Neuropharmacology 74 (2013) 135e146
Contents lists available at SciVerse ScienceDirect
Neuropharmacology journal homepage: www.elsevier.com/locate/neuropharm
Antagonists reversibly reverse chemical LTD induced by group I, group II and group III metabotropic glutamate receptors David Lodge a, *, Patrick Tidball a, Marion S. Mercier a, Sarah J. Lucas a, b, Lydia Hanna a, Laura Ceolin a, c, Minos Kritikos a, Stephen M. Fitzjohn a, d, John L. Sherwood a, d, Neil Bannister a, Arturas Volianskis a, David E. Jane a, Zuner A. Bortolotto a, Graham L. Collingridge a, e a
Centre for Synaptic Plasticity, School of Physiology and Pharmacology, Dorothy Hodgkin Building, University of Bristol, Bristol BS1 3NY, UK Laboratory of Neurobiology, National Institute of Environmental Health Sciences, 111 T.W. Alexander Drive, Research Triangle Park, NC 27709, USA c Institute of Functional Genomics, CNRS-UMR5203, INSERM-U661, Université Montpellier 1, Université Montpellier 2, 141 Rue de la Cardonille, 34000 Montpellier, France d Neuroscience Research Division, Lilly Research Centre, Eli Lilly & Co. Ltd., Erl Wood Manor, Windlesham, Surrey GU20 6PH, UK e Department of Brain and Cognitive Sciences, College of Natural Sciences, Seoul National University, Shilim, Gwanak, Seoul 151-746, Republic of Korea b
a r t i c l e i n f o
a b s t r a c t
Article history: Received 19 December 2012 Received in revised form 25 February 2013 Accepted 7 March 2013
Metabotropic glutamate (mGlu) receptors are implicated in many neurological and psychiatric diseases and are the targets of therapeutic agents currently in clinical development. Their activation has diverse effects in the central nervous system (CNS) that includes an involvement in synaptic plasticity. We previously reported that the brief exposure of hippocampal slices to dihydroxyphenylglycine (DHPG) can result in a long-term depression (LTD) of excitatory synaptic transmission. Surprisingly, this LTD could be fully reversed by mGlu receptor antagonists in a manner that was itself fully reversible upon washout of the antagonist. Here, 15 years after the discovery of DHPG-LTD and its reversible reversibility, we summarise these initial findings. We then present new data on DHPG-LTD, which demonstrates that evoked epileptiform activity triggered by activation of group I mGlu receptors can also be reversibly reversed by mGlu receptor antagonists. Furthermore, we show that the phenomenon of reversible reversibility is not specific to group I mGlu receptors. We report that activation of group II mGlu receptors in the temporo-ammonic pathway (TAP) and mossy fibre pathway within the hippocampus and in the cortical input to neurons of the lateral amygdala induces an LTD that is reversed by LY341495, a group II mGlu receptor antagonist. We also show that activation of group III mGlu8 receptors induces an LTD at lateral perforant path inputs to the dentate gyrus and that this LTD is reversed by MDCPG, an mGlu8 receptor antagonist. In conclusion, we have shown that activation of representative members of each of the three groups of mGlu receptors can induce forms of LTD than can be reversed by antagonists, and that in each case washout of the antagonist is associated with the re-establishment of the LTD. This article is part of the Special Issue entitled ‘Glutamate Receptor-Dependent Synaptic Plasticity’. Ó 2013 Elsevier Ltd. All rights reserved.
Keywords: Metabotropic glutamate receptors mGlu1 mGlu2 mGlu5 mGlu8 DHPG DCG-IV Long-term depression
1. Introduction Glutamate receptors comprise three main ionotropic subtypes, AMPA, NMDA and kainate (Collingridge et al., 2009), and a family of G-protein coupled metabotropic (mGlu) receptors (Pin and Duvoisin, 1995). The eight known mGlu receptors are further subdivided into three groups based on structural and functional criteria: group I (mGlu1 and mGlu5), group II (mGlu2 and mGlu3) and group III (mGlu4, mGlu6, mGlu7 and mGlu8). There is currently * Corresponding author. Tel.: þ44 1173313148; fax: þ44 1173313029. E-mail address:
[email protected] (D. Lodge). 0028-3908/$ e see front matter Ó 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.neuropharm.2013.03.011
considerable interest in mGlu receptors as targets for therapeutic agents (Nicoletti et al., 2011; Niswender and Conn, 2010). For example, negative allosteric modulators (NAMs) acting at mGlu5 receptors are being explored as potential treatments for epileptogenesis, Fragile-X syndrome, tardive dyskinesias and autism spectrum disorders (Bianchi et al., 2012; Gürkan and Hagerman, 2012; Michalon et al., 2012; Rylander et al., 2009). In addition, agonists acting at group II mGlu receptors have shown efficacy in man as therapies for anxiety and schizophrenia (Dunayevich et al., 2008; Grillon et al., 2003; Patil et al., 2007). Furthermore, mGlu5 receptors have been linked to the aetiology of Alzheimer’s disease (Hu et al., 2012). Therefore understanding both the functions of
136
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
Abbreviations 1S,3R-ACPD (1S,3R)-1-aminocyclopentane-1,3-dicarboxylic acid DCG-IV (2S,20 R,30 R)-2-(20 ,30 -dicarboxycyclopropyl)glycine DCPG (S)-3,4-dicarboxyphenylglycine DHPG (RS)-3,5-dihydroxyphenylglycine L-AP4 L-(þ)-2-amino-4-phosphonobutyric acid L-689,560 trans-2-carboxy-5,7-dichloro-4-phenylaminocarbonylamino-1,2,3,4-tetrahydroquinoline LY341495 (2S,10 S,20 S)-2-(9-xanthylmethyl)-2-(20 -carboxycyclopropyl)glycine LY367385 (S)-(þ)-a-amino-4-carboxy-2-methylbenzeneacetic acid LY379268 (1R,4R,5S,6R)-4-amino-2-oxabicyclo[3.1.0]hexane-4,6-dicarboxylic acid LY395756 (1SR,2SR,4RS,5RS,6SR)-2-amino-4-methylbicyclo[3.1.0]-hexane-2,6-dicarboxylic acid a-Methyl-4-carboxyphenylglycine MCPG MDCPG (RS)-a-methyl-3,4-dicarboxyphenylglycine MPEP 2-methyl-6-(phenylethynyl)-pyridine hydrochloride
mGlu receptors and the actions of ligands that affect these receptors is of relevance to the development of therapeutic agents active against these, and other, disorders. Fifteen years ago we described a form of LTD following the selective activation of group I mGlu receptors using the agonist, DHPG (Fitzjohn et al., 1998; Palmer et al., 1997). This form of synaptic plasticity, referred to hereafter as DHPG-LTD, has been extensively studied by many groups (e.g. Huang and Hsu, 2006; Huber et al., 2000; Massey and Bashir, 2007; Watabe et al., 2002; Xiao et al., 2001) A similar form of LTD has also been seen using the less selective mGlu receptor agonist, 1S,3R-ACPD, (O’Mara et al., 1995; Overstreet et al., 1997) although this is not invariably the case (Palmer et al., 1997). DHPG-LTD is mechanistically different from the other major form of chemical LTD that can be induced by activation of NMDA receptors (NMDA-LTD) (see Collingridge et al., 2010). However, there is still disagreement over some aspects of the pharmacology and biochemistry of the induction and expression mechanisms. For example, which group I receptor mediates the DHPG-LTD (Faas et al., 2002; Gladding et al., 2009; Kumar and Foster, 2007; Moult et al., 2006; Volk et al., 2006), which kinases and phosphatases are involved (Gallagher et al., 2004; Hou and Klann, 2004; Mockett et al., 2011; Moult et al., 2008; Schnabel et al., 1999) and the relative contributions of pre- and post-synaptic mechanisms (Fitzjohn et al., 2001; Moult et al., 2006; Qian and Noebels, 2006; Tan et al., 2003; Upreti et al., 2013; Watabe et al., 2002; Waung et al., 2008; Xiao et al., 2001) are all subject to controversy. These aspects of LTD are, however, not the subject of the present article but rather we concentrate on the pharmacology of the maintenance phase of LTD induced by the pharmacological activation of mGlu receptors. More specifically, at the same time as reporting DHPG-LTD, we noted the curious phenomenon that mGlu receptor antagonists, such as a-methyl-4-carboxyphenylglycine (MCPG), can reverse this plasticity long after it has been induced and that the plasticity is restored after washout of the antagonist (Fitzjohn et al., 1998; Palmer et al., 1997). Here we summarise these historical findings and present a series of new observations concerning the antagonist reversal of the maintained phase of group I mGlu receptortriggered LTD (mGlu receptor-LTD). We also demonstrate similar phenomena for other mGlu receptor subtypes in groups II (mGlu2 and mGlu3) and III (mGlu8). 2. Materials and methods 2.1. Animals and slice preparation Experiments were performed according UK Scientific Procedures Act, 1986 and EU Guidelines for Animal Care and conducted as described in previous publications (Ceolin
et al., 2011; Fitzjohn et al., 1999; Hanna et al., 2012; Lucas et al., 2012; Mercier et al., 2013; Sherwood et al., 2012). Briefly, electrophysiological recordings were obtained from slices prepared from rats and mice. Extracellular and whole cell experiments were performed as described in the text. Stimulating electrodes were placed in the Schaffer collaterals, the temporo-ammonic pathway, mossy fibre pathway, the external capsule or the lateral perforant path and recordings were made in the CA1 stratum radiatum, the CA1 stratum lacunosum moleculare, the CA3 stratum lucidum, the lateral amygdala or the outer third of the dentate gyrus stratum moleculare, respectively. 2.2. Analyses of electrophysiological recordings Data were captured and analysed using the WinLTP program (Anderson and Collingridge, 2007). In extracellular experiments, the effects of compounds on AMPA receptormediated synaptic transmission were quantified by measuring either initial slopes or peak amplitudes of AMPA receptor-mediated field excitatory post-synaptic potentials (fEPSPs) and normalised to baseline. Similarly, peak excitatory postsynaptic currents (EPSCs) were quantified in whole cell experiments. Changes in excitability due to DHPG application were quantified using coastline analyses, which are built into WinLTP (www.winltp.com). Data are presented both as single experiments and as mean values of experimental groups (S.E.M). 2.3. Chemicals Compounds were purchased from Tocris Cookson, Bristol, U.K.:-(2S,20 R,30 R)-2(20 ,30 -dicarboxycyclopropyl)glycine (DCG-IV) and trans-2-carboxy-5,7-dichloro-4phenylaminocarbonylamino-1,2,3,4-tetrahydroquinoline (L-689,560), or from Abcam, Cambridge, U.K.:- picrotoxin (PTX), (S)-3,4-dicarboxyphenylglycine (DCPG), (2S,1’S,20 S)-2-(9-xanthylmethyl)-2-(20 -carboxycyclopropylglycine (LY341495), (RS)3,5-dihydroxyphenylglycine (DHPG), D-2-amino-5-phosphono-pentanoate (D-AP5), (S)-(þ)-a-amino-4-carboxy-2-methylbenzeneacetic acid (LY367385), a-methyl-4carboxyphenylglycine (MCPG) and 2-methyl-6-(phenylethynyl)-pyridine hydrochloride (MPEP). (1R,4R,5S,6R)-4-amino-2-oxabicyclo[3.1.0]hexane-4,6-dicarboxylic acid (LY379268) and (1SR,2SR,4RS,5RS,6SR)-2-amino-4-methylbicyclo[3.1.0]-hexane-2,6-dicarboxylic acid (LY395756) were kind gifts of Dr James Monn, Eli Lilly & Co. and (RS)-a-methyl-3,4-dicarboxyphenylglycine (MDCPG) was synthesized in house. All other fine chemicals were purchased from Fisher Scientific or Sigma.
3. Results 3.1. Group I mGlu receptor-mediated LTD The original observation that MCPG (Palmer et al., 1997), and other mGlu receptor antagonists (Fitzjohn et al., 1998, 1999; Palmer et al., 1997), can reverse DHPG-LTD in a reversible manner is illustrated in Fig. 1. Fig. 1A shows a single example time-course plot. MCPG alone had little effect on the synaptic response when applied during baseline recording but, when applied after the induction of DHPG-LTD, it was able to fully reverse the synaptic depression. Significantly, DHPG-LTD was fully restored upon washout of MCPG. The consistency of this observation can be seen in the pooled timecourse data plot (Fig. 1B). The effect is extremely unlikely to be due
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
137
Fig. 1. Transient reversal of the expression of group I mGlu receptor mediated LTD by antagonists. A. Single example showing the lack of effect of (S)-MCPG (1 mM) on baseline but complete reversal of LTD induced by DHPG (100 mM). Note that upon washout of MCPG, LTD was re-instated. B. Pooled data (n ¼ 5). C. The ability of DHPG (100 mM) to potentiate depolarisations induced by NMDA (20 mM, 2 min applied at times indicated by arrowheads) is rapidly reversible, demonstrating rapid washout of DHPG. D. A single example to show the ability of a range of mGlu receptor antagonists (1 mM (S)-4CPG, 0.5 mM MTPG, 0.5 mM MPPG and 1 mM (S)-MCPG) to reverse the expression of DHPG-LTD. E. Reversal of LTD induced by three applications of DHPG (30 mM), by LY341495 (100 mM). F. Reversal of DHPG-LTD by LY393053 (n ¼ 5). Data shown in panels AeE are from hippocampal slices obtained from adult rats (4e10 week old), bathed in Mg2þ-free medium (AeD; (Palmer et al., 1997) and in 1 mM Mg2þ-containing medium (E; Fitzjohn et al., 1998). Data shown in panel F are from hippocampal slices obtained from juvenile (12e18 day old) rats, bathed in 1 mM Mg2þ-containing medium (Fitzjohn et al., 1999).
to a slow washout of DHPG for two reasons. First, in interleaved experiments the ability of DHPG to acutely potentiate depolarisations induced by NMDA was rapidly reversible (Fig. 1C). Second, as shown in Fig. 1A and B, the synaptic plasticity was restored following washout of MCPG. The reversal of DHPG-LTD was not unique to MCPG. In Fig. 1D, a range of phenylglycine antagonists, which have a varying degree of activity at the various mGlu receptor subtypes (Schoepp et al., 1999; Watkins and Collingridge, 1994), were compared. The spectrum of activity was consistent with the reversal effect being due to persistent activity of group I mGlu receptors. Note that, even when applied approximately 3 h after the induction of DHPG-LTD, MCPG fully reversed the LTD in a completely reversible manner. Fig. 1E shows an example of full reversal of a large DHPG-LTD, induced by three applications of the agonist, by LY341495 applied at a concentration (100 mM) that blocks all mGlu receptor subtypes (Fitzjohn et al., 1999). Fig. 1F shows pooled data of the effects of LY393053, an antagonist that has a degree of selectively towards group I mGlu receptors (Fitzjohn et al., 1998). The data illustrated in Fig. 1AeE were obtained from slices prepared from 4 to 10 week old rats whilst those in Fig. 1F were obtained from slices obtained from 12 to 18 day old rats. Therefore this phenomenon appears not to be unique to a particular developmental time point. The above observations with antagonists have been largely replicated in several other laboratories. Thus, there is general
agreement that persistent activation of group I mGlu receptors is involved in the maintenance of DHPG-LTD (Huang and Hsu, 2006; Palmer et al., 1997; Watabe et al., 2002), rather than other mGlu receptors (Palmer et al., 1997) or NMDA receptors (Huang and Hsu, 2006). Most studies, however, do not differentiate between the contribution of mGlu1 and mGlu5 receptor subtypes to the maintenance of this form of plasticity. By comparing the actions of LY367385 and MPEP, selective antagonists for mGlu1 and mGlu5 receptors, respectively (Schoepp et al., 1999), it was shown that either mGlu1 (Volk et al., 2006) or mGlu5 (Huang and Hsu, 2006; Ronesi et al., 2012) receptors provided the greater role in the maintenance of DHPG-LTD. We recently investigated this issue in parasagittal hippocampal slices with the CA3 field removed from 8 to 12 week old rats, using methods described previously, which include addition of GABAA (picrotoxin) and NMDA (L-689,560) receptor antagonists in the perfusate (Palmer et al., 1997). Bath application of DHPG 100 mM for 10 min reduced the fEPSP recorded in stratum radiatum of CA1 following submaximal stimulation of the Schaffer collateral/commissural input (Fig. 2). The acute phase was followed by a sustained reduction (i.e., DHPG-LTD), which lasted for as long as recordings were maintained (typically up to 2 h after washout of DHPG). During this sustained phase application of the mGlu1 antagonist, LY367385 (100 mM; applied 70 min after stopping the perfusion with DHPG) produced a partial reversal of the LTD, and following washout of LY367385, the DHPG-LTD was re-
138
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
A
a
b
D
c
a
b
c
0.4 mV
0.2 mV 140
20 ms
350
DHPG (100 µM)
120
LY367385 (100 µM)
Coastline (% Baseline)
fEPSP Slope (% Baseline)
5 ms
100 80
a
60
c
40
b
20 0
DHPG (100 µM)
300
LY367385 (100 µM)
250 200
a
150
b c
100 50 0
0
20
40
60
80
100
120
140
160
0
20
40
60
Time (min)
100
120
140
160
E 140
250
DHPG (100 µM)
120
LY367385 (100 µM)
100 80 60 40 20
(n = 6)
0 0
20
40
60
80
100
120
140
Coastline (% Baseline)
fEPSP Slope (% Baseline)
B
DHPG (100 µM)
LY367385 (100 µM)
200 150 100 50
(n = 6)
0
160
0
20
40
60
Time (min)
80
100
120
140
160
Time (min)
C
F 140
250
DHPG (100 µM)
120
MPEP (10 µM)
100 80 60 40 20
(n = 4)
0 0
20
40
60
80
100
120
Time (min)
Coastline (% Baseline)
fEPSP Slope (% Baseline)
80
Time (min)
DHPG (100 µM)
MPEP (10 µM)
200 150 100 50
(n = 4)
0 0
20
40
60
80
100
120
Time (min)
Fig. 2. The effects of the subtype-selective antagonists, LY367385 and MPEP, on the maintenance of DHPG-LTD and DHPG-induced epileptiform bursting. A. Representative example, and B. pooled data (n ¼ 6) showing that chemical LTD of Schaffer collateral-CA1 fEPSPs induced by DHPG (100 mM, 10 min), was transiently reversed by the mGlu1 selective antagonist LY367385 (100 mM, 20 min). Sample traces are taken from the time points indicated in the figure (in this and subsequent figures, traces are an average of 4 consecutive sweeps). LTD was 69 5%, 82 3% and 73 6% of baseline levels respectively before, during and after LY367385. C. Pooled data (n ¼ 4) illustrating equivalent experiment with the mGlu5 selective antagonist, MPEP (10 mM; 20 min). LTD was 81 3%, 85 3% and 80 4% of baseline before, during and after MPEP, respectively. D. Single example (same experiment as panel A) and E. pooled data (n ¼ 6) of ‘coastline’ measurement of fEPSP waveforms showing persistent enhancement of evoked epileptiform bursting by DHPG and its reversal by LY367385. Coastline values were 171 14%, 140 8% and 182 18% of baseline before, during and after LY367385, respectively. F. Pooled coastline plot (n ¼ 4) showing no effect of MPEP on the DHPG-induced enhancement of evoked epileptiform bursting with coastline values of 146 14%, 144 22% and 137 15% before, during and after application of MPEP, respectively.
established (Fig. 2A and B). Bath application of the mGlu5 antagonist, MPEP (10 mM), had a much smaller effect (Fig. 2C). The present data, supporting a major role for mGlu1 receptors in the maintenance of DHPG-LTD, are in agreement with those of Volk et al. (2006) but not with those of Huang and Hsu (2006) and Ronesi et al. (2012); the reason for this discrepancy is presently unknown. It should, however, be noted that heterodimers between mGlu1 and mGlu5 receptors (Doumazane et al., 2011) and variability in proportions of heterodimers and homodimers could potentially lead to experimental differences. During these experiments, it became clear that, in addition to inducing LTD of synaptic transmission, exposure to DHPG-induced a persistent enhancement of evoked epileptiform activity that followed the initial fEPSPs in the presence of picrotoxin (Fig. 2DeF). Increases in excitability following DHPG administration are well documented (Davies et al., 1995). For example, an enhancement of spontaneous epileptiform activity following DHPG administration in the CA3 region of disinhibited slices has been extensively described (e.g. Merlin and Wong, 1997 and reviewed by Bianchi
et al., 2012), but overt epileptiform activity was not observed in our earlier experiments (Fitzjohn et al., 1998, 1999; Palmer et al., 1997), possibly due to methodological differences. Persistent, non-synaptic bursting network activity has also been observed following DHPG application to hippocampal slices, in the absence of Ca2þ-dependent synaptic transmission (Piccinin et al., 2008). Our study of synaptically-evoked epileptiform activity in the CA1 region of the hippocampus showed that, similar to DHPG-LTD, the increase in population spike activity was maintained for up to 2 h after DHPG application, and was reduced by LY367385, but not by MPEP, in a fully reversible manner (Fig. 2DeF). The major role played by mGlu1 receptors in this form of plasticity parallels that for the maintenance of spontaneous DHPG-induced epileptiform activity in CA3 (Merlin, 2002). 3.2. Group II mGlu receptor-mediated LTD In addition to group I mGlu receptor-mediated LTD, there are several reports of both synaptic (e.g. Altinbilek and Manahan-
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
Vaughan, 2009; Huang et al., 1997; Manahan-Vaughan, 1997; Yokoi et al., 1996) and chemical (e.g. Huang et al., 1999; Pöschel and Manahan-Vaughan, 2005; Tzounopoulos et al., 1998; Wostrack and Dietrich, 2009) LTD induced by group II mGlu receptor activation in various hippocampal pathways. Group II mGlu receptormediated LTD has also been described in several other central pathways, including the prefrontal cortex (Otani et al., 1999, 2002), the striatum (Kahn et al., 2001), the nucleus accumbens (Robbe et al., 2002) and the amygdala (Lin et al., 2000; Lucas et al., 2012). In the majority of cases, the ability of mGlu receptor antagonists to affect the maintained phase of group II mGlu receptormediated LTD has not been explored, or not been observed. We have reported that the fEPSP in CA1 evoked by stimulation of the temporo-ammonic pathway (TAP) is inhibited by DCG-IV (0.3e1 mM) in a fully reversible manner (Ceolin et al., 2011; Hanna et al., 2012). By contrast, using these same methods in parasagittal slices from 12 to 16 day old rats, inhibition induced by bath application of a more potent mGlu2/3 agonist, LY379268 (0.3 mM; Fig. 3A and B) was maintained for several hours. Interestingly, LY341495 (0.3 mM), a concentration likely to act preferentially as an antagonist of group II mGlu receptors (Kingston et al., 1998), fully restored this LTD of the fEPSP back to baseline values (Fig. 3A and B). This antagonism was, however, at least partially reversible on washout of LY341495, suggesting a maintained chemical LTD mediated by group II mGlu receptors in this pathway in rat slices (Fig. 3A and B). We extended these observations to TAP stimulation in coronal slices from 10 to 16 week old CD-1 mice (Fig. 3C) and investigated the role of mGlu2 and mGlu3 receptors by comparing the effects of both LY379268 (0.3 mM) and LY341495 (0.3 mM) in mGlu3/ or mGlu2/ mice with those in their wild-type (WT) littermates (Linden et al., 2005). Fig. 3D illustrates the data showing
A
a
b
c
that specific activation of mGlu2 receptors is able to induce an LTD of the CA1 fEPSP in these knock-out mice comparable to that seen in WT animals (Fig. 3D) whereas activation of mGlu3 receptors alone induces a smaller LTD. In all three cases, LY341495 induced a reversible reversal of the LTD (Fig. 3D). To further investigate the role of mGlu2 receptors in chemical LTD, the effects of DCG-IV and of LY395756, a selective mGlu2 receptor agonist but an antagonist at mGlu3 receptors (Ceolin et al., 2011; Dominguez et al., 2005) were examined in the mossy fibre pathway to CA3 (Fig. 4). In a previous study we demonstrated that sensitivity to DCG-IV in the mossy fibre pathway was dependent on experimental conditions and in particular on slice orientation (Sherwood et al., 2012). We therefore confined this investigation of group II mGlu receptor-mediated LTD of mossy fibre transmission to transverse slices, in which sensitivity to DCG-IV (Fig. 4A) and facilitation to increased stimulation frequency (Fig. 4B) were used to identify the mossy fibre input to CA3 (Sherwood et al., 2012; Weisskopf et al., 1994). The reduction in fEPSP amplitude by DCG-IV (0.1 mM) outlasted the drug application (Fig. 4A), as reported previously with 1 mM DCG-IV (Wostrack and Dietrich, 2009). This chemical LTD was also elicited by the mGlu2 receptor agonist, LY395756, (10e30 mM; Fig. 4C), which in turn was reversibly reversed on application of LY341495 (0.3 mM; Fig. 4C). We have also extended our observations from hippocampal fields to amygdala neurons by studying the effect of the group II mGlu receptor antagonist, LY341495 (0.3 mM) on the LTD evoked by DCG-IV (1 mM) using patch-clamp recordings from neurons within the lateral amygdala of 5e6 week old mice (Lucas et al., 2012). The evoked EPSC remained depressed for at least 1 h after removal of DCG-IV (Fig. 5A and C) but LY341495 fully reversed the LTD in 6 out of 8 neurons examined (Fig. 5B and C). With respect to
C
d
139
a
b
c
d
0.4 mV
10 ms
LY379268
LY341495 (0.3 µM)
(0.03 µM)(0.3 µM) c
a
b 0
30
60
90
d
fEPSP Amp (% Baseline)
fEPSP Slope (% Baseline)
0.5 mV 180 160 140 120 100 80 60 40 20 0 -20
180 160 140 120 100 80 60 40 20 0
120 150 180 210 240 270 300 330 360 390
LY379268 (0.3 µM)
a
c
0
30
60
90
120
LY341495 (0.3 µM)
(0.03 µM)(0.3 µM)
(n = 4) 30
60
90
120 150 180 210 240 270 300 330 360 390
Time (min)
fEPSP Amp (% Baseline)
fEPSP Slope (% Baseline)
D
0
150
180
210
240
270
300
330
360
Time (min)
B LY379268
d
b
Time (min)
180 160 140 120 100 80 60 40 20 0
10 ms
LY341495 (0.3 µM)
WT (n = 12) 180 160 140 120 100 80 60 40 20 0
mGlu3-/- (n = 6)
LY379268 (0.3 µM)
0
30
60
90
120
mGlu2-/- (n = 6)
LY341495 (0.3 µM)
150
180
210
240
270
300
330
360
Time (min)
Fig. 3. The group II mGlu receptor antagonist, LY341495, reverses LY379268 induced LTD in the temporo-ammonic pathway of the hippocampus. A. Representative example, and B. pooled data (n ¼ 4) showing that chemical LTD of TAP-CA1 fEPSPs induced in rat hippocampal slices by LY379268 (0.03 and 0.3 mM, 40 min each) was reversed by the group II mGlu receptor antagonist, LY341495 (0.3 mM; 60 min). Sample traces are taken from the time points indicated in the figure. LTD was 23 9%, 102 23% and 70 21% of baseline levels before, during and after LY341495, respectively. C. Representative experiment from a CD-1 wild-type (WT) mouse and D. pooled data from transgenic and WT littermate mice showing that LTD induced by LY379268 (0.3 mM, 30 min) in hippocampal slices was transiently reversed by LY341495 (0.3 mM, 60 min). Specifically, the results from mGlu3/ mice (B; n ¼ 6) and WT mice (C; n ¼ 12) were very similar, whereas in mGlu2/ mice ( ; n ¼ 6) the LTD was significantly smaller. Thus the values were 29 3%, 35 5% and 74 5% of baseline before, 103 10%, 126 9% and 95 9% of baseline during and 52 7%, 58 5% and 76 8% of baseline after LY341495 application for mGlu3/, WT and mGlu2/ CD1 mice, respectively.
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
a
fEPSP Amp (% Baseline)
A
a&b
a&c 0.3 mV 5 ms
120
DCG-IV (0.1 µM)
100
a 80
b c
60
(n = 5)
40 0
10
20
30
40
50
60
70
80
a
A
90
100
EPSC Amp (% Baseline)
140
b
100 pA 10 ms
140
DCG-IV (1 µM)
120 100
a
80 60
c
40
b
20 0 0
10
20
30
40
110
i & ii i
ii
400 300 200
0.3 mV 5 ms
(n = 5)
EPSC Amp (% Baseline)
0.03
b
70
80
90
c 100 pA 10 ms
140
DCG-IV (1 µM)
120
LY341495 (0.3 µM)
100
a
80
c
60 40
b
20 0 0
10
20
30
100
40
50
60
70
80
90
Time (min) 0
30
60
90
a
a&b
a&c
a&d 0.3 mV 5 ms
120
LY395756 (10-30 µM)
LY341495 (0.3 µM)
100
a 80
c d (n = 5)
40 0
30
60
90
120
150
180
140
DCG-IV (1 µM)
120
210
240
Time (min) Fig. 4. The selective mGlu2 agonist, LY395756, induces reversible LTD in the mossy fibre pathway to CA3. A. Pooled data (n ¼ 5) showing that DCG-IV (0.1 mM, 20 min) induced an LTD of the evoked fEPSP. Following the acute reduction during DCG-IV application, the fEPSP remained at 77 8% of baseline 60 min after washout. B. Pooled data (n ¼ 5) illustrating the facilitation (365 21%) observed following switching from 0.033 Hz to 1.0 Hz characteristic of the MF-CA3 pathway in transverse hippocampal slices. The inset to the right shows sample traces taken at the points indicated in the figure. C. Pooled data (n ¼ 5) showing acute reduction of the fEPSP by LY395756 (10 (n ¼ 2) and 30 (n ¼ 3) mM; 20 min) and after 80 min of washout an LTD of 80 7%. The subsequent application of LY341495 (0.3 mM; 20 min) returned the fEPSP to 93 7% of baseline which reversed to 81 6% of baseline 70 min after washout of LY341495. Sample traces are taken from the time points indicated in the figure. In these latter experiments, the application of 3 mM DCG-IV depressed the transverse mossy fibre field response to 19 2% of the baseline (n ¼ 5, data not shown).
receptor subtype, we have previously reported, using knock-out mice, that both mGlu2 and mGlu3 receptors are separately capable of inducing this DCG-IV-LTD in amygdala neurons (Lucas et al., 2012).
LY341495 (0.3 µM)
100 80 60 40 20 0 0
b
60
DCG-IV alone (n=6) DCG-IV reversed with LY341495 (n=8)
C
Stimulation Number
EPSC Amp (% Baseline)
Field Response Amp (% Baseline)
60
1.00
500
C fEPSP Amp (% Baseline)
a
B
Hz
50
Time (min)
Time (min)
B
c
10
20
30
40
50
60
70
80
90
Time (min) Fig. 5. The group II mGlu receptor antagonist, LY341495, reverses DCG-IV-induced LTD in the lateral amygdala. A. Representative patch-clamp experiment showing that application of DCG-IV (1 mM, 20 min) induced an LTD of EPSC amplitude. Sample traces are taken from the time points indicated in the figure. B. Representative experiment showing that application of LY341495 (0.3 mM, 30 min) reversed the LTD induced by DCG-IV (1 mM, 20 min). C. Pooled data showing the effect of DCG-IV alone (n ¼ 6) and the reversal of the DCG-IV-induced LTD by LY341495 (n ¼ 8). In the latter data set, EPSC measurements were 56 8% and 79 10% of baseline before and during LY341495 application, respectively.
3.3. Group III mGlu receptor-mediated LTD Having demonstrated antagonist reversal of established chemical LTD in relation to group I and II mGlu receptors, it was of interest to see if this phenomenon extended to agonists of group III mGlu receptors. The paucity of subtype-selective and potent agonists for group III mGlu receptors has, thus far, limited the investigation of group III mGlu receptor subtypes in synaptic plasticity in general, and chemical LTD in particular. There is, however, some evidence for group III mGlu receptor-mediated LTD. For instance, the mGlu4 and mGlu8 preferring agonist, L-(þ)-2-amino-4phosphonobutyric acid (L-AP4) (Schoepp et al., 1999) has been
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
4. Discussion In the present study we have extended our analysis of DHPGLTD in CA1 and its reversal by group I mGlu receptor antagonists. In particular, we have shown that antagonists are able to reverse both the LTD of stimulus-evoked synaptic activity and stimulusevoked epileptiform discharges. Significantly, we have demonstrated that this unusual phenomenon of reversible reversal of LTD by antagonists also applies to members of mGlu receptor groups II (mGlu2 and mGlu3) and III (mGlu8). 4.1. Chemical forms of LTD induced by group II and III mGlu receptors We have extended the number of pathways where a chemical form of group II or III mGlu receptor mediated LTD can be observed and demonstrated that appropriate antagonists reduce the maintenance phase of the LTD with subsequent reversal on washout of the antagonist. With respect to group II mGlu receptors, we have found that activation of mGlu2 can induce LTD in the TAP. Interestingly, this LTD was consistently observed with LY379268 but not with DCG-IV. LY379268 is a more potent mGlu2/3 agonist than DCG-IV (Schoepp et al., 1999) and so whether LTD is induced or not could relate to the level of intrinsic activity of the agonist. However, this cannot be the sole explanation, since DCG-IV was capable of inducing LTD in both the MF input to the CA3 (Fig. 4) and the cortical input to the lateral amygdala (Fig. 5 and Lucas et al., 2012). Perhaps, the ability of an agonist to induce LTD depends upon both its intrinsic activity for the receptor and the receptor density at the particular pathway under investigation. Our studies in the TAP-CA1 and MF-CA3 pathways identified a primary role for mGlu2
a
A
b 0.5 mV
fEPSP Slope (% Baseline)
10 ms 140
MDCPG (30 µM)
120 100
a
80
b
60 40 20
(n = 13)
0 0
10
20
30
40
50
60
Time (min) a
B
b
c
d
fEPSP Slope (% Baseline)
0.5 mV 10 ms
140
DCPG (100 µM)
120
MDCPG (30 µM)
100
a
80
c
60
d
40
b
20 0 0
30
60
90
120
150
180
210
Time (min)
C fEPSP Slope (% Baseline)
shown to induce an LTD in ventral septal neurons in vitro (Tolchard et al., 2000) and in CA1 neurons in vivo (Naie et al., 2006). L-AP4 (200 mM) was also found to induce an LTD at mossy fibre synapses in interneurons but, significantly, not pyramidal neurons (Pelkey et al., 2005). The lack of effect with lower concentrations of L-AP4, selective for mGlu4 and mGlu8 receptors, indicated a role for mGlu7 receptors in this form of LTD. Although few studies have directly investigated plasticity following specific activation of mGlu8 receptors, incomplete recovery from the depressant actions of L-AP4 and DCPG, an mGlu8 receptor preferring agonist (Thomas et al., 2001), has been observed in the bed nucleus of stria terminalis (Gosnell et al., 2011; Grueter and Winder, 2005). During a recent investigation into the selectivity of DCPG as an mGlu8 receptor agonist (Mercier et al., 2013), we noticed that bath application of 100 mM DCPG induced a long lasting depression of the lateral perforant path (LPP) evoked fEPSP in the dentate gyrus. It was found that DCPG was a selective mGlu8 agonist at low concentrations but, at concentrations greater than 1 mM, DCPG also acted at other mGlu receptors, principally mGlu2 (Mercier et al., 2013). In order to avoid non-selective effects via mGlu2 receptors, experiments were conducted in a strain of Han Wistar rats (10 weeks old) known not to express mGlu2 receptors (Ceolin et al., 2011). We first demonstrated that the mGlu8 receptor antagonist, MDCPG (30 mM) (Mercier et al., 2013; Thomas et al., 2001), had no effect on baseline fEPSPs (Fig. 6A) and then that 30 min perfusion with DCPG (100 mM) induced a long lasting depression of fEPSPs (Fig. 6B and C), which we term DCPG-LTD. This novel form of chemical LTD was fully reversed to baseline levels by the application of 30 mM MDCPG and, as with the group I and II chemical LTD, washout of the antagonist re-instated the LTD (Fig. 6B and C). Again this indicates that activity of the mGlu receptor, in this case mGlu8, is required for the maintenance of DCPG-LTD, and extends this curious phenomenon to all three mGlu receptor subgroups.
141
140
DCPG (100 µM)
120
MDCPG (30 µM)
100 80 60 40 20
(n = 5)
0 0
30
60
90
120
150
180
210
Time (min) Fig. 6. The mGlu8 receptor antagonist, MDCPG, reverses DCPG induced LTD in the lateral perforant path input to the dentate gyrus. A. Pooled data (n ¼ 13 from 9 rats) showing no effect of mGlu8 receptor antagonist MDCPG (30 mM, 30 min) on baseline synaptic transmission in the LPP. Sample traces are taken from the time points indicated in the figure. B. Representative experiment and C. pooled data (n ¼ 5 from 4 rats) showing that LTD induced by the mGlu8 receptor agonist DCPG (100 mM, 30 min) was reversed by MDCPG (30 mM, 30 min) such that fEPSP slope was reduced to 63 5%, 98 3% and 72 5% of baseline before, during and after MDCPG application, respectively.
receptors in LTD, but also suggest a minor role for mGlu3 receptors. In the lateral amygdala, we have previously shown that both mGlu2 and mGlu3 receptors are capable of inducing LTD following application of DCG-IV (Lucas et al., 2012). Other reports (e.g. Johnson et al., 2011; Yokoi et al., 1996) emphasize the role of mGlu2 receptors as inducers of LTD in the CNS. Regarding group III mGlu receptors, we have shown that mGlu8 receptors can induce LTD, in this case in the LPP projection to the dentate gyrus. To achieve this effect it was necessary to apply a high concentration, relative to that needed to induce a transient synaptic
142
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
depression (Mercier et al., 2013). However, we are confident that the LTD is due, primarily at least, to activation of mGlu8 receptors, since we performed our experiments in a rat strain that is devoid of mGlu2 receptors, which primarily mediate the non-specific effects of DCPG (Mercier et al., 2013). The maintained phase of the LTD was reduced by MDCPG, which is a selective antagonist for mGlu8 receptors (Mercier et al., 2013). However, the possibility that mGlu2 receptors contribute to DCPG-LTD in wild-type tissue cannot be ruled out. Indeed, lower concentrations of DCPG (e.g., 30 mM) induced a persistent depression in wild-type but not in mGlu2/ mice (Mercier et al., 2013). 4.2. Reversal of mGlu receptor-mediated LTD by antagonists The major focus of the present study has been to investigate the reversal of the maintenance of LTD by mGlu receptor antagonists. Our main conclusion is that the reversible reversal of LTD is observed across all three subgroups and so is likely to relate to a general property of these G-protein coupled receptors. So what are the possible explanations for the reversal of these chemical forms of LTD by the antagonists? Broadly speaking there are two primary explanations: 1) the exogenous agonist remains in the tissue and is available for activation of the receptor for long periods of time or 2) the receptor remains active in the absence of the exogenous agonist. 4.3. Exogenous agonist remains in the tissue A simple explanation of the antagonist reversal of LTD is that the exogenous agonist remains bound to the receptor and exposure to the antagonist displaces the agonist from the receptor and it is then removed from the slice. For example, Wostrack and Dietrich (2009) reported that LY341495 (3 mM) reversed the maintained long lasting inhibition of the mossy fibre input to dentate gyrus following DCG-IV (1 mM) but did not observe any recovery from the effects of this antagonist, unlike the present experiments where recovery followed washout of LY341495 (0.3 mM). Wostrack and Dietrich (2009) interpreted their data as indicating that the agonist remained bound to the group II receptor until displaced by LY341495. Although this is an attractively simple idea, it is difficult to reconcile with the fact that the LTD of each of the three mGlu receptor groups re-appears after washout of the antagonist, with the fact that the depression of synaptic activity by the same concentration of a group III agonist is long lasting (LTD) at one synapse but not at another synapse in the same preparation (Pelkey et al., 2005) and with the fact that long lasting inhibition of afterhyperpolarisations by DHPG remains intact after multiple applications of MCPG (Young et al., 2012). To retain this hypothesis, it is therefore necessary to conjecture that the exogenous agonist is sequestered in some form within the tissue such that it is able to reactivate the receptor and the transduction processes underlying the LTD once the antagonist is washed out. One such possibility is that the agonist is internalised along with the receptor and the receptor may continue to signal (Shenoy and Lefkowitz, 2011). This in itself seems unlikely to be the entire explanation since the extracellularly applied antagonists are able to reverse the LTD, unless this is also internalised and therefore prevents the transduction mechanisms internally. Alternatively, it is also possible to envisage a situation where a bound agonistereceptor complex is continuously recycled back to the membrane where its actions would be blocked during antagonist application. Thus, arrestins or other mediators (Hong et al., 2009; Mundell et al., 2001) are involved in resensitisation as well as desensitisation of G-protein coupled receptors (Bockaert et al., 2004; Ferguson, 2001), although whether this can occur with agonist bound during the recycling event is unclear. Such a long
lasting recycling of the receptoreagonist complex could be envisaged for a high affinity agonist, such as LY379268, but seems less likely an explanation for the LTD induced by relatively low affinity agonists such as DHPG and DCPG. A second type of sequestering of an agonist might be uptake into glia and/or presynaptic terminals. Subsequent hetero-exchange with, for example, synaptically released glutamate, or release from terminals would provide a continued background of the exogenous agonist and hence activation of the respective mGlu receptor. Group II mGlu receptor agonists inhibit the cysteine-glutamate antiporter via astrocytic mGlu3 receptors (Baker et al., 2002) and stimulate GLAST and GLT1 as well as the glutamate astrocycte transporters (Aronica et al., 2003; D’Antoni et al., 2008) all increasing uptake of glutamatelike substrates. Any subsequent release of such substrates is well positioned to activate presynaptic mGlu2/3 receptors (Kalivas, 2009). DHPG is also known to increase expression of the glutamate transporter, EAAC1 (Ross et al., 2011) and along with several other phenylglycine compounds may be a substrate for glutamate uptake (Gochenauer and Robinson, 2001). Nevertheless with all the above putative schemes, it seems likely that any recycling exogenous agonist would diffuse from the slice in a finite time and not result in LTD that is capable of lasting for hours. 4.4. Receptors remain active in the absence of exogenous agonist One possibility is that endogenous glutamate maintains activity of the mGlu receptors. For example, exposure to the exogenous agonist could lead to a chronic increase in the perisynaptic glutamate concentration, or to an increase in the affinity and/or efficacy of the mGlu receptors or to redistribution of receptors into a zone of higher glutamate exposure. Potentially, one or more of these mechanisms could result in an increase in glutamate receptor tone sufficient to maintain LTD. In the case of DHPG, the induced epileptiform activity reported here and previously (Merlin and Wong, 1997) may increase ambient glutamate levels and hence lead to continued mGlu receptor activation. This would constitute a form of “autopotentiation,” i.e., transient agonist-mediated selective activation results in long lasting enhanced responsiveness of group I mGlu receptors, allowing endogenous glutamate to now sustain the activation of the receptors (Bianchi et al., 2012). However, if ambient levels of glutamate do increase, one might then expect several of the mGlu receptor subtypes to be activated and contribute to the LTD. This appears not to be the case since we found that LTD induced by an agonist for a particular mGlu receptor subtype was essentially reversed fully by antagonists selective for that particular receptor (Figs. 1e6) and, where tested, not by antagonists selective for the other mGlu receptors (e.g. (Palmer et al., 1997)). Simplistically, one would not expect this hypothesis to extend to LTD evoked by agonists selective for group II and III mGlu receptors. These are largely presynaptic autoreceptors (Nicoletti et al., 2011; Niswender and Conn, 2010) which tend to reduce extracellular glutamate levels and hence would not be expected to contribute to their maintained activation. mGlu3 receptors located on glia could, however, contribute to changes in glutamate dynamics, possibly leading to agonist release onto presynaptic mGlu2 receptors (see 3.1). This is, however, not likely to be the major explanation for group II agonist-LTD since in the mGlu3/ mice, the extent of the LTD remained similar to that in wild-type mice (Fig. 3). The possibility that the induction of mGlu receptor mediated LTD involves an increase in receptor affinity could underlie each of the forms of LTD studied here. In each case, the acute affect of mGlu receptor activation is a depression of synaptic transmission and so an increase in the receptor affinity sufficient to permit its activation by ambient levels of L-glutamate would lead to a persistent
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
reduction in synaptic strength. There is precedence for such a mechanism involving G-protein associated kainate receptors (Lauri et al., 2006). In neonatal tissue, an LTP stimulus leads to a rapid decrease in the affinity of kainate receptors such that they are no longer activated by ambient L-glutamate. This results in an increase in the probability of release. Conceivably, mGlu receptor mediated LTD could elicit the opposite effect; an increase in receptor affinity. This would need to be quite a substantial alteration, however, to account for the plasticity in tissue obtained from adult animals, where ambient L-glutamate concentrations are likely to be maintained at a very low level. Another possibility is that, after prolonged exposure to the exogenous agonist, the mGlu receptors become constitutively active, in a manner similar to that recently suggested for DHPGinduced epileptiform activity (Young et al., 2013). Prolonged or constitutive activity of G-proteins and downstream transduction processes per se is unlikely to be the explanation, since this would not be reversed by mGlu receptor antagonists. Constitutive activity of the receptor implies a conformational change such that even in the agonist-unbound state, the receptor continues to couple to intracellular G-proteins and the downstream transduction processes. Constitutive activity of G-protein coupled receptors is well
143
documented (Cotecchia et al., 1990; Milligan et al., 1995; Seifert and Wenzel-Seifert, 2002; Tiberi and Caron, 1994, reviewed by Milligan, 2003) and has been proposed for mGlu receptors (Ango et al., 2001; Hermans and Challiss, 2001; Joly et al., 1995; Prézeau et al., 1996, and see Bockaert et al., 2004). Several conformational changes may induce such constitutive activity. Firstly, G-protein coupled receptors can interact with scaffolding and interacting proteins forming large macromolecular complexes. Such complexes have been termed ‘signalosomes’ and are capable of constitutive activity (Bockaert et al., 2004). One example of such a functional interaction with group I mGlu receptors occurs with the Homer class of scaffolding proteins. Normally Homer3 is present in the submembrane domains of these mGlu receptors and prevents constitutive signalling. This negative influence is not present in non-neuronal cells and hence constitutive activity is more common in heterologous systems especially at high levels of expression (e.g. Tiberi and Caron, 1994). Neuronal synaptic activity, however, can lead to expression of Homer1a which competes with Homer3 and allows G-protein activation by group I mGlu receptors to occur in the absence of agonist (Ango et al., 2001; Bockaert et al., 2004; Kammermeier and Worley, 2007). Interestingly, this form of constitutive activity is blocked by inverse agonists, e.g. MPEP, but
(A) Baseline
Active Conformation
(B) Induction Inactive Conformation
(C) Expression
Endogenous Agonist
Exogenous Agonist
Exogenous Antagonist
Fig. 7. Hypothetical mechanism to explain mGlu receptor-LTD and its reversible reversal by antagonists. Baseline: mGlu receptors can bind synaptically released L-glutamate to become activated transiently or bind an antagonist that prevents their activation by L-glutamate (single monomers are shown for clarity). Under baseline conditions antagonists have little or no effect on synaptic strength and similarly the transient activation of mGlu receptors by L-glutamate has no discernible effect. Thus, the receptors display little or no constitutive activity (as indicated by the arrows favouring the inactive conformation of the receptor). Induction: Prolonged application of an exogenous agonist (such as DHPG) leads to a modification of the mGlu receptor such that there is an increased probability of it being in the constitutively active state. Expression: Upon washout of the exogenous agonist, the mGlu receptor remains in the constitutively active state, and actively depresses synaptic transmission. However, the constitutively active receptor is in equilibrium with the inactive conformation, which can bind the antagonist. Hence application of the antagonist leads to inhibition of the response but upon washout the equilibrium between inactive and constitutively active states is restored.
144
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
not by competitive antagonists, e.g. LY367385 (Ango et al., 2001; Carroll et al., 2001). Therefore, it seems unlikely that this mechanism is the explanation for the LTDs induced by the agonists from each of the three mGlu receptor groups reported here, because all were sensitive to competitive antagonists. Secondly, in many G-protein coupled receptors, the agonist recognition site is within the heptahelical domain, although this is not the case for the transmitter glutamate with mGlu receptors. Nevertheless, class C receptors, to which mGlu receptors belong, have an equivalent domain which controls G-protein coupling (Malherbe et al., 2006). This site and other sites on the cytoplasmic and extracellular loops (Halls, 2012; Seifert and Wenzel-Seifert, 2002) can be modulated in the absence of a glutamate-like agonist to become constitutively active (Jensen and BräunerOsborne, 2007; Jensen et al., 2002; Yamashita et al., 2004; Yanagawa et al., 2009). Negative and positive allosteric modulators bind to this region or the equivalent region in other mGlu receptors (Carroll et al., 2001; Goudet et al., 2004; Malherbe et al., 2006; Pagano et al., 2000). In constitutive activity via this mechanism, both the orthosteric agonists and antagonists would be expected to be less effective (Yanagawa et al., 2009). A further possibility is shown schematically in Fig. 7. In the presence of agonist the clam shaped glutamate recognition domain adopts a closed (active) conformation whereas in the presence of the antagonist the clam is held open and hence inactive. However, in the absence of agonist or antagonist, these G-protein coupled receptors generally adopt a closed (active) conformation with a low probability relative to an open (active) conformation. It is therefore possible that prolonged administration of an agonist can induce a long lasting change in the receptor protein such that the closed (active) conformation becomes thermodynamically more probable. This scheme still allows an orthosteric antagonist to bind and hence block the response but upon washout the high probability of the closed state remains. Such a mechanism leading to constitutive activity has been induced in the presence of Gd3þ ions and possibly Ca2þ ions for the mGlu1 receptor (Jensen et al., 2002). Such constitutive activity, being sensitive to both orthosteric and allosteric antagonists, could underlie the observations in the present study. Presently, however, there is insufficient evidence to distinguish between the above possible mechanisms and to determine whether there are common features in the maintenance of LTD by all three groups of mGlu receptors. Further investigation is required. 4.5. Therapeutic relevance of the antagonist reversal phenomenon Does the ability of mGlu receptor antagonists to reverse the sustained effects of mGlu receptor ligands have any functional significance? With respect to group I mGlu receptors, the ability of antagonists to reverse epileptiform activity could be targeted therapeutically. It will be interesting to determine whether interictal, and/or ictal events present in epileptic human tissue are sensitive to mGlu receptor antagonists. Interestingly, mGlu5 receptor antagonists have been reported to have efficacy in animal models of various diseases, including autism spectrum disorders, fragile-X syndrome and tardive dyskinesia (Bianchi et al., 2012; Gürkan and Hagerman, 2012; Michalon et al., 2012; Rylander et al., 2009). The putative therapeutic effects could be due to antagonist prevention of the repeated induction of a pathological form of neuronal plasticity, or to antagonists reversing prolonged expression of the pathological plasticity. Conceivably, in various disease states, the affected mGlu receptor pathway may become constitutively active, in which case a therapeutic effect could be achieved by inhibition of the constitutive activity.
Acknowledgements This work was supported by Eli Lilly, BBSRC, MRC and Wellcome Trust (all UK) and the WCU Program (Republic of Korea). References Altinbilek, B., Manahan-Vaughan, D., 2009. A specific role for group II metabotropic glutamate receptors in hippocampal long-term depression and spatial memory. Neuroscience 158, 149e158. Anderson, W.W., Collingridge, G.L., 2007. Capabilities of the WinLTP data acquisition program extending beyond basic LTP experimental functions. J. Neurosci. Methods 162, 346e356. Ango, F., Prézeau, L., Muller, T., Tu, J.C., Xiao, B., Worley, P., Pin, J.P., Bockaert, J., Fagni, L., 2001. Agonist-independent activation of metabotropic glutamate receptors by the intracellular protein Homer. Nature 411, 962e964. Aronica, E., Gorter, J.A., Ijlst-Keizers, H., Rozemuller, A.J., Yankaya, B., Leenstra, S., Troost, D., 2003. Expression and functional role of mGluR3 and mGluR5 in human astrocytes and glioma cells: opposite regulation of glutamate transporter proteins. Eur. J. Neurosci. 17, 2106e2118. Baker, D.A., Xi, Z.-X., Shen, H., Swanson, C.J., Kalivas, P.W., 2002. The origin and neuronal function of in vivo nonsynaptic glutamate. J. Neurosci. 22, 9134e9141. Bianchi, R., Wong, R.K.S., Merlin, L.R., 2012. Glutamate receptors in epilepsy: group I mGluR-mediated epileptogenesis. In: Noebels, J.L., Avoli, M., Rogawski, M.A., Olsen, R.W., Delgado-Escueta, A.V. (Eds.), Jasper’s Basic Mechanisms of the Epilepsies (Internet). National Center for Biotechnology Information (US), Bethesda (MD). Bockaert, J., Fagni, L., Dumuis, A., Marin, P., 2004. GPCR interacting proteins (GIP). Pharmacol. Ther. 103, 203e221. Carroll, F.Y., Stolle, A., Beart, P.M., Voerste, A., Brabet, I., Mauler, F., Joly, C., Antonicek, H., Bockaert, J., Müller, T., 2001. BAY36-7620: a potent noncompetitive mGlu1 receptor antagonist with inverse agonist activity. Mol. Pharmacol. 59, 965e973. Ceolin, L., Kantamneni, S., Barker, G.R.I., Hanna, L., Murray, L., Warburton, E.C., Robinson, E.S.J., Monn, J.A., Fitzjohn, S.M., Collingridge, G.L., Bortolotto, Z.A., Lodge, D., 2011. Study of novel selective mGlu2 agonist in the temporoammonic input to CA1 neurons reveals reduced mGlu2 receptor expression in a Wistar substrain with an anxiety-like phenotype. J. Neurosci. 31, 6721e6731. Collingridge, G.L., Olsen, R.W., Peters, J., Spedding, M., 2009. A nomenclature for ligand-gated ion channels. Neuropharmacology 56, 2e5. Collingridge, G.L., Peineau, S., Howland, J.G., Wang, Y.T., 2010. Long-term depression in the CNS. Nat. Rev. Neurosci. 11, 459e473. Cotecchia, S., Exum, S., Caron, M.G., Lefkowitz, R.J., 1990. Regions of the alpha 1adrenergic receptor involved in coupling to phosphatidylinositol hydrolysis and enhanced sensitivity of biological function. Proc. Natl. Acad. Sci. U.S.A. 87, 2896e2900. D’Antoni, S., Berretta, A., Bonaccorso, C.M., Bruno, V., Aronica, E., Nicoletti, F., Catania, M.V., 2008. Metabotropic glutamate receptors in glial cells. Neurochem. Res. 33, 2436e2443. Davies, C.H., Clarke, V.R.J., Jane, D.E., Collingridge, G.L., 1995. Pharmacology of postsynaptic metabotropic glutamate receptors in rat hippocampal CA1 pyramidal neurones. Br. J. Pharmacol. 116, 1859e1869. Dominguez, C., Prieto, L., Valli, M.J., Massey, S.M., Bures, M., Wright, R.A., Johnson, B.G., Andis, S.L., Kingston, A., Schoepp, D.D., 2005. Methyl substitution of 2-aminobicyclo [3.1. 0] hexane 2, 6-dicarboxylate (LY354740) determines functional activity at metabotropic glutamate receptors: identification of a subtype selective mGlu2 receptor agonist. J. Med. Chem. 48, 3605e3612. Doumazane, E., Scholler, P., Zwier, J.M., Trinquet, E., Rondard, P., Pin, J.P., 2011. A new approach to analyze cell surface protein complexes reveals specific heterodimeric metabotropic glutamate receptors. FASEB J. 25, 66e77. Dunayevich, E., Erickson, J., Levine, L., Landbloom, R., Schoepp, D.D., Tollefson, G.D., 2008. Efficacy and tolerability of an mGlu2/3 agonist in the treatment of generalized anxiety disorder. Neuropsychopharmacology 33, 1603e1610. Faas, G.C., Adwanikar, H., Gereau, R.W., Saggau, P., 2002. Modulation of presynaptic calcium transients by metabotropic glutamate receptor activation: a differential role in acute depression of synaptic transmission and long-term depression. J. Neurosci. 22, 6885e6890. Ferguson, S.S.G., 2001. Evolving concepts in G protein-coupled receptor endocytosis: the role in receptor desensitization and signaling. Pharmacol. Rev. 53, 1e24. Fitzjohn, S.M., Bortolotto, Z.A., Palmer, M.J., Doherty, A.J., Ornstein, P.L., Schoepp, D.D., Kingston, A.E., Lodge, D., Collingridge, G.L., 1998. The potent mGlu receptor antagonist LY341495 identifies roles for both cloned and novel mGlu receptors in hippocampal synaptic plasticity. Neuropharmacology 37, 1445e1458. Fitzjohn, S.M., Kingston, A.E., Lodge, D., Collingridge, G.L., 1999. DHPG-induced LTD in area CA1 of juvenile rat hippocampus; characterisation and sensitivity to novel mGlu receptor antagonists. Neuropharmacology 38, 1577e1583. Fitzjohn, S.M., Palmer, M.J., May, J.E., Neeson, A., Morris, S.A., Collingridge, G.L., 2001. A characterisation of long-term depression induced by metabotropic glutamate receptor activation in the rat hippocampus in vitro. J. Physiol. 537, 421e430. Gallagher, S.M., Daly, C.A., Bear, M.F., Huber, K.M., 2004. Extracellular signalregulated protein kinase activation is required for metabotropic glutamate
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146 receptor-dependent long-term depression in hippocampal area CA1. J. Neurosci. 24, 4859e4864. Gladding, C.M., Fitzjohn, S.M., Molnár, E., 2009. Metabotropic glutamate receptormediated long-term depression: molecular mechanisms. Pharmacol. Rev. 61, 395e412. Gochenauer, G.E., Robinson, M.B., 2001. Dibutyryl-cAMP (dbcAMP) up-regulates astrocytic chloride-dependent l-[3H]glutamate transport and expression of both system xc subunits. J. Neurochem. 78, 276e286. Gosnell, H.B., Silberman, Y., Grueter, B.A., Duvoisin, R.M., Raber, J., Winder, D.G., 2011. mGluR8 Modulates excitatory transmission in the bed nucleus of the stria terminalis in a stress-dependent manner. Neuropsychopharmacology 36, 1599e1607. Goudet, C., Gaven, F., Kniazeff, J., Vol, C., Liu, J., Cohen-Gonsaud, M., Acher, F., Prézeau, L., Pin, J.P., 2004. Heptahelical domain of metabotropic glutamate receptor 5 behaves like rhodopsin-like receptors. Proc. Natl. Acad. Sci. U.S.A. 101, 378e383. Grillon, C., Cordova, J., Levine, L.R., Morgan, C.A., 2003. Anxiolytic effects of a novel group II metabotropic glutamate receptor agonist (LY354740) in the fearpotentiated startle paradigm in humans. Psychopharmacology 168, 446e454. Grueter, B.A., Winder, D.G., 2005. Group II and III metabotropic glutamate receptors suppress excitatory synaptic transmission in the dorsolateral bed nucleus of the stria terminalis. Neuropsychopharmacology 30, 1302e1311. Gürkan, C.K., Hagerman, R.J., 2012. Targeted treatments in autism and fragile X syndrome. Res. Autism Spect. Dis. 6, 1311e1320. Halls, M.L., 2012. Constitutive formation of an RXFP1-signalosome: a novel paradigm in GPCR function and regulation. Br. J. Pharmacol. 165, 1644e1658. Hanna, L., Ceolin, L., Lucas, S., Monn, J., Johnson, B., Collingridge, G., Bortolotto, Z., Lodge, D., 2012. Differentiating the roles of mGlu2 and mGlu3 receptors using LY541850, an mGlu2 agonist/mGlu3 antagonist. Neuropharmacology, 1e8. Hermans, E., Challiss, R., 2001. Structural, signalling and regulatory properties of the group I metabotropic glutamate receptors: prototypic family C G-proteincoupled receptors. Biochem. J. 359, 465. Hong, Y.H., Kim, J.Y., Lee, J.H., Chae, H.G., Jang, S.S., Jeon, J.H., Kim, C.H., Kim, J., Kim, S.J., 2009. Agonist-induced internalization of mGluR1a is mediated by caveolin. J. Neurochem. 111, 61e71. Hou, L., Klann, E., 2004. Activation of the phosphoinositide 3-kinase-Aktmammalian target of rapamycin signaling pathway is required for metabotropic glutamate receptor-dependent long-term depression. J. Neurosci. 24, 6352e 6361. Hu, N.W., Ondrejcak, T., Rowan, M.J., 2012. Glutamate receptors in preclinical research on Alzheimer’s disease: update on recent advances. Pharmacol. Biochem. Behav. 100, 855e862. Huang, C.-C., Hsu, K.-S., 2006. Sustained activation of metabotropic glutamate receptor 5 and protein tyrosine phosphatases mediate the expression of (S)-3,5dihydroxyphenylglycine-induced long-term depression in the hippocampal CA1 region. J. Neurochem. 96, 179e194. Huang, L., Killbride, J., Rowan, M.J., Anwyl, R., 1999. Activation of mGluRII induces LTD via activation of protein kinase A and protein kinase C in the dentate gyrus of the hippocampus in vitro. Neuropharmacology 38, 73e83. Huang, L.Q., Rowan, M.J., Anwyl, R., 1997. mGluR II agonist inhibition of LTP induction, and mGluR II antagonist inhibition of LTD induction, in the dentate gyrus in vitro. Neuroreport 8, 687e693. Huber, K.M., Kayser, M.S., Bear, M.F., 2000. Role for rapid dendritic protein synthesis in hippocampal mGluR-dependent long-term depression. Science 288, 1254e 1257. Jensen, A.A., Bräuner-Osborne, H., 2007. Allosteric modulation of the calciumsensing receptor. Curr. Neuropharm. 5, 180e186. Jensen, A.A., Greenwood, J.R., Bräuner-Osborne, H., 2002. The dance of the clams: twists and turns in the family C GPCR homodimer. Trends Pharmacol. Sci. 23, 491e493. Johnson, K.A., Niswender, C.M., Conn, P.J., Xiang, Z., 2011. Activation of group II metabotropic glutamate receptors induces long-term depression of excitatory synaptic transmission in the substantia nigra pars reticulata. Neurosci. Lett. 504, 102e106. Joly, C., Gomeza, J., Brabet, I., Curry, K., Bockaert, J., Pin, J., 1995. Molecular, functional, and pharmacological characterization of the metabotropic glutamate receptor type 5 splice variants: comparison with mGluR1. J. Neurosci. 15, 3970e 3981. Kahn, L., Alonso, G., Robbe, D., Bockaert, J., Manzoni, O.J., 2001. Group 2 metabotropic glutamate receptors induced long term depression in mouse striatal slices. Neurosci. Lett. 316, 178e182. Kalivas, P.W., 2009. The glutamate homeostasis hypothesis of addiction. Nat. Rev. Neurosci. 10, 561e572. Kammermeier, P.J., Worley, P.F., 2007. Homer 1a uncouples metabotropic glutamate receptor 5 from postsynaptic effectors. Proc. Natl. Acad. Sci. U.S.A. 104, 6055e 6060. Kingston, A.E., Ornstein, P.L., Wright, R.A., Johnson, B.G., Mayne, N.G., Burnett, J.P., Belagaje, R., Wu, S., Schoepp, D.D., 1998. LY341495 is a nanomolar potent and selective antagonist of group II metabotropic glutamate receptors. Neuropharmacology 37, 1e12. Kumar, A., Foster, T.C., 2007. Shift in induction mechanisms underlies an agedependent increase in DHPG-induced synaptic depression at CA3 CA1 synapses. J. Neurophysiol. 98, 2729e2736. Lauri, S.E., Vesikansa, A., Segerstråle, M., Collingridge, G.L., Isaac, J.T.R., Taira, T., 2006. Functional maturation of CA1 synapses involves activity-dependent loss
145
of tonic kainate receptor-mediated inhibition of glutamate release. Neuron 50, 415e429. Lin, H.C., Wang, S.J., Luo, M.Z., Gean, P.W., 2000. Activation of group II metabotropic glutamate receptors induces long-term depression of synaptic transmission in the rat amygdala. J. Neurosci. 20, 9017e9024. Linden, A.-M., Shannon, H., Baez, M., Yu, J.L., Koester, A., Schoepp, D.D., 2005. Anxiolytic-like activity of the mGLU2/3 receptor agonist LY354740 in the elevated plus maze test is disrupted in metabotropic glutamate receptor 2 and 3 knock-out mice. Psychopharmacology 179, 284e291. Lucas, S.J., Bortolotto, Z.A., Collingridge, G.L., Lodge, D., 2012. Selective activation of either mGlu2 or mGlu3 receptors can induce LTD in the amygdala. Neuropharmacology, 6e11. Malherbe, P., Kratochwil, N., Mühlemann, A., Zenner, M.T., Fischer, C., Stahl, M., Gerber, P.R., Jaeschke, G., Porter, R.H.P., 2006. Comparison of the binding pockets of two chemically unrelated allosteric antagonists of the mGlu5 receptor and identification of crucial residues involved in the inverse agonism of MPEP. J. Neurochem. 98, 601e615. Manahan-Vaughan, D., 1997. Group 1 and 2 metabotropic glutamate receptors play differential roles in hippocampal long-term depression and long-term potentiation in freely moving rats. J. Neurosci. 17, 3303e3311. Massey, P.V., Bashir, Z.I., 2007. Long-term depression: multiple forms and implications for brain function. Trends Neurosci. 30, 176e184. Mercier, M.S., Lodge, D., Fang, G., Nicolas, C.S., Collett, V.J., Jane, D.E., Collingridge, G.L., Bortolotto, Z.A., 2013. Characterisation of an mGlu8 receptorselective agonist and antagonist in the lateral and medial perforant path inputs to the dentate gyrus. Neuropharmacology 67, 294e303. Merlin, L.R., 2002. Differential roles for mGluR1 and mGluR5 in the persistent prolongation of epileptiform bursts. J. Neurophysiol. 87, 621e625. Merlin, L.R., Wong, R.K., 1997. Role of group I metabotropic glutamate receptors in the patterning of epileptiform activities in vitro. J. Neurophysiol. 78, 539e544. Michalon, A., Sidorov, M., Ballard, T.M., Ozmen, L., Spooren, W., Wettstein, J.G., Jaeschke, G., Bear, M.F., Lindemann, L., 2012. Chronic pharmacological mGlu5 inhibition corrects fragile X in adult mice. Neuron 74, 49e56. Milligan, G., 2003. Constitutive activity and inverse agonists of G protein-coupled receptors: a current perspective. Mol. Pharmacol. 64, 1271e1276. Milligan, G., Bond, R.A., Lee, M., 1995. Inverse agonism: pharmacological curiosity or potential therapeutic strategy? Trends Pharmacol. Sci. 16, 10e13. Mockett, B.G., Guevremont, D., Wutte, M., Hulme, S.R., Williams, J.M., Abraham, W.C., 2011. Calcium/calmodulin-dependent protein kinase II mediates group I metabotropic glutamate receptor-dependent protein synthesis and long-term depression in rat hippocampus. J. Neurosci. 31, 7380e7391. Moult, P.R., Corrêa, S.A.L., Collingridge, G.L., Fitzjohn, S.M., Bashir, Z.I., 2008. Coactivation of p38 mitogen-activated protein kinase and protein tyrosine phosphatase underlies metabotropic glutamate receptor-dependent long-term depression. J. Physiol. 586, 2499e2510. Moult, P.R., Gladding, C.M., Sanderson, T.M., Fitzjohn, S.M., Bashir, Z.I., Molnar, E., Collingridge, G.L., 2006. Tyrosine phosphatases regulate AMPA receptor trafficking during metabotropic glutamate receptor-mediated long-term depression. J. Neurosci. 26, 2544e2554. Mundell, S.J., Matharu, A.L., Pula, G., Roberts, P.J., Kelly, E., 2001. Agonist-induced internalization of the metabotropic glutamate receptor 1a is arrestin-and dynamin-dependent. J. Neurochem. 78, 546e551. Naie, K., Gundimi, S., Siegmund, H., Heinemann, U., Manahan-Vaughan, D., 2006. Group III metabotropic glutamate receptor-mediated, chemically induced longterm depression differentially affects cell viability in the hippocampus. Eur. J. Pharmacol. 535, 104e113. Nicoletti, F., Bockaert, J., Collingridge, G.L., Conn, P.J., Ferraguti, F., Schoepp, D.D., Wroblewski, J.T., Pin, J.P., 2011. Metabotropic glutamate receptors: from the workbench to the bedside. Neuropharmacology 60, 1017e1041. Niswender, C.M., Conn, P.J., 2010. Metabotropic glutamate receptors: physiology, pharmacology, and disease. Annu. Rev. Pharmacol. Toxicol. 50, 295e322. O’Mara, S.M., Rowan, M.J., Anwyl, R., 1995. Metabotropic glutamate receptorinduced homosynaptic long-term depression and depotentiation in the dentate gyrus of the rat hippocampus in vitro. Neuropharmacology 34, 983e989. Otani, S., Auclair, N., Desce, J.M., Roisin, M.P., Crépel, F., 1999. Dopamine receptors and groups I and II mGluRs cooperate for long-term depression induction in rat prefrontal cortex through converging postsynaptic activation of MAP kinases. J. Neurosci. 19, 9788e9802. Otani, S., Daniel, H., Takita, M., Crépel, F., 2002. Long-term depression induced by postsynaptic group II metabotropic glutamate receptors linked to phospholipase C and intracellular calcium rises in rat prefrontal cortex. J. Neurosci. 22, 3434e3444. Overstreet, L.S., Pasternak, J.F., Colley, P.A., Slater, N.T., Trommer, B.L., 1997. Metabotropic glutamate receptor mediated long-term depression in developing hippocampus. Neuropharmacology 36, 831e844. Pagano, A., Rüegg, D., Litschig, S., Stoehr, N., Stierlin, C., Heinrich, M., Floersheim, P., Prezèau, L., Carroll, F., Pin, J.P., 2000. The non-competitive antagonists 2methyl-6-(phenylethynyl) pyridine and 7-hydroxyiminocyclopropan [b] chromen-1a-carboxylic acid ethyl ester interact with overlapping binding pockets in the transmembrane region of group I metabotropic glutamate receptors. J. Biol. Chem. 275, 33750e33758. Palmer, M.J., Irving, A.J., Seabrook, G.R., Jane, D.E., Collingridge, G.L., 1997. The group I mGlu receptor agonist DHPG induces a novel form of LTD in the CA1 region of the hippocampus. Neuropharmacology 36, 1517e1532.
146
D. Lodge et al. / Neuropharmacology 74 (2013) 135e146
Patil, S.T., Zhang, L., Martenyi, F., Lowe, S.L., Jackson, K.A., Andreev, B.V., Avedisova, A.S., Bardenstein, L.M., Gurovich, I.Y., Morozova, M.A., Mosolov, S.N., Neznanov, N.G., Reznik, A.M., Smulevich, A.B., Tochilov, V.A., Johnson, B.G., Monn, J.A., Schoepp, D.D., 2007. Activation of mGlu2/3 receptors as a new approach to treat schizophrenia: a randomized Phase 2 clinical trial. Nat. Med. 13, 1102e1107. Pelkey, K.A., Lavezzari, G., Racca, C., Roche, K.W., McBain, C.J., 2005. mGluR7 is a metaplastic switch controlling bidirectional plasticity of feedforward inhibition. Neuron 46, 89e102. Piccinin, S., Thuault, S.J., Doherty, A.J., Brown, J.T., Randall, A.D., Davies, C.H., Bortolotto, Z.A., Collingridge, G.L., 2008. The induction of long-term plasticity of non-synaptic, synchronized activity by the activation of group I mGluRs. Neuropharmacology 55, 459e463. Pin, J.P., Duvoisin, R., 1995. The metabotropic glutamate receptors: structure and functions. Neuropharmacology 34, 1e26. Pöschel, B., Manahan-Vaughan, D., 2005. Group II mGluR-induced long term depression in the dentate gyrus in vivo is NMDA receptor-independent and does not require protein synthesis. Neuropharmacology 49, 1e12. Prézeau, L., Gomeza, J., Ahern, S., Mary, S., Galvez, T., Bockaert, J., Pin, J.P., 1996. Changes in the carboxyl-terminal domain of metabotropic glutamate receptor 1 by alternative splicing generate receptors with differing agonist-independent activity. Mol. Pharmacol. 49, 422e429. Qian, J., Noebels, J.L., 2006. Exocytosis of vesicular zinc reveals persistent depression of neurotransmitter release during metabotropic glutamate receptor long-term depression at the hippocampal CA3-CA1 synapse. J. Neurosci. 26, 6089e6095. Robbe, D., Alonso, G., Chaumont, S., Bockaert, J., Manzoni, O.J., 2002. Role of p/qCa2þ channels in metabotropic glutamate receptor 2/3-dependent presynaptic long-term depression at nucleus accumbens synapses. J. Neurosci. 22, 4346e 4356. Ronesi, J.A., Collins, K.A., Hays, S.A., Tsai, N.-P., Guo, W., Birnbaum, S.G., Hu, J.-H., Worley, P.F., Gibson, J.R., Huber, K.M., 2012. Disrupted Homer scaffolds mediate abnormal mGluR5 function in a mouse model of fragile X syndrome. Nat. Neurosci. 15, 431e440. Ross, J.R., Porter, B.E., Buckley, P.T., Eberwine, J.H., Robinson, M.B., 2011. mRNA for the EAAC1 subtype of glutamate transporter is present in neuronal dendrites in vitro and dramatically increases in vivo after a seizure. Neurochem. Int. 58, 366e375. Rylander, D., Recchia, A., Mela, F., Dekundy, A., Danysz, W., Cenci, M.A., 2009. Pharmacological modulation of glutamate transmission in a rat model of LDOPA-induced dyskinesia: effects on motor behavior and striatal nuclear signaling. J. Pharmacol. Exp. Ther. 330, 227e235. Schnabel, R., Kilpatrick, I.C., Collingridge, G.L., 1999. An investigation into signal transduction mechanisms involved in DHPG-induced LTD in the CA1 region of the hippocampus. Neuropharmacology 38, 1585e1596. Schoepp, D.D., Jane, D.E., Monn, J.A., 1999. Pharmacological agents acting at subtypes of metabotropic glutamate receptors. Neuropharmacology 38, 1431e1476. Seifert, R., Wenzel-Seifert, K., 2002. Constitutive activity of G-protein-coupled receptors: cause of disease and common property of wild-type receptors. Naunyn-Schmiedeberg’s Arch. Pharmacol. 366, 381e416. Shenoy, S.K., Lefkowitz, R.J., 2011. b-arrestin-mediated receptor trafficking and signal transduction. Trends Pharmacol. Sci. 32, 521e533. Sherwood, J.L., Amici, M., Dargan, S.L., Culley, G.R., Fitzjohn, S.M., Jane, D.E., Collingridge, G.L., Lodge, D., Bortolotto, Z.A., 2012. Differences in kainate
receptor involvement in hippocampal mossy fibre long-term potentiation depending on slice orientation. Neurochem. Int. 61, 482e489. Tan, Y., Hori, N., Carpenter, D.O., 2003. The mechanism of presynaptic long-term depression mediated by group I metabotropic glutamate receptors. Cell. Mol. Neurobiol. 23, 187e203. Thomas, N.K., Wright, R.A., Howson, P.A., Kingston, A.E., Schoepp, D.D., Jane, D.E., 2001. (S)-3,4-DCPG, a potent and selective mGlu8a receptor agonist, activates metabotropic glutamate receptors on primary afferent terminals in the neonatal rat spinal cord. Neuropharmacology 40, 311e318. Tiberi, M., Caron, M.G., 1994. High agonist-independent activity is a distinguishing feature of the dopamine D1B receptor subtype. J. Biol. Chem. 269, 27925e 27931. Tolchard, S., Clarke, G., Collingridge, G.L., Fitzjohn, S.M., 2000. Modulation of synaptic transmission in the rat ventral septal area by the pharmacological activation of metabotropic glutamate receptors. Eur. J. Neurosci. 12, 1843e1847. Tzounopoulos, T., Janz, R., Südhof, T.C., Nicoll, R.A., Malenka, R.C., 1998. A role for cAMP in long-term depression at hippocampal mossy fiber synapses. Neuron 21, 837e845. Upreti, C., Zhang, X., Alford, S., Stanton, P.K., 2013. Role of presynaptic metabotropic glutamate receptors in the induction of long-term synaptic plasticity of vesicular release. Neuropharmacology 66, 31e39. Volk, L.J., Daly, C.A., Huber, K.M., 2006. Differential roles for group 1 mGluR subtypes in induction and expression of chemically induced hippocampal longterm depression. J. Neurophysiol. 95, 2427e2438. Watabe, A.M., Carlisle, H.J., O’Dell, T.J., 2002. Postsynaptic induction and presynaptic expression of group 1 mGluR-dependent LTD in the hippocampal CA1 region. J. Neurophysiol. 87, 1395e1403. Watkins, J., Collingridge, G., 1994. Phenylglycine derivatives as antagonists of metabotropic glutamate receptors. Trends Pharmacol. Sci. 15, 333e342. Waung, M.W., Pfeiffer, B.E., Nosyreva, E.D., Ronesi, J.A., Huber, K.M., 2008. Rapid translation of Arc/Arg3.1 selectively mediates mGluR-dependent LTD through persistent increases in AMPAR endocytosis rate. Neuron 59, 84e97. Weisskopf, M.G., Castillo, P.E., Zalutsky, R.A., Nicoll, R.A., 1994. Mediation of hippocampal mossy fiber long-term potentiation by cyclic AMP. Science (New York, NY) 265, 1878. Wostrack, M., Dietrich, D., 2009. Involvement of Group II mGluRs in mossy fiber LTD. Synapse 63, 1060e1068. Xiao, M.-Y., Zhou, Q., Nicoll, R.A., 2001. Metabotropic glutamate receptor activation causes a rapid redistribution of AMPA receptors. Neuropharmacology 41, 664e671. Yamashita, T., Kai, T., Terakita, A., Shichida, Y., 2004. A novel constitutively active mutation in the second cytoplasmic loop of metabotropic glutamate receptor. J. Neurochem. 91, 484e492. Yanagawa, M., Yamashita, T., Shichida, Y., 2009. Activation switch in the transmembrane domain of metabotropic glutamate receptor. Mol. Pharmacol. 76, 201e207. Yokoi, M., Kobayashi, K., Manabe, T., Takahashi, T., Sakaguchi, I., Katsuura, G., Shigemoto, R., Ohishi, H., Nomura, S., Nakamura, K., Nakao, K., Katsuki, M., Nakanishi, S., 1996. Impairment of hippocampal mossy fiber LTD in mice lacking mGluR2. Science 273, 645e647. Young, S.R., Chuang, S.-C., Zhao, W., Wong, R.K.S., Bianchi, R., 2013. Persistent receptor activity underlies group I mGluR-mediated cellular plasticity in CA3. J. Neurosci. 33 (6), 2526e2540.