Next Article in Journal
The Molecular Mechanisms Responsible for Tear Hyperosmolarity-Induced Pathological Changes in the Eyes of Dry Eye Disease Patients
Previous Article in Journal
FastCellpose: A Fast and Accurate Deep-Learning Framework for Segmentation of All Glomeruli in Mouse Whole-Kidney Microscopic Optical Images
Previous Article in Special Issue
Biochemical and Molecular Pathways in Neurodegenerative Diseases: An Integrated View
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Striatal Glutamate and Phosphodiesterases to Control L-DOPA-Induced Dyskinesia

by
Brik A. Kochoian
1,
Cassandra Bure
1 and
Stella M. Papa
1,2,*
1
Emory National Primate Research Center, Emory University, Atlanta, GA 30329, USA
2
Department of Neurology, Emory University School of Medicine, Atlanta, GA 30329, USA
*
Author to whom correspondence should be addressed.
Cells 2023, 12(23), 2754; https://doi.org/10.3390/cells12232754
Submission received: 1 November 2023 / Revised: 22 November 2023 / Accepted: 28 November 2023 / Published: 30 November 2023

Abstract

:
A large body of work during the past several decades has been focused on therapeutic strategies to control L-DOPA-induced dyskinesias (LIDs), common motor complications of long-term L-DOPA therapy in Parkinson’s disease (PD). Yet, LIDs remain a clinical challenge for the management of patients with advanced disease. Glutamatergic dysregulation of striatal projection neurons (SPNs) appears to be a key contributor to altered motor responses to L-DOPA. Targeting striatal hyperactivity at the glutamatergic neurotransmission level led to significant preclinical and clinical trials of a variety of antiglutamatergic agents. In fact, the only FDA-approved treatment for LIDs is amantadine, a drug with NMDAR antagonistic actions. Still, novel agents with improved pharmacological profiles are needed for LID therapy. Recently other therapeutic targets to reduce dysregulated SPN activity at the signal transduction level have emerged. In particular, mechanisms regulating the levels of cyclic nucleotides play a major role in the transduction of dopamine signals in SPNs. The phosphodiesterases (PDEs), a large family of enzymes that degrade cyclic nucleotides in a specific manner, are of special interest. We will review the research for antiglutamatergic and PDE inhibition strategies in view of the future development of novel LID therapies.

1. Introduction

Dopamine (DA) replacement therapy with L-DOPA in Parkinson’s disease (PD) has remained the most effective antiparkinsonian treatment since its inception more than half a century ago. However, long-term L-DOPA therapy leads to the development of motor complications that significantly reduce its efficacy. In particular, L-DOPA-induced dyskinesias (LIDs), which are involuntary movements, can affect patients after a few years of therapy. The development of LIDs largely depends on (1) prolonged non-physiologic, pulsatile treatment with L-DOPA or DA agonists and (2) progressive loss of the dopaminergic nigrostriatal system [1]. Other factors, such as abnormal DA release from serotonergic terminals and neuroinflammation, have been implicated in the development of LIDs [2,3]. Although the pathophysiology of LIDs is still not fully understood, it is well established that long-term therapy and progressive disease cause plasticity changes in striatal circuits, namely maladaptive plasticity, that underlie abnormal motor responses to dopaminergic stimulation [4,5,6]. In addition, chronic DA loss is associated with excitability changes in striatal projection neurons (SPNs) that suggest dysregulation of non-dopamine signaling mechanisms [7,8,9].
Several non-dopaminergic mechanisms may be at play to cause the maladaptive plasticity underlying altered SPN responses to DA. In particular, glutamatergic excitation mediated by corticostriatal and thalamostriatal afferents is a major contributor to the SPN plasticity changes. A large body of work supports that in advanced stages of PD, dysregulated glutamate signals lead to hyperexcitability and hyperactivity of SPNs [9,10]. In such a basal excitability state, DA replacement does not restore physiologic SPN outputs to basal ganglia circuits. Furthermore, DA action in hyperactive SPNs induces functional responses (changes in activity patterns) that are typically “unstable” and favor the appearance of LIDs [11,12]. Notably, reducing SPN hyperactivity by administration of NMDA receptor inhibitors stabilizes DA modulation of SPNs and improves LIDs [13,14,15].
DA signaling is mediated by G-protein-coupled receptors that are distributed in two subpopulations of SPNs, i.e., direct pathway SPNs (dSPNs) expressing DA type 1 receptors (D1Rs) and projecting to the internal pallidum and substantia nigra pars reticulata and indirect pathway SPNs (iSPNs) expressing DA type 2 receptors (D2Rs) and projecting to the external globus pallidum [16]. DA receptor activation regulates the synthesis of cyclic adenosine monophosphate (cAMP) and cyclic guanosine monophosphate (cGMP) (Figure 1). These cyclic nucleotides are second messengers that mediate SPN responses to dopaminergic modulation. DA exerts opposite effects on DA receptor subtypes, excitation on D1Rs and inhibition on D2Rs [17]. Therefore, cyclic nucleotides mediate opposite signals in D1R-expressing dSPNs and D2R-expressing iSPNs. However, measurements of global striatal levels of cAMP and cGMP show a decrease in rodent models of LIDs [18], which is difficult to explain by loss of DA modulation of cyclic nucleotide synthesis that is mediated differentially in dSPNs and iSPNs. cAMP and cGMP are catabolized by phosphodiesterases (PDEs) [19], and these enzymes may contribute to changes in cAMP and cGMP levels. Therefore, PDEs may play a role in molecular cascades associated with cyclic nucleotides, the messengers of DA signaling. Indeed, targeting striatal PDE activity with selective inhibitors may have therapeutic potential, particularly restoring physiological SPN responses to DA and thereby reducing LIDs. Several PDE inhibitors have shown anti-LID effects in animal models [20,21], but further studies are necessary to identify specific mechanisms and novel therapeutic targets. Therefore, this review is focused on key LID mechanisms based on maladaptive changes known to contribute to altered DA responses of SPNs, i.e., glutamatergic signals and cyclic nucleotide transduction pathways.

2. Dysregulation of Striatal Projection Neurons

2.1. Glutamate Signaling

Both dSPNs and iSPNs are affected by loss of dopaminergic innervation to the striatum in PD [22,23,24,25]. DA depletion leads to significant changes in SPN activity that include upregulated firing and excitability and plasticity changes [9,10,26,27,28]. Ex vivo studies using rodent models showed increased neuronal excitability in the striatum that is positively correlated with DA denervation [7]. Extracellular recordings in advanced parkinsonian non-human primates (NHPs) have shown that spontaneous SPN firing in the basal parkinsonian state is increased by more than 10-fold compared to normal [11,12]. This has also been shown in rat models [25,29,30,31]. Although these recordings in NHPs do not use high-cell-resolution methods, and possibly only iSPNs are hyperactive, the analysis of neuronal responses to L-DOPA suggests that dSPNs are also hyperactive. Similar activity patterns have been observed in patients with PD undergoing deep brain stimulation surgery [32], but due to some conflicting studies [33], the human data may require further analysis. Nevertheless, the available experimental data are highly compelling for upregulated activity of SPNs after DA depletion. This basal hyperactivity may contribute to abnormal L-DOPA responses in neurons that are highly sensitized following chronic loss of DA modulatory inputs [10]. Recordings in parkinsonian NHPs with reproducible LIDs have shown that SPNs exhibit “unstable” responses to DA replacement therapy [11]. The SPN activity changes developed at the onset of the “on” state (reversal of motor disability following DA replacement) are reversed at the peak of L-DOPA effect during the occurrence of dyskinesias [11]. Notably, unstable responses to DA are predominantly found in SPNs with significantly increased basal activity. Thus, the upregulated activity of SPNs underlying altered SPN responses to DA seems to play a role in LID mechanisms.
Alterations in the strength of excitatory synapses play a pivotal role in the acquisition of motor skills [34]. Synaptic plasticity mechanisms, including long-term potentiation (LTP), long-term depression (LTD), and synaptic depotentiation (SD), require the coordinated activation of multiple inputs onto SPNs [35]. Changes in corticostriatal plasticity have been found in models of PD [6,36], indicating a loss of bidirectional plasticity (LTP and LTD) in both d- and iSPNs, which exhibit only one type (unidirectional) [37]. Notably, in adult parkinsonian mice with stable dopaminergic lesions, iSPNs display LTP and dSPNs display LTD unidirectional corticostriatal plasticity [38]. However, in rodents chronically treated with L-DOPA and exhibiting LIDs, the unidirectional plasticity is reversed, i.e., iSPNs show only LTD and dSPNs show only LTP. Other studies in dyskinetic rats without SPN subtype differentiation showed loss of SD likely to be displayed by dSPNs according to pharmacological tests [39]. Using retrograde cell labeling, Belujon et al. 2010 showed that dyskinesias are correlated with an inability to de-depress established LTD in dSPNs [40]. Therefore, changes in excitability are intricately related to alterations in synaptic plasticity in both subtypes of SPNs. LTP, which is the unidirectional plasticity present in dSPNs in LID models, and the loss of SD in these neurons result in the strengthening of excitatory synapses, further heightening the excitability of these neurons in response to L-DOPA.
Glutamate plays a crucial role in the striatum by driving SPN firing and synaptic plasticity mechanisms. Glutamate signaling is mediated by ionotropic (NMDA, AMPA, and kainate) and metabotropic receptors (mGluRs). Ionotropic receptors are ligand-gated ion channels that produce a fast depolarization of the cell membrane and are implicated in various aspects of synaptic plasticity [35]. Conversely, mGluRs mediate signal transduction through phosphoinositide (PI) and cyclic nucleotide messenger systems that may produce slower and more prolonged excitatory or inhibitory responses. mGluRs are subdivided into three groups based on G-protein coupling and ligand profiles [41]. Group I receptors (mGluR1 and mGluR5) mediate excitatory responses, while groups II (mGluR2 and mGluR3) and III (mGluR4, mGluR6, mGluR7, and mGluR8) are inhibitory [41]. Group I receptors are postsynaptic on SPNs, while groups II and III are presynaptically located on corticostriatal terminals and GABAergic fibers [41]. The extended distribution of glutamate receptors with different functions emphasizes their importance in striatal circuit mechanisms.
In PD, DA loss significantly impacts striatal glutamatergic transmission, which can be interpreted as a compensatory mechanism in the SPN microcircuitry [41]. Importantly, these compensatory changes in glutamatergic activity within the striatum occur independently of DA or GABA levels [41], suggesting that glutamatergic function is primarily dysregulated [41]. In addition, chronic L-DOPA therapy has been shown to alter glutamate receptor expression [41], which may facilitate signal transmission. Altogether, data support that glutamatergic hyperactivity is associated with altered SPN responses to DA [14]. These changes underlie the emergence of involuntary movements in response to L-DOPA [41]. Table 1 summarizes the data relevant to the role of the glutamate system in LIDs.

2.2. Dopamine Signaling

DAR signaling is predominantly mediated through the cAMP/protein kinase A (PKA) molecular cascade [63] (Figure 1). D1R activation stimulates Gs/olf proteins, activating adenylyl cyclase (AC) to synthesize cAMP in dSPNs [64,65]. Conversely, D2R activation stimulates Gi/o proteins, inhibiting AC activity and thereby reducing cAMP levels in iSPNs. Notably, adenosine 2A receptors (A2AR) in iSPNs counteract D2Rs by activating AC, increasing cAMP synthesis, likely for preventing excessively low levels of cAMP. A2ARs and D2Rs form heteromers in the striatum, which leads to macromolecular complexes with different functionality than their individual receptors [66,67]. Of importance, recent work has further described the high proportion of receptors that display this interaction, emphasizing the development of A2AR-based antiparkinsonian therapy [68]. cAMP serves as a key modulator of various physiological processes that include PKA (regulations of Ca2+ levels and the function of various proteins) and cyclic nucleotide-gated channel (CNGC) mechanisms [69,70]. Activated PKA can also phosphorylate transcription factors, leading to gene expression changes affecting ion channels, synaptic plasticity, and neuronal excitability [69,70].
cGMP synthesis can be triggered through the activation of either membrane-bound particulate guanylyl cyclase (pGC) or soluble guanylyl cyclase (sGC) by natriuretic peptides and nitric oxide (NO), respectively [71,72]. NO is a gaseous free radical produced by GABAergic interneurons co-expressing neuronal NO synthase (nNOS), neuropeptide Y, and somatostatin [73] and requires concurrent NMDA receptor and D1R activation [74]. Upon binding to sGC, NO enhances GC activity, converting guanosine 5′-triphosphate (GTP) to cGMP. cGMP activates the cGMP/PKG molecular cascade and cGMP-gated ion channels [63]. Both cAMP and cGMP are crucial second messengers of DA signaling in SPNs.

2.3. Modulation of Dopamine Signaling by PDEs

Cyclic nucleotides are critically involved not only in dopaminergic but also glutamatergic signal transduction mechanisms. Drugs that increase cyclic nucleotide levels or reduce their degradation have been shown to facilitate corticostriatal transmission, while those that decrease cyclic nucleotide levels or increase their degradation can suppress it [75]. Interestingly, reductions in cGMP, but not cAMP, have been reported in animal models and patients with PD [76,77]. However, both striatal cAMP and cGMP levels are significantly lower at the peak of LID in rodent models [18,78]. cAMP and cGMP levels are regulated by PDEs, which are enzymes that hydrolyze the cyclic bond converting them into inactive 5′-AMP and 5′-GMP, respectively. Different isoforms of PDEs are expressed in distinct brain regions and cell types, allowing for precise modulation of cyclic nucleotide-mediated DA signaling. Numerous studies have demonstrated that PDE inhibitors can elevate intracellular nucleotide levels and facilitate membrane excitability and responsiveness of SPNs to corticostriatal inputs [75,79,80]. The function of PDEs in various neurological disorders, including PD, has recently received significant attention, even though the specific roles of PDEs in the pathophysiology of PD symptoms, including LIDs, remain to be investigated.

3. Targeting Glutamatergic Dysfunction

3.1. Ionotropic Glutamate Receptors

Following DA loss and chronic DA replacement therapy, various molecular, structural, and functional alterations occur in ionotropic glutamate receptors, including changes in their expression, composition, trafficking, and localization [81,82,83,84]. Striatal overexpression of NMDA and AMPA receptors has been shown in animal models and patients with PD [44,85]. Notably, there is an increased ratio of NMDA receptors (NMDARs) to AMPA receptors (AMPARs) following DA depletion as well as a higher ratio of GluN2A to GluN2B subunits after prolonged exposure to L-DOPA and the development of LIDs [42,46,47,48]. These changes may amplify excitatory glutamatergic transmission at corticostriatal synapses, a mechanism strongly implicated in LID pathophysiology [86,87,88,89,90].
Additionally, elevated levels of glutamate release and NMDAR activation have been observed in dyskinetic PD patients following L-DOPA administration [50]. Intrastriatal microinjection of AMPAR or NMDAR antagonists in parkinsonian and dyskinetic NHPs reduces basal SPN hyperactivity and counteracts the typical “unstable” response to L-DOPA, restoring stable SPN responses of non-dyskinetic parkinsonian NHPs [14]. Significantly, in these studies, the application of NMDAR antagonists across an extended area of the putamen in parkinsonian NHPs significantly reduces LIDs. Studies that tested the systemic administration of AMPA antagonists showed contradictory results likely because of different experimental paradigms [91,92]. Of interest, glycine (a co-agonist on NMDARs) antagonists improved the antiparkinsonian action of L-DOPA in a study of hemiparkinsonian NHPs and LIDs in parkinsonian rats and NHPs [93,94,95,96]. The systemic administration of selective NMDAR ion channel blockers and competitive antagonists significantly reduces dyskinesias [13,51,97]. These early tests used agents not viable for further development due to toxicity, but results served as proof of concept, and, thus, several non-competitive NMDAR antagonists were subsequently tested in rodent and primate models. Examples include amantadine, dextromethorphan, and memantine [98,99,100,101,102]. However, the pharmacological profiles of these drugs whose properties included partial antagonistic action, lower affinity, and multiple binding sites weakened their specific efficacy for controlling LIDs. In addition, a common problem with targeting NMDARs is their wide distribution in the brain, which favors dose-dependent off-target effects of selective agents [103,104]. One strategy to overcome this problem has been to develop subunit-selective inhibitors based on the predominant subunit composition of NMDARs in the striatum. Studies conducted with an NR2B-selective NMDAR antagonist, CP-101,606 (Pfizer, Inc., New York, NY, USA), have yielded inconsistent results in animals models, improving dyskinesia and augmenting the antiparkinsonian action of L-DOPA in some models and exacerbating LIDs in others [105,106]. Therefore, the experimental work provided evidence to support that therapeutic interventions targeting dysregulated glutamatergic signaling can reverse altered SPN responses to L-DOPA and control LIDs. Table 2 summarizes the relevant studies of agents targeting glutamatergic transmission.

3.2. Metabotropic Glutamate Receptors

The role of mGlu receptors (mGluRs) in LID pathophysiology has been extensively studied in search of more moderate effects that could have less toxicity. Within the mGluR family, mGluR5, a group I receptor, has garnered the most focus, whereas the role of mGluR1 remains less well defined, partly due to its lower presence in the striatum than mGluR5 [107,108]. Studies indicate that mGluR5 levels in the striatum correlate with LID development and are elevated in animal models and patients with LIDs [41,52,55]. Moreover, dyskinesia severity is linked to higher mGluR5 levels in parkinsonian NHPs with LIDs [53,54]. Genetic studies in mice that compared NMDAR-evoked responses between animals lacking mGluR1 and mGluR5 have shown a prominent role of mGluR5 in NMDAR signaling [109]. These findings highlight the specific role of mGluR5 in LID pathophysiology and have fueled interest in developing mGluR5 antagonists as therapeutic agents for LIDs.
Table 2. Antiglutamatergic pharmacotherapies.
Table 2. Antiglutamatergic pharmacotherapies.
ReceptorFindingsSpeciesReferences
NMDARNMDAR antagonists (e.g., amantadine, memantine, CP-101,606, and ramacemide) reduce LIDs (some studies also claim antiparkinsonian effects)RatPapa et al., 1995; Tronci et al., 2014 [51,98]
NHPSingh et al., 2018 [14]
HumanNutt et al., 2008; Shoulson et al., 2001 [104,110]
AMPARAMPAR antagonists reduce LIDsRatKobylecki et al., 2010 [91]
NHPKobylecki et al., 2010 [91]
AMPAR antagonists have no antidyskinetic effect
NHP
Human
Luquin et al., 1993 [92]
Eggert et al., 2010 [111]
Group I
mGluRs
mGluR5 NAMs reduced LIDsRat
NHP
Human
Rylander et al., 2009 [112]
Morin et al., 2010 [113]
McFarthing et al., 2019 [114]
mGluR1 antagonists modestly improved LIDsRat

Rylander et al., 2009 [112]
Group II
mGluRs
mGluR2/3 agonist improved LID durationRat

Zheng et al., 2020 [102]
Group III
mGluRs
mGluR4 PAM has antidyskinetic effectsRat
NHP
Calabrese et al., 2022 [62]
Charvin et al., 2018 [115]
mGluR4 PAMs “spare” L-DOPA, reducing dose requirementRat

Iderberg et al., 2015 [116]
Rat

Human
Le Poul et al., 2012 [117]
mGluR4 PAMs have antiparkinsonian effects but no effect on LIDs
Rascol et al., 2022 [118]
The mGluR5 negative allosteric modulator (NAM) 2-nethyl-6-(phenylethynyl)pyridine (MPEP) prevented the observed increase in mGluR5 expression following L-DOPA treatment in rodents [54,60] and reduced LIDs in parkinsonian NHPs [113]. Other mGluR5 NAMs, such as dipraglurant and mavoglurant, also demonstrated antidyskinetic effects without affecting the antiparkinsonian action of L-DOPA in MPTP-treated NHPs [119,120]. In contrast, mGluR1 antagonists have only modestly improved dyskinesia in animal models [112].
Preclinical studies also explored targeting presynaptic group II mGluRs to control LIDs by inhibiting glutamate release, thereby reducing the excitatory drive on SPNs and subthalamonigral terminals for improving parkinsonism [61,121,122]. However, translating this approach into effective LID treatments using mGluR2/3 agonists has yielded inconsistent results [90,112]. A more recent study compared amantadine to LY354740, an mGluR2/3 agonist, in parkinsonian rats and found that LY354740 weakly reduced LID duration but had no effect on LID intensity and interfered with the antiparkinsonian action of L-DOPA [102]. Consequently, group II agonists are unlikely to be as effective as group I mGluR5 antagonists for LID therapy.
Group III mGluRs, notably mGluR4, are emerging as therapeutic targets for LIDs. mGluR4 is also presynaptically expressed in corticostriatal synapses, reducing glutamate release. In vitro recordings of corticostriatal slices revealed that foliglurax decreases spontaneous glutamatergic transmission and, when co-administered with L-DOPA, restores bidirectional plasticity in the striatum [62]. Foliglurax, a selective positive allosteric modulator (PAM) at mGluR4, has antidyskinetic effects in parkinsonian rats and NHPs [115]. Foliglurax likely also modulates GABA release from iSPNs, which can mediate antiparkinsonian effects and reduce L-DOPA requirements [123]. Thus, some studies of mGluR4 PAMs in parkinsonian models also showed antiparkinsonian effects without a measurable impact on LIDs [116,117,124]. Nonetheless, further exploring mGluR4 PAMs also holds promise for improving LIDs indirectly by reducing L-DOPA dose requirements. Overall, mGluR5 and mGluR4 seem like better targets for LID treatment than mGluR1 and group II mGluRs (Table 2).

3.3. Recent Clinical Trials

Early clinical trials of NMDAR antagonists included ramacemide and CP-101,606, which were reported to decrease LIDs and augment the antiparkinsonian effects of L-DOPA [104,110]. Amantadine has been studied in multiple trials, showing a reduction of LID severity without affecting L-DOPA’s efficacy (ClinicalTrials.gov Identifiers: NCT02136914, NCT02274766). The antidyskinetic effect of amantadine is thought to primarily result from its inhibition of NMDARs, reducing glutamatergic hyperactivity [125], although it also has other effects, including preventing the abnormal reductions in cyclic nucleotides linked to LIDs in animal models of PD [78]. Of note, amantadine has also demonstrated anti-inflammatory effects in both in vitro and in vivo models of neuroinflammation and PD, leading to a reduction in levels of TNF-α, IL-1β, and NO [126,127]. Since neuroinflammation is thought to contribute to LIDs, the antidyskinetic effects of amantadine may be partially attributed to a reduction of neuroinflammation. Extended-release amantadine is the sole recently approved medication for the chronic treatment of LIDs, although the instant-release formulation has been used acutely for over 50 years [125]. It is important to note that amantadine can produce significant adverse events, including visual hallucinations, confusion, blurred vision, and gastrointestinal symptoms, which underscore the need for developing alternative treatments [125]. Glycine binds to the NMDAR as a co-agonist, contributing to opening of the ion channel, and the glycine receptor antagonist AV-101 (L-4-chlorokynurenine, VistaGen Therapeutics, Inc., South San Francisco, CA, USA) is in a Phase 2 trial for LID therapy (ClinicalTrials.gov Identifier: NCT04147949).
The mGluR5 NAM dipraglurant (ADX48621, Addex Therapeutics, Geneva, Switzerland) was tested in a Phase 2a trial (ClinicalTrials.gov Identifier: NCT04857359) for reducing peak-dose LIDs and showed positive effects [114]. However, other mGluR5 NAMs had only mild antidyskinetic effects, worsened motor function, and had increased adverse events [128,129,130]. Foliglurax (Prexton Therapeutics, Oss, The Netherlands), the first mGluR4 PAM evaluated clinically, failed to meet Phase 2 trial endpoints (ClinicalTrials.gov Identifier: NCT03162874) and showed no measurable improvement of LIDs [118]. Since the use of non-specific drugs targeting glutamate receptors typically produces off-target effects [104], clinical trials use narrow dose ranges to minimize adverse reactions at the expense of compromising efficacy [103]. Overall, antiglutamatergic pharmacotherapies have shown variable results for controlling LIDs (Table 2), with adverse events listed as the major limitation for their use [104,125].

4. Targeting Phosphodiesterases

4.1. Families and Properties of PDEs

The PDE superfamily, consisting of 21 genes and 11 subtypes with over 100 splice variants, controls cyclic nucleotide catabolism [19,131]. PDEs are an ideal target to manipulate signal transduction mechanisms in a region-specific manner since isoenzymes have distinct distributions in the brain [132,133]. Furthermore, PDEs exhibit different substrate affinities, selectively catabolizing cAMP, cGMP, or both, which enables precise control over the levels of a specific cyclic nucleotide. This specificity is particularly relevant to DA signaling pathways that may differentially utilize cAMP or cGMP, thereby offering a unique opportunity for developing precise drug therapies to modulate L-DOPA responses (Table 3). PDE isoforms with mRNA expression in the striatum include the dual-substrate enzymes PDE1B, PDE2A, and PDE10A that metabolize cAMP and cGMP as well as the cAMP-specific enzymes PDE3A, PDE3B, PDE4D, PDE7B, and PDE8B, along with the cGMP-specific enzyme PDE9A [133]. In recent years, PDE inhibitors have garnered attention for their potential therapeutic applications in various disorders that are associated with striatal pathophysiology, such as schizophrenia, Huntington’s disease, and PD [76,77,134,135]. We will review the most relevant research concerning PDE isoenzymes as potential targets for LID therapy.

4.2. Preclinical Research in Animal Models

4.2.1. Phosphodiesterase 10

PDE10A is expressed almost exclusively in SPNs (with very low mRNA expression found in other brain areas), localizing primarily in the membranes of dendrites and dendritic spines. This enzyme regulates the levels of cAMP and cGMP [133,136,137]. Given the restricted expression of PDE10A to the striatum, targeting PDE10A could have an improved safety profile compared with other PDEs that have a wide distribution in the CNS [69,138].
PDE10A expression and activity are decreased in the striatum and increased in the nucleus accumbens in parkinsonian rats [139]. Accordingly, cAMP levels were upregulated in the striatum and downregulated in the nucleus accumbens. However, cGMP levels were decreased in all DA-depleted regions [139], suggesting that other mechanisms intervene in striatal cGMP regulation in the context of PD [139]. Positron emission tomography (PET) imaging of (11)C-IMA107, a highly selective PDE10A radioligand, revealed lower PDE10A availability in the caudate, putamen, and globus pallidus in patients with PD than in healthy controls [140]. The decrease in PDE10A binding correlated with disease duration motor disability and LIDs [140].
Several studies that examined PDE10A inhibition in models of PD found antidyskinetic effects [20,141,142]. In parkinsonian rats, slight inhibition of PDE10A attenuated abnormal involuntary movements (AIMs; the rodent equivalent of LIDs) without interfering with cortically evoked activity of striatal neurons [142]. More marked inhibition of the enzyme using higher inhibitor doses resulted in AIM prolongation and increased cortically evoked activity [142]. These findings suggest that facilitating corticostriatal transmission through PDE10A inhibition is likely to contribute to AIMs. However, PDE10A inhibition activates iSPNs to a greater extent than dSPNs, suggesting a predominant action of PDE10A inhibition counteracting DA modulation in the striatopallidal pathway [69,143,144]. Therefore, the dose-dependent effect of PDE10A inhibition may be attributed to the selective activation of iSPNs at lower doses and non-selective activation of both iSPNs and dSPNs at higher doses. Given these findings, further analysis of the functional changes mediated by PDE10A inhibition with cell-type resolution is warranted.
Studies in NHPs showed that selective PDE10A inhibitors produce behavioral effects that are akin to D2R antagonism but distinct from typical antagonists, likely due to the combined D1R agonism-like effect [145,146]. A PDE10A inhibitor, MR1916 (Mochida Pharmaceuticals, Inc., Shinjuku City, Japan), reduced dyskinesia without interfering with L-DOPA’s efficacy in parkinsonian rats and NHPs [20]. In the primate study, the dose of MR1916 that produced the maximal antidyskinetic effect was not the highest dose tested, supporting the notion that the antidyskinetic properties of PDE10A inhibition are driven by facilitation of the striatopallidal, rather than striatonigral, pathway. Nevertheless, contradictory results have been reported regarding the role of PDE10A in PD and LIDs, and, thus, further investigation is needed to better understand the mechanisms and therapeutic potential of PDE10A inhibition.
Table 3. Properties of PDEs.
Table 3. Properties of PDEs.
FamilySubstrateRegulationCNS ExpressionInhibitor (s)Role in PD and LIDsReferences
PDE1cAMP/
cGMP
Cortex, hippocampus, cerebellum, striatum, nucleus accumbens, olfactory bulb, amygdala, thalamusVinpocetine, ITI-214, DSR-143136Upregulated in parkinsonian rats; genetic KO enhances responsiveness to DA agonists; inhibited by antiparkinsonian medications; antidyskinetic effect of inhibitors in NHPsSancesario et al., 2004; Lakics et al., 2010; Cenci et al., 2018; Enomoto et al., 2021; Kakkar et al., 1997; Kakkar et al., 1996; Polli et al., 1994; Reed et al., 2002 [77,133,147,148,149,150,151,152]
PDE2cAMP/
cGMP
cGMP stimulatedCortex, hippocampus, striatum, hypothalamus, amygdala, midbrainLu AF64280, BAY 60-7550Upregulated activity in dyskinetic ratsSancesario et al., 2014; Lakics et al., 2010; Doummar et al., 2020; Heckman et al., 2017; Salpietro et al., 2018; Simpson et al., 2010; Stephenson et al., 2009 [78,133,153,154,155,156,157]
PDE3cAMPcGMP
inhibited
Hippocampus, striatumCilostazol Lakics et al., 2010 [133]
PDE4cAMP Cortex, hippocampus, striatum, olfactory bulb, thalamus, hypothalamus, cerebellum, midbrainRolipram, ND1251, MK-0952, MEM1414, HT-0712, roflumilast, DenbufyllineModulates A2AR signaling; enhances A2AR- and D1R-mediated phosphorylation of DARPP-32; linked to LIDs in animal modelsNishi et al., 2008; Lakics et al., 2010; Casacchia et al., 1983; Cherry et al., 1999; Perez-Torres et al., 2000; Vignola et al., 2004; Yang et al., 2008 [69,133,158,159,160,161,162]
PDE5cGMP Cerebellum, spinal cord, cortex, hippocampusSildenafil, udenafil, vardenafil Lakics et al., 2010 [133]
PDE6cGMP Retinal rod, cone cells, pineal gland Lakics et al., 2010 [133]
PDE7cAMP Hippocampus, striatum, cerebellum, cortex, thalamus, hypothalamus, midbrainS14, BRL-50481Upregulated in degenerating DA cells; neuroprotective effects of inhibitorsLakics et al., 2010; Casacchia et al., 1983; Ciccocioppo et al., 2021; Morales-Garcia et al., 2015; Morales-Garcia et al., 2020; Morales-Garcia et al., 2011 [133,158,163,164,165,166]
PDE8cAMP Cortex, hippocampus, striatum, cerebellum, olfactory bulb, thalamus, hypothalamus, midbrainNon-chiral 9-benzyl-2-chloro-adenine derivatives, PF-04957325Potential role in motor performance and coordinationLakics et al., 2010; Golkowski et al., 2016; Kobayashi et al., 2003; Wu et al., 2022 [133,167,168,169]
PDE9cGMP Striatum, cerebellum, thalamus, hypothalamus, amygdala, olfactory bulb, cortex, hippocampus, midbrainPF-04447943, BI 409306, FRM-16606Inhibition enhanced antiparkinsonian action of L-DOPA in NHPs; no direct effect on LIDsMasilamoni et al., 2022; Lakics et al., 2010 [21,133]
PDE10cAMP/
cGMP
cAMP inhibitedStriatum (hippocampus, cortex, midbrain, and cerebellum: very low mRNA levels detected)MP-10, TAK-063, RO5545965, AMG 579, OMS824, PDM-042Inhibition shows antidyskinetic effects in animal models; regulates corticostriatal transmission; altered expression in PD patientsBeck et al., 2018; Addy et al., 2009; Lakics et al., 2010; Xie et al., 2006; Niccolini et al., 2015; Arakawa et al., 2020; Guimaraes et al., 2022; Schmidt et al., 2008; Uthayathas et al., 2014; Lenda et al., 2021 [20,103,133,136,140,141,142,145,146,170]
PDE11cAMP/
cGMP
Hippocampus Lakics et al., 2010 [133]

4.2.2. Phosphodiesterase 1

PDE1 is a family of enzymes that can also hydrolyze both cAMP and cGMP and is comprised of three genes: PDE1A, PDE1B, and PDE1C. PDE1B highly co-localizes with D1Rs in the frontal cortex and striatum [133], and, therefore, it is believed to primarily affect the function of dSPNs [151]. Studies of the enzyme levels showed upregulation in parkinsonian rats [77], while genetic knockout in mice enhanced responsiveness to DA agonists and increased locomotor activity [152]. A recent study reported that the PDE1 inhibitor DSR-143136 (Sumitomo Dainippon Pharma Co., Ltd., Osaka, Japan) reversed LIDs in parkinsonian NHPs without adversely affecting parkinsonian motor symptoms [148]. DSR-143136 is highly selective for PDE1B over PDE1A and PDE1C and exhibits brain penetration in rodents, lending credibility to its potential therapeutic application [148,171] and supporting further studies of PDE1 inhibition.

4.2.3. Phosphodiesterase 2

PDE2 is another dual-substrate enzyme, and one of its isoforms, PDE2A, is highly expressed in the striatum [133]. Immunoreactivity studies have shown that PDE2A is more abundant in the axons of SPNs than in their cell bodies [157]. Most research on PDE2 has focused on Alzheimer’s disease and schizophrenia, showing potential benefits for memory and cognitive function [154], which hypothetically could be applicable for the treatment of non-motor symptoms in PD. PDE2A is activated by cGMP binding to its allosteric site located in the N-terminal domain [157,172]. Notably, increased activity of cGMP-specific PDEs has been observed in dyskinetic rats, suggesting that inhibition of this enzymatic activity may be antidyskinetic [78]. Conflicting with this notion, mutations in PDE2A genes that result in reduced enzymatic activity have been associated with several other forms of dyskinesia, including chorea and pleiotropic fluctuation/paroxysmal dyskinesias [153,155]. Thus, the future of targeting PDE2A activity for LID therapy is uncertain. PDE11 is another dual-substrate enzyme, but its expression is predominantly in the hippocampus.

4.2.4. Phosphodiesterase 4

In the group of cAMP-selective PDEs expressed in the striatum, PDE3 has not been studied for motor effects in the context of PD. PDE4 is encoded by four genes: PDE4A, PDE4B, PDE4C, and PDE4D. PDE4B is the most abundant isoform in the CNS and is largely distributed through brain regions [133]. In the striatum, PDE4A, PDE4B, and PDE4D are expressed, but PDE4B is the most abundant [160,173]. PDE4B is expressed in both dSPNs and iSPNs, but immunohistochemical analysis revealed higher expression within the iSPNs [69].
Rolipram, a PDE4 inhibitor, weakly enhances cAMP/PKA signaling in the striatum both in vitro and in vivo [69]. This enhancement is aligned with adenosine A2A receptor (A2AR)-mediated signaling, which counteracts the action of D2Rs. Rolipram treatment increased A2AR-mediated phosphorylation of DARPP-32, a key signaling molecule, but not D1R-mediated phosphorylation of DARPP-32. Therefore, PDE4 inhibition is believed to primarily enhance adenosine signals in the striatal indirect pathway [69]. This effect of PDE4 inhibition counteracts DA activation of D2Rs, an effect similar to D2 receptor antagonism. Rolipram has also been reported to have protective effects for MPTP-induced neurodegeneration in models of PD [162].

4.2.5. Phosphodiesterase 7

PDE7 is another isoenzyme with specific affinity for cAMP, as shown by the effect of selective inhibitors significantly elevating cAMP levels in the CNS [163]. PDE7 has two known isoforms, PDE7A and PDE7B, each with distinct tissue distribution and subcellular localization. While mRNAs of both isoforms are expressed in the brain and peripherally throughout different body tissues, PDE7B predominates in the brain [133,174]. Based on upregulation of PDE7 in degenerating DA cells and microglia in various experimental models, studies of PDE7 inhibition have been focused mainly on neuroprotection [165]. S-14 (a non-selective inhibitor) and BRL50481 (a selective inhibitor) reduced microglial activation and neuronal cell loss [164,166]. These studies also reported improved apomorphine-induced asymmetric motor behavior, suggesting that these agents may potentiate the antiparkinsonian action of DA agonists. However, the rodent model used in this study was produced with lipopolysaccharide, an endotoxin that triggers neuronal damage via neuroinflammation. Therefore, the translatability of these results is unclear, and data have not yet been replicated in other models of PD. PDE8, which also specifically hydrolyzes cAMP and has some expression in the striatum [133,168], has not been investigated in the context of PD.

4.2.6. Phosphodiesterase 9

Three PDE isoenzymes (PDE5, PDE6, and PDE9) have specificity for cGMP, but PDE5 and PDE6 have very low striatal expression [175,176]. PDE9A is widely distributed in the brain and is abundant in the striatum [133]. Therefore, PDE9 inhibition is an experimental tool that can facilitate the study of mechanisms mediated by the cGMP-PKG signaling pathway. Several selective inhibitors have been used in parkinsonian models of PD and LIDs. The PDE9 inhibitor FRM-16606 (Forum Pharmaceuticals, Inc., Waltham, MA, USA) prolongs the antiparkinsonian action of L-DOPA with a slight increase in LIDs [21]. However, it is important to note that the LID increase was likely due to the longer duration of the L-DOPA response rather than a direct effect of the PDE9 inhibitor on LID mechanisms. PDE9 inhibition did not induce dyskinesias when administered without L-DOPA and did not augment LID scores at the peak of the L-DOPA response. The lack of antidyskinetic effects of PDE9 inhibition may be attributed to the different roles of the cAMP-PKA and cGMP-PKG pathways in LID mechanisms. Data also suggest that the NO-cGMP-PKG pathway is an important signaling mechanism in DA-mediated reversal of parkinsonian motor deficits. Striatal cGMP levels, but not cAMP levels, are decreased in parkinsonian animals [18,77], which may explain the synergistic effect of the PDE9 inhibitor with L-DOPA antiparkinsonian action. Of note, it is unclear how the NO-cGMP-PKG signaling pathway is modulated by D1R or D2R activation in the intact and DA-depleted striatum. Thus, PDE9-mediated mechanisms in distinct SPN subtypes and correlated motor effects remain to be investigated.

4.3. Recent Clinical Trials

While several PDE4 inhibitors have been investigated in clinical trials for neurological disorders, dose-limiting adverse events have often been determinant for low efficacy [161]. For example, rolipram at the doses tested did not improve the efficacy of L-DOPA or other dopaminergic drugs [158]. The extensive expression of PDE4 in the basal ganglia and its role in modulating adenosine and DA signaling pathways support further research for developing more selective and better-tolerated PDE4 inhibitors as potential therapeutic agents for PD.
Among numerous clinical trials that have explored PDE inhibitors for various cognitive and motor disorders, including antiparkinsonian effects [158], only one clinical trial was designed to determine the effects on LIDs. This ongoing trial of a selective PDE10A inhibitor, CPL500036 (Celon Pharma SA), is in Phase 2 (ClinicalTrials.gov Identifier: NCT05297201). Contrary to most experimental data on PDE10A inhibitors, the preclinical study that had investigated the efficacy of CPL500036 in hemiparkinsonian rats found antiparkinsonian-like effects, reversing impairments in contralateral forelimb function [170]. Notably, this study did not investigate CPL500036′s impact on LIDs and reported sedation in parkinsonian rats at therapeutic doses [170].

5. Concluding Remarks

Striatal maladaptive plasticity is highly involved in SPN responses to L-DOPA underlying dyskinesias. A large body of work has shown that glutamatergic dysregulation is a major contributor to the reduced homeostatic modulation of SPNs by DA replacement [81,82,83,84]. Ionotropic glutamate receptors undergo significant changes following DA loss and chronic replacement therapy, including alterations in expression, composition, and subunit ratios [44,47,48,85]. Such changes impact plasticity mechanisms at corticostriatal synapses. Metabotropic glutamate receptors also play a role, either modulating postsynaptic signaling pathways or regulating glutamate release [55]. On the other hand, recent data have also highlighted the role of altered DA signaling transduction in LID mechanisms. Changes in DA signal transduction pathways occur after chronic L-DOPA treatment, as indicated by the levels of cyclic nucleotides. Striatal cAMP and cGMP levels are lower in models of LIDs [18,78]. Therefore, there is a growing interest in the regulation of cyclic nucleotides, which is mediated by the catabolic PDE enzymes.
The review of preclinical and clinical work on glutamatergic agents lead to the conclusion that selective NMDAR antagonists and mGluR agents have antidyskinetic effects, but efficacy has been limited by the target characteristics (e.g., receptor distribution) and pharmacological profiles of available drugs [98,99,100,101,102,125]. Strategies to circumvent this relying on identifying receptor properties that could confer more region specificity are yet to be proven successful. Some examples would be activity-dependent binding agents or particular receptor biophysical features (binding kinetics and pH changes) [177,178,179]. Gene therapy provides an alternative approach for specific manipulation of striatal glutamate signaling. One study in rodents that assessed selective mGluR5 knockdown in dSPNs showed positive results [180]. Overall, there is compelling evidence for the link between striatal glutamate and LID mechanisms and to support further studies of this target to develop effective antidyskinetic therapies.
In turn, PDE inhibitors, including those selective for isoforms with dual- or single-substrate affinity, are of interest for LID treatment. PDE10A and PDE1B inhibition (dual action) have been studied in parkinsonian rats and NHPs, providing proof of concept that targeting cyclic nucleotide regulation may produce antidyskinetic effects [20,148]. PDE9 is the only isoenzyme that has specificity for cGMP (single action) and significant expression in the striatum. Selective PDE9 inhibition enhances the antiparkinsonian action of L-DOPA with little to no impact on LIDs [21]. Also important, PDEs with single-cAMP action have not yet been explored in the context of PD/LIDs. Therefore, the available data warrant further studies on DA signal transduction mechanisms and the role of different PDE targets in responses to DA replacement. Research into PDE inhibition is emerging as a new therapeutic avenue to control LIDs and improve the quality of life of patients with PD.

Author Contributions

All authors have contributed to writing and editing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NIH grants NS045962, NS110416, NS125502, NS126924, and NIH/ORIP OD011132 (S.M.P.) and NIH/NRSA NS124269 (B.A.K.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nutt, J.G.; Holford, N.H. The response to levodopa in Parkinson’s disease: Imposing pharmacological law and order. Ann. Neurol. 1996, 39, 561–573. [Google Scholar] [CrossRef] [PubMed]
  2. Corsi, S.; Stancampiano, R.; Carta, M. Serotonin/dopamine interaction in the induction and maintenance of L-DOPA-induced dyskinesia: An update. Prog. Brain Res. 2021, 261, 287–302. [Google Scholar] [CrossRef] [PubMed]
  3. Lopez-Lopez, A.; Labandeira, C.M.; Labandeira-Garcia, J.L.; Munoz, A. Rho kinase inhibitor fasudil reduces l-DOPA-induced dyskinesia in a rat model of Parkinson’s disease. Br. J. Pharmacol. 2020, 177, 5622–5641. [Google Scholar] [CrossRef] [PubMed]
  4. Calabresi, P.; Ghiglieri, V.; Mazzocchetti, P.; Corbelli, I.; Picconi, B. Levodopa-induced plasticity: A double-edged sword in Parkinson’s disease? Philos. Trans. R. Soc. Lond. B Biol. Sci. 2015, 370, 20140184. [Google Scholar] [CrossRef]
  5. Cenci, M.A.; Konradi, C. Maladaptive striatal plasticity in L-DOPA-induced dyskinesia. Prog. Brain. Res. 2010, 183, 209–233. [Google Scholar] [CrossRef] [PubMed]
  6. Cenci, M.A.; Lundblad, M. Post- versus presynaptic plasticity in L-DOPA-induced dyskinesia. J. Neurochem. 2006, 99, 381–392. [Google Scholar] [CrossRef]
  7. Calabresi, P.; Mercuri, N.B.; Sancesario, G.; Bernardi, G. Electrophysiology of dopamine-denervated striatal neurons. Implications for Parkinson’s disease. Brain 1993, 116 Pt 2, 433–452. [Google Scholar]
  8. Day, M.; Wokosin, D.; Plotkin, J.L.; Tian, X.; Surmeier, D.J. Differential excitability and modulation of striatal medium spiny neuron dendrites. J. Neurosci. 2008, 28, 11603–11614. [Google Scholar] [CrossRef]
  9. Surmeier, D.J.; Graves, S.M.; Shen, W. Dopaminergic modulation of striatal networks in health and Parkinson’s disease. Curr. Opin. Neurobiol. 2014, 29, 109–117. [Google Scholar] [CrossRef]
  10. Beck, G.; Singh, A.; Papa, S.M. Dysregulation of striatal projection neurons in Parkinson’s disease. J. Neural. Transm. 2018, 125, 449–460. [Google Scholar] [CrossRef]
  11. Liang, L.; DeLong, M.R.; Papa, S.M. Inversion of dopamine responses in striatal medium spiny neurons and involuntary movements. J. Neurosci. 2008, 28, 7537–7547. [Google Scholar] [CrossRef] [PubMed]
  12. Singh, A.; Liang, L.; Kaneoke, Y.; Cao, X.; Papa, S.M. Dopamine regulates distinctively the activity patterns of striatal output neurons in advanced parkinsonian primates. J. Neurophysiol. 2015, 113, 1533–1544. [Google Scholar] [CrossRef] [PubMed]
  13. Papa, S.M.; Chase, T.N. Levodopa-induced dyskinesias improved by a glutamate antagonist in Parkinsonian monkeys. Ann. Neurol. 1996, 39, 574–578. [Google Scholar] [CrossRef] [PubMed]
  14. Singh, A.; Jenkins, M.A.; Burke, K.J., Jr.; Beck, G.; Jenkins, A.; Scimemi, A.; Traynelis, S.F.; Papa, S.M. Glutamatergic Tuning of Hyperactive Striatal Projection Neurons Controls the Motor Response to Dopamine Replacement in Parkinsonian Primates. Cell Rep. 2018, 22, 941–952. [Google Scholar] [CrossRef] [PubMed]
  15. Gardoni, F.; Sgobio, C.; Pendolino, V.; Calabresi, P.; Di Luca, M.; Picconi, B. Targeting NR2A-containing NMDA receptors reduces L-DOPA-induced dyskinesias. Neurobiol. Aging 2012, 33, 2138–2144. [Google Scholar] [CrossRef]
  16. Gerfen, C.R. The neostriatal mosaic: Multiple levels of compartmental organization. Trends Neurosci. 1992, 15, 133–139. [Google Scholar] [CrossRef] [PubMed]
  17. Gerfen, C.R.; Surmeier, D.J. Modulation of striatal projection systems by dopamine. Annu. Rev. Neurosci. 2011, 34, 441–466. [Google Scholar] [CrossRef] [PubMed]
  18. Giorgi, M.; D’Angelo, V.; Esposito, Z.; Nuccetelli, V.; Sorge, R.; Martorana, A.; Stefani, A.; Bernardi, G.; Sancesario, G. Lowered cAMP and cGMP signalling in the brain during levodopa-induced dyskinesias in hemiparkinsonian rats: New aspects in the pathogenetic mechanisms. Eur. J. Neurosci. 2008, 28, 941–950. [Google Scholar] [CrossRef]
  19. Beavo, J.A. Cyclic nucleotide phosphodiesterases: Functional implications of multiple isoforms. Physiol. Rev. 1995, 75, 725–748. [Google Scholar] [CrossRef]
  20. Beck, G.; Maehara, S.; Chang, P.L.; Papa, S.M. A Selective Phosphodiesterase 10A Inhibitor Reduces L-Dopa-Induced Dyskinesias in Parkinsonian Monkeys. Mov. Disord. 2018, 33, 805–814. [Google Scholar] [CrossRef]
  21. Masilamoni, G.J.; Sinon, C.G.; Kochoian, B.A.; Singh, A.; McRiner, A.J.; Leventhal, L.; Papa, S.M. Phosphodiesterase 9 inhibition prolongs the antiparkinsonian action of l-DOPA in parkinsonian non-human primates. Neuropharmacology 2022, 212, 109060. [Google Scholar] [CrossRef] [PubMed]
  22. Bagetta, V.; Picconi, B.; Marinucci, S.; Sgobio, C.; Pendolino, V.; Ghiglieri, V.; Fusco, F.R.; Giampa, C.; Calabresi, P. Dopamine-dependent long-term depression is expressed in striatal spiny neurons of both direct and indirect pathways: Implications for Parkinson’s disease. J. Neurosci. 2011, 31, 12513–12522. [Google Scholar] [CrossRef] [PubMed]
  23. Fieblinger, T.; Graves, S.M.; Sebel, L.E.; Alcacer, C.; Plotkin, J.L.; Gertler, T.S.; Chan, C.S.; Heiman, M.; Greengard, P.; Cenci, M.A.; et al. Cell type-specific plasticity of striatal projection neurons in parkinsonism and L-DOPA-induced dyskinesia. Nat. Commun. 2014, 5, 5316. [Google Scholar] [CrossRef] [PubMed]
  24. Nelson, A.B.; Kreitzer, A.C. Reassessing models of basal ganglia function and dysfunction. Annu. Rev. Neurosci. 2014, 37, 117–135. [Google Scholar] [CrossRef] [PubMed]
  25. Tseng, K.Y.; Kasanetz, F.; Kargieman, L.; Riquelme, L.A.; Murer, M.G. Cortical slow oscillatory activity is reflected in the membrane potential and spike trains of striatal neurons in rats with chronic nigrostriatal lesions. J. Neurosci. 2001, 21, 6430–6439. [Google Scholar] [CrossRef] [PubMed]
  26. Calabresi, P.; Picconi, B.; Tozzi, A.; Ghiglieri, V.; Di Filippo, M. Direct and indirect pathways of basal ganglia: A critical reappraisal. Nat. Neurosci. 2014, 17, 1022–1030. [Google Scholar] [CrossRef] [PubMed]
  27. Calabresi, P.; Pisani, A.; Rothwell, J.; Ghiglieri, V.; Obeso, J.A.; Picconi, B. Hyperkinetic disorders and loss of synaptic downscaling. Nat. Neurosci. 2016, 19, 868–875. [Google Scholar] [CrossRef]
  28. DeLong, M.R. Primate models of movement disorders of basal ganglia origin. Trends Neurosci. 1990, 13, 281–285. [Google Scholar] [CrossRef]
  29. Gubellini, P.; Picconi, B.; Bari, M.; Battista, N.; Calabresi, P.; Centonze, D.; Bernardi, G.; Finazzi-Agro, A.; Maccarrone, M. Experimental parkinsonism alters endocannabinoid degradation: Implications for striatal glutamatergic transmission. J. Neurosci. 2002, 22, 6900–6907. [Google Scholar] [CrossRef]
  30. Kita, H.; Kita, T. Role of Striatum in the Pause and Burst Generation in the Globus Pallidus of 6-OHDA-Treated Rats. Front. Syst. Neurosci. 2011, 5, 42. [Google Scholar] [CrossRef]
  31. Sharott, A.; Vinciati, F.; Nakamura, K.C.; Magill, P.J. A Population of Indirect Pathway Striatal Projection Neurons Is Selectively Entrained to Parkinsonian Beta Oscillations. J. Neurosci. 2017, 37, 9977–9998. [Google Scholar] [CrossRef] [PubMed]
  32. Singh, A.; Mewes, K.; Gross, R.E.; DeLong, M.R.; Obeso, J.A.; Papa, S.M. Human striatal recordings reveal abnormal discharge of projection neurons in Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2016, 113, 9629–9634. [Google Scholar] [CrossRef] [PubMed]
  33. Deffains, M.; Iskhakova, L.; Katabi, S.; Haber, S.N.; Israel, Z.; Bergman, H. Subthalamic, not striatal, activity correlates with basal ganglia downstream activity in normal and parkinsonian monkeys. Elife 2016, 5, e16443. [Google Scholar] [CrossRef] [PubMed]
  34. Yin, H.H.; Mulcare, S.P.; Hilario, M.R.; Clouse, E.; Holloway, T.; Davis, M.I.; Hansson, A.C.; Lovinger, D.M.; Costa, R.M. Dynamic reorganization of striatal circuits during the acquisition and consolidation of a skill. Nat. Neurosci. 2009, 12, 333–341. [Google Scholar] [CrossRef]
  35. Calabresi, P.; Pisani, A.; Mercuri, N.B.; Bernardi, G. Long-term Potentiation in the Striatum is Unmasked by Removing the Voltage-dependent Magnesium Block of NMDA Receptor Channels. Eur. J. Neurosci. 1992, 4, 929–935. [Google Scholar] [CrossRef]
  36. Calabresi, P.; Giacomini, P.; Centonze, D.; Bernardi, G. Levodopa-induced dyskinesia: A pathological form of striatal synaptic plasticity? Ann. Neurol. 2000, 47, S60–S68, discussion S68–S69. [Google Scholar]
  37. Shen, W.; Flajolet, M.; Greengard, P.; Surmeier, D.J. Dichotomous dopaminergic control of striatal synaptic plasticity. Science 2008, 321, 848–851. [Google Scholar] [CrossRef]
  38. Thiele, S.L.; Chen, B.; Lo, C.; Gertler, T.S.; Warre, R.; Surmeier, J.D.; Brotchie, J.M.; Nash, J.E. Selective loss of bi-directional synaptic plasticity in the direct and indirect striatal output pathways accompanies generation of parkinsonism and l-DOPA induced dyskinesia in mouse models. Neurobiol. Dis. 2014, 71, 334–344. [Google Scholar] [CrossRef]
  39. Picconi, B.; Centonze, D.; Hakansson, K.; Bernardi, G.; Greengard, P.; Fisone, G.; Cenci, M.A.; Calabresi, P. Loss of bidirectional striatal synaptic plasticity in L-DOPA-induced dyskinesia. Nat. Neurosci. 2003, 6, 501–506. [Google Scholar] [CrossRef]
  40. Belujon, P.; Lodge, D.J.; Grace, A.A. Aberrant striatal plasticity is specifically associated with dyskinesia following levodopa treatment. Mov. Disord. 2010, 25, 1568–1576. [Google Scholar] [CrossRef]
  41. Sebastianutto, I.; Cenci, M.A. mGlu receptors in the treatment of Parkinson’s disease and L-DOPA-induced dyskinesia. Curr. Opin. Pharmacol. 2018, 38, 81–89. [Google Scholar] [CrossRef] [PubMed]
  42. Bagetta, V.; Sgobio, C.; Pendolino, V.; Del Papa, G.; Tozzi, A.; Ghiglieri, V.; Giampa, C.; Zianni, E.; Gardoni, F.; Calabresi, P.; et al. Rebalance of striatal NMDA/AMPA receptor ratio underlies the reduced emergence of dyskinesia during D2-like dopamine agonist treatment in experimental Parkinson’s disease. J. Neurosci. 2012, 32, 17921–17931. [Google Scholar] [CrossRef] [PubMed]
  43. Chase, T.N.; Oh, J.D. Striatal dopamine- and glutamate-mediated dysregulation in experimental parkinsonism. Trends Neurosci. 2000, 23, S86–S91. [Google Scholar] [CrossRef] [PubMed]
  44. Calon, F.; Morissette, M.; Ghribi, O.; Goulet, M.; Grondin, R.; Blanchet, P.J.; Bedard, P.J.; Di Paolo, T. Alteration of glutamate receptors in the striatum of dyskinetic 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine-treated monkeys following dopamine agonist treatment. Prog. Neuropsychopharmacol. Biol. Psychiatry 2002, 26, 127–138. [Google Scholar] [CrossRef] [PubMed]
  45. Calon, F.; Rajput, A.H.; Hornykiewicz, O.; Bedard, P.J.; Di Paolo, T. Levodopa-induced motor complications are associated with alterations of glutamate receptors in Parkinson’s disease. Neurobiol. Dis. 2003, 14, 404–416. [Google Scholar] [CrossRef] [PubMed]
  46. Gardoni, F.; Picconi, B.; Ghiglieri, V.; Polli, F.; Bagetta, V.; Bernardi, G.; Cattabeni, F.; Di Luca, M.; Calabresi, P. A critical interaction between NR2B and MAGUK in L-DOPA induced dyskinesia. J. Neurosci. 2006, 26, 2914–2922. [Google Scholar] [CrossRef] [PubMed]
  47. Hallett, P.J.; Dunah, A.W.; Ravenscroft, P.; Zhou, S.; Bezard, E.; Crossman, A.R.; Brotchie, J.M.; Standaert, D.G. Alterations of striatal NMDA receptor subunits associated with the development of dyskinesia in the MPTP-lesioned primate model of Parkinson’s disease. Neuropharmacology 2005, 48, 503–516. [Google Scholar] [CrossRef] [PubMed]
  48. Mellone, M.; Stanic, J.; Hernandez, L.F.; Iglesias, E.; Zianni, E.; Longhi, A.; Prigent, A.; Picconi, B.; Calabresi, P.; Hirsch, E.C.; et al. NMDA receptor GluN2A/GluN2B subunit ratio as synaptic trait of levodopa-induced dyskinesias: From experimental models to patients. Front. Cell Neurosci. 2015, 9, 245. [Google Scholar] [CrossRef]
  49. Nash, J.E.; Brotchie, J.M. Characterisation of striatal NMDA receptors involved in the generation of parkinsonian symptoms: Intrastriatal microinjection studies in the 6-OHDA-lesioned rat. Mov. Disord. 2002, 17, 455–466. [Google Scholar] [CrossRef]
  50. Ahmed, I.; Bose, S.K.; Pavese, N.; Ramlackhansingh, A.; Turkheimer, F.; Hotton, G.; Hammers, A.; Brooks, D.J. Glutamate NMDA receptor dysregulation in Parkinson’s disease with dyskinesias. Brain 2011, 134, 979–986. [Google Scholar] [CrossRef]
  51. Papa, S.M.; Boldry, R.C.; Engber, T.M.; Kask, A.M.; Chase, T.N. Reversal of levodopa-induced motor fluctuations in experimental parkinsonism by NMDA receptor blockade. Brain Res. 1995, 701, 13–18. [Google Scholar] [CrossRef] [PubMed]
  52. Crabbe, M.; Van der Perren, A.; Weerasekera, A.; Himmelreich, U.; Baekelandt, V.; Van Laere, K.; Casteels, C. Altered mGluR5 binding potential and glutamine concentration in the 6-OHDA rat model of acute Parkinson’s disease and levodopa-induced dyskinesia. Neurobiol. Aging 2018, 61, 82–92. [Google Scholar] [CrossRef] [PubMed]
  53. Morin, N.; Gregoire, L.; Morissette, M.; Desrayaud, S.; Gomez-Mancilla, B.; Gasparini, F.; Di Paolo, T. MPEP, an mGlu5 receptor antagonist, reduces the development of L-DOPA-induced motor complications in de novo parkinsonian monkeys: Biochemical correlates. Neuropharmacology 2013, 66, 355–364. [Google Scholar] [CrossRef] [PubMed]
  54. Morin, N.; Morissette, M.; Gregoire, L.; Gomez-Mancilla, B.; Gasparini, F.; Di Paolo, T. Chronic treatment with MPEP, an mGlu5 receptor antagonist, normalizes basal ganglia glutamate neurotransmission in L-DOPA-treated parkinsonian monkeys. Neuropharmacology 2013, 73, 216–231. [Google Scholar] [CrossRef] [PubMed]
  55. Ouattara, B.; Gregoire, L.; Morissette, M.; Gasparini, F.; Vranesic, I.; Bilbe, G.; Johns, D.R.; Rajput, A.; Hornykiewicz, O.; Rajput, A.H.; et al. Metabotropic glutamate receptor type 5 in levodopa-induced motor complications. Neurobiol. Aging 2011, 32, 1286–1295. [Google Scholar] [CrossRef] [PubMed]
  56. Armentero, M.T.; Fancellu, R.; Nappi, G.; Bramanti, P.; Blandini, F. Prolonged blockade of NMDA or mGluR5 glutamate receptors reduces nigrostriatal degeneration while inducing selective metabolic changes in the basal ganglia circuitry in a rodent model of Parkinson’s disease. Neurobiol. Dis. 2006, 22, 1–9. [Google Scholar] [CrossRef] [PubMed]
  57. Masilamoni, G.J.; Bogenpohl, J.W.; Alagille, D.; Delevich, K.; Tamagnan, G.; Votaw, J.R.; Wichmann, T.; Smith, Y. Metabotropic glutamate receptor 5 antagonist protects dopaminergic and noradrenergic neurons from degeneration in MPTP-treated monkeys. Brain 2011, 134, 2057–2073. [Google Scholar] [CrossRef]
  58. Sanchez-Pernaute, R.; Wang, J.Q.; Kuruppu, D.; Cao, L.; Tueckmantel, W.; Kozikowski, A.; Isacson, O.; Brownell, A.L. Enhanced binding of metabotropic glutamate receptor type 5 (mGluR5) PET tracers in the brain of parkinsonian primates. Neuroimage 2008, 42, 248–251. [Google Scholar] [CrossRef]
  59. Samadi, P.; Gregoire, L.; Morissette, M.; Calon, F.; Hadj Tahar, A.; Dridi, M.; Belanger, N.; Meltzer, L.T.; Bedard, P.J.; Di Paolo, T. mGluR5 metabotropic glutamate receptors and dyskinesias in MPTP monkeys. Neurobiol. Aging 2008, 29, 1040–1051. [Google Scholar] [CrossRef]
  60. Lea, P.M.t.; Faden, A.I. Metabotropic glutamate receptor subtype 5 antagonists MPEP and MTEP. CNS Drug Rev. 2006, 12, 149–166. [Google Scholar] [CrossRef]
  61. Pisani, A.; Bonsi, P.; Centonze, D.; Calabresi, P.; Bernardi, G. Activation of D2-like dopamine receptors reduces synaptic inputs to striatal cholinergic interneurons. J. Neurosci. 2000, 20, RC69. [Google Scholar] [CrossRef] [PubMed]
  62. Calabrese, V.; Picconi, B.; Heck, N.; Campanelli, F.; Natale, G.; Marino, G.; Sciaccaluga, M.; Ghiglieri, V.; Tozzi, A.; Anceaume, E.; et al. A positive allosteric modulator of mGlu4 receptors restores striatal plasticity in an animal model of l-Dopa-induced dyskinesia. Neuropharmacology 2022, 218, 109205. [Google Scholar] [CrossRef] [PubMed]
  63. Seifert, R.; Schneider, E.H.; Bahre, H. From canonical to non-canonical cyclic nucleotides as second messengers: Pharmacological implications. Pharmacol. Ther. 2015, 148, 154–184. [Google Scholar] [CrossRef] [PubMed]
  64. Beaulieu, J.M.; Gainetdinov, R.R. The physiology, signaling, and pharmacology of dopamine receptors. Pharmacol. Rev. 2011, 63, 182–217. [Google Scholar] [CrossRef]
  65. Sibley, D.R.; Monsma, F.J., Jr.; Shen, Y. Molecular neurobiology of dopaminergic receptors. Int. Rev. Neurobiol. 1993, 35, 391–415. [Google Scholar] [CrossRef]
  66. Franco, R.; Ferre, S.; Agnati, L.; Torvinen, M.; Gines, S.; Hillion, J.; Casado, V.; Lledo, P.; Zoli, M.; Lluis, C.; et al. Evidence for adenosine/dopamine receptor interactions: Indications for heteromerization. Neuropsychopharmacology 2000, 23, S50–S59. [Google Scholar] [CrossRef]
  67. Hillion, J.; Canals, M.; Torvinen, M.; Casado, V.; Scott, R.; Terasmaa, A.; Hansson, A.; Watson, S.; Olah, M.E.; Mallol, J.; et al. Coaggregation, cointernalization, and codesensitization of adenosine A2A receptors and dopamine D2 receptors. J. Biol. Chem. 2002, 277, 18091–18097. [Google Scholar] [CrossRef]
  68. Rivas-Santisteban, R.; Rico, A.J.; Munoz, A.; Rodriguez-Perez, A.I.; Reyes-Resina, I.; Navarro, G.; Labandeira-Garcia, J.L.; Lanciego, J.L.; Franco, R. Boolean analysis shows a high proportion of dopamine D(2) receptors interacting with adenosine A(2A) receptors in striatal medium spiny neurons of mouse and non-human primate models of Parkinson’s disease. Neurobiol. Dis. 2023, 188, 106341. [Google Scholar] [CrossRef]
  69. Nishi, A.; Kuroiwa, M.; Miller, D.B.; O’Callaghan, J.P.; Bateup, H.S.; Shuto, T.; Sotogaku, N.; Fukuda, T.; Heintz, N.; Greengard, P.; et al. Distinct roles of PDE4 and PDE10A in the regulation of cAMP/PKA signaling in the striatum. J. Neurosci. 2008, 28, 10460–10471. [Google Scholar] [CrossRef]
  70. Siuciak, J.A.; McCarthy, S.A.; Chapin, D.S.; Fujiwara, R.A.; James, L.C.; Williams, R.D.; Stock, J.L.; McNeish, J.D.; Strick, C.A.; Menniti, F.S.; et al. Genetic deletion of the striatum-enriched phosphodiesterase PDE10A: Evidence for altered striatal function. Neuropharmacology 2006, 51, 374–385. [Google Scholar] [CrossRef]
  71. Bredt, D.S. Nitric oxide signaling specificity--the heart of the problem. J. Cell Sci. 2003, 116, 9–15. [Google Scholar] [CrossRef] [PubMed]
  72. Garthwaite, J. Concepts of neural nitric oxide-mediated transmission. Eur. J. Neurosci. 2008, 27, 2783–2802. [Google Scholar] [CrossRef] [PubMed]
  73. Kawaguchi, Y.; Wilson, C.J.; Augood, S.J.; Emson, P.C. Striatal interneurones: Chemical, physiological and morphological characterization. Trends Neurosci. 1995, 18, 527–535. [Google Scholar] [CrossRef] [PubMed]
  74. West, A.R.; Tseng, K.Y. Nitric Oxide-Soluble Guanylyl Cyclase-Cyclic GMP Signaling in the Striatum: New Targets for the Treatment of Parkinson’s Disease? Front. Syst. Neurosci. 2011, 5, 55. [Google Scholar] [CrossRef] [PubMed]
  75. Padovan-Neto, F.E.; Sammut, S.; Chakroborty, S.; Dec, A.M.; Threlfell, S.; Campbell, P.W.; Mudrakola, V.; Harms, J.F.; Schmidt, C.J.; West, A.R. Facilitation of corticostriatal transmission following pharmacological inhibition of striatal phosphodiesterase 10A: Role of nitric oxide-soluble guanylyl cyclase-cGMP signaling pathways. J. Neurosci. 2015, 35, 5781–5791. [Google Scholar] [CrossRef] [PubMed]
  76. Nishino, N.; Kitamura, N.; Hashimoto, T.; Tanaka, C. Transmembrane signalling systems in the brain of patients with Parkinson’s disease. Rev. Neurosci. 1993, 4, 213–222. [Google Scholar] [CrossRef]
  77. Sancesario, G.; Giorgi, M.; D’Angelo, V.; Modica, A.; Martorana, A.; Morello, M.; Bengtson, C.P.; Bernardi, G. Down-regulation of nitrergic transmission in the rat striatum after chronic nigrostriatal deafferentation. Eur. J. Neurosci. 2004, 20, 989–1000. [Google Scholar] [CrossRef]
  78. Sancesario, G.; Morrone, L.A.; D’Angelo, V.; Castelli, V.; Ferrazzoli, D.; Sica, F.; Martorana, A.; Sorge, R.; Cavaliere, F.; Bernardi, G.; et al. Levodopa-induced dyskinesias are associated with transient down-regulation of cAMP and cGMP in the caudate-putamen of hemiparkinsonian rats: Reduced synthesis or increased catabolism? Neurochem. Int. 2014, 79, 44–56. [Google Scholar] [CrossRef]
  79. Threlfell, S.; Sammut, S.; Menniti, F.S.; Schmidt, C.J.; West, A.R. Inhibition of Phosphodiesterase 10A Increases the Responsiveness of Striatal Projection Neurons to Cortical Stimulation. J. Pharmacol. Exp. Ther. 2009, 328, 785–795. [Google Scholar] [CrossRef]
  80. West, A.R.; Grace, A.A. The nitric oxide-guanylyl cyclase signaling pathway modulates membrane activity States and electrophysiological properties of striatal medium spiny neurons recorded in vivo. J. Neurosci. 2004, 24, 1924–1935. [Google Scholar] [CrossRef]
  81. Calabresi, P.; Di Filippo, M.; Ghiglieri, V.; Tambasco, N.; Picconi, B. Levodopa-induced dyskinesias in patients with Parkinson’s disease: Filling the bench-to-bedside gap. Lancet Neurol. 2010, 9, 1106–1117. [Google Scholar] [CrossRef] [PubMed]
  82. Dunah, A.W.; Wang, Y.; Yasuda, R.P.; Kameyama, K.; Huganir, R.L.; Wolfe, B.B.; Standaert, D.G. Alterations in subunit expression, composition, and phosphorylation of striatal N-methyl-D-aspartate glutamate receptors in a rat 6-hydroxydopamine model of Parkinson’s disease. Mol. Pharmacol. 2000, 57, 342–352. [Google Scholar] [PubMed]
  83. Picconi, B.; Piccoli, G.; Calabresi, P. Synaptic dysfunction in Parkinson’s disease. Adv. Exp. Med. Biol. 2012, 970, 553–572. [Google Scholar] [CrossRef] [PubMed]
  84. Vastagh, C.; Gardoni, F.; Bagetta, V.; Stanic, J.; Zianni, E.; Giampa, C.; Picconi, B.; Calabresi, P.; Di Luca, M. N-methyl-D-aspartate (NMDA) receptor composition modulates dendritic spine morphology in striatal medium spiny neurons. J. Biol. Chem. 2012, 287, 18103–18114. [Google Scholar] [CrossRef] [PubMed]
  85. Sgambato-Faure, V.; Cenci, M.A. Glutamatergic mechanisms in the dyskinesias induced by pharmacological dopamine replacement and deep brain stimulation for the treatment of Parkinson’s disease. Prog. Neurobiol. 2012, 96, 69–86. [Google Scholar] [CrossRef] [PubMed]
  86. Calabresi, P.; Centonze, D.; Gubellini, P.; Marfia, G.A.; Pisani, A.; Sancesario, G.; Bernardi, G. Synaptic transmission in the striatum: From plasticity to neurodegeneration. Prog. Neurobiol. 2000, 61, 231–265. [Google Scholar] [CrossRef] [PubMed]
  87. Centonze, D.; Gubellini, P.; Rossi, S.; Picconi, B.; Pisani, A.; Bernardi, G.; Calabresi, P.; Baunez, C. Subthalamic nucleus lesion reverses motor abnormalities and striatal glutamatergic overactivity in experimental parkinsonism. Neuroscience 2005, 133, 831–840. [Google Scholar] [CrossRef] [PubMed]
  88. Gubellini, P.; Eusebio, A.; Oueslati, A.; Melon, C.; Kerkerian-Le Goff, L.; Salin, P. Chronic high-frequency stimulation of the subthalamic nucleus and L-DOPA treatment in experimental parkinsonism: Effects on motor behaviour and striatal glutamate transmission. Eur. J. Neurosci. 2006, 24, 1802–1814. [Google Scholar] [CrossRef]
  89. Picconi, B.; Centonze, D.; Rossi, S.; Bernardi, G.; Calabresi, P. Therapeutic doses of L-dopa reverse hypersensitivity of corticostriatal D2-dopamine receptors and glutamatergic overactivity in experimental parkinsonism. Brain 2004, 127, 1661–1669. [Google Scholar] [CrossRef]
  90. Picconi, B.; Pisani, A.; Centonze, D.; Battaglia, G.; Storto, M.; Nicoletti, F.; Bernardi, G.; Calabresi, P. Striatal metabotropic glutamate receptor function following experimental parkinsonism and chronic levodopa treatment. Brain 2002, 125, 2635–2645. [Google Scholar] [CrossRef]
  91. Kobylecki, C.; Cenci, M.A.; Crossman, A.R.; Ravenscroft, P. Calcium-permeable AMPA receptors are involved in the induction and expression of l-DOPA-induced dyskinesia in Parkinson’s disease. J. Neurochem. 2010, 114, 499–511. [Google Scholar] [CrossRef] [PubMed]
  92. Luquin, M.R.; Obeso, J.A.; Laguna, J.; Guillen, J.; Martinez-Lage, J.M. The AMPA receptor antagonist NBQX does not alter the motor response induced by selective dopamine agonists in MPTP-treated monkeys. Eur. J. Pharmacol. 1993, 235, 297–300. [Google Scholar] [CrossRef] [PubMed]
  93. Bourque, M.; Gregoire, L.; Patel, W.; Dickens, D.; Snodgrass, R.; Di Paolo, T. AV-101, a Pro-Drug Antagonist at the NMDA Receptor Glycine Site, Reduces L-Dopa Induced Dyskinesias in MPTP Monkeys. Cells 2022, 11, 3530. [Google Scholar] [CrossRef] [PubMed]
  94. Frouni, I.; Belliveau, S.; Maddaford, S.; Nuara, S.G.; Gourdon, J.C.; Huot, P. Effect of the glycine transporter 1 inhibitor ALX-5407 on dyskinesia, psychosis-like behaviours and parkinsonism in the MPTP-lesioned marmoset. Eur. J. Pharmacol. 2021, 910, 174452. [Google Scholar] [CrossRef]
  95. Frouni, I.; Kang, W.; Bedard, D.; Belliveau, S.; Kwan, C.; Hadj-Youssef, S.; Bourgeois-Cayer, E.; Ohlund, L.; Sleno, L.; Hamadjida, A.; et al. Effect of glycine transporter 1 inhibition with bitopertin on parkinsonism and L-DOPA induced dyskinesia in the 6-OHDA-lesioned rat. Eur. J. Pharmacol. 2022, 929, 175090. [Google Scholar] [CrossRef]
  96. Papa, S.M.; Auberson, Y.P.; Greenamyre, J.T. Prolongation of levodopa responses by glycineB antagonists in parkinsonian primates. Ann. Neurol. 2004, 56, 723–727. [Google Scholar] [CrossRef] [PubMed]
  97. Boldry, R.C.; Papa, S.M.; Kask, A.M.; Chase, T.N. MK-801 reverses effects of chronic levodopa on D1 and D2 dopamine agonist-induced rotational behavior. Brain Res. 1995, 692, 259–264. [Google Scholar] [CrossRef] [PubMed]
  98. Tronci, E.; Fidalgo, C.; Zianni, E.; Collu, M.; Stancampiano, R.; Morelli, M.; Gardoni, F.; Carta, M. Effect of memantine on L-DOPA-induced dyskinesia in the 6-OHDA-lesioned rat model of Parkinson’s disease. Neuroscience 2014, 265, 245–252. [Google Scholar] [CrossRef]
  99. Blanchet, P.J.; Konitsiotis, S.; Chase, T.N. Amantadine reduces levodopa-induced dyskinesias in parkinsonian monkeys. Mov. Disord. 1998, 13, 798–802. [Google Scholar] [CrossRef]
  100. Blanchet, P.J.; Papa, S.M.; Metman, L.V.; Mouradian, M.M.; Chase, T.N. Modulation of levodopa-induced motor response complications by NMDA antagonists in Parkinson’s disease. Neurosci. Biobehav. Rev. 1997, 21, 447–453. [Google Scholar] [CrossRef]
  101. Liu, C.T.; Kao, L.T.; Shih, J.H.; Chien, W.C.; Chiu, C.H.; Ma, K.H.; Huang, Y.S.; Cheng, C.Y.; Shiue, C.Y.; Li, I.H. The effect of dextromethorphan use in Parkinson’s disease: A 6-hydroxydopamine rat model and population-based study. Eur. J. Pharmacol. 2019, 862, 172639. [Google Scholar] [CrossRef] [PubMed]
  102. Zheng, C.; Xu, Y.; Chen, G.; Tan, Y.; Zeng, W.; Wang, J.; Cheng, C.; Yang, X.; Nie, S.; Zhang, Z.; et al. Distinct anti-dyskinetic effects of amantadine and group II metabotropic glutamate receptor agonist LY354740 in a rodent model: An electrophysiological perspective. Neurobiol. Dis. 2020, 139, 104807. [Google Scholar] [CrossRef] [PubMed]
  103. Addy, C.; Assaid, C.; Hreniuk, D.; Stroh, M.; Xu, Y.; Herring, W.J.; Ellenbogen, A.; Jinnah, H.A.; Kirby, L.; Leibowitz, M.T.; et al. Single-dose administration of MK-0657, an NR2B-selective NMDA antagonist, does not result in clinically meaningful improvement in motor function in patients with moderate Parkinson’s disease. J. Clin. Pharmacol. 2009, 49, 856–864. [Google Scholar] [CrossRef] [PubMed]
  104. Nutt, J.G.; Gunzler, S.A.; Kirchhoff, T.; Hogarth, P.; Weaver, J.L.; Krams, M.; Jamerson, B.; Menniti, F.S.; Landen, J.W. Effects of a NR2B selective NMDA glutamate antagonist, CP-101,606, on dyskinesia and Parkinsonism. Mov. Disord. 2008, 23, 1860–1866. [Google Scholar] [CrossRef] [PubMed]
  105. Nash, J.E.; Ravenscroft, P.; McGuire, S.; Crossman, A.R.; Menniti, F.S.; Brotchie, J.M. The NR2B-selective NMDA receptor antagonist CP-101,606 exacerbates L-DOPA-induced dyskinesia and provides mild potentiation of anti-parkinsonian effects of L-DOPA in the MPTP-lesioned marmoset model of Parkinson’s disease. Exp. Neurol. 2004, 188, 471–479. [Google Scholar] [CrossRef] [PubMed]
  106. Wessell, R.H.; Ahmed, S.M.; Menniti, F.S.; Dunbar, G.L.; Chase, T.N.; Oh, J.D. NR2B selective NMDA receptor antagonist CP-101,606 prevents levodopa-induced motor response alterations in hemi-parkinsonian rats. Neuropharmacology 2004, 47, 184–194. [Google Scholar] [CrossRef] [PubMed]
  107. Lujan, R.; Nusser, Z.; Roberts, J.D.; Shigemoto, R.; Somogyi, P. Perisynaptic location of metabotropic glutamate receptors mGluR1 and mGluR5 on dendrites and dendritic spines in the rat hippocampus. Eur. J. Neurosci. 1996, 8, 1488–1500. [Google Scholar] [CrossRef] [PubMed]
  108. Shigemoto, R.; Nomura, S.; Ohishi, H.; Sugihara, H.; Nakanishi, S.; Mizuno, N. Immunohistochemical localization of a metabotropic glutamate receptor, mGluR5, in the rat brain. Neurosci. Lett. 1993, 163, 53–57. [Google Scholar] [CrossRef]
  109. Pisani, A.; Gubellini, P.; Bonsi, P.; Conquet, F.; Picconi, B.; Centonze, D.; Bernardi, G.; Calabresi, P. Metabotropic glutamate receptor 5 mediates the potentiation of N-methyl-D-aspartate responses in medium spiny striatal neurons. Neuroscience 2001, 106, 579–587. [Google Scholar] [CrossRef]
  110. Shoulson, I.; Penney, J.; McDermott, M.; Schwid, S.; Kayson, E.; Chase, T.; Fahn, S.; Greenamyre, J.T.; Lang, A.; Siderowf, A.; et al. A randomized, controlled trial of remacemide for motor fluctuations in Parkinson’s disease. Neurology 2001, 56, 455–462. [Google Scholar] [CrossRef]
  111. Eggert, K.; Squillacote, D.; Barone, P.; Dodel, R.; Katzenschlager, R.; Emre, M.; Lees, A.J.; Rascol, O.; Poewe, W.; Tolosa, E.; et al. Safety and efficacy of perampanel in advanced Parkinson’s disease: A randomized, placebo-controlled study. Mov. Disord. 2010, 25, 896–905. [Google Scholar] [CrossRef] [PubMed]
  112. Rylander, D.; Recchia, A.; Mela, F.; Dekundy, A.; Danysz, W.; Cenci, M.A. Pharmacological modulation of glutamate transmission in a rat model of L-DOPA-induced dyskinesia: Effects on motor behavior and striatal nuclear signaling. J. Pharmacol. Exp. Ther. 2009, 330, 227–235. [Google Scholar] [CrossRef] [PubMed]
  113. Morin, N.; Gregoire, L.; Gomez-Mancilla, B.; Gasparini, F.; Di Paolo, T. Effect of the metabotropic glutamate receptor type 5 antagonists MPEP and MTEP in parkinsonian monkeys. Neuropharmacology 2010, 58, 981–986. [Google Scholar] [CrossRef] [PubMed]
  114. McFarthing, K.; Prakash, N.; Simuni, T. Clinical Trial Highlights—Dyskinesia. J. Parkinsons Dis. 2019, 9, 449–465. [Google Scholar] [CrossRef] [PubMed]
  115. Charvin, D.; Di Paolo, T.; Bezard, E.; Gregoire, L.; Takano, A.; Duvey, G.; Pioli, E.; Halldin, C.; Medori, R.; Conquet, F. An mGlu4-Positive Allosteric Modulator Alleviates Parkinsonism in Primates. Mov. Disord. 2018, 33, 1619–1631. [Google Scholar] [CrossRef]
  116. Iderberg, H.; Maslava, N.; Thompson, A.D.; Bubser, M.; Niswender, C.M.; Hopkins, C.R.; Lindsley, C.W.; Conn, P.J.; Jones, C.K.; Cenci, M.A. Pharmacological stimulation of metabotropic glutamate receptor type 4 in a rat model of Parkinson’s disease and L-DOPA-induced dyskinesia: Comparison between a positive allosteric modulator and an orthosteric agonist. Neuropharmacology 2015, 95, 121–129. [Google Scholar] [CrossRef] [PubMed]
  117. Le Poul, E.; Bolea, C.; Girard, F.; Poli, S.; Charvin, D.; Campo, B.; Bortoli, J.; Bessif, A.; Luo, B.; Koser, A.J.; et al. A potent and selective metabotropic glutamate receptor 4 positive allosteric modulator improves movement in rodent models of Parkinson’s disease. J. Pharmacol. Exp. Ther. 2012, 343, 167–177. [Google Scholar] [CrossRef]
  118. Rascol, O.; Medori, R.; Baayen, C.; Such, P.; Meulien, D.; Group, A.S. A Randomized, Double-Blind, Controlled Phase II Study of Foliglurax in Parkinson’s Disease. Mov. Disord. 2022, 37, 1088–1093. [Google Scholar] [CrossRef]
  119. Bezard, E.; Pioli, E.Y.; Li, Q.; Girard, F.; Mutel, V.; Keywood, C.; Tison, F.; Rascol, O.; Poli, S.M. The mGluR5 negative allosteric modulator dipraglurant reduces dyskinesia in the MPTP macaque model. Mov. Disord. 2014, 29, 1074–1079. [Google Scholar] [CrossRef]
  120. Gregoire, L.; Morin, N.; Ouattara, B.; Gasparini, F.; Bilbe, G.; Johns, D.; Vranesic, I.; Sahasranaman, S.; Gomez-Mancilla, B.; Di Paolo, T. The acute antiparkinsonian and antidyskinetic effect of AFQ056, a novel metabotropic glutamate receptor type 5 antagonist, in L-Dopa-treated parkinsonian monkeys. Parkinsonism Relat. Disord. 2011, 17, 270–276. [Google Scholar] [CrossRef]
  121. Conn, P.J.; Battaglia, G.; Marino, M.J.; Nicoletti, F. Metabotropic glutamate receptors in the basal ganglia motor circuit. Nat. Rev. Neurosci. 2005, 6, 787–798. [Google Scholar] [CrossRef] [PubMed]
  122. Pisani, A.; Calabresi, P.; Centonze, D.; Bernardi, G. Enhancement of NMDA responses by group I metabotropic glutamate receptor activation in striatal neurones. Br. J. Pharmacol. 1997, 120, 1007–1014. [Google Scholar] [CrossRef] [PubMed]
  123. Cuomo, D.; Martella, G.; Barabino, E.; Platania, P.; Vita, D.; Madeo, G.; Selvam, C.; Goudet, C.; Oueslati, N.; Pin, J.P.; et al. Metabotropic glutamate receptor subtype 4 selectively modulates both glutamate and GABA transmission in the striatum: Implications for Parkinson’s disease treatment. J. Neurochem. 2009, 109, 1096–1105. [Google Scholar] [CrossRef]
  124. Jones, C.K.; Bubser, M.; Thompson, A.D.; Dickerson, J.W.; Turle-Lorenzo, N.; Amalric, M.; Blobaum, A.L.; Bridges, T.M.; Morrison, R.D.; Jadhav, S.; et al. The metabotropic glutamate receptor 4-positive allosteric modulator VU0364770 produces efficacy alone and in combination with L-DOPA or an adenosine 2A antagonist in preclinical rodent models of Parkinson’s disease. J. Pharmacol. Exp. Ther. 2012, 340, 404–421. [Google Scholar] [CrossRef]
  125. Rascol, O.; Fabbri, M.; Poewe, W. Amantadine in the treatment of Parkinson’s disease and other movement disorders. Lancet Neurol. 2021, 20, 1048–1056. [Google Scholar] [CrossRef] [PubMed]
  126. Kim, J.H.; Lee, H.W.; Hwang, J.; Kim, J.; Lee, M.J.; Han, H.S.; Lee, W.H.; Suk, K. Microglia-inhibiting activity of Parkinson’s disease drug amantadine. Neurobiol. Aging 2012, 33, 2145–2159. [Google Scholar] [CrossRef] [PubMed]
  127. Ocal, O.; Cosar, A.; Naziroglu, M. Amantadine Attenuated Hypoxia-Induced Mitochondrial Oxidative Neurotoxicity, Apoptosis, and Inflammation via the Inhibition of TRPM2 and TRPV4 Channels. Mol. Neurobiol. 2022, 59, 3703–3720. [Google Scholar] [CrossRef]
  128. Tison, F.; Keywood, C.; Wakefield, M.; Durif, F.; Corvol, J.C.; Eggert, K.; Lew, M.; Isaacson, S.; Bezard, E.; Poli, S.M.; et al. A Phase 2A Trial of the Novel mGluR5-Negative Allosteric Modulator Dipraglurant for Levodopa-Induced Dyskinesia in Parkinson’s Disease. Mov. Disord. 2016, 31, 1373–1380. [Google Scholar] [CrossRef]
  129. Trenkwalder, C.; Stocchi, F.; Poewe, W.; Dronamraju, N.; Kenney, C.; Shah, A.; von Raison, F.; Graf, A. Mavoglurant in Parkinson’s patients with l-Dopa-induced dyskinesias: Two randomized phase 2 studies. Mov. Disord. 2016, 31, 1054–1058. [Google Scholar] [CrossRef]
  130. Wang, W.W.; Zhang, X.R.; Zhang, Z.R.; Wang, X.S.; Chen, J.; Chen, S.Y.; Xie, C.L. Effects of mGluR5 Antagonists on Parkinson’s Patients With L-Dopa-Induced Dyskinesia: A Systematic Review and Meta-Analysis of Randomized Controlled Trials. Front. Aging Neurosci. 2018, 10, 262. [Google Scholar] [CrossRef]
  131. Bollen, E.; Prickaerts, J. Phosphodiesterases in neurodegenerative disorders. IUBMB Life 2012, 64, 965–970. [Google Scholar] [CrossRef] [PubMed]
  132. Menniti, F.S.; Faraci, W.S.; Schmidt, C.J. Phosphodiesterases in the CNS: Targets for drug development. Nat. Rev. Drug Discov. 2006, 5, 660–670. [Google Scholar] [CrossRef] [PubMed]
  133. Lakics, V.; Karran, E.H.; Boess, F.G. Quantitative comparison of phosphodiesterase mRNA distribution in human brain and peripheral tissues. Neuropharmacology 2010, 59, 367–374. [Google Scholar] [CrossRef] [PubMed]
  134. Belmaker, R.H.; Ebstein, R.P.; Biederman, J.; Stern, R.; Berman, M.; van Praag, H.M. The effect of L-dopa and propranolol on human CSF cyclic nucleotides. Psychopharmacology 1978, 58, 307–310. [Google Scholar] [CrossRef]
  135. Volicer, L.; Beal, M.F.; Direnfeld, L.K.; Marquis, J.K.; Albert, M.L. CSF cyclic nucleotides and somatostatin in Parkinson’s disease. Neurology 1986, 36, 89–92. [Google Scholar] [CrossRef] [PubMed]
  136. Xie, Z.; Adamowicz, W.O.; Eldred, W.D.; Jakowski, A.B.; Kleiman, R.J.; Morton, D.G.; Stephenson, D.T.; Strick, C.A.; Williams, R.D.; Menniti, F.S. Cellular and subcellular localization of PDE10A, a striatum-enriched phosphodiesterase. Neuroscience 2006, 139, 597–607. [Google Scholar] [CrossRef] [PubMed]
  137. Coskran, T.M.; Morton, D.; Menniti, F.S.; Adamowicz, W.O.; Kleiman, R.J.; Ryan, A.M.; Strick, C.A.; Schmidt, C.J.; Stephenson, D.T. Immunohistochemical localization of phosphodiesterase 10A in multiple mammalian species. J. Histochem. Cytochem. 2006, 54, 1205–1213. [Google Scholar] [CrossRef]
  138. Nishi, A.; Snyder, G.L. Advanced research on dopamine signaling to develop drugs for the treatment of mental disorders: Biochemical and behavioral profiles of phosphodiesterase inhibition in dopaminergic neurotransmission. J. Pharmacol. Sci. 2010, 114, 6–16. [Google Scholar] [CrossRef]
  139. Giorgi, M.; Melchiorri, G.; Nuccetelli, V.; D’Angelo, V.; Martorana, A.; Sorge, R.; Castelli, V.; Bernardi, G.; Sancesario, G. PDE10A and PDE10A-dependent cAMP catabolism are dysregulated oppositely in striatum and nucleus accumbens after lesion of midbrain dopamine neurons in rat: A key step in parkinsonism physiopathology. Neurobiol. Dis. 2011, 43, 293–303. [Google Scholar] [CrossRef]
  140. Niccolini, F.; Haider, S.; Reis Marques, T.; Muhlert, N.; Tziortzi, A.C.; Searle, G.E.; Natesan, S.; Piccini, P.; Kapur, S.; Rabiner, E.A.; et al. Altered PDE10A expression detectable early before symptomatic onset in Huntington’s disease. Brain 2015, 138, 3016–3029. [Google Scholar] [CrossRef]
  141. Arakawa, K.; Yuge, N.; Maehara, S. Ameliorative effects of a phosphodiesterase 10A inhibitor, MR1916 on L-DOPA-induced dyskinesia in parkinsonian rats. Pharmacol. Rep. 2020, 72, 443–448. [Google Scholar] [CrossRef] [PubMed]
  142. Guimaraes, R.P.; Ribeiro, D.L.; Dos Santos, K.B.; Talarico, C.H.Z.; Godoy, L.D.; Padovan-Neto, F.E. Phosphodiesterase 10A Inhibition Modulates the Corticostriatal Activity and L-DOPA-Induced Dyskinesia. Pharmaceuticals 2022, 1, 947. [Google Scholar] [CrossRef] [PubMed]
  143. Polito, M.; Guiot, E.; Gangarossa, G.; Longueville, S.; Doulazmi, M.; Valjent, E.; Herve, D.; Girault, J.A.; Paupardin-Tritsch, D.; Castro, L.R.; et al. Selective Effects of PDE10A Inhibitors on Striatopallidal Neurons Require Phosphatase Inhibition by DARPP-32. eNeuro 2015, 2, ENEURO.0060-0015.2015. [Google Scholar] [CrossRef]
  144. Wilson, J.M.; Ogden, A.M.; Loomis, S.; Gilmour, G.; Baucum, A.J., 2nd; Belecky-Adams, T.L.; Merchant, K.M. Phosphodiesterase 10A inhibitor, MP-10 (PF-2545920), produces greater induction of c-Fos in dopamine D2 neurons than in D1 neurons in the neostriatum. Neuropharmacology 2015, 99, 379–386. [Google Scholar] [CrossRef] [PubMed]
  145. Schmidt, C.J.; Chapin, D.S.; Cianfrogna, J.; Corman, M.L.; Hajos, M.; Harms, J.F.; Hoffman, W.E.; Lebel, L.A.; McCarthy, S.A.; Nelson, F.R.; et al. Preclinical characterization of selective phosphodiesterase 10A inhibitors: A new therapeutic approach to the treatment of schizophrenia. J. Pharmacol. Exp. Ther. 2008, 325, 681–690. [Google Scholar] [CrossRef] [PubMed]
  146. Uthayathas, S.; Masilamoni, G.J.; Shaffer, C.L.; Schmidt, C.J.; Menniti, F.S.; Papa, S.M. Phosphodiesterase 10A inhibitor MP-10 effects in primates: Comparison with risperidone and mechanistic implications. Neuropharmacology 2014, 77, 257–267. [Google Scholar] [CrossRef] [PubMed]
  147. Cenci, M.A.; Jorntell, H.; Petersson, P. On the neuronal circuitry mediating L-DOPA-induced dyskinesia. J. Neural Transm. 2018, 125, 1157–1169. [Google Scholar] [CrossRef] [PubMed]
  148. Enomoto, T.; Nakako, T.; Goda, M.; Wada, E.; Kitamura, A.; Fujii, Y.; Ikeda, K. A novel phosphodiesterase 1 inhibitor reverses L-dopa-induced dyskinesia, but not motivation deficits, in monkeys. Pharmacol. Biochem. Behav. 2021, 205, 173183. [Google Scholar] [CrossRef]
  149. Kakkar, R.; Raju, R.V.; Rajput, A.H.; Sharma, R.K. Amantadine: An antiparkinsonian agent inhibits bovine brain 60 kDa calmodulin-dependent cyclic nucleotide phosphodiesterase isozyme. Brain Res. 1997, 749, 290–294. [Google Scholar] [CrossRef]
  150. Kakkar, R.; Raju, R.V.; Rajput, A.H.; Sharma, R.K. Inhibition of bovine brain calmodulin-dependent cyclic nucleotide phosphodiesterase isozymes by deprenyl. Life Sci. 1996, 59, PL337–PL341. [Google Scholar] [CrossRef]
  151. Polli, J.W.; Kincaid, R.L. Expression of a calmodulin-dependent phosphodiesterase isoform (PDE1B1) correlates with brain regions having extensive dopaminergic innervation. J. Neurosci. 1994, 14, 1251–1261. [Google Scholar] [CrossRef] [PubMed]
  152. Reed, T.M.; Repaske, D.R.; Snyder, G.L.; Greengard, P.; Vorhees, C.V. Phosphodiesterase 1B knock-out mice exhibit exaggerated locomotor hyperactivity and DARPP-32 phosphorylation in response to dopamine agonists and display impaired spatial learning. J. Neurosci. 2002, 22, 5188–5197. [Google Scholar] [CrossRef] [PubMed]
  153. Doummar, D.; Dentel, C.; Lyautey, R.; Metreau, J.; Keren, B.; Drouot, N.; Malherbe, L.; Bouilleret, V.; Courraud, J.; Valenti-Hirsch, M.P.; et al. Biallelic PDE2A variants: A new cause of syndromic paroxysmal dyskinesia. Eur. J. Hum. Genet. 2020, 28, 1403–1413. [Google Scholar] [CrossRef] [PubMed]
  154. Heckman, P.R.A.; Blokland, A.; Prickaerts, J. From Age-Related Cognitive Decline to Alzheimer’s Disease: A Translational Overview of the Potential Role for Phosphodiesterases. Adv. Neurobiol. 2017, 17, 135–168. [Google Scholar] [CrossRef] [PubMed]
  155. Salpietro, V.; Perez-Duenas, B.; Nakashima, K.; San Antonio-Arce, V.; Manole, A.; Efthymiou, S.; Vandrovcova, J.; Bettencourt, C.; Mencacci, N.E.; Klein, C.; et al. A homozygous loss-of-function mutation in PDE2A associated to early-onset hereditary chorea. Mov. Disord. 2018, 33, 482–488. [Google Scholar] [CrossRef] [PubMed]
  156. Simpson, E.H.; Kellendonk, C.; Kandel, E. A possible role for the striatum in the pathogenesis of the cognitive symptoms of schizophrenia. Neuron 2010, 65, 585–596. [Google Scholar] [CrossRef]
  157. Stephenson, D.T.; Coskran, T.M.; Wilhelms, M.B.; Adamowicz, W.O.; O’Donnell, M.M.; Muravnick, K.B.; Menniti, F.S.; Kleiman, R.J.; Morton, D. Immunohistochemical localization of phosphodiesterase 2A in multiple mammalian species. J. Histochem. Cytochem. 2009, 57, 933–949. [Google Scholar] [CrossRef] [PubMed]
  158. Casacchia, M.; Meco, G.; Castellana, F.; Bedini, L.; Cusimano, G.; Agnoli, A. Therapeutic use of a selective cAMP phosphodiesterase inhibitor (Rolipram) in Parkinson’s disease. Pharmacol. Res. Commun. 1983, 15, 329–334. [Google Scholar] [CrossRef]
  159. Cherry, J.A.; Davis, R.L. Cyclic AMP phosphodiesterases are localized in regions of the mouse brain associated with reinforcement, movement, and affect. J. Comp. Neurol. 1999, 407, 287–301. [Google Scholar] [CrossRef]
  160. Perez-Torres, S.; Miro, X.; Palacios, J.M.; Cortes, R.; Puigdomenech, P.; Mengod, G. Phosphodiesterase type 4 isozymes expression in human brain examined by in situ hybridization histochemistry and [3H]rolipram binding autoradiography. Comparison with monkey and rat brain. J. Chem. Neuroanat. 2000, 20, 349–374. [Google Scholar] [CrossRef]
  161. Vignola, A.M. PDE4 inhibitors in COPD—A more selective approach to treatment. Respir. Med. 2004, 98, 495–503. [Google Scholar] [CrossRef] [PubMed]
  162. Yang, L.; Calingasan, N.Y.; Lorenzo, B.J.; Beal, M.F. Attenuation of MPTP neurotoxicity by rolipram, a specific inhibitor of phosphodiesterase IV. Exp. Neurol. 2008, 211, 311–314. [Google Scholar] [CrossRef] [PubMed]
  163. Ciccocioppo, R.; Guglielmo, G.; Li, H.; Melis, M.; Caffino, L.; Shen, Q.; Domi, A.; Fumagalli, F.; Demopulos, G.A.; Gaitanaris, G.A. Selective inhibition of phosphodiesterase 7 enzymes reduces motivation for nicotine use through modulation of mesolimbic dopaminergic transmission. J. Neurosci. 2021, 41, 6128–6143. [Google Scholar] [CrossRef]
  164. Morales-Garcia, J.A.; Alonso-Gil, S.; Gil, C.; Martinez, A.; Santos, A.; Perez-Castillo, A. Phosphodiesterase 7 inhibition induces dopaminergic neurogenesis in hemiparkinsonian rats. Stem Cells Transl. Med. 2015, 4, 564–575. [Google Scholar] [CrossRef]
  165. Morales-Garcia, J.A.; Alonso-Gil, S.; Santos, A.; Perez-Castillo, A. Phosphodiesterase 7 Regulation in Cellular and Rodent Models of Parkinson’s Disease. Mol. Neurobiol. 2020, 57, 806–822. [Google Scholar] [CrossRef] [PubMed]
  166. Morales-Garcia, J.A.; Redondo, M.; Alonso-Gil, S.; Gil, C.; Perez, C.; Martinez, A.; Santos, A.; Perez-Castillo, A. Phosphodiesterase 7 inhibition preserves dopaminergic neurons in cellular and rodent models of Parkinson disease. PLoS ONE 2011, 6, e17240. [Google Scholar] [CrossRef] [PubMed]
  167. Golkowski, M.; Shimizu-Albergine, M.; Suh, H.W.; Beavo, J.A.; Ong, S.E. Studying mechanisms of cAMP and cyclic nucleotide phosphodiesterase signaling in Leydig cell function with phosphoproteomics. Cell Signal. 2016, 28, 764–778. [Google Scholar] [CrossRef] [PubMed]
  168. Kobayashi, T.; Gamanuma, M.; Sasaki, T.; Yamashita, Y.; Yuasa, K.; Kotera, J.; Omori, K. Molecular comparison of rat cyclic nucleotide phosphodiesterase 8 family: Unique expression of PDE8B in rat brain. Gene 2003, 319, 21–31. [Google Scholar] [CrossRef]
  169. Wu, X.N.; Zhou, Q.; Huang, Y.D.; Xie, X.; Li, Z.; Wu, Y.; Luo, H.B. Structure-based discovery of orally efficient inhibitors via unique interactions with H-pocket of PDE8 for the treatment of vascular dementia. Acta Pharm. Sin. B 2022, 12, 3103–3112. [Google Scholar] [CrossRef]
  170. Lenda, T.; Ossowska, K.; Berghauzen-Maciejewska, K.; Matloka, M.; Pieczykolan, J.; Wieczorek, M.; Konieczny, J. Antiparkinsonian-like effects of CPL500036, a novel selective inhibitor of phosphodiesterase 10A, in the unilateral rat model of Parkinson’s disease. Eur. J. Pharmacol. 2021, 910, 174460. [Google Scholar] [CrossRef]
  171. Cenci, M.A.; Crossman, A.R. Animal models of l-dopa-induced dyskinesia in Parkinson’s disease. Mov. Disord. 2018, 33, 889–899. [Google Scholar] [CrossRef] [PubMed]
  172. Sonnenburg, W.K.; Mullaney, P.J.; Beavo, J.A. Molecular cloning of a cyclic GMP-stimulated cyclic nucleotide phosphodiesterase cDNA. Identification and distribution of isozyme variants. J. Biol. Chem. 1991, 266, 17655–17661. [Google Scholar] [CrossRef] [PubMed]
  173. Ramirez, A.D.; Smith, S.M. Regulation of dopamine signaling in the striatum by phosphodiesterase inhibitors: Novel therapeutics to treat neurological and psychiatric disorders. Cent. Nerv. Syst. Agents Med. Chem. 2014, 14, 72–82. [Google Scholar] [CrossRef] [PubMed]
  174. Azevedo, M.F.; Faucz, F.R.; Bimpaki, E.; Horvath, A.; Levy, I.; de Alexandre, R.B.; Ahmad, F.; Manganiello, V.; Stratakis, C.A. Clinical and molecular genetics of the phosphodiesterases (PDEs). Endocr. Rev. 2014, 35, 195–233. [Google Scholar] [CrossRef] [PubMed]
  175. Reyes-Irisarri, E.; Markerink-Van Ittersum, M.; Mengod, G.; de Vente, J. Expression of the cGMP-specific phosphodiesterases 2 and 9 in normal and Alzheimer’s disease human brains. Eur. J. Neurosci. 2007, 25, 3332–3338. [Google Scholar] [CrossRef] [PubMed]
  176. Van Staveren, W.C.; Steinbusch, H.W.; Markerink-Van Ittersum, M.; Repaske, D.R.; Goy, M.F.; Kotera, J.; Omori, K.; Beavo, J.A.; De Vente, J. mRNA expression patterns of the cGMP-hydrolyzing phosphodiesterases types 2, 5, and 9 during development of the rat brain. J. Comp. Neurol. 2003, 467, 566–580. [Google Scholar] [CrossRef] [PubMed]
  177. Yuan, H.; Myers, S.J.; Wells, G.; Nicholson, K.L.; Swanger, S.A.; Lyuboslavsky, P.; Tahirovic, Y.A.; Menaldino, D.S.; Ganesh, T.; Wilson, L.J.; et al. Context-dependent GluN2B-selective inhibitors of NMDA receptor function are neuroprotective with minimal side effects. Neuron 2015, 85, 1305–1318. [Google Scholar] [CrossRef]
  178. Kew, J.N.; Trube, G.; Kemp, J.A. State-dependent NMDA receptor antagonism by Ro 8-4304, a novel NR2B selective, non-competitive, voltage-independent antagonist. Br. J. Pharmacol. 1998, 123, 463–472. [Google Scholar] [CrossRef]
  179. Vicini, S.; Wang, J.F.; Li, J.H.; Zhu, W.J.; Wang, Y.H.; Luo, J.H.; Wolfe, B.B.; Grayson, D.R. Functional and pharmacological differences between recombinant N-methyl-D-aspartate receptors. J. Neurophysiol. 1998, 79, 555–566. [Google Scholar] [CrossRef]
  180. Garcia-Montes, J.R.; Solis, O.; Enriquez-Traba, J.; Ruiz-DeDiego, I.; Drucker-Colin, R.; Moratalla, R. Genetic Knockdown of mGluR5 in Striatal D1R-Containing Neurons Attenuates L-DOPA-Induced Dyskinesia in Aphakia Mice. Mol. Neurobiol. 2019, 56, 4037–4050. [Google Scholar] [CrossRef]
Figure 1. Microcircuitry of striatal projection neurons. The scheme is an oversimplification of the microcircuitry. DA stimulation modulates the activity of the d- and iSPNs. DAR signaling is mediated by activation of cyclic nucleotide synthesis through two pathways (AC-cAMP and NO-GC-cGMP). PDEs are catabolic enzymes for cAMP and cGMP (PDE10, PDE1, PDE2, PDE4, PDE7, and PDE9 are abundant in the striatum). SPNs receive excitatory signals mediated by glutamate ionotropic (NMDAR and AMPAR) and metabotropic receptors (groups I–III), the latter also acting presynaptically to modulate glutamate release. Arrows represent activation, and perpendicular lines represent inactivation. Abbreviations: SPN = striatal projection neuron; PDE = phosphodiesterase; NOS = nitric oxide synthase; NO = nitric oxide; GC = guanylate cyclase; cGMP = cyclic guanosine monophosphate; PKG = protein kinase G; AC = adenylate cyclase; cAMP = cyclic adenosine monophosphate; PKA = protein kinase A; DARPP-32 = dopamine- and cAMP-regulated neuronal phosphoprotein; CREB = cAMP response element binding protein.
Figure 1. Microcircuitry of striatal projection neurons. The scheme is an oversimplification of the microcircuitry. DA stimulation modulates the activity of the d- and iSPNs. DAR signaling is mediated by activation of cyclic nucleotide synthesis through two pathways (AC-cAMP and NO-GC-cGMP). PDEs are catabolic enzymes for cAMP and cGMP (PDE10, PDE1, PDE2, PDE4, PDE7, and PDE9 are abundant in the striatum). SPNs receive excitatory signals mediated by glutamate ionotropic (NMDAR and AMPAR) and metabotropic receptors (groups I–III), the latter also acting presynaptically to modulate glutamate release. Arrows represent activation, and perpendicular lines represent inactivation. Abbreviations: SPN = striatal projection neuron; PDE = phosphodiesterase; NOS = nitric oxide synthase; NO = nitric oxide; GC = guanylate cyclase; cGMP = cyclic guanosine monophosphate; PKG = protein kinase G; AC = adenylate cyclase; cAMP = cyclic adenosine monophosphate; PKA = protein kinase A; DARPP-32 = dopamine- and cAMP-regulated neuronal phosphoprotein; CREB = cAMP response element binding protein.
Cells 12 02754 g001
Table 1. Glutamate transmission mechanisms.
Table 1. Glutamate transmission mechanisms.
ReceptorFindingsSpeciesReferences
NMDARIncreased ratio of NMDARs/AMPARsRat

Bagetta et al., 2012 [42]
Overexpression of receptors in the striatumRat
NHP
Human
Chase et al., 2000 [43]
Calon et al., 2002 [44]
Calon et al., 2003 [45]
Higher ratio of GluN2A/GluN2B subunits after prolonged exposure to L-DOPARat
NHP
Human
Gardoni et al., 2006 [46]
Hallett et al., 2005 [47]
Mellone et al., 2015 [48]
Elevated NMDAR activation and glutamate releaseRat
NHP
Human
Nash et al., 2002 [49]
Calon et al., 2002 [44]
Ahmed et al., 2011 [50]
NMDAR antagonists reduce basal SPN hyperactivity and “unstable” responses to L-DOPARat
NHP

Papa et al., 1995 [51]
Singh et al., 2018 [14]
AMPAROverexpression of receptors in the striatumRat
NHP
Human
Bagetta et al., 2012 [42]
Calon et al., 2002 [44]
Calon et al., 2003 [45]
AMPAR antagonists reduce basal SPN hyperactivity and “unstable” responses to L-DOPARat
NHP

Fieblinger et al., 2014 [23]
Singh et al., 2018 [14]
Group I
mGluRs
mGluR5 levels correlate with LID development and severityRat
NHP
Human
Crabbe et al., 2018 [52]
Morin et al., 2013; Morin et al., 2013 [53,54]
Ouattara et al., 2011 [55]
mGluR5 NAMs attenuate nigrostriatal degenerationRat
NHP

Armentero et al., 2006 [56]
Masilamoni et al., 2011 [57]
mGluR5 levels are elevated in LIDsRat
NHP
Human
Crabbe et al., 2018 [52]
Sanchez-Pernaute et al., 2008 [58]
Ouattara et al., 2011 [55]
mGluR5 binding and expression in the striatum increase following L-DOPARat
NHP
Human
Crabbe et al., 2018 [52]
Samadi et al., 2008 [59]
Ouattara et al., 2011 [55]
mGluR5 NAMs reduce mGluR5 expressionRat
NHP

Lea et al., 2006 [60]
Morin et al., 2013 [54]
Group II
mGluRs
mGluR2/3 reduce excitatory drive on SPNs and LIDsRat

Pisani et al., 2000 [61]
Group III
mGluRs
mGluR4 reduces glutamate neurotransmissionRat
NHP

Calabrese et al., 2022 [62]
Charvin et al., 2018 [62]
mGluR4 PAM decreases spontaneous glutamatergic transmission and restores bidirectional plasticityRat

Calabrese et al., 2022 [62]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kochoian, B.A.; Bure, C.; Papa, S.M. Targeting Striatal Glutamate and Phosphodiesterases to Control L-DOPA-Induced Dyskinesia. Cells 2023, 12, 2754. https://doi.org/10.3390/cells12232754

AMA Style

Kochoian BA, Bure C, Papa SM. Targeting Striatal Glutamate and Phosphodiesterases to Control L-DOPA-Induced Dyskinesia. Cells. 2023; 12(23):2754. https://doi.org/10.3390/cells12232754

Chicago/Turabian Style

Kochoian, Brik A., Cassandra Bure, and Stella M. Papa. 2023. "Targeting Striatal Glutamate and Phosphodiesterases to Control L-DOPA-Induced Dyskinesia" Cells 12, no. 23: 2754. https://doi.org/10.3390/cells12232754

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop