Next Article in Journal
Purification and Identification of Novel Myeloperoxidase Inhibitory Antioxidant Peptides from Tuna (Thunnas albacares) Protein Hydrolysates
Next Article in Special Issue
Combined Treatment with PI3K Inhibitors BYL-719 and CAL-101 Is a Promising Antiproliferative Strategy in Human Rhabdomyosarcoma Cells
Previous Article in Journal
Dating Sediments by EPR Using Al-h Centre: A Comparison between the Properties of Fine (4–11 µm) and Coarse (>63 µm) Quartz Grains
Previous Article in Special Issue
Advances in the Development of SARS-CoV-2 Mpro Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pathophysiological Role and Medicinal Chemistry of A2A Adenosine Receptor Antagonists in Alzheimer’s Disease

1
Department of Translational Medicine and for Romagna, University of Ferrara, 44121 Ferrara, Italy
2
Department of Chemical, Pharmaceutical and Agricultural Sciences, University of Ferrara, 44121 Ferrara, Italy
3
Molecular Recognition Section, Laboratory of Bioorganic Chemistry, National Institute of Diabetes and Digestive and Kidney Diseases, Bethesda, MD 20892, USA
4
Laboratory of Neuroimaging, National Institute on Alcohol Abuse and Alcoholism, National Institutes of Health, Bethesda, MD 20892, USA
5
National Institute on Drug Abuse, National Institutes of Health, Bethesda, MD 20892, USA
*
Authors to whom correspondence should be addressed.
Submission received: 18 March 2022 / Revised: 14 April 2022 / Accepted: 18 April 2022 / Published: 21 April 2022
(This article belongs to the Special Issue Featured Papers in Medicinal Chemistry)

Abstract

:
The A2A adenosine receptor is a protein belonging to a family of four GPCR adenosine receptors. It is involved in the regulation of several pathophysiological conditions in both the central nervous system and periphery. In the brain, its localization at pre- and postsynaptic level in striatum, cortex, hippocampus and its effects on glutamate release, microglia and astrocyte activation account for a crucial role in neurodegenerative diseases, including Alzheimer’s disease (AD). This ailment is considered the main form of dementia and is expected to exponentially increase in coming years. The pathological tracts of AD include amyloid peptide-β extracellular accumulation and tau hyperphosphorylation, causing neuronal cell death, cognitive deficit, and memory loss. Interestingly, in vitro and in vivo studies have demonstrated that A2A adenosine receptor antagonists may counteract each of these clinical signs, representing an important new strategy to fight a disease for which unfortunately only symptomatic drugs are available. This review offers a brief overview of the biological effects mediated by A2A adenosine receptors in AD animal and human studies and reports the state of the art of A2A adenosine receptor antagonists currently in clinical trials. As an original approach, it focuses on the crucial role of pharmacokinetics and ability to pass the blood–brain barrier in the discovery of new agents for treating CNS disorders. Considering that A2A receptor antagonist istradefylline is already commercially available for Parkinson’s disease treatment, if the proof of concept of these ligands in AD is confirmed and reinforced, it will be easier to offer a new hope for AD patients.

1. Introduction

Adenosine is a ubiquitous nucleoside derived largely from the hydrolysis of ATP, which acts as an extracellular autacoid to regulate a wide number of pathophysiological functions. The intracellular production of adenosine arises from the action of cytosolic 5′-nucleotidase on AMP or the action of SAH hydrolase on S-adenosyl-homocysteine (SAH) hydrolysis. Conversely, an extracellular pathway predominates when cells undergo increased metabolic activity or stress, e.g., in pathological conditions. In this case, the main source of adenosine is from the activity of specific enzymes known as ectonucleoside triphosphate diphosphohydrolases (CD39s), which convert ATP into AMP, and ecto-5′-nucleotidase (CD73), which dephosphorylates AMP to produce adenosine [1]. Upon generation, the extracellular nucleoside enters cells through both concentrative and equilibrative nucleoside transporters, CNTs (SLC28 family) and ENTs (SLC29 family), respectively. These proteins are crucial to fine-tuning adenosine levels, with the ENTs playing the more important role. CNT proteins increase adenosine from the extracellular to intracellular pools driven by the sodium gradient, while ENTs allow bidirectional transport, following the nucleoside concentration gradient across the membrane. Normally, adenosine produced extracellularly enters into cells by these transporters, while during hypoxia this flux can be reversed [2].
The amount of extracellular adenosine is also regulated by its metabolism through the enzyme adenosine deaminase (ADA), which transforms it into inosine. This type of regulation occurs also in the intracellular pool, where adenosine may be either converted into inosine by intracellular ADA or phosphorylated to AMP by adenosine kinase (ADK). Due to the different Km of these enzymes for adenosine, with ADK being more potent than ADA, under physiological conditions AMP generation is predominant, while in stressful environments its conversion to inosine principally occurs [3].
Adenosine effects are due to the stimulation of four transmembrane receptors, termed A1, A2A, A2B and A3 adenosine receptors (ARs) [4]. These AR subtypes belong to the family of G protein-coupled receptors, either Gi-(A1, A3) or Gs-(A2A, A2B) coupled. Altogether, they are balanced in modulating adenylyl cyclase activity, with inhibitory or stimulatory effects depending on Gi or Gs coupling, respectively. In addition, the A2A subtype in the striatum triggers recruitment of Golf proteins [5,6]. Furthermore, the ARs modulate mitogen-activated protein kinases (MAPK) p38, ERK1/2, and JNK1/2 as well as Akt phosphorylation, thus regulating important intracellular signal transduction pathways, potentially affecting a wide number of pathophysiological functions [7,8,9,10,11].
Although all ARs are potential pharmaceutical targets for a variety of human diseases, the A2A subtype is the one for which the understanding of biological relationships and the discovery of potent and selective ligands are further along a translational path. The first clinically approved selective A2AAR antagonist istradefylline, a caffeine analogue, is now on the market for Parkinson’s disease (PD) [12,13]. Therefore, this review will focus on a survey of recent literature on the A2A adenosine receptor as a druggable target for another important neurodegenerative disorder, i.e., Alzheimer’s disease (AD), considering the present technological state and directions for improving drug discovery and development from both biological and chemical perspectives.

2. A2A Receptors in AD

AD is a multifactorial neurodegenerative pathology that is experiencing exponential growth in the aging population, which is expected to triple by 2050 [14]. Its pathological indications include a progressive appearance of amnesic disabilities, associated with aphasia, apraxia, and agnosia. Within the scope of this disease are behavioral problems, consisting of problem solving and orientation disorders. Current pharmacological strategies to treat AD patients are to increase synaptic acetylcholine (ACh) concentrations, using the cholinesterase inhibitors donepezil, rivastigmine, galantamine, or to reduce glutamate excitotoxicity through memantine [15,16]. Unfortunately, these drugs offer merely symptomatic relief, and they do not impede disease progression, thus making the search for new therapeutic strategies urgent. Various novel pharmacological targets aim to accelerate the clearance of β amyloid (Aβ) peptide and tau protein, or to impede neurodegenerative molecular pathways [17,18]. This is an ambitious challenge, but a new drug targeting Aβ protein accumulation, named aducanumab, recently entered the market, with still controversial clinical efficacy [18,19,20]. However, due to the failure of numerous clinical trials with drugs directed towards Aβ protein, it is questionable whether β-amyloid is certainly an effective target. Another strategy for the treatment of AD concerns tau neurofibrillary tangles that are strongly correlated with local neurodegeneration and cognitive impairment in AD. As a result, several approaches for lowering tau pathogenicity have been suggested, and tau-targeted active and passive immunotherapies are now being investigated for safety and effectiveness in patients with early-stage AD. However, because no clinical trials have been finished to date, no assumptions about the therapeutic effects of therapies for tau pathology can be established [16]. Interestingly, caffeine, the most commonly ingested A2A receptor antagonist, as well as genetic elimination of A2A receptors in mice, decreased hippocampus tau hyperphosphorylation, fought neuroinflammation, and restored the associated memory loss [13].
Indeed, the A2A adenosine receptor may represent a novel and pleiotropic protein through which it is possible to alter the disease course. The A2A adenosine receptor has been extensively characterized biochemically and structurally, with many X-ray and cryo-EM structures now available [12]. A lengthy intracellular carboxy-terminal tail, where phosphorylation and palmitoylation activities may cause receptor desensitization and internalization, is a structural feature of the A2A subtype that other family members lack [21]. This subtype forms heteromers with other receptors, such as A1 adenosine and D2 dopamine receptors, that have different pharmacological properties compared to the receptor monomers [22]. In the central nervous system, the A2A receptor is found at the highest levels in striatum, olfactory tubercle, and immune system, while lower amounts are present in the cerebral cortex, hippocampus and vasculature [23,24]. In any case, low expression does not imply lesser importance; indeed, cerebrocortical A2A receptors account for most of the more relevant effects attributed to brain A2A receptors (e.g., mood, memory). Likewise, proper recognition must be given to the sparse but highly significant cortical A2A receptors.
A2A receptors in the brain are mostly detected in glutamatergic synapses, where the major consequence of their activation at the presynaptic level is reflected in increased glutamate release, which contributes to excitotoxicity. Indeed, under physiological conditions, hippocampal glutamate release is under the control of an inhibitory A1 receptor that, following adenosine increase, is counteracted by a presynaptic A2A receptor resulting in the release of this excitatory neurotransmitter, possibly through induction of PKA-dependent calcium currents [25,26]. Postsynaptically, the A2A receptor promotes long-term potentiation (LTP) by facilitating AMPA receptor-mediated currents, NMDA phosphorylation that is mGluR5-dependent and a Ca2+ increase [27,28,29,30]. The A2A receptor reduces glutamate uptake in astrocytes, through modulating the expression of glutamate transporters, GLT-1 and GLAST, and induces astrocytic proliferation and activation. In microglia it plays an important role in the induction of proliferation, activation, process retraction and the production of molecules involved in inflammation, such as NO, PGE2, IL1β and cyclooxygenase [31] (Figure 1). In addition to affecting the inflammatory response, A2A receptor promotes BDNF and GDNF secretion from microglia and astrocytes, respectively, in response to LPS [1].
It is widely recognized that glutamatergic synaptic failure and temporal lobe degeneration are the earliest signs of cognitive decline, preceding the production of Aβ plaques and tau-derived neurofibrillary tangles [32]. Indeed, individuals with mild cognitive impairment (MCI) and early AD show synaptic loss in the hippocampus and posterior cingulate gyrus, indicating that this event is the starting point for memory loss [33,34,35]. Therefore, A2A adenosine receptor engagement in glutamatergic synaptic physiology has suggested a relationship between this subtype and AD. The A2A adenosine receptor expression in both hippocampal neurons and astrocytes is elevated in aging, AD animal models, and AD patients, elevating glutamate release, calcium influx, the LTP-to-long-term depression transition, and cognitive decline [26,36,37,38,39,40,41,42,43,44,45,46]. Overall, this evidence indicates that the important contribution of the A2A adenosine receptor to synaptic and neuronal damage by increasing glutamate excitotoxicity and neuroinflammation, as evidenced by the neuroprotection provided by genetic or pharmacological A2A adenosine receptor blockade. Indeed, A2A receptor antagonism reduces hippocampus-dependent memory impairment and LTP changes in aged animals and AD models [36,47,48,49,50,51]. In addition, A2A adenosine receptor genetic knockdown can ameliorate synaptic damage present in AD models, and numerous studies have found that A2A adenosine receptor antagonists are effective at improving cognition after synaptic loss in AD animal models, providing a way to combat synaptic toxicity [40,52,53,54]. Adenosine levels in postmortem AD brains are higher in the parietal and temporal lobes than in the frontal cortex, implying enhanced A2A receptor activation in overexpressing areas [55]. A study of A2A receptor photoaffinity labeling in post-mortem striatal membranes provided the first hint that the A2A receptor could be enhanced in the AD brain [56]. Indeed, in comparison to healthy, post-mortem subjects, brains from AD patients had a higher density of A2A receptors. More interestingly, in comparison to both frontal grey matter and frontal white matter regions in AD but not in normal participants, the overexpression was observed primarily in the hippocampus and entorhinal areas, suggesting that the upregulation is likely connected to the existence of AD pathology. Indeed, A2A expression appears to replicate the same AD scenario that extends in the brain from hippocampal/entorhinal structures to other cortical areas, with the hippocampal/entorhinal cortex having the highest density of pathology. As a result, the frontal white matter region, where the AD pathology is less prominent, has the lowest A2A expression [24]. This leads to the idea that, once reliable PET ligands are validated, assessing A2A receptor density in diseased brain circuits may become an essential biomarker of vulnerability to detect development of AD [57].
A further crucial component of A2A adenosine receptor action in AD is its control of neuroinflammation via glial cells. The A2A adenosine receptor was shown to be overexpressed in the astrocytes of AD patients, and its genetic silencing in both young and old mice improved long-term memory [40,58,59]. A2A receptors in astrocytes were required for fine-tuning inhibitory and excitatory regulation of synaptic transmission through GABA and glutamate uptake. A1 and A2A receptors, which can occur together as heterodimers, controlled GABA transport in different ways, with the A1 blocking and the A2A boosting it [60]. Furthermore, glutamate uptake was hampered by A2A receptor activation, which is required for the decrease of synaptic glutamate transporters GLAST and GLT-I that is produced by Aβ peptide [61]. An A2A receptor antagonist or receptor knockout prevented the Aβ-induced decrease in glutamate transporters and astrogliosis. A2A receptor overexpression causes significant transcriptional alterations, primarily impacting genes involved in immune responses, angiogenesis, and cell activation. Treatment with SCH58261 (Figure 2), a selective A2A antagonist, restored the expression levels of numerous inflammatory and astrocytic activation-related genes. This reinforces the concept that blocking the A2A receptor might reverse some astrocytic disorders caused by aberrant A2A receptor expression, implying that receptor antagonists could be useful in A2A receptor-induced transcriptional dysregulation, inflammation, and astrogliosis. Altogether, these findings shed light on the possible influence of A2A receptor upregulation on astrocyte transcriptional dysregulation, paving the way for the development of A2A receptor antagonists as a potential therapy for AD [62]. Activated microglia, in addition to astrocytes, are relevant cells for neuroinflammation and, interestingly, they overexpress the A2A receptor. Microglia in AD also express the NMDA receptor, which is a crucial target in the fight against AD and might be counteracted by inhibiting the A2A receptor through their crosstalk in heteromeric structures [63,64]. Furthermore, by inhibiting A2A receptor-mediated cytokine production, this strategy may be able to reduce memory loss by attenuating neuroinflammation [65,66].
We already have proof that blocking A2A improves memory in humans. As a matter of fact, caffeine, the most commonly ingested A2A receptor antagonist, can ameliorate cognition providing the evidence for the A2A receptor’s relevance in AD through different strategies. Indeed, caffeine decreased hippocampal tau hyperphosphorylation, attenuated neuroinflammation and reversed the associated memory loss [12,50,53,82]. The Effect of CAFfeine on Cognition in Alzheimer’s Disease (CAFCA) is under investigation in the Phase 3 clinical trial NCT04570085. It is a multicenter, randomized, double-blind, placebo-controlled trial evaluating the effect of a 30-week caffeine treatment on cognition in AD at beginning to moderate stages (MMSE 16–24). The results of human research will help to clarify the function of caffeine in the prevention of AD. In addition, istradefylline, a potent and selective A2A antagonist, is already in use in the clinic for another neurodegenerative condition, PD. It was approved earlier in Japan and Korea, and in August 2019 in the United States after passing various clinical safety and efficacy tests (as Nouriast) [73,83,84,85]. As for its role in AD, low istradefylline doses improved spatial memory and habituation in animal models of AD, providing important confirmation that A2A receptor blockade could be a novel target for countering cognitive deficits in AD patients [59,71]. This finding shows that A2A receptor-targeting drugs, such as istradefylline and hopefully novel ones, have a great deal of promise for treating dementia.

3. Chemistry of A2A Receptor Antagonists in Clinical Trials

Many selective A2A antagonists have been reported and some have entered clinical trials (Table 1) [12,67,68,86] for either neurodegenerative diseases or cancer immunotherapy (including by co-administration with checkpoint inhibitors and other cancer immunotherapies). It is to be noted that A2B antagonists and mixed A2A/A2B antagonists are also being evaluated in cancer immunotherapy [80]. Although not covered in this review, A2B antagonists may also have neuroprotective properties [87]. The naturally occurring alkylxanthine caffeine, a weak non-selective adenosine antagonist, is in a Phase 3 clinical trial for AD in France [50]. Recently, the analogue of caffeine with full deuteration of the three methyl groups was shown to have a prolonged in vivo plasma half-life in the rat (from 1.9 h to 5 h) without losing adenosine receptor binding affinity [88].
Istradefylline 3 [72], a moderately potent (pKi 7.44) and selective A2A antagonist that is structurally related to caffeine (Figure 2), is the only such ligand that is FDA-approved for clinical use, i.e., as a co-therapy with L-DOPA in PD treatment. A total of 23 clinical trials have been carried out to probe the safety and efficacy of istradefylline in PD and also to study the effects of hepatic impairment and antibiotic rifampicin on its metabolism.
Unlike some of the more recently introduced A2A antagonists for cancer, tozadenant 6 (SYN115) was initially developed for treatment of PD and later for cocaine addiction [67,89,90,91]. Tozadenant has been studied both as a PD monotherapy and as a cotherapy with L-DOPA or dopamine agonists. Clinical efficacy in reducing off-time was observed without severe adverse effects in initial clinical trials, but a Phase 3 clinical trial turned disastrous with seven cases of sepsis out of 890 patients and six fatalities [86]. Therefore, its clinical development was terminated in 2018. Its potential use in treating cocaine addiction was previously studied [89], but there was insufficient clinical efficacy to proceed with this target. Tozadenant analogues were recently reported as A2A antagonists, including 18F-labeled radiotracers for potential PET imaging [92]. Vipadenant 7 [67] is a potent A2A antagonist (pKi 8.89) that was introduced for PD [74,93] and its development halted due to safety issues [68], but it later transitioned to use in cancer immunotherapy. Other 18F-labeled A2A antagonists related to vipadenant were reported by Lai et al. [94]. Preladenant 12 followed a similar developmental path, first being introduced for PD and later repurposed to cancer [68,95]. The 11C forms of a predecessor of preladenant, i.e., SCH442416 10, and preladenant itself, have been shown to be useful as A2A receptor PET imaging agents [94]. The use of structure-based approaches has accelerated the discovery of A2A antagonists. The earliest success of a structurally designed A2A antagonist was imaradenant 22. It was initially discovered using an inactive conformation-stabilized A2A receptor, enabled by the ability to readily determine the three-dimensional structure of A2A antagonist-bound receptors by strategic stabilizing mutations [96]. Early in its development it was positioned for use in treating attention deficit hyperactive disorder (ADHD), but it was repurposed for cancer immunotherapy. Taminadenant 24 and ciforadenant 25 are being developed for cancer immunotherapy [77,78], but ciforadenant has a long history of clinical development, initially for CNS application. An analogue of taminadenat (PBF-999, structure not shown) displayed dual A2A receptor and phosphodiesterase (PDE)10 activity and is now in a clinical trial for cancer [68,94].
The role of pharmacokinetics and ability to pass the blood–brain barrier (BBB) are important factors in the discovery of new agents for treating CNS disorders. There are now protocols for predicting the likelihood that a new compound will cross the BBB with appropriate metabolic stability, such as the multi-parameter optimization (MPO) score [97,98,99,100]. The MPO score estimation for CNS drugs, along with binding/functional affinity measurement could be useful tool to predict and design radiotracers for PET. Use of such algorithms will aid in the discovery of future A2A antagonists for the treatment of CNS disorders, such as PD and AD, or for treatment of peripheral conditions in which CNS penetration would cause side effects. Nevertheless, it must be kept in mind that multiple physicochemical and biochemical factors influence the suitability of a given compound for use as a CNS drug or imaging agent [100].
Table 2 shows the MPO scores calculated using standard commercial software of common A2A antagonists, including those in clinical trials and those for use as PET imaging agents [98,99]. This is a means of predicting which antagonists might be suitable for use in treatment of neurodegenerative diseases. For CNS drug candidates expected to be readily brain penetrant, the high clogBB (log([brain]:[blood]) value (>−0.2) and “no” substrate for the brain efflux pumps (e.g., P-gp), and a higher Intravenous CNS Profile Score [100] are preferable. A logP value <4 is one of many parameters predictive of effective oral dosing [100]. Thus, alkylxanthines caffeine and DMPX readily cross the BBB with clogBB values of −0.0323 and −0.180. Most of the compounds shown in Table 1 have the ability to cross the BBB to varying degrees, but an anti-tumor agent, inupadenant (clogBB −1.25), is not brain penetrant.
Furthermore, the application of in vivo imaging with PET ligands has become an important component in CNS drug development to determine target engagement and occupancy related to duration of drug action. Fortunately, there are numerous A2A antagonist PET ligands developed over decades, resulting in [18F]MNI-444 ([18F] 12) and [11C]preladenant ([18F] 11), which demonstrated high specific binding signal [82] as clinical imaging tools. The interaction of A2A antagonists and the D2 dopamine receptor was also studied in humans using istradefylline (40 mg QD) and PET imaging of dopamine receptors with [11C]raclopride [104].

4. Conclusions

This paper presents a brief overview of A2A receptor antagonist clinical use in the context of drug discovery for AD therapy. Many new A2A antagonists are being tested clinically, mostly for cancer, and some of these compounds are able to cross the BBB, which would enable their use in AD. Overexpression and functional upregulation of the A2A receptor is associated with early changes leading to progressive neurodegeneration. Thus, PET ligands that are validated for A2A receptor imaging in the human brain could be useful for the early clinical diagnosis of AD or other neurodegenerative conditions, which is an unmet medical need. Given the generality of the involvement of this receptor in brain injury and other pathology, the clinical development of A2A receptor antagonists for treatment of these conditions promises to be on the verge of becoming a reality.

Author Contributions

Conceptualization, S.M., K.A.J., S.G. and S.W.K.; writing—original draft preparation, A.T., M.N., R.R.S. and S.P., writing—review and editing, F.V., K.A.J., S.W.K. and N.D.V.; visualization, P.A.B. and K.V.; supervision, S.M., S.G. and K.A.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by University of Ferrara, NIDDK Intramural Res. Program (ZIADK031117, K.A.J.) and NIAAA Intramural Res. Program (Y1AA-3009, N.D.V.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Borea, P.A.; Gessi, S.; Merighi, S.; Vincenzi, F.; Varani, K. Pharmacology of Adenosine Receptors: The State of the Art. Physiol. Rev. 2018, 98, 1591–1625. [Google Scholar] [CrossRef] [PubMed]
  2. Pastor-Anglada, M.; Pérez-Torras, S. Who Is Who in Adenosine Transport. Front. Pharmacol. 2018, 9, 627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Borea, P.A.; Gessi, S.; Merighi, S.; Varani, K. Adenosine as a Multi-Signalling Guardian Angel in Human Diseases: When, Where and How Does it Exert its Protective Effects? Trends Pharmacol. Sci. 2016, 37, 419–434. [Google Scholar] [CrossRef] [PubMed]
  4. Borea, P.A.; Varani, K.; Vincenzi, F.; Baraldi, P.G.; Tabrizi, M.A.; Merighi, S.; Gessi, S. The A3 adenosine receptor: History and perspectives. Pharmacol. Rev. 2015, 67, 74–102. [Google Scholar] [CrossRef] [Green Version]
  5. Kull, B.; Svenningsson, P.; Fredholm, B.B. Adenosine A2A receptors are colocalized with and activate g(olf) in rat striatum. Mol. Pharmacol. 2000, 58, 771–777. [Google Scholar] [CrossRef] [Green Version]
  6. Schwindinger, W.F.; Mihalcik, L.J.; Giger, K.E.; Betz, K.S.; Stauffer, A.M.; Linden, J.; Herve, D.; Robishaw, J.D. Adenosine A2A receptor signaling and golf assembly show a specific requirement for the gamma7 subtype in the striatum. J. Biol. Chem. 2010, 285, 29787–29796. [Google Scholar] [CrossRef] [Green Version]
  7. Schulte, G.; Fredholm, B.B. Signalling from adenosine receptors to mitogen-activated protein kinases. Cell Signal. 2003, 15, 813–827. [Google Scholar] [CrossRef]
  8. Baraldi, P.G.; Tabrizi, M.A.; Gessi, S.; Borea, P.A. Adenosine receptor antagonists: Translating medicinal chemistry and pharmacology into clinical utility. Chem. Rev. 2008, 108, 238–263. [Google Scholar] [CrossRef]
  9. Ke, R.H.; Xiong, J.; Liu, Y.; Ye, Z.R. Adenosine A2a receptor induced gliosis via Akt/NF-kappaB pathway in vitro. Neurosci. Res. 2009, 65, 280–285. [Google Scholar] [CrossRef]
  10. Preti, D.; Baraldi, P.G.; Moorman, A.R.; Borea, P.A.; Varani, K. History and perspectives of A2A adenosine receptor antagonists as potential therapeutic agents. Med. Res. Rev. 2015, 35, 790–848. [Google Scholar] [CrossRef]
  11. Gessi, S.; Bencivenni, S.; Battistello, E.; Vincenzi, F.; Colotta, V.; Catarzi, D.; Varano, F.; Merighi, S.; Borea, P.A.; Varani, K. Inhibition of A2A Adenosine Receptor Signaling in Cancer Cells Proliferation by the Novel Antagonist TP455. Front. Pharmacol. 2017, 8, 888. [Google Scholar] [CrossRef] [Green Version]
  12. Jacobson, K.A.; Gao, Z.G.; Matricon, P.; Eddy, M.T.; Carlsson, J. Adenosine A2A receptor antagonists: From caffeine to selective non-xanthines. Br. J. Pharmacol. 2020, 1–16. [Google Scholar] [CrossRef]
  13. Merighi, S.; Borea, P.A.; Varani, K.; Vincenzi, F.; Jacobson, K.A.; Gessi, S. A2A Adenosine Receptor Antagonists in Neurodegenerative Diseases. Curr. Med. Chem. 2021, in press. [CrossRef]
  14. Alzheimer’s Association. Alzheimer’s Disease Facts and Figures. Alzheimer’s & Dementia. J. Alzheimer’s Assoc. 2020, 16, 391–460. [Google Scholar]
  15. Park, K.; Kim, E.; Han, H.; Shim, Y.; Kwon, J.; Ku, B.; Park, K.H.; Yi, H.A.; Kim, K.K.; Yang, D.W. Efficacy and Tolerability of Rivastigmine Patch Therapy in Patients with Mild-To-Moderate Alzheimer’s Dementia Associated with Minimal and Moderate Ischemic white Matter Hyperintensities: A Multicenter Prospective Open-Label Clinical Trial. PLoS ONE 2017, 12, e0182123. [Google Scholar] [CrossRef] [Green Version]
  16. Stoiljkovic, M.; Horvath, T.L.; Hajós, M. Therapy for Alzheimer’s disease: Missing targets and functional markers? Ageing Res. Rev. 2021, 68, 101318. [Google Scholar] [CrossRef]
  17. Johansson, M.; Stomrud, E.; Lindberg, O.; Westman, E.; Johansson, P.M.; van Westen, D.; Mattsson, N.; Hansson, O. Apathy and Anxiety Are Early Markers of Alzheimer’s Disease. Neurobiol. Aging 2020, 85, 74–82. [Google Scholar] [CrossRef]
  18. Cummings, J. New Approaches to Symptomatic Treatments for Alzheimer’s Disease. Mol. Neurodegener. 2021, 16, 2. [Google Scholar] [CrossRef]
  19. Long, J.M.; Holtzman, D.M. Alzheimer Disease: An Update on Pathobiology and Treatment Strategies. Cell 2019, 179, 312–339. [Google Scholar] [CrossRef]
  20. Mullard, A. FDA approval for Biogen’s aducanumab sparks Alzheimer disease firestorm. Nat. Rev. Drug Discov. 2021, 20, 496. [Google Scholar] [CrossRef]
  21. Fredholm, B.B.; Arslan, G.; Halldner, L.; Kull, B.; Schulte, G.; Wasserman, W. Structure and function of adenosine receptors and their genes. Naunyn Schmiedebergs Arch. Pharmacol. 2000, 362, 364–374. [Google Scholar] [CrossRef]
  22. Ferré, S.; Navarro, G.; Casadó, V.; Cortés, A.; Mallol, J.; Canela, E.I.; Lluís, C.; Franco, R. G protein-coupled receptor heteromers as new targets for drug development. Prog. Mol. Biol. Transl. Sci. 2010, 91, 41–52. [Google Scholar]
  23. Borea, P.A.; Gessi, S.; Merighi, S.; Vincenzi, F.; Varani, K. Pathological overproduction: The bad side of adenosine. Br. J. Pharmacol. 2017, 174, 1945–1960. [Google Scholar] [CrossRef] [Green Version]
  24. 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. Alzheimer’s Dis. 2021, 80, 1105–1117. [Google Scholar] [CrossRef]
  25. Lopes, L.V.; Cunha, R.A.; Kull, B.; Fredholm, B.B.; Ribeiro, J.A. Adenosine A2A receptor facilitation of hippocampal synaptic transmission is dependent on tonic A1 receptor inhibition. Neuroscience 2002, 112, 319–329. [Google Scholar] [CrossRef] [Green Version]
  26. Gonçalves, F.Q.; Lopes, J.P.; Silva, H.B.; Lemos, C.; Silva, A.C.; Gonçalves, N.; Tomé, Â.R.; Ferreira, S.G.; Canas, P.M.; Rial, D.; et al. Synaptic and memory dysfunction in a β-amyloid model of early Alzheimer’s disease depends on increased formation of ATP-derived extracellular adenosine. Neurobiol. Dis. 2019, 132, 104570. [Google Scholar] [CrossRef]
  27. Rebola, N.; Canas, P.M.; Oliveira, C.R.; Cunha, R.A. Different synaptic and subsynaptic localization of adenosine A2A receptors in the hippocampus and striatum of the rat. Neuroscience 2005, 132, 893–903. [Google Scholar] [CrossRef]
  28. Rebola, N.; Lujan, R.; Cunha, R.A.; Mulle, C. Adenosine A2A receptors are essential for long-term potentiation of NMDA-EPSCs at hippocampal mossy fiber synapses. Neuron 2008, 57, 121–134. [Google Scholar] [CrossRef] [Green Version]
  29. Tebano, M.T.; Martire, A.; Rebola, N.; Pepponi, R.; Domenici, M.R.; Gro, M.C.; Schwarzschild, M.A.; Chen, J.F.; Cunha, R.A.; Popoli, P. Adenosine A2A receptors and metabotropic glutamate 5 receptors are co-localized and functionally interact in the hippocampus: A possible key mechanism in the modulation of N-methyl-D-aspartate effects. J. Neurochem. 2005, 95, 1188–1200. [Google Scholar] [CrossRef] [Green Version]
  30. Costenla, A.R.; Diógenes, M.J.; Canas, P.M.; Rodrigues, R.J.; Nogueira, C.; Maroco, J.; Agostinho, P.M.; Ribeiro, J.A.; Cunha, R.A.; De Mendonça, A. Enhanced role of adenosine A2A receptors in the modulation of LTP in the rat hippocampus upon ageing. Eur. J. Neurosci. 2011, 34, 12–21. [Google Scholar] [CrossRef]
  31. Temido-Ferreira, M.; Coelho, J.E.; Pousinha, P.A.; Lopes, L.V. Novel Players in the Aging Synapse: Impact on Cognition. J. Caffeine Adenosine Res. 2019, 9, 104–127. [Google Scholar] [CrossRef] [PubMed]
  32. Terry, R.D.; Masliah, E.; Salmon, D.P.; Butters, N.; DeTeresa, R.; Hill, R.; Hansen, L.A.; Katzman, R. Physical basis of cognitive alterations in Alzheimer’s disease: Synapse loss is the major correlate of cognitive impairment. Ann. Neurol. 1991, 30, 572–580. [Google Scholar] [CrossRef] [PubMed]
  33. Selkoe, D.J. Alzheimer’s disease is a synaptic failure. Science 2002, 298, 789–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Scheff, S.W.; Price, D.A.; Schmitt, F.A.; DeKosky, S.T.; Mufson, E.J. Synaptic alterations in CA1 in mild Alzheimer disease and mild cognitive impairment. Neurology 2007, 68, 1501–1508. [Google Scholar] [CrossRef]
  35. Scheff, S.W.; Price, D.A.; Ansari, M.A.; Roberts, K.N.; Schmitt, F.A.; Ikonomovic, M.D.; Mufson, E.J. Synaptic change in the posterior cingulate gyrus in the progression of Alzheimer’s disease. J. Alzheimer’s Dis. 2015, 43, 1073–1090. [Google Scholar] [CrossRef] [Green Version]
  36. Arendash, G.W.; Cao, C. Caffeine and coffee as therapeutics against Alzheimer’s disease. J. Alzheimer’s Dis. 2010, 20, S117–S126. [Google Scholar] [CrossRef] [Green Version]
  37. Albasanz, J.L.; Perez, S.; Barrachina, M.; Ferrer, I.; Martín, M. Up-regulation of adenosine receptors in the frontal cortex in Alzheimer’s disease. Brain Pathol. 2008, 18, 211–219. [Google Scholar] [CrossRef]
  38. Canas, P.M.; Duarte, J.M.; Rodrigues, R.J.; Köfalvi, A.; Cunha, R.A. Modification upon aging of the density of presynaptic modulation systems in the hippocampus. Neurobiol. Aging 2009, 30, 1877–1884. [Google Scholar] [CrossRef] [Green Version]
  39. Espinosa, J.; Rocha, A.; Nunes, F.; Costa, M.S.; Schein, V.; Kazlauckas, V.; Kalinine, E.; Souza, D.O.; Cunha, R.A.; Porciúncula, L.O. Caffeine consumption prevents memory impairment, neuronal damage, and adenosine A2A receptors upregulation in the hippocampus of a rat model of sporadic dementia. J. Alzheimer’s Dis. 2013, 34, 509–518. [Google Scholar] [CrossRef]
  40. 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. Astrocytic adenosine receptor A2A and Gs-coupled signaling regulate memory. Nat. Neurosci. 2015, 18, 423–434. [Google Scholar] [CrossRef] [Green Version]
  41. Merighi, S.; Poloni, T.E.; Pelloni, L.; Pasquini, S.; Varani, K.; Vincenzi, F.; Borea, P.A.; Gessi, S. An Open Question: Is the A2A Adenosine Receptor a Novel Target for Alzheimer’s Disease Treatment? Front. Pharmacol. 2021, 12, 652455. [Google Scholar] [CrossRef]
  42. Gessi, S.; Poloni, T.E.; Negro, G.; Varani, K.; Pasquini, S.; Vincenzi, F.; Borea, P.A.; Merighi, S. A2A Adenosine Receptor as a Potential Biomarker and a Possible Therapeutic Target in Alzheimer’s Disease. Cells 2021, 10, 2344. [Google Scholar] [CrossRef]
  43. Lopes, L.V.; Cunha, R.A.; Ribeiro, J.A. Increase in the number, G protein coupling, and efficiency of facilitatory adenosine A2A receptors in the limbic cortex, but not striatum, of aged rats. J. Neurochem. 1999, 73, 1733–1738. [Google Scholar] [CrossRef]
  44. Rebola, N.; Sebastião, A.M.; de Mendonca, A.; Oliveira, C.R.; Ribeiro, J.A.; Cunha, R.A. Enhanced adenosine A2A2A receptor facilitation of synaptic transmission in the hippocampus of aged rats. J. Neurophysiol. 2003, 90, 1295–1303. [Google Scholar] [CrossRef]
  45. Viana da Silva, S.; 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]
  46. Pagnussat, N.; Almeida, A.S.; Marques, D.M.; Nunes, F.; Chenet, G.C.; Botton, P.H.S.; Mioranzza, S.; Loss, C.M.; Cunha, R.A.; Porciúncula, L.O. Adenosine A2A receptors are necessary and sufficient to trigger memory impairment in adult mice. Br. J. Pharmacol. 2015, 172, 3831–3845. [Google Scholar] [CrossRef] [Green Version]
  47. Costa, M.S.; Botton, P.H.; Mioranzza, S.; Souza, D.O.; Porciúncula, L.O. Caffeine prevents age-associated recognition memory decline and changes brain-derived neurotrophic factor and tirosine kinase receptor (TrkB) content in mice. Neuroscience 2008, 153, 1071–1078. [Google Scholar] [CrossRef]
  48. Prediger, R.D.S.; Batista, L.C.; Takahashi, R.N. Caffeine reverses age-related deficits in olfactory discrimination and social recognition memory in rats. Involvement of adenosine A1 and A2A receptors. Neurobiol. Aging 2005, 26, 957–964. [Google Scholar] [CrossRef]
  49. Dall’Igna, O.P.; Fett, P.; Gomes, M.W.; Souza, D.O.; Cunha, R.A.; Lara, D.R. Caffeine and adenosine A(2a) receptor antagonists prevent beta-amyloid (25–35)-induced cognitive deficits in mice. Exp. Neurol. 2007, 203, 241–245. [Google Scholar] [CrossRef]
  50. Laurent, C.; Eddarkaoui, S.; Derisbourg, M.; Leboucher, A.; Demeyer, D.; Carrier, S.; Schneider, M.; Hamdane, M.; Müller, C.E.; Buée, L. Beneficial effects of caffeine in a transgenic model of Alzheimer’s disease-like tau pathology. Neurobiol. Aging 2014, 35, 2079–2090. [Google Scholar] [CrossRef]
  51. Cunha, G.M.A.; Canas, P.M.; Melo, C.S.; Hockemeyer, J.; Müller, C.E.; Oliveira, C.R.; Cunha, R.A. Adenosine A2A receptor blockade prevents memory dysfunction caused by beta-amyloid peptides but not by scopolamine or MK-801. Exp. Neurol. 2008, 210, 776–781. [Google Scholar] [CrossRef]
  52. 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]
  53. Laurent, C.; Burnouf, S.; Ferry, B.; Batalha, V.L.; Coelho, J.E.; Baqi, Y.; Malik, E.; Mariciniak, E.; Parrot, S.; Van der Jeugd, A. A2A adenosine receptor deletion is protective in a mouse model of Tauopathy. Mol. Psychiatry 2016, 21, 97–107. [Google Scholar] [CrossRef] [Green Version]
  54. Temido-Ferreira, M.; Ferreira, D.G.; Batalha, V.L.; Marques-Morgado, I.; Coelho, I.E.; Pereira, P.; Gomes, R.; Pinto, A.; Carvalho, S.; Canas, P.M.; et al. Age-related shift in LTD is dependent on neuronal adenosine A 2A receptors interplay with mGluR5 and NMDA receptors. Mol. Psychiatry 2020, 25, 1876–1900. [Google Scholar] [CrossRef]
  55. Alonso-Andrés, P.; Albasanz, J.L.; Ferrer, I.; Martín, M. Purine-related metabolites and their converting enzymes are altered in frontal, parietal and temporal cortex at early stages of Alzheimer’s disease pathology. Brain Pathol. 2018, 28, 933–946. [Google Scholar] [CrossRef] [Green Version]
  56. Ji, X.-D.; Stiles, G.L.; van Galen, P.J.M.; Jacobson, K.A. Characterization of human striatal A2-adenosine receptors using radioligand binding and photoaffinity labeling. J. Receptor Res. 1992, 12, 149–169. [Google Scholar] [CrossRef]
  57. Moreira-de-Sá, A.; Lourenço, V.S.; Canas, P.M.; Cunha, R.A. Adenosine A2A Receptors as Biomarkers of Brain Diseases. Front. Neurosci. 2021, 15, 882. [Google Scholar] [CrossRef]
  58. Illes, P.; Rubini, P.; Ulrich, H.; Zhao, Y.; Tang, Y. Regulation of Microglial Functions by Purinergic Mechanisms in the Healthy and Diseased CNS. Cells 2020, 9, 1108. [Google Scholar] [CrossRef]
  59. Merighi, S.; Poloni, T.E.; Terrazzan, A.; Moretti, E.; Gessi, S.; Ferrari, D. Alzheimer and Purinergic Signaling: Just a Matter of Inflammation? Cells 2021, 10, 1267. [Google Scholar] [CrossRef]
  60. Cristovao-Ferreira, S.; Navarro, G.; Brugarolas, M.; Perez-Capote, K.; Vaz, S.H.; Fattorini, G.; Conti, F.; Lluism, C.; Ribeiro, J.A.; McCormick, P.J.; et al. A1R-A2AR heteromers coupled to Gs and Gi/0 proteins modulate GABA transport into astrocytes. Purinergic Signal. 2013, 9, 433–449. [Google Scholar] [CrossRef] [Green Version]
  61. Matos, M.; Augusto, E.; Machado, N.J.; Dos Santos-Rodrigues, A.; Cunha, R.A.; Agostinho, P. Astrocytic adenosine A2A receptors control the amyloid-beta peptide-induced decrease of glutamate uptake. J. Alzheimer’s Dis. 2012, 31, 555–567. [Google Scholar] [CrossRef] [PubMed]
  62. Paiva, I.; Carvalho, K.; Santos, P.; Cellai, L.; Pavlou, M.A.S.; Jain, G.; Gnad, T.; Pfeifer, A.; Vieau, D.; Fischer, A. A2AR-induced transcriptional deregulation in astrocytes: An in vitro study. Glia 2019, 67, 2329–2342. [Google Scholar] [CrossRef] [PubMed]
  63. Franco, R.; Rivas-Santisteban, R.; Casanovas, M.; Lillo, A.; Saura, C.A.; Navarro, G. Adenosine A2A receptor antagonists affects NMDA glutamate receptor function. Potential to address neurodegeneration in Alzheimer’s disease. Cells 2020, 9, 1075. [Google Scholar] [CrossRef] [PubMed]
  64. Franco, R.; Lillo, A.; Rivas-Santisteban, R.; Reyes-Resina, I.; Navarro, G. Microglial Adenosine Receptors: From Preconditioning to Modulating the M1/M2 Balance in Activated Cells. Cells 2021, 7, 1124. [Google Scholar] [CrossRef]
  65. Saura, J.; Angulo, E.; Ejarque, A.; Casado, V.; Tusell, J.M.; Moratalla, R.; Chen, J.-F.F.; Schwarzschild, M.A.; Lluis, C.; Franco, R.; et al. Adenosine A2A receptor stimulation potentiates nitric oxide release by activated microglia. J. Neurochem. 2005, 95, 919–929. [Google Scholar] [CrossRef]
  66. Colella, M.; Zinni, M.; Pansiot, J.; Cassanello, M.; Mairesse, J.; Ramenghi, L.; Baud, O. Modulation of Microglial Activation by Adenosine A2a Receptor in Animal Models of Perinatal Brain Injury. Front. Neurol. 2018, 9, 605. [Google Scholar] [CrossRef]
  67. Pourcher, E.; Huot, P. Adenosine A2A Receptor Antagonists for the Treatment of Motor Symptoms in Parkinson’s Disease. Mov. Disord. Clin. Pract. 2015, 2, 331–340. [Google Scholar] [CrossRef] [Green Version]
  68. Yu, F.; Zhu, C.; Xie, Q.; Wang, Y. Adenosine A2A Receptor Antagonists for Cancer Immunotherapy: Miniperspective. J. Med. Chem. 2020, 63, 12196–12212. [Google Scholar] [CrossRef]
  69. 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. 2021, 68, 113907. [Google Scholar] [CrossRef]
  70. Sun, M.J.; Liu, F.; Zhao, Y.F.; Wu, X.A. In Vivo Positron Emission Tomography Imaging of Adenosine A2A Receptors. Front. Pharmacol. 2020, 11, 599857. [Google Scholar] [CrossRef]
  71. 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]
  72. Torti, M.; Vacca, L.; Stocchi, F. Istradefylline for the treatment of Parkinson’s disease: Is it a promising strategy? Expert Opin. Pharmacother. 2018, 19, 1821–1828. [Google Scholar] [CrossRef]
  73. 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]
  74. 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 2014, 13, 767–776. [Google Scholar] [CrossRef]
  75. 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]
  76. Voronova, V.; Peskov, K.; Kosinsky, Y.; Helmlinger, G.; Chu, L.; Borodovsky, A.; Woessner, R.; Sachsenmeier, K.; Shao, W.; Kumar, R.; et al. Evaluation of Combination Strategies for the A2AR Inhibitor AZD4635 Across Tumor Microenvironment Conditions via a Systems Pharmacology Model. Front. Immunol. 2021, 12, 16. [Google Scholar] [CrossRef]
  77. Mediavilla-Varela, M.; Castro, J.; Chiappori, A.; Noyes, D.; Hernandez, D.C.; Allard, B.; Stagg, J.; Antonia, S.J. A novel antagonist of the immune checkpoint protein adenosine A2A receptor restores tumor-infiltrating lymphocyte activity in the context of the tumor microenvironment. Neoplasia 2017, 19, 530–536. [Google Scholar] [CrossRef]
  78. Willingham, S.B.; Ho, P.Y.; Hotson, A.; Hill, C.; Piccione, E.C.; Hsieh, J.; Liu, L.; Buggy, J.J.; McCaffery, I.; Miller, R.A. A2AR antagonism with CPI-444 induces antitumor responses and augments efficacy to anti-PD-L1 and anti-CTLA-4 in preclinical models. Cancer Immunol. Res. 2018, 6, 1136–1149. [Google Scholar] [CrossRef] [Green Version]
  79. Buisseret, L.; Rottey, S.; De Bono, J.S.; Migeotte, A.; Delafontaine, B.; Manickavasagar, T.; Martinoli, C.; Wald, N.; Rossetti, M.; Gangolli, E.A.; et al. Phase 1 trial of the adenosine A2A receptor antagonist inupadenant (EOS-850): Update on tolerability, and antitumor activity potentially associated with the expression of the A2A receptor within the tumor. J. Clin. Oncol. 2021, 39, 2562. [Google Scholar] [CrossRef]
  80. Franco, R.; Rivas-Santisteban, R.; Navarro, G.; Reyes-Resina, I. Adenosine Receptor Antagonists to Combat Cancer and to Boost Anti-Cancer Chemotherapy and Immunotherapy. Cells 2021, 10, 2831. [Google Scholar] [CrossRef]
  81. Chen, K.X.; Zhou, K.; Hu, B.; Xiao, M.; Chen, S. Fused Ring Derivative as A2A Receptor Inhibitor. U.S. Patent WO2019/062803A1, 30 July 2019. [Google Scholar]
  82. Carvalho, K.; Faivre, E.; Pietrowski, M.J.; Marques, X.; Gomez-Murcia, V.; Deleau, A.; Huin, V.; Hansen, J.N.; Kozlov, S.; Danis, C.; et al. Exacerbation of C1q dysregulation, synaptic loss and memory deficits in tau pathology linked to neuronal adenosine A2A receptor. Brain 2019, 142, 3636–3654. [Google Scholar] [CrossRef] [PubMed]
  83. Kondo, T.; Mizuno, Y. Japanese Istradefylline Study Group A long-term study of istradefylline safety and efficacy in patients with Parkinson disease. Clin. Neuropharmacol. 2015, 38, 41–46. [Google Scholar] [CrossRef] [PubMed]
  84. Saki, M.; Yamada, K.; Koshimura, E.; Sasaki, K.; Kanda, T. In vitro pharmacological profile of the A2A receptor antagonist istradefylline. Naunyn Schmiedebergs Arch. Pharmacol. 2013, 386, 963–972. [Google Scholar] [CrossRef] [PubMed]
  85. Mizuno, Y.; Kondo, T. Adenosine A2A receptor antagonist istradefylline reduces daily OFF time in Parkinson’s disease. Mov. Disord. 2013, 28, 1138–1141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. LeWitt, P.A.; Aradi, S.D.; Hauser, R.A.; Rascol, O. The challenge of developing adenosine A(2A) antagonists for Parkinson disease: Istradefylline, preladenant, and tozadenant. Parkinsonism Relat. Disord. 2020, 80, S54–S63. [Google Scholar] [CrossRef] [PubMed]
  87. Fusco, I.; Ugolini, F.; Lana, D.; Coppi, E.; Dettori, I.; Gaviano, L.; Nosi, D.; Cherchi, F.; Pedata, F.; Giovannini, M.G.; et al. The Selective Antagonism of Adenosine A2B Receptors Reduces the Synaptic Failure and Neuronal Death Induced by Oxygen and Glucose Deprivation in Rat CA1 Hippocampus In Vitro. Front. Pharmacol. 2018, 9, 399. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Parente, R.M.; Tarantino, P.M.; Sippy, B.C.; Burdock, G.A. Pharmacokinetic, pharmacological, and genotoxic evaluation of deuterated caffeine. Food Chem. Toxicol. 2022, 160, 112774. [Google Scholar] [CrossRef]
  89. Lane, S.; Green, C.; Steinberg, J.; Ma, L.; Schmitz, J.; Rathnayaka, N.; Bandak, S.; Ferre, S.; Moeller, F. Cardiovascular and Subjective Effects of the Novel Adenosine A(2A) Receptor Antagonist SYN115 in Cocaine Dependent Individuals. J. Addict. Res. Ther. 2012, 9 (Suppl. S1). [Google Scholar] [CrossRef] [Green Version]
  90. Kieburtz, K.; Olanow, C.W.; Krishnaswami, J.; Resburg, C.; Kerwood, F.; Glass, A.; Kenney, C. A Phase 3 Study of Tozadenant (TOZ-PD) as a Maintenance Therapy for Patients With Parkinson’s Disease Experiencing Motor Fluctuations: Characterization of Study Population (P2.045). Neurology 2018, 90 (Suppl. P2.045). [Google Scholar]
  91. Shang, P.; Baker, M.; Banks, S.; Hong, S.I.; Choi, D.S. Emerging Nondopaminergic Medications for Parkinson’s Disease: Focusing on A2A Receptor Antagonists and GLP1 Receptor Agonists. J. Mov. Disord. 2021, 14, 193–203. [Google Scholar] [CrossRef]
  92. Renk, D.R.; Skraban, M.; Bier, D.; Schulze, A.; Wabbals, E.; Wedekind, F.; Neumaier, F.; Neumaier, B.; Holschbach, M. Design, synthesis and biological evaluation of Tozadenant analogues as adenosine A2A receptor ligands. Eur. J. Med. Chem. 2021, 214, 113214. [Google Scholar] [CrossRef]
  93. Gillespie, R.J.; Bamford, S.J.; Botting, R.; Comer, M.; Denny, S.; Gaur, S.; Griffin, M.; Jordan, A.M.; Knight, A.R.; Lerpiniere, J.; et al. Antagonists of the human A2A adenosine receptor. 4. Design, synthesis, and preclinical evaluation of 7-aryltriazolo[4,5-d]- pyrimidines. J. Med. Chem. 2009, 52, 33–47. [Google Scholar] [CrossRef]
  94. Lai, T.H.; Schröder, S.; Toussaint, M.; Dukić-Stefanović, S.; Kranz, M.; Ludwig, F.A.; Fischer, S.; Steinbach, J.; Deuther-Conrad, W.; Brust, P.; et al. Development of 18F-Labeled Radiotracers for PET Imaging of the Adenosine A2A Receptor: Synthesis, Radiolabeling and Preliminary Biological Evaluation. Int. J. Mol. Sci. 2021, 22, 2285. [Google Scholar] [CrossRef]
  95. Hodgson, R.A.; Bertorelli, R.; Varty, G.B.; Lachowicz, J.E.; Forlani, A.; Fredduzzi, S.; Cohen-Williams, M.E.; Higgins, G.A.; Impagnatiello, F.; Nicolussi, E.; et al. Characterization of the potent and highly selective A2A receptor antagonists preladenant and SCH 412348 [7-[2-[4-2,4-Difluorophenyl]-1-piperazinyl]ethyl]-2-(2-furanyl)-7H-pyrazolo[4,3-e][1,2,4]- triazolo[1,5-c]pyrimidin-5-amine] in rodent models of movement disorders and depression. J. Pharmacol. Exp. Ther. 2009, 330, 294–303. [Google Scholar]
  96. Congreve, M.; de Graaf, C.; Swain, N.A.; Tate, C.G. Impact of GPCR Structures on Drug Discovery. Cell 2020, 181, 81–91. [Google Scholar] [CrossRef]
  97. Wager, T.T.; Hou, X.; Verhoest, P.R.; Villalobos, A. Moving beyond Rules: The Development of a Central Nervous System Multiparameter Optimization (CNS MPO) Approach to Enable Alignment of Druglike Properties. ACS Chem. Neurosci. 2010, 1, 435–449. [Google Scholar] [CrossRef] [Green Version]
  98. Segall, M.D. Multi-parameter optimization: Identifying high quality compounds with a balance of properties. Curr. Pharm. Des. 2012, 18, 1292–1310. [Google Scholar] [CrossRef] [Green Version]
  99. Available online: https://www.optibrium.com/downloads/MPO_Explorer_Profile_Builder.pdf (accessed on 17 February 2022).
  100. Yusof, I.; Shah, F.; Hashimoto, T.; Segall, M.D.; Greene, N. Finding the rules for successful drug optimisation. Drug Discov. Today 2014, 19, 680–687. [Google Scholar] [CrossRef]
  101. Baraldi, S.; Baraldi, P.G.; Oliva, P.; Toti, K.S.; Ciancetta, A.; Jacobson, K.A. Chapter 5. A2A adenosine receptor: Structures, modeling and medicinal chemistry. In The Adenosine Receptors, The Receptors; Varani, K., Ed.; Springer: Berlin/Heidelberg, Germany, 2018; Volume 34, pp. 91–136. [Google Scholar]
  102. Haanes, K.A.; Labastida-Ramírez, A.; Chan, K.Y.; de Vries, R.; Shook, B.; Jackson, P.; Zhang, J.; Flores, C.M.; Danser, A.H.J.; Villalón, C.M.; et al. Characterization of the trigeminovascular actions of several adenosine A2A receptor antagonists in an in vivo rat model of migraine. J. Headache Pain 2018, 19, 41. [Google Scholar] [CrossRef]
  103. Lai, T.H.; Toussaint, M.; Teodoro, R.; Dukić-Stefanović, S.; Gündel, D.; Ludwig, F.A.; Wenzel, B.; Schröder, S.; Sattler, B.; Moldovan, R.P.; et al. Improved in vivo PET imaging of the adenosine A2A receptor in the brain using [18F]FLUDA, a deuterated radiotracer with high metabolic stability. Eur. J. Nucl. Med. Mol. Imaging 2021, 48, 2727–2736. [Google Scholar] [CrossRef]
  104. Okita, K.; Kato, K.; Shigemoto, Y.; Sato, N.; Matsumoto, T.; Matsuda, H. Effects of an Adenosine A(2A) Receptor Antagonist on Striatal Dopamine D2-Type Receptor Availability: A Randomized Control Study Using Positron Emission Tomography. Front. Neurosci. 2021, 15, 1173. [Google Scholar] [CrossRef]
Figure 1. The picture represents the physiological effect of A2A receptor activation on post-synaptic neurons, microglia, and astrocytes. In more detail, its post-synaptic stimulation leads to mGluR5 activation and NMDA phosphorylation thus increasing calcium influx, while in microglia and astrocytes it raises activation and proliferation with an increase in neuroinflammation.
Figure 1. The picture represents the physiological effect of A2A receptor activation on post-synaptic neurons, microglia, and astrocytes. In more detail, its post-synaptic stimulation leads to mGluR5 activation and NMDA phosphorylation thus increasing calcium influx, while in microglia and astrocytes it raises activation and proliferation with an increase in neuroinflammation.
Molecules 27 02680 g001
Figure 2. Structures of representative A2A receptor antagonists 127, including those listed in Table 1. Some of these pharmacological probes and clinical candidate compounds have been radiolabeled for PET. Other compounds not listed in Table 1 are described in the literature [67,68,69]. Additional dedicated PET ligands are shown in Figure 3. Compound 26 is one of a series of A2A antagonists discovered by AdoRx Therapeutics [68]. * Indicates the location of a positron emitting isotope (11C or 18F) when used as a PET tracer [70].
Figure 2. Structures of representative A2A receptor antagonists 127, including those listed in Table 1. Some of these pharmacological probes and clinical candidate compounds have been radiolabeled for PET. Other compounds not listed in Table 1 are described in the literature [67,68,69]. Additional dedicated PET ligands are shown in Figure 3. Compound 26 is one of a series of A2A antagonists discovered by AdoRx Therapeutics [68]. * Indicates the location of a positron emitting isotope (11C or 18F) when used as a PET tracer [70].
Molecules 27 02680 g002
Figure 3. Structures of additional A2A receptor ligands (antagonists 2836) that have been radiolabeled for PET imaging, as described in in Sun et al. [70] * Indicates the location of a positron emitting isotope (11C or 18F). DMPX 28 is shown for comparison as an early, slightly selective A2A antagonist that is too low in affinity to be used a PET tracer. In addition, the only synthetic A2A agonist approved for human use (regadenoson 37, used in unlabeled form for myocardial perfusion imaging) is shown.
Figure 3. Structures of additional A2A receptor ligands (antagonists 2836) that have been radiolabeled for PET imaging, as described in in Sun et al. [70] * Indicates the location of a positron emitting isotope (11C or 18F). DMPX 28 is shown for comparison as an early, slightly selective A2A antagonist that is too low in affinity to be used a PET tracer. In addition, the only synthetic A2A agonist approved for human use (regadenoson 37, used in unlabeled form for myocardial perfusion imaging) is shown.
Molecules 27 02680 g003
Table 1. A2A receptor antagonists that are or have been in clinical trials for neurodegenerative diseases and/or for cancer immunotherapy. This list does not include compounds that display mixed selectivity for A2A and A2B receptors.
Table 1. A2A receptor antagonists that are or have been in clinical trials for neurodegenerative diseases and/or for cancer immunotherapy. This list does not include compounds that display mixed selectivity for A2A and A2B receptors.
Compound
(hA2AR Affinity, pKi)
Company or Sponsor
(Country)
Condition
(Reference, Dose)
Clinicaltrials.Gov
Number (Phase)
Caffeine (1), (5.0, non-selective)University Hospital, Lille (France)AD
[50] 200 mg BID
NCT04570085 (3)
Istradefylline (3, KW-6002), (7.44)Kyowa Hakko Bio (Japan)PD
[71,72] 20 or 40 mg BID
NCT00250393 (2)
NCT00955526, 6002–009 (3)
NCT01968031 (3)
Tozadenant (6, SYN-115), (8.30)Hoffmann-La Roche (The Switzerland); Biotie Therapies (Finland)PD
[73,74] 120 or 180 mg BID
NCT01283594 (2)
NCT03051607 (3)
Vipadenant (7, BIIB014), (8.89)Vernalis (UK); Biogen Idec, RedoxTherapies (Juno Therapeutics) (US)PD
[67,75] 30 or 100 mg
QD
NCT00438607 (2)
NCT00442780 (2)
Preladenant
(12, MK-3814, SCH 420814), (8.96)
Merck (US)PD, antipsychotic drug side effects, cancer
[67] 5 mg BID
NCT01155479, PARADYSE (3)
NCT00686699, P04628 (2)
NCT0309916 (1)
Imaradenant
(22, AZD4635, HTL1071), (8.77)
Sosei Heptares (Japan, UK), AstraZeneca (UK)cancer
[76] 75 mg QD
NCT03980821 (1)
NCT03381274 (1/2)
NCT02740985 (1)
NCT04089553 (2)
Taminadenant (24, NIR178, PBF-509), (7.92)Palobiofarma SL (Spain), Novartis (The Switzerland)PD, cancer
[77] 80–640 mg BID
NCT02111330 (1)
NCT02403193, AdenONCO (1/2)
NCT03207867 (3)
NCT04895748 (1)
Ciforadenant (25, CPI-444, V81444), (8.45)Bristol-Myers Squibb (US); Corvus Pharmaceuticals (US); Vernalis (UK)cancer
[78] 100 or 200 mg BID
NCT03454451 (1)
NCT03337698 (1/2)
NCT03549000 (1)
Inupadenant (27, EOS100850, EOS-850)iTeos (Belgium, US)cancer [79] 20 mg QD–160 mg BIDNCT03873883 (1)
NCT05117177 (1)
TT-10
(structure not disclosed)
Tarus Therapeutics (US)cancer [80] 10–200 mg BIDNCT04969315 (1/2)
CS3005
(structure not disclosed)
CStone (China)Cancer [81]NCT04233060 (1)
Abbreviations: AD, Alzheimer’s disease; PD, Parkinson’s disease; PD-1, programmed cell death 1.
Table 2. Calculated MPO parameters of representative A2A antagonists and one agonist 37 as shown in Figure 1 and Figure 2, including compounds that are for use as PET tracers. The probabilistic scores indicative of CNS penetration were calculated for each small molecule with the StarDrop software package (Optibrium Inc., Cambridge, UK) [98,99]. Six physicochemical parameters that affect BBB passage are included in this calculation: molecular weight, most basic center’s pKa, calculated logP (clogP), calculated logD at pH 7.4 (clogD), topological polar surface area (tPSA), and hydrogen bond donor number. The BBB log([brain]:[blood]) (clogBB) parameter is a single indicator of probability of crossing the BBB at pharmacologically significant levels.
Table 2. Calculated MPO parameters of representative A2A antagonists and one agonist 37 as shown in Figure 1 and Figure 2, including compounds that are for use as PET tracers. The probabilistic scores indicative of CNS penetration were calculated for each small molecule with the StarDrop software package (Optibrium Inc., Cambridge, UK) [98,99]. Six physicochemical parameters that affect BBB passage are included in this calculation: molecular weight, most basic center’s pKa, calculated logP (clogP), calculated logD at pH 7.4 (clogD), topological polar surface area (tPSA), and hydrogen bond donor number. The BBB log([brain]:[blood]) (clogBB) parameter is a single indicator of probability of crossing the BBB at pharmacologically significant levels.
NumberCompound [Reference]Intravenous CNS Scoring Profile ScorePredicted BBB Log([Brain]:[Blood])P-gp
Category
logP
1caffeine [12]0.355−0.0323no0.0231
2CSC [10]0.238−0.539no2.24
3istradefylline [73]0.129−0.999yes2.16
4MSX-2 [10]0.119−1.25no1.12
6tozadenant [74,92]0.0786−0.891yes2.47
7vipadenant [75]0.103−0.388yes2.12
8ZM241,385 [70]0.101−1.10yes2.13
9SCH58241 [70]0.0783−0.644yes2.60
10SCH442416 [94]0.0926−0.779yes2.78
11MRS5425 [70]0.0747−0.773yes3.08
12preladenant [95]0.105−0.592yes2.41
13MNI-444 [70]0.0964−0.495yes2.73
14ATL444 [101]0.163−1.17no1.16
15ANR94 [101]0.147−0.953no0.958
16imidazopyridine [101]0.1130.0684no3.47
17TPP455 [11]0.0753−0.876no3.14
18JNJ40255293 [10]0.121−0.219yes2.72
19JNJ41501798 [102]0.0200−0.600no5.14
20LuAA47070 [101]0.122−0.672no2.87
21imaradenant [76]0.0833−0.508no3.04
23ST-1535 [10]0.0767−1.07yes1.46
24taminadenant [77]0.153−1.01no1.02
25ciforadenenant [78]0.0507−0.965yes2.21
262-(oxazol-2-yl)pyrimidine [68]0.0723−0.502no2.64
27inupadenant [79]0.0994−1.25yes2.66
28DMPX [70]0.308−0.180no0.629
29KF18446 [70]0.227−0.947no1.34
30KF21213 [70]0.183−0.826no2.34
31KF17837 [70]0.178−1.02yes2.55
32KF19631 [70]0.183−0.525yes1.74
33BS-DMPX [70]0.192−0.603no2.21
34IS-DMPX [70]0.185−0.616no2.26
35FLUDA [103]0.0674−0.774yes3.08
36MDMPC [70]0.0934−1.43yes0.678
37regadenoson [101]0.0551−1.30yes−1.35
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Merighi, S.; Borea, P.A.; Varani, K.; Vincenzi, F.; Travagli, A.; Nigro, M.; Pasquini, S.; Suresh, R.R.; Kim, S.W.; Volkow, N.D.; et al. Pathophysiological Role and Medicinal Chemistry of A2A Adenosine Receptor Antagonists in Alzheimer’s Disease. Molecules 2022, 27, 2680. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27092680

AMA Style

Merighi S, Borea PA, Varani K, Vincenzi F, Travagli A, Nigro M, Pasquini S, Suresh RR, Kim SW, Volkow ND, et al. Pathophysiological Role and Medicinal Chemistry of A2A Adenosine Receptor Antagonists in Alzheimer’s Disease. Molecules. 2022; 27(9):2680. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27092680

Chicago/Turabian Style

Merighi, Stefania, Pier Andrea Borea, Katia Varani, Fabrizio Vincenzi, Alessia Travagli, Manuela Nigro, Silvia Pasquini, R. Rama Suresh, Sung Won Kim, Nora D. Volkow, and et al. 2022. "Pathophysiological Role and Medicinal Chemistry of A2A Adenosine Receptor Antagonists in Alzheimer’s Disease" Molecules 27, no. 9: 2680. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27092680

Article Metrics

Back to TopTop