous trace) was fitted by the model's CC response (grey trace) by minimizing the squared deviation. The difference between the two traces is shown on the top. The best fit was achieved with Rm = 59 kΩ cm2, Ri = 10 Ω cm, and Cm = 0.91 μF cm−2, giving Rin = 303 MΩ. Ac, similarly, the fit to a VC response was best with Rm = 61 kΩ cm2, Ri = 200 Ω cm, and Cm = 0.85 μF cm−2, giving Rin = 314 MΩ. Bb and c, fitting between the responses recorded from the IN (solid traces) and the model CC and VC response (grey traces). The best fits were achieved with: Rm = 78 and 107 kΩ cm2, Ri = 113 and 87 Ω cm, and Cm = 0.84 and 0.91 μF cm−2, giving Rin of 591 and 555 MΩ, for CC and VC responses, respectively. The vertical glitch on the residual is due to the fact that the first 3 ms were not fitted (see Methods).

The compartmental models were then used to produce responses which were fitted to the experimental responses. Optimal fits between the model-produced and experimental responses (grey and black traces in Fig. 1) were achieved by adjusting the free parameters Rm, Cm, Ri and Racc. In the examples presented in Fig. 1, the differences between the simulated and experimental responses (residuals, plotted above the charging curves) were small, without obvious trends.

Parameter values giving optimal fits

Whereas the two neurons in Fig. 1 had clearly different morphology, they had similar total membrane surface. This means that the differences in steady-state responses shown in Fig. 1 could not be attributed to differences in membrane surface, but rather to differences in Rm, Ri and/or diameters of dendrites.

The best-fitting values for Rm were similar in CC and VC modes, but about 50% less in the TC compared to the IN. This difference was similar to the difference in Rin between the two cells, suggesting that Rm was the main factor determining Rin.

Whereas Cm was similar for the two recording modes and for the two cell types, the estimate for Ri differed considerably in CC and VC modes and between the two cells. Whether this indicates that Ri was not reliably estimated in one or both modes, or that incorrect model assumptions had a different impact in the two recording modes, was tested with simulations (next section).

The differences and similarities between the TCs and INs shown in Fig. 1 were supported by the results from all 15 TCs and three INs (Fig. 2). Figure 2 present the data from both CC and VC fits, given as medians with 10%–90% confidence intervals for the TCs, and the individual values for the three INs.

In TCs, both Rin and Rm values were distributed over a narrow range (Fig. 2A and B). The Rm values for INs were clearly outside this range. In agreement with the idea that the Rin difference between the two cell groups was mainly due to a difference in Rm, we found a clear correlation between Rin and Rm (Fig. 2C). We found no correlation between Rin and membrane surface areas (not shown). This was expected because membrane surface was similar in TCs and INs (Fig. 2D). It is interesting that Rm may also be the most important factor for the differences in Rin within the TC group, because Rin and Rm were correlated within this group, but Rin and membrane area were not.

TCs and INs had similar values for Cm (Fig. 2E). However, some TCs had Cm values > 1.5 μF cm−2, which is unlikely for neuronal membranes (Gentet et al. 2000). Cm and Rm also correlated negatively (Fig. 2F), suggesting that the large Cm values may be due to reconstruction errors (Roth & Häusser, 2001). This possibility was tested in our simulation experiments (next section).

The Ri estimates for both TCs and INs overlapped (Fig. 2G). In TCs, however, the range was large, supporting the suggestion that Ri can not be accurately determined in the TCs because of their short dendrites (Briska et al. 2003). In both cell types, however, Ri probably added very little to the somatic Rin because dividing Rm by the surface area produced resistance values very close to the Rin values with only small deviations from the line of identity (Fig. 2H). Such tight linear relationship between Rin and Rm/area suggests that the cells were electrotonically compact at steady state.

The parameter values that gave optimal fits in this study are similar to those reported by Briska et al. (2003) in the adult rat. Moreover, our larger sample allows us to clearly show a separate distribution of Rm values between the two cell types and a very narrow distribution of Rin and Rm in the TC population.

Simulations to evaluate potential sources of errors in Rm, Cm and Ri estimates

Because models are only approximate representations of real neurons, we wanted to test the reliability of our parameter estimates. To do this we used our original models to perform a series of parameter fits to voltage and current responses that were produced by models that were slightly modified versions of the original models (‘modified models’ in Fig. 3A and see Methods for details).

Five types of errors were examined by including in the modified models: (1) inaccuracies in the length and (2) diameter of the dendrites; (3) inaccuracies due to the experimental noise; and (4) violations of the assumptions of linear and time-independent conductances with (5) uniform distribution. For each type of error, the ranges of values explored are indicated in the table at the top of Fig. 3A (for details see Methods). We describe the results of the simulations first for Rm and Cm and then for Ri separately.

Errors of Rm and Cm estimates due to surface errors and experimental noise

All error types were included in the simulations to allow interactions between the different model errors. Initially, we found only weak correlations between model errors and the errors of the estimated parameters. However, by multiplying the errors of the dendritic length and diameter to obtain a surface error, we found strong and almost linear correlations between the surface error and the errors of Rm and Cm estimates (Fig. 3B). Based on these correlations, we could estimate that the 3.3% larger diameter of TC we found by comparing measurements in living and mounted slices (see Methods) would give only a 3.3% underestimation of 1/Rm and Cm.

In Fig. 3B, the sensitivity of the parameter estimates to the remaining three model errors (noise, active and non-uniform conductances) can be assessed from the variability around the identity lines. This variability was clearly smaller in VC than CC, showing that the Rm and Cm estimates were less sensitive to these three model errors in VC compared to CC.

To find which of those three model errors contributed the most to errors in Rm and Cm estimates, we adjusted the Rm and Cm errors by normalizing them by the surface errors. In Fig. 3C we plotted these normalized Rm and Cm estimates as a function of the s.d. of the noise. The results show that the errors of the surface normalized Rm and Cm estimates (i.e. the spread around the 1.0 value) were small at low noise levels despite the presence of active and non-uniform conductances. This means that, over the parameter ranges explored, the Cm and Rm errors were considerably more influenced by the noise level than by active and non-uniform conductances. At the average noise levels that were recorded experimentally (vertical dotted line), the Rm and Cm errors were substantial in CC but only in the order of 5% in VC. Increasing the number of averaged traces would of course reduce the noise contribution to the parameter errors also in CC (see, e.g. Roth & Häusser, 2001)). However, with the same number of traces in the averages, VC showed smaller parameter errors.

Errors of the Rm and Cm estimates due to active conductances

A major concern when passive neuronal models are used for parameter estimation is that the charging curves from the real neurons are shaped by voltage- and time-dependent conductances. We studied the I–V response of our neurons in the voltage range where we applied the test pulses and, under simple visual inspection, these appeared linear and symmetrical around the resting potential (Fig. 4A). However, when a modified model with voltage- and time-dependent conductances (low threshold calcium conductance (Gt), 0.2 mS cm−2; hyperpolarization activated conductance (Gh), 0.2 mS cm−2; see Methods) was used to simulate similar voltage and current responses (Fig. 4B), we also obtained rather linear I–V curves (Fig. 4C).

We quantified deviations from perfect linearity using R2 (see Methods), which gives the value 1.0 at perfect linearity. We compared the distributions of the R2 values for the I–V curves produced by the passive models to which only experimental noise was added (‘passive’) to the R2 values of the I–V curves produced by the modified models and the R2 values of the experimental I–V curves (‘exp.’ in Fig. 4D). The means and s.d. for the various R2 distributions are displayed in Fig. 4D. As shown, the R2 obtained using the modified models with active conductances (Gt, 0.2 mS cm−2 and Gh, 0.2 mS cm−2) was not significantly different from the passive models, whose R2 value was a little less than 1.0 due to the experimental noise. In fact, we had to increase the active conductances to 2.0 mS cm−2 (G.t, 1.0 mS cm−2 and G.h, 1.0 mS cm−2) get an R2 value that was significantly different from the passive models. This was true only for CC mode and in VC mode the R2 value was not significantly different from the passive models even at this conductance level.

Although it was possible to find other combinations of Gt and Gh, and other holding potentials, which made the I–V plots less linear, we want to emphasize that the I–V plots could often be close to linear even in the presence of some active conductance. Therefore, an important question is whether the amounts of active conductances not detectable on I–V plots could give a serious error of Rm and Cm estimates.

In order to explore this question we tested a large range of Gt and Gh. Figure 4E shows the Rm and Cm errors as a function of the sum of Gt and Gh. The Cm errors were very little influenced by the active conductances in VC. In CC, however, the errors were larger, as indicated by the increasing deviation from 1.0 with increasing amount of active conductances. The Rm errors were generally larger than Cm errors. With 0.4 mS cm−2 active conductance (amount where the I–V plots were not significantly non-linear, Fig. 4D), many Rm errors were > 10%.

Obviously, some of the CC charging curves in Fig. 4B cannot be well fitted with a purely passive model. Thus, in addition to the linearity of the I–V plots, we examined the fitting error (sum of squared deviations; SSQ) at different levels of Gt and Gh. The SSQ should be a more sensitive statistic because the shape of the charging curve is evaluated, not only the steady-state level. For all the neurons, we simulated charging curves similar to those shown in Fig. 4B. These synthetic curves were then fitted with our original models. An SSQ was obtained, which was normalized to the SSQ from the fit between the responses from the original models and the passive models with only experimental noise added. If the original models were correct, the latter SSQ should represent the best possible SSQ and give an expectation of 1.0 (indicated by a dotted line in Fig. 4F).

As shown in Fig. 4F, the relative SSQ errors increased with increasing active conductances (sum of Gt and Gh), and were significantly different from the passive models when the sum of active conductances was > 0.05 mS cm−2. Also plotted for comparison, the average SSQ from the fit to the experimental data (arrow) was not significantly different from the value of 1.0. These results suggest that, with the type of currents included in this study, the SSQ error (fitting error) is more sensitive to active conductances than the linearity of the I–V plot. Furthermore, with these data we can estimate that at the conductance level of 0.05 mS cm−2, the errors of Cm and Rm was < 5% (Fig. 4E, VC data). Thus, the errors in Rm and Cm estimates introduced by active conductances were small and not larger than the errors introduced by noise and dimensional errors.

Errors of the Ri estimate

The Ri estimate is very sensitive to errors in the determination of Racc. We found that the commonly used approach to estimate Racc (dividing the step command voltage by the initial peak current in VC) systematically overestimated Racc by 4%–6% because of an underestimation of the peak current. This relatively small inaccuracy in Racc measurement resulted in a large (30%–80%) underestimation of Ri. Therefore, we included Racc as a free variable (together with Rm, Cm and Ri) in our models. As described below, by doing so we were able to obtain reliable Ri estimates for the INs.

In simulations with all model errors (as in Fig. 3), Ri errors correlated positively with length error and negatively with the error of the cross-sectional area (squared diameter) of the dendrites. When the Ri estimation error was displayed as a function of the summed errors in dendritic length and cross-sectional area (x-axis in Fig. 5), a clear correlation could be seen for the INs but not for the TCs. The fact that the Ri cannot be accurately estimated for TCs is attributed to their shorter dendrites and their lower total axial resistance.

Figure 5. Ri was sensitive to errors in model dimensions.

Figure 5

The error of the Ri estimate in INs (•) correlated well with errors of the reconstructed morphology, even with all the other model errors (see Fig. 3) present. For TCs, the Ri estimate was poorly constrained (○). The Ri error in INs increased almost linearly with the length error and inverse of the squared diameter error (i.e. the error of the inverse cross-sectional area of the dendrites).

The present simulations therefore suggest that for neurons with long dendrites such as INs, Ri can be determined with good accuracy, over a large range of noise levels and active and non-uniform conductances, as long as the morphology of the neuron is accurately reconstructed. Furthermore, using the results displayed in Fig. 5, we can estimate directly the errors of Ri estimates at least for INs. With dimensional errors of about 5%, the error on Ri would be < 10%.

Converting synaptic conductances into somatic voltages; role of Rm and morphology

The purpose of finding accurate parameter estimates was to investigate how short-lasting synaptic conductance changes are converted to somatic voltages. In this context, we asked one simple question: what would be the somatic voltage response to a single excitatory synapse? We measured the excitatory postsynaptic potentials (EPSPs) resulting from activation of one fast excitatory synaptic conductance (0.5 nS, see Methods) and repositioned the synapse at 50 different, randomly selected positions in the dendritic tree. Our previous simulations showed that the passive parameters were always more accurate in VC than in CC mode. Therefore, for the present series of simulations each cell was given their respective VC passive parameters.

Peak somatic EPSP

The amplitudes of the EPSPs were, of course, much larger at the synaptic site (Fig. 6A) than at the soma (Fig. 6B). We will focus on the somatic EPSP (EPSPsoma) because of its direct role in bringing the membrane closer to firing threshold. For each cell, the EPSPsoma from the 50 dendritic locations were first rank-ordered with respect to increasing amplitude. For each rank, the values were averaged and a population mean was obtained for TCs and INs separately. In Fig. 6CF, the population means are shown as open circles for TCs and filled circles for INs.

Figure 6. Simulating dendritic synaptic conductances.

Figure 6

Fifty randomly chosen locations (○) along the dendrites of TCs and INs were tested with a 0.5 nS excitatory synapse (see Methods). As expected, the EPSPs at the synapse (A) were larger than the EPSPs measured at soma (B). Note, however, that the peak amplitudes of the EPSPsoma were relatively similar in the INs and the TCs. C, for each cell, the EPSPsoma from the 50 dendritic locations for the synapse were rank-ordered with respect to the peak amplitude. For each rank, a population average for all INs (•) and TCs (○) was obtained. The vertical bars give standard deviations. The right plot, which is the same as the left plot but with one-quarter of the membrane resistance, shows that the median peak amplitude is not very sensitive to membrane resistance in the tested interval. D, same as C, for the integral of EPSPsoma. The integral was larger for almost all dendritic positions in INs compared to TCs. The INs displayed larger EPSPsoma than the TCs even after a large reduction in Rm (right plot). E, for the cells illustrated in A, the EPSPsoma induced by a 20 nS synaptic conductance. F, at this conductance level, EPSPsoma was smaller in the INs than in the TCs at almost all dendritic locations. G, with further increase in the conductance of the synapse (data connected by solid lines), the median EPSPsoma still increased much less in INs (•) than in TCs (○). However, when the increases in total conductance were obtained by simultaneous activation of many small (0.5 nS), spatially distributed synapses, the EPSPsoma reached much larger amplitudes in both cell types (data connected by dashed lines).

In each TC, EPSPsoma from different dendritic locations had very similar peak amplitude with a coefficient of variation (CV) of only 12% on average (see also Neubig & Destexhe, 2001; Briska et al. 2003). Furthermore, the different TCs were also very similar to each other; the mean EPSPsoma varied only 24% between the different TCs. Because the INs have much longer dendrites than the TCs they have been thought to filter their EPSPs much more than the TCs. We found, however, only moderate differences between the TCs and INs, with an average EPSPsoma in INs only 12% smaller than the average EPSPsoma in TCs.

For the INs, the peak amplitude of the EPSPsoma from different dendritic locations varied with a CV of 36% on average. This may indicate that dendritic location is slightly more important for the EPSPsoma in INs than in TCs. However, the largest contribution to the variation in INs was from a few very large EPSPs, probably from synapses close to the soma (largest peak values in left panel of Fig. 6C). When the synapses located farthest away from the soma (10% smallest EPSPsoma) were considered, then the effect of dendritic location was similar in both cell types: the 10% smallest EPSPsoma were 85% and 69% of the median EPSPsoma for TCs and INs, respectively.

Finally, because Rm is the most flexible passive parameter and may vary with changes in, for example, temperature and network activity, we wanted to test how sensitive the peak amplitude of EPSPsoma was to changes in Rm. We tested a 75% reduction in Rm. As shown in Fig. 6C (right panel), such a reduction decreased the peak EPSPsoma by only 24% in TCs and 20% in INs. This result confirms that the peak amplitude of the EPSPsoma is relatively insensitive to changes in Rm (Spruston et al. 1993; Jaffe & Carnevale, 1999; Briska et al. 2003) suggesting that at a given synaptic conductance, the peak amplitude of the EPSPsoma in LGN cells is mainly determined by their cell geometry.

Integral of EPSPsoma

Because several EPSPs would be needed to depolarize the soma above the threshold for action potential, depolarization over time may be a more relevant measure than the peak amplitude for assessing the contribution of an EPSP to soma depolarization. Consequently we also measured the integral of the EPSPsoma.

We found that similar to the peak EPSPsoma, the integral (Fig. 6D, left panel) varied little in the individual TC (on average 11%) or between the TCs (26%). The CV of the EPSPsoma in the individual IN was a little larger than for the TCs with a value of 21%. For the INs, the integral measurements made it even more obvious than the peak amplitude measurements that the long dendrites of the INs did not prevent synaptic inputs from all dendritic locations from influencing somatic voltage. Indeed, the EPSPsoma in INs were typically larger than in TCs (compare the two curves in left panel of Fig. 6D). The 10% smallest EPSPsoma were 86% and 77% of the median EPSPsoma for TCs and INs, respectively.

As expected, the integrals of the EPSPsoma were more sensitive to changes in Rm than the peak amplitude. A decrease of 75% in Rm (Fig. 6D, right panel) reduced the integrals by 73% and 68% in TCs and INs, respectively. However, the difference between the median and the 10% smallest EPSPsoma remained small (the 10% smallest EPSPssoma were 85% and 74% of the median for TCs and INs, respectively).

Voltage saturation at synaptic sites

In the simulations above we used a synaptic conductance of 0.5 nS because activation of such a small conductance resulted in a moderate reduction of the driving potential at the synaptic site. However, because of a lack of precise knowledge about the size of the unitary synaptic conductances and to account for possibility of several simultaneously activated synapses on the same dendrite, we also tested a range of synaptic conductances.

When the single synaptic conductances were increased by 100%, the peak EPSPsoma increased on average only by 82% and 70% in TCs and INs, respectively (not shown). This sublinearity, attributed to the reduction of the voltage driving force, was more pronounced for INs than for TCs. For example, with a synaptic conductance of 20 nS, most EPSPsoma were smaller in the IN than in the TC (Fig. 6E and F).

With even larger synaptic conductances (stepwise increase up to 100 nS in Fig. 6G), the loss of voltage driving force clearly put an upper limit to how large an EPSPsoma could be generated from a small dendritic segment; around 8 and 2.5 mV in TCs and INs, respectively (Fig. 6G data points connected with solid lines). This indicates that the INs were more susceptible to saturation.

Because the most remote synapses saturated at smaller conductances than the more proximal, the 10% smallest peak EPSPsoma were only 30% (TCs) and 19% (INs) of the median EPSPsoma with 100 nS conductance. Similar results (with slightly smaller differences between TCs and INs) were obtained when we measured the integral of the EPSPsoma rather than the peak (not shown).

For comparison we tested again the effect of a large increase in the total synaptic conductance but this time by simultaneous activation of many, small (0.5 nS) spatially distributed conductances. This resulted in a much more linear relationship between the total synaptic conductance and the peak EPSPsoma (Fig. 6G data connected with dashed lines), and in much larger EPSPsoma in both cell types. In fact, by spatially distributing many small synaptic conductances, the INs converted the synaptic signals to soma potentials almost as efficiently as the TCs. Our data also indicate that the INs were more sensitive to the spatial distribution of synaptic conductances than the TCs. With 100 nS total synaptic conductance, the EPSPsoma produced by the distributed synaptic conductances was six times larger than the single-site EPSPsoma for the TCs and 13 times larger for the INs.

Active dendrites

One problem in the study of integration properties is that we do not know in detail the amount and distribution of active conductances in the dendrites. When we tested large synaptic conductances (> 2.0 nS), or several synapses, together with a combination of 0.2 mS cm−2 for Gh and Gt, we obtained results that differed in several ways from the passive simulation described above. More importantly, we found that the results were very sensitive to the exact combination of these factors (not shown).

However, as long as Gh and Gt were < 0.2 mS cm−2 (i.e. below the values found to be incompatible with our charging curves, see Fig. 4F) and the synapses were small (1.0 nS), our main conclusions were not influenced by the active conductances. With 0.2 mS cm−2Gh and 0.2 mS cm−2Gt, the different synaptic locations gave very similar EPSPsoma, as for the passive models. The integral was similar in the individual TC (CV of 10% on average) and between the TCs (CV of 24%). For the INs, the CVs were slightly larger than for the TCs, with an average value of 20%. In both cell types, the smallest 10% of the peak and the integral EPSPsoma were > 72% of the median EPSPsoma.

Discussion

Conversion of small synaptic conductances to somatic voltages

Our study suggests that despite large differences in dendritic morphology, the TCs and INs have similar ability to convert small, short-lasting synaptic conductance changes to somatic voltages. This conclusion is different from other investigations that have emphasized the more local signal processing that may occur in the INs due to their long dendrites (Bloomfield & Sherman, 1989; Sherman, 2004). Part of the explanation may be attributed to differences in methodology and the use of tight-seal patch electrodes rather than penetrating electrodes. With tight-seal patch electrodes, higher Rm values are obtained (present study and Briska et al. 2003) and this may compensate at least partly for the attenuating effect of long dendrites. An additional explanation is that earlier investigations have often focused on voltage attenuation, usually defined as the ratio between somatic and dendritic voltage, whereas we focused on the absolute amplitude of the somatic potentials. We made this choice because the influence on somatic voltage is easy to interpret. The exact voltage at the synapse may be important, particularly if it reaches the equilibrium potential for the synapse or, as in the case of INs, can trigger transmitter release by the dendrites. The ratio between dendritic and somatic voltage, however, does not have any known physiological effect.

One main finding was that the integrals of the smallest EPSPsoma produced by single, small synaptic conductances were larger in the INs than in the TCs (Fig. 6D), illustrating clearly that long dendritic cables, like in the INs, do not prevent distal synapses from influencing the somatic voltage. Because INs have larger dendritic fields than TCs (Perreault et al. 2003), the fact that even distal synaptic inputs can significantly depolarize the soma suggests that INs have larger integration fields.

We also provide new data that suggest that TCs are a particularly homogeneous population with respect to integration properties. We found that the ability to convert synaptic conductances to soma potentials was remarkably similar among TCs, possibly giving all synapses on all TCs an equal weight. For the integral of the EPSPsoma, the low variation fits well with the previously described low variability in Rin (Zhu & Heggelund, 2001; Perreault et al. 2003). This is not so surprising because Rin is the main determinant of the EPSP integral. However, the variation in the peak amplitude between the different TCs was also small (24%) despite the fact that this measure depends mostly on dendritic diameter, branching pattern and Ri (i.e. factors that contribute little to Rin).

Interference with active membrane properties

It is important that our conclusions are limited to small, short-lasting synaptic conductance changes, because with larger and longer conductances, the results become very sensitive to the level of voltage- and time-dependent dendritic conductances. The knowledge about such dendritic properties is not sufficiently accurate to make model-based predictions of the voltage trajectories close to action potential threshold.

However, with small synaptic conductances our conclusions were almost unaffected by Ih and It as long as these were within levels compatible with the shape of our experimental charging curves. The main reason is probably that the synaptic voltage transients in the dendrites were too fast and local to influence Ih and It sufficiently to significantly change EPSPsoma. The linear behaviour of the membrane in response to single synaptic inputs is supported by results from other cell types (Langmoen & Andersen, 1983; Gasparini & Magee, 2006).

Accuracy of the models

The conclusions in this study depend on the accuracy of the models. Therefore an important part of this article was to explore the influence of different model errors on the accuracy of the parameter estimates. In addition to the finding that VC mode gave more accurate parameter estimates, several useful results came out of these simulations. These are discussed below.

Rm

Using morphology-based models together with tight-seal recording techniques, Briska et al. (2003) have recently suggested Rm values around 30 kΩ cm2 for two TCs and 90 kΩ cm2 for two INs. These values are considerably higher than earlier estimates based on more simplified morphologies (Bloomfield et al. 1987; Bloomfield & Sherman, 1989). In the present study, Rm was relatively similar within the population of 15 TCs, with 80% of the cells lying between 20 and 53 kΩ cm2 (compatible with Briska et al. (2003)). By providing a range of Rm values within the TC population, we show that the INs had Rm values clearly outside this range.

Our biggest concern when estimating Rm was that our experimental curves were contaminated by active currents even when our I–V plots were close to linear. We found, however, that a passive model can capture Rm even in neurons with moderate amounts of active currents.

Our upper level of Gh and Gt that were compatible with the original charging curves (0.025 mS cm−2 of Gh and 0.025 mS cm−2 of Gt) were multiplied with the cell surfaces, obtaining 3.8 nS for INs and 3.2 nS for TCs, for both Gh and Gt. These values are compatible with the values from other reports (Gh, 1 nS for INs and 3–15 nS for TC; Gt, 8.4 nS for both cell types; Crunelli et al. 1989; Zhu et al. 1999a,b). So the range of the conductances we tested are relevant and would not severely bias our Rm estimates.

Cm

Our median value for Cm was 1.2 μF cm−2 and our simulations indicated that this value was probably correct within 20% of the true value (Fig. 3C, VC data). This fits well with Cm values close to 1.0 μF cm−2 from measurements in cells where surface area can be reliably estimated (Curtis & Cole, 1938; Hodgkin & Huxley, 1952; Gentet et al. 2000).

We found that the most important error in the estimation of this parameter was from error in the measurement of the membrane area. The relatively large range of Cm values for the TCs (Fig. 2G), may suggest that the surface area was inaccurate for some cells. Such a hypothesis would also fit with the negative correlation we found between Cm and Rm (Fig. 2H). However, as reported in the Methods, shrinkage did not significantly change the cell dimensions. An alternative possibility is that those few cells that display high Cm might have been connected to other TCs via gap junctions (Parenti et al. 2000; Hughes et al. 2002, 2004).

Ri

The only passive parameter whose accuracy was obviously cell-type dependent was Ri, which was more accurate for INs than for TCs. The Ri estimate in INs was influenced by inaccuracies in dendritic diameter and length. While this was expected from cable theory, we did not expect this influence to be so obvious when the active currents, non-uniform Rm and noise were present at the same time (Fig. 5). We thus conclude that the latter three factors introduced less error than morphometric inaccuracies, at least in the ranges explored in this study.

Is there an important role for dendritic voltage saturation?

INs and TCs differed in their ability to convert single large synaptic conductances to somatic voltage (see Fig. 6G). In the INs, no single synapse was able to depolarize the soma by more than ∼2.5 mV (on average in individual cells). In contrast, in the TCs, the upper limit for how large an EPSPsoma could be generated from a single synapse was much higher (∼8.0 mV).

It has been suggested that the transmission of synaptic signals from dendrites to soma is more limited by passive cable properties in INs than in TCs (Briska et al. 2003; Sherman, 2004). We suggest that this is the case only for particularly large synaptic conductances. Furthermore, we do not think that this is due to voltage attenuation but to voltage saturation in the IN dendrites which more easily approach the reversal potential for the synapses. Dendritic voltage saturation has been suggested to be important for the input–output function of other cell types as well (Cushing et al. 2005; Golding et al. 2005).

In theory, dendritic voltage saturation could be investigated experimentally with dendritic recordings; however, due to the low probability of encountering INs in the LGN this has proven to be difficult. Hopefully, the use of new transgenic animals where glutamic acid decarboxylase isoform 67 (GAD67)-expressing INs are selectively labelled (Tamamaki et al. 2003) will make this possible.

Acknowledgments

The present experiments were performed in the laboratory of Dr Paul Heggelund and were supported by grants from the Norwegian Research Council. We would like to thank Drs Per Andersen, Joel Glover and Paul Heggelund for their helpful comments on an earlier version of the manuscript.

References

  1. Ambros-Ingerson J, Holmes WR. Analysis and comparison of morphological reconstructions of hippocampal field CA1 pyramidal cells. Hippocampus. 2005;15:302–315. doi: 10.1002/hipo.20051. [DOI] [PubMed] [Google Scholar]
  2. Barrett JN, Crill WE. Specific membrane properties of cat motoneurones. J Physiol. 1974;239:301–324. doi: 10.1113/jphysiol.1974.sp010570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Bloomfield SA, Hamos JE, Sherman SM. Passive cable properties and morphological correlates of neurones in the lateral geniculate nucleus of the cat. J Physiol. 1987;383:653–692. doi: 10.1113/jphysiol.1987.sp016435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Bloomfield SA, Sherman SM. Dendritic current flow in relay cells and interneurons of the cat's lateral geniculate nucleus. Proc Natl Acad Sci U S A. 1989;86:3911–3914. doi: 10.1073/pnas.86.10.3911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Briska AM, Uhlrich DJ, Lytton WW. Computer model of passive signal integration based on whole-cell in vitro studies of rat lateral geniculate nucleus. Eur J Neurosci. 2003;17:1531–1541. doi: 10.1046/j.1460-9568.2003.02579.x. [DOI] [PubMed] [Google Scholar]
  6. Clements JD, Redman SJ. Cable properties of cat spinal motoneurones measured by combining voltage clamp, current clamp and intracellular staining. J Physiol. 1989;409:63–87. doi: 10.1113/jphysiol.1989.sp017485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Coulter DA, Huguenard JR, Prince DA. Calcium currents in rat thalamocortical relay neurones: kinetic properties of the transient, low-threshold current. J Physiol. 1989;41:587–604. doi: 10.1113/jphysiol.1989.sp017705. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Crunelli V, Lightowler S, Pollard CE. A T-type Ca2+ current underlies low-threshold Ca2+ potentials in cells of the cat and rat lateral geniculate nucleus. J Physiol. 1989;413:543–561. doi: 10.1113/jphysiol.1989.sp017668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Curtis HJ, Cole KS. Transverse electrical impedance of the squid giant axon. J Gen Physiol. 1938;21:757–765. doi: 10.1085/jgp.21.6.757. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Cushing S, Bui T, Rose PK. Effect of nonlinear summation of synaptic currents on the input-output properties of spinal motoneurons. J Neurophysiol. 2005;94:3465–3478. doi: 10.1152/jn.00439.2005. [DOI] [PubMed] [Google Scholar]
  11. Destexhe A, Bal T, McCormick DA, Sejnowski TJ. Ionic mechanisms underlying synchronized oscillations and propagating waves in a model of ferret thalamic slices. J Neurophysiol. 1996;76:2049–2070. doi: 10.1152/jn.1996.76.3.2049. [DOI] [PubMed] [Google Scholar]
  12. Gasparini S, Magee JC. State-dependent dendritic computation in hippocampal CA1 pyramidal neurons. J Neurosci. 2006;26:2088–2100. doi: 10.1523/JNEUROSCI.4428-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Gentet LJ, Stuart GJ, Clements JD. Direct measurement of specific membrane capacitance in neurons. Biophys J. 2000;79:314–320. doi: 10.1016/S0006-3495(00)76293-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Golding NL, Mickus TJ, Katz Y, Kath WL, Spruston N. Factors mediating powerful voltage attenuation along CA1 pyramidal neuron dendrites. J Physiol. 2005;568:69–82. doi: 10.1113/jphysiol.2005.086793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Hernandez-Cruz A, Pape HC. Identification of two calcium currents in acutely dissociated neurons from the rat lateral geniculate nucleus. J Neurophysiol. 1989;61:1270–1283. doi: 10.1152/jn.1989.61.6.1270. [DOI] [PubMed] [Google Scholar]
  16. Hines M. A program for simulation of nerve equations with branching geometries. Int J Biomed Comput. 1989;24:55–68. doi: 10.1016/0020-7101(89)90007-x. [DOI] [PubMed] [Google Scholar]
  17. Hines ML, Carnevale NT. The NEURON simulation environment. Neural Comput. 1997;9:1179–1209. doi: 10.1162/neco.1997.9.6.1179. [DOI] [PubMed] [Google Scholar]
  18. Hodgkin AL, Huxley AF. A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol. 1952;117:500–544. doi: 10.1113/jphysiol.1952.sp004764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Holmes WR, Ambros-Ingerson J, Grover LM. Fitting experimental data to models that use morphological data from public databases. J Comput Neurosci. 2006;20:349–365. doi: 10.1007/s10827-006-7189-8. [DOI] [PubMed] [Google Scholar]
  20. Horikawa K, Armstrong WE. A versatile means of intracellular labeling: injection of biocytin and its detection with avidin conjugates. J Neurosci Methods. 1988;25:1–11. doi: 10.1016/0165-0270(88)90114-8. [DOI] [PubMed] [Google Scholar]
  21. Hughes SW, Blethyn KL, Cope DW, Crunelli V. Properties and origin of spikelets in thalamocortical neurons in vitro. Neuroscience. 2002;110:395–401. doi: 10.1016/s0306-4522(01)00577-2. [DOI] [PubMed] [Google Scholar]
  22. Hughes SW, Lörincz M, Cope DW, Blethyn KL, Kékesi KA, Parri HR, Juhász G, Crunelli V. Synchronized oscillations at alpha and tetha frequencies in the lateral geniculate nucleus. Neuron. 2004;42:253–268. doi: 10.1016/s0896-6273(04)00191-6. [DOI] [PubMed] [Google Scholar]
  23. Huguenard JR, Prince DA. A novel T-type current underlies prolonged Ca2+-dependent burst firing in GABAergic neurons of rat thalamic reticular nucleus. J Neurosci. 1992;12:3804–3817. doi: 10.1523/JNEUROSCI.12-10-03804.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Jackson MB. Cable analysis with the whole-cell patch clamp -theory and experiment. Biophys J. 1992;61:756–766. doi: 10.1016/S0006-3495(92)81880-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Jaffe DB, Carnevale NT. Passive normalization of synaptic integration influenced by dendritic architecture. J Neurophysiol. 1999;82:3268–3285. doi: 10.1152/jn.1999.82.6.3268. [DOI] [PubMed] [Google Scholar]
  26. Jahnsen H, Llinas R. Electrophysiological properties of guinea-pig thalamic neurones: an in vitro study. J Physiol. 1984a;349:205–226. doi: 10.1113/jphysiol.1984.sp015153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Jahnsen H, Llinas R. Ionic basis for the electro-responsiveness and oscillatory properties of guinea-pig thalamic neurones in vitro. J Physiol. 1984b;349:227–247. doi: 10.1113/jphysiol.1984.sp015154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Langmoen IA, Andersen P. Summation of excitatory postsynaptic potentials in hippocampal pyramidal cells. J Neurophysiol. 1983;50:1320–1329. doi: 10.1152/jn.1983.50.6.1320. [DOI] [PubMed] [Google Scholar]
  29. London M, Meunier C, Segev I. Signal transfer in passive dendrites with nonuniform membrane conductance. J Neurosci. 1999;19:8219–8233. doi: 10.1523/JNEUROSCI.19-19-08219.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. McCormick DA, Pape HC. Acetylcholine inhibits identified interneurons in the cat lateral geniculate nucleus. Nature. 1988;334:246–248. doi: 10.1038/334246a0. [DOI] [PubMed] [Google Scholar]
  31. McCormick DA, Pape HC. Properties of a hyperpolarization-activated cation current and its role in rhythmic oscillation in thalamic relay neurones. J Physiol. 1990;431:291–318. doi: 10.1113/jphysiol.1990.sp018331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Major G, Larkman AU, Jonas P, Sakmann B, Jack JJ. Detailed passive cable models of whole-cell recorded CA3 pyramidal neurons in rat hippocampal slices. J Neurosci. 1994;14:4613–4638. doi: 10.1523/JNEUROSCI.14-08-04613.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Neubig M, Destexhe A. Dendritic organization in thalamocortical neurons and state-dependent functions of inhibitory synaptic inputs. Thalamus Relat Syst. 2001;1:3952. [Google Scholar]
  34. Pape HC, Budde T, Mager R, Kisvarday ZF. Prevention of Ca2+-mediated action potentials in GABAergic local circuit neurones of rat thalamus by a transient K+ current. J Physiol. 1994;478:403–422. doi: 10.1113/jphysiol.1994.sp020261. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Pape HC, McCormick DA. Noradrenaline and serotonin selectively modulate thalamic burst firing by enhancing a hyperpolarization-activated cation current. Nature. 1989;340:715–718. doi: 10.1038/340715a0. [DOI] [PubMed] [Google Scholar]
  36. Pape HC, McCormick DA. Electrophysiological and pharmacological properties of interneurons in the cat dorsal lateral geniculate nucleus. Neuroscience. 1995;68:1105–1125. doi: 10.1016/0306-4522(95)00205-w. [DOI] [PubMed] [Google Scholar]
  37. Parenti R, Gulisano M, Zappala A, Cicirata F. Expression of connexin36 mRNA in adult rodent brain. Neuroreport. 2000;11:1497–1502. [PubMed] [Google Scholar]
  38. Perreault M-C, Qin Y, Heggelund P, Zhu JJ. Postnatal development of GABAergic signalling in the rat lateral geniculate nucleus: presynaptic dendritic mechanisms. J Physiol. 2003;546:137–148. doi: 10.1113/jphysiol.2002.030643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Perreault M-C, Raastad M, Heggelund P. Postnatal changes in the electrotonic size of the interneurons of the rat lateral geniculate nucleus. Soc Neurosci Abstract. 2000:447.11. [Google Scholar]
  40. Press WH, Flannery BP, Teukolsky SA, Vetterling WT. The Art of Scientific Computing. Cambridge: Cambridge University Press; 1986. Numerical recipes. [Google Scholar]
  41. Rall W, Burke RE, Holmes WR, Jack JJ, Redman SJ, Segev I. Matching dendritic neuron models to experimental data. Physiol Rev. 1992;72:159–186. doi: 10.1152/physrev.1992.72.suppl_4.S159. [DOI] [PubMed] [Google Scholar]
  42. Rocha M, Sur M. Rapid acquisition of dendritic spines by visual thalamic neurons after blockade of N-methyl-D-aspartate receptors. Proc Natl Acad Sci U S A. 1995;92:8026–8030. doi: 10.1073/pnas.92.17.8026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Roth A, Häusser M. Compartmental models of rat cerebellar Purkinje cells based on simultaneous somatic and dendritic patch-clamp recordings. J Physiol. 2001;535:445–472. doi: 10.1111/j.1469-7793.2001.00445.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Sherman SM. Interneurons and triadic circuitry of the thalamus. Trends Neurosci. 2004;27:670–675. doi: 10.1016/j.tins.2004.08.003. [DOI] [PubMed] [Google Scholar]
  45. Spruston N, Jaffe DB, Williams SH, Johnston D. Voltage- and space-clamp errors associated with the measurement of electrotonically remote synaptic events. J Neurophysiol. 1993;70:781–802. doi: 10.1152/jn.1993.70.2.781. [DOI] [PubMed] [Google Scholar]
  46. Steuber V, De Schutter E, Jaeger D. Passive models of neurons in the deep cerebellar nuclei: the effect of reconstruction errors. Neurocomputing. 2004;58–60:563–568. [Google Scholar]
  47. Stuart G, Spruston N. Determinants of voltage attenuation in neocortical pyramidal neuron dendrites. J. Neurosci. 1998;18:3501–3510. doi: 10.1523/JNEUROSCI.18-10-03501.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Suzuki S, Rogawski MA. T-type calcium channels mediate the transition between tonic and phasic firing in thalamic neurons. Proc Natl Acad Sci U S A. 1989;86:7228–7732. doi: 10.1073/pnas.86.18.7228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Szilagyi T, De Schutter E. Effects of variability in anatomical reconstruction techniques on models of synaptic integration by dendrites: a comparison of three Internet archives. Eur J Neurosci. 2004;19:66. doi: 10.1111/j.1460-9568.2004.03222.x. [DOI] [PubMed] [Google Scholar]
  50. Tamamaki N, Yanagawa Y, Tomioka R, Miyazaki J, Obata K, Kaneko T. Green fluorescent protein expression and colocalization with calretinin, parvalbumin, and somatostatin in the GAD67-GFP knock-in mouse. J Comp Neurol. 2003;467:60–79. doi: 10.1002/cne.10905. [DOI] [PubMed] [Google Scholar]
  51. Williams SR, Turner JP, Anderson CM, Crunelli V. Electrophysiological and morphological properties of interneurones in the rat dorsal lateral geniculate nucleus in vitro. J Physiol. 1996;490:129–147. doi: 10.1113/jphysiol.1996.sp021131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Zhu J, Heggelund P. Muscarinic regulation of dendritic and axonal outputs of rat thalamic interneurons: a new cellular mechanism for uncoupling distal dendrites. J Neurosci. 2001;21:1148–1159. doi: 10.1523/JNEUROSCI.21-04-01148.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Zhu JJ, Uhlrich DJ. Nicotinic receptor-mediated responses in relay cells and interneurons in the rat lateral geniculate nucleus. Neuroscience. 1997;80:191–202. doi: 10.1016/s0306-4522(97)00095-x. [DOI] [PubMed] [Google Scholar]
  54. Zhu JJ, Uhlrich DJ, Lytton WW. Burst firing in identified rat geniculate interneurons. Neuroscience. 1999a;91:1445–1460. doi: 10.1016/s0306-4522(98)00665-4. [DOI] [PubMed] [Google Scholar]
  55. Zhu JJ, Uhlrich DJ, Lytton WW. Properties of a hyperpolarization-activated cation current in interneurons in the rat lateral geniculate nucleus. Neuroscience. 1999b;92:445–457. doi: 10.1016/s0306-4522(98)00759-3. [DOI] [PubMed] [Google Scholar]

Articles from The Journal of Physiology are provided here courtesy of The Physiological Society

RESOURCES