Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

Review Article Biased agonism and allosteric modulation of metabotropic 5

Phuc N.H. Trinh, Lauren T. May, Katie Leach and Karen J. Gregory Drug Discovery Biology, Monash Institute of Pharmaceutical Sciences and Department of Pharmacology, Monash University, Parkville, VIC, Australia Correspondence: Karen Gregory ([email protected])

Metabotropic glutamate receptors belong to class C G-protein-coupled receptors and con- sist of eight subtypes that are ubiquitously expressed throughout the central nervous sys- tem. In recent years, the metabotropic glutamate receptor subtype 5 (mGlu5) has emerged as a promising target for a broad range of psychiatric and neurological disorders. Drug dis- covery programs targetting mGlu5 are primarily focused on development of allosteric mod- ulators that interact with sites distinct from the endogenous agonist glutamate. Significant efforts have seen mGlu5 allosteric modulators progress into clinical trials; however, recent failures due to lack of efficacy or adverse effects indicate a need for a better understand- ing of the functional consequences of mGlu5 allosteric modulation. Biased agonism is an interrelated phenomenon to allosterism, describing how different ligands acting through the same receptor can differentially influence signaling to distinct transducers and pathways. Emerging evidence demonstrates that allosteric modulators can induce biased pharmacol- ogy at the level of intrinsic agonism as well as through differential modulation of orthosteric agonist-signaling pathways. Here, we present key considerations in the discovery and devel- opment of mGlu5 allosteric modulators and the opportunities and pitfalls offered by biased agonism and modulation.

Introduction Glutamate is the principal excitatory neurotransmitter in the mammalian central nervous system (CNS) and elicits its activity via the ionotropic and metabotropic glutamate (mGlu) receptors. The ionotropic glutamate receptors are ligand-gated ion channels, while the mGlu receptors belong to the class C G-protein-coupled receptors (GPCRs) – a group that also includes calcium sensing receptor (CaSR), γ-aminobutyric acid receptor B (GABAB) and numerous taste, pheromone, and orphan receptors. Class C GPCRs share the characteristic 7 transmembrane (7TM) α-helical bundle architecture separated by al- ternating intracellular and extracellular loops. The 7TM domain links an intracellular carboxy-terminal domain and a large bilobed extracellular amino-terminus, termed the Venus flytrap (VFT) domain [1-3], where endogenous ligands bind (Figure 1A). Furthermore, the mGlu receptors are also known for their obligatory constitutive dimerization [4] mediated via a disulfide bond connecting the top of the VFT domains [5,6]. Evidence suggests that mGlu receptors can form homo- or heterodimers with other sub- types [7-12] and unrelated GPCRs [13-17]. Glutamate and other orthosteric agonists typically bind in a cleft between two lobes that comprise the VFT domain to induce active ‘closed–closed’ (both clefts of the VFT dimer closed) or ‘open–closed’ (one cleft open, one cleft closed) conformations [18]. Conforma- Received: 16 August 2018 tional rearrangements within the VFT domains are transmitted to the 7TM domain via a cysteine-rich Revised: 03 October 2018 region, to modulate G-protein coupling, subsequently triggering intracellular signaling [19]. Based on Accepted: 05 October 2018 sequence homology, preferred signal transduction pathways and pharmacological profiles, the mGlu re- Version of Record published: ceptors are subclassified into group I (mGlu1 and mGlu5 (metabotropic glutamate receptor subtype 5)), 02 November 2018 group II (mGlu2 and mGlu3), and group III (mGlu4,mGlu6,mGlu7,andmGlu8) [20]. Group I mGlu

c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 2323 Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

Figure 1. Schematic representation of orthosteric agonism and allosteric modulation. (A) Glutamate binds to the cleft between two lobes of VFT domain, whereas allosteric modulators (e.g. PAM, NAM, and NAL) bind to the common allosteric site in the 7TM domain. Simultaneous binding of an orthosteric and allosteric ligand at a receptor can have two dominant effects; affinity modulation (orange arrow) and/or efficacy modulation (green arrow).B ( ) Quantification of allosteric

effects with operational model of allosterism [97]. In the equation, KA and KB denote the equilibrium dissociation constant of the orthosteric ligand A and allosteric ligand B, respectively. The parameter α governs allosteric modulation of binding affinity. Allosteric modulation of orthosteric ligand efficacy is incorporated by the empirical parameter, β which scales from zero to infinity. The param-

eters τA and τB reflect the ability of the orthosteric and allosteric ligands, respectively to promote receptor activation and incorporate the intrinsic efficacy of each ligand, the receptor density, and the efficiency of stimulus-responsem coupling.E denotes the maximal system response and n is the slope factor of the transducer function that links receptor occupancy to the final observed response. Abbreviations: NAL, neutral allosteric ligand; NAM, negative ; PAM, positive allosteric modulators.

members are primarily located postsynaptically and preferentially couple to Gαq/11, resulting in activation of phos- pholipase C, production of inositol 1,4,5-trisphosphate and diacylglycerol, and subsequent mobilization of intracellu- 2+ lar calcium (iCa ). The group II and group III receptors, with the exception of mGlu6, which is exclusively localized to retina, are typically found presynaptically and functionally linked to Gαi/o and the inhibition of adenylyl cyclase activity [20]. Despite being members of the same group, mGlu1 and mGlu5 receptors have distinct functional roles within the 2+ brain [21], which have been linked to their iCa mobilization profiles. For instance, mGlu5 activation stimulates 2+ 2+ 2+ iCa oscillations, whereas mGlu1 primarily elicits sustained and non-oscillatory Ca responses [22,23]. The iCa oscillations induced by mGlu5 occur via a dynamic mechanism involving phosphorylation and dephosphorylation of the receptor at Ser839 [24,25]. In addition to coupling to Gαq/11, a recent study indicated that purified mGlu5 activates Gαs upon glutamate stimulation to increase intracellular cyclic adenosine monophosphate (cAMP) [26]. Further, mGlu5 activates additional non-Gαq/11-dependent signaling cascades and regulates gene expression via its expression on intracellular membranes of the endoplasmic reticulum and nucleus [27,28]. Importantly, the distinct signaling profiles of intracellular and cell surface mGlu5 result in different physiological effects [29,30]. For example, activation of cell surface mGlu5 leads to long-term potentiation and long-term depression, whereas intracellular mGlu5 mediates only long-term depression [31]. mGlu5 is expressed ubiquitously throughout the postsynaptic densities of the cerebral cortex, corpus stria- tum, olfactory bulb, hippocampus, caudate nucleus, and nucleus accumbens. Further, mGlu5 is also expressed in non-neuronal cells throughout the brain, including astrocytes and microglia [32-36]. mGlu5 mediates diverse ef- fects including synaptic transmission and plasticity, and proliferative, growth and survival responses of neurons and glial cells in the CNS [20,37-47] and periphery [48]. Thus, mGlu5 has emerged as an attractive drug target for a

2324 c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

broad range of psychiatric, neurological, and neurodevelopmental disorders. Multiple excellent reviews have exten- sively covered the therapeutic potential of mGlu5 in schizophrenia, addiction, anxiety, depression, Parkinson’s and Alzheimer’s diseases, fragile X syndrome (FXS), autism spectrum disorders, chronic pain, and gastroesophageal reflux disease (GERD) [49-51]. Promisingly, several mGlu5 inhibitors (i.e. [52-54], /RG7090 [55-58], /AFQ056 [59-63], raseglurant/ADX10059 [64-66], and /ADX48621 [67]) advanced to clini- cal trials. Despite significant promise in preclinical settings, mGlu5-targetted compounds have yet to reach the market, attributable to unanticipated adverse effects or lack of efficacy in phase II or III trials. These failures suggest there are unresolved problems and/or unappreciated complexity in targetting mGlu5. Herein, we review the promise and chal- lenges of targetting mGlu5 with allosteric modulators in light of emerging evidence surrounding the phenomenon of biased agonism.

Targetting mGlu5 receptors: allosteric modulators GPCRs constitute the largest family of human membrane receptors with more than 800 members [1], and as such are long-standing therapeutic targets, accounting for approximately 34% of all drugs approved by the US Food and Drug Administration [68]. The prevailing approach for targetting GPCRs has focused on the endogenous ligand binding site, termed orthosteric site, to mimic or block the endogenous ligand activity in a competitive manner. Several mGlu5 orthosteric ligands (i.e. L-quisqualate [69,70], (S)-DHPG [71,72], (S)-2-chloro-5-hydroxyphenylglycine ((S)-CHPG) [70,73], and (S)-α-methyl-4-carboxyphenylglycine [70,74,75]) have been discovered. However, orthosteric ligands largely retain affinity and potency at other mGlu receptor subtypes, likely owing to high conservation of orthosteric sites. Thus, many research groups have sought to target secondary binding pockets, referred to as allosteric sites, that are topographically distinct and show greater diversity between receptor subtypes, offering the potential for improved subtype selectivity. For mGlu5, a wealth of site-directed mutagenesis studies (reviewed in [76]) and recent co-crystal structures with allosteric modulators [77-79] have defined one allosteric site within the 7TM bundle (Figure 1A). Ad- ditional sites have also been proposed within the 7TM domain; however, the locations remain unknown. Therefore, for mGlu5 and other Class C GPCRs, these 7TM allosteric sites are considerably distant from the orthosteric pock- ets within the VFTs. Four allosteric modulators of GPCRs have now entered the market: cinacalcet, evocalcet, and etelcalcitide (allosteric potentiators of the CaSR for the treatment of hyperparathyroidism or parathyroid carcinoma) [80-82] and maraviroc (an allosteric inhibitor of the chemokine receptor 5 for advanced HIV disease) [83]. Discov- ery and characterization of mGlu5 allosteric modulators have gained significant attention offering novel avenues for targetted drug development for CNS disorders.

Overview of allosteric modulation Allosterism arises as a consequence of the conformational flexibility of GPCRs. The classic model of allosterism, the Monod–Wyman–Changeux model, posits that allosteric proteins exist in a dynamic equilibrium between multiple conformational states, even in the absence of ligand. Further, allosteric proteins possess multiple binding sites, and binding of a ligand (or other entity) to any of these sites favors a subset of possible conformational states [84]. In doing so, allosteric modulators engender GPCR conformational states that are less stable in the presence of an orthosteric ligand alone, influencing the interactions with transducers. Allosteric modulators are classified based on the magni- tude and direction of their effect on orthosteric ligand affinity and/or efficacy, a property referred to as cooperativity. Positive allosteric modulators (PAMs) enhance the binding and/or activity of orthosteric ligands, whereas negative allosteric modulators (NAMs) inhibit the binding or response to orthosteric agonist. Last, neutral allosteric ligands (NALs) bind to the receptor, but do not alter the binding and/or function of the orthosteric ligand [85,86]. A continuing challenge for allosteric modulator-based drug discovery is the ability to detect, validate, and quan- tify the pharmacology of allosteric modulators. All current methods (e.g. radioligand binding and functional assays) that assess allosteric interactions require the availability of a probe to provide a relevant readout of receptor-based activity. When screening for allosteric interactions, it is imperative to consider both methods to avoid missing any behaviors of allosteric ligands [87,88]. Binding assays have the advantage of direct validation of an allosteric mode of action and reveal the site of interaction of allosteric ligands [89,90]; however, they do not assess efficacy modulation. For functional assays, the most relevant probe is the endogenous agonist; however, it is not always feasible to use the endogenous agonist. Thus, surrogate orthosteric probes are often required. Functional assays can readily detect a wider spectrum of allosteric behaviors and allow for detection of modulation of both affinity and efficacy. Further, when designing experiments to study allosterism, it is worth considering probe dependence and saturation of effect, two basic allosteric modulator properties. Probe dependence describes the phenomenon where the degree and di- rection of cooperativity is determined by the chemical nature of the orthosteric ligand (probe). Cooperativity might

c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 2325 Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

beapparentwithoneprobe,buttheobservedinteractionmaybenegligibleorinanopposingdirectionwhenusing another [91-93]. A second characteristic is the saturability of effect; that is, the influence of allosteric modulators is limited by the cooperativity between orthosteric and allosteric sites. The degree of cooperativity introduces a ceiling onthemagnitudeoftheallostericeffect,andassuch,nofurthermodulationwillbeobservedwhensaturatingcon- centrations of allosteric modulators are present. Therefore, allosteric modulators may have reduced risk in the event of drug overdose [89,94]. Allosteric interactions can be quantified by numerous models, namely the allosteric ternary complex model (ATCM) [95], an allosteric ‘two-state’ model (ATSM) [96], or the operational model of allosterism [89,97]. Of these, the ATSM and operational model of allosterism offer the ability to quantify both affinity and effi- cacy modulation. However, the ATSM contains many parameters that cannot be readily fitted to experimental data; therefore, it is more feasible and practical to use the operational model of allosterism (Figure 1B), which combines the ATCM with the classic Black-Leff operational model of agonism [11]. The major advantage of this model is the capability of quantifying modulator efficacy as well as delineating modulator effects on orthosteric agonist affinity compared with efficacy [11,97-99], providing valuable information to guide lead-discovery programs.

mGlu5 allosteric pharmacology Allosteric modulators of mGlu5 are attractive putative therapeutics for a number of reasons. As previously men- tioned, allosteric modulators may achieve subtype selectivity due to sequence divergence between allosteric pock- ets across receptor subtypes, which is important given that there are eight mGlu subtypes. Further, in the absence of intrinsic efficacy, allosteric modulators have the potential to maintain the spatial and temporal profile of en- dogenous neurotransmitters. Pure allosteric modulators only elicit an effect when and where the endogenous ag- onist is present, thereby tuning up, or down, the existing responses in proportion to normal physiological tone. In contrast, orthosteric ligands continuously stimulate GPCR activity, or block the action of the endogenous ag- onist, for the duration of their presence. This profile may be particularly detrimental in the CNS where recep- tor signaling must be precisely controlled. It is worth noting that some allosteric modulators may possess intrin- sic efficacy to either activate or inhibit GPCR-mediated signaling in the absence of orthosteric ligands and are re- ferred to as agonist-PAMs (ago-PAMs) or inverse agonist-NAMs. Last, allosteric modulators can have differential cooperativity with respect to orthosteric agonist affinity compared with efficacy. For instance, the mGlu1 NAM, 7-(hydroxyimino)cyclopropa[b]chromen-1a-carboxylate ethyl ester (CPCCOEt), completely inhibits glutamate ef- ficacy in chinese hamster ovary (CHO) cells but does not impact glutamate binding, suggesting that CPCCOEt has neutral cooperativity with respect to glutamate affinity and negative cooperativity at the level of efficacy [100]. Discovery of mGlu5 allosteric ligands has been particularly successful with multiple diverse chemotypes disclosed covering the full spectrum of pharmacology including, pure PAMs, ago-PAMs, NALs, partial and full NAMs, largely 2+ classified based on glutamate-mediated iCa mobilization assays (see Table 1, Figure 2). The first mGlu5 allosteric modulators discovered were NAMs, with subsequent intense discovery programs by both academic and industrial researchers yielding diverse chemotypes [20,101,102]. Inhibition of mGlu5 has been confirmed as a potential mech- anism for treatment of depression, anxiety, Parkinson’s disease levodopa-induced dyskinesias (PD-LID), FXS and Alzheimer’s disease, with mGlu5 NAMs showing efficacy in preclinical models [44,55,57,103-108]. Despite significant efforts in preclinical settings, several mGlu5 NAMs were reported to have detrimental effects, such as cognitive im- pairments and psychotomimetic effects in rodents [53,109-112]. Moreover, enthusiasm for mGlu5 NAMs has waned given recent failures in clinical trials for depression and FXS, although ongoing studies are being pursued for PD-LID and dystonia [59,106]. Recent studies have shown that the partial mGlu5 NAMs (where cooperativity between the allosteric ligand and glutamate is limited), 2-(2-(3-methoxyphenyl)ethynyl)-5-methylpyridine (M-5MPEP) and 2-(2-(5-bromopyridin-3-yl)ethynyl)-5-methylpyridine, reduce cocaine-mediated addiction behaviors and have comparable anxiolytic and antidepressant activity to the full NAM, 3-((2-Methyl-4-thiazolyl)ethynyl)pyridine (MTEP) [112]. Importantly, the partial NAMs did not exhibit psychotomimetic adverse effects. N,N-diethyl-5-((3-fluorophenyl)ethynyl)picolinamide (VU0477573), another partial mGlu5 NAM, displays robust efficacy in the anxiolytic behavioral model as comparable with MTEP [113]. Therefore, partial NAM activity may provide one means to achieve a desirable therapeutic window for mGlu5 NAMs via fine-tuning rather than complete blockade of mGlu5 activity. Enhancement or activation of mGlu5 is a promising strategy to treat psychosis, cognitive disorders and drug addiction, with mGlu5 PAMs demonstrating efficacy in preclinical models [110,114-118]. However, there is growing evidence for CNS-related adverse effect liability for certain mGlu5 PAMs and ago-PAMs, such as the seizure liability caused by N-cyclobutyl-5-((3-fluorophenyl)ethynyl)picolinamide (VU0403602) [119]

2326 c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

Table 1 Allosteric pharmacology and bias evidence for selected mGlu5 allosteric ligands

Ligand name Pharmacology Biological/clinical significance References

PAMs and PAM-agonists (ago-PAMs) 2+ 5PAM523 PAM: glu efficacy for iCa (CHO-mGlu5 & rat cortical Antipsychotic-like activity with adverse [120,121] 3 astrocytes); [ H]quisqualate binding (CHO-mGlu5); effects (seizures, neurotoxicity) in DHPG efficacy for NMDA-evoked currents (rat rodents hippocampus) 2+ ADX47273 PAM: glu, quisqualate efficacy for iCa and IP1 Antipsychotic-like and procognitive [115,121,139] (HEK-mGlu5 & rat cortical astrocytes); effects 3 [ H]quisqualate binding (CHO-mGlu5); DHPG efficacy for NMDA-evoked currents (rat hippocampal CA1 pyramidal cells) 2+ CDPPB PAM: glu, quisqualate efficacy for iCa and IP1 Antipsychotic-like and procognitive, HD, [91,99,114,120,139-142] (CHO-mGlu5 & rat astrocytes) no reported adverse effects in rodents 2+ Ago-PAM: glu efficacy for iCa (CHO-mGlu5); pERK1/2 (BACHD mice); p-Akt (mouse striatal neurons) Bias: agonism (relative to DHPG): IP1 > pERK1/2 2+ and iCa (HEK-mGlu5 & mouse cortical neurons) CPPHA PAM: glu, quisqualate, DHPG efficacy for iCa2+ and Unsuitable for in vivo studies due to [99,137,139,143,144] IP1 (CHO-mGlu5 and rat astrocytes) poor brain impenetrance Ago-PAM: glu efficacy for pERK1/2 (HEK-mGlu5) Bias: agonism (relative to glu): iCa2+ > pERK1/2 (HEK-mGlu5), ago-NAM: DHPG - pERK1/2 (rat astrocytes) 2+ DPFE PAM: glu, DHPG efficacy for iCa ,IP1,pERK1/2 Antipsychotic-like effects, no reported [91,117] (HEK-mGlu5 and mouse cortical neurons) adverse effects in rodents 2+ Bias: agonism (relative to DHPG): IP1 >> iCa , biphasic pERK1/2 (mouse cortical neurons) 2+ VU0360172 PAM: glu, DHPG efficacy for iCa (HEK-mGlu5) Antipsychotic-like effects, no known [91,99] Ago-PAM: glu, DHPG efficacy for IP1,pERK1/2 adverse effects in rodents (HEK-mGlu5 & mouse cortical neurons) 2+ Bias: agonism (relative to DHPG): IP1 >> iCa and pERK1/2 (mouse cortical neurons) VU0409551 (JNJ-46778212) PAM: glu efficacy for iCa2+ and pERK1/2 Antipsychotic-like, HD, no reported [121,138,145,146] (HEK-mGlu5 and cortical astrocytes), DHPG-LTD adverse effects in mice (hippocampal slices) Bias: agonism (relative to DHPG): IP1 > pERK1/2 (mouse cortical neurons); no modulation of DHPG efficacy for NMDA-evoked currents (rat hippocampus) 2+ VU0424465 Ago-PAM: glu, DHPG efficacy for iCa ,IP1,and Seizurogenic and neurotoxic in rats [26,91,119] cAMP accumulation, pERK1/2 (HEK-mGlu5,rat cortical astrocytes, neurons), DHPG efficacy for hippocampal LTD Bias: agonism (relative to DHPG): IP1  pERK1/2 2+ >>> iCa (HEK- mGlu5 and mouse cortical neurons) NAMs and partial NAMs 2+ Basimglurant/RG7090 NAM: quisqualate efficacy for iCa and IP1 Depression, FXS [55-58] (HEK-mGlu5); inverse agonism: IP1 2+ CTEP NAM: quisqualate efficacy for iCa and IP1 Alzheimer’s disease, FXS [108,147,148] (HEK-mGlu5); inverse agonism: IP1 2+ Dipraglurant/ADX48621 NAM: glu efficacy for iCa (HEK- mGlu5) PD-LID [67,149] 2+ Fenobam NAM: quisqualate efficacy for iCa and IP1 FXS [53,54] (HEK-mGlu5); inverse agonism: IP1 2+ GET73 NAM: quisqualate, glu, CHPG efficacy for iCa ,IP1, Alcohol dependence, anxiety [150,151] pCREB (rat cortical astrocytes and neurons, hippocampal slices) 2+ Mavoglurant/AFQ056 NAM: glu efficacy for iCa (L(tk-)-mGlu5)andIP1 FXS, PD-LID, GERD, HD [59-63,104,152] (CHO- mGlu5) MTEP NAM: quisqualate, glu efficacy for iCa2+ (L(tk-)-, Alzheimer’s disease, addiction, anxiety, [113,153,154] HEK-mGlu5); inverse agonism: IP1 depression, epilepsy, FXS, GERD, pain, HD, PD-LID, psychotomimetic, and cognitive adverse effects

Continued over

c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 2327 Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

Table 1 Allosteric pharmacology and bias evidence for selected mGlu5 allosteric ligands (Continued)

Ligand name Pharmacology Biological/clinical significance References

MPEP NAM: quisqualate and DHPG efficacy for IP1 [99,139,140,155-160] 2+ (CHO-mGlu5); glu efficacy for iCa and pERK1/2 (HEK-mGlu5); glu, DHPG and quisqualate induced 2+ iCa oscillations and IP1 (rat astrocytes); inverse agonism: IP1 M-5MPEP NAM: DHPG, and quisqualate mediated iCa2+ Addiction, anxiety, depression, no [99,112,139,159] oscillations (rat cortical astrocytes); glu efficacy for psychotomimetic effects pERK1/2 (HEK-mGlu5) Partial NAM: glu efficacy for iCa2+; quisqualate and DHPG efficacy for IP1 (HEK-mGlu5, rat cortical astrocytes and cortex membranes)

VU0477573 NAM: glu efficacy for pERK1/2 (HEK-mGlu5) Anxiety disorders [113] 2+ Partial NAM: glu efficacy for iCa and IP1 (HEK-mGlu5 and cortical neurons) NALs 2+ 5MPEP NAL: glu, DHPG efficacy for iCa ,IP1 (rat cortical Pharmacological tool compound [139,144,159,161] astrocytes) 2+ BMS-984923 NAL: glu, DHPG efficacy for iCa (HEK-mGlu5 and Alzheimer’s disease [162,163] rat cortical neurons)

Underlined text in the ‘biological/clinical significance’ column denotes human clinical evidence associated with these allosteric ligands. The following abbreviations are used in the table: 5PAM523, 5-fluoro-2-(61-oxadiazol-5-yl)pyridine; CPPHA, N-[4-Chloro-2-[(1,3-dihydro-1,3-dioxo-2H-isoindol-2-yl)methyl]phenyl]-2-hydroxybenzamide; pCREB, phosphorylation of cAMP response element binding protein; CTEP, 2-Chloro-4-[2-[2,5-dimethyl-1-[4-(trifluoromethoxy)phenyl]-1H-imidazol-4-yl]ethynyl]pyridine; DPFE, 1-(4-(2,4-difluorophenyl)piperazin-1-yl)-2-((4-fluorobenzyl)oxy)ethenone; HD, Huntington’s disease; IP1, inositol monophosphate; L(tk-), mouse fibroblast; NMDA, N-methyl-D-aspartate; VU0409551, ((4-fluorophenyl)(2-(phenoxymethyl)-6,7-dihydrooxazolo[5,4-c]pyridin-5(4H)-yl)methanone); VU0360172, N-cyclobutyl-6-((3-fluorophenyl)ethynyl)picolinamide; VU0424465, (R)-5-((3-fluorophenyl)ethynyl)-N-(3-hydroxy-3-methylbutan-2-yl)picolinamide.

and (R)-5-((3-fluorophenyl)ethynyl)-N-(3-hydroxy-3-methylbutan-2-yl)picolinamide (VU0424465) [119] and fore- brain neurotoxicity induced by 5-fluoro-2-(61-oxadiazol-5-yl)pyridine (5PAM523) [120]. The observed side effects have been attributed to excessive mGlu5 activation by PAMs with intrinsic efficacy, such that it may be prudent to optimize mGlu5 PAMs devoid of intrinsic agonist activity. However, some ‘pure’ mGlu5 PAMs can still cause seizures and cytotoxicity after chronic administration [120], highlighting that intrin- sic agonism may not be solely responsible for ontarget adverse effects. Indeed, the recent discovery of ((4-fluorophenyl)(2-(phenoxymethyl)-6,7-dihydrooxazolo[5,4-c]pyridin-5(4H)-yl)methanone) (VU0409551) chal- lenges this idea and suggests that not all mGlu5 ago-PAMs are equal. VU0409551 has ago-PAM activity with glu- tamate in extracellular signal-regulated kinases 1 and 2 (ERK1/2) phosphorylation assays in recombinant cells but 2+ is a pure PAM in iCa mobilization assays [91,121]. Further, VU0409551 is distinguished from other mGlu5 PAM chemotypes in slice electrophysiology preparations from rat hippocampus where it has no efficacy for enhancement of N-methyl-D-aspartate (NMDA) receptor dependent long-term potentiation [121]. Collectively, these data demon- 2+ strate that reliance on iCa mobilization assays to classify mGlu5 PAMshaslikelyresultedinanunderappreciation of the full scope of mGlu5 PAM activity and the optimal pharmacological profile to achieve therapeutic benefit.

Emerging concept: mGlu5 allosteric modulators engender biased signaling While discovery of mGlu5 allosteric modulators has been successful, these drugs have failed to progress to market. Oneofthemajorunderlyingissuesisarelianceonlimitedin vitro characterization and therefore a failure to capture thecompletepharmacologicalprofileofaparticularcompound.Overrecentdecades,ithasbecomeincreasinglyap- preciated that different ligands acting at the same receptor preferentially stabilize different receptor conformations, such that each ligand may engage a subset of intracellular transducers, to the relative exclusion of others (Figure 3) [122,123]. Now referred to as biased agonism, this phenomenon has also been known as functional selectivity [124],

2328 c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

Figure 2. Representative chemotypes for mGlu5 allosteric modulators

Figure 3. Schematic depicting the concept of biased agonism and allosteric modulation The binding of endogenous ligand (gray square) activates a receptor and results in intracellular signal transduction. The surrogate orthosteric agonist (green circle) may favor stimulation of one response over another when compared with the endogenous ligand. The binding of an allosteric modulator (blue) may modulate functional responses to an agonist equally across all signaling pathways. However, the interaction of another modulator (orange) may differentially modulate the response to the same agonist in a pathway dependent fashion. In this scenario, response 1 is potentiated whereas response 2 is inhibited.

c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 2329 Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

stimulus trafficking [125], and ligand-directed signaling [126]. Biased agonism and allosteric modulation are inter- linked concepts, offering the potential to fine-tune GPCR activity to preference therapeutically beneficial signaling pathways and avoid those associated with adverse effects. This concept is attractive because it could provide a new av- enue for the development of drugs that are not only ‘receptor subtype-selective’,but also ‘pathway-selective’.Although biased agonism offers great clinical potential, there are significant challenges to translation. First, there is a limited understanding of specific or combined pathways leading to therapeutically relevant outcomes, and hence the rational design of biased ligands remains difficult [85]. The identification of signaling pathways responsible for therapeutic compared with adverse effects is not straightforward. Therefore, selection of appropriate end point experiments is complex since the desirable signaling profile for most drug targets has not yet been well-defined. Second, biased ago- nism is highly dependent on the ligand and cellular background in which it acts (ligand and system bias, respectively). A particular bias profile in a heterologous system does not guarantee the same profile will be observed in endoge- nous systems or in vivo [127]. Third, since ligand bias is influenced by cellular context (context-dependent), biased agonism may conceivably change with alterations in membrane composition, proteins and signaling partners as a consequence of disease progression or the natural differences between different tissue/organ environments [128]. For example, in the setting of Alzheimer’s disease accumulation of pathological protein aggregates (e.g. amyloid β and hyperphosphorylated tau) and inflammation can change how cells respond to mGlu5 stimulation [129-131]. In recent years, the concept of biased agonism has been extended to include biased allosteric modulation. A notable example is revealed by an autoantibody that targets the CaSR, which enhances Gαq/11 coupling to inositol phosphate accumulation but inhibits ERK1/2 phosphorylation [132]. Such modulators engendering biased agonism offer great potential to dissect the physiology linked to therapeutic and adverse effects and to tailor receptor activity for improved therapeutic effects. Despite the challenges, the development of ligands with desired selectivity toward particular signaling pathways is of considerable interest. The ability to experimentally detect ligand bias has necessitated the development of methods to quantify bias in order to guide structure-activity studies and drug candidate selection. Historically, comparisons of agonist potency or maximal effects between different pathways were used as indicators of biased agonism; however, these approaches are suboptimal. One such method to quantify bias is the calculation of ‘transduction coefficients’ [133]. This approach, which is based on the Black and Leff (1983) operational model of agonism [11], can be applied to concentration-response curves to obtain a single parameter that describes bias between signaling pathways in a system independent manner [123]. However, even if transduction coefficients are used to establish true bias, it is still imperative to consider that agonist responsiveness may be influenced by ligand-independent factors, such as differing receptor-transducer coupling efficiencies and observational bias resulting from differing assay conditions and sensi- tivities. Therefore, a critical aspect for biased quantification is the need to exclude both system and observational bias by comparing bias factors to a reference ligand and pathway.

Biased agonism in mGlu receptors Biased agonism has been demonstrated for orthosteric agonists at mGlu1 [72,134], mGlu4,mGlu7,andmGlu8 [135]. For instance, when compared with two synthetic orthosteric agonists (L-2-amino-4-phosphonobutyrate and (2S)-2-amino-4-([[4-(carboxymethoxy)phenyl](hydroxy)methyl](hydroxy)phosphoryl)butanoic acid), glutamate is 2+ biased toward iCa mobilization over G-protein-coupled inwardly rectifying potassium channel activation at mGlu4 and mGlu8.Intriguingly,thereverseistrueatmGlu7 [135]. At closely related mGlu1, orthosteric agonists includ- ing DHPG, a surrogate agonist often used in native tissue studies of mGlu5 allosteric modulators, are biased relative to glutamate [72]. The potential for surrogate orthosteric agonists to induce biased agonism relative to the endoge- nous ligand is a critical consideration for allosteric modulator discovery programs and a potential confound when classifying modulator pharmacology.

Biased mGlu5 allosteric modulators: an emerging concept Biased agonism does not apply exclusively to orthosteric agonists but is also observed for mGlu5 allosteric modu- lators, with respect to both agonism and cooperativity. Comparison of the allosteric agonist activity of early mGlu5 ago-PAMs, N-(1,3-diphenyl-1H-pyrazolo-5-yl)-4-nitrobenzamide (VU29), and CDPPB, with glutamate, showed that in recombinant cells VU29 and CDPPB have higher intrinsic efficacy for signaling to ERK1/2 phosphorylation rela- tive to iCa2+ mobilization, while glutamate has higher intrinsic efficacy for iCa2+ mobilization over ERK1/2 phos- 2+ phorylation [99]. Multiple reports have since shown compounds classified as pure mGlu5 PAMs based on iCa mobilization to be ago-PAMs with respect to ERK1/2 phosphorylation [91,121,136]. Recent bias profiling for di- verseago-PAMsacrossmultiplemeasuresofmGlu5 activity in both recombinant and native cells revealed distinct

2330 c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

bias fingerprints that are postulated to be linked to different preclinical profiles. VU0424465, an mGlu5 ago-PAM with known adverse effect liability, showed the most extreme biased profile in both cell types relative to DHPG [91]. 1-(4-(2,4-difluorophenyl)piperazin-1-yl)-2-((4-fluorobenzyl)oxy)ethenone (DPFE) and VU0409551, which have not been associated with adverse effects, were also biased agonists relative to DHPG, but their bias fingerprints were dis- tinct from VU0424465 and from one another. Beyond biased agonism, several studies have revealed bias of mGlu5 allosteric modulators at the level of cooperativity, referred to as biased modulation (Figure 3). For instance, in rat cortical astrocytes N-[4-Chloro-2-[(1,3-dihydro-1,3-dioxo-2H-isoindol-2-yl)methyl]phenyl]-2-hydroxybenzamide is a PAM in 2+ iCa mobilization but a NAM for ERK1/2 phosphorylation [137]. More recently, diverse mGlu5 PAMs and ago-PAMs with different chemical scaffolds have exhibited differential cooperativity with DHPG depending on the 2+ measure of receptor activation (inositol monophosphate (IP1) accumulation, ERK1/2 phosphorylation, or iCa mobilization) in primary neuronal cultures [91]. Importantly, the differences in cooperativity were not uniform for all PAMs; N-cyclobutyl-6-((3-fluorophenyl)ethynyl)picolinamide (VU0360172) had similar cooperativity with DHPG between iCa2+ mobilization and pERK1/2, whereas DPFE was a PAM for iCa2+ but neutral or negative for pERK1/2 [91]. Therefore, relative to VU0360172, DPFE is a biased allosteric modulator of mGlu5. Unappreciated bias may also be a contributing factor to unanticipated adverse effects or non-selective actions, with a recent study demonstrating that the apparent selectivity of allosteric ligands for class C GPCRs is largely driven by cooperativity and bias [88]. Furthermore, bias fingerprints are distinct between recombinant and native cells, highlighting the contribution of cellular context, which can confound classification of allosteric modulator pharmacology. An excellent example is VU0409551, which does not potentiate NMDA receptor dependent synaptic plasticity in the hippocampus of wild type animals [121], but does enhance hippocampal NMDA receptor activity in the serine racemase knockout mouse model of schizophrenia [138]. The increasing evidence for biased modulation by diverse mGlu5 PAM chemotypes underscores the paucity of our understanding about chemical structure and functional consequences of mGlu5 allosteric modulation. These knowledge gaps are largely attributable to minimal characterization of the effects of allosteric ligands on different receptor-mediated behaviors. Thus, to identify biased mGlu5 allosteric modulators, future drug discovery campaigns should consider multiple measures of mGlu5 activation. Collectively, the emerging concept of biased allosteric modulation highlights the inherent complexity and challenges for allosteric modulator discovery and translation, but if harnessed effectively has the potential to provide new avenues for the development highly selective agents with improved therapeutic indices. Concluding remarks Allosteric modulators offer preferred therapeutic modalities within the context of CNS disorders, and as such have been the focus of intense research efforts by academia and industry. While several mGlu5 PAMs and NAMs have demonstrated robust efficacy in preclinical models of multiple diseases, such as schizophrenia, cognitive dysfunc- tion, anxiety, depression, Parkinson’s and Alzheimer’s diseases, clinical failures highlight key knowledge gaps that persist with respect to our understanding of the full scope of allosteric modulator pharmacology. Robust quantifica- tion, detection, and validation remain critical elements to successful allosteric modulator discovery programs. In this respect, biased agonism and modulation offer both opportunities and challenges. Unappreciated bias likely under- scoresobservationsthatnotallPAMsorNAMsarethesamewhentranslatingfrom in vitro systems to in vivo models. With an enhanced understanding of the intracellular signaling and regulatory partners engaged by mGlu5,howthese are perturbed in the disease context, and how different partners contribute to beneficial compared with adverse ef- fects,wemaybeabletorationallydesignanddevelopmoreselectiveandeffectivemGlu5 targetting therapeutics for multiple CNS disorders.

Competing interests The author declares that there are no competing interests associated with the manuscript.

Abbreviations 5PAM523, 5-fluoro-2-(61-oxadiazol-5-yl)pyridine; 7TM, seven transmembrane; Ago-PAM, agonist-PAM; ATCM, allosteric ternary complex model; ATSM, allosteric two-state model; cAMP, cyclic adenosine monophosphate; CaSR, calcium sensing receptor; CDPPB, 3-cyano-N-(1,3-diphenyl-1H-pyrazol-5-yl)benzamide; CHO, chi- nese hamster ovary; CHPG, (S)-2-chloro-5-hydroxyphenylglycine; CNS, central nervous system; CPCCOEt, 7-(hydroxyimino)cyclopropa[b]chromen-1a-carboxylate ethyl ester; DHPG, (S)-3,5-; DPFE, 1-(4-(2,4-difluorophenyl)piperazin-1-yl)-2-((4-fluorobenzyl)oxy)ethenone; ERK1/2, extracellu- lar signal-regulated kinases 1 and 2; FXS, fragile X syndrome; GABAB, γ-aminobutyric acid receptor B;

c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 2331 Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

GERD, gastroesophageal reflux disease; GPCR, G-protein-coupled receptors;2+ iCa , intracellular cal-

cium; HEK, human embryonic kidney; IP1, inositol monophosphate; LTD, long-term depression; M-5MPEP, 2-(2-(3-methoxyphenyl)ethynyl)-5-methylpyridine; mGlu, metabotropic glutamate; mGlu5, metabotropic glutamate receptor subtype 5; MPEP, 2-methyl-6-(phenylethynyl) pyridine; MTEP, 3-((2-Methyl-4-thiazolyl)ethynyl)pyridine; NAL, neutral allosteric ligand; NAM, negative allosteric modulator; NMDA, N-methyl-D-aspartate; PAM, positive al- losteric modulator; pCREB, phosphorylated cAMP response element binding protein; PD-LID, Parkinson’s disease levodopa-induced dyskinesias; VFT, Venus flytrap; VU0360172, N-cyclobutyl-6-((3-fluorophenyl)ethynyl)picolinamide; VU0409551, ((4-fluorophenyl)(2-(phenoxymethyl)-6,7-dihydrooxazolo[5,4-c]pyridin-5(4H)-yl)methanone); VU0424465, (R)-5-((3-fluorophenyl)ethynyl)-N-(3-hydroxy-3-methylbutan-2-yl)picolinamide; VU0477573, N,N-diethyl-5-((3-fluorophenyl)ethynyl)picolinamide; VU29, N-(1,3-diphenyl-1H-pyrazolo-5-yl)-4-nitrobenzamide.

References 1 Fredriksson, R., Lagerstrom, M.C., Lundin, L.G. and Schioth, H.B. (2003) The G-protein-coupled receptors in the human genome form five main families. Phylogenetic analysis, paralogon groups, and fingerprints. Mol. Pharmacol. 63, 1256–1272, https://doi.org/10.1124/mol.63.6.1256 2 Lagerstrom, M.C. and Schioth, H.B. (2008) Structural diversity of G protein-coupled receptors and significance for drug discovery. Nat. Rev. Drug Discov. 7, 339–357, https://doi.org/10.1038/nrd2518 3 Rosenbaum, D.M., Rasmussen, S.G. and Kobilka, B.K. (2009) The structure and function of G-protein-coupled receptors. Nature 459, 356–363, https://doi.org/10.1038/nature08144 4 El Moustaine, D., Granier, S., Doumazane, E., Scholler, P., Rahmeh, R., Bron, P. et al. (2012) Distinct roles of metabotropic glutamate receptor dimerization in agonist activation and G-protein coupling. Proc. Natl. Acad. Sci. U.S.A. 109, 16342–16347, https://doi.org/10.1073/pnas.1205838109 5 Kunishima, N., Shimada, Y., Tsuji, Y., Sato, T., Yamamoto, M., Kumasaka, T. et al. (2000) Structural basis of glutamate recognition by a dimeric metabotropic glutamate receptor. Nature 407, 971–977, https://doi.org/10.1038/35039564 6 Muto, T., Tsuchiya, D., Morikawa, K. and Jingami, H. (2007) Structures of the extracellular regions of the group II/III metabotropic glutamate receptors. Proc. Natl. Acad. Sci. U.S.A. 104, 3759–3764, https://doi.org/10.1073/pnas.0611577104 7 Goudet, C., Kniazeff, J., Hlavackova, V., Malhaire, F., Maurel, D., Acher, F. et al. (2005) Asymmetric functioning of dimeric metabotropic glutamate receptors disclosed by positive allosteric modulators. J. Biol. Chem. 280, 24380–24385, https://doi.org/10.1074/jbc.M502642200 8 Doumazane, E., Scholler, P., Zwier, J.M., Trinquet, E., Rondard, P. and Pin, J.P. (2011) A new approach to analyze cell surface protein complexes reveals specific heterodimeric metabotropic glutamate receptors. FASEB J. 25, 66–77, https://doi.org/10.1096/fj.10-163147 9 Kammermeier, P.J. (2012) Functional and pharmacological characteristics of metabotropic glutamate receptors 2/4 heterodimers. Mol. Pharmacol. 82, 438–447, https://doi.org/10.1124/mol.112.078501 10 Sevastyanova, T.N. and Kammermeier, P.J. (2014) Cooperative signaling between homodimers of metabotropic glutamate receptors 1 and 5. Mol. Pharmacol. 86, 492–504, https://doi.org/10.1124/mol.114.093468 11 Black, J.W. and Leff, P. (1983) Operational models of pharmacological agonism. Proc. Royal Soc. Lond. B Biol. Sci. 220, 141–162, https://doi.org/10.1098/rspb.1983.0093 12 Yin, S., Noetzel, M.J., Johnson, K.A., Zamorano, R., Jalan-Sakrikar, N., Gregory, K.J. et al. (2014) Selective actions of novel allosteric modulators reveal functional heteromers of metabotropic glutamate receptors in the CNS. J. Neurosci. 34, 79–94, https://doi.org/10.1523/JNEUROSCI.1129-13.2014 13 Ciruela, F., Escriche, M., Burgueno, J., Angulo, E., Casado, V., Soloviev, M.M. et al. (2001) Metabotropic glutamate 1alpha and adenosine A1 receptors assemble into functionally interacting complexes. J. Biol. Chem. 276, 18345–18351, https://doi.org/10.1074/jbc.M006960200 14 Gama, L., Wilt, S.G. and Breitwieser, G.E. (2001) Heterodimerization of calcium sensing receptors with metabotropic glutamate receptors in neurons. J. Biol. Chem. 276, 39053–39059, https://doi.org/10.1074/jbc.M105662200 15 Gonzalez-Maeso, J., Ang, R.L., Yuen, T., Chan, P., Weisstaub, N.V., Lopez-Gimenez, J.F. et al. (2008) Identification of a serotonin/glutamate receptor complex implicated in psychosis. Nature 452, 93–97, https://doi.org/10.1038/nature06612 16 Cabello, N., Gandia, J., Bertarelli, D.C., Watanabe, M., Lluis, C., Franco, R. et al. (2009) Metabotropic glutamate type 5, dopamine D2 and adenosine A2a receptors form higher-order oligomers in living cells. J. Neurochem. 109, 1497–1507, https://doi.org/10.1111/j.1471-4159.2009.06078.x 17 Kniazeff, J., Prezeau, L., Rondard, P., Pin, J.P. and Goudet, C. (2011) Dimers and beyond: the functional puzzles of class C GPCRs. Pharmacol. Ther. 130, 9–25, https://doi.org/10.1016/j.pharmthera.2011.01.006 18 Parnot, C. and Kobilka, B. (2004) Toward understanding GPCR dimers. Nat. Struct. Mol. Biol. 11, 691–692, https://doi.org/10.1038/nsmb0804-691 19 Rondard, P., Liu, J., Huang, S., Malhaire, F., Vol, C., Pinault, A. et al. (2006) Coupling of agonist binding to effector domain activation in metabotropic glutamate-like receptors. J. Biol. Chem. 281, 24653–24661, https://doi.org/10.1074/jbc.M602277200 20 Niswender, C.M. and Conn, P.J. (2010) Metabotropic glutamate receptors: physiology, pharmacology, and disease. Annu. Rev. Pharmacol. Toxicol. 50, 295–322, https://doi.org/10.1146/annurev.pharmtox.011008.145533 21 Bonsi, P., Platania, P., Martella, G., Madeo, G., Vita, D., Tassone, A. et al. (2008) Distinct roles of group I mGlu receptors in striatal function. Neuropharmacology 55, 392–395, https://doi.org/10.1016/j.neuropharm.2008.05.020 22 Nash, M.S., Young, K.W., Challiss, R.A. and Nahorski, S.R. (2001) Intracellular signalling. Receptor-specific messenger oscillations. Nature 413, 381–382, https://doi.org/10.1038/35096643

2332 c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

23 Nash, M.S., Schell, M.J., Atkinson, P.J., Johnston, N.R., Nahorski, S.R. and Challiss, R.A.J. (2002) Determinants of metabotropic glutamate receptor-5-mediated Ca2+ and inositol 1,4,5-trisphosphate oscillation frequency: receptor density versus agonist concentration. J. Biol. Chem. 277, 35947–35960, https://doi.org/10.1074/jbc.M205622200 24 Kawabata, S., Tsutsumi, R., Kohara, A., Yamaguchi, T., Nakanishi, S. and Okada, M. (1996) Control of calcium oscillations by phosphorylation of metabotropic glutamate receptors. Nature 383, 89–92, https://doi.org/10.1038/383089a0 25 Kim, C.H., Braud, S., Isaac, J.T.R. and Roche, K.W. (2005) Protein kinase C phosphorylation of the metabotropic glutamate receptor mGluR5 on serine 839 regulates Ca2+ oscillations. J. Biol. Chem. 280, 25409–25415, https://doi.org/10.1074/jbc.M502644200 26 Nasrallah, C., Rottier, K., Marcellin, R., Compan, V., Font, J., Llebaria, A. et al. (2018) Direct coupling of detergent purified human mGlu(5) receptor to the heterotrimeric G proteins Gq and Gs. Sci. Rep. 8, 4407, https://doi.org/10.1038/s41598-018-22729-4 27 Jong, Y-JI, Kumar, V., Kingston, A.E., Romano, C. and O’Malley, K.L. (2005) Functional metabotropic glutamate receptors on nuclei from brain and primary cultured striatal neurons: role of transporters in delivering ligand. J. Biol. Chem. 280, 30469–30480, https://doi.org/10.1074/jbc.M501775200 28 Kumar, V., Jong, Y-JI and O’Malley, K.L. (2008) Activated nuclear metabotropic glutamate receptor mGlu5 couples to nuclear Gq/11 proteins to generate inositol 1,4,5-trisphosphate-mediated nuclear Ca2+ release. J. Biol. Chem. 283, 14072–14083, https://doi.org/10.1074/jbc.M708551200 29 Jong, Y.J., Kumar, V. and O’Malley, K.L. (2009) Intracellular metabotropic glutamate receptor 5 (mGluR5) activates signaling cascades distinct from cell surface counterparts. J. Biol. Chem. 284, 35827–35838, https://doi.org/10.1074/jbc.M109.046276 30 Kumar, V., Fahey, P.G., Jong, Y.J., Ramanan, N. and O’Malley, K.L. (2012) Activation of intracellular metabotropic glutamate receptor 5 in striatal neurons leads to up-regulation of genes associated with sustained synaptic transmission including Arc/Arg3.1 protein. J. Biol. Chem. 287, 5412–5425, https://doi.org/10.1074/jbc.M111.301366 31 Purgert, C.A., Izumi, Y., Jong, Y.J., Kumar, V., Zorumski, C.F. and Malley, K.L. (2014) Intracellular mGluR5 can mediate synaptic plasticity in the hippocampus. J. Neurosci. 34, 4589, https://doi.org/10.1523/JNEUROSCI.3451-13.2014 32 Shigemoto, R., Nomura, S., Ohishi, H., Sugihara, H., Nakanishi, S. and Mizuno, N. (1993) Immunohistochemical localization of a metabotropic glutamate receptor, mGluR5, in the rat brain. Neurosci. Lett. 163, 53–57, https://doi.org/10.1016/0304-3940(93)90227-C 33 Biber, K., Laurie, D.J., Berthele, A., Sommer, B., Tolle, T.R., Gebicke-Harter, P.J. et al. (1999) Expression and signaling of group I metabotropic glutamate receptors in astrocytes and microglia. J. Neurochem. 72, 1671–1680, https://doi.org/10.1046/j.1471-4159.1999.721671.x 34 Schools, G.P. and Kimelberg, H.K. (1999) mGluR3 and mGluR5 are the predominant metabotropic glutamate receptor mRNAs expressed in hippocampal astrocytes acutely isolated from young rats. J. Neurosci. Res. 58, 533–543, https://doi.org/10.1002/(SICI)1097-4547(19991115)58:4%3c533::AID-JNR6%3e3.0.CO;2-G 35 Crawford, J.H., Wainwright, A., Heavens, R., Pollock, J., Martin, D.J., Scott, R.H. et al. (2000) Mobilisation of intracellular Ca2+ by mGluR5 metabotropic glutamate receptor activation in neonatal rat cultured dorsal root ganglia neurones. Neuropharmacology 39, 621–630, https://doi.org/10.1016/S0028-3908(99)00167-7 36 Ferraguti, F. and Shigemoto, R. (2006) Metabotropic glutamate receptors. Cell Tissue Res. 326, 483–504, https://doi.org/10.1007/s00441-006-0266-5 37 Conn, P.J. and Pin, J.-P. (1997) Pharmacology and functions of metabotropic glutamate receptors. Annu. Rev. Pharmacol. Toxicol. 37, 205–237, https://doi.org/10.1146/annurev.pharmtox.37.1.205 38 Ballester-Rosado, C.J., Albright, M.J., Wu, C.S., Liao, C.C., Zhu, J., Xu, J. et al. (2010) mGluR5 in cortical excitatory neurons exerts both cell-autonomous and -nonautonomous influences on cortical somatosensory circuit formation. J. Neurosci. 30, 16896–16909, https://doi.org/10.1523/JNEUROSCI.2462-10.2010 39 Bortolotto, Z.A., Collett, V.J., Conquet, F., Jia, Z., van der Putten, H. and Collingridge, G.L. (2005) The regulation of hippocampal LTP by the molecular switch, a form of metaplasticity, requires mGlu5 receptors. Neuropharmacology 49, 13–25, https://doi.org/10.1016/j.neuropharm.2005.05.020 40 Jia, Z., Lu, Y., Henderson, J., Taverna, F., Romano, C., Abramow-Newerly, W. et al. (1998) Selective abolition of the NMDA component of long-term potentiation in mice lacking mGluR5. Learn Mem. 5, 331–343 41 Xu, J., Zhu, Y., Contractor, A. and Heinemann, S.F. (2009) mGluR5 has a critical role in inhibitory learning. J. Neurosci. 29, 3676–3684, https://doi.org/10.1523/JNEUROSCI.5716-08.2009 42 Galik, J., Youn, D.H., Kolaj, M. and Randic, M. (2008) Involvement of group I metabotropic glutamate receptors and glutamate transporters in the slow excitatory synaptic transmission in the spinal cord dorsal horn. Neuroscience 154, 1372–1387, https://doi.org/10.1016/j.neuroscience.2008.04.059 43 Wijetunge, L.S., Till, S.M., Gillingwater, T.H., Ingham, C.A. and Kind, P.C. (2008) mGluR5 regulates glutamate-dependent development of the mouse somatosensory cortex. J. Neurosci. 28, 13028–13037, https://doi.org/10.1523/JNEUROSCI.2600-08.2008 44 Hamilton, A., Esseltine, J.L., DeVries, R.A., Cregan, S.P. and Ferguson, S.S. (2014) Metabotropic glutamate receptor 5 knockout reduces cognitive impairment and pathogenesis in a mouse model of Alzheimer’s disease. Mol. Brain 7, 40, https://doi.org/10.1186/1756-6606-7-40 45 She, W.C., Quairiaux, C., Albright, M.J., Wang, Y.C., Sanchez, D.E., Chang, P.S. et al. (2009) Roles of mGluR5 in synaptic function and plasticity ofthe mouse thalamocortical pathway. Eur. J. Neurosci. 29, 1379–1396, https://doi.org/10.1111/j.1460-9568.2009.06696.x 46 Black, Y.D., Xiao, D., Pellegrino, D., Kachroo, A., Brownell, A.L. and Schwarzschild, M.A. (2010) Protective effect of metabotropic glutamate mGluR5 receptor elimination in a 6-hydroxydopamine model of Parkinson’s disease. Neurosci. Lett. 486, 161–165, https://doi.org/10.1016/j.neulet.2010.09.043 47 Chen, C.C., Lu, H.C. and Brumberg, J.C. (2012) mGluR5 knockout mice display increased dendritic spine densities. Neurosci. Lett. 524, 65–68, https://doi.org/10.1016/j.neulet.2012.07.014 48 Julio-Pieper, M., Flor, P.J., Dinan, T.G. and Cryan, J.F. (2011) Exciting times beyond the brain: metabotropic glutamate receptors in peripheral and non-neural tissues. Pharmacol. Rev. 63, 35–58, https://doi.org/10.1124/pr.110.004036

c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 2333 Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

49 Nickols, H.H. and Conn, P.J. (2014) Development of allosteric modulators of GPCRs for treatment of CNS disorders. Neurobiol. Dis. 61, 55–71, https://doi.org/10.1016/j.nbd.2013.09.013 50 Bruno, V., Caraci, F., Copani, A., Matrisciano, F., Nicoletti, F. and Battaglia, G. (2017) The impact of metabotropic glutamate receptors into active neurodegenerative processes: a ‘dark side’ in the development of new symptomatic treatments for neurologic and psychiatric disorders. Neuropharmacology 115, 180–192, https://doi.org/10.1016/j.neuropharm.2016.04.044 51 Nicoletti, F., Bockaert, J., Collingridge, G.L., Conn, P.J., Ferraguti, F., Schoepp, D.D. et al. (2011) Metabotropic glutamate receptors: from the workbench to the bedside. Neuropharmacology 60, 1017–1041, https://doi.org/10.1016/j.neuropharm.2010.10.022 52 Pecknold, J.C., McClure, D.J., Appeltauer, L., Wrzesinski, L. and Allan, T. (1982) Treatment of anxiety using fenobam (a nonbenzodiazepine) in a double-blind standard (diazepam) placebo-controlled study. J. Clin. Psychopharmacol. 2, 129–133, https://doi.org/10.1097/00004714-198204000-00010 53 Porter, R.H., Jaeschke, G., Spooren, W., Ballard, T.M., Buttelmann, B., Kolczewski, S. et al. (2005) Fenobam: a clinically validated nonbenzodiazepine anxiolytic is a potent, selective, and noncompetitive mGlu5 receptor antagonist with inverse agonist activity. J. Pharmacol. Exp. Ther. 315, 711–721, https://doi.org/10.1124/jpet.105.089839 54 Berry-Kravis, E., Hessl, D., Coffey, S., Hervey, C., Schneider, A., Yuhas, J. et al. (2009) A pilot open label, single dose trial of fenobam in adults with fragile X syndrome. J. Med. Genet. 46, 266–271, https://doi.org/10.1136/jmg.2008.063701 55 Lindemann, L., Porter, R.H., Scharf, S.H., Kuennecke, B., Bruns, A., von Kienlin, M. et al. (2015) Pharmacology of basimglurant (RO4917523, RG7090), a unique metabotropic glutamate receptor 5 negative allosteric modulator in clinical development for depression. J. Pharmacol. Exp. Ther. 353, 213–233, https://doi.org/10.1124/jpet.114.222463 56 Youssef, E.A., Berry-Kravis, E., Czech, C., Hagerman, R.J., Hessl, D., Wong, C.Y. et al. (2018) Effect of the mGluR5-NAM Basimglurant on behavior in adolescents and adults with fragile X syndrome in a randomized, double-Blind, placebo-controlled trial: FragXis Phase 2 results. Neuropsychopharmacology 43, 503–512, https://doi.org/10.1038/npp.2017.177 57 Jaeschke, G., Kolczewski, S., Spooren, W., Vieira, E., Bitter-Stoll, N., Boissin, P. et al. (2015) Metabotropic glutamate receptor 5 negative allosteric modulators: discovery of 2-chloro-4-[1-(4-fluorophenyl)-2,5-dimethyl-1H-imidazol-4-ylethynyl]pyridine (basimglurant, RO4917523), a promising novel medicine for psychiatric diseases. J. Med. Chem. 58, 1358–1371, https://doi.org/10.1021/jm501642c 58 Quiroz, J.A., Tamburri, P., Deptula, D., Banken, L., Beyer, U., Rabbia, M. et al. (2016) Efficacy and safety of Basimglurant as adjunctive therapy for major depression: a randomized clinical trial. JAMA Psychiatry 73, 675–684, https://doi.org/10.1001/jamapsychiatry.2016.0838 59 Kumar, R., Hauser, R.A., Mostillo, J., Dronamraju, N., Graf, A., Merschhemke, M. et al. (2016) Mavoglurant (AFQ056) in combination with increased levodopa dosages in Parkinson’s disease patients. Int. J. Neurosci. 126, 20–24, https://doi.org/10.3109/00207454.2013.841685 60 Stocchi, F., Rascol, O., Destee, A., Hattori, N., Hauser, R.A., Lang, A.E. et al. (2013) AFQ056 in Parkinson patients with levodopa-induced dyskinesia: 13-week, randomized, dose-finding study. Mov. Disord. 28, 1838–1846, https://doi.org/10.1002/mds.25561 61 Reilmann, R., Rouzade-Dominguez, M.L., Saft, C., Sussmuth, S.D., Priller, J., Rosser, A. et al. (2015) A randomized, placebo-controlled trial of AFQ056 for the treatment of chorea in Huntington’s disease. Mov. Disord. 30, 427–431, https://doi.org/10.1002/mds.26174 62 Rouzade-Dominguez, M.L., Pezous, N., David, O.J., Tutuian, R., Bruley des Varannes, S., Tack, J. et al. (2017) The selective metabotropic glutamate receptor 5 antagonist mavoglurant (AFQ056) reduces the incidence of reflux episodes in dogs and patients with moderate to severe gastroesophageal reflux disease. Neurogastroenterol. Motil. 29, https://doi.org/10.1111/nmo.13058 63 Berry-Kravis, E., Des Portes, V., Hagerman, R., Jacquemont, S., Charles, P., Visootsak, J. et al. (2016) Mavoglurant in fragile X syndrome: resultsof two randomized, double-blind, placebo-controlled trials. Sci. Transl. Med. 8, 321ra5, https://doi.org/10.1126/scitranslmed.aab4109 64 Keywood, C., Wakefield, M. and Tack, J. (2009) A proof-of-concept study evaluating the effect of ADX10059, a metabotropic glutamate receptor-5 negative allosteric modulator, on acid exposure and symptoms in gastro-oesophageal reflux disease. Gut 58, 1192–1199, https://doi.org/10.1136/gut.2008.162040 65 Zerbib, F., Keywood, C. and Strabach, G. (2010) Efficacy, tolerability and pharmacokinetics of a modified release formulation of ADX10059, a negative allosteric modulator of metabotropic glutamate receptor 5: an esophageal pH-impedance study in healthy subjects. Neurogastroenterol. Motil. 22, 859–865, e231, https://doi.org/10.1111/j.1365-2982.2010.01484.x 66 Zerbib, F., Bruley des Varannes, S., Roman, S., Tutuian, R., Galmiche, J.P., Mion, F. et al. (2011) Randomised clinical trial: effects of monotherapy with ADX10059, a mGluR5 inhibitor, on symptoms and reflux events in patients with gastro-oesophageal reflux disease. Aliment. Pharmacol. Ther. 33, 911–921, https://doi.org/10.1111/j.1365-2036.2011.04596.x 67 Tison, F., Keywood, C., Wakefield, M., Durif, F., Corvol, J.C., Eggert, K. et al. (2016) A phase 2A trial of the novel mGluR5 negative allosteric modulator Dipraglurant for levodopa-induced dyskinesia in Parkinson’s disease. Mov. Disord. 31, 1373–1380, https://doi.org/10.1002/mds.26659 68 Hauser, A.S., Attwood, M.M., Rask-Andersen, M., Schioth, H.B. and Gloriam, D.E. (2017) Trends in GPCR drug discovery: new agents, targets and indications. Nat. Rev. Drug Discov. 16, 829–842, https://doi.org/10.1038/nrd.2017.178 69 Porter, R.H., Roberts, P.J., Jane, D.E. and Watkins, J.C. (1992) (S)-homoquisqualate: a potent agonist at the glutamate metabotropic receptor. Br. J. Pharmacol. 106, 509–510, https://doi.org/10.1111/j.1476-5381.1992.tb14366.x 70 Mutel, V., Ellis, G.J., Adam, G., Chaboz, S., Nilly, A., Messer, J. et al. (2000) Characterization of [(3)H]Quisqualate binding to recombinant rat metabotropic glutamate 1a and 5a receptors and to rat and human brain sections. J. Neurochem. 75, 2590–2601, https://doi.org/10.1046/j.1471-4159.2000.0752590.x 71 Wisniewski, K. and Car, H. (2002) (S)-3,5-DHPG: a review. CNS Drug Rev. 8, 101–116, https://doi.org/10.1111/j.1527-3458.2002.tb00218.x 72 Hathaway, H.A., Pshenichkin, S., Grajkowska, E., Gelb, T., Emery, A.C., Wolfe, B.B. et al. (2015) Pharmacological characterization of mGlu1 receptors in cerebellar granule cells reveals biased agonism. Neuropharmacology 93, 199–208, https://doi.org/10.1016/j.neuropharm.2015.02.007

2334 c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

73 Doherty, A.J., Palmer, M.J., Henley, J.M., Collingridge, G.L. and Jane, D.E. (1997) (RS)-2-chloro-5-hydroxyphenylglycine (CHPG) activates mGlu5, but no mGlu1, receptors expressed in CHO cells and potentiates NMDA responses in the hippocampus. Neuropharmacology 36, 265–267, https://doi.org/10.1016/S0028-3908(97)00001-4 74 Doherty, A.J., Collingridge, G.L. and Jane, D.E. (1999) Antagonist activity of alpha-substituted 4-carboxyphenylglycine analogues at group I metabotropic glutamate receptors expressed in CHO cells. Br.J.Pharmacol.126, 205–210, https://doi.org/10.1038/sj.bjp.0702297 75 Kingston, A.E., Burnett, J.P., Mayne, N.G. and Lodge, D. (1995) Pharmacological analysis of 4-carboxyphenylglycine derivatives: comparison of effects on mGluR1 alpha and mGluR5a subtypes. Neuropharmacology 34, 887–894, https://doi.org/10.1016/0028-3908(95)00069-I 76 Gregory, K.J. and Conn, P.J. (2015) Molecular insights into metabotropic glutamate receptor allosteric modulation. Mol. Pharmacol. 88, 188–202, https://doi.org/10.1124/mol.114.097220 77 Dore, A.S., Okrasa, K., Patel, J.C., Serrano-Vega, M., Bennett, K., Cooke, R.M. et al. (2014) Structure of class C GPCR metabotropic glutamate receptor 5 transmembrane domain. Nature 511, 557–562, https://doi.org/10.1038/nature13396 78 Wu, H., Wang, C., Gregory, K.J., Han, G.W., Cho, H.P., Xia, Y. et al. (2014) Structure of a class C GPCR metabotropic glutamate receptor 1 bound to an allosteric modulator. Science 344, 58–64, https://doi.org/10.1126/science.1249489 79 Christopher, J.A., Aves, S.J., Bennett, K.A., Dore, A.S., Errey, J.C., Jazayeri, A. et al. (2015) Fragment and structure-based drug discovery for aclassC GPCR: discovery of the mGlu5 negative allosteric modulator HTL14242 (3-Chloro-5-[6-(5-fluoropyridin-2-yl)pyrimidin-4-yl]benzonitrile). J. Med. Chem. 58, 6653–6664, https://doi.org/10.1021/acs.jmedchem.5b00892 80 Lindberg, J.S., Culleton, B., Wong, G., Borah, M.F., Clark, R.V., Shapiro, W.B. et al. (2005) Cinacalcet HCl, an oral calcimimetic agent for the treatment of secondary hyperparathyroidism in hemodialysis and peritoneal dialysis: a randomized, double-blind, multicenter study. J. Am. Soc. Nephrol. 16, 800–807, https://doi.org/10.1681/ASN.2004060512 81 Fukagawa, M., Shimazaki, R. and Akizawa, T. (2018) Head-to-head comparison of the new calcimimetic agent evocalcet with cinacalcet in Japanese hemodialysis patients with secondary hyperparathyroidism. Kidney Int. 94, 818–825, https://doi.org/10.1016/j.kint.2018.05.013 82 Fukagawa, M., Yokoyama, K., Shigematsu, T., Akiba, T., Fujii, A., Kuramoto, T. et al. (2017) A phase 3, multicentre, randomized, double-blind, placebo-controlled, parallel-group study to evaluate the efficacy and safety of etelcalcetide (ONO-5163/AMG 416), a novel intravenous calcimimetic, for secondary hyperparathyroidism in Japanese haemodialysis patients. Nephrol. Dial. Transplant. 32, 1723–1730 83 Dorr, P., Westby, M., Dobbs, S., Griffin, P., Irvine, B., Macartney, M. et al. (2005) Maraviroc (UK-427,857), a potent, orally bioavailable, and selective small-molecule inhibitor of chemokine receptor CCR5 with broad-spectrum anti-human immunodeficiency virus type 1 activity. Antimicrob. Agents Chemother. 49, 4721–4732, https://doi.org/10.1128/AAC.49.11.4721-4732.2005 84 Canals, M., Sexton, P.M. and Christopoulos, A. (2011) Allostery in GPCRs: ‘MWC’ revisited. Trends Biochem. Sci. 36, 663–672, https://doi.org/10.1016/j.tibs.2011.08.005 85 Wootten, D., Christopoulos, A. and Sexton, P.M. (2013) Emerging paradigms in GPCR allostery: implications for drug discovery. Nat. Rev. Drug Discov. 12, 630–644, https://doi.org/10.1038/nrd4052 86 Christopoulos, A. (2014) Advances in G protein-coupled receptor allostery: from function to structure. Mol. Pharmacol. 86, 463–478, https://doi.org/10.1124/mol.114.094342 87 Price, M.R., Baillie, G.L., Thomas, A., Stevenson, L.A., Easson, M., Goodwin, R. et al. (2005) Allosteric modulation of the cannabinoid CB1 receptor. Mol. Pharmacol. 68, 1484–1495, https://doi.org/10.1124/mol.105.016162 88 Hellyer, S.D., Albold, S., Wang, T., Chen, A.N.Y., May, L.T., Leach, K. et al. (2018) ‘Selective’ class C G Protein-coupled receptor modulators are neutral or biased mGlu5 allosteric ligands. Mol. Pharmacol. 93, 504–514, https://doi.org/10.1124/mol.117.111518 89 May, L.T., Leach, K., Sexton, P.M. and Christopoulos, A. (2007) Allosteric modulation of G protein-coupled receptors. Annu. Rev. Pharmacol. Toxicol. 47, 1–51, https://doi.org/10.1146/annurev.pharmtox.47.120505.105159 90 Klein, M.T., Vinson, P.N. and Niswender, C.M. (2013) Approaches for probing allosteric interactions at 7 transmembrane spanning receptors. Prog. Mol. Biol. Transl. Sci. 115, 1–59, https://doi.org/10.1016/B978-0-12-394587-7.00001-4 91 Sengmany, K., Singh, J., Stewart, G.D., Conn, P.J., Christopoulos, A. and Gregory, K.J. (2017) Biased allosteric agonism and modulation of metabotropic glutamate receptor 5: implications for optimizing preclinical neuroscience drug discovery. Neuropharmacology 115, 60–72, https://doi.org/10.1016/j.neuropharm.2016.07.001 92 Valant, C., Felder, C.C., Sexton, P.M. and Christopoulos, A. (2012) Probe dependence in the allosteric modulation of a G protein-coupled receptor: implications for detection and validation of allosteric ligand effects. Mol. Pharmacol. 81, 41–52, https://doi.org/10.1124/mol.111.074872 93 Koole, C., Wootten, D., Simms, J., Valant, C., Sridhar, R., Woodman, O.L. et al. (2010) Allosteric ligands of the glucagon-like peptide 1 receptor (GLP-1R) differentially modulate endogenous and exogenous peptide responses in a pathway-selective manner: implications for drug screening. Mol. Pharmacol. 78, 456–465, https://doi.org/10.1124/mol.110.065664 94 Christopoulos, A. and Kenakin, T. (2002) G protein-coupled receptor allosterism and complexing. Pharmacol. Rev. 54, 323–374, https://doi.org/10.1124/pr.54.2.323 95 Ehlert, F.J. (1988) Estimation of the affinities of allosteric ligands using radioligand binding and pharmacological null methods. Mol. Pharmacol. 33, 187 96 Hall, D.A. (2000) Modeling the functional effects of allosteric modulators at pharmacological receptors: an extension of the two-state model of receptor activation. Mol. Pharmacol. 58, 1412–1423, https://doi.org/10.1124/mol.58.6.1412 97 Leach, K., Sexton, P.M. and Christopoulos, A. (2007) Allosteric GPCR modulators: taking advantage of permissive receptor pharmacology. Trends Pharmacol. Sci. 28, 382–389, https://doi.org/10.1016/j.tips.2007.06.004 98 Keov, P., Sexton, P.M. and Christopoulos, A. (2011) Allosteric modulation of G protein-coupled receptors: a pharmacological perspective. Neuropharmacology 60, 24–35, https://doi.org/10.1016/j.neuropharm.2010.07.010

c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 2335 Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

99 Gregory, K.J., Noetzel, M.J., Rook, J.M., Vinson, P.N., Stauffer, S.R., Rodriguez, A.L. et al. (2012) Investigating metabotropic glutamate receptor 5 allosteric modulator cooperativity, affinity, and agonism: enriching structure-function studies and structure-activity relationships. Mol. Pharmacol. 82, 860–875, https://doi.org/10.1124/mol.112.080531 100 Litschig, S., Gasparini, F., Rueegg, D., Stoehr, N., Flor, P.J., Vranesic, I. et al. (1999) CPCCOEt, a noncompetitive metabotropic glutamate receptor 1 antagonist, inhibits receptor signaling without affecting glutamate binding. Mol. Pharmacol. 55, 453–461 101 Emmitte, K.A. (2013) mGlu5 negative allosteric modulators: a patent review (2010-2012). Expert Opin. Ther. Pat. 23, 393–408, https://doi.org/10.1517/13543776.2013.760544 102 Rocher, J.P., Bonnet, B., Bolea, C., Lutjens, R., Le Poul, E., Poli, S. et al. (2011) mGluR5 negative allosteric modulators overview: a medicinal chemistry approach towards a series of novel therapeutic agents. Curr. Top. Med. Chem. 11, 680–695, https://doi.org/10.2174/1568026611109060680 103 Felts, A.S., Rodriguez, A.L., Morrison, R.D., Venable, D.F., Manka, J.T., Bates, B.S. et al. (2013) Discovery of VU0409106: a negative allosteric modulator of mGlu5 with activity in a mouse model of anxiety. Bioorg. Med. Chem. Lett. 23, 5779–5785, https://doi.org/10.1016/j.bmcl.2013.09.001 104 Levenga, J., Hayashi, S., de Vrij, F.M., Koekkoek, S.K., van der Linde, H.C., Nieuwenhuizen, I. et al. (2011) AFQ056, a new mGluR5 antagonist for treatment of fragile X syndrome. Neurobiol. Dis. 42, 311–317, https://doi.org/10.1016/j.nbd.2011.01.022 105 Pop, A.S., Gomez-Mancilla, B., Neri, G., Willemsen, R. and Gasparini, F. (2014) Fragile X syndrome: a preclinical review on metabotropic glutamate receptor 5 (mGluR5) antagonists and drug development. Psychopharmacology (Berl.) 231, 1217–1226, https://doi.org/10.1007/s00213-013-3330-3 106 Rascol, O., Fox, S., Gasparini, F., Kenney, C., Di Paolo, T. and Gomez-Mancilla, B. (2014) Use of metabotropic glutamate 5-receptor antagonists for treatment of levodopa-induced dyskinesias. Relat. Disord. 20, 947–956, https://doi.org/10.1016/j.parkreldis.2014.05.003 107 Picconi, B. and Calabresi, P. (2014) Targeting metabotropic glutamate receptors as a new strategy against levodopa-induced dyskinesia in Parkinson’s disease? Mov. Disord. 29, 715–719, https://doi.org/10.1002/mds.25851 108 Hamilton, A., Vasefi, M., Vander Tuin, C., McQuaid, R.J., Anisman, H. and Ferguson, S.S. (2016) Chronic pharmacological mGluR5 inhibition prevents cognitive impairment and reduces pathogenesis in an Alzheimer’s disease mouse model. Cell Rep. 15, 1859–1865, https://doi.org/10.1016/j.celrep.2016.04.077 109 Campbell, U.C., Lalwani, K., Hernandez, L., Kinney, G.G., Conn, P.J. and Bristow, L.J. (2004) The mGluR5 antagonist 2-methyl-6-(phenylethynyl)-pyridine (MPEP) potentiates PCP-induced cognitive deficits in rats. Psychopharmacology (Berl.) 175, 310–318, https://doi.org/10.1007/s00213-004-1827-5 110 Rodriguez, A.L., Grier, M.D., Jones, C.K., Herman, E.J., Kane, A.S., Smith, R.L. et al. (2010) Discovery of novel allosteric modulators of metabotropic glutamate receptor subtype 5 reveals chemical and functional diversity and in vivo activity in rat behavioral models of anxiolytic and antipsychotic activity. Mol. Pharmacol. 78, 1105–1123, https://doi.org/10.1124/mol.110.067207 111 Abou Farha, K., Bruggeman, R. and Balje-Volkers, C. (2014) Metabotropic glutamate receptor 5 negative modulation in phase I clinical trial: potential impact of circadian rhythm on the neuropsychiatric adverse reactions-do hallucinations matter? ISRN Psychiatry 2014, 652750, https://doi.org/10.1155/2014/652750 112 Gould, R.W., Amato, R.J., Bubser, M., Joffe, M.E., Nedelcovych, M.T., Thompson, A.D. et al. (2016) Partial mGlu(5) negative allosteric modulators attenuate cocaine-mediated behaviors and lack psychotomimetic-like effects. Neuropsychopharmacology 41, 1166–1178, https://doi.org/10.1038/npp.2015.265 113 Nickols, H.H., Yuh, J.P., Gregory, K.J., Morrison, R.D., Bates, B.S., Stauffer, S.R. et al. (2016) VU0477573: partial negative allosteric modulator of the subtype 5 metabotropic glutamate receptor with in vivo efficacy. J. Pharmacol. Exp. Ther. 356, 123–136, https://doi.org/10.1124/jpet.115.226597 114 Kinney, G.G., O’Brien, J.A., Lemaire, W., Burno, M., Bickel, D.J., Clements, M.K. et al. (2005) A novel selective positive allosteric modulator of metabotropic glutamate receptor subtype 5 has in vivo activity and antipsychotic-like effects in rat behavioral models. J. Pharmacol. Exp. Ther. 313, 199–206, https://doi.org/10.1124/jpet.104.079244 115 Liu, F., Grauer, S., Kelley, C., Navarra, R., Graf, R., Zhang, G. et al. (2008) ADX47273 [S-(4-fluoro-phenyl)-{3-[3-(4-fluoro-phenyl)-[1,2,4]-oxadiazol-5-yl]-piperidin-1- yl}-methanone]: a novel metabotropic glutamate receptor 5-selective positive allosteric modulator with preclinical antipsychotic-like and procognitive activities. J. Pharmacol. Exp. Ther. 327, 827–839, https://doi.org/10.1124/jpet.108.136580 116 Reichel, C.M., Schwendt, M., McGinty, J.F., Olive, M.F. and See, R.E. (2011) Loss of object recognition memory produced by extended access to methamphetamine self-administration is reversed by positive allosteric modulation of metabotropic glutamate receptor 5. Neuropsychopharmacology 36, 782–792, https://doi.org/10.1038/npp.2010.212 117 Gregory, K.J., Herman, E.J., Ramsey, A.J., Hammond, A.S., Byun, N.E., Stauffer, S.R. et al. (2013) N-aryl piperazine metabotropic glutamate receptor 5 positive allosteric modulators possess efficacy in preclinical models of NMDA hypofunction and cognitive enhancement. J. Pharmacol. Exp. Ther. 347, 438–457, https://doi.org/10.1124/jpet.113.206623 118 LaCrosse, A.L., Taylor, S.B., Nemirovsky, N.E., Gass, J.T. and Olive, M.F. (2015) mGluR5 positive and negative allosteric modulators differentially affect dendritic spine density and morphology in the prefrontal cortex. CNS Neurol. Disord. Drug Targets 14, 476–485, https://doi.org/10.2174/1871527314666150429112849 119 Rook, J.M., Noetzel, M.J., Pouliot, W.A., Bridges, T.M., Vinson, P.N., Cho, H.P. et al. (2013) Unique signaling profiles of positive allosteric modulators of metabotropic glutamate receptor subtype 5 determine differences in in vivo activity. Biol. Psychiatry 73, 501–509, https://doi.org/10.1016/j.biopsych.2012.09.012 120 Parmentier-Batteur, S., Hutson, P.H., Menzel, K., Uslaner, J.M., Mattson, B.A., O’Brien, J.A. et al. (2014) Mechanism based neurotoxicity of mGlu5 positive allosteric modulators–development challenges for a promising novel antipsychotic target. Neuropharmacology 82, 161–173, https://doi.org/10.1016/j.neuropharm.2012.12.003 121 Rook, J.M., Xiang, Z., Lv, X., Ghoshal, A., Dickerson, J.W., Bridges, T.M. et al. (2015) Biased mGlu5 positive allosteric modulators provide in vivo efficacy without potentiating mGlu5 modulation of NMDAR currents. Neuron 86, 1029–1040, https://doi.org/10.1016/j.neuron.2015.03.063

2336 c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

122 Kenakin, T. (2011) Functional selectivity and biased receptor signaling. J. Pharmacol. Exp. Ther. 336, 296–302, https://doi.org/10.1124/jpet.110.173948 123 Kenakin, T. and Christopoulos, A. (2013) Signalling bias in new drug discovery: detection, quantification and therapeutic impact. Nat. Rev. Drug Discov. 12, 205–216, https://doi.org/10.1038/nrd3954 124 Lawler, C.P., Prioleau, C., Lewis, M.M., Mak, C., Jiang, D., Schetz, J.A. et al. (1999) Interactions of the novel antipsychotic aripiprazole (OPC-14597) with dopamine and serotonin receptor subtypes. Neuropsychopharmacology 20, 612–627, https://doi.org/10.1016/S0893-133X(98)00099-2 125 Kenakin, T. (1995) Agonist-receptor efficacy. II. Agonist trafficking of receptor signals. Trends in Pharmacol. Sci. 16, 232–238, https://doi.org/10.1016/S0165-6147(00)89032-X 126 Michel, M.C. and Alewijnse, A.E. (2007) Ligand-directed signaling: 50 ways to find a lover. Mol. Pharmacol. 72, 1097–1099, https://doi.org/10.1124/mol.107.040923 127 Kenakin, T. (2015) The effective application of biased signaling to new drug discovery. Mol. Pharmacol. 88, 1055, https://doi.org/10.1124/mol.115.099770 128 Wootten, D., Christopoulos, A., Marti-Solano, M., Babu, M.M. and Sexton, P.M. (2018) Mechanisms of signalling and biased agonism in G protein-coupled receptors. Nat. Rev. Mol. Cell Biol. 19, 638–653, https://doi.org/10.1038/s41580-018-0049-3 129 Liu, F., Zhou, R., Yan, H., Yin, H., Wu, X., Tan, Y. et al. (2014) Metabotropic glutamate receptor 5 modulates calcium oscillation and innate immune response induced by lipopolysaccharide in microglial cell. Neuroscience 281, 24–34, https://doi.org/10.1016/j.neuroscience.2014.09.034 130 Haas, L.T. and Strittmatter, S.M. (2016) Oligomers of amyloid beta prevent physiological activation of the cellular prion protein-metabotropic glutamate receptor 5 complex by glutamate in Alzheimer disease. J. Biol. Chem. 291, 17112–17121, https://doi.org/10.1074/jbc.M116.720664 131 Haas, L.T., Salazar, S.V., Kostylev, M.A., Um, J.W., Kaufman, A.C. and Strittmatter, S.M. (2016) Metabotropic glutamate receptor 5 couples cellular prion protein to intracellular signalling in Alzheimer’s disease. Brain 139, 526–546, https://doi.org/10.1093/brain/awv356 132 Makita, N., Sato, J., Manaka, K., Shoji, Y., Oishi, A., Hashimoto, M. et al. (2007) An acquired hypocalciuric hypercalcemia autoantibody induces allosteric transition among active human Ca-sensing receptor conformations. Proc. Natl. Acad. Sci. U.S.A. 104, 5443–5448, https://doi.org/10.1073/pnas.0701290104 133 Kenakin, T., Watson, C., Muniz-Medina, V., Christopoulos, A. and Novick, S. (2012) A simple method for quantifying functional selectivity and agonist bias. ACS Chem. Neurosci. 3, 193–203, https://doi.org/10.1021/cn200111m 134 Emery, A.C., DiRaddo, J.O., Miller, E., Hathaway, H.A., Pshenichkin, S., Takoudjou, G.R. et al. (2012) Ligand bias at metabotropic glutamate 1a receptors: molecular determinants that distinguish β-Arrestin-mediated from G Protein-mediated signaling. Mol. Pharmacol. 82, 291–301, https://doi.org/10.1124/mol.112.078444 135 Jalan-Sakrikar, N., Field, J.R., Klar, R., Mattmann, M.E., Gregory, K.J., Zamorano, R. et al. (2014) Identification of positive allosteric modulators VU0155094 (ML397) and VU0422288 (ML396) reveals new insights into the biology of metabotropic glutamate receptor 7. ACS Chem. Neurosci. 5, 1221–1237, https://doi.org/10.1021/cn500153z 136 Noetzel, M.J., Rook, J.M., Vinson, P.N., Cho, H.P., Days, E., Zhou, Y. et al. (2012) Functional impact of allosteric agonist activity of selective positive allosteric modulators of metabotropic glutamate receptor subtype 5 in regulating central nervous system function. Mol. Pharmacol. 81, 120–133, https://doi.org/10.1124/mol.111.075184 137 Zhang, Y., Rodriguez, A.L. and Conn, P.J. (2005) Allosteric potentiators of metabotropic glutamate receptor subtype 5 have differential effectson different signaling pathways in cortical astrocytes. J. Pharmacol. Exp. Ther. 315, 1212–1219, https://doi.org/10.1124/jpet.105.090308 138 Balu, D.T., Li, Y., Takagi, S., Presti, K.T., Ramikie, T.S., Rook, J.M. et al. (2016) An mGlu5 positive allosteric modulator rescues the neuroplasticity deficits in a genetic model of NMDA receptor hypofunction in schizophrenia. Neuropsychopharmacology 41, 2052, https://doi.org/10.1038/npp.2016.2 139 Bradley, S.J., Langmead, C.J., Watson, J.M. and Challiss, R.A. (2011) Quantitative analysis reveals multiple mechanisms of allosteric modulation of the mGlu5 receptor in rat astroglia. Mol. Pharmacol. 79, 874–885, https://doi.org/10.1124/mol.110.068882 140 Chen, Y., Nong, Y., Goudet, C., Hemstapat, K., de Paulis, T., Pin, J.P. et al. (2007) Interaction of novel positive allosteric modulators of metabotropic glutamate receptor 5 with the negative allosteric antagonist site is required for potentiation of receptor responses. Mol. Pharmacol. 71, 1389–1398, https://doi.org/10.1124/mol.106.032425 141 Doria, J.G., de Souza, J.M., Andrade, J.N., Rodrigues, H.A., Guimaraes, I.M., Carvalho, T.G. et al. (2015) The mGluR5 positive allosteric modulator, CDPPB, ameliorates pathology and phenotypic signs of a mouse model of Huntington’s disease. Neurobiol. Dis. 73, 163–173, https://doi.org/10.1016/j.nbd.2014.08.021 142 Doria, J.G., Silva, F.R., Souza, J.M., Vieira, L.B., Carvalho, T.G., Reis, H.J. et al. (2013) Metabotropic glutamate receptor 5 positive allosteric modulators are neuroprotective in a mouse model of Huntington’s disease. Br. J. Pharmacol. 169, 909–921, https://doi.org/10.1111/bph.12164 143 O’Brien, J.A., Lemaire, W., Wittmann, M., Jacobson, M.A., Ha, S.N., Wisnoski, D.D. et al. (2004) A novel selective allosteric modulator potentiates the activity of native metabotropic glutamate receptor subtype 5 in rat forebrain. J. Pharmacol. Exp. Ther. 309, 568–577, https://doi.org/10.1124/jpet.103.061747 144 Chen, Y., Goudet, C., Pin, J.P. and Conn, P.J. (2008) N-{4-Chloro-2-[(1,3-dioxo-1,3-dihydro-2H-isoindol-2-yl)methyl]phenyl}-2-hydroxybe nzamide (CPPHA) acts through a novel site as a positive allosteric modulator of group 1 metabotropic glutamate receptors. Mol. Pharmacol. 73, 909–918, https://doi.org/10.1124/mol.107.040097 145 Conde-Ceide, S., Martinez-Viturro, C.M., Alcazar, J., Garcia-Barrantes, P.M., Lavreysen, H., Mackie, C. et al. (2015) Discovery of VU0409551/JNJ-46778212: an mGlu5 positive allosteric modulator clinical candidate targeting schizophrenia. ACS Med. Chem. Lett. 6, 716–720, https://doi.org/10.1021/acsmedchemlett.5b00181 146 Doria, J.G., de Souza, J.M., Silva, F.R., Olmo, I.G., Carvalho, T.G., Alves-Silva, J. et al. (2018) The mGluR5 positive allosteric modulator VU0409551 improves synaptic plasticity and memory of a mouse model of Huntington’s disease. J. Neurochem., https://doi.org/10.1111/jnc.14555

c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society 2337 Clinical Science (2018) 132 2323–2338 https://doi.org/10.1042/CS20180374

147 Lindemann, L., Jaeschke, G., Michalon, A., Vieira, E., Honer, M., Spooren, W. et al. (2011) CTEP: a novel, potent, long-acting, and orally bioavailable metabotropic glutamate receptor 5 inhibitor. J. Pharmacol. Exp. Ther. 339, 474, https://doi.org/10.1124/jpet.111.185660 148 Michalon, A., Sidorov, M., Ballard, T.M., Ozmen, L., Spooren, W., Wettstein, J.G. et al. (2012) Chronic pharmacological mGlu5 inhibition corrects fragile Xinadultmice.Neuron 74, 49–56, https://doi.org/10.1016/j.neuron.2012.03.009 149 Chae, E., Shin, Y.J., Ryu, E.J., Ji, M.K., Ryune Cho, N., Lee, K.H. et al. (2013) Discovery of biological evaluation of pyrazole/imidazole amides asmGlu5 receptor negative allosteric modulators. Bioorg. Med. Chem. Lett. 23, 2134–2139, https://doi.org/10.1016/j.bmcl.2013.01.116 150 Ferraro, L., Loche, A., Beggiato, S., Tomasini, M.C., Antonelli, T., Colombo, G. et al. (2013) The new compound GET73, N-[(4-trifluoromethyl)benzyl]4-methoxybutyramide, regulates hippocampal aminoacidergic transmission possibly via an allosteric modulation of mGlu5 receptor. Behavioural evidence of its ‘anti-alcohol’ and anxiolytic properties. Curr. Med. Chem. 20, 3339–3357, https://doi.org/10.2174/09298673113209990167 151 Beggiato, S., Borelli, A.C., Tomasini, M.C., Castelli, M.P., Pintori, N., Cacciaglia, R. et al. (2018) In vitro functional characterization of GET73 as possible negative allosteric modulator of metabotropic glutamate receptor 5. Front. Pharmacol. 9, 327, https://doi.org/10.3389/fphar.2018.00327 152 Vranesic, I., Ofner, S., Flor, P.J., Bilbe, G., Bouhelal, R., Enz, A. et al. (2014) AFQ056/mavoglurant, a novel clinically effective mGluR5 antagonist: identification, SAR and pharmacological characterization. Bioorg. Med. Chem. 22, 5790–5803, https://doi.org/10.1016/j.bmc.2014.09.033 153 Nagel, J., Greco, S., Parsons, C.G., Flik, G., Tober, C., Klein, K.-U. et al. (2015) Brain concentrations of mGluR5 negative allosteric modulator MTEP in relation to receptor occupancy – Comparison to MPEP. Pharmacol. Rep. 67, 624–630, https://doi.org/10.1016/j.pharep.2015.01.004 154 Cosford, N.D.P., Roppe, J., Tehrani, L., Schweiger, E.J., Seiders, T.J., Chaudary, A. et al. (2003) [3H]-Methoxymethyl-MTEP and [3H]-Methoxy-PEPy: potent and selective radioligands for the metabotropic glutamate subtype 5 (mGlu5) receptor. Bioorg. Med. Chem. Lett. 13, 351–354, https://doi.org/10.1016/S0960-894X(02)00997-6 155 Gasparini, F., Lingenhohl, K., Stoehr, N., Flor, P.J., Heinrich, M., Vranesic, I. et al. (1999) 2-Methyl-6-(phenylethynyl)-pyridine (MPEP), a potent, selective and systemically active mGlu5 receptor antagonist. Neuropharmacology 38, 1493–1503, https://doi.org/10.1016/S0028-3908(99)00082-9 156 Homayoun, H., Stefani, M.R., Adams, B.W., Tamagan, G.D. and Moghaddam, B. (2004) Functional interaction between NMDA and mGlu5 receptors: effects on working memory, instrumental learning, motor behaviors, and dopamine release. Neuropsychopharmacology 29, 1259–1269, https://doi.org/10.1038/sj.npp.1300417 157 Um, J.W., Kaufman, A.C., Kostylev, M., Heiss, J.K., Stagi, M., Takahashi, H. et al. (2013) Metabotropic glutamate receptor 5 is a co-receptor for Alzheimer aβ oligomer bound to cellular prion protein. Neuron 79, 887–902, https://doi.org/10.1016/j.neuron.2013.06.036 158 Koros, E., Rosenbrock, H., Birk, G., Weiss, C. and Sams-Dodd, F. (2007) The selective mGlu5 receptor antagonist MTEP, similar to NMDA receptor antagonists, induces social isolation in rats. Neuropsychopharmacology 32, 562–576, https://doi.org/10.1038/sj.npp.1301133 159 Rodriguez, A.L., Nong, Y., Sekaran, N.K., Alagille, D., Tamagnan, G.D. and Conn, P.J. (2005) A close structural analog of 2-methyl-6-(phenylethynyl)-pyridine acts as a neutral allosteric site ligand on metabotropic glutamate receptor subtype 5 and blocks the effects of multiple allosteric modulators. Mol. Pharmacol. 68, 1793–1802 160 Hovelsø, N., Sotty, F., Montezinho, L.P., Pinheiro, P.S., Herrik, K.F. and Mørk, A. (2012) Therapeutic potential of metabotropic glutamate receptor modulators. Curr. Neuropharmacol. 10, 12–48, https://doi.org/10.2174/157015912799362805 161 Ayala, J.E., Chen, Y., Banko, J.L., Sheffler, D.J., Williams, R., Telk, A.N. et al. (2009) mGluR5 positive allosteric modulators facilitate both hippocampal LTP and LTD and enhance spatial learning. Neuropsychopharmacology 34, 2057–2071, https://doi.org/10.1038/npp.2009.30 162 Haas, L.T., Salazar, S.V., Smith, L.M., Zhao, H.R., Cox, T.O., Herber, C.S. et al. (2017) Silent allosteric modulation of mGluR5 maintains glutamate signaling while rescuing Alzheimer’s mouse phenotypes. Cell Rep. 20, 76–88, https://doi.org/10.1016/j.celrep.2017.06.023 163 Huang, H., Degnan, A.P., Balakrishnan, A., Easton, A., Gulianello, M., Huang, Y. et al. (2016) Oxazolidinone-based allosteric modulators of mGluR5: defining molecular switches to create a pharmacological tool box. Bioorg. Med. Chem. Lett. 26, 4165–4169, https://doi.org/10.1016/j.bmcl.2016.07.065

2338 c 2018 The Author(s). Published by Portland Press Limited on behalf of the Biochemical Society