U.S. flag

An official website of the United States government

NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.

Madame Curie Bioscience Database [Internet]. Austin (TX): Landes Bioscience; 2000-2013.

Cover of Madame Curie Bioscience Database

Madame Curie Bioscience Database [Internet].

Show details

Neuroprotective Activity of Metabotropic Glutamate Receptor Ligands

, , , , and .

Metabotropic glutamate receptors form a family of currently eight subtypes (mGluR1 to −8), subdivided into three groups (I-III). Activation of group-II (mGluR2 and −3) or group-III metabotropic glutamate receptors (mGluR4, −6, −7 and −8) has been established to be neuroprotective in vitro and in vivo. In contrast, group-I mGluRs (mGluR1 and −5) need to be antagonized in order to evoke protection. Initially, all neuroprotective mGluR ligands were analogues of L-glutamate. Those compounds were valuable to demonstrate protection in vitro, but showed limited applicability in animal models, particularly in chronic tests, due to low blood-brain-barrier penetration. Recently, systemically active and more potent and selective ligands became available, e.g., the group-II mGluR agonists LY354740 and LY379268 or group-I antagonists like MPEP (mGluR5-selective) and BAY36–7620 (mGluR1-selective). This new generation of pharmacological agents allows a more stringent assessment of the role of individual mGluR-subtypes or groups of receptors in various nervous system disorders, including ischaemia-induced brain damage, traumatic brain injury, Huntington's- and Parkinson's-like pathology or epilepsy. Moreover, the use of genetically modified animals (e.g., knock-out mice) is starting to shed light on specific functions of mGluR-subtypes in experimental neuropathologies.

Introduction

The neurotoxicity of excitatory amino acids such as glutamate and some of its analogs, e.g., kainate and NMDA is well established in the central nervous systems.1 Glutamate, the transmitter of the vast majority of excitatory synapses in the mammalian brain, activates ionotropic and metabotropic receptors (iGluRs and mGluRs), which mediate the physiological and the toxic effects of the transmitter. Activation of iGluRs results in fast excitatory synaptic transmission via cation channel gating while mGluRs play a modulatory role by controlling membrane enzymes and second messengers. Molecular and physiological diversity of iGluRs was reviewed recently.2,3 The G protein-coupled mGluRs form a family of currently eight subtypes (mGluR1 to −8), subdivided into three groups (I-III) on the basis of their amino acid sequence identities, pharmacological profiles and signal transduction pathways.4–6 Group-I mGluRs (mGluR1 and mGluR5) are positively coupled to the phosphoinositide/Ca2+ cascade. Group-II (mGluR2 and mGluR3) and group-III (mGluR4, mGluR6, mGluR7 and mGluR8) receptors are both negatively coupled to adenylate cyclase in heterologous expression systems. The mGluR diversity is further increased by variants generated by alternative splicing at the intracellular C-terminal region. Splice variants for mGluR1 (termed mGluR1a, −b, −c, −d, −e),4,7–9 mGluR4 (a, b),10 mGluR5 (a, b),11,12 mGluR7 (a, b)13 and mGluR8 (a, b)14 have been cloned and pharmacologically characterized. Little pharmacological or physiological differences between such mGluR splice variants, most likely generated from the same receptor gene, are reported to date.

The three groups of mGluRs can be discriminated pharmacologically with the use of selective agonists. 3,5-DHPG selectively activates group-I mGluRs, while (2R, 4R)-APDC and LY-354740 are examples for group-II selective agonists; L-AP4, (R,S)-PPG, L-SOP and close analogues are selective agonists for group-III mGluRs.6,13,15–21

The individual mGluR-subtypes show a wide but distinct regional distribution throughout the mammalian nervous system. Subtypes of all three groups are highly expressed in neocortical layers, hippocampus, basal ganglia, thalamus/hypothalamus, cerebellum and spinal cord. Only mGluR6 appears to be exclusively restricted to retinal ON bipolar cells, where it couples to a cGMP-phosphodiesterase and amplifies visual transmission.5 The precise subcellular localization of each individual mGluR subtype (see Fig. 1) has been studied extensively throughout the mammalian nervous system, mostly in hippocampus, basal ganglia, cerebellum, spinal cord, and retina. Group-I mGluRs, as shown by numerous studies, are found postsynaptically with the highest receptor density at perisynaptic locations outside the postsynaptic membrane specializations.22,23 Similarly, it was shown that postsynaptic mGluR2/3 staining was also concentrated on the periphery of synaptic specializations.24 However, work by Shigemoto et al23 and Lujan et al25 demonstrated that mGluR2 immunoreactivity was located primarily in the presynaptic terminals; most of the receptors were found distant from the release sites. mGluR3 is highly expressed in glial cells,26–28 but the role of these glial receptors is as yet undetermined. The group-III receptors mGluR4, −7 and −8 are localized presynaptically in, or near, the active zones and are thought to mediate presynaptic depression of glutamatergic synaptic potentials, most likely via inhibition of voltage-gated calcium entry and regulation of glutamate release.6,23,29

Figure 1. Glial, Pre-, and Postsynaptic localization of metabotropic glutamate receptor subtypes (mGluR1–8).

Figure 1

Glial, Pre-, and Postsynaptic localization of metabotropic glutamate receptor subtypes (mGluR1–8). Coupling of mGluR1 and −5 (group-I) to postsynaptic phospholipase C (PLC) via G_q protein is indicated. Group-II mGluRs (subtypes −2 (more...)

Here we discuss how activation or inhibition of distinct groups or subtypes of metabotropic glutamate receptors result in neuroprotective effects. Synaptic localization and signal transduction of the individual receptor subtypes will be considered to explain possible mechanisms of neuroprotection; and finally we will discuss the potential use of mGluR-selective compounds in the treatment of acute and chronic degenerative disorders of the nervous system.

Chemical Structures and Receptor Profile of Neuroprotective mGluR Ligands

Group-I Preferring Antagonists

The first generation of selective antagonists described for metabotropic glutamate receptors were phenyl glycine derivatives, e.g., 4CPG, MCPG, and 4C3HPG (Fig. 2). Those compounds were widely examined in model systems of neurodegeneration/neuroprotection (see below). A very frequently used compound is MCPG which is an equally potent antagonist of mGluR1 and mGluR2 (IC50 = 20–500 μM).9,30 MCPG is less potent at mGluR5 (IC50 = 200–1000mM), and showed little antagonist activity at mGluR3, −4, −6, −7, and −8 (IC50 > 1000mM).9 4CPG is a more potent antagonist for mGluR1 (IC50 = 20–80mM) with little effects on mGluR2, −4, −5 (IC50 >500mM), and no test results available for mGluR3, −6, −7, and −8;9,30 4C3HPG shows a very similar profile but in addition to the effects of 4CPG it is an agonist at mGluR2 (EC50 = 20mM).30

Figure 2. Chemical structures and abbreviated names of most commonly used group-I mGluR antagonists.

Figure 2

Chemical structures and abbreviated names of most commonly used group-I mGluR antagonists. For all compounds shown, neuroprotective effects were reported.

The second generation of group-I mGluR antagonists showed higher potency and/or selectivity. LY367366 antagonizes mGluR1 and −5 (IC50 = 3–6 μM), interacts with group-II and -III to a lower extent (IC50 > 10mM) and shows no activity at iGluRs; the compound LY367385 shows a very similar profile except, it doesn't antagonize mGluR5 (Fig. 2).9,31 AIDA is a mGluR1-selective antagonist (IC50 = 4–200mM) with little activity at mGluR2, −4, and −5 (IC50 > 1000mM); tests at mGluR3, −6, −7 and −8 were not performed.9,30

More recently, amino acid-unrelated compounds with a high degree of selectivity for mGluR1 or mGluR5 were described. MPEP is a very potent and selective antagonist at mGluR5 with an IC50 of 32 nM, but it has no activity at all other mGluRs up to 10mM; at this concentration it's also inactive at a broad selection of iGluRs (Fig. 2).32 CPCCOEt is a very selective antagonist for mGluR1 (IC50 = 6.5mM) which shows up to 100 mM no activity at mGluR2, −4, −5, −7, and −8 (mGluR3 and −6 were not tested).9,33 Similarly, BAY36–7620 was very recently described as a highly selective mGluR1 antagonist showing higher potency than CPCCOEt (IC50 = 0.16mM; Fig. 2).34

Agonists for Metabotropic Glutamate Receptors

The first mGluR agonist used on neuronal tissue were quisqualate, ibotenate, and trans-ACPD or its active isomer 1S,3R-ACPD.9 Although important insights into neuroprotective/neurodegenerative activity of mGluR agonists were gained (e.g., activation of group-I amplifies neuronal death),35 more direct evidence for an involvement of specific groups/subtypes of mGluRs arose via the use of more selective agonists. The glutamate analogues L-CCG-I and DCG-IV (Fig. 3) were used to demonstrate group-II mGluRs as potential targets for neuroprotective drugs.36 Both compounds show very potent agonist activity at mGluR2 and mGluR3 (EC50-values = 0.1–0.9 mM), however at higher concentrations L-CCG-I is also an agonist at mGluR1, −4, −5, −6, −7, and −8 (EC50-values = 2–9 μM), while DCG-IV antagonizes all group-III mGluRs (IC50 = 22–40 mM); DCG-IV is also a potent agonist of NMDA receptors.9,37 Considerably improved compounds were published more recently. 2R,4R-APDC shows agonist EC50-values at mGluR2 and mGluR3 of 0.3–0.4 μM. The compound has no appreciable iGluR activity and doesn't activate or antagonize any group-I or -III mGluR up to 100 μM. Thus, 2R,4R-APDC is extremely specific for group-II receptors.9 The compounds LY354740 and LY379268 show much higher potency and equally good specificity for mGluR2 and mGluR3 (EC50-values = 3–20 nM), when compared to 2R,4R-APDC (Fig. 3).9,20 NAAG is an interesting molecule, because it activates only mGluR3 (EC50 = 65 μM) but not mGluR2 (EC50 > 300 mM), and it doesn't show interaction with group-I and group-III receptors.38

Figure 3. Chemical structures and abbreviated names of most commonly used group-II and -III mGluR agonists.

Figure 3

Chemical structures and abbreviated names of most commonly used group-II and -III mGluR agonists. For all compounds shown, neuroprotective effects were reported.

An involvement of group-III mGluRs in neuroprotection has primarily been demonstrated by the use of the phosphono amino acids L-AP4 and (R,S)-PPG (Fig. 3). Both compounds are selective activators of group-III mGluRs with no appreciable activity at other mGluRs or iGluRs. The EC50-values of L-AP4 at mGluR4 and −6 were reported to be 0.2–1 mM, 150–500 μM at mGluR7 and 0.06–0.9 μM at mGluR8. (R,S)-PPG is slightly less potent with EC50-values at mGluR4, −6, −7, and −8 receptors of 5.2 mM, 4.7 mM, 185 mM and 0.2 μM, respectively.5,9,18,21 The pharmacology of L-SOP at group-III mGluRs is very similar to (R,S)-PPG, but characterization of L-SOP at other receptors is not available to date.9

Physico-Chemical and Pharmacokinetic Properties of Neuroprotective mGluR Ligands

The investigation of the respective roles of each mGluR subtype in neurodegeneration and neuroprotection requires the use of ligands which are not only group- or subtype-selective but which also possess the adequate physico-chemical and pharmacokinetic properties needed for their use in the different experimental models.

General Considerations

Glutamate and aspartate are the two acidic representatives among the natural amino acids. The presence of a distal carboxylic acid enhances their polar properties in comparison to the other amino acids. These polar properties limit considerably their capacity to cross membranes by passive diffusion. Therefore, glutamate is actively transported through biological membranes by specific transporter proteins. Up to date, almost all mGluR-selective ligands are amino acid derivatives. The only non-amino acid compounds discussed in this report (see Figs. 2 and 3) are MPEP, CPCCOEt and BAY36–7620 which are all group-I antagonists.32,34,39,40 All other structures shown in Figure 2 and 3 share the same functional groups as glutamate, namely the amino acid functionality and the distal carboxylic acid (or phosphate moiety in the case of group-III ligands). Despite the structural analogy to glutamic acid the majority of these molecules fails to be actively transported through the cell membranes by the specific glutamate transporter systems and as a consequence have a poor oral bioavailability and do not cross the blood-brain-barrier (BBB). The two reported exceptions, LY354740 and its derivative LY379268, are orally bioavailable, enter the brain and display potent agonist activity at mGluR2, and −3.20,41 In addition, these two ligands have a good water solubility, which is generally the case for glutamate analogs and of advantage for the formulation and the use in in vitro neuroprotection assays. The recently published non-amino acid compounds MPEP and BAY 36–7620 also demonstrate oral bioavailability and good BBB penetration32,39,42 in contrast to the earlier published mGluR1 antagonist CPCCOEt, which neither shows bioavailability nor BBB penetration.

In Vivo Dosage and Side Effects

Most glutamate analogs such as MCPG, 4C3HPG and AIDA can only by studied in animal disease models after central administration (i.c.v.). The relatively low potency and lack of selectivity of such compounds usually doesn't allow conclusions on the side effect profile of antagonizing specific receptors. In contrast, MPEP, the mGluR5-selective antagonist at in vitro and in vivo conditions can easily be studied in animal models because it reaches the brain via oral and intravenous application (p.o. and i.v.).32,42 Furthermore, MPEP shows a large window between the doses achieving the beneficial effects in animal models for inflammatory pain (10–100 mg/kg, p.o.) or anxiety (1–10 mg/kg, p.o.) and the dose inducing the first side effects (> 100 mg/kg, p.o.) such as inhibition of spontaneous locomotor activity.32,42,43

In the case of group-II mGluRs, two orally active compounds have been investigated: LY354740 and LY379268. Both compounds are nanomolar potent agonists with similar potency at mGluR2 and mGluR3 and the investigations in vivo show that both compounds penetrate the BBB. In animal models, LY354740 showed anxiolytic like effects after oral administration (0.5–10 mg/kg, p.o.) without the sedative or motor impairment side effects of the classical benzodiazepine anxiolytics. Other group-II agonists, discussed here, are not bioavailable and do not cross the BBB, therefore have to be centrally administered.

None of the currently known group-III-selective agonists, e.g., L-AP4, L-SOP or (R,S)-PPG, show oral bioavailability or BBB penetration. Therefore, in vivo experiments necessitate a central administration of the drugs. The presence of a very polar acidic moiety such as the phosphate/phosphonate group is certainly a key element for the poor bioavailability of these molecules. In vivo (i.c.v. administration) and in vitro investigations using (R,S)-PPG in mGluR4- and mGluR7-deficient mice showed that mobilization of mGluR4 occurred to achieve neuroprotective effects (10 nmol, i.c.v.)21,44 whereas activation of mGluR7 produces anticonvulsive effects (Herman van der Putten et al, personal communication). Interestingly, the sedative and respiratory side-effects induced by central administration of higher doses of PPG (200–2000 nmol, i.c.v. in mice)21,45 are also seen in mGluR4- or mGluR7-deficient animals, clearly indicating that these side-effects are mediated by an interaction of (R,S)-PPG with target(s) distinct from mGluR4 and mGluR7 (Herman van der Putten, personal communication).

Group-I mGluRs as Targets for Neuroprotective Drugs

General and Introductory Considerations on Group-I Receptors

The role of mGluR1 and mGluR5 in neurodegeneration/neuroprotection has been debated for several years because of the contrasting effects of agonists (such as DHPG) in different in vitro models of neuronal toxicity.46 However, the evidence that subtype-selective mGluR1 or −5 antagonists are consistently neuroprotective, suggests that endogenous activation of group-I receptors is permissive to cell death. mGluR1 and −5 are both coupled to polyphosphoinositide hydrolysis, but they differ for the kinetics of IP3-induced Ca2+ mobilization. In transfected cells, activation of mGluR1 induces a single-peaked Ca2+ response, whereas activation of mGluR5 promotes oscillatory increases in cytosolic free Ca2+ due to the PKC-dependent phosphorylation of a specific threonine residue that is just distal to the 7th trans-membrane domain of mGluR5.47 Whether or not a similar difference exists under native conditions is unclear at present. Activation of both receptors regulates the activity of a variety of Ca2+ and K channels and turns on a not-yet-defined voltage-gated inward current,48,49 with the net effect of enhancing neuronal excitability. Particular attention deserves the description of a positive modulation of NMDA-gated ion currents by mGluR5, which provides the most likely substrate for the permissive action of mGluR5 activators on neuronal toxicity.50–55 Interestingly, NMDA receptor activation can also potentiate mGluR5 responses by inducing a calcineurin-dependent dephosphorylation of a PKC phosphorylation site that participates in mGlu5 desensitization.56,57 The Shank family of postsynaptic density proteins may cross-link Homer and PSD-95, thus allowing the functional coupling between mGluR5 and NMDA receptors in the same region of the dendritic spine.58 Studies performed in transfected oocytes suggest that only NMDA receptors containing NR2A or −2B subunits are positively modulated by mGluR5, whereas NMDA receptors containing the NR2C (or −2D) subunit are not. This may help explain why in mature cerebellar granule cells (that predominantly express the NR2C subunit) activation of group-I mGluRs enhances NMDA responses only when the NR2C subunit is knocked down by antisense oligonucleotides.59 A question that awaits a final answer is whether or not group-I mGluR subtypes are presynaptically localized and modulate neurotransmitter release. This issue is fundamental because the extent of excitotoxic degeneration is strictly related to glutamate and GABA release in a number of disorders, such as ischemia-induced neurodegeneration and temporal lobe epilepsy. A series of studies carried out in the laboratory of J. Sanchez-Prieto indicate that a first application of group-I mGluR agonists facilitates glutamate release in cortical or hippocampal synaptosomes,60–65 whereas a second drug application inhibits release. This functional change in the modulation of glutamate release depends on the PKC-dependent phosphorylation of the receptor (presumably mGluR5), leading to a switch in the G-protein coupling. A similar scenario is observed in cultured cortical cells, where a first exposure to DHPG amplifies NMDA toxicity, whereas a second exposure is neuroprotective.66 As a secondary release in endogenous glutamate contributes to the progression of NMDA toxicity, the functional state of ‘presynaptic’ mGluR5 might be determinant for the final effect of group-I mGluR agonists on excitotoxic neuronal death. Electron microscopy studies have shown that group-I mGluRs are preferentially localized in the peripheral portion of postsynaptic densities rather than in presynaptic terminals.22,23 However, the evidence that mGluR5 is found in the axon when co-transfected with Homer-1a raises the intriguing possibility that homer proteins regulate the dendritic or axon targeting of mGluR5.67 Homer-1a is encoded by an early gene, which is switched on by neuronal hyperactivity. It will be interesting to examine whether changes in the subcellular distribution of mGluR5 are associated with synaptic hyperactivity that occurs during epilepsy or brain ischemia.

Effect of Group-I mGluR Agonists on Excitotoxic Neuronal Death

mGluR1/5 agonists may either amplify excitotoxic neuronal death or produce neuroprotection, depending on the neuronal type and the paradigm of toxicity. Possible explanations for contrasting data include (i) the absence or presence of the NR2C subunit in NMDA receptors; (ii) the existence of an activity-dependent “switch” between facilitatory and inhibitory mGluR subtypes (see above); and (iii) a role for glial cells expressing mGluR5. These aspects have been discussed in detail in a recent review.46 It is noteworthy that excitotoxic neuronal death incorporates features of necrosis and apoptosis, and that apoptosis by trophic deprivation contributes to the overall neuronal death in brain ischemia.68 One expects that, in experimental model of brain ischemia, pharmacological activation of group-I mGluRs amplifies necrotic death by further increasing cytosolic free Ca2+, but supports the survival of neurons undergoing apoptosis by trophic deprivation through the same mechanism.69 Therefore one can easily conclude that pharmacological activation of group-I mGluR subtypes is not valuable in the experimental treatment of acute or chronic neurodegenerative disorders.

Neuroprotective Activity of mGluR1 Antagonists

A battery of competitive or non-competitive mGluR1 antagonists has been tested in in vitro and in vivo models of excitotoxic neuronal death. AIDA, LY367385, 4CPG, 4C3HPG, and CPCCOEt protect mixed murine cortical cultures against NMDA toxicity,31,70,71 and are also effective in in vivo models of excitotoxic death.31,71,72 LY367385, 4C3HPG and AIDA are also neuroprotective in the gerbil model of global ischemia and in murine cortical cultures and rat organotypic hippocampal cultures subjected to oxygen-glucose deprivation.31,73,74 mGluR1 is preferentially localized in GABAergic neurons and its activation depresses inhibitory synaptic transmission.22,75–78 Hence, it has been hypothesized that mGluR1 antagonists are neuroprotective by enhancing GABAergic transmission. Initial evidence in this line has been provided by the observation that local infusion of AIDA increases the basal output of GABA in the gerbil hippocampus.74 In murine cortical cultures, neuroprotection by LY367385 or CPCCOEt (but not by MPEP) is occluded by GABA or by SKF89976A (an inhibitor of GABA transporter) and is abolished by a cocktail of GABA-A and GABA-B receptor antagonists.71 The same cocktail of antagonists prevents the protective activity of CPCCOEt against neurodegeneration induced by NMDA in in vivo studies.71 A role for GABAergic transmission is strengthened by the evidence that LY367385 and CPCCOEt enable NMDA to enhance GABA release in the striatum, and that activation of mGluR1 inhibits GABAergic IPSCs in cortico-striatal slices.71 Taken together, these data suggest that mGluR1 antagonists are neuroprotective by removing a tonic inhibitory control exerted by mGluR1 on GABA release. The finding that mGluR1 is localized also on GABAergic nerve terminals in the striatum is consistent with this hypothesis.78 mGluR1 antagonists might be beneficial in CNS disorders characterized by an impairment of inhibitory synaptic transmission, such as Ammon's horn sclerosis.79 Interestingly, the very recently introduced mGluR1 antagonist BAY36–7620 was shown to be protective in epilepsy, stroke and trauma models upon systemic administration.34,39

Neuroprotective Effects of mGluR5 Antagonists

The development of subtype-selective mGluR5-specific antagonists has allowed to establish that endogenous activation of mGluR5 facilitates cell death in a variety of models of neurodegeneration. MPEP and its ancestors SIB-1757 and SIB-1893 (which all behave as non-competitive mGluR5 antagonists) are potent and effective in attenuating NMDA toxicity in culture, and at least MPEP can also reduce excitotoxic neuronal death in the rat striatum.80 As opposed to mGluR1 antagonists, these effects do no involve changes in GABAergic transmission71 and might be explained with the ability of mGluR5 to positively modulate NMDA-gated ion currents (see above). In cortical cultures, neuroprotection by MPEP, SIB-1757 or SIB-1893 is observed at concentrations <10 μM, and, therefore, cannot be ascribed to any non-specific interaction with NMDA receptors.32,42,80,81 In addition, the neuroprotective action of MPEP is not mimicked by its isomer iso-MPEP, which fails to antagonize mGluR5 although it shares most of the structural and physico-chemical features of MPEP.80 No data are yet published on the effect of MPEP in experimental models of hypoxic/ischemic neuronal death. The expectation is unclear, because activation of mGluR5 is known to support the survival of developing neurons,80 and apoptosis by trophic deprivation contributes to the overall neuronal death in brain ischemia (see above).

A series of studies outlines a potential use of mGluR5 antagonists in the experimental therapy of chronic neurodegenerative disorders. In cultured cortical cells, MPEP and its analogs are neuroprotective against β-amyloid toxicity at concentrations lower than those required for neuroprotection against excitotoxic death. In these studies, β-amyloid peptide is applied to the cultures in the presence of MK-801 and DNQX, thus excluding any involvement of ionotropic glutamate receptors in the neuroprotective activity of mGlu5 receptor antagonists.80 Interestingly, inhibition of mGluR1 by AIDA does not reduce but rather exacerbates ß-amyloid toxicity in cortical cells.83 Thus, at least in cultured neurons, endogenous activation of mGluR5 is required for the engagement of a death pathway in neurons exposed to ß-amyloid peptide. This encourages the study of mGluR5 antagonists in animal models of Alzheimer's disease. Finally, mGluR5 antagonists may be promising in the treatment of Parkinson's disease (PD). An optimal anti-Parkinsonian drug should combine neuroprotective properties with the ability to relief the cardinal symptoms of the disease (i.e., bradykinesia, rigidity and tremor). Recent unpublished data show that systemically administered MPEP or SIB-1893 (both at 10 mg/kg, i.p.) protect nigro-striatal dopaminergic terminals against metamphetamine toxicity and reduce the amount of reactive oxygen species produced by metamphetamine in the striatum of freely moving animals (G. Battaglia et al, manuscript in preparation). On the other hand, activation of mGluR5 induces direct excitation, and selectively potentiates NMDA currents in neurons of the subthalamic nucleus.53 As an increased activity of the subthalamic nucleus is implicated in the pathophysiology of bradykinesia, one can predict that mGluR5 antagonists can improve Parkinsonian symptoms in experimental animals and humans. It will be interesting to test mGluR5 antagonists in classical experimental models of Parkinsonism, such as the MPTP model in mice or monkeys.

Neuroprotection Mediated by Group-II mGluRs

mGluR2/3 Agonists Are Protective in Several Experimental Paradigms of Neurodegeneration

In the last decade several reports have shown that pharmacological activation of group-II mGluRs is neuroprotective in a variety of models of neuronal degeneration, including neuronal cultures, brain slices and in vivo models of excitotoxicity. The non-selective group-II mGluRs agonists, DCG-IV, 4C3HPG, L-CCG-I and 1S,3R-ACPD, protect neurons against NMDA toxicity in mixed cultures of mouse cerebral cortex.36,66,84–89 This effect is substantially reduced by the group-II mGluR antagonists, PCCG-IV or MCCG-I, suggesting that activation of mGluR2/3 is responsible for neuroprotection and is supported by the finding that the highly selective mGluR2/3 agonist, 2R,4R-APDC, is neuroprotective in the same model.90,91 DCG-IV or L-CCG-I are also neuroprotective against kainate toxicity in cultured cortical cells,36 although Gottron et al92 found that, in the same cultures, protection by DCG-IV is restricted to the small percentage of neurons that respond to kainate with an enhanced influx of Co2 . Neuroprotection by group-II mGluR agonists has also been shown in primary cultures of cerebellar granule cells,93 as well as in primary cultures of mesencephalic neurons,94 challenged with excitotoxins. Recently, the potent, highly selective and systemically active group-II mGluR agonists, LY354740, LY379268 and LY389795, have been shown to prevent excitotoxicity in both rat and mouse cortical neuronal cultures (see Kingston et al and D'Onofrio et al for support; but see also Behrens et al).89,95–97 Interestingly, the effect of these compounds is enhanced by the presence of glial cells, suggesting an involvement of glial mGluR3 in the mechanism of neuroprotection.96 Accordingly, NAAG, an endogenous selective mGluR3 agonist, protects cultured cortical neurons against NMDA toxicity.38,86

Selective and non-selective group-II mGluR agonists exert neuroprotective activity also in further in vitro models of neuronal injury, i.e., oxygen-glucose deprivation- and staurosporine-induced neuronal death, in which neurons follow the apoptotic pathway of degeneration.87,95,98,99

Group-II mGluR agonists have also been shown to be neuroprotective in in vivo models of neurodegeneration. DCG-IV, infused i.c.v., protects vulnerable neurons against local or systemic injection of kainate.100 In addition, LY379268, locally or systemically injected, protects against NMDA-induced striatal GABAergic neuronal loss in rats.89 Intracerebroventricular injection of 4C3HPG protects hippocampal CA1 neurons in a gerbil model of global ischemia (BCAO), both when administered prior to and after the induction of ischemia.101 The same degree of neuroprotection in hippocampal CA1 neurons, in the BCAO model, has been observed after systemic injection of LY354740 or LY379268.102,103 In contrast, the infarct size following middle cerebral artery occlusion (MCAO), induced by endothelin-1 in rat, is unaffected by the drugs, suggesting that group-II mGluR agonists are likely to have more utility in global than in focal cerebral ischemia.103,104 LY379268 exerts also neuroprotective activity in a neonatal rat model of hypoxia-ischemia.105

Recently, group-II mGluR agonists have been shown to possess anti-seizure activity. LY354740 is anticonvulsant in mice when systemically administered in both the ACPD-induced limbic seizure model and the pentetrazole- and picrotoxin-induced seizures.20,106,107 Moreover, group-II mGluR agonists attenuate both in vitro and in vivo models of traumatic neuronal injury, showing an additive effect with NMDA- and group-I mGluR antagonists.108 In addition, L-CCG-I and 4C3HPG inhibit excitotoxic phenomena mediated by kainate on spinal cord motor neurons, a model resembling amyotropic lateral sclerosis.109

Neuroprotection via mGluR2/3 Involves Glial-Neuronal Interactions

The mechanisms underlying group-II mGluR-mediated neuroprotection has been examined in cultured cortical cells exposed to a brief pulse with NMDA, a model of excitotoxicity in which the damage is induced by the influx of extracellular Ca2+ through the NMDA receptor channel during the pulse, and is amplified by the endogenous glutamate secondarily released after the pulse.110 Group-II mGluRs are preferentially localized in the preterminal region of the axon, far from the active zone of neurotransmitter release, and their activation inhibits the release of glutamate, but only in response to high concentrations of glutamate that spread back to the most remote region of the axon. Moreover, mGluR3 is also expressed by astrocytes throughout the brain.

It has been initially suggested that activation of mGluR2/3 is neuroprotective by reducing the release of endogenous glutamate. However, an additional mechanism might be responsible for neuroprotection since the protein synthesis inhibitor, cycloheximide, prevents the neuroprotective activity of group-II mGluR agonists in cortical cell cultures.85,95,96 Thus, it has been proposed that group-II mGluR activation triggers a specific program that requires new protein synthesis and that is likely to occur in glial cells, as the conditioned medium collected from pure cultures of astrocytes, transiently exposed to DCG-IV, L-CCG-I, 4C3HPG, LY379268 (which activate both mGluR2 and mGluR3) or to NAAG (a selective mGluR3 agonist), is highly neuroprotective when transferred to mixed cortical cultures already challenged with NMDA.85 Moreover, the neuroprotective activity of group-II mGluR agonists is almost completely lost in the absence of glial cells.95,96 This effect is likely to be mediated by the production and release from astrocytes of transforming growth factor-β (TGF-β), induced by the activation of glial mGluR3.86

In recombinant cells, mGluR2/3 are negatively coupled to adenylate cyclase through a Gi protein. However, it has been shown that in mixed cortical cultures exposed to NMDA, neither forskolin nor dibutyryl-cAMP counteract the neuroprotective activity of group-II mGluR agonists, excluding an involvement of the a subunits of the Gi protein coupled to group-II mGluRs. Therefore, to explain how activation of Gi-linked receptors leads to the production of neurotrophic factors, attention has been focused on the intracellular pathways activated by the βγ subunits released from the Gi protein (Fig. 4). A possible link is represented by the activation of mitogen-activated protein (MAP) kinase and phosphatidylinositol (PI)-3-kinase which occurs both in primary cultures and in recombinant cells (Fig. 4).89 Accordingly, blockade of these pathways leads not only to a reduction of the TGF-βproduction induced by group-II mGluR activation, but also reverses the neuroprotective activity of mGluR2/3 agonists, an effect which also occurs in in vivo models of excitotoxicity.89 MAP kinases, P90rsk and P70S6k, which regulate gene expression and protein synthesis, could represent the downstream effectors of the de novo synthesis of TGF-β.

Figure 4. Neuroprotection via activation of mGluR3.

Figure 4

Neuroprotection via activation of mGluR3. Schematic representation of the cascade of events leading to TGF-ß secretion by astrocytes. Protein kinases (e.g., TAK1) and SMAD proteins are likely to transduce TGF-ß signals from cell surface (more...)

However, the proposed mechanism of glial-neuronal interaction mediated by TGF-_ production in response to mGluR2/3 activation, observed both in cortical cultures challenged with NMDA and in rat striatum infused with NMDA, is not found in hippocampus after BCAO.111 As Gi-linked receptors, such as group-II mGluRs and A1 adenosine receptors, have been shown to induce in cultured astrocytes the release of other factors, such as nerve growth factor and S-100b protein, other than TGF-β,112 it cannot be excluded that in different brain regions other factors modulate the neuroprotective activity of group-II mGluRs.

mGluR2/3 Agonists and Potential Applicability in Acute and Chronic Neurodegeneration

Group-II mGluRs may be considered as a potential target for drugs aimed at reducing the progression of neuronal degeneration (Fig. 4). A safe neuroprotective drug with a favorable therapeutic window is particularly needed in the experimental therapy of ischemic brain damage. Group-II mGluR agonists might meet these criteria because: (i) they are neuroprotective against in vitro degeneration induced by oxygen-glucose deprivation and in in vivo models of global ischemia; (ii) they do not interfere with the normal excitatory synaptic transmission, in contrast to NMDA or AMPA receptor antagonists; (iii) they protect cultured neurons when applied after a toxic pulse with NMDA, at times at which NMDA antagonists are no longer protective, and they are even active in vivo when applied after the induction of an ischemic damage; (iiii) by locally increasing the production of neurotrophic factors such as TGF-β, they may provide a broad spectrum mechanism of protection, as TGF-β is known to exert neuroprotective activities against neuronal degeneration induced by excitotoxins, oxygen-glucose deprivation, β-amyloid peptide and the HIV capside protein gp120,86,98,99,113–124 and (iiiii) mGluR3 is localized on the vascular side of astrocytes, in proximity of endothelial cells, so they can be easily reached by drugs present in the blood stream and able to cross the blood-brain-barrier.

Recently, group-II mGluRs have been proposed as targets for the therapy of PD. Overactive glutamatergic afferents from the subthalamic nucleus could cause both an excitotoxic loss of dopaminergic neurons in the substantia nigra, and hyperactivation of GABAergic neurons in the globus pallidus, which leads to a reduction of motor activity.125 Drugs, which can reduce glutamate release by acting at presynaptic group-II mGluRs might at one time delay the degeneration of substantia nigra neurons and improve motor activity.126,127

Furthermore, it's interesting to add that group-II mGluR agonists could interfere with the pathophysiology of neuropsychiatric disorders such as anxiety and schizophrenia, and they may also exert anti-addictive effects.20,106,128–134

Role of Group-III mGluRs in Neuroprotection

Effects of Group-III mGluR Ligands in Experimental Paradigms of Neurodegeneration

There are numerous reports that describe activation of group-III mGluRs to be neuroprotective in vitro, and more recently, several in vivo observations support these findings. The agonists L-AP4, L-SOP and (RS)-PPG promote survival of rat cerebellar granule cells and protect cultured cortical and cerebellar neurons against toxic insults, such as prolonged β-amyloid peptide exposure, transient iGluR activation or mechanical damage.21,44,98,99,136–138 In rat hippocampal slices exposed to a severe hypoxic/hypoglycaemic insult, (RS)-PPG improves the recovery of population spike amplitude in CA1, a parameter for functional synaptic transmission and neuronal viability. Such acutely isolated hippocampal slices are only viable for a few hours, and electrophysiogical recordings can only be performed in a limited time window after the damaging insult when most probably excitotoxicity is still the predominant component in the pathophysiology of hypoxic/hypoglycaemic neuronal damage.45,139

In a recent study we found (RS)-PPG to be neuroprotective against striatal lesions induced by local infusion of NMDA or quinolinic acid into the rat caudate nucleus and, to our knowledge, this provided the first in vivo evidence that activation of group-III mGluRs is neuroprotective in animal models.21 The use of such in vivo excitotoxic injury models, to produce neuronal depletion, reactive gliosis and alterations of neurotransmitter levels, has been highly valuable for examining pathological patterns reminiscent of Huntington's disease (HD). Even if the primary cause of HD is unrelated, excitotoxic injury mediated by iGluR activation may play a role in progressive neuronal depletion.140

Extending the studies to further in vivo models of neurodegeneration, we and others confirmed neuroprotective effects of (RS)-PPG in rat and also mouse models of excitotoxicity (in vivo and in vitro), but, neither in focal cerebral ischaemia in mice, nor in global cerebral ischaemia in gerbils, nor in global cerebral ischaemia in rats (RS)-PPG had any significant influence on the extent of neuronal damage.44,45

Neuroprotection via Group-III mGluRs Is Mediated by mGluR4 and/or mGluR7 and Involves a Presynaptic Regulation of Transmitter Release

In contrast to group-II mGluRs that require glial components for neuroprotection, Group-III agonists are equally protective in mixed cortical cultures (containing astrocytes and neurons, see Fig. 5) and in pure cultures of neurons. Therefore, this neuroprotection is not expected to be mediated by glial factors, but rather involves one or more receptor subtype(s) expressed on neuronal structures. Recently, we have searched for the identity of the neuroprotective group-III mGluR subtype using mixed cortical cultures. This model is particularly suitable for the study because cortical neurons of mixed cultures express all currently known group-III mGluR subtypes.137 A possible role in neuroprotection for one or more group-III mGluR subtype(s) was initially suggested by the use of racemic (RS)-PPG in in vitro neuroprotection paradigms.21 More recently, the stereoisomers of (RS)-PPG were separated and it was found that all protective activity is harbored in the (+)-isomer.44,141 (+)-PPG is neuroprotective against NMDA toxicity with an EC50 value of about 5 μM. This value coincides with that found for the activation of recombinant mGluR4, but differs by at least 25-fold from the potency of PPG at mGluR7 and −8.21 Furthermore, there is strong evidence for a critical role of mGluR4 in mediating neuroprotection via the use of cortical cultures prepared from mGluR4 subtype-deficient mice (−/−, Fig. 5B), where all group-III agonists [i.e., L-AP4, (RS)-PPG and L-SOP] failed to protect against NMDA toxicity. This was in contrast to the protective effects of group-III agonists in wild-type (+/+) and heterozygous neurons and did not reflect a general refractoriness of −/− knockout neurons to mechanisms of protection, because the group-I/II mGluR ligands 4C3HPG, CPCCOEt and MPEP, retained their protective activity in those mGluR4 −/− cultures.32,33,44,142

Figure 5. Phase-contrast photomicrographs of cultured cortical cells from CD1 (wild-type, wt, A) and mGluR4-deficient (knock-out, ko, B) mice.

Figure 5

Phase-contrast photomicrographs of cultured cortical cells from CD1 (wild-type, wt, A) and mGluR4-deficient (knock-out, ko, B) mice. Dissociated cortical neurons prepared from fetal mice at 14–16 days of gestation after 13 days of in vitro culture (more...)

We extended the study to the in vivo model of excitotoxic degeneration by unilaterally injecting NMDA ± (RS)-PPG into the caudate nucleus of wild-type or mGluR4 −/− mice. This brain region has been selected because it receives an extensive glutamatergic innervation from the cerebral cortex. Low doses of (RS)-PPG (10 nmol), which should preferentially activate mGluR4 over mGluR7 receptors, were neuroprotective in +/+ mice, but were totally inactive in mGluR4 −/− mice. In contrast, the −/− mice were partially protected with higher doses of (RS)-PPG (100 nmol),44 which are expected to recruit mGluR7 (unpublished observations by Herman van der Putten, based on experiments with mGluR7 −/− mice). The prominent role of mGluR4 and possibly mGluR7 in mediating group-III agonist-induced neuroprotection is also consistent with the high expression levels and broad distribution of mGluR4 and mGluR7 in many regions of mammalian brain, including basal ganglia, cortical areas and hippocampus.23,143–145

Inhibition of glutamate release by presynaptic mGluR4 and mGluR7 may represent a common mechanism of neuroprotection in vitro and in vivo. Accordingly, an enhanced release of endogenous glutamate has been shown to facilitate the progression of NMDA toxicity in cortical cultures,110 and we have found that cortical cultures from mGluR4 −/− mice were more vulnerable to low concentrations of NMDA, and showed higher extracellular glutamate levels, as compared to wild-type +/+ cultures. Moreover, in vivo microdialysis studies showed that intrastriatal infusion of NMDA increased extracellular glutamate levels to a greater extent in mGluR4 −/− than in +/+ mice, supporting the hypothesis that the mGluR4 subtype is necessary for the maintenance of the homeostasis of extracellular glutamate levels.44 In addition to the regulation of presynaptic glutamate release, an inhibition of NMDA receptors by postsynaptic group-III mGluRs (hypothetical) via a protein phosphorylation cascade may also be involved in their neuroprotective effects.146 Furthermore, group-III mGluRs also exert protection in primary hippocampal neurons by modulation of the free radical nitric oxide and the cascade of programmed cell death.147 Thus, activation of group-III mGluRs could open several novel strategies to interfere with the progressive course of neurodegenerative disorders.

mGluR4/7 Activators May Become Applicable in Basal Ganglia Disorders and Other Chronic Degenerative Diseases

The broadest analysis of group-III mGluR activation in animal models of neuronal damage was recently preformed by Henrich-Noack et al.45 The results, as summarized above, support the notion that group-III mGluR agonists are a quite valuable class of drugs against pathologies closely related to excitotoxic cell damage, but are most probably not effective enough when damaged brain tissue has progressed into a multifactorial pathology as it appears after an ischaemic challenge, e.g., in human stroke. This view is supported by several clinical stroke trials that have evaluated the potential of glutamate transmission-modifying drugs, and to date, the results of these attempts have been disappointing.148,149

In amyotropic lateral sclerosis (ALS) and Huntington's disease (HD), circumstantial evidence suggests that excitotoxicity may contribute to the pathogenic process. In fact, it was recently shown that an anti-glutamate drug, riluzole, provides some therapeutic benefit in the treatment of ALS and clinical trials aiming at neuroprotection in HD are in progress.148 HD and Parkinson's disease (PD) are examples of neurodegenerative disorders where mitochondrial dysfunction may sensitize populations of basal ganglia neurons to excitotoxicity from synaptic glutamate, and recent evidence strongly suggests that the group-III receptors mGluR4 and mGluR7 modulate basal ganglia function at multiple sites.143,144 Activation of mGluR4 and/or mGluR7 may increase viability of selective cell populations, e.g., GABAergic-projecting neurons in the striatum which are selectively lost in HD. Alternatively, modulation of group-III mGluR activity may balance basal ganglia circuits under pathological conditions, for instance in PD where subthalamic nuclei are overactive.144 Thus, reduction of glutamatergic transmission by group-III mGluR activation may well be a novel and viable mechanism for experimental therapy of ALS, HD, and PD.

Evidence for an involvement of group-III mGluRs, or glutamatergic transmission in general, in Alzheimer's disease (AD) is more indirect and speculative than for ALS, HD, and PD. Experimental evidence, however, shows that various group-III agonists are highly protective against programmed cell death in cultured cortical cells induced by prolonged β-amyloid peptide exposure.80,98 On the other hand, no in vivo data from transgenic models of AD have been published to date that would support a role of group-III receptors in AD.

Finally, it's very interesting to mention the effectiveness of group-III agonists in various experimental animal models of epilepsy.21,150–152

Conclusions and Outlook

The early in vitro neuroprotection studies with the first mGluR-selective ligands such as MCPG, 4C3HPG, DCG-IV, and L-AP4 established a major principle: activation of group-II or group-III mGluRs is neuroprotective while antagonist activity at group-I mGluRs generally results in protection. Subsequent in vivo work was supportive for this principle and laid the ground to define possible target diseases for mGluR compounds. Based on our current knowledge, human stroke as target for mGluR ligands has to be considered unlikely due to no, or weak, activity of the so far tested compounds in focal ischemia models in rodents. On the other hand, a role for mGluRs in chronic neurodegenerative disorders, such as ALS, HD, PD or AD, receives increasingly more evidence. Clearly, the gold-standard animal models for those diseases are generally performed with transgenic mice expressing relevant disease genes, e.g., expanded trinucleotide repeat-containing huntingtin locus in the case of HD. Up to date, no studies with the improved and systemically active pharmacological tools such as LY354740, MPEP, or BAY36–7620 were reported in those models. Once available, those studies are likely to create excitement towards clinical neuroprotection trials with mGluR-selective compounds.

References

1.
Lipton SA, Rosenberg PA. Excitatory amino acids as a final common pathway for neurologic disorders. N Engl J Med. 1994;330:613–622. [PubMed: 7905600]
2.
Hollmann M, Heinemann S. Cloned glutamate receptors. Annu Rev Neurosci. 1994;17:31–108. [PubMed: 8210177]
3.
Dingledine R, Borges K, Bowie D. et al. The glutamate receptor ion channels. Pharmacol Rev. 1999;51:7–61. [PubMed: 10049997]
4.
Tanabe Y, Masu M, Ishii T. et al. A family of metabotropic glutamate receptors. Neuron. 1992;8:169–179. [PubMed: 1309649]
5.
Nakanishi S. Metabotropic glutamate receptors: Synaptic transmission, modulation, and plasticity. Neuron. 1994;13:1031–1037. [PubMed: 7946343]
6.
Conn PJ, Pin J -P. Pharmacology and functions of metabotropic glutamate receptors. Annu Rev Pharmacol Toxicol. 1997;37:205–237. [PubMed: 9131252]
7.
Laurie DJ, Boddeke HW, Hiltscher R. et al. HmGlu1d, a novel splice variant of the human type I metabotropic glutamate receptor. Eur J Pharmacol. 1996;296:R1–R3. [PubMed: 8838462]
8.
Flor PJ, Gomeza J, Tones MA. et al. The C-terminal domain of the mGluR1 metabotropic glutamate receptor affects sensitivity to agonists. J Neurochem. 1996;67:58–63. [PubMed: 8667026]
9.
Schoepp DD, Jane DE, Monn AJ. Pharmacological agents acting at subtypes of metabotropic glutamate receptors. Neuropharmacology. 1999;38:1431–1476. [PubMed: 10530808]
10.
Iversen L, Mulvihill E, Haldeman B. et al. Changes in metabotropic glutamate receptor mRNA levels following global ischemia: increase of a putative presynaptic subtype (mGluR4) in highly vulnerable rat brain areas. J Neurochem. 1994;63:626–633. [PubMed: 8035186]
11.
Minakami R, Katsuki F, Yamamoto T. et al. Molecular cloning and the functional expression of two isoforms of human metabotropic glutamate receptor subtype 5. Biochem Biophys Res Commun. 1994;199:1136–1143. [PubMed: 7908515]
12.
Joly C, Gomeza J, Brabet I. et al. Molecular, functional, and pharmacological characterization of the metabotropic glutamate receptor type 5 splice variants: Comparison with mGluR1. J Neurosci. 1995;15:3970–81. [PMC free article: PMC6578186] [PubMed: 7751958]
13.
Flor PJ, van der Putten H, Rüegg D. et al. A novel splice variant of a metabotropic glutamate receptor, human mGluR7b. Neuropharmacology. 1997;36:153–159. [PubMed: 9144652]
14.
Corti C, Rimland JM, Corsi M. et al. Isolation and analysis of splice variants of the rat metabotropic glutamate receptors, mGluR7 and mGluR8. Soc Neurosci Abstr. 1997;23:939.
15.
Okamoto N, Hori S, Akazawa C. et al. Molecular characterization of a new metabotropic glutamate receptor mGluR7 coupled to inhibitory cyclic AMP signal transduction. J Biol Chem. 1994;269:1231–1236. [PubMed: 8288585]
16.
Duvoisin RM, Zhang C, Ramonell K. A novel metabotropic glutamate receptor expressed in the retina and olfactory bulb. J Neurosci. 1995;15:3075–3083. [PMC free article: PMC6577763] [PubMed: 7722646]
17.
Flor PJ, Lindauer K, Püttner I. et al. Molecular cloning, functional expression and pharmacological characterization of the human metabotropic glutamate receptor type 2. Eur J Neurosci. 1995;7:622–629. [PubMed: 7620613]
18.
Flor PJ, Lukic S, Rüegg D. et al. Molecular cloning, functional expression and pharmacological characterization of the human metabotropic glutamate receptor type 4. Neuropharmacology. 1995;34:149–155. [PubMed: 7617140]
19.
Johansen PA, Chase LA. et al. Type 4a metabotropic glutamate receptor: identification of new potent agonists and differentiation from the L-(+)-2-amino-4-phosphonobutanoic acid-sensitive receptor in the lateral perforant pathway in rats. Mol Pharmacol. 1995;48:140–149. [PubMed: 7623768]
20.
Monn JA, Valli MJ, Massey SM. et al. Design, synthesis, and pharmacological characterization of (+)-2-aminobicyclo[3.1.0]hexane-2,6-dicarboxylic acid (LY354740): A potent, selective, and orally active group 2 metabotropic glutamate receptor agonist possessing anticonvulsant and anxiolytic properties. J Med Chem. 1997;40:528–537. [PubMed: 9046344]
21.
Gasparini F, Bruno V, Battaglia G. et al. (R,S)-4-Phosphonophenylglycine, a potent and selective group III metabotropic glutamate receptor agonist, is anticonvulsive and neuroprotective in vivo. J Pharmacol Exp Ther. 1999;289:1678–1687. [PubMed: 10336568]
22.
Baude A, Nusser Z, Roberts JD. et al. The metabotropic glutamate receptor (mGluR1 alpha) is concentrated at perisynaptic membrane of neuronal subpopulations as detected by Immunogold reaction. Neuron. 1993;11:771–787. [PubMed: 8104433]
23.
Shigemoto R, Kinoshita A, Wada E. et al. Differential presynaptic localization of metabotropic glutamate receptor subtypes in the rat hippocampus. J Neurosci. 1997;17:7503–7522. [PMC free article: PMC6573434] [PubMed: 9295396]
24.
Petralia RS, Wang YX, Niedzielski AS. et al. The metabotropic glutamate receptors, mGluR2 and mGluR3, show unique postsynaptic, presynaptic and glial localizations. Neuroscience. 1996;71:949–976. [PubMed: 8684625]
25.
Lujan R, Roberts JD, Shigemoto R. et al. Differential plasma membrane distribution of metabotropic glutamate receptors mGluR1 alpha, mGluR2 and mGluR5, relative to neurotransmitter release sites. J Chem Neuroanat. 1997;13:219–241. [PubMed: 9412905]
26.
Ohishi H, Shigemoto R, Nakanishi S. et al. Distribution of the mRNA for a metabotropic glutamate receptor (mGluR3) in the rat brain: An in situ hybridization study. J Comp Neurol. 1993;335:252–266. [PubMed: 8227517]
27.
Ohishi H, Ogawa-Meguro R, Shigemoto R. et al. Immunohistochemical localization of metabotropic glutamate receptors, mGluR2 and mGluR3, in rat cerebellar cortex. Neuron. 1994;13:55–66. [PubMed: 8043281]
28.
Mineff E, Valtschanoff J. Metabotropic glutamate receptors 2 and 3 expressed by astrocytes in rat ventrobasal thalamus. Neurosci Lett. 1999;270:95–98. [PubMed: 10462106]
29.
Trombley PQ, Westbrook GL. L-AP4 inhibits calcium currents and synaptic transmission via a G-protein-coupled glutamate receptor. J Neurosci. 1992;12:2043–2050. [PMC free article: PMC6575922] [PubMed: 1318954]
30.
Bräuner-Osborne H, Egebjerg J, Nielsen EO. et al. Ligands for glutamate receptors: Design and therapeutic prospects. J Med Chem. 2000;43:2609–2645. [PubMed: 10893301]
31.
Bruno V, Battaglia G, Kingston A. et al. Neuroprotective activity of the potent and selective mGlu1a metabotropic glutamate receptor antagonist, (+)-2-methyl−4-carboxyphenylglycine (LY367385): Comparison with LY367366, a broader spectrum antagonist with equal affinity for mGlu1a and mGlu5 receptors. Neuropharmacology. 1999;38:199–207. [PubMed: 10218860]
32.
Gasparini F, Lingenhöhl K, Stoehr N. et al. 2-Methyl−6-(phenylethynyl)-pyridine (MPEP), a potent, selective and systemically active mGlu5 receptor antagonist. Neuropharmacology. 1999;38:1493–1503. [PubMed: 10530811]
33.
Litschig S, Gasparini F, Rueegg D. et al. CPCCOEt, a noncompetitive metabotropic glutamate receptor 1 antagonist, inhibits receptor signaling without affecting glutamate binding. Mol Pharmacol. 1999;55:453–461. [PubMed: 10051528]
34.
Carroll FY, Stolle A, Beart PM. et al. BAY36–7620: A potent non-competitive mGlu1 receptor antagonist with inverse agonist activity. Mol Pharmacol. 2001;59:965–973. [PMC free article: PMC2573968] [PubMed: 11306677]
35.
Bruno V, Copani A, Knöpfel T. et al. Activation of metabotropic glutamate receptors coupled to inositol phospholipid hydrolysis amplifies NMDA-induced neuronal degeneration in cultured cortical cells. Neuropharmacology. 1995;34:1089–1098. [PubMed: 8532158]
36.
Bruno V, Battaglia G, Copani A. et al. Activation of class II or class III metabotropic glutamate receptors protects cultured cortical neurons against excitotoxic degeneration. Eur J Neurosci. 1995;7:1906–1913. [PubMed: 8528465]
37.
Brabet I, Parmentier M -L, De Colle C. et al. Comparative effect of L-CCG-I, DCG-IV and γ-carboxyl-L-glutamate on all cloned metabotropic glutamate receptor subtypes. Neuropharmacology. 1998;37:1043–1051. [PubMed: 9833633]
38.
Wroblewska B, Wroblewski JT, Pschenichkin S. et al. N-Acetylaspartylglutamate selectively activates mGluR3 receptors in transfected cells. J Neurochem. 1997;69:174–181. [PubMed: 9202308]
39.
Müller T, De Vry J, Schreiber R. et al. BAY 36–7620, a selective metabotropic glutamate receptor-1 antagonist with anticonvulsant and neuroprotective properties. Soc Neurosci Abs. 2000;26:16–50.
40.
Annoura H, Fukunaga A, Uesugi M. et al. A novel class of antagonists for metabotropic glutamate receptors, 7-(hydroxyimino)cyclopropa[b]chromen-1a-carboxylates. Bioorg Med Chem Lett. 1996;6:7763–7766.
41.
Monn JA, Valli MJ, Massey SM. et al. Synthesis, pharmacological characterization, and molecular modeling of heterobicyclic amino acids related to (+)-2-aminobicyclo[3.1.0]hexane-2,6-dicarboxylic acid (LY354740): Identification of two new potent, selective, and systemically active agonists for group II metabotropic glutamate receptors. J Med Chem. 1999;42:1027–1040. [PubMed: 10090786]
42.
Spooren W P J M, Gasparini F, Salt TE. et al. Novel allosteric antagonists shed light on mglu5 receptors and CNS disorders. Trends Pharmacol Sci. 2001;22:331–337. [PubMed: 11431019]
43.
Walker K, Bowes M, Panesar M. et al. Metabotropic glutamate receptor subtype 5 (mGlu5) and nociceptive function. I. Selective blockade of mGlu5 receptors in models of acute, persistent and chronic pain. Neuropharmacology. 2001;40:1–9. [PubMed: 11077065]
44.
Bruno V, Battaglia G, Ksiazek I. et al. Selective activation of mGlu4 metabotropic glutamate receptors is protective against excitotoxic neuronal death. J Neurosci. 2000;20:6413–6420. [PMC free article: PMC6772963] [PubMed: 10964947]
45.
Henrich-Noack P, Flor PJ, Sabelhaus CF. et al. Distinct influence of the group III metabotropic glutamate receptor agonist (R,S)-4-phosphonophenylglycine [(R,S)-PPG] on different forms of neuronal damage. Neuropharmacology. 2000;39:911–917. [PubMed: 10699457]
46.
Nicoletti F, Bruno V, Catania MV. et al. Group-I metabotropic glutamate receptors: Hypotheses to explain their dual role in neurotoxicity and neuroprotection. Neuropharmacology. 1999;38:1477–1484. [PubMed: 10530809]
47.
Kawabata S, Tsutsumi R, Kohara A. et al. Control of calcium oscillations by phosphorylation of metabotropic glutamate receptors. Nature. 1996;383:89–92. [PubMed: 8779726]
48.
Pin JP, Duvoisin R. The metabotropic glutamate receptors: structure and functions. Neuropharmacology. 1995;34:1–26. [PubMed: 7623957]
49.
Chuang SC, Bianchi R, Wong RK. Group-I mGluR activation turns on a voltage-gated inward current in hippocampal pyramidal cells. J Neurophysiol. 2000;83:2844–2853. [PubMed: 10805682]
50.
Pisani A, Calabresi P, Centonze D. et al. Enhancement of NMDA responses by group-I metabotropic glutamate receptor activation in striatal neurones. Br J Pharmacol. 1997;120:1007–1014. [PMC free article: PMC1564563] [PubMed: 9134210]
51.
Jia Z, Lu Y, Henderson J. et al. Selective abolition of the NMDA component of long-term potentiation in mice lacking mGluR5. Learn Mem. 1998;5:331–343. [PMC free article: PMC311253] [PubMed: 10454358]
52.
Ugolini A, Corsi M, Bordi F. Potentiation of NMDA and AMPA responses by the specific mGluR5 agonist CHPG in spinal cord motoneurons. Neuropharmacology. 1999;38:1569–1576. [PubMed: 10530818]
53.
Awad H, Hubert GW, Smith Y. et al. Activation of metabotropic glutamate receptor 5 has direct excitatory effects and potentiates NMDA receptor currents in neurons of the subthalamic nucleus. J Neurosci. 2000;20:7871–7879. [PMC free article: PMC6772731] [PubMed: 11050106]
54.
Doherty AJ, Palmer MJ, Bortolotto ZA. et al. A novel, competitive mGlu(5) receptor antagonist (LY344545) blocks DHPG-induced potentiation of NMDA responses but not the induction of LTP in rat hippocampal slices. Br J Pharmacol. 2000;131:239–244. [PMC free article: PMC1572327] [PubMed: 10991916]
55.
Salt TE, Binns KE. Contributions of mGlu1 and mGlu5 receptors to interactions with N-methyl-D-aspartate receptor-mediated responses and nociceptive sensory responses of rat thalamic neurons. Neuroscience. 2000;100:375–380. [PubMed: 11008175]
56.
Alagarsamy S, Marino MJ, Rouse ST. et al. Activation of NMDA receptors reverses desensitization of mGluR5 in native and recombinant systems. Nat Neurosci. 1999;2:234–240. [PubMed: 10195215]
57.
De Blasi A, Conn PJ, Pin JP. et al. Molecular determinant of metabotropic glutamate receptor signaling. Trends Pharmacol Sci. 2001;22:114–120. [PubMed: 11239574]
58.
Tu JC, Xiao B, Naisbitt S. et al. Coupling of mGluR/Homer and PSD-95 coplexes by the Shank family of postsynaptic density proteins. Neuron. 1999;23:583–592. [PubMed: 10433269]
59.
Pizzi M, Boroni F, Moraitis K. et al. Reversal of glutamate excitotoxicity by activation of PKC-associated metabotropic glutamate receptors in cerebellar granule cells relies on NR2C subunit expression. Eur J Neurosci. 1999;11:1–8. [PubMed: 10383638]
60.
Herrero I, Miras-Portugal MT, Sanchez-Prieto J. Positive feed-back of glutamate exocytosis by metabotropic presynaptic receptor stimulation. Nature. 1992;360:163–166. [PubMed: 1359425]
61.
Herrero I, Miras-Potugal MT, Sanchez-Prieto J. Rapid desensitization of the presynaptic metabotropic receptor that facilitates glutamate exocytosis. Eur J Neurosci. 1994;6:115–120. [PubMed: 7907519]
62.
Herrero I, Miras-Portugal MT, Sanchez-Prieto J. Functional switch from facilitation to inhibition in the presynapticcontrol of glutamate release by metabotropic glutamate receptors. J Biol Chem. 1998;273:1951–1958. [PubMed: 9442030]
63.
Rodriguez-Moreno A, Sistiaga A, Lerma J. et al. Switch from facilitation to inhibition of excitatory synaptic transmission by the desensitization of group-I mGluRs in the rat hippocampus. Neuron. 1998;21:1477–1486. [PubMed: 9883739]
64.
Sistiaga A, Herrero I, Conquet F. et al. The metabotropic glutamate receptor 1 is not involved in the facilitation of glutamate release in cerebrocortical nerve terminals. Neuropharmacology. 1998;3:1485–1492. [PubMed: 9886671]
65.
Sistiaga A, Sanchez-Prieto J. Protein phosphatase 1 and 2A inhibitors prolong the switch in the control of glutamate release by group I metabotropic glutamate receptors: characterization of the inhibitory pathway. J Neurochem. 2000;75:1566–1574. [PubMed: 10987837]
66.
Bruno V, Battaglia G, Copani A. et al. An activity-dependent switch from facilitation to inhibition in the control of excitotoxicity by group-I metabotropic glutamate receptors. Eur J Neurosci. 2001;13:1469–1478. [PubMed: 11328342]
67.
Ango F, Pin JP, Tu JC. et al. Dendritic and axonal targeting of type 5 metabotropic glutamate receptor is regulated by homer 1 protein and neuonal excitation. J Neurosci. 2000;20:8710–8716. [PMC free article: PMC6773061] [PubMed: 11102477]
68.
Lee JM, Zipfel GJ, Choi DW. The changing landscape of ischaemic brain injury mechanisms. Nature. 1999;399 (Suppl):A7–A14. [PubMed: 10392575]
69.
Allen JW, Knoblach SM, Faden AI. Activation of group I metabotropic glutamate receptors reduces neuronal apoptosis but increases necrotic cell death in vitro. Cell Death Differ. 2000;7:470–476. [PubMed: 10800080]
70.
Strasser U, Lobner D, Berhens μ M. et al. Antagonists of group I mGluRs attenuate exicitotoxic neuronal death in cortical cultures. Eur J Neurosci. 1998;10:2848–2855. [PubMed: 9758154]
71.
Battaglia G, Bruno V, Pisani A. et al. Selective blockade of type-I metabotropic glutamate receptors induces neuroprotection by enhancing GABAergic transmission Mol Cell Neurosci 2001; in press. [PubMed: 11414795]
72.
Orlando LR, Alsdorf SA, Penney J B Jr. et al. The role of group I and group II metabotropic glutamate receptors in modulation of striatal NMDA and quinolinic acid toxicity. Exp Neurol. 2001;167:196–204. [PubMed: 11161608]
73.
Pellegrini-Giampietro DE, Cozzi A, Peruginelli F. et al. 1-Aminoindan-1,5-dicarboxylic acid and (S)-(+)-2-(3'-carboxyicyclo(1.1.1.)pentyl)-glycine, two mGlu1 receptor-preferring antagonists, reduce neuronal death in in vitro and in vivo models of cerebral ischemia. Eur J Neurosci. 1999;11:3637–3647. [PubMed: 10564371]
74.
Pellegrini-Giampietro DE, Peruginelli F, Meli E. et al. Protection with metabotropic glutamate 1 receptor antagonists in models of ischemic neuronal death: time-course and mechanisms. Neuropharmacology. 1999;38:1607–1619. [PubMed: 10530822]
75.
Gereau IV RW, Conn PJ. Multiple presynaptic metabotropic glutamate receptors modulate excitatory and inhibitory synaptic transmission in hippocampal area CA1. J Neurosci. 1995;15:6879–6889. [PMC free article: PMC6578030] [PubMed: 7472445]
76.
Morishita W, Kirov SA, Alger BE. Evidence for metabotropic glutamate receptor activation in the induction of depolarization-induced suppression of inhibition in hippocampal CA1. J Neurosci. 1998;18:4870–4882. [PMC free article: PMC6792551] [PubMed: 9634553]
77.
Bushell TJ, Lee CC, Shigemoto R. et al. Modulation of synaptic transmission and differential localisation of mGluRs in cultured hippocampal autapses. Neuropharmacology. 1999;38:1553–1567. [PubMed: 10530817]
78.
Paquet M, Smith Y. Subsynaptic localization of group I metabotropic glutamate receptors at glutamatergic synapses in monkey striatum. Soc Neurosci Abs. 2000;26:740.13.
79.
McNamara JO. Emerging insights into the genesis of epilepsy. Nature. 1999;399 (Suppl.):A15–A22. [PubMed: 10392576]
80.
Bruno V, Ksiazek I, Battaglia G. et al. Selective blockade of metabotropic glutamate receptor subtype 5 is neuroprotective. Neuropharmacology. 2000;39:2223–2230. [PubMed: 10974306]
81.
O'Leary DM, Movsesyan V, Vicini S. et al. Selective mGluR5 antagonists MPEP and SIB-1893 decrease NMDA or glutamate-mediated neuronal toxicity through actions that reflect NMDA receptor antagonism. Br J Pharmacol. 2000;131:1429–1437. [PMC free article: PMC1572472] [PubMed: 11090117]
82.
Copani A, Casabona G, Bruno V. et al. The metabotropic glutamate receptor mGlu5 controls the onset of developmental apoptosis in cultured cerebellar neurons. Eur J Neurosci. 1998;10:2173–2184. [PubMed: 9753103]
83.
Allen JW, Eldadah BA, Faden AI. Beta-amyloid-induced apoptosis of cerebellar granule cells and cortical neurons: Exacerbation by selective inhibition of group-I metabotropic glutamate receptors. Neuropharmacology. 1999;38:1243–1252. [PubMed: 10462136]
84.
Bruno V, Copani A, Battaglia G. et al. Protective effect of the metabotropic glutamate receptor agonist, DCG-IV, against excitotoxic neuronal death. Eur J Pharmacol. 1994;256:109–112. [PubMed: 7517889]
85.
Bruno V, Sureda FX, Storto M. et al. The neuroprotective activity of group-II metabotropic glutamate receptors requires new protein synthesis and involves a glial-neuronal signaling. J Neurosci. 1997;17:1891–1897. [PMC free article: PMC6793767] [PubMed: 9045718]
86.
Bruno V, Battaglia G, Casabona G. et al. Neuroprotection by glial metabotropic glutamate receptors is mediated by transforming growth factor-beta. J Neurosci. 1998;18:9594–9600. [PMC free article: PMC6793276] [PubMed: 9822720]
87.
Buisson A, Choi DW. The inhibitory mGluR agonist, S-4-carboxy-3-hydroxy-phenylglycine selectively attenuates NMDA neurotoxicity and oxygen-glucose deprivation-induced neuronal death. Neuropharmacology. 1995;34:1081–1087. [PubMed: 8532157]
88.
Buisson A, Nicole O, Docagne F. et al. Up-regulation of a serine protease inhibitor in astrocytes mediates the neuroprotective activity of transforming growth factor-β1. FASEB J. 1998;12:1683–1691. [PubMed: 9837858]
89.
D'Onofrio M, Cuomo L, Battaglia G. et al. Neuroprotection mediated by group-II metabotropic glutamate receptors involves the activation of the MAP kinase pathway and requires the induction of TGF-β1. J Neurochem. 2001;78:435–445. [PubMed: 11483646]
90.
Thomsen C, Bruno V, Nicoletti F. et al. (2S,1'S,2'S,3'R)-2-(2'-Carboxy-3'-phenylcyclopropyl) glycine, a potent and selective antagonist of type 2 metabotropic glutamate receptors. Mol Pharmacol. 1996;50:6–9. [PubMed: 8700119]
91.
Battaglia G, Bruno V, Ngomba RT. et al. Selective activation of group-II metabotropic glutamate receptors is protective against excitotoxic neuronal death. Eur J Pharmacol. 1998;356:271–274. [PubMed: 9774259]
92.
Gottron F, Turetsky D, Choi D. SMI-32 antibody against non-phosphorylated neurofilaments identifies a subpopulation of cultured cortical neurons hypersensitive to kainate toxicity. Neurosci Lett. 1995;194:1–4. [PubMed: 7478186]
93.
Pizzi M, Fallacara C, Arrighi V. et al. Attenuation of excitatory amino acid toxicity by metabotropic glutamate receptor agonists and aniracetam in primary cultures of cerebellar granule cells. J Neurochem. 1993;61:683–689. [PubMed: 8101561]
94.
Ambrosini A, Bresciani L, Fracchia S. et al. Metabotropic glutamate receptors negatively coupled to adenylate cyclase inhibit N-methyl-D-aspartate receptor activity and prevent neurotoxicity in mesencephalic neurons in vitro. Mol Pharmacol. 1995;47:1057–1064. [PubMed: 7746273]
95.
Kingston AE, O'Neill MJ, Bond A. et al. Neuroprotective action of novel and potent ligands of group-I and group-II metabotropic glutamate receptors. Ann NY Acad Sci. 1999;890:438–439. [PubMed: 10668448]
96.
Kingston AE, O'Neill MJ, Lam A. et al. Neuroprotection by metabotropic glutamate receptor agonists: LY354740, LY379268 and LY389795. Eur J Pharmacol. 1999;377:155–165. [PubMed: 10456425]
97.
Behrens μ M, Strasser U, Heidinger V. et al. Selective activation of group II mGluRs with LY354740 does not prevent neuronal excitotoxicity. Neuropharmacology. 1999;38:1621–1630. [PubMed: 10530823]
98.
Copani A, Bruno V, Battaglia G. et al. Activation of metabotropic glutamate receptors protects cultured neurons against apoptosis induced by beta-amyloid peptide. Mol Pharmacol. 1995;47:890–897. [PubMed: 7746277]
99.
Copani A, Bruno V, Barresi V. et al. Activation of metabotropic glutamate receptors prevents neuronal apoptosis in culture. J Neurochem. 1995;64:101–108. [PubMed: 7798903]
100.
Miyamoto M, Ishida M, Kwak S. et al. Agonists for metabotropic glutamate receptors in the rat delay recovery from halothane anesthesia. Eur J Pharmacol. 1994;260:99–102. [PubMed: 7957632]
101.
Henrich-Noack P, Hatton CD, Reymann KG. The mGlu receptor ligand (S)-4C3HPG protects neurons after global ischaemia in gerbils. Neuroreport. 1998;9:985–988. [PubMed: 9601654]
102.
Bond A, O'Neill MJ, Hicks CA. et al. Neuroprotective effects of a systemically active group II metabotropic glutamate receptor agonist LY354740 in a gerbil model of global ischaemia. Neuroreport. 1998;9:1191–1193. [PubMed: 9601692]
103.
Bond A, Ragumoorthy N, Monn JA. et al. LY379268, a potent and selective group II metabotropic glutamate receptor agonist, is neuroprotective in gerbil global, but not focal, cerebral ischaemia. Neurosci Lett. 1999;273:191–194. [PubMed: 10515191]
104.
Lam AG, Soriano MA, Monn JA. et al. Effects of the selective metabotropic glutamate agonist LY354740 in a rat model of permanent ischaemia. Neurosci Lett. 1998;254:121–123. [PubMed: 9779935]
105.
Cai Z, Xiao F, Fratkin JD. et al. Protection of neonatal rat brain from hypoxic-ischemic injury by LY379268, a group II metabotropic glutamate receptor agonist. Neuroreport. 1999;10:3927–3931. [PubMed: 10716235]
106.
Klodzinska A, Chojnacka-Wojcik E, Pilc A. Selective group II glutamate metabotropic receptor agonist LY354740 attenuates pentetrazole- and picrotoxin-induced seizures. Pol J Pharmacol. 1999;51:543–545. [PubMed: 10817535]
107.
Klodzinska A, Bijak M, Chojnacka-Wojcik E. et al. Roles of group II metabotropic glutamate receptors in modulation of seizure activity. Naunyn Schmiedebergs Arch Pharmacol. 2000;361:283–288. [PubMed: 10731041]
108.
Allen JW, Ivanova SA, Fan L. et al. Group II metabotropic glutamate receptor activation attenuates traumatic neuronal injury and improves neurological recovery after traumatic brain injury. J Pharmacol Exp Ther. 1999;290:112–120. [PubMed: 10381766]
109.
Pizzi M, Benarese M, Boroni F. et al. Neuroprotection by metabotropic glutamate receptor agonists on kainate-induced degeneration of motor neurons in spinal cord slices from adult rat. Neuropharmacology. 2000;39:903–910. [PubMed: 10699456]
110.
Monyer H, Giffard RG, Hartley DM. et al. Oxygen or glucose deprivation-induced neuronal injury in cortical cell cultures is reduced by tetanus toxin. Neuron. 1992;8:967–973. [PubMed: 1350203]
111.
Bond A, Jones NM, Hicks CA. et al. Neuroprotective effects of LY379268, a selective mGlu2/3 receptor agonist: investigations into possible mechanism of action in vivo. J Pharmacol Exp Ther. 2000;294:800–809. [PubMed: 10945827]
112.
Ciccarelli R, Di Iorio P, Bruno V. et al. Activation of A1 adenosine and mGlu3 metabotropic glutamate receptors enhances the release of nerve growth factor and S-100b protein from cultured astrocytes. Glia. 1999;27:275–281. [PubMed: 10457374]
113.
Chao CC, Hu S, Kravitz FH. et al. Transforming growth factor-beta protects human neurons against beta-amyloid-induced injury. Mol Chem Neuropathol. 1994;23:159–178. [PubMed: 7702706]
114.
Prehn J H M, Bindokas VP, Marcuccilli CJ. et al. Regulation of neuronal Bcl-2 protein expression and calcium homeostasis by transforming growth factor type beta confers wide-ranging protection on rat hippocampal neurons. Proc Natl Acad Sci. 1994;91:12599–12603. [PMC free article: PMC45486] [PubMed: 7809085]
115.
Prehn JH, Miller RJ. Opposite effects of TGF-β1 on rapidly and slowly triggered excitotoxic injury. Neuropharmacology. 1996;35:249–256. [PubMed: 8783198]
116.
Prehn J H M, Bindokas VP, Jordan J. et al. Protective effect of transforming growth factor-β1 on β-amyloid neurotoxicity in rat hippocampal neurons. Mol Pharmacol. 1996;49:319–328. [PubMed: 8632765]
117.
Copani A, Condorelli F, Caruso A. et al. Mitotic signalling by b-amyloid causes neuronal death. FASEB J. 1999;13:2225–2234. [PubMed: 10593870]
118.
Henrich-Noack P, Prehn JH, Krieglstein J. TGF-β1 protects hippocampal neurons against degeneration caused by transient global ischemia. Dose-response relationship and potential neuroprotective mechanisms. Stroke. 1996;27:1609–1614. [PubMed: 8784137]
119.
Meucci O, Miller RJ. Gp120-induced neurotoxicity in hippocampal pyramidal neuron cultures: protective action of TGF-β1. J Neurosci. 1996;16:4080–4088. [PMC free article: PMC2689548] [PubMed: 8753870]
120.
Tomoda T, Shirasawa T, Yahagi YI. et al. Transforming growth factor-beta is a survival factor for neonate cortical neurons: coincident expression of type-1 receptors in developing cerebral cortices. Dev Biol. 1996;179:79–90. [PubMed: 8873755]
121.
Ren RF, Hawver DB, Kim RS. et al. Transforming growth factor-bprotects human hNT cells from degeneration induced by β-amyloid peptide: involvement of the TGF-βtype II receptor. Mol Brain Res. 1997;48:315–322. [PubMed: 9332729]
122.
Scorziello A, Florio T, Bajetto A. et al. TGF-β1 prevents gp120-induced impairment of Ca2+ homeostasis and rescues cortical neurons from apoptotic death. J Neurosci Res. 1997;49:600–607. [PubMed: 9302081]
123.
Flanders KC, Ren RF, Lippa CF. Transforming growth factor-βs in neurodegenerative disease. Prog Neurobiol. 1998;54:71–85. [PubMed: 9460794]
124.
Ruocco A, Nicole O, Docagne F. et al. A transforming growth factor-beta antagonist unmasks the neuroprotective role of this endogenous cytokine in excitotoxic and ischemic brain injury. J Cereb Blood Flow Metab. 1999;19:1345–1353. [PubMed: 10598939]
125.
Albin RL, Young AB, Penney JB. The functional anatomy of basal ganglia disorders. Trends Neurosci. 1989;12:366–375. [PubMed: 2479133]
126.
Konieczny J, Ossowska K, Wolfarth et al. LY354740, a group II metabotropic glutamate receptor agonist with potential anti-Parkinsonian properties in rats. Naunyn-Schmiedebergs Arch Pharmacol. 1998;358:500–502. [PubMed: 9826074]
127.
Wolfarth S, Konieczny J, Lorenc-Koci E. et al. The role of metabotropic glutamate receptor (mGluR) ligands in parkinsonian muscle rigidity. Amino Acids. 2000;19:95–101. [PubMed: 11026478]
128.
Helton DR, Tizzano JP, Monn JA. et al. LY354740: A metabotropic glutamate receptor agonist which ameliorates symptoms of nicotine withdrawal in rats. Neuropharmacology. 1997;36:1511–1516. [PubMed: 9517421]
129.
Helton DR, Tizzano JP, Monn JA. et al. Anxiolytic and side-effect profile of LY354740: A potent, highly selective, orally active agonist for group II metabotropic glutamate receptors. J Pharmacol Exp Ther. 1998;284:651–660. [PubMed: 9454811]
130.
Vandergriff J, Rasmussen K. The selective mGlu2/3 receptor agonist LY354740 attenuates morphine-withdrawal-induced activation of locus coeruleus neurons and behavioral signs of morphine withdrawal. Neuropharmacology. 1999;38:217–222. [PubMed: 10218862]
131.
Popik P, Kozela E, Pilc A. A selective agonist of group II glutamate metabotropic receptors, LY354740, inhibits tolerance to analgesic effects of morphine in mice. Br J Pharmacol. 2000;130:1425–1431. [PMC free article: PMC1572198] [PubMed: 10903986]
132.
Shekhar A, Keim SR. LY354740, a potent group II metabotropic glutamate receptor agonist prevents lactate-induced panic-like response in panic-prone rats. Neuropharmacology. 2000;39:1139–1146. [PubMed: 10760357]
133.
Moghaddam B, Adams BW. Reversal of phencyclidine effects by a group II metabotropic glutamate receptor agonist in rats. Science. 1998;281:1349–1352. [PubMed: 9721099]
134.
Neugebauer V, Zinebi F, Russell R. et al. Cocaine and kindling alter the sensitivity of group II and III metabotropic glutamate receptors in the central amygdala. J Neurophysiol. 2000;84:759–770. [PubMed: 10938303]
135.
Graham ME, Burgoyne RD. Activation of metabotropic glutamate receptors by L-AP4 stimulates survival of rat cerebellar granule cells in culture. Eur J Pharmacol. 1994;288:115–123. [PubMed: 7705463]
136.
Bruno V, Copani A, Bonanno L. et al. Activation of group III metabotropic glutamate receptors is neuroprotective in cortical cultures. Eur J Pharmacol. 1996;310:61–66. [PubMed: 8880068]
137.
Faden AI, Ivanova SA, Yakovlev AG. et al. Neuroprotective effects of group III mGluR in traumatic neuronal injury. J Neurotrauma. 1997;14:885–895. [PubMed: 9475370]
138.
Lafon-Cazal M, Fagni L, Guiraud MJ. et al. mGluR7-like metabotropic glutamate receptors inhibit NMDA-mediated excitotoxicity in cultured mouse cerebellar granule neurons. Eur J Neurosci. 1999;11:663–672. [PubMed: 10051767]
139.
Sabelhaus CF, Schroder UH, Breder J. et al. Neuroprotection against hypoxic/hypoglycaemic injury after the insult by the group III metabotropic glutamate receptor agonist (R, S)-4-phosphonophenylglycine. Br J Pharmacol. 2000;131:655–658. [PMC free article: PMC1572399] [PubMed: 11030711]
140.
DiFiglia M. Excitotoxic injury of the neostriatum: a model for Huntington's disease. TINS. 1990;13:286–289. [PubMed: 1695405]
141.
Gasparini F, Inderbitzin W, Francotte E. et al. (+)-4-Phosphonophenylglycine (PPG) a new group III selective metabotropic glutamate receptor agonist. Bioorg Med Chem Lett. 2000;10:1241–1244. [PubMed: 10866390]
142.
Pekhletski R, Gerlai R, Oversteet LS. et al. Impaired cerebellar synaptic plasticity and motor performance in mice lacking the mGluR4 subtype of metabotropic glutamate receptor. J Neurosci. 1996;16:6364–6373. [PMC free article: PMC6578923] [PubMed: 8815915]
143.
Bradley SR, Standaert DG, Rhodes KJ. et al. Immunohistochemical localization of subtype 4a metabotropic glutamate receptors in the rat and mouse basal ganglia. J Comp Neurol. 1999;407:33–46. [PubMed: 10213186]
144.
Kosinski CM, Risso Bradley S, Conn PJ. et al. Localization of metabotropic glutamate receptor 7 mRNA and mGluR7a protein in the rat basal ganglia. J Comp Neurol. 1999;415:266–84. [PubMed: 10545164]
145.
Thomsen C, Hampson DR. Contribution of metabotropic glutamate receptor mGluR4 to L-2-[3H]amino-4-phosphonobutyrate binding in mouse brain. J Neurochem. 1999;72:835–840. [PubMed: 9930760]
146.
Martin G, Nie Z, Siggins GR. Metabotropic glutamate receptors regulate N-methyl-D-aspartate-mediated synaptic transmission in nucleus accumbens. J Neurophysiol. 1997;78:3028–3038. [PubMed: 9405522]
147.
Maiese K, Vincent A, Lin SH. et al. Group I and group III metabotropic glutamate receptor subtypes provide enhanced neuroprotection. J Neurosci Res. 2000;62:257–272. [PubMed: 11020218]
148.
Doble A. The role of excitotoxicity in neurodegenerative disease: Implications for therapy. Pharmacology & Therapeutics. 1999;81:163–221. [PubMed: 10334661]
149.
Wood PL. NMDA antagonists for stroke and head trauma: current status. Exp Opin Invest Drugs. 1998;7:1505–1508. [PubMed: 15992048]
150.
Abdul-Ghani AS, Attwell PJ, Singh-Kent N. et al. Anti-epileptogenic and anticonvulsant activity of L-2-amino-4-phosphonobutyrate, a presynaptic glutamate receptor agonist. Brain Res. 1997;755:202–212. [PubMed: 9175888]
151.
Chapman AG, Nanan K, Yip P, Meldrum BS. Anticonvulsant activity of a metabotropic glutamate receptor 8 preferential agonist, (R,S)-4-phosphonophenylglycine. Eur J Pharmacol. 1999;383:23–27. [PubMed: 10556677]
152.
Tang E, Yip PK, Chapman AG. et al. Prolonged anticonvulsant action of glutamate metabotropic receptor agonists in inferior colliculus of genetically epilepsy-prone rats. Eur J Pharmacol. 1997;327:109–115. [PubMed: 9200548]
Copyright © 2000-2013, Landes Bioscience.
Bookshelf ID: NBK6556

Views

  • PubReader
  • Print View
  • Cite this Page

Related information

  • PMC
    PubMed Central citations
  • PubMed
    Links to PubMed

Recent Activity

Your browsing activity is empty.

Activity recording is turned off.

Turn recording back on

See more...