Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2001 Dec 1;537(Pt 2):421–430. doi: 10.1111/j.1469-7793.2001.00421.x

A characterisation of long-term depression induced by metabotropic glutamate receptor activation in the rat hippocampus in vitro

Stephen M Fitzjohn 1, Mary J Palmer 1, Jolyon E R May 1, Anne Neeson 1, Stephen A C Morris 1, Graham L Collingridge 1
PMCID: PMC2278956  PMID: 11731575

Abstract

  1. In the CA1 region of hippocampal slices prepared from juvenile (12- to 18-day-old) rats, activation of group I metabotropic l-glutamate (mGlu) receptors by the specific agonist (RS)-3,5-dihydroxyphenylglycine (DHPG) induces a form of long-term depression (LTD) of excitatory synaptic transmission.

  2. We have used a variety of electrophysiological techniques applied to CA1 neurones in hippocampal slices and from pyramidal cells in dissociated hippocampal cultures to investigate the Ca2+ dependence and locus of expression of DHPG-induced LTD.

  3. In patch-clamp experiments from hippocampal slices, bath application of DHPG induced a depression of synaptically evoked responses that persisted for the duration of the recording (up to 2 h after commencing washout of DHPG) in 27 of 29 neurones investigated.

  4. DHPG-induced LTD was associated with an increase in both the paired-pulse facilitation ratio and the coefficient of variation of EPSCs.

  5. Using dendritic recording, there was a decrease in EPSC success rate (number of trials that elicited a detectable response) but no change in potency (mean EPSC amplitude excluding failures) associated with DHPG-induced LTD.

  6. In experiments using dissociated hippocampal cultures, application of DHPG elicited a persistent decrease in the frequency of tetrodotoxin-resistant miniature EPSCs but no change in the amplitude of such events.

  7. DHPG-induced LTD was not blocked by intracellular application of the calcium chelator BAPTA. It was also unaffected when intracellular calcium stores were depleted by perfusion with thapsigargin. Furthermore, when synaptic transmission was blocked by perfusing with Ca2+-free medium, DHPG application reliably induced LTD.

  8. These data suggest that DHPG-induced LTD is Ca2+ independent and is expressed presynaptically.


The metabotropic l-glutamate (mGlu) receptors are G-protein coupled receptors that have been implicated in a variety of brain functions and pathological conditions (Conn & Pin, 1997; Anwyl, 1999). There are currently eight known subtypes of mGlu receptors, which are divided into three groups based on sequence homology, pharmacology and signal transduction mechanisms. There is evidence that mGlu receptors have a role in the induction of various forms of synaptic plasticity such as long-term potentiation and long-term depression (LTD) (Anwyl, 1999; Bortolotto et al. 1999).

Synaptic activation of mGlu receptors has been shown to be involved in LTD induction under a variety of experimental conditions (Stanton et al. 1991; Bashir et al. 1993; Bolshakov & Siegelbaum, 1994; Oliet et al. 1997; Fitzjohn et al. 1998a; Kemp & Bashir, 1999; Huber et al. 2000). However, mGlu receptor-dependent LTD often co-exists at the same synapses with NMDA receptor-dependent LTD (e.g. Oliet et al. 1997). One strategy to selectively activate mGlu receptor-dependent LTD is to apply a specific mGlu receptor agonist. The broad spectrum mGlu receptor agonist 1-aminocyclopentane-1,3-dicarboxylic acid (ACPD) has been shown to induce LTD in both the dentate gyrus (O'Mara et al. 1995) and CA1 region of the hippocampus (Bolshakov & Siegelbaum, 1994; Overstreet et al. 1997). Although several studies have failed to observe LTD induced by ACPD (Baskys & Malenka, 1991; Vignes et al. 1995; Harvey et al. 1996; Palmer et al. 1997), a more reproducible form of LTD induced by the group I-specific mGlu receptor agonist (RS)-3,5-dihydroxyphenylglycine (DHPG) has been reported recently (e.g. Palmer et al. 1997; Fitzjohn et al. 1998a, 1999; Camodeca et al. 1999; Schnabel et al. 1999; Tan et al. 1999; Wetzel et al. 1999; Huber et al. 2000). The ability to chemically induce LTD should help elucidate the molecular basis of this form of synaptic plasticity, as has been the case for NMDA receptor-dependent LTD (Lee et al. 1998).

The mechanism of expression of DHPG-induced LTD is unknown. Although electrically induced mGlu receptor-dependent LTD has been shown to be expressed presynaptically (Oliet et al. 1997), a recent report suggested that DHPG-induced LTD is expressed postsynaptically (Snyder et al. 2000). In the present study we have utilised electrophysiological recording of synaptic transmission in the hippocampus to investigate the locus of expression of this form of synaptic plasticity. Some of this work has been presented previously in abstract form (Fitzjohn et al. 1996, 1998b; Palmer & Collingridge, 1999).

METHODS

Hippocampal slices

Young (12- to 18-day-old) Wistar rats were killed by decapitation and 400 μm transverse hippocampal slices prepared. Slices were maintained in artificial cerebrospinal fluid (aCSF) that comprised (mm): 124 NaCl, 3 KCl, 26 NaHCO3, 1.4 NaH2PO4, 1 MgSO4, 2 CaCl2, 10 d-glucose, bubbled with 95 % O2-5 % CO2, pH 7.4.

Somatic recording

Slices were maintained in a submerged recording chamber at room temperature (20-24 °C) perfused with aCSF at 2 ml min−1. Whole-cell recordings were made blind from stratum pyramidale in area CA1. Patch electrodes had a resistance of 4-7 MΩ and the intracellular solution comprised (mm): 130 caesium methane sulphonate, 5 Hepes, 5 QX-314, 1 NaCl, 1 MgCl2, 4 Mg-ATP, 0.3 Na2-GTP, 0.05 EGTA, 0.035 CaCl2, pH 7.35. In experiments using the calcium chelator BAPTA, EGTA was replaced with 10 mm BAPTA free acid and the caesium methane sulphonate concentration was reduced to maintain osmolarity. Cells were voltage clamped at -80 mV and EPSCs evoked by electrical stimulation of the Schaffer collateral-commissural pathway at a frequency of 0.066 Hz. Inhibitory postsynaptic potentials were blocked by inclusion of 50 μm picrotoxin in the aCSF (to block GABAA receptors) and the use of Cs+ and QX314 in the intracellular solution (to block GABAB receptor-mediated activation of K+ channels). Recordings were made using an Axopatch 200B amplifier. Data were filtered at 5 kHz, digitised at 10 kHz, and collected and analysed both on- and off-line using the LTP program (Anderson & Collingridge, 2001; http://www.ltp-program.com). Two forms of analysis, paired-pulse facilitation and coefficient of variation, were performed to investigate the locus of expression of DHPG-induced LTD.

Extracellular recording

Extracellular recordings of field EPSPs (fEPSPs) were performed from the stratum radiatum of area CA1 of hippocampal slices using a glass microelectrode filled with 4 m NaCl (resistance 1-3 MΩ). Slices were maintained in a submerged recording chamber and perfused at a rate of 2 ml min−1 with aCSF at room temperature. Responses were evoked by stimulation of the Schaffer collateral-commissural pathway at a frequency of 0.033 Hz. Stimulation strength was set so as to give a slope value 50 % of that where a population spike was first observed. Recordings were made using an Axopatch 200B amplifier. Data were filtered at 5 kHz, digitised at 10 kHz and recorded and analysed both on- and off-line using LTP software. The slope of the rising phase of fEPSPs (20-80 % of peak response amplitude) was used as a measure of synaptic transmission.

Dendritic recording

Dendritic recordings were used for a failures analysis since there is less dendritic filtering and hence small EPSCs are more readily distinguished from true failures (Isaac et al. 1998). Slices were maintained at room temperature in a submerged recording chamber, perfused at a rate of 2 ml min−1 with aCSF of composition as above with the addition of ascorbic acid (2 mm) and increased glucose (15 mm). Whole-cell recordings were obtained from the proximal apical dendrites of CA1 pyramidal cells under visual control using infrared illumination and differential interference contrast optics, as described previously (Isaac et al. 1998). Patch electrodes of resistance 7-9 MΩ were filled with an intracellular solution of composition (mm): 130 caesium methane sulphonate, 10 Hepes, 5 QX-314, 8 NaCl, 0.5 EGTA, 4 Mg-ATP, 0.3 Na2-GTP, pH 7.25. Recordings were obtained from apical dendrites between 30 and 50 μm from soma. Cells were voltage clamped at -70 mV and the Schaffer collateral- commissural pathway stimulated at a frequency of 0.5 Hz using a fine monopolar electrode, placed 30-50 μm from the recording site. All experiments were carried out in the presence of 50 μm picrotoxin. Minimal stimulation was used to evoke a mixture of EPSCs and failures of synaptic transmission, with the stimulus intensity set to produce a success rate of around 70 %. Under the conditions of the present experiments, fibre failures do not contribute significantly to the synaptic failures (Isaac et al. 1998). Recordings were made using an Axopatch-1B amplifier. Data were filtered at 5 kHz, digitised at 10 kHz, and collected and analysed using the LTP program. A failures analysis was performed to investigate the locus of DHPG-induced LTD.

Paired-pulse facilitation

To measure paired-pulse facilitation (PPF), two stimuli were delivered with an inter-stimulus interval of 50 ms in place of single stimuli for two 1 min periods during the baseline, for 1 min during DHPG application, and for 1 min every 5 min following washout of DHPG. The paired-pulse ratio (PPR) was calculated as the ratio of the amplitude of the second EPSC to the first EPSC. To ensure that the measurement of the peak amplitude of the second EPSC was not contaminated by residual current from the first EPSC, this residual component was first removed by subtracting the average response to a single stimulation obtained in the minute prior to recording PPF from the average response to two stimuli.

Coefficient of variation

The coefficient of variation (CV) was calculated as described by Kullmann (1994). The mean and standard deviation were calculated for the EPSC amplitudes recorded during successive 5 min epochs (s.d.EPSC and MeanEPSC). The standard deviation of the background noise was also calculated for each 5 min epoch using a period immediately prior to electrical stimulation (s.d.Noise). The CV for each epoch was calculated as (s.d.EPSC - s.d.Noise)/MeanEPSC.

Failures analysis

EPSC amplitude was estimated by measuring the average current over two time windows (3 ms wide), one positioned immediately before the stimulus artefact and the other over the peak of the EPSC. Failures were visually identified and average EPSC amplitude (average of successes and failures), potency (average EPSC amplitude excluding failures) and success rate calculated for successive epochs throughout each experiment.

Hippocampal cultures

Cultures of dissociated hippocampal neurones were prepared from P3-P5 rats (Richmond et al. 1996). Rats were killed by decapitation and the brain rapidly removed. The CA1-CA3 region of the hippocampus was isolated and cells dissociated by trypsin digestion and mechanical dissociation before subsequent plating on 22 mm glass coverslips coated with poly-l-ornithine (25 μg ml−1; Sigma, UK) and matrigel (1:50 dilution; Becton Dickinson, UK). Cultures were maintained at 37 °C in a humidified atmosphere of 95 % O2-5 % CO2 in medium comprising minimal essential medium, 38.9 mm glucose, 2 mm glutamine, 15 mm Hepes, 100 μg ml−1 bovine transferrin, 30 μg ml−1 insulin, 0.1 μg ml−1 biotin, 1.5 μg ml−1 vitamin B12, 2 μg ml−1 gentamicin, 10 % fetal calf serum. After 24 h, medium was supplemented with cytosine arabinofuranoside (2.5 μm) to prevent glial cell proliferation. Culture medium was changed three times per week and after 10 days the concentration of fetal calf serum was reduced to 5 %. Cultures were used 6-14 days after plating. For experiments, cells were perfused continuously with Hepes buffered saline (HBS) at approximately 2 ml min−1 at room temperature. HBS comprised (mm): 119 NaCl, 5 KCl, 25 Hepes, 33 glucose, 2 CaCl2, 2 MgCl2, 0.0005 tetrodotoxin, 0.1 picrotoxin, pH 7.4, 305 mosmol l−1. Whole-cell recordings were made from the soma of pyramidal neurones voltage clamped at -70 mV using an Axopatch-1D amplifier. Intracellular solution comprised (mm): 110 caesium methane sulfonate, 40 Hepes, 0.6 EGTA, 10 NaCl, 4 Mg-ATP, 0.3 Na2-GTP, pH 7.2, 290 mosmol l−1. After obtaining a stable 10 min baseline period, cells were perfused locally with DHPG-containing HBS (100 μm for 5 min) via a glass capillary tube positioned near to the recorded cell. Application of DHPG was preceded and succeeded by 5 min local perfusion with HBS. Access resistance was monitored throughout using LTP software.

Data were filtered at 5 kHz and collected continuously on videotape for off-line analysis and subsequently digitised at 10 kHz and converted to Axon binary files using Axoscope software (Axon Instruments, USA). mEPSC detection was performed using MiniAnalysis software (Synaptosoft Inc., USA). Events were detected by setting the threshold value for detection at three times the level of the root mean squares noise, followed by visual confirmation of mEPSC detection. For construction of cumulative probability plots, 100 successive events were used from the period immediately preceding application of DHPG and from the period 30 min after commencing washout of DHPG. For pooled cumulative probability plots, data were normalised to the median value in the pre-DHPG period. Statistical significance was tested using the Kolmorogov-Smirnov test.

Drugs

QX-134 and BAPTA were obtained from Sigma, UK. RS-DHPG (Tocris, UK) was dissolved in H2O at a stock concentration of 100 times test concentration. Thapsigargin (Calbiochem, UK) was dissolved in DMSO at a stock concentration of 10 mm. Picrotoxin (Sigma, UK) was prepared as a stock solution of 5 mm in equimolar NaOH, except for use in hippocampal culture experiments where picrotoxin was dissolved directly in HBS. All stock solutions were stored as frozen aliquots and diluted to test concentration in aCSF. Neither DMSO (0.001 %) nor NaOH (50 μm) applied alone affected synaptic transmission when tested (data not shown).

Data analysis

All data are presented as means ±s.e.m. Unless stated otherwise, statistical significance was assessed using ANOVA.

RESULTS

Somatic recording from slices

Application of DHPG (30 μm for 20 min) induced LTD of synaptic transmission in 13/13 cells voltage clamped at -80 mV (Fig. 1). During application of DHPG, responses were depressed to 40 ± 4 % of control (n = 13) and after washout of DHPG the depression persisted for the time that the recording was maintained (minimum time 30 min, maximum time 120 min after commencing washout). Thirty minutes after washout of DHPG, responses were depressed to 56 ± 11 % of control (n = 13; P < 0.05). The change in EPSC amplitude was not associated with a change in access resistance, cell input resistance or holding current (Fig. 1). In three experiments, cells were voltage clamped at -30 mV for the period of DHPG application only. In these three cells the level of mGlu-LTD observed was similar to that seen when cells were voltage clamped at -80 mV throughout (responses depressed to 60 ± 10 % of control 30 min after commencing washout of DHPG; n = 3; data not shown).

Figure 1. DHPG induces LTD.

Figure 1

A, a representative experiment using somatic recording in which DHPG (30 μm) application induced LTD. In this and all subsequent figures DHPG application is indicated by a black bar. Example traces are taken from the time points indicated. Each point represents the mean of four successive EPSCs. B, holding current (•) and access resistance (○) remained unchanged during this experiment. C and D, pooled data from 13 cells. Data are presented as means ±s.e.m.

PPF and CV were analysed in seven of the cells voltage clamped at -80 mV. LTD induced by DHPG was associated with an increase in PPF (Fig. 2A–C). During application of DHPG, responses decreased to 37 ± 7 % of control (P < 0.05) whilst the PPR increased from 2.0 ± 0.1 to 2.6 ± 0.2 (P < 0.05). After 30 min of DHPG washout, responses were 56 ± 13 % of control (P < 0.05) and the PPR remained elevated at 2.8 ± 0.4 (P < 0.05 compared to baseline). In these seven cells, CV also increased with application of DHPG (Fig. 2D). Baseline CV was 0.22 ± 0.03, which increased to 0.35 ± 0.04 during DHPG application and remained elevated at 0.29 ± 0.03 30 min after commencing washout of DHPG (P < 0.05).

Figure 2. DHPG induces an increase in PPF and CV.

Figure 2

A, examples of EPSCs induced by paired pulses taken from a representative experiment (i) during baseline, (ii) during DHPG application and (iii) following DHPG washout. B–D, pooled data for seven cells showing changes in paired-pulse ratio, EPSC amplitude and coefficient of variation induced by DHPG application (30 μm for 20 min).

Dendritic recording from slices

We used dendritic recording for a failures analysis since this provides an enhanced signal to noise ratio and thus an unambiguous identification of failures (Isaac et al. 1998). Using dendritic whole-cell recording with minimal stimulation, application of DHPG (100 μm for 10 min) elicited mGlu-LTD in six of eight cells tested (Fig. 3). The DHPG-induced LTD was associated with a significant decrease in success rate but with no significant change in potency. During application of DHPG, average EPSC amplitude was depressed to 35 ± 10 % of control (P < 0.05; n = 6). This depression was associated with a decrease in success rate from 66 ± 6 % to 27 ± 5 % (P < 0.05) but no significant change in potency (-14 ± 4 to -10 ± 2 pA). Fifteen minutes after commencing washout of DHPG, average EPSC amplitude remained depressed at 38 ± 13 % of control (P < 0.05), success rate was 27 ± 7 % (P < 0.05 compared with baseline), but potency was not significantly changed (-14 ± 4 pA).

Figure 3. DHPG induces an increase in failure rate.

Figure 3

A, a representative experiment using dendritic recording and minimal stimulation, in which DHPG (100 μm) induced an increase in the failure rate of EPSCs but had little effect on potency. B, EPSC amplitude histogram for the experiment shown in A. Points around 0 mV represent failures. Bin width = 2 pA. C, 10 consecutive EPSCs (i and ii) and mean EPSCs (responses and failures; iii and iv) taken from baseline (i and iii) and DHPG washout periods (ii and iv) for a different cell. The mean EPSCs are scaled in v to show the lack of change in response kinetics. D and E, pooled data (n = 6) for cells showing success rate and potency changes during DHPG application and 15 min after DHPG washout (*P < 0.05 compared with baseline). F, plot of average EPSC amplitude ratio (DHPG washout/baseline) versus success rate ratio (DHPG washout/baseline) for cells in which LTD was induced, showing that the LTD can be fully accounted for by the decrease in success rate.

For the duration of each experiment, access resistance remained constant (baseline 40 ± 3 MΩ; DHPG washout 44 ± 4 MΩ). The rise and decay phases of mean EPSCs were fitted with single exponential functions to give time constants (τ) which were not significantly changed by DHPG application (baseline τrise= 1.4 ± 0.2 ms, τdecay= 8.5 ± 0.6 ms; LTD τrise= 1.5 ± 0.3 ms, τdecay= 9.1 ± 0.8 ms; n = 6).

The observation that potency did not decrease despite large decreases in success rate was surprising given the likelihood that multiple synapses were being activated (Hessler et al. 1993; Rosenmund et al. 1993; Allen & Stevens, 1994). When the change in potency was compared with that predicted from the Poisson equation for the change in success rate for each cell (Isaac et al. 1998) it was found that in the majority of cases (5/6) the actual potency ratio (LTD/baseline) was greater than the predicted ratio. One possible explanation for this observation is that DHPG induces not only a decrease in success rate, which causes LTD, but also a small increase in potency via another mechanism.

Hippocampal cultures

Application of DHPG (100 μm for 10 min) resulted in a decrease in the frequency of tetrodotoxin-resistant mEPSCs (Fig. 4). Thirty minutes after commencing washout of DHPG, mEPSC frequency was reduced to 61 ± 8 % of that recorded prior to application of DHPG (n = 5; P < 0.05). mEPSC amplitude was, however, not affected by application of DHPG; amplitude immediately prior to, and 30 min after commencing washout of, DHPG was 45 ± 8 and 49 ± 10 pA, respectively (n = 5; P > 0.05). These changes were accompanied by a shift in the cumulative probability plot for inter-event interval (P < 0.05, Kolmorogov-Smirnov test; Fig. 4E and F) but no change in the corresponding plots for mEPSC amplitude (P > 0.05).

Figure 4. DHPG application elicits a decrease in mEPSC frequency but no change in amplitude.

Figure 4

A, trace showing experiment where DHPG (100 μm for 10 min) was applied to cultured hippocampal neurones. Downward deflections indicate mEPSCs. Voltage steps used to measure access resistance have been blanked. B and C, traces from experiment shown in A on expanded time base taken from the period immediately prior to (B) and 30 min after (C) application of DHPG. Lower panel shows region of trace expanded from area indicated. D, plot showing decrease in the frequency of mEPSCs after application of DHPG in experiment shown in A. Data were divided into 10 s bins. E, cumulative probability plots for the experiment in A, showing no change in mEPSC amplitude (left) but a shift in inter-event interval (right; baseline period •, post-DHPG ○). F, pooled cumulative probability plots for five such cells.

Role of calcium in DHPG-induced LTD

To investigate the role of postsynaptic Ca2+ in DHPG-induced LTD, somatic recordings were made from CA1 neurones in hippocampal slices using an intracellular solution containing 10 mm of the calcium chelator BAPTA (experiments were interleaved with experiments using non-BAPTA containing solution, Fig. 1). Cells were dialysed for at least 1 h to ensure complete dialysis of BAPTA into the cell. Application of DHPG (30 μm for 20 min) induced LTD of synaptic transmission in cells voltage clamped at -80 mV. Thus, 30 min after commencing washout of DHPG, responses were 53 ± 15 % of control (n = 5; Fig. 5A).

Figure 5. DHPG-induced LTD is independent of calcium.

Figure 5

A, pooled data (n = 5) showing the effects of DHPG application (30 μm) in single cells dialysed with BAPTA-containing intracellular solution voltage clamped at -80 mV. B, pooled data (n = 5) for extracellular experiments where DHPG (50 μm; black bar) was applied in the presence of thapsigargin (10 μm; hatched bar). C, single extracellular experiment where DHPG (50 μm; black bar) was applied in the presence of Ca2+-free medium (open bars). Additional applications of Ca2+-free medium were performed without DHPG application prior to and after DHPG application. Traces to the right are taken from the time points indicated. D, pooled data showing the effects of DHPG application in the absence of extracellular Ca2+ (n = 5). E, pooled data from experiments as in D showing lack of long-term effect of Ca2+-free application (n = 5).

As group I mGlu receptors couple to release of Ca2+ from intracellular stores that may occur at the presynaptic site, we then investigated whether depletion of these stores by thapsigargin blocked DHPG-induced LTD using extracellular recordings (Fig. 5B). In these experiments, 40 min after application of DHPG (50 μm for 20 min) in the presence of thapsigargin (10 μm), responses were depressed to 67 ± 10 % of control (n = 5; P < 0.05). Finally, we performed experiments where DHPG was applied along with removal of extracellular Ca2+. Application of Ca2+-free aCSF, which completely abolished synaptic transmission, failed to prevent the induction of LTD by DHPG (50 μm for 20 min; Fig. 5C and D). Forty minutes after washout of DHPG (35 min after reintroduction of Ca2+) responses were 69 ± 6 % of control (n = 5; P < 0.05). In each experiment, additional applications of Ca2+-free media for similar periods (40 min) without DHPG application, both prior to and following the application of DHPG, did not lead to any long-term effects on synaptic transmission (Fig. 5C and E). Thirty-five minutes after reintroduction of Ca2+ (without DHPG application) responses were 98 ± 6 % of control (both applications of Ca2+-free medium from five experiments, P > 0.05).

DISCUSSION

The principal findings of the present study are that DHPG-induced LTD is induced via a Ca2+-independent mechanism and is expressed presynaptically.

On the mechanism of induction of DHPG-induced LTD

It has been shown previously that DHPG-induced LTD does not require electrical stimulation of afferent pathways for its induction (Fitzjohn et al. 1999). We now show that the induction is independent of alterations in postsynaptic Ca2+, since extensive loading of neurones with BAPTA failed to inhibit the effect. Indeed it is unlikely that Ca2+ is necessary for the effect since perfusion with Ca2+-free medium, which abolished synaptic transmission, or treatment with thapsigargin, which completely blocks mGlu receptor-mediated Ca2+ mobilisation in hippocampal neurones in culture (Rae et al. 2000) and is active in hippocampal slices (Harvey & Collingridge, 1992), also failed to prevent DHPG-induced LTD. These findings do not, however, distinguish between Ca2+-independent induction mechanisms in the pre- or postsynaptic neurone.

On the mechanism of expression of DHPG-induced LTD

Collectively, the findings that DHPG-induced LTD is associated with (1) an increase in paired-pulse facilitation, (2) an increase in the coefficient of variation of EPSC amplitude, (3) a decrease in success rate but not potency and (4) a decrease in mEPSC frequency but not amplitude, suggest strongly that this form of plasticity is due to a decrease in release probability (Bekkers & Stevens, 1990; Manabe et al. 1993; Kullman, 1994; Isaac et al. 1998). Recently, it has been suggested that DHPG-induced LTD involves the internalisation of AMPA receptors (Snyder et al. 2000). Changes in the coefficient of variation of EPSC amplitude, failure rate and mEPSC frequency could be explained by a postsynaptic mechanism if DHPG resulted in the silencing of synapses by causing internalisation of the entire AMPA receptor complement. In this context, it has been shown that NMDA receptor-dependent LTD involves the silencing of synapses (Lüthi et al. 1999) and internalisation of AMPA receptors (Carroll et al. 1999). However, postsynaptic silencing of synapses is unlikely to explain DHPG-induced LTD since the effect was associated with an increase in the level of paired-pulse facilitation. Thus a decrease in the number of AMPA receptor-containing synapses should not affect the level of paired-pulse facilitation, unless AMPA receptors are preferentially removed from sites where the probability of glutamate release is higher than average. This situation might arise if DHPG-induced LTD required both activation of mGlu receptors and synaptic activation of AMPA receptors. However, the finding that DHPG-induced LTD does not require either electrical stimulation of afferent pathways (Fitzjohn et al. 1999) or Ca2+-dependent synaptic transmission does not support this explanation. Alternatively, other Ca2+-independent methods for removing AMPA receptors from high probability synapses may exist. However, our results are most simply explained by a presynaptic mechanism of expression. This does not exclude the possibility that a form of DHPG-induced LTD that is expressed postsynaptically can co-exist at these synapses under certain conditions.

From induction to expression of DHPG-induced LTD

How is this form of presynaptic LTD induced? If the induction is postsynaptic then the process requires a Ca2+-independent, retrograde signalling mechanism. Alternatively, the entire process may occur presynaptically. Thus although group I mGlu receptors are predominantly located postsynaptically (Lujan et al. 1997) they have also been observed presynaptically (Romano et al. 1995). Furthermore, activation of group I mGlu receptors has been shown to modulate glutamate release from hippocampal synaptoneurosomes (Rodriguez-Moreno et al. 1998) and produces an acute depression of synaptic transmission that is also most readily explained by a presynaptic locus (Manzoni & Bockaert, 1995; Gereau & Conn, 1995).

The physiological significance of DHPG-induced LTD

There are intense investigations into the mechanisms of LTD in relation to synaptic plastic mechanisms involved in development and learning and memory. The most intensely studied is an NMDA receptor-dependent form of LTD (Dudek & Bear, 1992; Mulkey & Malenka, 1992), which is expressed postsynaptically (e.g. Oliet et al. 1997; Lüthi et al. 1999; Carroll et al. 1999). The ability to induce this form of LTD with a specific NMDA receptor agonist is enabling the molecular basis of this form of plasticity to be elucidated (Lee et al. 1998). Likewise, the ability to induce mGlu receptor-dependent LTD using a specific mGlu receptor agonist should similarly facilitate the understanding of its molecular basis. The physiological importance of DHPG-induced LTD is indicated by the findings that mGlu receptor-dependent LTD can be induced synaptically (Stanton et al. 1991; Bashir et al. 1993; Bolshakov & Siegelbaum, 1994; Oliet et al. 1997; Fitzjohn et al. 1998a; Kemp & Bashir, 1999; Huber et al. 2000). It is likely, however, that there is more than one form of mGlu receptor-dependent LTD. For example the form investigated here differs significantly from that studied by Oliet et al. (1997). The latter was input specific and dependent on postsynaptic Ca2+, and could not be induced when GABAA receptor-mediated synaptic inhibition was blocked. It also required artificially high concentrations of divalent cations (4 mm Ca2+ plus 4 mm Mg2+). Furthermore, it was blocked by a protein kinase C (PKC) inhibitor whereas DHPG-induced LTD at CA1 synapses is resistant to a variety of PKC inhibitors (Schnabel et al. 1999, 2001). Further work is necessary to determine the conditions required to induce synaptically the form of LTD that can be selectively induced by DHPG.

The presence of pre- and postsynaptic forms of LTD at the same synapses confers a greater flexibility in the long-term modulation of synaptic strength. NMDA receptor-dependent LTD senses both pre- and postsynaptic activity in a Hebbian manner whereas DHPG-induced LTD does not require active participation of the postsynaptic cell.

Acknowledgments

This work was supported by the MRC, the BBSRC, Eli Lilly and the European Union (Biomed BMH4-CT96-0228).

Stephen M. Fitzjohn and Mary J. Palmer contributed equally to this work.

References

  1. Allen C, Stevens CF. An evaluation of causes of unreliability of synaptic transmission. Proceedings of the National Academy of Sciences of the USA. 1994;91:10380–10383. doi: 10.1073/pnas.91.22.10380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Anderson WW, Collingridge GL. The LTP program: a data acquisition program for on-line analysis of long-term potentiation and other synaptic events. Journal of Neuroscience Methods. 2001;108:71–83. doi: 10.1016/s0165-0270(01)00374-0. [DOI] [PubMed] [Google Scholar]
  3. Anwyl R. Metabotropic glutamate receptors: electrophysiological properties and role in plasticity. Brain Research Reviews. 1999;29:93–120. doi: 10.1016/s0165-0173(98)00050-2. [DOI] [PubMed] [Google Scholar]
  4. Bashir ZI, Jane DE, Sunter DC, Watkins JC, Collingridge GL. Metabotropic glutamate receptors contribute to the induction of long-term depression in the CA1 region of the hippocampus. European Journal of Pharmacology. 1993;239:265–266. doi: 10.1016/0014-2999(93)91009-c. [DOI] [PubMed] [Google Scholar]
  5. Baskys A, Malenka RC. Agonists at metabotropic glutamate receptors presynaptically inhibit EPSCs in neonatal rat hippocampus. Journal of Physiology. 1991;444:687–701. doi: 10.1113/jphysiol.1991.sp018901. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bekkers JM, Stevens CF. Presynaptic mechanism for long-term potentiation in the hippocampus. Nature. 1990;346:724–729. doi: 10.1038/346724a0. [DOI] [PubMed] [Google Scholar]
  7. Bolshakov VY, Siegelbaum SA. Postsynaptic induction and presynaptic expression of hippocampal long-term depression. Science. 1994;264:1148–1152. doi: 10.1126/science.7909958. [DOI] [PubMed] [Google Scholar]
  8. Bortolotto ZA, Fitzjohn SM, Collingridge GL. Roles of metabotropic glutamate receptors in LTP and LTD in the hippocampus. Current Opinion in Neurobiology. 1999;9:299–304. doi: 10.1016/s0959-4388(99)80044-0. [DOI] [PubMed] [Google Scholar]
  9. Camodeca N, Breakwell NA, Rowan MJ, Anwyl R. Induction of LTD by activation of group I mGluR in the dentate gyrus in vitro. Neuropharmacology. 1999;38:1597–1606. doi: 10.1016/s0028-3908(99)00093-3. [DOI] [PubMed] [Google Scholar]
  10. Carroll RC, Lissin DV, von Zastrow M, Nicoll RA, Malenka RC. Rapid redistribution of glutamate receptors contributes to long-term depression in hippocampal cultures. Nature Neuroscience. 1999;2:454–460. doi: 10.1038/8123. [DOI] [PubMed] [Google Scholar]
  11. Conn PJ, Pin J-P. Pharmacology and functions of metabotropic glutamate receptors. Annual Review of Pharmacology and Toxicology. 1997;37:205–237. doi: 10.1146/annurev.pharmtox.37.1.205. [DOI] [PubMed] [Google Scholar]
  12. Dudek SM, Bear MF. Homosynaptic long-term depression in area CA1 of hippocampus and effects of N-methyl-D-aspartate receptor blockade. Proceedings of the National Academy of Sciences of the USA. 1992;89:4363–4567. doi: 10.1073/pnas.89.10.4363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Fitzjohn SM, Bortolotto ZA, Palmer MJ, Doherty AJ, Ornstein PL, Schoepp DD, Kingston AE, Lodge D, Collingridge GL. The potent mGlu receptor antagonist LY341495 identifies roles for both cloned and novel mGlu receptors in hippocampal synaptic plasticity. Neuropharmacology. 1998a;37:1445–1458. doi: 10.1016/s0028-3908(98)00145-2. [DOI] [PubMed] [Google Scholar]
  14. Fitzjohn SM, Kingston AE, Lodge D, Collingridge GL. DHPG-induced LTD in area CA1 of juvenile rat hippocampus; characterisation and sensitivity to novel mGlu receptor antagonists. Neuropharmacology. 1999;38:1577–1583. doi: 10.1016/s0028-3908(99)00123-9. [DOI] [PubMed] [Google Scholar]
  15. Fitzjohn SM, Lodge D, Collingridge GL. Evidence for a presynaptic locus of expression of LTD induced by group I mGluRs in the CA1 region of the immature rat hippocampus. Society for Neuroscience Abstracts. 1996;22 686.3. [Google Scholar]
  16. Fitzjohn SM, Morris SAC, Collingridge GL. Long-lasting synaptic depression induced by mGlu5 receptor activation is independent of calcium. Society for Neuroscience Abstracts. 1998b;24 810.2. [Google Scholar]
  17. Gereau RW, Conn PJ. Multiple presynaptic metabotropic glutamate receptors modulate excitatory and inhibitory synaptic transmission in hippocampal area CA1. Journal of Neuroscience. 1995;15:6879–6889. doi: 10.1523/JNEUROSCI.15-10-06879.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Harvey J, Collingridge GL. Thapsigargin blocks the induction of long-term potentiation in rat hippocampal slices. Neuroscience Letters. 1992;139:197–200. doi: 10.1016/0304-3940(92)90551-h. [DOI] [PubMed] [Google Scholar]
  19. Harvey J, Palmer MJ, Irving AJ, Clarke VRJ, Collingridge GL. NMDA receptor dependence of mGluR-mediated depression of synaptic transmission in the CA1 region of the rat hippocampus. British Journal of Pharmacology. 1996;119:1239–1247. doi: 10.1111/j.1476-5381.1996.tb16028.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Hessler NA, Shirke AM, Malinow R. The probability of transmitter release at a mammalian central synapse. Nature. 1993;366:569–572. doi: 10.1038/366569a0. [DOI] [PubMed] [Google Scholar]
  21. Huber KM, Kayser MS, Bear MF. Role for rapid dendritic protein synthesis in hippocampal mGluR-dependent long-term depression. Science. 2000;288:1254–1257. doi: 10.1126/science.288.5469.1254. [DOI] [PubMed] [Google Scholar]
  22. Isaac JTR, Lüthi A, Palmer MJ, Anderson WW, Benke TA, Collingridge GL. An investigation of the expression mechanism of LTP of AMPA receptor-mediated synaptic transmission at hippocampal CA1 synapses using failures analysis and dendritic receordings. Neuropharmacology. 1998;37:1399–1410. doi: 10.1016/s0028-3908(98)00140-3. [DOI] [PubMed] [Google Scholar]
  23. Kemp N, Bashir ZI. Induction of LTD in the adult hippocampus by the synaptic activation of AMPA/kainate and metabotropic glutamate receptors. Neuropharmacology. 1999;38:495–504. doi: 10.1016/s0028-3908(98)00222-6. [DOI] [PubMed] [Google Scholar]
  24. Kullmann DM. Amplitude fluctuations of dual-component EPSCs in hippocampal pyramidal cells: Implications for long-term potentiation. Neuron. 1994;12:1111–1120. doi: 10.1016/0896-6273(94)90318-2. [DOI] [PubMed] [Google Scholar]
  25. Lee HK, Kameyama K, Huganir RL, Bear MF. NMDA induces long-term synaptic depression and dephosphorylation of the GluR1 subunit of AMPA receptors in hippocampus. Neuron. 1998;21:1151–1162. doi: 10.1016/s0896-6273(00)80632-7. [DOI] [PubMed] [Google Scholar]
  26. Lujan R, Roberts JD, Shigemoto R, Ohishi H, Somogyi P. Differential plasma membrane distribution of metabotropic glutamate receptors mGluR1 α, mGluR2 and mGluR5, relative to neurotransmitter release sites. Journal of Chemical Neuroanatomy. 1997;13:219–241. doi: 10.1016/s0891-0618(97)00051-3. [DOI] [PubMed] [Google Scholar]
  27. Lüthi A, Chittajallu R, Duprat F, Palmer MJ, Benke TA, Kidd FL, Henley JM, Isaac JT, Collingridge GL. Hippocampal LTD expression involves a pool of AMPARs regulated by the NSF-GluR2 interaction. Neuron. 1999;24:389–399. doi: 10.1016/s0896-6273(00)80852-1. [DOI] [PubMed] [Google Scholar]
  28. Manabe T, Wyllie DJA, Perkel DJ, Nicoll RA. Modulation of synaptic transmission and long-term potentiation: Effects on paired pulse facilitation and EPSC variance in the CA1 region of the hippocampus. Journal of Neurophysiology. 1993;70:1451–1459. doi: 10.1152/jn.1993.70.4.1451. [DOI] [PubMed] [Google Scholar]
  29. Manzoni O, Bockaert J. Metabotropic glutamate receptors inhibiting excitatory synapses in the CA1 area of rat hippocampus. European Jounal of Neuroscience. 1995;7:2518–2523. doi: 10.1111/j.1460-9568.1995.tb01051.x. [DOI] [PubMed] [Google Scholar]
  30. Mulkey RM, Malenka RC. Mechanisms underlying induction of homosynaptic long-term depression in area CA1 of the hippocampus. Neuron. 1992;9:967–975. doi: 10.1016/0896-6273(92)90248-c. [DOI] [PubMed] [Google Scholar]
  31. Oliet SHR, Malenka RC, Nicoll RA. Two distinct forms of long-term depression coexist in CA1 hippocampal pyramidal cells. Neuron. 1997;18:969–982. doi: 10.1016/s0896-6273(00)80336-0. [DOI] [PubMed] [Google Scholar]
  32. O'Mara SM, Rowan MJ, Anwyl R. Metabotropic glutamate receptor-induced homosynaptic long-term depression and depotentiation in the dentate gyrus of the rat hippocampus in vitro. Neuropharmacology. 1995;34:983–989. doi: 10.1016/0028-3908(95)00062-b. [DOI] [PubMed] [Google Scholar]
  33. Overstreet LS, Pasternak JF, Colley PA, Slater NT, Trommer BL. Metabotropic glutamate receptor mediated long-term depression in developing hippocampus. Neuropharmacology. 1997;36:831–844. doi: 10.1016/s0028-3908(97)00031-2. [DOI] [PubMed] [Google Scholar]
  34. Palmer MJ, Collingridge GL. Metabotropic glutamate receptor-induced LTD in the hippocampus is associated with an increase in failure rate. Society for Neuroscience Abstracts. 1999;25 494.1. [Google Scholar]
  35. Palmer MJ, Irving AJ, Seabrook GR, Jane DE, Collingridge GL. The group I mGlu receptor agonist DHPG induces a novel form of LTD in the CA1 region of the hippocampus. Neuropharmacology. 1997;36:1517–1532. doi: 10.1016/s0028-3908(97)00181-0. [DOI] [PubMed] [Google Scholar]
  36. Rae MG, Martin D J, Collingridge GL, Irving AJ. Role of Ca2+ stores in metabotropic L-glutamate receptor-mediated supralinear Ca2+ signaling in rat hippocampal neurons. Journal of Neuroscience. 2000;20:8628–8636. doi: 10.1523/JNEUROSCI.20-23-08628.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Richmond SA, Irving AJ, Molnar E, McIlhinney RAJ, Michelangeli F, Henley JM, Collingridge GL. Localization of the glutamate receptor subunit GluR1 on the surface of living and within cultured hippocampal neurons. Neuroscience. 1996;75:69–82. doi: 10.1016/0306-4522(96)00217-5. [DOI] [PubMed] [Google Scholar]
  38. Rodriguez-Moreno A, Sistiaga A, Lerma J, Sanchez-Prieto J. Switch from facilitation to inhibition of excitatory synaptic transmission by group I mGluR desensitization. Neuron. 1998;21:1477–1486. doi: 10.1016/s0896-6273(00)80665-0. [DOI] [PubMed] [Google Scholar]
  39. Romano C, Sesma MA, McDonald, C T, O'Malley K, Vanden Pol AN, Olney JW. Distribution of metabotropic glutamate receptor mGluR5 immunoreactivity in rat brain. Journal of Comparative Neurology. 1995;355:455–469. doi: 10.1002/cne.903550310. [DOI] [PubMed] [Google Scholar]
  40. Rosenmund C, Clements JD, Westbrook GL. Nonuniform probability of glutamate release at a hippocampal synapse. Science. 1993;262:754–757. doi: 10.1126/science.7901909. [DOI] [PubMed] [Google Scholar]
  41. Schnabel R, Kilpatrick IC, Collingridge GL. An investigation into signal transduction mechanisms involved in DHPG-induced LTD in the CA1 region of the hippocampus. Neuropharmacology. 1999;38:1585–1596. doi: 10.1016/s0028-3908(99)00062-3. [DOI] [PubMed] [Google Scholar]
  42. Schnabel R, Kilpatrick IC, Collingridge GL. Protein phosphatase inhibitors facilitate DHPG-induced LTD in the CA1 region of the hippocampus. British Journal of Pharmacology. 2001;132:1095–1101. doi: 10.1038/sj.bjp.0703905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Snyder EM, Huber KM, Dong X, Fallon JR, Bear MF. Group 1 metabotropic glutamate receptor activation initiates internalization of AMPA receptors in cultured hippocampal neurons. Society for Neuroscience Abstracts. 2000;26 134.15. [Google Scholar]
  44. Stanton PK, Chattarji S, Sejnowski TJ. 2-Amino-3-phosphonoproprionic acid, an inhibitor of glutamate-stimulated phosphoinositide turnover, blocks induction of homosynaptic long-term depression, but not potentiation in rat hippocampus. Neuroscience Letters. 1991;127:61–66. doi: 10.1016/0304-3940(91)90895-z. [DOI] [PubMed] [Google Scholar]
  45. Tan Y, Hyland J, Hori N, Carpenter DO. The mechanism of presynaptic long-lasting depression mediated by group I metabotropic glutamate receptors. Society for Neuroscience Abstracts. 1999;25 doi: 10.1023/a:1022949922364. 494.2. [DOI] [PubMed] [Google Scholar]
  46. Vignes M, Clarke VRJ, Davies CH, Chambers A, Jane DE, Watkins JC, Collingridge GL. Pharmacological evidence for an involvement of group II and group III mGluRs in the presynaptic regulation of excitatory synaptic responses in the CA1 region of rat hippocampal slices. Neuropharmacology. 1995;34:973–982. doi: 10.1016/0028-3908(95)00093-l. [DOI] [PubMed] [Google Scholar]
  47. Wetzel W, Behnisch T, Jäger T, Sokolov M, Reymann KG, Balschun D. mGluR5 activation induces late LTD and impairs spatial learning in the adult rat. Society for Neuroscience Abstracts. 1999;25 290.11. [Google Scholar]

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

RESOURCES