Next Article in Journal
Synthesis, Herbicidal Activity, Crop Safety and Soil Degradation of Pyrimidine- and Triazine-Substituted Chlorsulfuron Derivatives
Next Article in Special Issue
A2A Adenosine Receptor Antagonists: Are Triazolotriazine and Purine Scaffolds Interchangeable?
Previous Article in Journal
Amphipolar, Amphiphilic 2,4-diarylpyrano[2,3-b]indoles as Turn-ON Luminophores in Acidic and Basic Media
Previous Article in Special Issue
Development of Bicyclo[3.1.0]hexane-Based A3 Receptor Ligands: Closing the Gaps in the Structure–Affinity Relationships
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Pharmacological Potential of Adenosine A2A Receptor Antagonists for Treating Parkinson’s Disease

1
Kyowa Kirin Co., Ltd., Tokyo 100-0004, Japan
2
Molecular Neuropharmacology Laboratory, Wenzhou Medical University, Wenzhou 325015, China
3
Kyowa Kirin International, Galashiels TD1 1QH, UK
4
Kyowa Kirin Inc., Bedminster, NJ 07921, USA
5
Institute of Pharmaceutical Science, Kings College London, London SE1 9NH, UK
*
Author to whom correspondence should be addressed.
Submission received: 16 February 2022 / Revised: 30 March 2022 / Accepted: 31 March 2022 / Published: 6 April 2022

Abstract

:
The adenosine A2A receptor subtype is recognized as a non-dopaminergic pharmacological target for the treatment of neurodegenerative disorders, notably Parkinson’s disease (PD). The selective A2A receptor antagonist istradefylline is approved in the US and Japan as an adjunctive treatment to levodopa/decarboxylase inhibitors in adults with PD experiencing OFF episodes or a wearing-off phenomenon; however, the full potential of this drug class remains to be explored. In this article, we review the pharmacology of adenosine A2A receptor antagonists from the perspective of the treatment of both motor and non-motor symptoms of PD and their potential for disease modification.

1. Introduction

The identification, characterization and classification of adenosine receptors have opened new vistas in pharmacology for treating a wide range of peripheral and central disease states [1,2,3,4,5]. This is particularly true of the potential use of centrally acting adenosine A2A receptor antagonists for the treatment of neurological and psychiatric illnesses where current therapies provide incomplete symptom relief and where there are no proven strategies that slow or prevent disease progression [6,7,8,9].
Adenosine receptors are widely expressed throughout the central nervous system and peripheral tissues or cells (for recent reviews, see Chen, 2014 [7]; Mori 2021 [10]) [7,10,11,12,13,14,15,16,17,18,19,20]. They belong to the family of G protein-coupled receptors (GPCRs), which possess seven transmembrane alpha helices, with three intracellular loops and three extracellular loops [11,21]. Adenosine receptors were initially cloned in the 1970s and 1980s, and then classified into four receptor subtypes based on their pharmacological profiles to inhibit (A1, A3) or stimulate (A2A, A2B) adenylate cyclase (AC) activity [11]. In particular, adenosine A2A receptors have a highly selective localization in brain [12] and are one of the best-characterized GPCRs [22,23]. As discussed in detail below, the selective localization of A2A receptors to the basal ganglia and the limbic brain region together with their ability to modulate the function of a range of other neurotransmitter systems offers a non-dopaminergic pharmacological approach to the treatment of Parkinson’s disease (PD), and potentially common neuropsychiatric disorders, such as depression and cognitive decline [4,6,9]. In addition, the positioning of A2A receptors on both neurons and glial cells and the ability to modulate pathogenic processes such as excessive glutamate release and the aggregation of toxic protein species identifies them as relevant to providing neuroprotective or disease modifying therapies in a range of neurodegenerative illnesses [17,20]. This has been demonstrated by the ability of A2A receptor antagonists to prevent neuronal death in preclinical models relevant to PD, Alzheimer’s disease (AD) and Huntington’s chorea [24,25,26,27].
In the following sections, we will describe in detail the rationale for the use of adenosine A2A antagonists, notably istradefylline for the treatment of PD, as well as the potential for treatment of a range of neuropsychiatric disorders and for application as a neuroprotective strategy. In addition, we will briefly expand on the range of degenerative disease states that might also be targets for therapy with A2A receptor antagonists.

2. Adenosine A2A Receptors and Parkinson’s Disease

To date, the exploitation of the central pharmacology of the A2A receptor has centered on the use of antagonist molecules for the treatment of PD, and for that reason, it will form the backbone of this review. However, since PD is characterized by both motor and non-motor symptoms, including a range of neuropsychiatric symptoms (depression, cognitive decline, sleep disturbance), this may hint at the potential of adenosinergic modulation as a target for central nervous system disorders in general [28,29,30]. A limitation with the current symptomatic therapies for PD is that they are almost entirely based on dopamine-replacement therapy (levodopa, dopamine agonists, enzymatic inhibitors) that shows a loss of efficacy (response fluctuations such as ‘wearing off’) with disease duration and severity, and where significant side effects occur (dyskinesia, hallucinations, impulse control disorders, sleep disturbance) [31]. In addition, most non-motor symptoms are only partially responsive to dopaminergic drugs, leaving a high degree of unmet need [8]. However, the widespread adaptive changes that occur within the basal ganglia as a result of the loss of dopaminergic neurons affect multiple neuronal systems that are non-dopaminergic in nature, offering potential new targets for treating both motor and non-motor symptoms [30,32].
In terms of other non-dopaminergic approaches, only the non-specific and weak N-methyl-D-aspartate (NMDA) antagonist amantadine and anticholinergic drugs such as benzhexol have been used in humans to treat motor symptoms, and both have undesirable adverse effects in PD [33,34]. More recently, the 5-HT2A inverse agonist pimavanserin has been approved for the treatment of psychosis in PD, supporting the potential for non-dopaminergic approaches [35]. However, despite considerable preclinical and clinical investigation, most non-dopaminergic drugs examined in clinical trials for improving motor symptoms in PD (serotoninergic drugs, noradrenergic antagonists, glutamate antagonists) have failed to show efficacy or have shown undesirable side effects [36,37]. It is for this reason that adenosine A2A receptor antagonists are of such interest as they may offer an alternative for treatment of motor and non-motor symptoms balanced with an acceptable adverse event profile and the potential for disease modifying activity.

2.1. Basal Ganglia Expression of A2A Receptors

The limited distribution of A2A receptors to the striatum, external globus pallidus (GPe), nucleus accumbens, and olfactory tubercle [38,39,40,41] has important implications for the treatment of PD as the most commonly occurring basal ganglia disorder [12,20,42]. Normal motor function is mediated in the striatum by the balance in activity between the direct and indirect GABAergic striatal output pathways, which in turn alters the activity of the striato-thalamo-cortical loops that control voluntary movement [43,44]. The direct output pathway (referred to as the ‘Go’ pathway) comprises spiny projection neurons (SPNs) that express dopamine D1 receptors and substance P [12,43,45]. The indirect output pathway (referred to as the ‘NoGo’ pathway) expresses dopamine D2 receptors and preproenkephalin [12,43,45]. Importantly, A2A receptors are located predominantly on the axon terminal and cell of the ‘NoGo’ pathway which projects from the striatum to the GPe, and they are segregated from the ‘Go’ pathway [12,13,46,47,48].
This unique targeted distribution is key to understanding the potential of A2A receptor antagonism in the management of PD because they are uniquely positioned to selectively modulate the activity of the ‘NoGo’ motor pathway at the level of the striatum and GPe [49,50]. Gs-coupled A2A receptors are positively coupled to AC [7] and thus to the 3′,5′-cyclic adenosine monophosphate (cAMP)/protein kinase A (PKA) pathway [7], whose activation facilitates GABA release in the GPe [14,51]. An A2A receptor agonist (CGS21680) or an AC activator significantly increased cAMP accumulation, subsequently activating the frequency of miniature inhibitory postsynaptic currents (mIPSCs) in GP slices [49]. CGS21680-induced cAMP accumulation was blocked by the adenosine A2A receptor selective antagonist KF17837, and CGS21680-induced enhancement of the mIPSCs frequency, which are striatal GABAergic outputs, was inhibited by AC or PKA inhibitors. Also, in the striatum, adenosine A2A receptor-induced inhibition of intrastriatal recurrent GABAergic activity, which cause excitation of the SPN cell body, has been revealed, and this phenomenon has been mimicked by cAMP analogues [52]. As a consequence, the entire GABAergic output activity (i.e., GABA release) of the SPNs is regulated by adenosine A2A receptor mediated signaling through cAMP/PKA-dependent mechanisms [12].
In PD, the progressive loss of nigrostriatal neurons causes a decline in dopaminergic transmission with the indirect pathway to GPe becoming overactive [15,16], resulting in an excessive ’NoGo’ signal from the GPe to the thalamus. A2A receptor activation also enhances the ‘NoGo’ signal while D2 receptor activation reduces the ‘NoGo’ signal [12]. While there remains much debate about the role of A2A-D2 chimeric receptors in modulating D2 receptor function [12,20,53,54], adenosine activation of A2A receptors opposes the effects of drugs that act through D2 receptors (levodopa, dopamine agonists) that are used to relieve the motor symptoms of PD, acting as a ‘brake’ on their effects. This implies that an adenosine A2A receptor antagonist should act to ‘release the brake’ and to increase the effects of dopaminergic medications (Figure 1). This is demonstrated in experimental models of PD by the ability of A2A antagonism to increase the improvement in motor performance produced by levodopa and dopamine agonists in 6-OHDA lesioned rats and in MPTP-treated primates [55,56,57].
The direct pathway expresses dopamine D1 receptors and relays motor signals from the dopaminergic neurons of the striatum through the globus pallidus internal (GPi) and substantia nigra pars reticulata (SNr) to relieve the inhibition of the thalamus and signal to the motor cortex to initiate movement. In this way, dopamine signaling in the direct pathway functions like the “gas” powering a car. The indirect pathway functions to control unwanted movement and functions like the “brake” in a car. Neurons of the indirect pathway express both dopamine D2 receptors as well as adenosine A2A receptors. Upon activation of the indirect pathway, signals relayed through the globus pallidus external (GPe) to the GPi/SNR complex via the subthalamic nucleus leads to thalamic inhibition and reduction in the motor output signals to the motor cortex. For the indirect pathway, dopamine D2 receptors suppress activation; in contrast, adenosine A2A receptors activate the pathway. In Parkinson’s disease, dopamine signaling to both the direct and indirect pathway is reduced, causing slow and dysregulated movements Furthermore, abnormal activation of the indirect pathway can be enhanced via adenosine A2A receptors overexpressed in progression of the disease, causing further dysfunction of movement.
In addition, the abundant expression of A2A receptors on the striatopallidal SPN [17] means that they are also uniquely positioned to integrate cortical glutamate signals and nigrostriatal/cortical dopamine signals to control striatal synaptic plasticity through a post-synaptic functional interaction with dopamine D2 receptors [58,59,60] and NMDA receptors [61,62], metabotropic glutamate 5 receptors [63,64,65], cannabinoid CB1 receptors [66,67], and presynaptic modulation of glutamate release at cortico-striatal projection terminals [68,69,70]. This complexity of activity may also contribute to the ability of adenosine A2A receptors to modulate motor function but also to exert actions relevant to non-motor symptoms of PD, such as cognition, which will be explored in subsequent sections of this review [71,72].
One other component of the expression of A2A receptors relevant to PD that requires mentioning is the increase in A2A receptor expression seen in the caudate nucleus/putamen and GPe of those with the disease at post-mortem and also in patients by PET imaging [71,73,74]. This upregulation of A2A receptor expression appears to occur early in the course of the illness, as it was found in the putamen of patients with Braak stage 1–2 disease [75]. The increased receptor expression was also more marked in those patients exhibiting dyskinesia. While it remains to be elucidated whether increased A2A receptor expression is a cause or effect of dyskinesia and/or drug treatment, these data imply that some functional change has occurred in adenosinergic signaling in the basal ganglia, and this is reflected in some experimental studies [12,76,77]. Dopamine depletion by mitochondrial neurotoxins such as 6-OHDA and MPTP increased ATP release, upregulated ecto-5′-nucleotidase (CD73) and A2A receptors in striatal synaptosomes [78] and in the striatum of MPTP-treated mice [79]. While the precise interpretation of these findings is not clear, preclinical data argue for an intervention in PD sooner rather than later using a selective adenosine A2A antagonist [80,81].

2.2. Relevance of A2A Receptor State to Mechanism of Action of Adenosinergic Agents in Managing the Motor Symptoms of Parkinson’s Disease

As discussed above, the therapeutic concept of A2A receptor antagonists developed to date is through receptor blockade to release D2 function suppressing ‘NoGo’ signals of the SPN. However, it is now increasingly understood that interpretation of the mode of action for A2A receptor agents in PD therapy should be discussed in terms of the receptor structure and formation to differentiate agonists, antagonists, and inverse agonists [81,82]. Several crystal structures of adenosine A2A receptors have been resolved in a high-affinity antagonist (ZM241385)-bound state [83] and in the agonist (UK-432097)-bound state [84]. Studies show that A2A agonists bind to the ligand binding pocket at the extracellular surface of adenosine A2A receptors where there are hydrophobic side chains required for ligand recognition) inside the third and seventh transmembrane helical domains (H3 and H7) [84]. Three distinct activation states of adenosine A2A receptors have been reported: the inactive state [83], the intermediately active state [84], and the active state [23]. Adenosine binds to the intermediate state and activates the adenosine A2A receptor through interactions with H3 and H7 (Ser277, His278) in the ligand-binding pocket [82]. Inverse agonists (e.g., ZM241385) reduce basal activity by binding to the inactive form of the adenosine A2A receptor [81]. Inverse agonists do not interact with H3 and H7; rather, they sterically prevent the conformational change in H5, thereby locking the receptor in its inactive state [81,82,85]. By contrast, neutral antagonists bind to the active state conformation of the adenosine A2A receptor and suppress the agonist binding without any structural hindrances, [83] and allosteric modulators bind to sites outside of the orthosteric ligand-binding site [81,82,85,86]. Recently, the crystal structure of adenosine A2A receptors bound to a partial agonist has also been resolved [87].
Adenosine A2A receptors are considered to exist in equilibrium between active and inactive forms in the two-state model of GPCRs [81]. (Figure 2). Whereas endogenous adenosine and A2A receptor agonists bind with higher affinity to the active form, inverse agonists have higher affinity for the inactive form—pushing the equilibrium towards the inactive state [12,76,77,80,81]. In the inactive state, adenosine A2A receptors possess constitutive activity even in the absence of adenosine [81] and A2A receptor inverse agonists act to further reduce cAMP accumulation generated by constitutive adenosine A2A receptor activity [12,81]. By contrast, neutral antagonists bind to active forms of the adenosine A2A receptor, thereby competitively antagonizing the effects of adenosine [12,81]. Thus, A2A receptor antagonists such as istradefylline act in PD by modulating the D2 receptor-bearing indirect pathway activity [12,81]. By contrast, inverse agonism not only releases D2 function but also directly affects neuronal activity in the same direction as the D2 receptor, manifesting as further reduced cAMP production in the indirect pathway [88]. Again, it is the functional interpretation of these data that is important. They suggest that, while an adenosine, A2A antagonist will antagonize the braking effect of adenosine on motor control via suppression of cAMP-dependent ‘NoGo’ pathway activity, it will not remove the constitutive receptor activity, which would continue to oppose the action of drugs acting through D2 dopamine receptors. In contrast, an inverse agonist would remove both the effects of adenosine on the A2A receptor and its constitutive activity, having greater efficacy in normalizing the activity of the ‘NoGo’ pathway in PD by further suppressing basal levels of pallidal GABA release and also by enhancing dopaminergic drug action. This all suggests that an inverse agonist would be more likely to have an effect on motor dysfunction in PD when used as monotherapy than occurs with a classical antagonist, such as istradefylline [12]. This idea is supported by a recent proof-of-concept study of the antagonist KW-6356, which also shows inverse agonist activity [89,90].
Endogenous ligand adenosine binds with higher affinity to the active form of Gs-coupled A2A receptors, and activates cAMP/PKA pathway via AC. This adenosine-induced cAMP accumulation activates PKA, following activation of multiple downstream targets, such as aPKC, CREB, and uORF5. Neutral antagonists bind to active forms of the adenosine A2A receptor, thereby competitively antagonizing adenosine binding. In the inactive state, adenosine A2A receptors possess constitutive activity with basal level production in cAMP, even in the absence of adenosine binding. Inverse agonists have higher affinity for the inactive form of the adenosine A2A receptor than the active—pushing the equilibrium towards the inactive state. Thus, inverse agonists act to further reduce cAMP accumulation from the basal level generated by constitutive adenosine A2A receptor activity. In addition, inverse agonists can also antagonize the effects of adenosine. A2AR: adenosine A2A receptor, AC: adenylate cyclase, aPKC: atypical protein kinase C, cAMP: 3′,5′-cyclic adenosine monophosphate, CREB: cAMP response element-binding protein, GABA: γ-aminobutyric acid, PKA: protein kinase A, uORF5: upstream open reading frame 5 + indicates activation of signal transduction in absence of ligand, indicates suppression of constitutive activity by inverse agonist.

3. Clinical Actions of Adenosine A2A Antagonists in Parkinson’s Disease

3.1. Motor Symptoms

The underlying science base for an effect of adenosine A2A antagonists on the motor symptoms of PD is strong and is based on the selective localization of the receptor and its functional significance to basal ganglia activity [12,20]. Taken along with the evidence from animal models of PD, an improvement of motor disability would be predicted when adenosine A2A antagonists are used in combination with levodopa therapy [32,91]. Not surprisingly, a number of selective adenosine A2A receptor antagonists have been developed [92] with the aim of increasing the duration of motor improvement seen during the time when symptoms are adequately controlled by medication (ON time) and decreasing the periods of inadequate control of PD symptoms (OFF time) in patients receiving levodopa but where insufficient efficacy is occurring [18,32,93]. Surprisingly, however, clinical development has proved challenging. Two candidate molecules, vipadenant and BIIB 014, have undergone only limited human testing [94,95,96], with vipadenant development discontinued due to toxicological concerns. Three other compounds—istradefylline [97,98,99,100,101,102,103,104], preladenant [105,106] and tozadenant [107]—have been examined in phase II and phase III clinical trials, and while positive efficacy results were obtained in all phase II studies, tozadenant was discontinued due to toxicity concerns [108], and only istradefylline has successfully reached commercialization [18]. To date, istradefylline has received approval in the US and Japan for the treatment of adult PD patients experiencing OFF time who are currently taking levodopa (plus a decarboxylase inhibitor) [109,110]. The efficacy of istradefylline was evaluated in a pooled analysis of eight multicenter, phase IIb/III double blind, placebo-controlled trials of patients experiencing motor fluctuations taking 20 or 40 mg/day of istradefylline [93]. The overall effect size in reducing OFF time (−0.38 to −0.82 h) was in the same range as those achieved by other adjuncts to levodopa therapy—such as MAO-B inhibitors and COMT-inhibitors [93].
There are many reasons why the clinical studies of istradefylline and other adenosine antagonists has been challenging—for example, trial design, trial population selection, placebo effects—and these have been discussed these in detail elsewhere [18,19]. The concepts of ON and OFF time and other motor endpoints are tailored to reflect the effects of dopaminergic replacement therapy, and there are no clinical endpoints available that fully capture the effects of non-dopaminergic approaches. It has also been suggested that the population of PD patients studied is an issue that may be key to the problems encountered in clinical development [19]. Istradefylline was taken into clinical trials mainly using later stage patients with PD experiencing significant ‘wearing off’ time but who were already treated with optimized dopaminergic therapy—in other words, further improvement could not be achieved using levodopa, dopamine agonists, MAO-B inhibitors or COMT inhibitors. Thus, the effects observed were ‘in addition’ to dopaminergic medications, reflecting the non-dopaminergic nature of the drug in an already difficult to treat patient population. It is important to stress that all the preclinical science shows that istradefylline has its maximal effect in models of PD when used in combination with low or threshold doses of levodopa and shows little increase in motor performance when used with optimized, high doses of levodopa. This suggests that a new generation of adenosine A2a receptor antagonists should be used at an earlier point in the progression of PD and in a manner where they are employed to be ‘levodopa sparing’ by avoiding further increases in levodopa dosage when the drug starts to exhibit a loss of efficacy. The other question that needs to be addressed is whether adenosine A2A antagonists would be effective as early monotherapy, and this has received little attention so far [111,112]. As discussed above, it may be more relevant to utilize an antagonist/inverse agonist for these studies as and when one becomes available for use in PD [113,114,115].

3.2. Non-Motor Symptoms

Non-motor symptoms of PD are a major area of unmet need in treating the illness and probably outweigh the importance of further improvements in the control of motor function [116,117]. A2A adenosine receptors are present in key ‘non-motor’ areas of the brain (e.g., nucleus accumbens, olfactory tubercle, amygdala, hippocampus, hypothalamus, thalamus, and cerebellum [17]) and play a role in the pathophysiology of sleep disturbances and mood disorders such as depression and anxiety which are common non-motor features of pre-motor and both early and late-stage clinical PD [28,118]. Importantly, preclinical investigation has delineated a role for adenosine signaling in attempting to address the development of novel approaches to their treatment [119,120,121]. To illustrate the progress being made, we will look at studies relevant to sleep–wake cycles, mood and cognition [119,120,121].

3.2.1. Role of A2A Receptors in Sleep–Wake Cycle Disturbance

Alterations to the sleep–wake cycle are a common non-motor symptom in PD and one that is often exacerbated by the use of dopaminergic therapies [122]. Nocturnal sleep disturbance is the most frequent manifestation, but excessive daytime somnolence also commonly occurs. Adenosine is a potent endogenous regulator of the sleep–wake cycle with adenosine levels increasing in the brain during periods of prolonged wakefulness and decreasing during sleep [123]. Excessive daytime somnolence in PD may be a result of over-activation of A2A receptors in the nucleus accumbens [32]. The potential for the use of adenosine antagonists in controlling disturbances in the sleep–wake cycle is shown by looking at the arousal effects of caffeine, a non-selective adenosine receptor antagonist. These are clearly mediated through the A2A receptor as caffeine’s ability to induce wakefulness is lost when A2A receptors are inactivated or deleted in the nucleus accumbens of mice but not when A1 receptor function is disrupted [120,122,124].
Two small open-label studies of istradefylline have investigated its impact on excessive daytime somnolence and nocturnal sleep. The first, an open-label study, evaluated patients over 12 weeks using the Epworth Sleepiness Scale (ESS), PD sleep scale (PDSS)-2, and PD Questionnaire (PDQ-8) at baseline and over a three-month period [125]. The study found that, in addition to improving motor scores, istradefylline decreased ESS scores; however, no impact on the PDSS-2 and PDQ-8 score were observed [125]. In a second study assessing istradefylline use over a month, patients taking istradefylline exhibited a decrease in ESS without a change in the total PD Sleep Scale (PDSS) [126]. These findings show that istradefylline may have an impact on sleep disturbances in PD patients, though additional large-scale studies are needed.

3.2.2. Role of A2A Receptors in Mood Disorders

Mood disorders are a common feature of PD, with depression affecting at least 35% of the patient population [111,112,127]. Depression can be secondary to the changes in quality of life brought on by a diagnosis of PD but may also occur directly due to the progression of pathological change that occurs as the illness advances [128,129]. Loss of dopaminergic input from the substantia nigra to the caudate nucleus and putamen disrupts the activity of the striato-thalamo-cortical loops, altering not only motor function but also the affective pathways controlling limbic and cortical function [32,130].
A role for A2A receptor activity in mediating symptoms of depression has been demonstrated using rodent models of depression (the forced swim test, tail suspension tests and learned helplessness) and examining the effects of adenosine A2A antagonist drugs [32,131,132]. For example, istradefylline decreased the immobility time in both the forced swim test and tail suspension tests [131]. When istradefylline was co-administered with fluoxetine, an additive effect was observed in reversing depressive behaviors [131]. Acute as well as chronic oral administration of istradefylline improved depression-like behavior in the rat learned-helplessness model [132]. The role of A2A receptors in anxiety in rodent models has also been investigated but the results are difficult to interpret [32]. A2A receptor knockout mice exhibit anxiety-like behaviors, but the effects seem dependent on the brain area investigated [133]. Some studies also report that A2A antagonists are able to ameliorate anxiety-like behaviors in rodents [117]. This is clearly an area relevant to mood changes in PD and where an adenosine A2A antagonist may be useful, but it is one that requires further study both at the preclinical level and evaluation in man.
Two small open-label studies have evaluated the efficacy of istradefylline for the treatment of depression or mood disorder in PD and are suggestive of some improvement. Patients with motor fluctuations not taking an antidepressant at baseline were evaluated using the Patient Health Questionnaire (PHQ) for depression. Eleven of fourteen patients had at least mild depression at baseline, and after 8 weeks of istradefylline (20–40 mg/day) treatment, five of fourteen patients showed an improvement in PHQ scores, with a trend to better improvement in those with worse (higher) PHQ scores at baseline [134]. In a second study, 30 patients with depressive symptoms not taking an antidepressant were studied at baseline and again after 12 weeks on istradefylline (20–40 mg/day). Significant improvement was seen in the Snaith–Hamilton Pleasure Scale Japanese version (SHAPS-J), Apathy Scale, and Beck Depression Inventory—2nd edition (BDI) [135]. Interestingly, while motor symptoms also improved with istradefylline administration, this was not correlated with the relief of depressive symptoms, indicating that A2A antagonism had an ameliorative effect on mood disorders independent of its effect on motor activity.

3.3. Role of A2A Receptors in Cognitive Impairment

There are multiple longitudinal studies showing an inverse relationship between consumption of caffeine and decreased memory impairments associated with aging as well as a reduced risk of developing dementia and AD [136,137,138,139,140,141]. Plasma levels of caffeine in subjects with mild cognitive impairment (MCI) who later progressed to dementia were lower than those whose cognition remained stable [142,143]. Furthermore, in a cross-sectional study in drug naïve PD patients, coffee drinking was significantly associated with a reduced severity of the mood/cognition domain of the non-motor symptom scale (NMSS) [144]. Accumulating evidence from preclinical studies in normal rodents and non-human primates also demonstrate that A2A receptor antagonists can enhance working memory [145], reversal learning [146], set-shifting [147], goal-directed behavior [148], and Pavlovian conditioning [133]. Consistent with increased A2A receptor expression that occurs in the brain in PD, recent studies have demonstrated that optogenetic stimulation of A2A receptor signaling [116] or transgenic overexpression of hippocampal A2A receptors [149] was sufficient to drive memory impairments in young animals. Overexpression also triggers HPA-axis dysfunction and a reduction in hippocampal glucocorticoid receptor levels [150].
Cognitive change occurs in both early PD impacting executive function (e.g., planning and decision making), working memory and procedural learning [151,152], and in late-stage illness with cognitive decline and frank dementia [152,153]. Early cognitive impairment in PD may relate to the dysfunction of the dopaminergic system and loss of LTP [29,154]. Dopamine replacement therapy may improve some aspects of cognition but not all while over-dosage may worsen cognition [155,156,157]. Dementia in PD is currently treated with cholinesterase inhibitors such as rivastigmine that provide only modest benefit [29,154]. Clearly there is a need for other pharmacological approaches to treatment, and epidemiological studies of caffeine consumption in aging populations suggest a role for adenosine receptors.
In rats, dopamine depletion in the prefrontal cortex (PFC) produces cognitive deficits, particularly associated with executive dysfunction [158,159], visuo-spatial, and non-spatial working memory that resemble those seen in PD [160]. Istradefylline reversed the cognitive deficits, and this was associated with increased dopamine levels in the PFC [161]. Istradefylline also improved spatial working memory in the MPTP-treated primate model, providing further preclinical data to suggest a potential benefit in man [162]. In rodents, A2A receptor antagonists reversed the deficits in working memory and motor sequence learning induced by the intracerebral injection of α-synuclein [163,164]—a protein species closely associated with neuronal toxicity in PD. Similarly, impairment of LTP associated with α-synuclein exposure was reversed in A2A receptor knockout mice or following A2A receptor blockade through an NMDA receptor-dependent mechanism.
Collectively, there is a convergence of clinical, epidemiological and experimental evidence supporting the A2A receptor as a potential therapeutic target for improving cognitive impairments in both AD and PD.

4. Potential Neuroprotective Effects of A2A Receptor Antagonism in PD

Uncovering a neuroprotective or disease modifying therapy for PD equates to the search for the holy grail. The promising activity of a range of compounds in a wide spectrum of preclinical models of the illness has stubbornly refused to translate into a clinical effect that meets with regulatory approval. Adenosine A2A receptors may provide another tantalizing lead to follow based on epidemiological data and findings from laboratory studies investigating the protective effects A2A receptor knockout and A2A antagonists against a variety of toxic insults to dopaminergic neurons.
The initial clue for the potential neuroprotective effect of A2A receptor antagonists came from a prospective epidemiological study (The Honolulu Heart Program) carried out over a 30-year period. Daily consumption of 784 mg/day or more of coffee during mid-life was shown to reduce the risk for developing PD at the age of 65 by fivefold compared to non-coffee drinkers after age- and smoking adjustments [165]. Since then, systemic analysis from over 20 observational studies has firmly established that regular human consumption of caffeine is associated with a fivefold reduction in risk for developing PD [165,166,167,168,169]. A meta-analysis of 13 studies, involving over 900,000 participants, found a non-linear relationship between coffee intake and the risk of PD with maximum protection seen at approximately three cups per day [170]. More recently, the Harvard Biomarkers Study conducted a cross-sectional, case–control study which confirmed the robustness of lower caffeine intake as a factor inversely associated with the risk of PD [171].
While these data sound compelling, there are several important caveats on this inverse relationship between coffee consumption and risk of developing PD which need to be considered:
(1)
The effect may not be due to caffeine—however, the consumption of decaffeinated coffee was not associated with a reduced risk of developing PD in the Health Professional Follow-up Study [166].
(2)
The effect is not universal in the population—it is strong and consistent in men in the Health Professional Follow-up Study [169] and in post-menopausal women who have never used hormone replacement therapy in the Cancer Prevention Study II Nutrition Cohort but uncertain in women and post-menopausal women who have used hormone replacement therapy at some point [172].
(3)
The effect could be due to early premotor features of PD causing a reduction in caffeine consumption—however, the persistence of these predictive associations in lag analyses that exclude PD cases that occurred within 2–6 years from dietary survey suggests this is unlikely.
(4)
The effect may reflect genetic predisposition to PD—there is emerging evidence that links dietary [173,174,175] and plasma [176]) caffeine exposure as being more negatively associated with the likelihood of developing PD among carriers of a pathogenic LRRK2 mutation than among PD patients who do not carry LRRK2 mutation [173,176].
The neuroprotective potential of A2A receptor antagonists is further substantiated by mounting evidence from experimental studies that demonstrate A2A receptor blockade confers protection against the degeneration of dopaminergic neurons induced by the toxins, MPTP, 6-OHDA and rotenone [24,177,178,179,180,181,182] and by intracerebral injection of toxic α-synuclein particles [183]. In studies using caffeine, protection against MPTP-induced dopaminergic toxicity in mice was observed after either acute or chronic treatment [24,178,184,185], and importantly, when caffeine was administered after the onset of the neurodegenerative process, reflecting how it might be used in a clinical setting [186]. A2A receptor knock out also prevented the loss of dopaminergic neurons caused by the intra-striatal injection of preformed A53T α-synuclein fibrils or by transgenic overexpression of the human A53T and A30P α-synuclein mutations associated with familial forms of PD [187].
These animal investigations provide a neurobiological basis for the inverse relationship between caffeine consumption and the reduced risk of developing PD established by the epidemiological studies. Collectively, these findings define aberrant A2A receptor signaling as a critical pathogenic mechanism of α-synuclein-triggered neurodegeneration in PD models and support the clinical potential for caffeine and A2A receptor antagonists as potential disease-modifying approaches to treatment. US approval and availability of istradefylline opens up a new opportunity for long-term phase IV clinical studies and/or real-world data analysis of istradefylline in PD that monitor the course of the illness and evaluate the potential disease modifying effects of istradefylline.

5. Summary and Conclusions

It is now understood that PD is more than a motor disorder, and the effects of A2A antagonism on the non-motor symptoms of this multifactorial disease are likely to take greater prominence in treatment decisions. Just as importantly, adenosine antagonism avoids the inevitability of the dopaminergic side effect profile that currently limits the classic dopamine replacement therapies. The potential for neuroprotection also adds another important element that requires further exploration. There is now momentum to understand the role of A2A receptors in a variety of other neurodegenerative diseases (Box 1) and its selective localization provides an attractive therapeutic target. Increased A2A receptor expression has been reported in the brain of patients with AD, amyotrophic lateral sclerosis, Huntington’s disease and multiple sclerosis [188,189,190,191,192,193,194,195].
In addition, A2A receptor antagonists show beneficial effects in animal models of each neurodegenerative disease. Within this context, the A2A receptor has become a seminal example of utilizing high-resolution structures of GPCRs for the rational design of diverse drug molecules. The introduction of istradefylline as the first adenosinergic drug for use in PD opened a new era in the pharmacological treatment of the illness. The continued development of new A2A receptor compounds (partial and inverse agonists as well as antagonists) opens exciting avenues for neurodegenerative disease management.
Box 1. Adenosine A2A receptors in other neurological diseases.
Interest in A2A receptors as a therapeutic target for other neurodegenerative diseases including Alzheimer’s disease (AD), amyotrophic lateral sclerosis (ALS), Huntington’s disease (HD) and multiple sclerosis (MS) has grown exponentially in recent years. Both agonists and antagonists to this receptor have been investigated in the respective animal models of disease and support a potential role for pharmacological manipulation of A2A receptors in each of the pathologies.
Alzheimer’s Disease (AD)
  • A2A receptors are increased in both patients and animal models of AD [188,189,190,191].
  • Istradefylline reduces memory deficits in aging mice with amyloid pathology [196].
  • Age-related shift in LTD is dependent on neuronal A2A receptors [189].
  • A2A receptor knockout mice slow the decrease in Aβ/Tau toxicity [188,197].
Amyotrophic Lateral Sclerosis (ALS)
  • A2A receptors are increased in the spinal cord of patients [192].
  • A2A receptors are increased at the early symptomatic stage in a mouse model of ALS [192], but decreased at the end stage [198].
  • Istradefylline significantly delayed disease progression [192] and rescued LTP impairment adenosine A2A receptors [199] in the mouse model of ALS.
  • Partial genetic ablation of A2A receptors significantly delayed disease progression in SOD1G93A mice [192].
  • The timing by which the alterations of the adenosinergic system occur during ALS pathogenesis in patients and animal models is a key factor to completely understand its contribution to disease progression and to identify the proper therapeutic window for putative treatments.
Huntington’s disease (HD)
  • A2A receptors are decreased in the striatum of HD patients [200].
  • There is a 50% reduction in A2A receptor striatal binding in patients with HD [201].
  • A2A receptors are decreased at the presymptomatic and symptomatic stage in animal models of HD [200,202].
  • A reduction in radiotracer binding for A2A receptors is found in the lesioned side compared with the intact side in a rat model of HD [203].
  • Several A2A receptor antagonists show multiple beneficial effects in chemical- and lesion-induced HD models [204,205].
  • A2A receptor knockout mice have worse survival and motor behaviors in the Tg mouse model of HD [202].
  • However, the effect of A2A receptors blockade in HD mice appears to be complex and could be detrimental because it compromises the function of BDNF [202].
Multiple sclerosis (MS)
  • A2A receptors are increased in lymphocytes from patients with MS and animal models [193,194].
  • A2A receptors are increased in the brain of secondary progressive MS [195].
  • A2A receptors are highly expressed on infiltrating macrophages [206].
  • A2A receptor antagonist (SCH58261) protects against experimental autoimmune encephalomyelitis (EAE) in mice [207].
  • A2A receptor agonist (CGS21680) inhibits the EAE progression in mice [208].
  • A2A receptor knockout mice show more severe EAE pathology and neurological deficits in EAE model mice [209].
Summary of adenosine A2A receptors and neurological disease
Molecules 27 02366 i001
We dedicate this paper to the contribution of Dr Kenneth Jacobson to the field of adenosine receptor science. Over a decades-long and distinctive research career at NIH, Dr. Kenneth Jacobson has led the pioneering work in design, chemically synthesizing, and characterizing pharmacologically new agonists and antagonists for all four subtypes of adenosine receptors, including A2A receptor agonists and antagonists. Through homology to known structures (X-ray and cryo-electron microscopy), Dr. Jacobson contributed to exploration and determination of three-dimensional structures of the A2A receptor (in both agonist- and antagonist-bound states). Dr. Jacobson’s lab has also designed, synthesized and provided functionalized congeners of inhibitors (or antagonists) of the A2A receptor and the enzyme CD73, of adenosine receptor ligands coupling to nanoparticles, such as quantum dots and gold particles, for tissue-specific targeting.

Funding

Some of the authors of this review are employed by Kyowa Kirin Co. Ltd. (A.M. and S.U.), Kyowa Kirin International (C.D.) and Kyowa Kirin, Inc. (S.M.K.), but no author received specific remuneration for this article.

Acknowledgments

Medical writing assistance (referencing and editing) was provided by Anita Chadha-Patel and Elizabeth Hocking of ACP Clinical Communications, Ltd. (Hertfordshire, UK) and funded by Kyowa Kirin, Inc. (Bedminster, NJ, USA).

Conflicts of Interest

Peter Jenner has received honoraria for consultancy and advisory boards from Kyowa Kirin Co., Abbvie, Adamas, Bial, Britannia Pharmaceuticals, Eisai, FP Pharmaceuticals, Hoffman La Roche, Profile Pharma, Teva, UCB, Worldwide Clinical Trials, Zambon. Akihisa Mori and Shinichi Uchida are employed by Kyowa Kirin Co Ltd., Tokyo, Japan. Cecile Durlach is employed by Kyowa Kirin International and Shelby King is employed by Kyowa Kirin, Inc., Bedminster, NJ, USA.

Abbreviations

ACadenylate cyclase
ADAlzheimer’s disease
ALSamyotrophic lateral sclerosis
cAMP3′,5′-cyclic adenosine monophosphate
BDIBeck Depression Inventory
EAEexperimental autoimmune encephalo-myelitis
ESSEpworth Sleepiness Scale
GPeexternal global pallidus
GPglobus pallidus
GPCRG protein-coupled receptor
HDHuntington’s disease
MCImild cognitive impairment
mIPSCsminiature inhibitory postsynaptic currents
MPTP1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine
MSmultiple sclerosis
NMDAN-methyl-D-aspartate
NMSSnon-motor symptom scale
PDParkinson’s disease
PDQ-8PD Questionnaire 8-items
PDSSPD sleep scale
PFCprefrontal cortex
PHQPatient Health Questionnaire
PKAprotein kinase A
SHAPS-JSnaith–Hamilton Pleasure Scale Japanese version
SPNspiny projection neuron
6-OHDA6-hydroxydopamine

References

  1. IJzerman, A.P.; Jacobson, K.A.; Müller, C.E.; Cronstein, B.N.; Cunha, R.A. International Union of Basic and Clinical Pharmacology. CXII: Adenosine Receptors: A Further Update. Pharmacol. Rev. 2022, 74, 340–372. [Google Scholar] [CrossRef] [PubMed]
  2. Saini, A.; Patel, R.; Gaba, S.; Singh, G.; Gupta, G.D.; Monga, V. Adenosine receptor antagonists: Recent advances and therapeutic perspective. Eur. J. Med. Chem. 2022, 227, 113907. [Google Scholar] [CrossRef] [PubMed]
  3. Huang, Z.; Xie, N.; Illes, P.; Virgilio, F.D.; Ulrich, H.; Semyanov, A.; Verkhratsky, A.; Sperlagh, B.; Yu, S.G.; Huang, C.; et al. From purines to purinergic signalling: Molecular functions and human diseases. Signal Transduct. Target. Ther. 2021, 6, 162. [Google Scholar] [CrossRef] [PubMed]
  4. Atif, M.; Alsrhani, A.; Naz, F.; Imran, M.; Ullah, M.I.; Alameen, A.A.M.; Gondal, T.A.; Raza, Q. Targeting Adenosine Receptors in Neurological Diseases. Cell. Reprogramming 2021, 23, 57–72. [Google Scholar] [CrossRef]
  5. Salmaso, V.; Jacobson, K.A. Purinergic Signaling: Impact of GPCR Structures on Rational Drug Design. ChemMedChem 2020, 15, 1958–1973. [Google Scholar] [CrossRef]
  6. Chen, J.F.; Schwarzschild, M.A. Do caffeine and more selective adenosine A2A receptor antagonists protect against dopaminergic neurodegeneration in Parkinson’s disease? Parkinsonism Relat. Disord. 2020, 80, S45–S53. [Google Scholar] [CrossRef]
  7. Chen, J.F.; Lee, C.F.; Chern, Y. Adenosine receptor neurobiology: Overview. Int. Rev. Neurobiol. 2014, 119, 1–49. [Google Scholar]
  8. LeWitt, P.A.; Chaudhuri, K.R. Unmet needs in Parkinson disease: Motor and non-motor. Parkinsonism Relat. Disord. 2020, 80, S7–S12. [Google Scholar] [CrossRef]
  9. Müller, T.; Mueller, B.K.; Riederer, P. Perspective: Treatment for Disease Modification in Chronic Neurodegeneration. Cells 2021, 10, 873. [Google Scholar] [CrossRef]
  10. Mori, A.; Cross, B.; Uchida, S.; Walker, J.K.; Ristuccia, R. How Are Adenosine and Adenosine A2A Receptors Involved in the Pathophysiology of Amyotrophic Lateral Sclerosis? Biomedicines 2021, 9, 1027. [Google Scholar] [CrossRef]
  11. Fredholm, B.B.; IJzerman, A.P.; Jacobson, K.A.; Linden, J.; Müller, C.E. International Union of Basic and Clinical Pharmacology. LXXXI. Nomenclature and classification of adenosine receptors—An update. Pharmacol. Rev. 2011, 63, 1–34. [Google Scholar] [CrossRef]
  12. Mori, A. How do adenosine A2A receptors regulate motor function? Parkinsonism Relat. Disord. 2020, 80, S13–S20. [Google Scholar] [CrossRef]
  13. Jenner, P.; Mori, A.; Hauser, R.; Morelli, M.; Fredholm, B.B.; Chen, J.F. Adenosine, adenosine A2A antagonists, and Parkinson’s disease. Parkinsonism Relat. Disord. 2009, 15, 406–413. [Google Scholar] [CrossRef]
  14. Fredholm, B.B.; Chern, Y.; Franco, R.; Sitkovsky, M. Aspects of the general biology of adenosine A2A signaling. Prog. Neurobiol. 2007, 83, 263–276. [Google Scholar] [CrossRef]
  15. Obeso, J.A.; Rodriguez-Oroz, M.C.; Rodriguez, M.; Lanciego, J.L.; Artieda, J.; Gonzalo, N.; Olanow, C.W. Pathophysiology of the basal ganglia in Parkinson’s disease. Trends Neurosci. 2000, 23, S8–S19. [Google Scholar] [CrossRef]
  16. Obeso, J.A.; Marin, C.; Rodriguez-Oroz, C.; Blesa, J.; Benitez-Temino, B.; Mena-Segovia, J.; Rodriguez, M.; Olanow, C.W. The basal ganglia in Parkinson’s disease: Current concepts and unexplained observations. Ann. Neurol. 2008, 64, S30–S46. [Google Scholar] [CrossRef]
  17. Svenningsson, P.; Moine, C.L.; Fisone, G.; Fredholm, B.B. Distribution, biochemistry and function of striatal adenosine A2A receptors. Prog. Neurobiol. 1999, 59, 355–396. [Google Scholar] [CrossRef]
  18. LeWitt, P.A.; Aradi, S.D.; Hauser, R.A.; Rascol, O. The challenge of developing adenosine A2A antagonists for Parkinson disease: Istradefylline, preladenant, and tozadenant. Parkinsonism Relat. Disord. 2020, 80, S54–S63. [Google Scholar] [CrossRef]
  19. Jenner, P.; Mori, A.; Aradi, S.D.; Hauser, R.A. Istradefylline—A first generation adenosine A2A antagonist for the treatment of Parkinson’s disease. Expert Rev. Neurother. 2021, 21, 317–333. [Google Scholar] [CrossRef]
  20. Fredholm, B.B.; Svenningsson, P. Why target brain adenosine receptors? A historical perspective. Parkinsonism Relat. Disord. 2020, 80, S3–S6. [Google Scholar] [CrossRef]
  21. Fredholm, B.B.; IJzerman, A.P.; Jacobson, K.A.; Klotz, K.N.; Linden, J. International Union of Pharmacology. XXV. Nomenclature and classification of adenosine receptors. Pharmacol. Rev. 2001, 53, 527–552. [Google Scholar] [PubMed]
  22. Hino, T.; Arakawa, T.; Iwanari, H.; Yurugi-Kobayashi, T.; Ikeda-Suno, C.; Nakada-Nakura, Y.; Kusano-Arai, O.; Weyand, S.; Shimamura, T.; Nomura, N.; et al. G-protein-coupled receptor inactivation by an allosteric inverse-agonist antibody. Nature 2012, 482, 237–240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Carpenter, B.; Lebon, G. Human Adenosine A2A Receptor: Molecular Mechanism of Ligand Binding and Activation. Front. Pharmacol. 2017, 8, 898. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Chen, J.F.; Xu, K.; Petzer, J.P.; Staal, R.; Xu, Y.H.; Beilstein, M.; Sonsalla, P.K.; Castagnoli, K.; Castagnoli, N., Jr.; Schwarzschild, M.A. Neuroprotection by caffeine and A2A adenosine receptor inactivation in a model of Parkinson’s disease. J. Neurosci. 2001, 21, RC143. [Google Scholar] [CrossRef] [Green Version]
  25. Carta, A.R.; Kachroo, A.; Schintu, N.; Xu, K.; Schwarzschild, M.A.; Wardas, J.; Morelli, M. Inactivation of neuronal forebrain A2A receptors protects dopaminergic neurons in a mouse model of Parkinson’s disease. J. Neurochem. 2009, 111, 1478–1489. [Google Scholar] [CrossRef] [Green Version]
  26. Rahman, A. The role of adenosine in Alzheimer’s disease. Curr. Neuropharmacol. 2009, 7, 207–216. [Google Scholar] [CrossRef]
  27. Wiprich, M.T.; Bonan, C.D. Purinergic Signaling in the Pathophysiology and Treatment of Huntingtons Disease. Front. Neurosci. 2021, 15, 657338. [Google Scholar] [CrossRef]
  28. Schapira, A.H.V.; Chaudhuri, K.R.; Jenner, P. Non-motor features of Parkinson disease. Nat. Rev. Neurosci. 2017, 18, 435–450. [Google Scholar] [CrossRef]
  29. Chaudhuri, K.R.; Schapira, A.H. Non-motor symptoms of Parkinson’s disease: Dopaminergic pathophysiology and treatment. Lancet Neurol. 2009, 8, 464–474. [Google Scholar] [CrossRef]
  30. Rukavina, K.; Batzu, L.; Leta, V.; Chaudhuri, K.R. New approaches to treatments for sleep, pain and autonomic failure in Parkinson’s disease—Pharmacological therapies. Neuropharmacology 2022, 208, 108959. [Google Scholar] [CrossRef]
  31. Armstrong, M.J.; Okun, M.S. Diagnosis and Treatment of Parkinson Disease: A Review. JAMA 2020, 323, 548–560. [Google Scholar] [CrossRef]
  32. Jenner, P.; Mori, A.; Kanda, T. Can adenosine A2A receptor antagonists be used to treat cognitive impairment, depression or excessive sleepiness in Parkinson’s disease? Parkinsonism Relat. Disord. 2020, 80, S28–S36. [Google Scholar] [CrossRef]
  33. 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]
  34. Barrett, M.J.; Sargent, L.; Nawaz, H.; Weintraub, D.; Price, E.T.; Willis, A.W. Antimuscarinic Anticholinergic Medications in Parkinson Disease: To Prescribe or Deprescribe? Mov. Disord. Clin. Pract. 2021, 8, 1181–1188. [Google Scholar] [CrossRef]
  35. Isaacson, S.H.; Ballard, C.G.; Kreitzman, D.L.; Coate, B.; Norton, J.C.; Fernandez, H.H.; Ilic, T.V.; Azulay, J.P.; Ferreira, J.J.; Abler, V.; et al. Efficacy results of pimavanserin from a multi-center, open-label extension study in Parkinson’s disease psychosis patients. Parkinsonism Relat. Disord. 2021, 87, 25–31. [Google Scholar] [CrossRef]
  36. Freitas, M.E.; Fox, S.H. Nondopaminergic treatments for Parkinson’s disease: Current and future prospects. Neurodegener. Dis. Manag. 2016, 6, 249–268. [Google Scholar] [CrossRef] [Green Version]
  37. Jing, X.Z.; Yuan, X.Z.; Luo, X.; Zhang, S.Y.; Wang, X.P. An Update on Nondopaminergic Treatments for Motor and Non-motor symptoms of Parkinson’s disease. Curr. Neuropharmacol. 2022. [Google Scholar] [CrossRef]
  38. Jarvis, M.F.; Williams, M. Direct autoradiographic localization of adenosine A2 receptors in the rat brain using the A2-selective agonist, [3H]CGS 21680. Eur. J. Pharmacol. 1989, 168, 243–246. [Google Scholar] [CrossRef]
  39. Rosin, D.L.; Robeva, A.; Woodard, R.L.; Guyenet, P.G.; Linden, J. Immunohistochemical localization of adenosine A2A receptors in the rat central nervous system. J. Comp. Neurol. 1998, 401, 163–186. [Google Scholar] [CrossRef]
  40. Svenningsson, P.; Hall, H.; Sedvall, G.; Fredholm, B.B. Distribution of adenosine receptors in the postmortem human brain: An extended autoradiographic study. Synapse 1997, 27, 322–335. [Google Scholar] [CrossRef]
  41. Ishiwata, K.; Mishina, M.; Kimura, Y.; Oda, K.; Sasaki, T.; Ishii, K. First visualization of adenosine A2A receptors in the human brain by positron emission tomography with [11C]TMSX. Synapse 2005, 55, 133–136. [Google Scholar] [CrossRef]
  42. Schiffmann, S.N.; Fisone, G.; Moresco, R.; Cunha, R.A.; Ferre, S. Adenosine A2A receptors and basal ganglia physiology. Prog. Neurobiol. 2007, 83, 277–292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Alexander, G.E.; Crutcher, M.D. Functional architecture of basal ganglia circuits: Neural substrates of parallel processing. Trends Neurosci. 1990, 13, 266–271. [Google Scholar] [CrossRef]
  44. DeLong, M.R. Primate models of movement disorders of basal ganglia origin. Trends Neurosci. 1990, 13, 281–285. [Google Scholar] [CrossRef]
  45. Gerfen, C.R.; Engber, T.M.; Mahan, L.C.; Susel, Z.; Chase, T.N.; Monsma, F.J., Jr.; Sibley, D.R. D1 and D2 dopamine receptor-regulated gene expression of striatonigral and striatopallidal neurons. Science 1990, 250, 1429–1432. [Google Scholar] [CrossRef] [PubMed]
  46. Mori, A.; Shindou, T. Modulation of GABAergic transmission in the striatopallidal system by adenosine A2A receptors: A potential mechanism for the antiparkinsonian effects of A2A antagonists. Neurology 2003, 61, S44–S48. [Google Scholar] [CrossRef] [PubMed]
  47. Cui, G.; Jun, S.B.; Jin, X.; Pham, M.D.; Vogel, S.S.; Lovinger, D.M.; Costa, R.M. Concurrent activation of striatal direct and indirect pathways during action initiation. Nature 2013, 494, 238–242. [Google Scholar] [CrossRef]
  48. Schiffmann, S.N.; Jacobs, O.; Vanderhaeghen, J.J. Striatal restricted adenosine A2 receptor (RDC8) is expressed by enkephalin but not by substance P neurons: An in situ hybridization histochemistry study. J. Neurochem. 1991, 57, 1062–1067. [Google Scholar] [CrossRef]
  49. Shindou, T.; Nonaka, H.; Richardson, P.J.; Mori, A.; Kase, H.; Ichimura, M. Presynaptic adenosine A2A receptors enhance GABAergic synaptic transmission via a cyclic AMP dependent mechanism in the rat globus pallidus. Br. J. Pharmacol. 2002, 136, 296–302. [Google Scholar] [CrossRef] [Green Version]
  50. Shindou, T.; Mori, A.; Kase, H.; Ichimura, M. Adenosine A2A receptor enhances GABAA-mediated IPSCs in the rat globus pallidus. J. Physiol. 2001, 532, 423–434. [Google Scholar] [CrossRef]
  51. Morales-Figueroa, G.E.; Rivera-Ramirez, N.; Gonzalez-Pantoja, R.; Escamilla-Sanchez, J.; Garcia-Hernandez, U.; Galvan, E.J.; Arias-Montano, J.A. Adenosine A2A and histamine H3 receptors interact at the cAMP/PKA pathway to modulate depolarization-evoked [3H]-GABA release from rat striato-pallidal terminals. Purinergic Signal. 2019, 15, 85–93. [Google Scholar] [CrossRef]
  52. Mori, A.; Shindou, T.; Ichimura, M.; Nonaka, H.; Kase, H. The role of adenosine A2A receptors in regulating GABAergic synaptic transmission in striatal medium spiny neurons. J. Neurosci. 1996, 16, 605–611. [Google Scholar] [CrossRef]
  53. Ferré, S.; Bonaventura, J.; Zhu, W.; Hatcher-Solis, C.; Taura, J.; Quiroz, C.; Cai, N.S.; Moreno, E.; Casadó-Anguera, V.; Kravitz, A.V.; et al. Essential Control of the Function of the Striatopallidal Neuron by Pre-coupled Complexes of Adenosine A2A-Dopamine D2 Receptor Heterotetramers and Adenylyl Cyclase. Front. Pharmacol. 2018, 9, 243. [Google Scholar] [CrossRef] [Green Version]
  54. Ferré, S.; Quiroz, C.; Woods, A.S.; Cunha, R.; Popoli, P.; Ciruela, F.; Lluis, C.; Franco, R.; Azdad, K.; Schiffmann, S.N. An update on adenosine A2A-dopamine D2 receptor interactions: Implications for the function of G protein-coupled receptors. Curr. Pharm. Des. 2008, 14, 1468–1474. [Google Scholar] [CrossRef] [Green Version]
  55. Chase, T.N.; Bibbiani, F.; Bara-Jimenez, W.; Dimitrova, T.; Oh-Lee, J.D. Translating A2A antagonist KW6002 from animal models to parkinsonian patients. Neurology 2003, 61, S107–S111. [Google Scholar] [CrossRef]
  56. Uchida, S.; Tashiro, T.; Kawai-Uchida, M.; Mori, A.; Jenner, P.; Kanda, T. Adenosine A2A-receptor antagonist istradefylline enhances the motor response of L-DOPA without worsening dyskinesia in MPTP-treated common marmosets. J. Pharmacol. Sci. 2014, 124, 480–485. [Google Scholar] [CrossRef] [Green Version]
  57. Uchida, S.; Soshiroda, K.; Okita, E.; Kawai-Uchida, M.; Mori, A.; Jenner, P.; Kanda, T. The adenosine A2A receptor antagonist, istradefylline enhances the anti-parkinsonian activity of low doses of dopamine agonists in MPTP-treated common marmosets. Eur. J. Pharmacol. 2015, 747, 160–165. [Google Scholar] [CrossRef]
  58. Canals, M.; Marcellino, D.; Fanelli, F.; Ciruela, F.; Benedetti, P.d.; Goldberg, S.R.; Neve, K.; Fuxe, K.; Agnati, L.F.; Woods, A.S.; et al. Adenosine A2A-dopamine D2 receptor-receptor heteromerization: Qualitative and quantitative assessment by fluorescence and bioluminescence energy transfer. J. Biol. Chem. 2003, 278, 46741–46749. [Google Scholar] [CrossRef] [Green Version]
  59. 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] [Green Version]
  60. Fuxe, K.; Agnati, L.F.; Jacobsen, K.; Hillion, J.; Canals, M.; Torvinen, M.; Tinner-Staines, B.; Staines, W.; Rosin, D.; Terasmaa, A.; et al. Receptor heteromerization in adenosine A2A receptor signaling: Relevance for striatal function and Parkinson’s disease. Neurology 2003, 61, S19–S23. [Google Scholar] [CrossRef]
  61. Gerevich, Z.; Wirkner, K.; Illes, P. Adenosine A2A receptors inhibit the N-methyl-D-aspartate component of excitatory synaptic currents in rat striatal neurons. Eur. J. Pharmacol. 2002, 451, 161–164. [Google Scholar] [CrossRef]
  62. Higley, M.J.; Sabatini, B.L. Competitive regulation of synaptic Ca2+ influx by D2 dopamine and A2A adenosine receptors. Nat. Neurosci. 2010, 13, 958–966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Ferre, S.; Karcz-Kubicha, M.; Hope, B.T.; Popoli, P.; Burgueno, J.; Gutierrez, M.A.; Casado, V.; Fuxe, K.; Goldberg, S.R.; Lluis, C.; et al. Synergistic interaction between adenosine A2A and glutamate mGlu5 receptors: Implications for striatal neuronal function. Proc. Natl. Acad. Sci. USA 2002, 99, 11940–11945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Coccurello, R.; Breysse, N.; Amalric, M. Simultaneous blockade of adenosine A2A and metabotropic glutamate mGlu5 receptors increase their efficacy in reversing Parkinsonian deficits in rats. Neuropsychopharmacology 2004, 29, 1451–1461. [Google Scholar] [CrossRef] [PubMed]
  65. Kachroo, A.; Orlando, L.R.; Grandy, D.K.; Chen, J.F.; Young, A.B.; Schwarzschild, M.A. Interactions between metabotropic glutamate 5 and adenosine A2A receptors in normal and parkinsonian mice. J. Neurosci. Off. J. Soc. Neurosci. 2005, 25, 10414–10419. [Google Scholar] [CrossRef]
  66. Lerner, T.N.; Horne, E.A.; Stella, N.; Kreitzer, A.C. Endocannabinoid signaling mediates psychomotor activation by adenosine A2A antagonists. J. Neuroscience 2010, 30, 2160–2164. [Google Scholar] [CrossRef]
  67. Ferre, S.; Lluis, C.; Justinova, Z.; Quiroz, C.; Orru, M.; Navarro, G.; Canela, E.I.; Franco, R.; Goldberg, S.R. Adenosine-cannabinoid receptor interactions. Implications for striatal function. Br. J. Pharmacol. 2010, 160, 443–453. [Google Scholar] [CrossRef] [Green Version]
  68. Rosin, D.L.; Hettinger, B.D.; Lee, A.; Linden, J. Anatomy of adenosine A2A receptors in brain: Morphological substrates for integration of striatal function. Neurology 2003, 61, S12–S18. [Google Scholar] [CrossRef]
  69. Rebola, N.; Rodrigues, R.J.; Lopes, L.V.; Richardson, P.J.; Oliveira, C.R.; Cunha, R.A. Adenosine A1 and A2A receptors are co-expressed in pyramidal neurons and co-localized in glutamatergic nerve terminals of the rat hippocampus. Neuroscience 2005, 133, 79–83. [Google Scholar] [CrossRef]
  70. Ciruela, F.; Casado, V.; Rodrigues, R.J.; Lujan, R.; Burgueno, J.; Canals, M.; Borycz, J.; Rebola, N.; Goldberg, S.R.; Mallol, J.; et al. Presynaptic control of striatal glutamatergic neurotransmission by adenosine A1-A2A receptor heteromers. J. Neuroscience 2006, 26, 2080–2087. [Google Scholar] [CrossRef] [Green Version]
  71. Ramlackhansingh, A.F.; Bose, S.K.; Ahmed, I.; Turkheimer, F.E.; Pavese, N.; Brooks, D.J. Adenosine 2A receptor availability in dyskinetic and nondyskinetic patients with Parkinson disease. Neurology 2011, 76, 1811–1816. [Google Scholar] [CrossRef] [Green Version]
  72. Morelli, M.; Paolo, T.D.; Wardas, J.; Calon, F.; Xiao, D.; Schwarzschild, M.A. Role of adenosine A2A receptors in parkinsonian motor impairment and l-DOPA-induced motor complications. Prog. Neurobiol. 2007, 83, 293–309. [Google Scholar] [CrossRef]
  73. Calon, F.; Dridi, M.; Hornykiewicz, O.; Bedard, P.J.; Rajput, A.H.; Paolo, T.D. Increased adenosine A2A receptors in the brain of Parkinson’s disease patients with dyskinesias. Brain 2004, 127, 1075–1084. [Google Scholar] [CrossRef] [Green Version]
  74. Mishina, M.; Ishiwata, K.; Naganawa, M.; Kimura, Y.; Kitamura, S.; Suzuki, M.; Hashimoto, M.; Ishibashi, K.; Oda, K.; Sakata, M.; et al. Adenosine A(2A) receptors measured with [C]TMSX PET in the striata of Parkinson’s disease patients. PLoS ONE 2011, 6, e17338. [Google Scholar] [CrossRef]
  75. Villar-Menendez, I.; Porta, S.; Buira, S.P.; Pereira-Veiga, T.; Diaz-Sanchez, S.; Albasanz, J.L.; Ferrer, I.; Martin, M.; Barrachina, M. Increased striatal adenosine A2A receptor levels is an early event in Parkinson’s disease-related pathology and it is potentially regulated by miR-34b. Neurobiol. Dis. 2014, 69, 206–214. [Google Scholar] [CrossRef]
  76. Mori, A. Mode of action of adenosine A2A receptor antagonists as symptomatic treatment for Parkinson’s disease. Int. Rev. Neurobiol. 2014, 119, 87–116. [Google Scholar]
  77. Xu, K.; Bastia, E.; Schwarzschild, M. Therapeutic potential of adenosine A2A receptor antagonists in Parkinson’s disease. Pharmacol. Ther. 2005, 105, 267–310. [Google Scholar] [CrossRef]
  78. Carmo, M.; Gonçalves, F.Q.; Canas, P.M.; Oses, J.P.; Fernandes, F.D.; Duarte, F.V. Enhanced ATP release and CD73-mediated adenosine formation sustain adenosine A2A receptor over-activation in a rat model of Parkinson’s disease. Br. J. Pharmacol. 2019, 176, 3666–3680. [Google Scholar] [CrossRef]
  79. Meng, F.; Guo, Z.; Hu, Y.; Mai, W.; Zhang, Z.; Zhang, B.; Ge, Q.; Lou, H.; Guo, F.; Chen, J.; et al. CD73-derived adenosine controls inflammation and neurodegeneration by modulating dopamine signalling. Brain 2019, 142, 700–718. [Google Scholar] [CrossRef]
  80. Rodríguez, D.; Gao, Z.G.; Moss, S.M.; Jacobson, K.A.; Carlsson, J. Molecular docking screening using agonist-bound GPCR structures: Probing the A2A adenosine receptor. J. Chem. Inf. Modeling 2015, 55, 550–563. [Google Scholar] [CrossRef] [Green Version]
  81. Bennett, K.A.; Tehan, B.G.; Lebon, G.; Tate, C.G.; Weir, M.; Marshall, F.H.; Langmead, C.J. Pharmacology and Structure of Isolated Conformations of the Adenosine A2A Receptor Define Ligand Efficacy. Mol. Pharmacol. 2013, 83, 949–958. [Google Scholar] [CrossRef] [Green Version]
  82. Lebon, G.; Warne, T.; Edwards, P.C.; Bennett, K.; Langmead, C.J.; Leslie, A.G.; Tate, C.G. Agonist-bound adenosine A2A receptor structures reveal common features of GPCR activation. Nature 2011, 474, 521–525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Jaakola, V.P.; Griffith, M.T.; Hanson, M.A.; Cherezov, V.; Chien, E.Y.; Lane, J.R.; Ijzerman, A.P.; Stevens, R.C. The 2.6 angstrom crystal structure of a human A2A adenosine receptor bound to an antagonist. Science 2008, 322, 1211–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Xu, F.; Wu, H.; Katritch, V.; Han, G.W.; Jacobson, K.A.; Gao, Z.G.; Cherezov, V.; Stevens, R.C. Structure of an agonist-bound human A2A adenosine receptor. Science 2011, 332, 322–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Doré, A.S.; Robertson, N.; Errey, J.C.; Ng, I.; Hollenstein, K.; Tehan, B.; Hurrell, E.; Bennett, K.; Congreve, M.; Magnani, F.; et al. Structure of the adenosine A2A receptor in complex with ZM241385 and the xanthines XAC and caffeine. Structure 2011, 19, 1283–1293. [Google Scholar] [CrossRef] [Green Version]
  86. Caliman, A.D.; Miao, Y.; McCammon, J.A. Mapping the allosteric sites of the A2A adenosine receptor. Chem. Biol. Drug Des. 2018, 91, 5–16. [Google Scholar] [CrossRef]
  87. Amelia, T.; Veldhoven, J.P.D.v.; Falsini, M.; Liu, R.; Heitman, L.H.; Westen, G.J.P.v.; Segala, E.; Verdon, G.; Cheng, R.K.Y.; Cooke, R.M.; et al. Crystal Structure and Subsequent Ligand Design of a Nonriboside Partial Agonist Bound to the Adenosine A2A Receptor. J. Med. Chem. 2021, 64, 3827–3842. [Google Scholar] [CrossRef]
  88. Mishra, A.; Singh, S.; Shukla, S. Physiological and Functional Basis of Dopamine Receptors and Their Role in Neurogenesis: Possible Implication for Parkinson’s disease. J. Exp. Neurosci. 2018, 12, 1179069518779829. [Google Scholar] [CrossRef] [Green Version]
  89. Kyowa Kirin North American Pharmaceutical Pipeline. Available online: https://kkna.kyowakirin.com/wp-content/uploads/Kyowa-Kirin-North-American-Pipeline.pdf (accessed on 30 March 2022).
  90. Kyowa Hakko Kirin Announces Results of Early Phase 2 Trial of KW-6356 for Parkinson’s Disease at IAPRD. Available online: https://www.kyowakirin.com/media_center/news_releases/2018/e20180820_01.html (accessed on 30 March 2022).
  91. Pinna, A.; Serra, M.; Marongiu, J.; Morelli, M. Pharmacological interactions between adenosine A2A receptor antagonists and different neurotransmitter systems. Parkinsonism Relat. Disord. 2020, 80, S37–S44. [Google Scholar] [CrossRef]
  92. Shook, B.C.; Jackson, P.F. Adenosine A2A Receptor Antagonists and Parkinson’s Disease. Acs. Chem. Neurosci. 2011, 2, 555–567. [Google Scholar] [CrossRef] [Green Version]
  93. Hauser, R.A.; Hattori, N.; Fernandez, H.; Isaacson, S.H.; Mochizuki, H.; Rascol, O.; Stocchi, F.; Li, J.; Mori, A.; Nakajima, Y.; et al. Efficacy of Istradefylline, an Adenosine A2A Receptor Antagonist, as Adjunctive Therapy to Levodopa in Parkinson’s Disease: A Pooled Analysis of 8 Phase 2b/3 Trials. J. Parkinsons Dis. 2021, 11, 1663–1675. [Google Scholar] [CrossRef]
  94. Brooks, D.J.; Papapetropoulos, S.; Vandenhende, F.; Tomic, D.; He, P.; Coppell, A.; O’Neill, G. An open-label, positron emission tomography study to assess adenosine A2A brain receptor occupancy of vipadenant (BIIB014) at steady-state levels in healthy male volunteers. Clin. Neuropharmacol. 2010, 33, 55–60. [Google Scholar] [CrossRef]
  95. Pawsey, S.; Donaldson, K.; Warrington, S.; Tranter, E. A phase I single and repeated dose pharmacokinetic study of oral V81444, a selective non-xanthine adenosine A2A receptor antagonist. J. Neurol. Sci. 2013, 333, e135. [Google Scholar] [CrossRef]
  96. Pinna, A. Adenosine A2A receptor antagonists in Parkinson’s disease: Progress in clinical trials from the newly approved istradefylline to drugs in early development and those already discontinued. CNS Drugs 2014, 28, 455–474. [Google Scholar] [CrossRef]
  97. Mizuno, Y.; Hasegawa, K.; Kondo, T.; Kuno, S.; Yamamoto, M.; Japanese Istradefylline Study Group. Clinical efficacy of istradefylline (KW-6002) in Parkinson’s disease: A randomized, controlled study. Mov. Disord. 2010, 25, 1437–1443. [Google Scholar] [CrossRef]
  98. Fernandez, H.H.; Greeley, D.R.; Zweig, R.M.; Wojcieszek, J.; Mori, A.; Sussman, N.M. Istradefylline as monotherapy for Parkinson disease: Results of the 6002-US-051 trial. Parkinsonism Relat. Disord. 2010, 16, 16–20. [Google Scholar] [CrossRef]
  99. Stacy, M.; Silver, D.; Mendis, T.; Sutton, J.; Mori, A.; Chaikin, P.; Sussman, N.M. A 12-week, placebo-controlled study (6002-US-006) of istradefylline in Parkinson disease. Neurology 2008, 70, 2233–2240. [Google Scholar] [CrossRef]
  100. LeWitt, P.A.; Guttman, M.; Tetrud, J.W.; Tuite, P.J.; Mori, A.; Chaikin, P.; Sussman, N.M. Adenosine A2A receptor antagonist istradefylline (KW-6002) reduces “off” time in Parkinson’s disease: A double-blind, randomized, multicenter clinical trial (6002-US-005). Ann. Neurol. 2008, 63, 295–302. [Google Scholar] [CrossRef]
  101. Hauser, R.A.; Shulman, L.M.; Trugman, J.M.; Roberts, J.W.; Mori, A.; Ballerini, R.; Sussman, N.M. Study of istradefylline in patients with Parkinson’s disease on levodopa with motor fluctuations. Mov. Disord. 2008, 23, 2177–2185. [Google Scholar] [CrossRef]
  102. Pourcher, E.; Fernandez, H.H.; Stacy, M.; Mori, A.; Ballerini, R.; Chaikin, P. Istradefylline for Parkinson’s disease patients experiencing motor fluctuations: Results of the KW-6002-US-018 study. Parkinsonism Relat. Disord. 2012, 18, 178–184. [Google Scholar] [CrossRef]
  103. Guttman, M. Efficacy of istradefylline in Parkinson’s disease patients treated with levodopa with motor response complications: Results of the KW-6002 US-018 study. Mov. Disord. 2006, 21, S585. [Google Scholar]
  104. Hauser, R.A.; Hubble, J.P.; Truong, D.D.; Istradefylline US-001 Study Group. Randomized trial of the adenosine A2A receptor antagonist istradefylline in advanced PD. Neurology 2003, 61, 297–303. [Google Scholar] [CrossRef]
  105. Hauser, R.A.; Cantillon, M.; Pourcher, E.; Micheli, F.; Mok, V.; Onofrj, M.; Huyck, S.; Wolski, K. Preladenant in patients with Parkinson’s disease and motor fluctuations: A phase 2, double-blind, randomised trial. Lancet Neurol. 2011, 10, 221–229. [Google Scholar] [CrossRef]
  106. Hauser, R.A.; Stocchi, F.; Rascol, O.; Huyck, S.B.; Capece, R.; Ho, T.W.; Sklar, P.; Lines, C.; Michelson, D.; Hewitt, D. Preladenant as an Adjunctive Therapy with Levodopa in Parkinson Disease: Two Randomized Clinical Trials and Lessons Learned. JAMA Neurol. 2015, 72, 1491–1500. [Google Scholar] [CrossRef] [Green Version]
  107. Hauser, R.A.; Olanow, C.W.; Kieburtz, K.D.; Pourcher, E.; Docu-Axelerad, A.; Lew, M.; Kozyolkin, O.; Neale, A.; Resburg, C.; Meya, U.; et al. Tozadenant (SYN115) in patients with Parkinson’s disease who have motor fluctuations on levodopa: A phase 2b, double-blind, randomised trial. Lancet Neurol. 2014, 13, 767–776. [Google Scholar] [CrossRef]
  108. BRIEF-Acorda Discontinues Tozadenant Development Program. Available online: https://www.reuters.com/article/brief-acorda-discontinues-tozadenant-dev/brief-acorda-discontinues-tozadenant-development-program-idUSFWN1NQ0IG (accessed on 30 March 2022).
  109. Nourianz (Istradefylline) Tablets, for Oral Use (Prescribing Information). Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2019/022075s000lbl.pdf (accessed on 30 March 2022).
  110. Report on the Deliberation Results. Available online: https://www.pmda.go.jp/files/000153870.pdf (accessed on 30 March 2022).
  111. Weintraub, D.; Mamikonyan, E. The Neuropsychiatry of Parkinson Disease: A Perfect Storm. Am. J. Geriatr. Psychiatry 2019, 27, 998–1018. [Google Scholar] [CrossRef] [PubMed]
  112. Weintraub, D.; Aarsland, D.; Chaudhuri, K.R.; Dobkin, R.D.; Leentjens, A.F.; Rodriguez-Violante, M.; Schrag, A. The neuropsychiatry of Parkinson’s disease: Advances and challenges. Lancet Neurol. 2022, 21, 89–102. [Google Scholar] [CrossRef]
  113. Chen, J.F.; Cunha, R.A. The belated US FDA approval of the adenosine A2A receptor antagonist istradefylline for treatment of Parkinson’s disease. Purinergic Signal. 2020, 16, 167–174. [Google Scholar] [CrossRef] [PubMed]
  114. Kaster, M.P.; Machado, N.J.; Silva, H.B.; Nunes, A.; Ardais, A.P.; Santana, M.; Baqi, Y.; Müller, C.E.; Rodrigues, A.L.; Porciúncula, L.O.; et al. Caffeine acts through neuronal adenosine A2A receptors to prevent mood and memory dysfunction triggered by chronic stress. Proc. Natl. Acad. Sci. USA 2015, 112, 7833–7838. [Google Scholar] [CrossRef] [Green Version]
  115. Ren, X.; Chen, J.F. Caffeine and Parkinson’s Disease: Multiple Benefits and Emerging Mechanisms. Front. Neurosci. 2020, 14, 602697. [Google Scholar] [CrossRef]
  116. Li, P.; Rial, D.; Canas, P.M.; Yoo, J.H.; Li, W.; Zhou, X.; Wang, Y.; Westen, G.J.v.; Payen, M.P.; Augusto, E.; et al. Optogenetic activation of intracellular adenosine A2A receptor signaling in the hippocampus is sufficient to trigger CREB phosphorylation and impair memory. Mol. Psychiatry 2015, 20, 1339–1349. [Google Scholar] [CrossRef] [Green Version]
  117. Batalha, V.L.; Pego, J.M.; Fontinha, B.M.; Costenla, A.R.; Valadas, J.S.; Baqi, Y.; Radjainia, H.; Müller, C.E.; Sebastião, A.M.; Lopes, L.V. Adenosine A2A receptor blockade reverts hippocampal stress-induced deficits and restores corticosterone circadian oscillation. Mol. Psychiatry 2013, 18, 320–331. [Google Scholar] [CrossRef] [Green Version]
  118. Todorova, A.; Jenner, P.; Chaudhuri, K.R. Non-motor Parkinson’s: Integral to motor Parkinson’s, yet often neglected. Pract. Neurol. 2014, 14, 310–322. [Google Scholar] [CrossRef]
  119. Chahine, L.M.; Amara, A.W.; Videnovic, A. A systematic review of the literature on disorders of sleep and wakefulness in Parkinson’s disease from 2005 to 2015. Sleep Med. Rev. 2017, 35, 33–50. [Google Scholar] [CrossRef]
  120. Oishi, Y.; Xu, Q.; Wang, L.; Zhang, B.J.; Takahashi, K.; Takata, Y.; Luo, Y.J.; Cherasse, Y.; Schiffmann, S.N.; d’Exaerde, A.d.K.; et al. Slow-wave sleep is controlled by a subset of nucleus accumbens core neurons in mice. Nat. Commun. 2017, 8, 734. [Google Scholar] [CrossRef] [Green Version]
  121. Wang, Y.Q.; Li, R.; Wang, D.R.; Cherasse, Y.; Zhang, Z.; Zhang, M.Q.; Lavielle, O.; McEown, K.; Schiffmann, S.N.; d’Exaerde, A.d.K.; et al. Adenosine A2A receptors in the olfactory bulb suppress rapid eye movement sleep in rodents. Brain Struct. Funct. 2017, 222, 1351–1366. [Google Scholar] [CrossRef]
  122. Lazarus, M.; Shen, H.Y.; Cherasse, Y.; Qu, W.M.; Huang, Z.L.; Bass, C.E.; Winsky-Sommerer, R.; Semba, K.; Fredholm, B.B.; Boison, D.; et al. Arousal effect of caffeine depends on adenosine A2A receptors in the shell of the nucleus accumbens. J. Neurosci. 2011, 31, 10067–10075. [Google Scholar] [CrossRef] [Green Version]
  123. Porkka-Heiskanen, T.; Strecker, R.E.; Thakkar, M.; Bjorkum, A.A.; Greene, R.W.; McCarley, R.W. Adenosine: A mediator of the sleep-inducing effects of prolonged wakefulness. Science 1997, 276, 1265–1268. [Google Scholar] [CrossRef] [Green Version]
  124. Huang, Z.L.; Qu, W.M.; Eguchi, N.; Chen, J.F.; Schwarzschild, M.A.; Fredholm, B.B.; Urade, Y.; Hayaishi, O. Adenosine A2A, but not A1, receptors mediate the arousal effect of caffeine. Nat. Neurosci. 2005, 8, 858–859. [Google Scholar] [CrossRef]
  125. Suzuki, K.; Miyamoto, M.; Miyamoto, T.; Uchiyama, T.; Watanabe, Y.; Suzuki, S.; Kadowaki, T.; Fujita, H.; Matsubara, T.; Sakuramoto, H.; et al. Istradefylline improves daytime sleepiness in patients with Parkinson’s disease: An open-label, 3-month study. J. Neurol. Sci. 2017, 380, 230–233. [Google Scholar] [CrossRef]
  126. Matsuura, K.; Kajikawa, H.; Tabei, K.I.; Satoh, M.; Kida, H.; Nakamura, N.; Tomimoto, H. The effectiveness of istradefylline for the treatment of gait deficits and sleepiness in patients with Parkinson’s disease. Neurosci. Lett. 2018, 662, 158–161. [Google Scholar] [CrossRef]
  127. Aarsland, D.; Påhlhagen, S.; Ballard, C.G.; Ehrt, U.; Svenningsson, P. Depression in Parkinson disease—Epidemiology, mechanisms and management. Nat. Reviews Neurol. 2011, 8, 35–47. [Google Scholar] [CrossRef] [PubMed]
  128. Remy, P.; Doder, M.; Lees, A.; Turjanski, N.; Brooks, D. Depression in Parkinson’s disease: Loss of dopamine and noradrenaline innervation in the limbic system. Brain 2005, 128, 1314–1322. [Google Scholar] [CrossRef] [Green Version]
  129. Santangelo, G.; Vitale, C.; Picillo, M.; Cuoco, S.; Moccia, M.; Pezzella, D.; Erro, R.; Longo, K.; Vicidomini, C.; Pellecchia, M.T.; et al. Apathy and striatal dopamine transporter levels in de-novo, untreated Parkinson’s disease patients. Parkinsonism Relat. Disord. 2015, 21, 489–493. [Google Scholar] [CrossRef]
  130. Heron, C.L.; Plant, O.; Manohar, S.; Ang, Y.S.; Jackson, M.; Lennox, G.; Hu, M.T.; Husain, M. Distinct effects of apathy and dopamine on effort-based decision-making in Parkinson’s disease. Brain 2018, 141, 1455–1469. [Google Scholar] [CrossRef]
  131. Yamada, K.; Kobayashi, M.; Mori, A.; Jenner, P.; Kanda, T. Antidepressant-like activity of the adenosine A(2A) receptor antagonist, istradefylline (KW-6002), in the forced swim test and the tail suspension test in rodents. Pharmacol. Biochem. Behav. 2013, 114, 23–30. [Google Scholar] [CrossRef]
  132. Yamada, K.; Kobayashi, M.; Shiozaki, S.; Ohta, T.; Mori, A.; Jenner, P.; Kanda, T. Antidepressant activity of the adenosine A2A receptor antagonist, istradefylline (KW-6002) on learned helplessness in rats. Psychopharmacology 2014, 231, 2839–2849. [Google Scholar] [CrossRef]
  133. Wei, C.J.; Augusto, E.; Gomes, C.A.; Singer, P.; Wang, Y.; Boison, D.; Cunha, R.A.; Yee, B.K.; Chen, J.F. Regulation of fear responses by striatal and extrastriatal adenosine A2A receptors in forebrain. Biol. Psychiatry 2014, 75, 855–863. [Google Scholar] [CrossRef] [Green Version]
  134. Ito, H.; Fukutake, S.; Odake, S.; Kawada, J.; Iwanaga, S.; Kamei, T. Clinical Efficacy of Istradefylline for Depression in Parkinson’s Disease. J. Neurol. Neurosci. 2018, 9, 261–263. [Google Scholar] [CrossRef]
  135. Nagayama, H.; Kano, O.; Murakami, H.; Ono, K.; Hamada, M.; Toda, T.; Sengoku, R.; Shimo, Y.; Hattori, N. Effect of istradefylline on mood disorders in Parkinson’s disease. J. Neurol. Sci. 2019, 396, 78–83. [Google Scholar] [CrossRef] [PubMed]
  136. Hameleers, P.A.; Boxtel, M.P.V.; Hogervorst, E.; Riedel, W.J.; Houx, P.J.; Buntinx, F.; Jolles, J. Habitual caffeine consumption and its relation to memory, attention, planning capacity and psychomotor performance across multiple age groups. Hum. Psychopharmacol. 2000, 15, 573–581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Vanboxtel, M.; Schmitt, J.; Bosma, H.; Jolles, J. The effects of habitual caffeine use on cognitive change: A longitudinal perspective. Pharmacol. Biochem. Behav. 2003, 75, 921–927. [Google Scholar] [CrossRef]
  138. Lindsay, J.; Laurin, D.; Verreault, R.; Hebert, R.; Helliwell, B.; Hill, G.B.; McDowell, I. Risk factors for Alzheimer’s disease: A prospective analysis from the Canadian Study of Health and Aging. Am. J. Epidemiol. 2002, 156, 445–453. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Gelder, B.M.v.; Buijsse, B.; Tijhuis, M.; Kalmijn, S.; Giampaoli, S.; Nissinen, A.; Kromhout, D. Coffee consumption is inversely associated with cognitive decline in elderly European men: The FINE Study. Eur. J. Clin. Nutr. 2007, 61, 226–232. [Google Scholar] [CrossRef] [Green Version]
  140. Ritchie, K.; Carrière, I.; Portet, F.; Mendonca, A.D.; Dartigues, J.F.; Rouaud, O.; Barberger-Gateau, P.; Ancelin, M.L. The neuro-protective effects of caffeine: A prospective population study (the Three City Study). Neurology 2007, 69, 536–545. [Google Scholar] [CrossRef]
  141. Eskelinen, M.H.; Ngandu, T.; Tuomilehto, J.; Soininen, H.; Kivipelto, M. Midlife coffee and tea drinking and the risk of late-life dementia: A population-based CAIDE study. J. Alzheimers Dis. 2009, 16, 85–91. [Google Scholar] [CrossRef] [Green Version]
  142. Cao, C.; Cirrito, J.R.; Lin, X.; Wang, L.; Verges, D.K.; Dickson, A.; Mamcarz, M.; Zhang, C.; Mori, T.; Arendash, G.W.; et al. Caffeine suppresses amyloid-beta levels in plasma and brain of Alzheimer’s disease transgenic mice. J. Alzheimers Dis. 2009, 17, 681–697. [Google Scholar] [CrossRef] [Green Version]
  143. Woods, L.T.; Ajit, D.; Camden, J.M.; Erb, L.; Weisman, G.A. Purinergic receptors as potential therapeutic targets in Alzheimer’s disease. Neuropharmacology 2016, 104, 169–179. [Google Scholar] [CrossRef] [Green Version]
  144. Cho, B.H.; Choi, S.M.; Kim, J.T.; Kim, B.C. Association of coffee consumption and non-motor symptoms in drug-naive, early-stage Parkinson’s disease. Parkinsonism Relat. Disord. 2018, 50, 42–47. [Google Scholar] [CrossRef]
  145. Zhou, S.J.; Zhu, M.E.; Shu, D.; Du, X.P.; Song, X.H.; Wang, X.T.; Zheng, R.Y.; Cai, X.H.; Chen, J.F.; He, J.C. Preferential enhancement of working memory in mice lacking adenosine A(2A) receptors. Brain Res. 2009, 1303, 74–83. [Google Scholar] [CrossRef]
  146. Wei, C.J.; Singer, P.; Coelho, J.; Boison, D.; Feldon, J.; Yee, B.K.; Chen, J.F. Selective inactivation of adenosine A2A receptors in striatal neurons enhances working memory and reversal learning. Learn. Mem. 2011, 18, 459–474. [Google Scholar] [CrossRef] [Green Version]
  147. Mingote, S.; Font, L.; Farrar, A.M.; Vontell, R.; Worden, L.T.; Stopper, C.M.; Port, R.G.; Sink, K.S.; Bunce, J.G.; Chrobak, J.J.; et al. Nucleus accumbens adenosine A2A receptors regulate exertion of effort by acting on the ventral striatopallidal pathway. J. Neurosci. 2008, 28, 9037–9046. [Google Scholar] [CrossRef] [Green Version]
  148. Li, Y.; He, Y.; Chen, M.; Pu, Z.; Chen, L.; Li, P.; Li, B.; Li, H.; Huang, Z.L.; Li, Z.; et al. Optogenetic Activation of Adenosine A2A Receptor Signaling in the Dorsomedial Striatopallidal Neurons Suppresses Goal-Directed Behavior. Neuropsychopharmacology 2016, 41, 1003–1013. [Google Scholar] [CrossRef] [Green Version]
  149. Giménez-Llort, L.; Schiffmann, S.N.; Shmidt, T.; Canela, L.; Camón, L.; Wassholm, M.; Canals, M.; Terasmaa, A.; Fernández-Teruel, A.; Tobeña, A.; et al. Working memory deficits in transgenic rats overexpressing human adenosine A2A receptors in the brain. Neurobiol. Learn. Mem. 2007, 87, 42–56. [Google Scholar] [CrossRef] [Green Version]
  150. Batalha, V.L.; Ferreira, D.G.; Coelho, J.E.; Valadas, J.S.; Gomes, R.; Temido-Ferreira, M.; Shmidt, T.; Baqi, Y.; Buée, L.; Müller, C.E.; et al. The caffeine-binding adenosine A2A receptor induces age-like HPA-axis dysfunction by targeting glucocorticoid receptor function. Sci. Rep. 2016, 6, 31493. [Google Scholar] [CrossRef]
  151. Knowlton, B.J.; Mangels, J.A.; Squire, L.R. A neostriatal habit learning system in humans. Science 1996, 273, 1399–1402. [Google Scholar] [CrossRef] [Green Version]
  152. Kehagia, A.A.; Barker, R.A.; Robbins, T.W. Neuropsychological and clinical heterogeneity of cognitive impairment and dementia in patients with Parkinson’s disease. Lancet Neurol. 2010, 9, 1200–1213. [Google Scholar] [CrossRef]
  153. Aarsland, D.; Creese, B.; Politis, M.; Chaudhuri, K.R.; Ffytche, D.H.; Weintraub, D.; Ballard, C. Cognitive decline in Parkinson disease. Nat. Reviews. Neurol. 2017, 13, 217–231. [Google Scholar] [CrossRef] [Green Version]
  154. Laar, T.v.; Deyn, P.P.D.; Aarsland, D.; Barone, P.; Galvin, J.E. Effects of cholinesterase inhibitors in Parkinson’s disease dementia: A review of clinical data. CNS Neurosci. Ther. 2011, 17, 428–441. [Google Scholar] [CrossRef]
  155. Taylor, A.E.; Saint-Cyr, J.A.; Lang, A.E. Parkinson’s disease. Cognitive changes in relation to treatment response. Brain 1987, 110, 35–51. [Google Scholar] [CrossRef]
  156. Gotham, A.M.; Brown, R.G.; Marsden, C.D. ‘Frontal’ cognitive function in patients with Parkinson’s disease ‘on’ and ‘off’ levodopa. Brain 1988, 111, 299–321. [Google Scholar] [CrossRef] [PubMed]
  157. Lange, K.W.; Robbins, T.W.; Marsden, C.D.; James, M.; Owen, A.M.; Paul, G.M. L-dopa withdrawal in Parkinson’s disease selectively impairs cognitive performance in tests sensitive to frontal lobe dysfunction. Psychopharmacology 1992, 107, 394–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Clinton, S.M.; Sucharski, I.L.; Finlay, J.M. Desipramine attenuates working memory impairments induced by partial loss of catecholamines in the rat medial prefrontal cortex. Psychopharmacology 2006, 183, 404–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Brozoski, T.J.; Brown, R.M.; Rosvold, H.E.; Goldman, P.S. Cognitive deficit caused by regional depletion of dopamine in prefrontal cortex of rhesus monkey. Science 1979, 205, 929–932. [Google Scholar] [CrossRef]
  160. Decamp, E.; Schneider, J.S. Attention and executive function deficits in chronic low-dose MPTP-treated non-human primates. Eur. J. Neurosci. 2004, 20, 1371–1378. [Google Scholar] [CrossRef]
  161. Horita, T.K.; Kobayashi, M.; Mori, A.; Jenner, P.; Kanda, T. Effects of the adenosine A2A antagonist istradefylline on cognitive performance in rats with a 6-OHDA lesion in prefrontal cortex. Psychopharmacology 2013, 230, 345–352. [Google Scholar] [CrossRef]
  162. Li, Z.; Chen, X.; Wang, T.; Gao, Y.; Li, F.; Chen, L.; Xue, J.; He, Y.; Li, Y.; Guo, W.; et al. The Corticostriatal Adenosine A2A Receptor Controls Maintenance and Retrieval of Spatial Working Memory. Biol. Psychiatry 2018, 83, 530–541. [Google Scholar] [CrossRef]
  163. Ko, W.K.D.; Camus, S.M.; Li, Q.; Yang, J.; McGuire, S.; Pioli, E.Y.; Bezard, E. An evaluation of istradefylline treatment on Parkinsonian motor and cognitive deficits in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-treated macaque models. Neuropharmacology 2016, 110, 48–58. [Google Scholar] [CrossRef]
  164. He, Y.; Huang, L.; Wang, K.; Pan, X.; Cai, Q.; Zhang, F.; Yang, J.; Fang, G.; Zhao, X.; You, F.; et al. α-Synuclein Selectively Impairs Motor Sequence Learning and Value Sensitivity: Reversal by the Adenosine A2A Receptor Antagonists. Cereb. Cortex 2021, 32, 808–823. [Google Scholar] [CrossRef]
  165. Ross, G.W.; Abbott, R.D.; Petrovitch, H.; Morens, D.M.; Grandinetti, A.; Tung, K.H.; Tanner, C.M.; Masaki, K.; Blanchette, P.L.; Curb, J.D.; et al. Association of coffee and caffeine intake with the risk of Parkinson disease. JAMA 2000, 283, 2674–2679. [Google Scholar] [CrossRef]
  166. Ascherio, A.; Zhang, S.M.; Hernan, M.A.; Kawachi, I.; Colditz, G.A.; Speizer, F.E.; Willett, W.C. Prospective study of caffeine consumption and risk of Parkinson’s disease in men and women. Ann. Neurol. 2001, 50, 56–63. [Google Scholar] [CrossRef]
  167. Ascherio, A.; Chen, H.; Schwarzschild, M.A.; Zhang, S.M.; Colditz, G.A.; Speizer, F.E. Caffeine, postmenopausal estrogen, and risk of Parkinson’s disease. Neurology 2003, 60, 790–795. [Google Scholar] [CrossRef]
  168. Saaksjarvi, K.; Knekt, P.; Rissanen, H.; Laaksonen, M.A.; Reunanen, A.; Mannisto, S. Prospective study of coffee consumption and risk of Parkinson’s disease. Eur. J. Clin. Nutr. 2008, 62, 908–915. [Google Scholar] [CrossRef]
  169. Grosso, G.; Godos, J.; Galvano, F.; Giovannucci, E.L. Coffee, caffeine, and health outcomes: An umbrella review. Annu. Rev. Nutr. 2017, 37, 131–156. [Google Scholar] [CrossRef] [Green Version]
  170. Qi, H.; Li, S. Dose-response meta-analysis on coffee, tea and caffeine consumption with risk of Parkinson’s disease. Geriatr. Gerontol. Int. 2014, 14, 430–439. [Google Scholar] [CrossRef]
  171. Bakshi, R.; Macklin, E.A.; Hung, A.Y.; Hayes, M.T.; Hyman, B.T.; Wills, A.M.; Gomperts, S.N.; Growdon, J.H.; Ascherio, A.; Scherzer, C.R.; et al. Associations of Lower Caffeine Intake and Plasma Urate Levels with Idiopathic Parkinson’s Disease in the Harvard Biomarkers Study. J. Parkinsons Dis. 2020, 10, 505–510. [Google Scholar] [CrossRef]
  172. Palacios, N.; Gao, X.; McCullough, M.L.; Schwarzschild, M.A.; Shah, R.; Gapstur, S.; Ascherio, A. Caffeine and risk of Parkinson’s disease in a large cohort of men and women. Mov. Disord. 2012, 27, 1276–1282. [Google Scholar] [CrossRef] [Green Version]
  173. Kumar, P.M.; Paing, S.S.; Li, H.; Pavanni, R.; Yuen, Y.; Zhao, Y.; Tan, E.K. Differential effect of caffeine intake in subjects with genetic susceptibility to Parkinson’s Disease. Sci. Rep. 2015, 5, 15492. [Google Scholar] [CrossRef] [Green Version]
  174. Tanner, C.M.; Meng, C.; Marder, K.; Bressman, S.; Saunders-Pullman, R.; Alcalay, R.; Tolosa, E.; Brice, A.; Goldman, S.; Schuele, B.; et al. LRRK2 Cohort-Consortium, Caffeinated Drinks, LRRK2 Genotype and PD [abstract]. Mov. Disord. 2017, 32 (Supp. 2), S398. [Google Scholar]
  175. Yahalom, G.; Rigbi, A.; Israeli-Korn, S.; Krohn, L.; Rudakou, U.; Ruskey, J.A.; Benshimol, L.; Tsafnat, T.; Gan-Or, Z.; Hassin-Baer, S.; et al. Age at Onset of Parkinson’s Disease Among Ashkenazi Jewish Patients: Contribution of Environmental Factors, LRRK2 p.G2019S and GBA p.N370S Mutations. J. Parkinson’s Dis. 2020, 10, 1123–1132. [Google Scholar] [CrossRef]
  176. Crotty, G.F.; Wang, J.; Montalban, M.; Davis, S.; Alkabsh, J.; Bakshi, R.; Chen, X.; Ascherio, A.; Macklin, E.; Maciuca, R.; et al. Metabolomic Analysis Identifies Caffeine and its Metabolites as Plasma Markers of Resistance to Parkinson’s Disease Among LRRK2 Mutation Carriers in the LRRK2 Cohort Consortium (LCC) [abstract]. Mov. Disord. Clin. Pract. 2020, 7, S183. [Google Scholar]
  177. Ikeda, K.; Kurokawa, M.; Aoyama, S.; Kuwana, Y. Neuroprotection by adenosine A2A receptor blockade in experimental models of Parkinson’s disease. J. Neurochem. 2002, 80, 262–270. [Google Scholar] [CrossRef] [Green Version]
  178. Xu, K.; Xu, Y.H.; Chen, J.F.; Schwarzschild, M.A. Caffeine’s neuroprotection against 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine toxicity shows no tolerance to chronic caffeine administration in mice. Neurosci. Lett. 2002, 322, 13–16. [Google Scholar] [CrossRef]
  179. Aguiar, L.M.; Nobre, H.V., Jr.; Macedo, D.S.; Oliveira, A.A.; Freitas, R.M.; Vasconcelos, S.M.; Cunha, G.M.; Sousa, F.C.; Viana, G.S. Neuroprotective effects of caffeine in the model of 6-hydroxydopamine lesion in rats. Pharmacol. Biochem. Behav. 2006, 84, 415–419. [Google Scholar] [CrossRef] [PubMed]
  180. Souza, M.F.d.; Bispo, J.M.M.; Leal, P.C.; Gois, A.M.d.; Santos, J.R.D. Commentary: Adenosine A(2A) Receptor Blockade Prevents Rotenone-Induced Motor Impairment in a Rat Model of Parkinsonism. Front. Behav. Neurosci. 2017, 11, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Fathalla, A.M.; Soliman, A.M.; Moustafa, A.A. Selective A2A receptors blockade reduces degeneration of substantia nigra dopamine neurons in a rotenone-induced rat model of Parkinson’s disease: A histological study. Neurosci. Lett. 2017, 643, 89–96. [Google Scholar] [CrossRef]
  182. Kalda, A.; Yu, L.; Oztas, E.; Chen, J.F. Novel neuroprotection by caffeine and adenosine A2A receptor antagonists in animal models of Parkinson’s disease. J. Neurol. Sci. 2006, 248, 9–15. [Google Scholar] [CrossRef]
  183. Luan, Y.; Ren, X.; Zheng, W.; Zeng, Z.; Guo, Y.; Hou, Z.; Guo, W.; Chen, X.; Li, F.; Chen, J.F. Chronic Caffeine Treatment Protects Against alpha-Synucleinopathy by Reestablishing Autophagy Activity in the Mouse Striatum. Front. Neurosci. 2018, 12, 301. [Google Scholar] [CrossRef]
  184. Karuppagounder, S.S.; Uthaythas, S.; Govindarajulu, M.; Ramesh, S.; Parameshwaran, K.; Dhanasekaran, M. Caffeine, a natural methylxanthine nutraceutical, exerts dopaminergic neuroprotection. Neurochem. Int. 2021, 148, 105066. [Google Scholar] [CrossRef]
  185. Bagga, P.; Chugani, A.N.; Patel, A.B. Neuroprotective effects of caffeine in MPTP model of Parkinson’s disease: A (13)C NMR study. Neurochem. Int. 2016, 92, 25–34. [Google Scholar] [CrossRef]
  186. Sonsalla, P.K.; Wong, L.Y.; Harris, S.L.; Richardson, J.R.; Khobahy, I.; Li, W.; Gadad, B.S.; German, D.C. Delayed caffeine treatment prevents nigral dopamine neuron loss in a progressive rat model of Parkinson’s disease. Exp. Neurol. 2012, 234, 482–487. [Google Scholar] [CrossRef] [Green Version]
  187. Kachroo, A.; Schwarzschild, M.A. Adenosine A2A receptor gene disruption protects in an alpha-synuclein model of Parkinson’s disease. Ann. Neurol. 2012, 71, 278–282. [Google Scholar] [CrossRef] [Green Version]
  188. Orr, A.G.; Hsiao, E.C.; Wang, M.M.; Ho, K.; Kim, D.H.; Wang, X.; Guo, W.; Kang, J.; Yu, G.Q.; Adame, A.; et al. Astrocytic adenosine receptor A2A and Gs-coupled signaling regulate memory. Nat. Neurosci. 2015, 18, 423–434. [Google Scholar] [CrossRef] [Green Version]
  189. Temido-Ferreira, M.; Ferreira, D.G.; Batalha, V.L.; Marques-Morgado, I.; Coelho, J.E.; Pereira, P.; Gomes, R.; Pinto, A.; Carvalho, S.; Canas, P.M.; et al. Age-related shift in LTD is dependent on neuronal adenosine A2A receptors interplay with mGluR5 and NMDA receptors. Mol. Psychiatry 2020, 25, 1876–1900. [Google Scholar] [CrossRef]
  190. Silva, S.V.d.; Haberl, M.G.; Zhang, P.; Bethge, P.; Lemos, C.; Gonçalves, N.; Gorlewicz, A.; Malezieux, M.; Gonçalves, F.Q.; Grosjean, N.; et al. Early synaptic deficits in the APP/PS1 mouse model of Alzheimer’s disease involve neuronal adenosine A2A receptors. Nat. Commun. 2016, 7, 11915. [Google Scholar] [CrossRef]
  191. Merighi, S.; Battistello, E.; Casetta, I.; Gragnaniello, D.; Poloni, T.E.; Medici, V.; Cirrincione, A.; Varani, K.; Vincenzi, F.; Borea, P.A.; et al. Upregulation of Cortical A2A Adenosine Receptors Is Reflected in Platelets of Patients with Alzheimer’s Disease. J. Alzheimers Dis. 2021, 80, 1105–1117. [Google Scholar] [CrossRef]
  192. Ng, S.K.; Higashimori, H.; Tolman, M.; Yang, Y. Suppression of adenosine 2A receptor (A2AR)-mediated adenosine signaling improves disease phenotypes in a mouse model of amyotrophic lateral sclerosis. Exp. Neurol. 2015, 267, 115–122. [Google Scholar] [CrossRef] [Green Version]
  193. Mills, J.H.; Kim, D.G.; Krenz, A.; Chen, J.F.; Bynoe, M.S. A2A adenosine receptor signaling in lymphocytes and the central nervous system regulates inflammation during experimental autoimmune encephalomyelitis. J. Immunol. 2012, 188, 5713–5722. [Google Scholar] [CrossRef] [Green Version]
  194. Vincenzi, F.; Corciulo, C.; Targa, M.; Merighi, S.; Gessi, S.; Casetta, I.; Gentile, M.; Granieri, E.; Borea, P.A.; Varani, K. Multiple sclerosis lymphocytes upregulate A2A adenosine receptors that are antiinflammatory when stimulated. Eur. J. Immunol. 2013, 43, 2206–2216. [Google Scholar] [CrossRef]
  195. Rissanen, E.; Virta, J.R.; Paavilainen, T.; Tuisku, J.; Helin, S.; Luoto, P.; Parkkola, R.; Rinne, J.O.; Airas, L. Adenosine A2A receptors in secondary progressive multiple sclerosis: A [(11)C]TMSX brain PET study. J. Cereb. Blood Flow Metab. 2013, 33, 1394–1401. [Google Scholar] [CrossRef] [Green Version]
  196. Orr, A.G.; Lo, I.; Schumacher, H.; Ho, K.; Gill, M.; Guo, W.; Kim, D.H.; Knox, A.; Saito, T.; Saido, T.C.; et al. Istradefylline reduces memory deficits in aging mice with amyloid pathology. Neurobiol. Dis. 2018, 110, 29–36. [Google Scholar] [CrossRef] [PubMed]
  197. Laurent, C.; Burnouf, S.; Ferry, B.; Batalha, V.L.; Coelho, J.E.; Baqi, Y.; Malik, E.; Mariciniak, E.; Parrot, S.; Jeugd, A.V.d.; et al. A2A adenosine receptor deletion is protective in a mouse model of Tauopathy. Mol. Psychiatry 2016, 21, 97–107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Potenza, R.L.; Armida, M.; Ferrante, A.; Pèzzola, A.; Matteucci, A.; Puopolo, M.; Popoli, P. Effects of chronic caffeine intake in a mouse model of amyotrophic lateral sclerosis. J. Neurosci. Res. 2013, 91, 585–592. [Google Scholar] [CrossRef] [PubMed]
  199. Rei, N.; Rombo, D.M.; Ferreira, M.F.; Baqi, Y.; Müller, C.E.; Ribeiro, J.A.; Sebastião, A.M.; Vaz, S.H. Hippocampal synaptic dysfunction in the SOD1(G93A) mouse model of Amyotrophic Lateral Sclerosis: Reversal by adenosine A2AR blockade. Neuropharmacology 2020, 171, 108106. [Google Scholar] [CrossRef] [PubMed]
  200. Villar-Menéndez, I.; Blanch, M.; Tyebji, S.; Pereira-Veiga, T.; Albasanz, J.L.; Martín, M.; Ferrer, I.; Pérez-Navarro, E.; Barrachina, M. Increased 5-methylcytosine and decreased 5-hydroxymethylcytosine levels are associated with reduced striatal A2AR levels in Huntington’s disease. Neuromolecular Med. 2013, 15, 295–309. [Google Scholar] [CrossRef]
  201. Tavares, A.; Barret, O.; Seibyl, J.; Tamagnan, G. Imaging studies with A2A receptor antagonists. In The Adenosinergic System: A Non-Dopaminergic Target in Parkinson’s Disease; Morelli, M., Simola, N., Wardas, J., Eds.; Springer: New York, NY, USA, 2015; Volume 10, pp. 206–232. [Google Scholar]
  202. Mievis, S.; Blum, D.; Ledent, C. A2A receptor knockout worsens survival and motor behaviour in a transgenic mouse model of Huntington’s disease. Neurobiol. Dis. 2011, 41, 570–576. [Google Scholar] [CrossRef]
  203. Ishiwata, K.; Ogi, N.; Hayakawa, N.; Oda, K.; Nagaoka, T.; Toyama, H.; Suzuki, F.; Endo, K.; Tanaka, A.; Senda, M. Adenosine A2A receptor imaging with [11C]KF18446 PET in the rat brain after quinolinic acid lesion: Comparison with the dopamine receptor imaging. Ann. Nucl. Med. 2002, 16, 467–475. [Google Scholar] [CrossRef]
  204. Scattoni, M.L.; Valanzano, A.; Pezzola, A.; March, Z.D.; Fusco, F.R.; Popoli, P.; Calamandrei, G. Adenosine A2A receptor blockade before striatal excitotoxic lesions prevents long term behavioural disturbances in the quinolinic rat model of Huntington’s disease. Behav. Brain Res. 2007, 176, 216–221. [Google Scholar] [CrossRef]
  205. Galluzzo, M.; Pintor, A.; Pèzzola, A.; Grieco, R.; Borsini, F.; Popoli, P. Behavioural and neurochemical characterization of the adenosine A2A receptor antagonist ST1535. Eur. J. Pharmacol. 2008, 579, 149–152. [Google Scholar] [CrossRef]
  206. Cekic, C.; Linden, J. Purinergic regulation of the immune system. Nat. Rev. Immunol. 2016, 16, 177–192. [Google Scholar] [CrossRef]
  207. Chen, Y.; Zhang, Z.X.; Zheng, L.P.; Wang, L.; Liu, Y.F.; Yin, W.Y.; Chen, Y.Y.; Wang, X.S.; Hou, S.T.; Chen, J.F.; et al. The adenosine A2A receptor antagonist SCH58261 reduces macrophage/microglia activation and protects against experimental autoimmune encephalomyelitis in mice. Neurochem. Int. 2019, 129, 104490. [Google Scholar] [CrossRef]
  208. Liu, Y.; Zou, H.; Zhao, P.; Sun, B.; Wang, J.; Kong, Q.; Mu, L.; Zhao, S.; Wang, G.; Wang, D.; et al. Activation of the adenosine A2A receptor attenuates experimental autoimmune encephalomyelitis and is associated with increased intracellular calcium levels. Neuroscience 2016, 330, 150–161. [Google Scholar] [CrossRef]
  209. Yao, S.Q.; Li, Z.Z.; Huang, Q.Y.; Li, F.; Wang, Z.W.; Augusto, E.; He, J.C.; Wang, X.T.; Chen, J.F.; Zheng, R.Y. Genetic inactivation of the adenosine A2A receptor exacerbates brain damage in mice with experimental autoimmune encephalomyelitis. J. Neurochem. 2012, 123, 100–112. [Google Scholar] [CrossRef]
Figure 1. The direct and indirect pathways work together to control movement.
Figure 1. The direct and indirect pathways work together to control movement.
Molecules 27 02366 g001
Figure 2. Two-state model of adenosine A2A receptor and its major intracellular signal transduction.
Figure 2. Two-state model of adenosine A2A receptor and its major intracellular signal transduction.
Molecules 27 02366 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mori, A.; Chen, J.-F.; Uchida, S.; Durlach, C.; King, S.M.; Jenner, P. The Pharmacological Potential of Adenosine A2A Receptor Antagonists for Treating Parkinson’s Disease. Molecules 2022, 27, 2366. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27072366

AMA Style

Mori A, Chen J-F, Uchida S, Durlach C, King SM, Jenner P. The Pharmacological Potential of Adenosine A2A Receptor Antagonists for Treating Parkinson’s Disease. Molecules. 2022; 27(7):2366. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27072366

Chicago/Turabian Style

Mori, Akihisa, Jiang-Fan Chen, Shinichi Uchida, Cecile Durlach, Shelby M. King, and Peter Jenner. 2022. "The Pharmacological Potential of Adenosine A2A Receptor Antagonists for Treating Parkinson’s Disease" Molecules 27, no. 7: 2366. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27072366

Article Metrics

Back to TopTop