Introduction

In central neurons, changes in intrinsic neuronal excitability have been shown to occur in parallel with synaptic modifications, thus affecting synergistically synaptic strength and dendritic integration in the post-synaptic neuron1,2. Induction of Long-Term synaptic Potentiation (LTP) is associated with an increase in the firing probability of the postsynaptic neuron3,4,5,6,7, whereas induction of Long-Term synaptic Depression (LTD) is associated with a reduced firing probability in response to the test input4,5,8.

Induction of synaptic plasticity has also been associated with non-synergistic (i.e. homeostatic) modifications of intrinsic excitability9,10. One of the key players in CA1 pyramidal cells is the hyperpolarization-activated cationic h-current (I h), a major determinant of input resistance and intrinsic neuronal excitability11,12. Very large LTP (≈+300%) was found to up-regulate I h to counteract excessive synaptic excitation13,14 whereas large LTD (≈−60%) was found to down-regulate I h to counteract excessive synaptic depression15.

These two homeostatic regulations of I h are not compatible with the synergistic changes in excitability and synaptic strength reported earlier4,5,16. The discrepancy for the LTP side was resolved by showing that h-channel regulation depended on LTP amplitude17. In fact, it was shown in this study that physiological LTP (i.e. +20–50%) produced a decrease in I h seen as an increase in Rin whereas extreme LTP (i.e. +200–300%) produced an increase in I h and the two extrema were linked by a continuum of synergistic and homeostatic plasticity. However, the discrepancy still remained for LTD. We examined whether a similar continuum also exists for homeostatic and synergistic changes in intrinsic neuronal excitability for the LTD side.

We show here that Rin also depends on the magnitude of LTD, with a decrease following induction of moderate LTD but an increase after induction of strong LTD. This dependence of Rin on the magnitude of LTD is abolished by the h-channel blocker ZD-7288. The decrease in Rin (due to an up-regulation of I h) is mediated by NMDA receptor activation whereas the increase in Rin (i.e. due to a down-regulation of I h) is mediated by activation of mGluR1. We show here that induction of LTD in the presence of the NMDA receptor antagonist D-AP5 enhanced neuronal excitability whereas LTD induction in the presence of the mGluR antagonist LY341485 diminished excitability of CA1 pyramidal neurons. We conclude that intrinsic plasticity induced by LTD also describes a continuum between synergistic and homeostatic plasticity in CA1 pyramidal neurons, involving different sets of glutamate receptors.

Results

LTD magnitude determines changes in Rin in CA1 pyramidal neurons

All experiments were performed in the presence of the GABA receptor antagonist PiTx (100 µM). EPSPs were evoked in CA1 pyramidal neurons recorded in whole-cell configuration by stimulating the Schaffer collaterals at 0.1 Hz. After obtaining a stable base line, Long-Term Depression (LTD) of synaptic transmission was induced by stimulation of the Schaffer collaterals at 3 Hz during 3 or 5 min. Input resistance (Rin) measured with large hyperpolarizing current pulses to recruit h-current (−120 pA, 800 ms) was found to be reduced to ~95.7% of the control value (n = 33, t-test p < 0.01) following LTD induction (Fig. 1A). But more interestingly a negative correlation was observed between the normalized Rin and the synaptic change (y = −0.198x-91.1; r = 0.36; p < 0.05; Fig. 1B). This negative correlation was further confirmed by the difference in the mean Rin change observed after 3 or 5 min at 3 Hz (after 3 min at 3 Hz: 93 ± 3%, n = 11 for a mean EPSP change of −9 ± 4%; after 5 min at 3 Hz: 97 ± 2% for an EPSP change of −31 ± 3%; Fig. 1B).

Figure 1
figure 1

LTD magnitude determines changes in Rin. (A) Time-courses of EPSP slope (top) and Rin (bottom) in a single experiment showing a reduction in Rin following induction of LTD. Scale bars: top, 2 mV & 20 ms; bottom, 5 mV & 200 ms. (B) Plot of Rin as a function of EPSP change induced by 3 Hz stimulation for 3 or 5 minutes. Note the negative correlation (y = 0.198x + 91.096, r = 0.36; p < 0.01). (C) Top, EPSP slope time course pooled over sixteen experiments. Arrows indicate 3 Hz stimulation episodes. Representative EPSP traces in control (a), after the first stimulation episode (b) and after the third (c). Scale bars: 1 mV, 10 ms. Bottom, time course of apparent input resistance (Rin) after each stimulation episode. Rin is reduced after the first 3 Hz stimulation (a, b) and increased for the third (c). Stars indicate statistical significance (p < 0.05) Top, representative Rin traces in control (a), after the first stimulation episode (b) and after the third (c). Scale bars: 10 mV, 100 ms. (D) Normalized Rin as a function of EPSP change for each cell. Note that correlations for each cell are all negatively oriented. (E) Normalized Rin changes versus normalized LTD level for each stimulation episode. A significant linear negative correlation was observed (y = −0.414x + 85.487; r = 0.66; p < 0.001).

To confirm the correlation observed with one train of 3 Hz stimulation, synaptic depressions of larger magnitudes were induced by repeated episodes of 3 Hz stimulation with ten minutes intervals. A progressive decrease in EPSP slope and a parallel changes in apparent Rin were observed (Fig. 1C). While Rin decreased after the first stimulation episode (see also Fig. 1A), it progressively increased after each stimulation episode (Fig. 1C). The analysis of the trajectories of individual cells showed in all cases an anti-correlation (Fig. 1D). The plot of Rin versus EPSP change revealed a significant anti-correlation (r = 0.66; p < 0.001; Fig. 1E). Rin was reduced to 90 ± 3% for moderate LTD (<20%) but increased to 116 ± 4% for large LTD (>50%; Fig. 1E). As previously reported for LTP17, modulation of Rin was not associated with significant change in Vm following induction of LTD (−62.3 ± 0.9 mV in control and −61.8 ± 0.6 mV after the 3rd episode of 3 Hz stimulation, p > 0.1; Supplementary Figure 1). In conclusion, the magnitude of LTD determines the polarity of Rin change in CA1 pyramidal neurons.

Temporal stability of synaptic transmission and Rin

In order to eliminate any non-specific Rin changes, we repeated the same protocol with 0.1 Hz stimulation to test the temporal stability of synaptic strength and Rin in CA1 pyramidal neurons. No changes in EPSP slope were observed after 1 (Fig. 2A & B) or several repetitive episodes of 0.1 Hz stimulation (−4 ± 3% of control EPSP slope; Fig. 2C). Furthermore Rin remained unchanged throughout the experiment (103 ± 1%; Fig. 2C). Finally, no linear correlation was observed between normalized Rin and EPSP slope at the level of individual cells (Fig. 2D) or all taken together (r = 0.05; p > 0.05; Fig. 2E).

Figure 2
figure 2

Temporal stability. (A) Time-courses of EPSP slope (top) and Rin (bottom) after 0.1 Hz stimulation. Scale bars; top, 2 mV & 20 ms; bottom, 10 mV & 200 ms. (B) Plot of Rin versus EPSP change. No correlation was observed (y = 0.021x + 101.31, r = 0.10). (C) Top, EPSP slope time-course pooled over ten experiments. Arrows indicate 0.1 Hz stimulation episodes. Bottom, time-course of Rin. Representative Rin traces in control (a), and after stimulation episodes (b, c & d). Scale bars: 10 mV, 100 ms. No statistically significant changes were observed. (D) Correlation between Rin and EPSP change for each cell after 0.1 Hz stimulation. (E) Normalized Rin versus normalized EPSP change for all episodes of 0.1 Hz stimulation (blue dots). Compared to the control condition (grey dots, see Fig. 1B), no linear correlation was observed (y = −0.027x + 102.79; r = 0.05, p > 0.05).

Regulation of I h is responsible for changes in Rin

Rin is mainly governed by the h-current in CA1 pyramidal neurons. We therefore tested the role of I h in the observed changes in Rin. We repeated the same protocol in the presence of the pharmacological blocker of h-channels ZD-7288 (1 µM). This concentration of ZD-7288 has been shown to block I h without altering excitatory synaptic transmission18. In the presence of ZD-7288, stimulation of the Schaffer collaterals at 3 Hz still induced LTD (−23 ± 9%, n = 8; Fig. 3A) but Rin remained unchanged (98 ± 2% of control Rin, n = 8, Fig. 3A & B). Similarly, no change in Rin occurred following induction of incremental LTD by repeated low frequency stimulation at 3 Hz (Fig. 3C & D). No linear correlation was found between LTD magnitude and Rin changes in the presence of ZD-7288 (r = 0.18; p > 0.05; Fig. 3E). These results indicate that I h is directly involved in the bidirectional regulation of Rin following induction of LTD.

Figure 3
figure 3

Blockade of h-channels ZD-7288 impairs regulation of Rin. (A) Time-courses of EPSP slope (top) and Rin (bottom) following 3 Hz stimulation for 5 min in the presence of ZD-72288. Scale bars: top, 2 mV & 20 ms; bottom, 5 mV & 200 ms. (B) Plot of Rin as a function of EPSP change. Note the lack of correlation (y = 0.072 + 100.61; r = 0.31). (C) Top, EPSP slope time-course pooled over nine experiments with ZD7288 (1 µM) in the bath. Note that ZD7288 does not alter synaptic plasticity. Bottom, Rin time course. Representative traces in control (a), after the first stimulation episode (b) and after the third (c). Scale bars: 10 mV, 100 ms. No Rin changes were observed. (D) Negative correlations between Rin changes and LTD levels for each cell in the presence of ZD-7288. (E) Normalized Rin versus EPSP change levels in control (grey dots) and in the presence of ZD-7288 (blue dots). No significant linear correlation was observed (y = 0.05x + 100.6; r = 0.18; p > 0.05).

To further confirm the implication of h-channels, the sag produced by activation of I h was analysed. The sag was found to decrease after the 1st episode of 3 Hz stimulation and remained reduced by ~15% thereafter (Supplementary Figure 2A). Interestingly, a significant correlation between the normalized sag change and the magnitude of LTD was observed (Supplementary Figure 2B). But, surprisingly, no increase in the sag amplitude was observed following induction of LTD. We thus developed a simplified model of hippocampal neuron in which h conductance (Gh) increased from 0 to 10 nS (Supplementary Figure 2C). Importantly, while Rin provided a good description of changes in Gh, the sag increased when Gh increased in the 0–2 nS range but it was found to decrease when Gh increased in the 2–10 nS range. This result indicates that the sag is a not an index appropriate for evaluating activity-dependent regulation of h-channels.

Reduction of Rin depends on NMDAR

Induction of LTD requires both N-methyl-D-aspartate receptor (NMDAR)19,20,21 and/or metabotropic glutamate receptor (mGluR)15,22,23,24. To dissect the role of NMDAR in the regulation in Rin, we applied the specific antagonist D-AP5 (50 µM) in the bath. In the presence of D-AP5, the magnitude of LTD induced by the first episode of 3 Hz stimulation was found to be reduced (93 ± 7% of control EPSP slope, n = 6 versus 72 ± 4%, n = 16, in control condition; Fig. 4A & B). In contrast with what was observed in control conditions, Rin was increased in 5 out of 6 cells after the first stimulation episode (110 ± 4% of control Rin, Fig. 4A & B).

Figure 4
figure 4

Blockade of NMDAR prevents decrease in Rin. (A) Time-courses of EPSP slope (top) and Rin (bottom) following 3 Hz stimulation for 5 min in the presence of D-AP5. Scale bars: top, 2 mV & 20 ms; bottom, 5 mV & 200 ms. (B) Plot of Rin as a function of EPSP change. (C) Top, EPSP slope time-course induced by repetitive 3 Hz stimulation in the presence of 50 µM D-AP5 in the bath. Bottom, corresponding Rin time-course after each stimulation episode. Rin is increased from the first episode to the last one. Stars indicate statistical significance (p < 0.05). Representative Rin traces in control (a), after the first stimulation episode (b) and after the third (c). Scale bars: 10 mV, 100 ms. (D) Individual linear correlations between Rin changes and EPSP slope modifications induced by 3 Hz stimulation in the presence of D-AP5. (E) Normalized Rin changes versus normalized LTD size for control condition (grey dots) and in presence of D-AP5 (purple dots). Note the shift to higher values of Rin when D-AP5 is present during the plasticity induction (correlation: y = −0.34x + 108.7, r = 0.47; p < 0.05).

The following stimulation episodes produced, however, comparable levels of LTD and Rin was found to augment up to 130% after the last episode of 3 Hz stimulation (128 ± 1%; Fig. 4C & D). Compared to the control situation, the plot of normalized Rin against EPSP change in D-AP5 indicates an upward shift of the linear anti-correlation (r = 0.47; p < 0.05; Fig. 4E). These data suggest that NMDARs are implicated in the down-regulation of Rin observed for moderate LTD. The remaining increase in Rin might result from the stimulation of mGluRs.

Enhancement of Rin depends on mGluR1

We next tested whether mGluRs were implicated in the up-regulation of Rin. We first applied the mGluR1/5 agonist DHPG (50–100 µM) during 5 min23. DHPG induced synaptic LTD (74 ± 6% of the control EPSP slope, n = 12; Fig. 5A). Interestingly, this mGluR-induced LTD was associated with a long-lasting increase of Rin (116 ± 4% of the control Rin; Fig. 5A & B).

Figure 5
figure 5

Blockade of mGluRs prevents increase in Rin. (A) Top, time-course of synaptic changes induced by bath application of 50–100 µM DHPG during 5 minutes (pooled data from 12 cells). Upper traces, representative examples of EPSPs before and after DHPG. Scale bars: top, 2 mV & 20 ms. Bottom, normalized Rin changes induced by DHPG. Upper traces, representative traces. Scale bars: 10 mV, 100 ms. (B) Plot of Rin as a function of EPSP changes. No correlation is observed (r = 0.06). (C) Top, time-course of synaptic changes induced by 3 Hz stimulation in the presence of 100 µM LY341495. Upper traces, representative examples of EPSPs before and after 3 Hz stimulation. Scale bars: 5 mV & 30 ms. Bottom, normalized Rin changes induced by 3 Hz stimulation in the presence of 100 µM LY341495. Scale bars: 10 mV & 50 ms. (D) Plot, of Rin as a function of EPSP changes (linear correlation, y = −0.203x + 85.1, r = 0.64; p < 0.01).

Next, we induced LTD with 3 Hz stimulation of the Schaffer collaterals in the presence of the broad spectrum mGluR antagonist, LY341495 (100 µM). In this condition, synaptic LTD was still induced by 3 Hz stimulation (68 ± 12% of control EPSP slope, n = 7 after the 3rd episode of stimulation; Fig. 5C) but importantly Rin was found to be reduced (to 85 ± 7% of control Rin, n = 7 after the 3rd episode of 3 Hz stimulation; Fig. 5C). Furthermore, the plot of normalized Rin against EPSP change in LY341495 was found to follow a linear anti-correlation (r = 0.64; Fig. 5D). In contrast, Rin was found to be still enhanced when LTD was induced in the presence of the specific mGluR5 antagonist, MPEP (10 µM; Supplementary Figure 3), suggesting that mGluR1 and not mGluR5 is responsible for the increase in Rin.

In conclusion, the stimulation of NMDARs induces a decrease in Rin (i.e. up-regulation of I h) whereas the stimulation of mGluR1 is responsible for an increase in Rin (i.e. down-regulation of I h).

Changes in excitability associated with LTD

Next, we tested whether these changes in Rin were associated with changes in intrinsic excitability following induction of LTD. To better dissect the implication of bidirectional changes in Rin we pharmacologically isolated the mGluR- and NMDAR-mediated component of Rin changes associated with LTD induced by 3 Hz stimulation of the Schaffer collateral for 10 min with either D-AP5 or LY341495 in the bath. Consistent with the increase in Rin after 3 Hz stimulation in the presence of D-AP5, excitability was found to be increased following LTD induction in D-AP5 (Fig. 6A & B). Conversely, in the presence of LY341495 excitability was found to be significantly reduced following induction of LTD (Fig. 6C & D).

Figure 6
figure 6

LTD is associated with bidirectional changes in excitability. (A) & (B) LTD induced in the presence of D-AP5 is associated with an increased excitability. (A) Top, time-course of LTD induced by 3 Hz stimulation of the glutamatergic inputs for 10 minutes in the presence of 50 µM D-AP5. Middle, time-course of the increased Rin. Bottom, time-course of Vm. (B) Top, representative example of firing induced by a current step of 75 pA before and after LTD induction with 3 Hz stimulation in the presence of D-AP5. Scale bars: 20 mV & 100 ms. Bottom, input-output curves before (black) and after (purple) LTD induction. Stars indicate significant change (p < 0.05). (C) & (D) LTD induced in the presence of LY341495 is associated with a decreased excitability. (C) Top, time-course of LTD induced by 3 Hz stimulation of the glutamatergic inputs for 10 minutes in the presence of 100 µM LY341495. Middle, time-course of the increased Rin. Bottom, time-course of Vm. (D) Top, representative example of firing induced by a current step of 75 pA before and after LTD induction with 3 Hz stimulation in the presence of LY341495. Scale bars: 20 mV & 200 ms. Bottom, input-output curves before (black) and after (dark green) LTD induction. Stars indicate significant change (p < 0.05).

In conclusion, LTD induced with 3 Hz stimulation activates NMDAR and mGluR that in turn regulate both Rin and intrinsic excitability in CA1 pyramidal cells.

Discussion

We show here that in CA1 pyramidal neurons, LTD magnitude determines the changes in input resistance (Rin) and hence, the direction of I h regulation. Moderate LTD induces an increase in I h (seen as a decrease in Rin) while strong LTD results in a decrease of I h (i.e. an increase in Rin). LTD induction in the presence of the NMDA receptor antagonist D-AP5 suppressed the reduction in Rin, suggesting that it is mediated by NMDA receptors (Fig. 7A). In contrast, LTD induced by activation of mGluR1/5 with DHPG is associated with an increase in Rin (i.e. decrease in I h). Furthermore, LTD induced in the presence of the mGluR antagonist LY341495 suppressed the increase in Rin and left it reduced by ~15%. However, no reduction in Rin was observed when LTD was induced in the presence of the mGluR5 antagonist, MPEP, suggesting that activation of mGluR1 and not mGluR5 triggers an increase in Rin (Fig. 7A). Finally, excitability was found to be increased when LTD was induced in the presence of D-AP5 whereas it was reduced when LTD was induced in the presence of LY341495. These results suggest that changes in intrinsic excitability follow a single learning rule linking synergistic changes induced by synaptic modification in the physiological range to homeostatic changes induced by large synaptic modification (Fig. 7B). Thus, our results bring strong evidence for fast compensatory processes in Hebbian plasticity25.

Figure 7
figure 7

The continuum rule: linking synergistic plasticity with homeostatic plasticity. (A) Summary of the glutamate-receptor induced changes in Rin as a function of synaptic changes. Green curve illustrates the NMDA-dependent change in Rin. Red curve shows the mGluR-dependent change in Rin. Black curve illustrates the sum of the green and red curves. (B) Unifying rule for intrinsic plasticity. In a physiological range (defined by the green square), modulation of neuronal activity results in a conjugated modification in intrinsic excitability (synergistic). Out of this range, persisting increases (right) or decreases (left) in synaptic efficacy induce compensatory changes in intrinsic excitability (homeostatic). Adapted from Campanac et al. 2008 and from the present study.

LTD induces NMDAR-dependent up-regulation of I h

Our results show that a single episode of 3 Hz stimulation for 3–5 min decreases Rin in CA1 pyramidal neurons. Blocking I h with ZD-7288 prevents changes in Rin following 3 Hz stimulation, indicating that Rin is decreased through an increase of I h. This component could be isolated by blocking mGluRs with LY341495. Because a reduction of Rin causes a decrease in intrinsic excitability12, this regulation is functionally synergic to the long-lasting depression of synaptic transmission. Such a Hebbian regulation of neuronal excitability has already been reported following LTD induction in CA1 neurons4,5,16, but this had not been reported in previous studies in which very large LTD was induced15. We show that in the presence of the NMDA receptor antagonist D-AP5, no decrease in Rin was observed. Rather, Rin was enhanced, indicating that stimulation of non-NMDA receptors triggers the down-regulation of h-channel activity.

From Hebbian to homeostatic

Increasing LTD magnitude through repetition of 3 Hz stimulation episodes revealed that Rin could be regulated in the other direction. In fact, after the 3rd or 4th stimulation episode, large LTD was induced and Rin was found to be increased. This increase in Rin was prevented by the presence of ZD-7288 in the bath indicating that it was due to the down-regulation of I h. A reduction of I h has already been demonstrated in CA1 pyramidal cells following LTD induction of large magnitude15. This regulation is supposed to counteract the reduction in synaptic efficiency in a homeostatic manner. In fact, other experimental studies have shown that sensory deprivation or chronic inactivity leads to the down-regulation of I h in pyramidal neurons of the barrel cortex26 or the hippocampus27. The down-regulation of I h could be induced by stimulation of group I mGluR. We indeed show that DHPG induced LTD associated with an increase in Rin. In addition, we show that Rin diminished when LTD was induced by 3 Hz stimulation in the presence of the broad spectrum mGluR antagonist LY341495 but not in the presence of the specific mGluR5 antagonist MPEP, suggesting that activation of mGluR1 mediates the homeostatic increase in Rin. These results are consistent with the mGluR-dependent increase in both Rin and intrinsic excitability reported by Brager & Johnson (2007) and suggest that two sets of receptors might be able to up-regulate and down-regulate h-channel activity depending on the magnitude of synaptic modification.

mGluR5 has been shown to mediate enhanced excitability induced by stimulation of glutamatergic inputs in L5 pyramidal neurons28 and in hippocampal parvalbumin-positive basket cells7. In these cases, the changes in excitability were synergistic to synaptic modification. Here, we show that stimulation of mGluR1 appears as the main factor responsible for the switch of synergistic to homeostatic regulation of intrinsic excitability.

Pharmacological11 or activity-dependent15 reduction in h-channel activity is usually associated with a hyperpolarizing shift in membrane potential. No change in membrane potential was, however, observed in the experiments reported here (see also Campanac et al., 2008). The apparent discrepancy with the results of Brager & Johnston (2007) might be due to the much larger increase in input resistance obtained in this study following LTD induction (+100% versus + 20% in our case).

Compared to Hebbian plasticity, homeostatic regulation is generally considered as a slow process. In fact, most of the regulations of intrinsic excitability reported so far have been obtained with manipulating network activity for 2–3 days10,27,29,30,31. Here, we report induction of homeostatic plasticity of intrinsic excitability that can be induced in parallel with Hebbian synaptic plasticity on a much faster time-scale. Such rapid compensatory processes are thought to be necessary to stabilize neuronal activity32,33.

Bidirectional regulation of I h has already been revealed following LTP induction in CA1 pyramidal neurons17. The present study not only reconciles contradictive experimental results4,15 but it also shows that Hebbian and homeostatic regulations of I h occur in the same neuron after LTD induction and follow a single rule establishing a continuum between functionally opposite forms of intrinsic plasticity that target h-channels (Fig. 7B)34.

Mechanisms of h-channels regulation

The existence of a learning rule linking synergistic and homeostatic changes implies multiple modes of h-channel regulation. Although further experimental investigations will be required, the mechanisms of molecular regulation of h-channels are multiple35. Activity of h-channels can be regulated by a change in their density (i.e. by insertion or removal of HCN subunits), by a change in the distribution of h-channels at the surface of the neuron36 or by changes in their sensitivity to cyclic nucleotides37. Trip8b (Tetratricopeptide-Repeat containing Rab8b-interacting protein) has been identified as an important binding partner of HCN38. Interestingly, Trip8b undergo alternative splicing and its isoforms have been demonstrated to differently affect I h density39,40 and the sensitivity of h-channels to cyclic AMP40,41. In fact, while most isoforms of Trip8b enhance expression of dendritic HCN subunits39,40, some Trip8b isoforms, however, suppress HCN subunit expression42. As dendrites are able to locally translate mRNA following LTD43 and promote alternative splicing44, Trip8b isoforms offer an attractive mechanism to explain the bidirectional regulation of I h. Interestingly, it has recently been shown that h-channel upregulation that normally occurs after induction of large LTP13 is absent in Trip8b knock-out mice45. Similar experiments should be conducted on the LTD side.

A remaining question is: what is the molecular link between activation of NMDAR/mGluR1 and the regulation of h-channels? The activation of different protein kinases such as Ca2+/CaMKII or PKC results in the modulation of h-channel activity in response to different patterns of neuronal activity13,15,46 but precise data on the regulation of Trip8b by either NMDAR or mGluR1 through Ca2+/CaMKII or PKC are still missing today.

Experimental procedures

Slice preparation

Hippocampal slices were obtained from 14- to 20- day-old rats according to institutional guidelines for the care and use of laboratory animals (Directive 86/609/EEC and French National Research Council) an approved by the local health authority (# D1305508, Préfecture des Bouches-du-Rhône, Marseille). Rats were deeply anaesthetized with chloral hydrate (intraperitoneal 400 mg/kg) and killed by decapitation. Slices (350 µm) were cut in a solution containing a reduced concentration of sodium (in mM: 280 sucrose, 26 NaHCO3, 10 D-glucose, 1.3 KCl, 1 CaCl2, and 10 MgCl2) on a vibratome (Leica VT1000S) and were maintained for 1 h at room temperature in oxygenated (95% O2/5% CO2) Artificial Cerebro-Spinal Fluid (ACSF; in mM: 125 NaCl, 2.5 KCl, 0.8 NaH2PO4, 26 NaHCO3, 3 CaCl2, 2 MgCl2, and 10 D-glucose) with foetal bovine serum (4%). For recording, each slice was transferred to a temperature-controlled (30 °C) chamber with oxygenated ACSF. GABAA channels were blocked with picrotoxin (PiTX, 100 µM) and the CA3 area was surgically removed.

Electrophysiology

Neurons were identified with an Olympus BX 50WI microscope using infrared video microscopy and Differential Interference Contrast (DIC) × 60 optics.

Whole-cell recordings were made from CA1 pyramidal neurons with electrodes filled with a solution containing the following (in mM): 120 K-gluconate, 20 KCl, 10 HEPES, 0.5 EGTA, 2 MgCl2 6H2O, and 2 Na2ATP. Stimulating pipettes filled with extracellular saline were placed in the stratum radiatum to stimulate the Schaffer collaterals.

In control and test conditions, Excitatory Post-Synaptic Potentials (EPSPs) were elicited at 0.1 Hz by a digital stimulator (NEURO DATA PG4000, Instruments corp.) or by pCLAMP (Molecular devices). LTD was induced with continuous shocks delivered at 3 Hz during 5 min. Apparent input resistance was tested by current injection (−120 pA; 800 ms). Series resistance was monitored throughout the recording and only experiments with stable resistance were kept (changes <10%).

Intrinsic excitability has been measured before and after 3 Hz stimulation with input-output curves consisting in plotting spike number in response to incrementing steps of current pulses27,28. Changes in membrane potential (Vm) were measured in the absence of any holding current.

Drugs

Drugs were bath applied. Picrotoxin (PiTx) was purchased from Sigma. [4-(N-ethyl-N-phenylamino)-1,2-dimethyl-6-(methylamino) pyrimidinium chloride] (ZD-7288), D-(-)-2-Amino-5-phosphonopentanoic acid (D-AP5), 3,5-Dihydroxyphenylglycine (DHPG), (2 S)-2-Amino-2-[(1 S,2 S)-2-carboxycycloprop-1-yl]-3-(xanth-9-yl) propanoic acid (LY341495) and 2-methyl-6-(phenylethylyl)pyridine (MPEP) were purchased from Tocris Bioscience.

Data acquisition and analysis

Recordings were obtained using an Axoclamp-2B (Molecular Devices) or a MultiClamp 700B (Molecular Devices) amplifier and pClamp10 software. Data were sampled at 10 kHz, filtered at 3 kHz, and digitized by a Digidata1322A (Molecular Devices). All data analyses were performed with custom written software in Igor Pro 6 (Wavemetrics).

Apparent input resistance was determined by the subtraction of the steady-state voltage change during hyperpolarizing current injection from the baseline.

Pooled data are presented as mean ± SEM. Statistical comparisons were made using Wilcoxon or Mann-Whitney test as appropriate with Sigma Plot software. Statistical correlations were tested using Spearman test. Data were considered as significant when p < 0.05.

Modelling

A simple Hodgkin-Huxley-type model of hippocampal neuron was developed under LabView (LabView 7). The model had no dimension and included only the h conductance with parameters taken from Campanac et al., 2008. The leak resistance was set to 1 GΩ. The h-current was given by:

$${{\rm{I}}}_{{\rm{h}}}={{\rm{G}}}_{{\rm{h}}}\,\ast \,({{\rm{V}}}_{{\rm{m}}}-{{\rm{E}}}_{{\rm{h}}})$$
(1)

with Vm the membrane potential, Eh = −37.7 mV, and the h-conductance given by the following equation:

$${{\rm{G}}}_{{\rm{h}}}={{\rm{G}}}_{h,\max }\,\ast \,{\rm{n}}$$
(2)

The activation and deactivation time constants were determined by fitting experimental data from Campanac et al. (2008). The following differential equation was solved,

$${\rm{dn}}({\rm{V}},{\rm{t}})/{\rm{dt}}={{\rm{\alpha }}}_{{\rm{n}}}({\rm{V}})\,\ast \,[1-{\rm{n}}({\rm{V}},{\rm{t}})]$$
(3)

This equation corresponds to:

$${\rm{d}}({\rm{V}},{\rm{t}})/{\rm{dt}}=[{{\rm{n}}}_{\infty }({\rm{V}})-{\rm{n}}({\rm{V}},{\rm{t}})]/{\rm{\tau }}({\rm{V}})$$
(4)
$${{\rm{n}}}_{\infty }({\rm{V}})={{\rm{\alpha }}}_{{\rm{n}}}({\rm{V}})/[{{\rm{\alpha }}}_{{\rm{n}}}({\rm{V}})+{{\rm{\beta }}}_{{\rm{n}}}({\rm{V}})]$$
(5)
$${\rm{\tau }}({\rm{V}})=1\,{\rm{for}}\,{\rm{V}} > -30\,{\rm{mV}}$$
(6)

otherwise

$${\rm{\tau }}({\rm{V}})=1/[{{\rm{\alpha }}}_{{\rm{n}}}({\rm{V}})+{{\rm{\beta }}}_{{\rm{n}}}({\rm{V}})]$$
(7)
$${\alpha }_{{\rm{n}}}=0.0204/(1+\exp [({\rm{V}}+98.68)/13.24])$$
(8)
$${\beta }_{{\rm{n}}}=0.0176/(1+\exp [-({\rm{V}}+57.96)/13.2])$$
(9)

where n(V) is the steady-state activation parameter and τ(V) the activation time constant.