Correspondence to: Angel Alonso, Department of Neurology and Neurosurgery, Montreal Neurological Institute, McGill University, 3801 University Street, Montréal, Québec, H3A 2B4 Canada. Fax: (514) 398-8106; E-mail:mdao{at}musica.mcgill.ca.
![]() |
Abstract |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
The functional and biophysical properties of a sustained, or "persistent," Na+ current (INaP) responsible for the generation of subthreshold oscillatory activity in entorhinal cortex layer-II principal neurons (the "stellate cells") were investigated with whole-cell, patch-clamp experiments. Both acutely dissociated cells and slices derived from adult rat entorhinal cortex were used. INaP , activated by either slow voltage ramps or long-lasting depolarizing pulses, was prominent in both isolated and, especially, in situ neurons. The analysis of the gating properties of the transient Na+ current (INaT) in the same neurons revealed that the resulting time-independent "window" current (INaTW) had both amplitude and voltage dependence not compatible with those of the observed INaP , thus implying the existence of an alternative mechanism of persistent Na+-current generation. The tetrodotoxin-sensitive Na+ currents evoked by slow voltage ramps decreased in amplitude with decreasing ramp slopes, thus suggesting that a time-dependent inactivation was taking place during ramp depolarizations. When ramps were preceded by increasingly positive, long-lasting voltage prepulses, INaP was progressively, and eventually completely, inactivated. The V1/2 of INaP steady state inactivation was approximately -49 mV. The time dependence of the development of the inactivation was also studied by varying the duration of the inactivating prepulse: time constants ranging from ~6.8 to ~2.6 s, depending on the voltage level, were revealed. Moreover, the activation and inactivation properties of INaP were such as to generate, within a relatively broad membrane-voltage range, a really persistent window current (INaPW). Significantly, INaPW was maximal at about the same voltage level at which subthreshold oscillations are expressed by the stellate cells. Indeed, at -50 mV, the INaPW was shown to contribute to >80% of the persistent Na+ current that sustains the subthreshold oscillations, whereas only the remaining part can be attributed to a classical Hodgkin-Huxley INaTW. Finally, the single-channel bases of INaP slow inactivation and INaPW generation were investigated in cell-attached experiments. Both phenomena were found to be underlain by repetitive, relatively prolonged late channel openings that appeared to undergo inactivation in a nearly irreversible manner at high depolarization levels (-10 mV), but not at more negative potentials (-40 mV).
Key Words: persistent Na+ current, window current, stellate cells, oscillations, patch clamp
![]() |
INTRODUCTION |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
The so-called persistent sodium current (INaP)1 is known to be expressed, together with the classical, transient sodium current (INaT), by numerous mammalian neuronal types. Its basic features include persistence during prolonged depolarizations, lower threshold of activation than INaT, and low amplitude (the underlying conductance normally representing 0.22% of the total sodium conductance;
The entorhinal cortex (EC) has proven to be a particularly interesting neuronal system for the study of INaP functions. In the stellate cells of EC layer II, which give rise to the main cortical projection to the hippocampus (
For the above considerations, a detailed knowledge of the biophysical properties of INaP expressed by EC layer II neurons seems of great interest since it would help to understand how this current specifically influences the physiological behavior of these cells. In this study, we have characterized INaP in both acutely isolated and in situ EC layer II neurons. Our results revealed the existence of some interesting biophysical properties of INaP that had not been thoroughly investigated yet, including: (a) a slow voltage- and time-dependent inactivation occurring with voltage-dependent time constants in the order of seconds; (b) a full inactivation at sufficiently positive potentials; and (c) a truly persistent, or "window," current arising from the particular activation and steady state inactivation properties of the corresponding conductance. Moreover, we describe the single-channel events that account for all of the above-mentioned macroscopic phenomena.
![]() |
MATERIALS AND METHODS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Slice and Cell Preparation
Young-adult Long-Evans rats (P25P35) were killed by decapitation. The brain was quickly removed under hypothermic conditions, blocked on the stage of a vibratome (Pelco), and submerged in an ice-cold cutting solution containing (mmol/liter): 115 NaCl, 5 KCl, 4 MgCl2, 1 CaCl2, 20 PIPES, and 25 D-glucose, pH 7.4 with NaOH, bubbled with pure O2. Osmotic pressure () of the latter solution, as measured with a Micro Osmette 5004 osmometer, was typically 320 mOsm. Horizontal slices of the retrohippocampal region were cut at 350400 µm. For in situ recordings, slices were stored at room temperature in the above solution until use. For recordings on isolated neurons, the layer II of medial entorhinal cortex was dissected from each slice (
Whole-Cell Recordings on Slices
The recording chamber was mounted on the stage of an upright microscope (see below). Slices were laid onto the bottom of the chamber and perfused with an extracellular solution containing (mmol/liter): 34 NaCl, 26 NaHCO3, 80 tetraethylammonium (TEA)-Cl, 5 KCl, 3 CsCl, 2 CaCl2, 3 MgCl2, 2 BaCl2, 2 CoCl2, 0.4 CdCl2, 4 4-aminopyridine (4-AP), 10 glucose, pH 7.4 when bubbled with 95% O2, 5% CO2 (
320 mOsm). The association of Co2+ and Cd2+ was found to depress residual inward rectification insensitive to the application of tetrodotoxin (TTx; Sigma-Aldrich Canada Ltd.) more effectively than Cd2+ alone in slices (but not in freshly dissociated neurons; see below). Preliminary experiments were performed in in situ neurons in the absence of extracellular Co2+ and Ba2+ and in the presence of 0.2 rather than 0.4 mM Cd2+. Both average peak amplitude and current density of ramp (50 mV/s)-evoked, TTx-subtracted INaPs recorded under the former and latter ionic conditions (-179.3 ± 120.7 vs. -195.8 ± 100.0 pA, respectively, and -17.7 ± 29.1 vs. -12.9 ± 10.6 pA/pF, respectively, n = 54 and 8, respectively) were not significantly different (P = 0.68 and 0.63, respectively). Hence, possible enhancing effects of nonphysiological extracellular divalent cations on INaP amplitude (
adjusted to ~290 mOsm with mannitol). When filled with the above solution, the patch pipettes had a resistance of 35 M
. Slices were observed with an Axioskop microscope (Carl Zeiss, Inc.) equipped with a 40x water-immersion objective lens and differential-contrast optics. A near-infrared charge-coupled device camera (XC-75; Sony Corp.) was also connected to the microscope, and used to improve cell visualization for identification of neuron types and during the approaching and patching procedures. With this equipment, the principal cells of EC layer II were easily distinguished based on their somato-dendritic shape (stellate cells:
) and the whole-cell configuration were obtained by suction (
(n = 54). Cell capacitance was evaluated online by canceling the fast component of whole-cell capacitive transients evoked by -10-mV voltage steps with the amplifier compensation section, and reading out the corresponding value. Voltage-clamp recordings were performed at room temperature (~22°C) using an Axopatch 1D amplifier (Axon Instruments). The general holding potential was -80 mV.
Whole-Cell Recordings on Isolated Neurons
The recording chamber was mounted on the stage of an inverted microscope (see below). After seeding into the chamber, dissociated cells were perfused with a standard HEPES buffer containing (mmol/liter): 140 NaCl, 5 KCl, 10 HEPES (free acid), 2 CaCl2, 2 MgCl2, 25 glucose, pH 7.4 with NaOH, bubbled with pure O2 (
320 mOsm). After wash-out of cell debris, cell perfusion was switched to a solution suitable for Na+-current isolation containing (mmol/liter): 100 NaCl, 40 TEA-Cl, 10 HEPES (free acid), 2 CaCl2, 3 MgCl2, 0.2 CdCl2, 5 4-AP, 25 glucose, pH 7.4 with NaOH, bubbled with pure O2 (
318 mOsm). The intrapipette solution was the same as described in the previous paragraph. Cells were observed at 400x with an Axiovert 100 microscope (Carl Zeiss, Inc.). After tight-seal formation (>100 G
) and the establishment of the whole-cell configuration, series resistance was on average 12.0 ± 4.5 M
(n = 38), and was always compensated by ~70%. The remaining procedures and experimental conditions were the same as described in the previous paragraph.
Single-Channel Recordings
Single-channel, cell-attached experiments were performed in acutely isolated neurons. After seeding into the recording chamber, cells where initially perfused with the same solution as described in the previous paragraph. The pipette solution contained (mmol/liter): 130 NaCl, 35 TEA-Cl, 10 HEPES-Na, 2 CaCl2, 2 MgCl2, 5 4-AP, pH 7.4 with HCl (
338 mOsm). Single-channel patch pipettes had resistances ranging from 10 to 35 M
when filled with the above solution, and were always coated with Sylgard® (Dow Corning Corp.) from the shoulder to a point as close as possible to the tip so as to minimize stray pipette capacitance. After obtaining the cell-attached configuration, the extracellular perfusion was switched to a high-potassium solution containing: 140 K-acetate, 5 NaCl, 10 HEPES (free acid), 4 MgCl2, 0.2 CdCl2, 25 glucose, pH 7.4 with KOH (
320 mOsm) so as to hold the neuron resting membrane potential at or near 0 mV. Recordings were performed at room temperature with an Axopatch 200B amplifier (Axon Instruments). Capacitive transients and linear current leakage were minimized online by acting on the respective built-in compensation sections of the amplifier. Long-duration (20-s) depolarizing voltage steps were delivered one every 40 s from a holding potential of -80 or -100 mV.
Data Acquisition
Voltage protocols were commanded and current signals were acquired with a Pentium PC interfaced to an Axon TL1 interface, using the Clampex program of the pClamp 6.0.2 software (Axon Instruments). Current signals were filtered online (using the amplifier's built-in low pass filter) and digitized at different frequencies according to the specific experimental aim. Filtering and acquisition frequencies were 5 and 20 kHz, respectively, for INaT recordings; 0.11 and 0.6710 kHz (depending on the protocol duration), respectively, for INaP recordings; 1 and 2 kHz, respectively, for single-channel recordings. In all of the voltage protocols applied, cell-membrane potential was kept at the holding level for 15 (in whole-cell experiments) or 20 s (in single-channel experiments) between the end of each sweep and the beginning of the subsequent sweep (or of the conditioning prepulse preceding it, when applied). This avoided the development of cumulative voltage-dependent inactivation of INaP during consecutive acquisition cycles.
Data Analysis
Whole-cell recordings were analyzed by means of the Clampfit program (Axon Instruments). Offline leak subtraction was performed on INaT - (but not INaP-) protocol traces. Current density was calculated by dividing the peak current amplitude by cell capacitance, estimated as explained above. Conductance values were calculated from Na+-current amplitudes by applying the extended Ohm's law in the form: GNa = INa/(V VNa), where VNa is the nominal (Nernst) Na+ reversal potential. Data were fitted with exponential functions, I = (Ai · exp(-t/i) + C, using Clampfit, or with Boltzmann functions, G = Gmax/{1 + exp[(V - V1/2)/k)]}, using Origin 3.06 (MicroCal Software).
Single-channel recordings were analyzed using Clampfit, Fetchan, and pStat (Axon Instruments). Residual capacitive transients were nullified by offline subtracting fits of average blank traces. Residual leakage currents were carefully measured in every single sweep at trace stretches devoid of any channel openings, and digitally subtracted. Channel dwell times were determined using a standard threshold routine of the Fetchan program. Ensemble-average traces were fitted with single exponential functions, I = A · exp(-t/) + C, using Clampfit. Dwell-time histograms were fitted with double exponential functions, N = A1 · exp(-t/
1) + A2 · exp(-t/
2), using pStat.
Average values were expressed as mean ± SD, unless otherwise explicitly stated. Statistical significance was evaluated by means of the two-tail Student's t test for unpaired data.
Modeling INaP
For a phenomenological description of INaP activation and slow inactivation, a simple Hodgkin-Huxley model was assumed. We applied the basic relationship:
![]() |
(1) |
where
![]() |
(2) |
and m and h are the probabilities of the activating and inactivation particles, respectively, to be in the permissive position. INaP activation was assumed to be instantaneous, and m(V) was derived directly from the GNaP activation curve. h
(V) was derived directly from the GNaP steady state inactivation curve. The transitions of the inactivating particle, h, were modeled according to the following first-order kinetic scheme:
from which it follows:
![]() |
(3) |
where
![]() |
(4) |
and
![]() |
(5) |
Numerical values for the rate constants, and ß, were derived from the experimental values of time constants of inactivation and recovery from inactivation (
h) and from the h
curve by applying Equation 4 and Equation 5. After obtaining the analytical functions describing the voltage dependence of the rate constants (see RESULTS), the time course of INaPs activated in response to various voltage protocols was numerically reconstructed on the basis of the Equation 1 and Equation 2, and Equation 3 in its differential form. The simulation programs were compiled using QuickBASIC 4.5 (Microsoft Corp.).
![]() |
RESULTS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
General Properties of INaP in Acutely Dissociated and In Situ Neurons
In acutely dissociated EC layer II principal neurons, a prominent INaP could be evoked by delivering long-lasting depolarizing pulses. Figure 1 A illustrates current traces from a representative neuron. INaP was blocked by 1 µM TTx, and was therefore routinely isolated via TTx subtraction. INaP threshold of activation was at about -65 mV, and peak at -30 mV. Average INaP absolute amplitude (derived by averaging the data points between 400 and 500 ms from the pulse onset) and current density (calculated as explained in MATERIALS AND METHODS) were -96.5 ± 61.5 pA and -12.0 ± 7.0 pA/pF, respectively, at the peak of the current-voltage (IV) relationship (n = 5). The average INaP IV relationship showed a linear region from -30 to -5 mV, the linear best fitting of which returned a zero-current level at +63.0 mV (not shown). This value compares favorably with the theoretical Nernst Na+ reversal potential calculated for our ionic conditions (VNa = +61.0 mV). The voltage dependence of the conductance underlying INaP (GNaP, calculated as explained in the MATERIALS AND METHODS) is shown in the average plot of Figure 1 A, inset. Boltzmann fitting to data points returned a half-activation voltage, V1/2, of -44.4 mV, and a slope factor, k, of -5.2 mV. The ratio between the peak GNaP value found in each cell and the maximal value of the conductance underlying the transient Na+ current (INaT) expressed by the same cell was also calculated, and averaged 0.0187 ± 0.0097 (n = 5).
|
To quickly explore the whole voltage range of INaP activation, ramp protocols were then used. Slow ramps at 50 mV/s were initially selected since they allowed full inactivation of fast-decaying Na+-current component(s). Figure 1 B shows the currents evoked by such a protocol in a representative acutely dissociated neuron, both in control conditions and in the presence of 1-µM TTx (inset). Offline digital subtraction returned the TTx-sensitive INaP in isolation (Figure 1 B). The continuous IV relationship thus obtained showed a threshold at -70/-60 mV and a peak at -40/-30 mV. Noteworthy in both pulse and ramp protocols, INaP activation was accompanied by an evident increase in current noise, especially at voltage levels close to the peak of the IV relationship (Figure 1A and Figure B), consistent with the relatively high conductance (~20 pS) characterizing the channels responsible for INaP generation in EC layer II neurons (
Ramp protocols were also used in experiments performed in in situ neurons. In this situation, TTx subtraction always returned prominent INaPs in isolation, whose IV relationship closely resembled that of INaPs in acutely dissociated neurons (Figure 1 C). INaP amplitude, measured at the peak of the IV relationship, was significantly higher in in situ than in isolated neurons (-179.3 ± 120.7 pA, n = 54, vs. -65.6 ± 37.9 pA, n = 38; P < 5 x 10-7), whereas the current density did not significantly differ in the two situations (-17.7 ± 29.1 pA/pF, n = 54, vs. -16.5 ± 8.6 pA/pF, n = 38; P = 0.8). These findings strongly suggest that the channels responsible for INaP are located not only on the soma, but also on neuronal processes severed by the dissociation procedure. Activation curves of INaPs recorded in both in situ and isolated neurons were also constructed. Conductance values were derived from INaPs by applying the extended Ohm's law (see MATERIALS AND METHODS), and the resulting activation curves were fitted with single Boltzmann functions (Figure 1 D). Average half-activation potentials and slope factors were very similar in in situ neurons (V1/2 = -51.3 ± 3.9 mV, k = -4.0 ± 0.7 mV, n = 39) and isolated neurons (V1/2 = -48.7 ± 4.7 mV, k = -4.4 ± 0.9 mV, n = 19). These values compare favorably with those obtained from step protocols (see above), and are also in good agreement with the activation parameters previously reported for INaPs expressed in other neuronal systems (
Given the effectiveness of TTx subtraction in isolating INaPs in both isolated and in situ neurons, this procedure was routinely used in our study. All of the data presented from this point on are from TTx-subtracted currents.
INaP Is Not a Window Current Generated by Transient Na+ Channels
It is well known that a noninactivating, "window" current (INaTW) can be generated by the gating properties of fast, transient Na+ channels (
|
|
INaP Inactivates in a Time- and Voltage-dependent Manner
Whereas a depolarization-activated conductance, once maximally recruited, would be expected to maintain a steady value at more positive voltage levels, GNaP , as mentioned above, consistently showed some degree of decline from its maximum. A possible explanation for this observation is the existence of a time-dependent inactivation of GNaP acting during the ramp (
|
The above data clearly pointed to the existence of a time- and voltage-dependent inactivation of INaP. However, inactivation properties of voltage-dependent channels have been shown to be possibly affected by the composition of intracellular milieu, and in particular by intracellular nonphysiological halogenic anions (
We then investigated the issue of INaP inactivation in further detail by analyzing the effects of variable prepulse potentials on ramp-activated INaPs. The protocol employed is illustrated in Figure 4 A1. 50-mV/s voltage ramps were preceded by very-long-lasting (15 s) conditioning prepulses at various voltage levels (Vcond). When Vconds of -90 to -50 mV were used, the ramp started from the same voltage level as Vcond itself, rather than from a fixed, negative voltage level: in this way, the possible occurrence of recovery from inactivation during the initial part of the ramp was avoided. At more positive Vconds, the ramp started from -50 mV, so as to preserve the voltage region of INaP peak. Currents recorded with the above protocol in a representative in situ neuron are shown in Figure 4 A2. INaP peak amplitude turned out to markedly depend on the conditioning potential. Average, normalized current traces obtained from seven in situ neurons are depicted in Figure 4 B1. It can be observed that voltage-dependent steady state inactivation of INaP was nearly complete at about -20 mV. The average plot of INaP's voltage dependence of inactivation (Figure 4 B2) could be fitted by a single Boltzmann function, with V1/2 at about -49 mV and a slope factor, k, of ~10 mV. In addition, we also constructed an activation plot from the average INaP derived from the same neuron pool, and fitted it with a single Boltzmann function (Figure 4 B2). Note that, importantly, GNaP activation and steady state inactivation functions overlapped over a wide voltage range. Due to this phenomenon, a significant window conductance (GNaPW) is expected to arise from GNaP. The predicted voltage dependence of GNaPW is depicted in Figure 4 B2 (dotted line), and will be compared with relevant experimental data later on in the paper.
|
Time Dependence of INaP Inactivation and Recovery from Inactivation
The kinetic properties of INaP voltage-dependent inactivation were then further characterized. Time dependence of inactivation was first analyzed by means of prepulse-ramp protocols (Figure 5 B); the voltage ramp eliciting INaP was preceded by a prepulse at various voltage levels (from -60 to -20 mV), which was made to vary in duration from 0 to up to 20 s. Currents recorded in response to such a protocol in a representative in situ neuron are shown in Figure 5 A. Average, normalized peak-current amplitudes were used for constructing plots of time dependence of inactivation, each one referring to a specific conditioning potential (Figure 5 C). These plots could be best fitted with single exponential functions; the time constants were slow and ranged from ~6.8 to ~2.6 s, depending on the conditioning potential.
|
The time course of INaP recovery from inactivation was also investigated. The voltage protocols applied (Figure 6 B) consisted of a first 10-s prepulse at -30 mV that substantially inactivated INaP , followed by a second prepulse at -90 or -80 mV of variable duration (from 0 to 10 s), and by the standard voltage ramp. Currents recorded in response to such a protocol in a representative in situ neuron are shown in Figure 6 A. Average, normalized peak-current amplitudes were used for constructing plots of time dependence of recovery from inactivation, for both recovery potentials (Figure 6 C). These plots could be best fitted with single exponential functions, with time constants of ~5.2 (-80 mV), and 4.7 s (-90 mV).
|
A Major, Persistent Window Current Is Generated as a Consequence of GNaP Gating Properties
As mentioned above and illustrated in Figure 4 B2, the wide overlapping of INaP activation and steady state inactivation curves is expected to result in a prominent window current (INaPW) distinct from the classical window current (INaTW) predicted on the basis of the gating properties of the fast, transient Na+ conductance. To test this prediction, voltage protocols consisting of very-long-lasting depolarizing pulses were applied to in situ neurons so as to try to uncover steady current components in INaP. Figure 7 A1 shows average, TTx-subtracted Na+ currents obtained from five neurons in response to 15-s voltage steps at -60 to -10 mV. After an initial phase displaying fast and intermediate-speed decay components, a slower decaying current component, which was identified as the INaP under study, became evident. When the last 14 s of the current trace were considered, the decay phase of INaP could be best fitted by a single exponential function, with voltage-dependent time constants very similar to those determined for the time-dependent inactivation of INaP revealed by ramp protocols (see above). In addition to the decaying component, fittings returned a steady (offset) component (Iss) whose amplitude also displayed a marked voltage dependence. When plotted as a function of test potential (Figure 7 B), the normalized Iss amplitude closely paralleled that of the expected INaPW. Therefore, a steady current component of INaP can be directly demonstrated whose voltage-dependent behavior fits that predicted for the time-independent INaPW.
|
On the basis of the above data, we then estimated the relative contribution of INaPW and INaTW to the total, noninactivating Na+ current generated by EC layer II cells in a subthreshold region of membrane voltages, where it is known to sustain theta-like membrane-potential oscillations lasting for indefinitely long periods (
Modeling of INaP Slow Inactivation
To further clarify the basis of the experimentally observed decline in GNaP at positive potentials (Figure 2 D), a theoretical reconstruction of the biophysical properties of INaP inactivation was carried out. A simple Hodgkin-Huxley model, considering a single inactivation gate switching between two energy states, was considered in order to give account for the monoexponential time course of INaP decay and recovery from inactivation. This reconstruction was merely phenomenological and was given no mechanistic meaning since single-channel data clearly indicated different features of the underlying elementary events (see below). The time constants of INaP inactivation and recovery from inactivation and the data on voltage dependence of INaP steady state inactivation were processed to derive numerical values for the rate constants of the inactivation-gate transitions, as explained in MATERIALS AND METHODS. Figure 8 B shows the voltage dependence of the values thus obtained for the two rate constants, and ß. The plots were then best fitted with the empirical function,
(or ß) = (a · Vm + b)/{1 - exp[(Vm + b/a)/k]}, where Vm is the membrane voltage. The numerical values returned by the fittings for the a, b, and k coefficients, in both the
and ß plots, are indicated in the legend to Figure 8. The voltage-dependence functions thus obtained for
and ß were then used to derive the predicted voltage dependence of INaP inactivation-gating time constants (see MATERIALS AND METHODS). The concordance between the reconstructed, theoretical function and the real data is shown in Figure 8 A.
|
The kinetic parameters obtained in the above manner were then used to verify the possible effects of INaP slow inactivation on the results of voltage-clamp ramp protocols. Reconstructed INaPs evoked by simulated voltage ramps of variable slope are illustrated in Figure 8 C. It is apparent that progressively reducing the depolarization rate causes a decrease in INaP peak amplitude, and the appearance of increasing discrepancies between the true INaP voltage dependence and that measured in the late part of the ramp protocol. Figure 8 E shows the voltage dependence of GNaP as obtained in response to a simulated 50 mV/s-ramp protocol: the reconstructed GNaP declined at voltages positive to about -30 mV very similarly to the experimentally observed GNaP (see Figure 2 D), whereas GNaP voltage dependence of activation, as measured by considering only the early part of the ramp protocol, was negligibly affected. Moreover, the plot of reconstructed INaP's peak amplitude as a function of the inverse of ramp slope (Figure 8 D) could be fitted, with some approximation, with a single exponential function, with a slope constant () similar to the slow slope constant observed in experimental plots (see Figure 3 B), and no sign of any fast, early component. The latter results further confirm the adequacy of the ramp protocol we routinely used for characterizing the biophysical properties of INaP, and in particular the slow inactivation of this current.
Finally, a simulated protocol of steady state inactivation revealed that, during a ramp preceded by a prepulse that fully inactivates INaP, a significant recovery from inactivation can occur provided that the ramp starts at sufficiently negative levels and is sufficiently slow. For instance, during a 25-mV/s ramp starting at -80 mV, 17% of the total current can recover from inactivation (not shown). By contrast, voltage protocols similar to those we employed for the study of INaP voltage-dependent inactivation (see above) determined no appreciable ramp-dependent recovery from inactivation of INaP .
Single-Channel Bases of INaP Slow Inactivation and INaPW Generation
The single-channel basis of INaP in EC layer II neurons has been described elsewhere (
Recordings from cell-attached patches in acutely isolated EC layer II neurons frequently revealed the presence of a persistent Na+-dependent channel activity that proved to remain stable even over prolonged periods of time (tens of minutes). When very-long-lasting (20 s) depolarizing steps at -10 mV were commanded from a holding potential of -80 mV, multiple, repetitive single-channel openings were observed that tended to cluster preferentially at the beginning of the test pulse. Typical examples of this channel activity for a series of consecutive 20-s test pulses is shown in Figure 9 A1. A detail of some prolonged, late channel openings is also provided by Figure 9 A1, inset. Ensemble averaging of multiple traces was then carried out for each patch (Figure 9 A2). Due to the very long overall duration of the recording cycle required for every single sweep (40 s), only a limited number of traces could be recorded in each patch (15 on average), what explains the low signal-to-noise ratio of ensemble-average traces. In all cases, however, the averaged currents showed a noticeable trend to decay towards zero, and this decay could be properly fitted by a single exponential function (Figure 9 A2, blank trace). The time constant of average-current decay was 2.66 ± 0.52 s in six patches at -10 mV, a value that compares favorably with those found in whole-cell protocols on macroscopic INaP inactivation at the most positive voltage levels tested (see Figure 8 A). Ensemble averaging of all of the available sweeps recorded from the same six patches returned a better signal-to-noise ratio (Figure 9 B). The decay time constant of the resulting average current was 2.46 s, again in good agreement with the data obtained from the analysis of both individual patches and whole-cell recordings.
|
The open-time distribution of the single-channel activity evoked by 20-s depolarizing steps at -10 mV was then investigated. Figure 10 A shows the open-time distribution found in the same patch as illustrated in Figure 9 A. As in this case, all plots were best fitted by double-exponential functions, with average time constants of 3.35 ± 0.86 and 21.07 ± 17.74 ms, and an average ratio of the slow vs. fast exponential-component weight (W = A · ) of 0.118 ± 0.048 (n = 6). It seems important to point out at this time that: (a) even the faster of the two time constants exceeds by more than six times the mean open time found in classical, transient Na+ channel openings at approximately the same test voltage level (
|
Long-lasting depolarizing protocols at more negative test-voltage levels were also applied in cell-attached recordings so as to investigate the possible bases of INaPW generation. A typical example of the recordings obtained at the test voltage of -40 mV is shown in Figure 11 A. Again, multiple, repetitive single-channel openings were observed and these were able to generate a measurable inward current in ensemble-average traces (Figure 11 B). Note, however, that at these more negative voltage levels channel openings were more widely distributed over the entire 20-s sweeps. Consequently, at -40 mV the ensemble average-current decayed at a slower rate (with a time constant of 4.33 ms) than at -10 mV. In addition, this decay was towards a steady value (C in exponential fittings; see MATERIALS AND METHODS) that was higher than zero and represented 25.1% of the current's total amplitude coefficient (namely A + C). These data are in good agreement with the whole-cell data on INaP inactivation and INaPW generation illustrated above. They are also consistent with the idea that the same single-channel events can account for INaP as well as a steady, nondecaying Na+-current component (namely INaPW) generated within a limited voltage window, where it represents a substantial fraction of total INaP .
|
Finally, the analysis of open-time distribution at -40 mV also revealed the existence of two exponential components (Figure 10 B), with average time constants of 1.33 ± 0.14 and 6.63 ± 3.47 ms, and an average W2/W1 ratio of 0.069 ± 0.042 (n = 3). At this potential, therefore, the discrepancy between mean open times and time constants of inactivation of both ensemble-average currents and macroscopic INaP was even bigger than at -10 mV.
The observation that the channel mean open times are far exceeded by the time constants of ensemble-average current inactivation clearly indicates that the latter do not reflect average channel-opening lifetimes in a simple model considering two open states each undergoing one single closure process (with rate constant b) towards an absorbing state, like:
with the index i being either 1 or 2, and with ai >> bi. Rather, alternative models considering late first openings and/or late reopenings must be taken into account. In an extreme situation, late channel openings such those of Figure 9 A1 might all be first openings of as many distinct channels inactivating towards an absorbing state, in a model qualitatively similar to that depicted in Scheme I, but in which the kinetics of the macroscopic inactivation process would rather reflect the rate constants of the closed-to-open reactions, ai. This interpretation, which resembles the classical Aldrich-Corey-Stevens model of the kinetics of transient Na+ channels (
A nonnegligible rate constant () for the reaction Ii
Oi should also be considered for more negative voltages, at which a major INaPW is produced. The analytical derivation of relaxation time constants from rate constants (see
i >>
i, and
1
2 (or if only one inactivated state exists, communicating with one of the two open states), one slow time constant of ensemble-average current inactivation would be produced, which would faithfully reflect 1/
. Other possible kinetic schemes consider that an inactivated state is reached from a closed rather than a conducting state. The evaluation of the relationships between transition rate constants in such a scheme and the time constants of ensemble-average current inactivation would require the study of closed-time distribution. This could not be reliably accomplished from our data since in no patches in which slow inactivation was studied could the presence of one single channel be assumed. Despite these limitations in providing precise kinetic schemes, our data clearly demonstrate the importance of late channel (re)openings, rather than of early, very-long-lasting openings, for the generation of slow kinetic components in macroscopic INaP.
![]() |
DISCUSSION |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
The present study provides a biophysical characterization of the INaP expressed by rat EC layer II principal neurons. The major issues we addressed deal with the mechanism of generation of INaP , the possible influence of the voltage protocols employed for the study of INaP biophysical properties, and the real degree of INaP persistence over time and specific voltage windows.
In general, the hypotheses on generation of persistent Na+ currents consider two main possibilities: (a) INaP simply derives from an incomplete steady inactivation of transient Na+ channels over a narrow voltage window, due to the partial superimposition of activation and steady state inactivation voltage-dependence curves of the corresponding conductance; (b) INaP is the result of channel openings that do not functionally behave according to the properties of transient Na+ channels, but derive from a rare and atypical gating modality of the same channels (
Another interesting issue raised by our data relates to the definition itself of INaP. Due to their practicality, ramp protocols are the most widely used means for eliciting and isolating INaP. However, the employment of such protocols presupposes that they can adequately reproduce the voltage dependence of INaP without significantly recruiting any other kinetically different (i.e., faster-decaying) Na+-current component. This is often implicitly assumed rather than demonstrated. Moreover, the notion of "slow voltage ramp" suitable for INaP activation varies considerably among different experimental works (
Moreover, our data show that, importantly, the process of INaP slow inactivation is a potential source of distortions in the measurements of INaP biophysical parameters (amplitude, voltage dependence of activation, reversal) when this current is elicited by voltage ramps. All these considerations point to the importance of accurately choosing the voltage protocol applied for the study of INaP in each specific experimental situation. We propose that some previously reported, atypical biophysical features of INaP , particularly regarding its nonsigmoidal voltage dependence of activation (
The inactivation properties of what, according to our operative definition, can be considered as INaP have been characterized in detail in our study. Our experiments indicate that, in our preparation, the steady state inactivation of the conductance underlying INaP (GNaP) (a) has a voltage dependence that extends over a wide voltage window, and (b) reaches a nearly complete level at -20 to -10 mV. The former of these features, together with the relative position of the GNaP activation curve along the voltage axis, is expected to give rise to a major, time-independent, "window" current over a limited voltage range. This window current (INaPW), clearly different from that arising from the voltage-dependence properties of the transient Na+ conductance (INaTW), could also be directly demonstrated both as nonzero baselines at the beginning of ramp protocols on voltage dependence of inactivation (see Figure 4), and as offsets, or pedestals, in exponentially decaying INaPs elicited by long-lasting voltage steps (Figure 7). The peak of the observed INaPW occurred at voltage levels very close to those at which INaP-dependent, theta-like subthreshold membrane-potential oscillations are generated by EC stellate cells (
INaP slow inactivation and recovery from inactivation were found to occur with voltage-dependent time constants in the order of a few seconds. We worked out an analytical reconstruction of the kinetics of these processes which, if introduced into suitable neuronal models, should allow us to make predictions on the effects of slow voltage-dependent inactivation on INaP modulation of membrane-voltage events. We demonstrated that INaP inactivation can considerably affect the apparent current maximal amplitude during the delivery of slow depolarizing ramps. Therefore, it is conceivable that the recruitment of INaP and its impact onto membrane-voltage dynamics can be significantly influenced by the speed of membrane depolarization. For instance, the ability of INaP to bring the membrane potential towards threshold for action-potential firing may be expected to be higher in response to a step depolarizing current injection than to slower or sustained depolarizations. This is consistent with the role of INaP, demonstrated in various central nervous system neurons (
The question can then be raised whether the biophysical properties we describe here for the INaP expressed by EC stellate cells represent a general feature of this current in various neuronal populations. Slow time-dependent inactivation of INaP has been previously reported in neocortical pyramidal neurons (
Finally, the present study provides an insight into the fine mechanisms underlying the complex biophysical features displayed by INaP. Our previous work has already clarified the nature and elementary properties of the single-channel events responsible for INaP generation in EC principal neurons (
In conclusion, the INaP expressed by EC layer II principal neurons is a prominent current operating in a subthreshold range of membrane potentials, most of which is generated by a process independent of the classical gating behavior of the transient Na+ conductance. It also displays complex and previously nonrecognized biophysical characteristics that appear to be tailored to the specific role this current is known to play in the generation of the sub- and near-threshold membrane-potential events typical of the same neurons. The concept of "persistent Na+ current" may turn out to be susceptible to some critical revision also in other experimental situations, with reference to both the existence of multiple and heterogeneous functional components, and the expression of kinetic and voltage-dependence properties that might influence their impact onto neuronal function in previously unforeseen ways.
![]() |
Footnotes |
---|
Portions of this work were previously published in abstract form (Magistretti, J., and A. Alonso. 1998. NYAS Meetings. Abstr. No. PI29; Alonso, A., and J. Magistretti. 1998. Soc. Neurosci. Abstr. 24:2035).
![]() |
Acknowledgements |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
J. Magistretti thanks the Human Frontier Science Program Organization (HFSPO), the Istituto Nazionale Neurologico "Carlo Besta," and Dr. Marco de Curtis for support.
This work has been funded by grants from the Medical Research Council of Canada, HFSPO, and the North Atlantic Treaty Organization to A. Alonso.
Submitted: 2 June 1999
Revised: 4 August 1999
Accepted: 9 August 1999
1used in this paper: 4-AP, 4-aminopyridine; EC, entorhinal cortex; INaP and GNaP, persistent Na+ current and conductance; INaPW and GNaPW, window current and conductance resulting from GNaP; INaT and GNaT, transient Na+ current and conductance; INaTW and GNaTW, window current and conductance resulting from GNaT; IV, currentvoltage; TEA, tetraethylammonium; TTx, tetrodotoxin
![]() |
References |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Adey, W.R., Sunderland, S., Dunlop, C.W. 1957. The entorhinal area: electrophysiological studies of its interrelations with rhinencephalic structures and the brainstem. Electroencephalogr. Clin. Neurophysiol. 9:309-324.
Aldrich, R.W., Corey, D.P., Stevens, C.F. 1983. A reinterpretation of mammalian sodium channel gating based on single channel recording. Nature. 306:436-441[Medline].
Alonso, A., de Curtis, M., Llinás, R. 1990. Postsynaptic Hebbian and non-Hebbian long-term potentiation of synaptic efficacy in the cortex in slices and in the isolated adult guinea pig brain. Proc. Natl. Acad. Sci. USA. 87:9280-9284[Abstract].
Alonso, A., García-Austt, E. 1987a. Neuronal sources of theta rhythm in the entorhinal cortex of the rat. I. Laminar distribution of theta field potentials. Exp. Brain Res. 67:493-501[Medline].
Alonso, A., García-Austt, E. 1987b. Neuronal sources of theta rhythm in the entorhinal cortex of the rat. II. Phase relations between unit discharges and theta field potentials. Exp. Brain Res. 67:502-509[Medline].
Alonso, A., Klink, R. 1993. Differential electroresponsiveness of stellate and pyramidal-like cells of medial entorhinal cortex layer II. J. Neurophysiol. 70:128-143
Alonso, A., Llinás, R.R. 1989. Subthreshold Na+-dependent theta-like rhythmicity in stellate cells of entorhinal cortex layer II. Nature. 342:175-177[Medline].
Alzheimer, C., Schwindt, P.C., Crill, W.E. 1993a. Postnatal development of a persistent Na+ current in pyramidal neurons from rat sensorimotor cortex. J. Neurophysiol. 69:290-292
Alzheimer, C., Schwindt, P.C., Crill, W.E. 1993b. Modal gating of Na+ channels as a mechanism of persistent Na+ current in pyramidal neurons from rat and cat sensorimotor cortex. J. Neurosci. 13:660-673[Abstract].
Amitai, Y. 1994. Membrane potential oscillations underlying firing patterns in neocortical neurons. Neuroscience. 63:151-161[Medline].
Arispe, N.J., Argibay, J.A., Rojas, L.V. 1984. Sodium currents in skeletal muscle fibres from the toad Bufo marinus. Q. J. Exp. Physiol. 69:507-519[Medline].
Baker, M.D., Bostock, H. 1997. Low-threshold, persistent sodium current in rat large dorsal root ganglion neurons in culture. J. Neurophysiol. 77:1503-1513
Brown, A.M., Schwindt, P.C., Crill, W.E. 1994. Different voltage dependence of transient and persistent Na+ currents is compatible with modal-gating hypothesis for sodium channels. J. Neurophysiol. 71:2562-2565
Buzsáki, G. 1996. The hippocampal-neocortical dialogue. Cereb. Cortex. 6:81-92[Abstract].
Cepeda, C., Chandler, S.H., Shumate, L.W., Levine, M.S. 1995. Persistent Na1 conductance in medium-sized neostriatal neurons: characterization using infrared videomicroscopy and whole cell patch-clamp recordings. J. Neurophysiol. 74:1343-1348
Chandler, W.K., Meves, H. 1970. Sodium and potassium currents in squid axons perfused with fluoride solutions. J. Physiol. 211:623-652[Medline].
Chao, T.I., Alzheimer, C. 1995. Do neurons from rat neostriatum express both a TTX-sensitive and a TTX-insensitive slow Na+ current? J. Neurophysiol. 74:934-941
Colquhoun, D., Hawkes, A.G. 1977. Relaxation and fluctuations of membrane currents that flow through drug-operated channels. Proc. R. Soc. Lond. B Biol. Sci. 199:231-262[Medline].
Colquhoun, D., Hawkes, A.G. 1981. On the stochastic properties of single ion channels. Proc. R. Soc. Lond. B Biol. Sci. 211:205-235[Medline].
Connors, B.W., Gutnick, M.J., Prince, D.A. 1982. Electrophysiological properties of neocortical neurons in vitro. J. Neurophysiol. 48:1302-1320
Crill, W.E. 1996. Persistant sodium current in mammalian central neurons. Annu. Rev. Physiol. 58:349-362[Medline].
Cummins, T.R., Howe, J.R., Waxman, S.G. 1998. Slow closed-state inactivation: a novel mechanism underlying ramp currents in cells expressing the hNE/PN1 sodium channel. J. Neurosci. 18:9607-9619
Cummins, T.R., Xia, Y., Haddad, G.G. 1994. Functional properties of rat and human neocortical voltage-sensitive sodium currents. J. Neurophysiol. 71:1052-1064[Abstract].
D'Angelo, E., De Filippi, G., Rossi, P., Taglietti, V. 1998. Ionic mechanism of electroresponsiveness in cerebellar granule cells implicates the action of a persistent sodium current. J. Neurophysiol. 80:493-503
Deisz, R.A., Fortin, G., Zieglgänsberger, W. 1991. Voltage dependence of excitatory postsynaptic potentials of rat neocortical neurons. J. Neurophysiol. 65:371-382
Dickson, C.T., Alonso, A. 1997. Muscarinic induction of synchronous population activity in the entorhinal cortex. J. Neurosci. 17:6729-6744
Doyère, V., Laroche, S. 1992. Linear relationship between the maintenance of hippocampal long-term potentiation and retention of an associative memory. Hippocampus. 2:39-48[Medline].
Fan, S., Stewart, M., Wong, R.K. 1994. Differences in voltage-dependent sodium currents exhibited by superficial and deep layer neurons of pig entorhinal cortex. J. Neurophysiol. 71:1986-1991
Fleidervish, I.A., Gutnick, M.J. 1996. Kinetics of slow inactivation of persistent sodium current in layer V neurons of mouse neocortical slices. J. Neurophysiol. 76:2125-2130
Franceschetti, S., Guatteo, E., Panzica, F., Sancini, G., Wanke, E., Avanzini, G. 1995. Ionic mechanisms underlying burst firing in pyramidal neurons: intracellular study in rat sensorimotor cortex. Brain Res. 696:127-139[Medline].
French, C.R., Gage, P.W. 1985. A threshold sodium current in pyramidal cells in rat hippocampus. Neurosci. Lett. 56:289-293[Medline].
French, C.R., Sah, P., Buckett, K.J., Gage, P.W. 1990. A voltage-dependent persistent sodium current in mammalian hippocampal neurons. J. Gen. Physiol. 95:1139-1157[Abstract].
Greenstein, Y.J., Pavlides, C., Winson, J. 1988. Long-term potentiation in the dentate gyrus is preferentially induced at theta rhythm periodicity. Brain Res. 438:331-334[Medline].
Hamill, O.P., Marty, A., Neher, E., Sakmann, B., Sigworth, F.J. 1981. Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches. Pflügers Arch. 391:85-100[Medline].
Hodgkin, A.L., Huxley, A.F. 1952. A quantitative description of membrane current and its application to conduction and excitation in nerve. J. Physiol. 117:500-544[Medline].
Holmes, J.E., Adey, W.R. 1960. Electrical activity of the entorhinal cortex during conditioned behaviour. Am. J. Physiol. 199:741-744.
Hölscher, C., Anwyl, R., Rowan, M.J. 1997. Stimulation on the positive phase of hippocampal theta rhythm induces long-term potentiation that can be depotentiated by stimulation on the negative phase in area CA1 in vivo. J. Neurosci. 17:6470-6477
Hotson, J.R., Prince, D.A., Schwartzkroin, P.A. 1979. Anomalous inward rectification in hippocampal neurons. J. Neurophysiol. 42:889-895
Huerta, P.T., Lisman, J.E. 1996. Low-frequency stimulation at the troughs of 2-oscillation induces long-term depression of previously potentiated CA1 synapses. J. Neurophysiol. 75:877-884
Huguenard, J.R., Hamill, O.P., Prince, D.A. 1988. Developmental changes in Na+ conductances in rat neocortical neurons: appearance of a slowly inactivating component. J. Neurophysiol. 59:778-795
Jahnsen, H. 1986. Extracellular activation and membrane conductances of neurones in the guinea-pig deep cerebellar nuclei in vitro. J. Physiol. 372:149-168[Abstract].
Jahnsen, H., Llinás, R. 1984. Ionic basis for the electro-responsiveness and oscillatory properties of guinea-pig thalamic neurones in vitro. J. Physiol. 349:227-247[Abstract].
Ju, Y.K., Saint, D.A., Gage, P.W. 1994. Inactivation-resistant channels underlying the persistent sodium current in rat ventricular myocytes. Proc. R. Soc. Lond. B Biol. Sci. 256:163-168[Medline].
Kay, A.R., Miles, R., Wong, R.K. 1986. Intracellular fluoride alters the kinetic properties of calcium currents facilitating the investigation of synaptic events in hippocampal neurons. J. Neurosci. 6:2915-2920[Abstract].
Kay, A.R., Sugimori, M., Llinás, R. 1998. Kinetic and stochastic properties of a persistent sodium current in mature guinea pig cerebellar Purkinje cells. J. Neurophysiol. 80:1167-1179
Klink, R., Alonso, A. 1993. Ionic mechanisms for the subthreshold oscillations and differential electroresponsiveness of medial entorhinal cortex layer II neurons. J. Neurophysiol. 70:144-157
Klink, R., Alonso, A. 1997. Morphological characteristics of layer II projection neurons in the rat medial entorhinal cortex. Hippocampus. 7:571-583[Medline].
Larson, J., Lynch, G. 1986. Induction of synaptic potentiation in hippocampus by patterned stimulation involves two events. Science. 232:985-988[Medline].
Larson, J., Wong, D., Lynch, G. 1986. Patterned stimulation at the theta frequency is optimal for the induction of hippocampal long-term potentiation. Brain Res. 368:347-350[Medline].
Lipowsky, R., Gillessen, T., Alzheimer, C. 1996. Dendritic Na+ channels amplify EPSPs in hippocampal CA1 pyramidal cells. J. Neurophysiol. 76:2181-2191
Llinás, R.R., Alonso, A. 1992. Electrophysiology of the mammillary complex in vitro. I. Tuberomammillary and lateral mammillary neurons. J. Neurophysiol. 68:1307-1320
Llinás, R., Sugimori, M. 1980. Electrophysiological properties of in vitro Purkinje cell somata in mammalian cerebellar slices. J. Physiol. 305:171-195[Abstract].
Ma, J.Y., Catterall, W.A., Scheuer, T. 1997. Persistent sodium currents through brain sodium channels induced by G protein betagamma subunits. Neuron. 19:443-452[Medline].
Magistretti, J., de Curtis, M. 1998. Low-voltage activated T-type calcium currents are differently expressed in superficial and deep layers of guinea pig piriform cortex. J. Neurophysiol. 79:808-816
Magistretti, J., Ragsdale, D.S., Alonso, A. 1999. High conductance sustained single channel activity responsible for the low threshold persistent Na+ current in entorhinal cortex neurons. J. Neurosci. 19:7334-7341
Masukawa, L.M., Hansen, A.J., Shepherd, G. 1991. Distribution of single-channel conductances in cultured rat hippocampal neurons. Cell. Mol. Neurobiol 11:231-243[Medline].
Mitchell, S.J., Ranck, J.B., Jr. 1980. Generation of theta rhythm in medial entorhinal cortex of freely moving rats. Brain Res. 189:49-66[Medline].
Nisenbaum, E.S., Wilson, C.J., Foehring, R.C., Surmeier, D.J. 1996. Isolation and characterization of a persistent potassium current in neostriatal neurons. J. Neurophysiol. 76:1180-1194
Pape, H.C., Driesang, R.B. 1998. Ionic mechanisms of intrinsic oscillations in neurons of the basolateral amygdaloid complex. J. Neurophysiol. 79:217-226
Parri, H.R., Crunelli, V. 1998. Sodium current in rat and cat thalamocortical neurons: role of a non-inactivating component in tonic and burst firing. J. Neurosci. 18:854-867
Pennartz, C.M., Bierlaagh, M.A., Geurtsen, A.M. 1997. Cellular mechanisms underlying spontaneous firing in rat suprachiasmatic nucleus: involvement of a slowly inactivating component of sodium current. J. Neurophysiol. 78:4811-4825.
Ragsdale, D.S., Avoli, M. 1998. Sodium channels as molecular targets for antiepileptic drugs. Brain Res. Brain Res. Rev. 26:16-28[Medline].
Ramón y Cajal, S. 1902. Sobre un ganglio especial de la corteza esfeno-occipital. Trab. del Lab. de invest. Biol. Univ. Madrid. 1:189201.
Sah, P., Gibb, A.J., Gage, P.W. 1988. The sodium current underlying action potentials in guinea pig hippocampal CA1 neurons. J. Gen. Physiol. 91:373-398[Abstract].
Schwindt, P.C., Crill, W.E. 1995. Amplification of synaptic current by persistent sodium conductance in apical dendrite of neocortical neurons. J. Neurophysiol. 74:2220-2224
Segal, M.M. 1994. Endogenous bursts underlie seizurelike activity in solitary excitatory hippocampal neurons in microcultures. J. Neurophysiol. 72:1874-1884
Stafstrom, C.E., Schwindt, P.C., Chubb, M.C., Crill, W.E. 1985. Properties of persistent sodium conductance and calcium conductance of layer V neurons from cat sensorimotor cortex in vitro. J. Neurophysiol. 53:153-170
Stafstrom, C.E., Schwindt, P.C., Crill, W.E. 1982. Negative slope conductance due to a persistent subthreshold sodium current in cat neocortical neurons in vitro. Brain Res. 236:221-226[Medline].
Steward, O., Scoville, S.A. 1976. Cells of origin of entorhinal cortical afferents to the hippocampus and fascia dentata of the rat. J. Comp. Neurol. 169:347-370[Medline].
Stuart, G., Sakmann, B. 1995. Amplification of EPSPs by axosomatic sodium channels in neocortical pyramidal neurons. Neuron. 15:1065-1076[Medline].
Taylor, C.P. 1993. Na+ currents that fail to inactivate. Trends Neurosci 16:455-460[Medline].
Taylor, C.P., Meldrum, B.S. 1995. Na+ channels as targets for neuroprotective drugs. Trends Pharmacol. Sci. 16:309-316[Medline].
Uteshev, V., Stevens, D.R., Haas, H.L. 1995. A persistent sodium current in acutely isolated histaminergic neurons from rat hypothalamus. Neuroscience. 66:143-149[Medline].
van der Linden, S., F.H. Lopes da Silva,. 1998. Comparison of the electrophysiology and morphology of layers III and II neurons of the rat medial entorhinal cortex in vitro. Eur. J. Neurosci. 104:1479-1489.
White, J.A., Alonso, A., Kay, A.R. 1993. A heart-like Na+ current in the medial entorhinal cortex. Neuron 11:1037-1047[Medline].