Next Article in Journal
Cannabinoids in Medicine: A Multifaceted Exploration of Types, Therapeutic Applications, and Emerging Opportunities in Neurodegenerative Diseases and Cancer Therapy
Next Article in Special Issue
Namodenoson Inhibits the Growth of Pancreatic Carcinoma via Deregulation of the Wnt/β-catenin, NF-κB, and RAS Signaling Pathways
Previous Article in Journal
Zeb1 Regulates the Function of Lympho-Myeloid Primed Progenitors after Transplantation
Previous Article in Special Issue
Functional Interaction between Adenosine A2A and mGlu5 Receptors Mediates STEP Phosphatase Activation and Promotes STEP/mGlu5R Binding in Mouse Hippocampus and Neuroblastoma Cell Line
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pharmacology of Adenosine Receptors: Recent Advancements

1
Department of Translational Medicine, University of Ferrara, 44121 Ferrara, Italy
2
Department of Chemical, Pharmaceutical and Agricultural Sciences, University of Ferrara, 44121 Ferrara, Italy
3
University of Ferrara, 44121 Ferrara, Italy
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(9), 1387; https://doi.org/10.3390/biom13091387
Submission received: 14 August 2023 / Revised: 5 September 2023 / Accepted: 12 September 2023 / Published: 14 September 2023

Abstract

:
Adenosine receptors (ARs) are widely acknowledged pharmacological targets yet are still underutilized in clinical practice. Their ubiquitous distribution in almost all cells and tissues of the body makes them, on the one hand, excellent candidates for numerous diseases, and on the other hand, intrinsically challenging to exploit selectively and in a site-specific manner. This review endeavors to comprehensively depict the substantial advancements witnessed in recent years concerning the development of drugs that modulate ARs. Through preclinical and clinical research, it has become evident that the modulation of ARs holds promise for the treatment of numerous diseases, including central nervous system disorders, cardiovascular and metabolic conditions, inflammatory and autoimmune diseases, and cancer. The latest studies discussed herein shed light on novel mechanisms through which ARs exert control over pathophysiological states. They also introduce new ligands and innovative strategies for receptor activation, presenting compelling evidence of efficacy along with the implicated signaling pathways. Collectively, these emerging insights underscore a promising trajectory toward harnessing the therapeutic potential of these multifaceted targets.

1. Introduction

Adenosine is a biochemical molecule of paramount importance that performs a diverse range of physiological roles in the organism. It is a nucleoside composed of a purine base, adenine, linked to a ribofuranose moiety through a β-N9-glycosidic bond. This molecule is involved in a wide range of biological functions in virtually every system, organ, and tissue of the body. Endogenous adenosine triphosphate (ATP) can be extensively released during processes such as necrosis, apoptosis, inflammation, hypoxia, and mechanical injury. Extracellular adenosine primarily originates from ATP through the actions of the enzymes, nucleoside triphosphate diphosphohydrolase-1 (NTPDase1 or CD39) and ecto-5′-nucleotidase (CD73). However, other enzymatic processes can also play a role in producing extracellular adenosine. These processes encompass the conversion of ATP by nucleotide pyrophosphatase/phosphodiesterase-1 (NPP1), NTPDase2 and 3, adenylate kinase-1 (AK1), followed by the hydrolysis of the resulting adenosine monophosphate (AMP) by CD73 (Figure 1). Inside the cell, the metabolism of adenosine is predominantly regulated by enzymes such as adenosine kinase (ADK) and adenosine deaminase (ADA). In addition, adenosine can also be generated through the hydrolysis of S-adenosylhomocysteine (SAH) by SAH hydrolase (SAHH) [1].
Adenosine, ATP, and other purine and pyrimidine nucleotides signal through purinergic receptors, also known as purinoceptors. These membrane receptors are divided into P1 purinoceptors, which respond to adenosine and are thus generally referred to as adenosine receptors (ARs), and P2 purinoceptors, which respond to ATP and other nucleotides [2]. Currently, four main subtypes of ARs have been identified, namely, A1, A2A, A2B, and A3 ARs. A2A and A2B ARs are Gs-protein-coupled receptors, whereas A1 and A3 ARs are Gi-coupled receptors, although A2B and A3 ARs may also interact with Gq proteins (for a comprehensive review of their signal transduction, refer to [3]). These receptors are distributed throughout the body, including the central nervous system (CNS), cardiovascular system, peripheral organs, and immune system. In the central nervous system, adenosine plays an important role in modulating neurotransmission. Activation of ARs in the brain leads to a decrease in neuronal activity, promoting sedation and sleep [4]. Additionally, adenosine has protective effects on the brain by reducing inflammation and preventing neuronal damage. In the cardiovascular system, adenosine contributes to the regulation of blood flow and arterial pressure. Adenosine induces vasodilation of the arteries, thereby increasing blood flow and tissue oxygenation. Furthermore, adenosine also inhibits the contraction of smooth muscle cells in coronary arteries, protecting the heart from ischemia and heart attacks. Adenosine also plays a significant role in regulating the immune response. Activation of ARs on immune cells reduces the production of inflammatory cytokines, thereby decreasing inflammation and the activity of T lymphocytes [5]. This mechanism of immune regulation is crucial for maintaining a balance between an effective immune response and an excessive immune reaction.
Given its significance in various systems of the body, understanding the mechanisms of adenosine and its receptors unveils new therapeutic perspectives for a broad spectrum of disorders and medical conditions. The purpose of this review is to gather and critically analyze the latest information on the therapeutic potential of adenosine and its receptors (Figure 2). By synthesizing the current knowledge, we aim to provide a comprehensive understanding of the diverse therapeutic possibilities offered by adenosine ligands in various physiological systems and pathological conditions. Additionally, we strive to investigate whether some of the challenges that have hindered the clinical translation of compounds modulating the adenosine system have been, at least partially, resolved.

2. ARs in CNS Diseases

Adenosine, derived from ATP catabolism or directly released by neurons and glia, has neuromodulatory actions and regulates numerous physiological functions at the CNS level [6]. The A1AR is widely distributed in the CNS and has an inhibitory action on neuronal activities. Activation of A1ARs reduces neuronal excitability and can have sedative, analgesic, and anticonvulsant effects. Selective A1AR agonists or positive allosteric modulators (PAM) have been studied as potential treatments for epilepsy and other neurological conditions characterized by neuronal hyperactivity [7]. The A2AAR is abundant in the basal ganglia and plays an important role in regulating motor tone. Activation of A2AARs can have both neuroprotective and harmful effects, depending on the context [8]. Some compounds acting on A2AARs have been studied as possible pharmacological treatments for neurodegenerative diseases and neuropsychiatric disorders. The A2BARs have been less studied compared to other subtypes, but they are believed to play a role in neuroprotection and the regulation of the neuroinflammatory response. The A3ARs are mainly expressed by microglial cells, and their activation can influence the inflammatory response and neuroprotection.

2.1. Recent Advancements in AR Modulation for Pain Management

Neuropathic pain is a widespread and poorly managed problem, resulting from nerve injury and inflammation causing central sensitization and amplified pain. Adenosine and AR ligands effectively reduce neuropathic pain in preclinical models by activating A1ARs and/or A3ARs, and both A2AAR agonists and antagonists have shown efficacy in pain models [4,9]. However, finding a safe and successful way to utilize this pathway for clinical pain treatment remains a challenge. Numerous clinical trials investigating analgesic agents acting as agonists at the A1 and A2A ARs were discontinued due to either insufficient efficacy or the occurrence of side effects [10]. The potential CNS and cardiovascular side effects associated with A1 or A2A AR agonists could restrict their dose range and overall utility [11].
Recently, several strategies have been implemented to overcome these limitations. These include the development of partial agonists, PAMs, biased agonists, and innovative drug delivery systems. PAMs can augment the responsiveness of ARs to endogenous adenosine specifically within the localized regions of its elevated production [7,12,13]. MIPS521, an A1AR PAM, demonstrated in vivo analgesic effectiveness by modulating the elevated endogenous adenosine levels observed in the rat spinal cord during neuropathic pain states. The authors also reported elegant Gaussian accelerated molecular dynamic simulations that offer mechanistic insights into the positive cooperativity of MIPS521. These simulations suggest that the cooperativity is likely achieved by stabilizing the A1AR-Gi2 complex, which facilitates the formation of, and delays relaxation from, TM6 and TM7 in the ‘G protein-bound-like’ conformation [14]. To further support the role of A1ARs in neuropathic pain, in resiniferatoxin-induced neuropathy, Kan et al. found a reduced activity of the transmembrane isoform of prostatic acid phosphatase, which hydrolyzes extracellular AMP into adenosine, and downregulation of A1ARs. Low levels of adenosine associated with a low expression of A1ARs contributed to the development of mechanical allodynia [15]. Recently, the selective activation of Gαob, one of the six Gαi/o subtypes, by the A1AR agonist, benzyloxy-cyclopentyladenosine, was demonstrated to produce analgesic effects in an in vivo model of chronic neuropathic pain without inducing sedation, bradycardia, hypotension, or respiratory depression [16]. This breakthrough represents a significant advancement in creating new research tools and drugs based on the untapped potential of biased and Gα-selective agonists.
Studies using A2AARs agonists delivered to the injured nerve site demonstrated effective suppression of neuropathic pain and neuroinflammation [17]. A single peri-sciatic nerve administration of the A2AARs agonist, ATL313, effectively suppressed ongoing neuropathic pain in rats [18]. This effect appears to be facilitated, at least in part, by the sustained activation of protein kinase A (PKA), the release of the anti-inflammatory cytokine interleukin (IL)-10, the decreased release of IL-1β and nitric oxide, as well as the reduced expression of markers associated with monocytes/macrophages. However, several studies also have indicated that inhibiting A2AARs could alleviate pain in various acute and neuropathic models [19,20,21]. The inconsistencies in the reported effects of A2AARs might arise from the possible conflicting roles that A2AARs play in the periphery versus the CNS.
The A3AR has also emerged as a potential therapeutic target for neuropathic pain [22]. Therapeutic intervention with selective A3AR agonists could take advantage of the activation of an AR subtype that is less prone to developing side effects compared to A1 and A2A ARs. Durante et al. recently provided further insights into the mode of action of A3ARs in reducing neuropathic pain. They demonstrated that, while A3AR activation led to a reduction of neuropathic pain in wild-type (WT) mice, Rag-knockout (KO) mice lacking T and B cells did not respond to A3AR agonist treatment. The anti-allodynic effect of A3AR activation was reinstated in Rag-KO mice through the adoptive transfer of CD4+ T cells from WT mice. Furthermore, the activation of A3AR on CD4+ T cells resulted in the release of IL-10, which reduced dorsal root ganglion excitability and regulated neuronal hypersensitivity [23].

2.2. Recent Advancements in AR Modulation for Neurodegenerative Diseases

Neurodegenerative diseases represent a group of pathologies characterized by the progressive loss of neurons and cerebral functions, often leading to severe disabilities and cognitive decline. Among these pathologies, Alzheimer’s disease (AD) and Parkinson’s disease (PD), together with autoimmune Multiple Sclerosis (MS), are the most well-known and studied, also in connection with the involvement of ARs.
Preclinical and clinical studies have suggested that modulating ARs could have beneficial effects on neurodegenerative diseases [6,24]. The significant interest in recent years stems from the fact that adenosine is capable of influencing synaptic transmission, neuroinflammation, energy metabolism, sleep–wake cycle, and stress response, all of which are processes altered in neurodegenerative diseases.

2.2.1. Alzheimer’s Disease

Dysregulation of A2AARs appears to play a significant role in neurodegenerative processes, particularly in AD and aging-related cognitive disorders [25]. A2AARs play a role in synaptic remodeling during development and aging, with increased expression in aged individuals and animal models [26,27]. In AD patients and related rodent models, A2AAR expression has been found significantly elevated in the cortex [28], hippocampus [29,30], glial cells [31], and platelets [32]. Activation or overexpression of A2AARs has been linked to memory deficits and other aging-like phenotypes [33], while pharmacological blockade or deletion of A2AARs has been shown to mitigate synaptotoxicity and memory deficits induced by β-amyloid (Aβ) peptides in various AD and tauopathy models [34,35,36].
Recent preclinical studies have further corroborated the involvement of A2AARs in AD. Dias and colleagues suggested the role of A2AARs in regulating Ca2+ dynamics in astrocytes, where the antagonist, SCH58261, controlled ATP-evoked Ca2+ responses, an effect blunted by Aβ1-42 peptides [37]. Aβ1-42-induced synaptic and memory deficits were not encountered in CD73-KO mice, strongly linking cognitive impairment and synaptic dysfunction to ATP-derived adenosine [38].
Playing the role of an AR antagonist, caffeine has been researched regarding its potential as a protective agent against neurodegenerative disorders. Strong epidemiological and experimental evidence substantiates the notion that regular and prolonged caffeine intake can restore synaptic plasticity and alleviate cognitive deterioration in conditions of altered allostatic states, such as AD [39]. A recent and comprehensive work by Paiva et al. performed in mice revealed that chronic caffeine consumption has widespread and diverse effects on the hippocampus, impacting multiple biological processes simultaneously, including epigenomic, proteomic, and metabolomic levels, improving the signal-to-noise ratio during information encoding [40]. In the Tg4-42 mouse model of AD, long-term caffeine consumption resulted in reduced hippocampal neuron loss, improved learning and memory deficits, and enhanced neurogenesis, with no impact on extracellular Aβ levels [41].

2.2.2. Parkinson’s Disease

The approval of istradefylline (also known as KW-6002) as an add-on treatment to levodopa/carbidopa for adult PD patients experiencing “off” episodes signifies a major breakthrough in the application of drugs that interact with the adenosinergic system [42]. Although istradefylline has been approved in the USA and Japan but not in the European Union, this milestone has not only spotlighted the therapeutic potential of the A2AAR antagonists but also paved the way for an array of research endeavors and possible therapeutic applications related to this class of compounds [43].
Primarily located in the putamen, caudate, nucleus accumbens, and external globus pallidus, A2AARs interact with dopamine D2 receptors (D2Rs) within the indirect basal ganglia pathway. This explains its potential to modulate motor symptoms in PD. As recently demonstrated, extracellular adenosine, via A2AARs, increases PKA activity in striatal indirect spiny projection neurons and restricts the dopamine-induced rise of PKA activity in striatal direct spiny projection, both actions resulting in reduced locomotion [44]. Thus, adenosine and dopamine appear to form a counterbalancing system, potentially aiding fine motor control.
A recent alternative to istradefylline is KW-6356, a novel A2AAR antagonist/inverse agonist. When tested on 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-treated common marmosets, KW-6356 demonstrated a remarkable ability to effectively reverse motor disability. Notably, its anti-parkinsonian activity was found to be superior to that of istradefylline, all the while avoiding significant induction of dyskinesia [45]. In the same model, KW-6356 exhibited the ability to augment the anti-Parkinsonian effects of different doses of L-DOPA [46].
Extracellular adenosine is predominantly produced through the catabolism of ATP, facilitated by the ectonucleotidase CD73. This makes CD73 an attractive potential target for pharmacological interventions in PD [47]. By curbing the production of adenosine from CD73, neuroinflammation driven by microglia was significantly reduced, which in turn enhanced the survival of dopaminergic neurons and improved motor function in models of PD. The A2AARs mediated this process, amplifying inflammation by antagonizing the dopamine-mediated anti-inflammatory responses [48]. During the pre-symptomatic phase of a rat model of PD induced by 6-hydroxydopamine, there was an observed increase in ATP release from striatal nerve terminals. This increase led to a rise in adenosine levels through the action of CD73, and subsequently, the activation of A2AARs, influencing corticostriatal long-term potentiation [49]. These findings indicate that the activation of A2AARs plays a crucial role in the abnormal synaptic plasticity associated with the onset of motor symptoms in PD.
Numerous studies have shown that coffee consumption may lower the risk of developing PD [50,51]. It is believed that these beneficial effects are predominantly due to the action of caffeine as an antagonist of the A2AARs. Recently, Ishibashi and colleagues sought to determine the occupancy rates of striatal A2AARs by caffeine, following coffee consumption in individuals with Parkinson’s disease. The study’s conclusion highlighted that a significant A2AAR occupancy could be achieved by consuming a cup of coffee, which is approximately equivalent to 100 mg of caffeine [52].

2.2.3. Multiple Sclerosis

A1AR activation has been shown to have mostly beneficial effects on MS, by decreasing inflammatory response and promoting remyelination. However, it also increases the permeability of the blood–brain barrier, which makes this treatment approach somewhat uncertain [53]. A2AAR is associated with anti-inflammatory effects and can influence the course of MS, with early A2AAR activation reducing disease severity [54,55] but late activation worsening it [56,57]. Recently, in the experimental autoimmune encephalomyelitis (EAE) model, Zheng et al. found that A2AAR expression increased in the choroid plexus (CP), leading to enhanced CP gateway activity at day 12 post-immunization. Treatment with the A2AAR antagonist, KW6002, or focal knock-down of CP-A2AARs reduced T cell trafficking across the CP and alleviated EAE pathology. In cultured CP epithelium, A2AAR activation increased the permeability of the CP and facilitated lymphocyte migration [58].
The role of the A2BARs in MS also remains elusive. A recent work by Coppi et al. demonstrated that A2BAR activation inhibits oligodendrocyte precursor cell maturation by reducing voltage-dependent K+ currents. Silencing A2BARs in cells led to increased cell maturation, decreased sphingosine kinase 1 expression, and enhanced sphingosine-1-phosphate lyase levels [59].

2.3. Recent Advancements in AR Modulation for Brain Injury

In the context of brain injury, whether of traumatic, ischemic, or chemical origin, activation of ARs has been shown to exert both neuroprotective and neurotoxic effects, primarily through the regulation of excitotoxicity, inflammation, blood flow, and neuronal survival [60,61].

2.3.1. Traumatic Brain Injury

In traumatic brain injury (TBI), recent works further explored the role of A2A and A3ARs. TBI can lead to dysregulated fear memory, contributing to post-traumatic stress disorder and anxiety. In a craniocerebral trauma model, the A2AAR agonist, CGS21680, further enhanced fear memory, while the A2AAR antagonist, ZM241385, reduced freezing levels. Genetic knockdown of neuronal A2AAR in specific hippocampal regions (CA1, CA3, and DG) reduced fear memory after TBI, with the KO in the DG region having the most significant impact [62]. Farr et al. investigated the effect of the selective A3AR agonist, MRS5980, on the pathological outcomes and cognitive function in CD1 male mouse models of TBI. MRS5980 reduced secondary tissue injury, brain infarction, and cognitive impairment, specifically linked to reduced activation of the NFκB and the MAPK pathways, as well as the inhibition of the downstream NOD-like receptor pyrin domain-containing 3 (NLRP3) inflammasome activation. Additionally, the use of MRS5980 led to a decrease in the influx of CD4+ and CD8+ T cells caused by TBI [63].

2.3.2. Cerebral Ischemia

All subtypes of ARs have been studied for their role in cerebral ischemia. A1AR agonists and A2AAR antagonists exhibit a neuroprotective effect immediately after the insult, while A2A and possibly A2B and A3 AR agonists control inflammation and infiltration in the hours and days following brain ischemia, providing protection [64]. As well as in other therapeutic applications, the use of A1AR full agonists is hampered by their cardiovascular side effects. Using A1AR partial agonists, which are potentially less likely to cause side effects but still have similar effectiveness to full agonists, Martire et al. demonstrated a significant improvement in synaptic transmission in mice subjected to oxygen-glucose deprivation [65]. A recent elegant study aimed to explore the potential of A1ARs as imaging biomarkers and treatment targets for stroke using positron emission tomography. After transient middle cerebral artery occlusion, A1ARs were found to be overexpressed in microglia and infiltrated leukocytes in the ischemic area. Treatment with the A1AR agonist, ENBA, reduced the proliferation of microglia and macrophages. Additionally, A1AR activation led to a decrease in brain lesion size, as measured by T2W-MRI, and improved neurological outcomes, including motor, sensory, and reflex responses [66].
The neuroprotective effect of A2AAR antagonists was recently corroborated. The targeted inactivation of endothelial A2AARs resulted in a reduction of ischemic brain injury and improvement in post-stroke outcomes. These beneficial effects were achieved, at least partially, by exerting anti-inflammatory effects through the blockade of NLRP3 inflammasome activity, leading to decreased levels of cleaved caspase 1 and IL-1β expression [67]. In rat striatal slices exposed to oxygen and glucose deprivation (OGD), the A2AAR antagonist, SCH58261, notably reduced ionic imbalances and the occurrence of anoxic depolarization in medium-spiny neurons. On the other hand, the activation of A2AARs appeared to worsen the damage caused by OGD, potentially by inhibiting K+ channels [68]. Although few studies have investigated the role of A2BARs in cerebral ischemia, the findings of a recent work suggested that A2BAR activation may represent a new and interesting pharmacological approach. In a rat model of focal ischemia induced by transient middle cerebral artery occlusion, the A2BAR agonist, BAY60-6583, improved neurological deficits, significantly reduced brain damage in the cortex and striatum, reduced the activation of microglia and alterations in astrocytes, decreased the expression of TNF-α, increased the expression of IL-10, and reduced the infiltration of blood cells in the ischemic cortex [69]. Regarding A3AR, its stimulation with the selective agonist, IB-MECA, improved memory deficits caused by chronic cerebral ischemia in mice, reduced ERK phosphorylation and GFAP expression, and upregulated MAP-2 and IFN-β [70]. In a study using nonhuman primates with transient middle cerebral artery occlusion, the dual A1/A3 agonist, AST-004, exhibited a reduction in ischemic damage, thereby demonstrating the potential to simultaneously target the two neuroprotective AR subtypes [71].

2.3.3. Chemotherapy-Induced Neurotoxicity

Chemotherapeutic agents often cause neurological impairments. Two recent studies explored the potential neuroprotective role of A2AAR blockade or A3AR activation in cisplatin-induced neurotoxicities. Oliveros et al. revealed a significant increase in A2AAR and related signaling molecules in the adult mouse hippocampus following cisplatin treatment. They observed that blocking the A2AARs with KW-6002 (istradefylline) prevented the negative effects of cisplatin on neural progenitor cell proliferation and dendrite development in newly generated neurons. Additionally, the inhibition of A2AARs improved memory and reduced anxiety-like behavior in the treated mice [72]. A protective effect on cisplatin-induced neurotoxicity was found activating A3ARs with the agonist, MRS5980. The compound normalized the expression of different genes that were altered by cisplatin, thereby preventing mitochondrial dysfunction and oxidative stress. Moreover, it upregulated genes associated with repair pathways. MRS5980 also successfully reversed cisplatin-induced cognitive impairment, neuropathy, and sensorimotor deficits [73].

2.4. Recent Advancements in AR Modulation for Epilepsy

The fact that adenosine functions as an endogenous anticonvulsant and seizure terminator has been known for decades. The anticonvulsant effects of adenosine are mainly mediated by A1ARs due to this receptor’s high affinity for adenosine and its predominant expression in the seizure-prone limbic system [74]. They primarily function by blocking N-type calcium channels and inducing neuronal hyperpolarization. They also suppress neuronal hyperexcitability and maintain an inhibitory tonus in the brain. Increased levels of adenosine and A1AR expression were observed following seizures in temporal lobe epilepsy patients, which is believed to be a protective feedback mechanism to limit seizure duration and intensity [75]. A2AARs are known for their excitatory effects, enhancing NMDA receptor function and increasing glutamate release. They tend to have a proconvulsive role, though some studies suggest an anticonvulsive effect. The roles of A2B and A3 ARs in epilepsy are not clearly characterized, though it is noted that the activation of A3ARs can counteract the inhibitory effects of A1ARs [76].
As already stated, targeting A1ARs with full agonists is often associated with cardiovascular side effects. Recently, Sagu et al. proposed an intriguing alternative strategy to target A1ARs for neurological disorders. A1ARs form a complex with the neuronal protein neurabin and the regulator of G protein signaling 4 (RGS4), a protein that inhibits G protein signaling. They developed a peptide that blocks the interaction between A1ARs and neurabin, increasing the A1AR signaling and consequently reducing kainate-induced seizures and neuronal injury. A protective effect of this peptide was also demonstrated in an AD mouse model of spontaneous seizures, where it reduced epileptic spike frequency [77]. A distinct approach has been employed involving deep brain stimulation (DBS), which has proven to be an effective therapy for patients with epilepsy resistant to conventional drugs. It has been demonstrated that the A1AR antagonist, DPCPX, reversed the impact of DBS on interictal epileptic discharges in a model of status epilepticus induced by pilocarpine. Furthermore, DBS inhibited the overexpression of ADK, a crucial negative regulator of adenosine, and the downregulation of A1ARs [78].
A role for A2AARs in sudden unexpected death in epilepsy (SUDEP) has been recently suggested. In a boosted-KA model of SUDEP using genetically modified ADK knockdown mice, the A2AAR antagonist, SCH58261, increased theta and beta oscillations. Additionally, it partially restored the KA injection-induced suppression of gamma oscillation in the nucleus tractus solitarius of epileptic WT mice [79].

2.5. Recent Advancements in AR Modulation for Neuropsychiatric Disorders

In the context of neuropsychiatric diseases, the latest studies have primarily focused on the A2AARs. This is because the A2AAR plays a significant role in regulating the function of essential neurotransmitters, the dysregulation of which is commonly observed in various neuropsychiatric disorders. A2AARs form functional complexes with dopamine D2Rs and are involved in controlling glutamate and GABA release. Furthermore, A2AARs also play a role in neuroinflammation, which is increasingly recognized as a critical factor in many neuropsychiatric disorders [80,81].

2.5.1. Depression

A subpopulation of lateral septum GABAergic neurons expressing A2AARs was identified as mediating depressive symptoms through direct projections to the lateral habenula and dorsomedial hypothalamus. Additionally, A2AAR expression was found to be upregulated in the lateral septum in two male mouse models of repeated stress-induced depression, suggesting that A2AAR antagonists could have antidepressant potential [82]. A2AAR expression was also found to be upregulated in rats with sevoflurane-induced depression. Activation of A2AARs led to decreased ERK phosphorylation, reduced synaptic plasticity, and the induction of depressive-like behavior [83].

2.5.2. Anxiety

Two recent studies have investigated the effect of chronic or acute caffeine intake on anxiety. Mice exposed to chronic caffeine intake in their drinking water exhibited heightened anxiety-like behavior and improved memory function. Memory enhancement caused by caffeine was prevented when dorsal hippocampal A2AARs were disrupted, while the impact of caffeine on anxiety was blocked when ventral hippocampal A2AARs were deleted. Optogenetic activation of dorsal or ventral hippocampal A2AARs reversed the behavioral changes induced by caffeine [84]. Rats with high anxiety-like behavior, following acute caffeine intake, displayed reduced risk-taking in the multivariate concentric square field test, along with increased BDNF expression in the hippocampus and lower A2AAR mRNA expression in the caudate putamen [85].

2.5.3. Schizophrenia

The adenosine hypothesis of schizophrenia suggests that the hyperdopaminergic state typically associated with the condition could be caused by either reduced levels of adenosine in the brain or changes in the density and functional interaction of A2AARs with D2Rs [86,87]. A recent study provides evidence supporting the latter mechanism, using the phencyclidine (PCP) mouse model and A2AAR-KO mice. A2AAR-KO mice exhibited reduced prepulse inhibition, a characteristic sensory gating impairment seen in schizophrenia, and upregulation of striatal D2Rs without changes in A2AAR expression in PCP-treated animals. Furthermore, PCP-treated animals showed a significant reduction in striatal A2AAR-D2R heteromers, an effect counteracted by sub-chronic doses of antipsychotic drugs haloperidol or clozapine. Finally, in the caudate nucleus of postmortem brain samples from individuals with schizophrenia, a substantial reduction in A2AAR-D2R heteromers was observed [88]. In a subsequent study, the same authors found an increase in A2AAR-D2R heteromerization following the exposure of mammalian cells to haloperidol or aripiprazole, and a reduction with clozapine. Using computational binding models, distinctive molecular signatures for each drug were highlighted, explaining their differing effects on heteromerization [89]. In rats treated with methylphenidate to induce mania-like behavior, the A2AAR antagonist, SCH58261, reduced locomotor hyperactivity, risk-taking behavior, dopamine, and glutamate levels. It also suppressed PKC-α expression and modulated Akt/GSK-3β/β-catenin axis, indicating the potential of A2AAR as a therapeutic target for mania-like behavior treatment [90].

2.6. Recent Advancements in AR Modulation for Sleep Disorders

Adenosine plays a critical role in the homeostatic regulation of sleep and wakefulness. A1ARs promote sleep by inhibiting wake-promoting neurons and disinhibiting sleep-active neurons, while also mediating homeostatic sleep pressure through astrocytic gliotransmission. A2AARs promote sleep by inhibiting the major arousal systems in the brain, and their inhibition is the main reason for the wake-promoting effects of caffeine [91].
Recently, some noteworthy studies have provided further clarity on the role and involvement of adenosine in sleep. Using a genetically encoded GPCR-activation-based sensor for adenosine, Peng et al. discovered that, in the mouse basal forebrain, the extracellular adenosine concentration was higher during wakefulness compared to non-rapid eye movement (NREM) sleep. A significant increase in adenosine levels was observed during REM sleep, and adenosine concentrations changed rapidly during transitions between different brain states, indicating a release dependent on neural activity. The authors found a correlation between the activation of glutamatergic neurons and changes in extracellular adenosine concentration. When these neurons were ablated, there was a reduced increase in extracellular adenosine during wakefulness and REM sleep, leading to increased wakefulness and impaired sleep homeostasis [92]. In an elegant work, Jagannath and colleagues reveal a regulatory mechanism involving adenosine that enables the coordination of sleep and circadian processes to optimize sleep/wake timing in mice. In particular, adenosine influenced the circadian clock through A1/A2A AR signaling, which activates pathways that play a crucial role in regulating the clock genes Per1 and Per2 [93].
To further support the sleep-promoting effect of A1ARs, their correlation with the hypnotic effect of rosmarinic acid was investigated. In mice, rosmarinic acid decreased neuronal activity in wake-promoting brain regions and increased activity in the sleep-promoting region, resulting in reduced sleep fragmentation and decreased time to enter NREM sleep. These effects were demonstrated to be mediated by its agonistic binding to A1ARs [94].
To harness the potential of A2AARs, a PAM was explored as a means to improve insomnia while minimizing cardiovascular side effects often seen with direct agonists. The compound, named A2ARPAM-1, was found to effectively alleviate insomnia linked to mania- or schizophrenia-like behaviors in mice. Unlike diazepam, it did not lead to abnormal sleep patterns [95]. Furthermore, activating A2AARs in the olfactory tubercle promoted NREM sleep [96]. A recent study proposed a new understanding of sleep regulation involving astrocytes within the ventrolateral preoptic nucleus. According to this research, astrocytes release ATP in this brain region, which is then converted into adenosine by tissue-nonspecific alkaline phosphatase. Adenosine subsequently inhibits local GABAergic neurons, resulting in augmented excitability GABAergic projection neurons that facilitate sleep [97].

2.7. Recent Advancements in AR Modulation for Eye Diseases

In the mammalian eye, ATP and adenosine are crucial for vascular remodeling, retinal function, and neurovascular coupling. A1 and A3 ARs are generally considered to have protective effects on the retina, while A2AARs modulate neuroinflammation. Due to these roles, there have been developments of A1 and A3 AR agonists, as well as A2AAR agonists or antagonists, as potential treatments for eye diseases such as glaucoma, diabetic retinopathy, and age-related macular degeneration [98].
In a model of glaucoma, the intravitreal injection of the selective A3AR agonist, 2-Cl-IB-MECA, reversed the alterations induced by ocular hypertension, preserved retinal ganglion cell (RGC) function, improved retrograde axonal transport, and enhanced optic nerve structure [99]. In a follow-up study, a biodegradable intraocular implant with a porous structure, loaded with 2-Cl-IB-MECA, was developed. The primary goal was to avoid multiple intravitreous injections. The A3AR agonist, when released from the implant, effectively maintained its efficacy in reducing retinal cell death and promoting the survival of RGCs induced by transient ischemia [100]. Supporting the therapeutic potential of A2AAR antagonists, a recent study investigated the effects of KW6002 on retinal injury induced by the mitochondrial oxidative phosphorylation uncoupler CCCP. KW6002 treatment partially reversed CCCP-induced reduction in retinal thickness and retinal ganglia cell number by increasing mitochondrial content and reducing apoptosis of retinal ganglia cells. Additionally, KW6002 reversed the alterations in the competing endogenous RNA network caused by CCCP treatment [101]. In Thy1-YFPH transgenic mice, A2AARs were studied for their role in regulating the morphogenesis of three types of RGCs during postnatal development and neonatal inflammation. KW6002 had bidirectional effects on dendritic complexity of Type I and III RGCs and altered their morphologies. Under neonatal inflammation, KW6002 increased the proportion of Type I and II RGCs with specific changes in their morphology [102]. Table 1 shows the main recent studies investigating the effects of AR ligands in in vivo models of CNS pathology.

3. ARs in Cardiovascular and Metabolic Diseases

Adenosine plays a multifaceted role in cardiovascular function, acting as a regulator that finely adjusts various processes. It helps maintain a balanced cellular energy state and enhances the cells’ ability to withstand stress and injury. Through the interactions with all its receptor subtypes, adenosine influences every key aspect of cardiovascular function. This includes regulating the heart rate and contractility, controlling the conduction of electrical impulses within the heart, modulating autonomic control of the heart, ensuring adequate coronary blood flow, participating in cardiovascular growth and remodeling processes, and providing protection to the heart and blood vessels against harmful insults [103]. Moreover, ARs can influence glucose and lipid homeostasis. Therefore, agonists and antagonists of ARs have been explored as potential treatments for atherosclerosis, diabetes, obesity, and non-alcoholic fatty liver disease in preclinical and clinical studies [104,105].

3.1. Recent Advancements in AR Modulation for Ischemic Heart Disease

Adenosine has been extensively studied as a mediator of cardiac protection during ischemia-reperfusion [106]. Adenosine is also involved in adaptive preconditioning responses that induce prolonged shifts in stress resistance and limit later remodeling changes and heart failure progression. A recent study investigated the effects and the mechanisms associated with remote tissue compression in a mouse model of myocardial infarction. It was found that rhythmic compression on the forelimb, known as remote conditioning, served as a novel cardioprotective intervention. The study unveiled that the transmission of cardioprotective signals from the compressed limb to the heart relied on the release of adenosine, which acted on A2A and A2B ARs and modulated the cAMP/PKA/NF-κB axis [107]. In a unique porcine model of circulatory arrest and extracorporeal cardiopulmonary resuscitation, it has been shown that the use of the specific A2AAR agonist, ATL1223, significantly reduced the severity of systemic ischemia-reperfusion injury, and ameliorated renal, hepatic, and cardiac injury [108]. A different strategy to harness the cardioprotective effects of adenosine against ischemia and reperfusion injury involves prolonging the presence of its elevated concentrations in the extracellular space. The termination of extracellular adenosine signaling occurs when it is taken up into cells by equilibrative nucleoside transporters (ENTs). In a study by Ruan et al., mice exposed to myocardial ischemia and reperfusion injury were treated with the nonspecific ENT inhibitor dipyridamole, resulting in a reduction of myocardial injury. The specific deletion of A2BAR in myeloid cells and the myocyte-specific deletion of ENT1 revealed an unanticipated role for myocyte-specific ENT1 in cardioprotection, enhancing myeloid-dependent A2BAR signaling during reperfusion [109].

3.2. Recent Advancements in AR Modulation for Hypertension

ARs play a role in blood pressure regulation. While A1ARs cause vasoconstriction in the renal microcirculation, aorta, and mesenteric arteries, while also stimulating sodium reabsorption in renal tubules, A2AARs induce vasodilation in various vascular beds. The activation of the A2BARs can have complex effects on blood pressure, with both vasodilatory and vasoconstrictive actions depending on specific conditions and physiological context [110].
A1AR-induced vascular contractions in mesenteric arteries and aorta were enhanced in L-NAME hypertensive mice, where a higher receptor expression was identified. Cyp4A appeared to play a role in the altered vascular responses of A1ARs in mesenteric arteries [111].
The effects of salt diets on blood pressure in salt-sensitive hypertension with A1, A2A, or A2B AR-KOs were recently investigated. A2AAR-KOs showed higher blood pressure, while A1AR-KOs and A2BAR-KOs had lower blood pressure on the 4% salt diet. While both sexes of A2AAR-KOs were more salt-sensitive, female A1AR-KOs and A2BAR-KOs were less susceptible to salt-induced stroke and had improved survival [112].
A2AAR plays a crucial role in the function of brown adipose tissue (BAT) and pathological cardiac remodeling. An endocrine role of BAT in hypertensive cardiac remodeling through the A2AAR/fibroblast growth factor 21 (FGF21) pathway has been proposed. It was found that dysfunctional interscapular BAT caused by A2AAR-KO leads to accelerated cardiac remodeling in hypertension compared to WT mice. The FGF21, induced by the AMPK/PGC1α pathway in brown adipocytes, is necessary for A2AAR-mediated inhibition of hypertensive cardiac remodeling. Administration of recombinant FGF21 improves cardiac remodeling in hypertensive mice lacking interscapular BAT. Additionally, specific A2AAR-KO in brown adipocytes inhibits FGF21 production and accelerates cardiac damage in hypertension [113].

3.3. Recent Advancements in AR Modulation for the Regulation of Angiogenesis

The role of A2AAR activation as a potent angiogenic stimulus is widely recognized [114]. However, the role of the other ARs remains less clear.
A novel anti-angiogenic mechanism based on adenosine production acting on A2BARs has been proposed. Mesenchymal stem cells stimulated with pro-inflammatory cytokines secreted anti-angiogenic extracellular vesicles (EVs) that were enriched in CD73. These EVs inhibited endothelial cell migration and reduced vascularization in in vivo models. The anti-migratory effect of EVs is attributed to oxidative stress induced by NADPH oxidase 2 activation, triggered by adenosine produced by EVs through CD73 activity, via the activation of A2BARs [115].
In fetal intrauterine growth restriction (IUGR) placenta, where abnormal angiogenesis is significant, a recent report revealed reduced adenosine concentration and downregulated expression of A2AARs. Furthermore, the activation of A2AARs in IUGR mice placenta promoted angiogenin-dependent angiogenesis through the phosphorylation of STAT3 and Akt [116]. The administration of adenosine in the diet of piglets with IUGR resulted in an increase in average birth weight and placental efficiency and promoted angiogenesis [117].

3.4. Recent Advancements in AR Modulation for Metabolic Diseases

Fascinating new discoveries regarding the enigmatic A2BAR have been revealed in a study that examined its potential in combating obesity and aging. Mice with specific deletion of A2BARs in skeletal muscle showed signs of sarcopenia, decreased muscle strength, and reduced energy expenditure (EE), whereas deletion of A2BARs in adipose tissue worsened age-related effects and decreased EE in BAT. Pharmacological activation of A2BARs mitigated obesity induced by a high-fat diet by positively influencing whole-body EE. Additionally, A2BAR treatment led to increased muscle mass and force, enhanced the thermogenic capacity of BAT, and promoted browning of white adipose tissue (WAT). A2BAR expression was associated with increased EE in human BAT and browning of WAT, suggesting that individuals with low A2BAR levels might have been more susceptible to obesity. In human myocytes, A2BAR activation improved muscle quality, including fiber composition, oxidative metabolism, glucose uptake, and energy utilization [118]. A2AAR has also recently been associated with adipose browning. Kong and colleagues identified complement C3a receptor and C5a receptor as important regulators of adipocyte browning and energy balance. The loss of these receptors promotes the accumulation of regulatory T cells (Tregs), which in turn produce adenosine, which is then converted to inosine. Inosine activates A2AARs, promoting adipocyte browning and attenuating diet-induced obesity [119].
A1ARs are widely expressed in adipose tissue and play a significant role in glucose homeostasis. In preclinical studies, A1AR agonists reduced lipolysis, improved insulin resistance in rats fed a high-fat diet, lowered plasma triglycerides and cholesterol levels, and enhanced glucose uptake in skeletal muscles. The A1AR antagonist, BW-1433, improved glucose tolerance and increased lipolysis [104]. Recent research revealed that insulin resistance induced by a high-sucrose diet is linked to higher levels of A1, A2A, and A2B ARs in the skeletal muscle, an increase in A1ARs in the liver and adipose tissue, and a decrease in A2BARs in the liver. When A1ARs were blocked in adipose tissue, insulin signaling improved in control animals, but it had a negative effect on insulin signaling in animals on the high-sucrose diet. A2A or A2B ARs antagonists were found to reverse the impaired insulin signaling in the skeletal muscle of rats on the high-sucrose diet. However, they did not have any significant impact on insulin signaling in the liver or adipose tissue [120]. Table 2 summarizes the most recent in vivo studies using AR ligands concerning cardiovascular and metabolic diseases.

4. ARs in Inflammation and Autoimmunity

Adenosine exerts significant control over the inflammatory process. The immunoregulatory effects of adenosine and its receptors, primarily anti-inflammatory in nature, contribute to an overall tissue-protective action [5]. All immune cells of the innate system express the four subtypes of ARs. When A2A, A2B, and A3 ARs are activated in macrophages, they limit the production of various pro-inflammatory mediators while promoting the release of anti-inflammatory ones. Adenosine also plays a regulatory role in dendritic cells (DCs), with A2AARs reducing pro-inflammatory cytokines. In mast cells, A2BARs trigger degranulation, while A3ARs display anti-inflammatory properties. Neutrophils express all four ARs: activation of A1 and A3 ARs promotes chemotaxis and phagocytosis, while A2A and A2B ARs inhibit neutrophil trafficking and effector functions. Adenosine, produced by regulatory Tregs, reduces NF-κB activation in T effector cells through stimulation of A2AARs. Moreover, adenosine modulates B cell functions, with all four receptor subtypes being expressed [121]. Platelets express only A2A and A2B ARs, and when A2AAR are activated, they inhibit the secretion of pro-inflammatory mediators, reduce cell activation, and decrease P-selectin expression [122]. New research highlights and corroborates the importance of AR modulation in the regulation of inflammation and in autoimmunity diseases.

4.1. Recent Advancements in AR Modulation for Autoimmunity Diseases

The adenosine-based targeting of certain widely used drugs for treating rheumatic diseases, particularly methotrexate, is well recognized [123]. Human synoviocytes exhibit significant expression of both A2A and A3 ARs, and their activation induces an anti-inflammatory response [124]. Furthermore, A2A and A3 ARs have been found overexpressed in immune cells from rheumatoid arthritis [125,126], systemic lupus erythematosus [127], and ankylosing spondylitis [128].
A2A or A2B AR stimulation has shown a beneficial effect in various preclinical models of arthritis [126,129,130,131]. Recently, it has been shown that the A2AAR agonist, CGS21680, inhibited arthritis development and redirected the differentiation of autoreactive CD4+ T cells away from the germinal center T follicular helper lineage. In addition, CGS21680 treatment prevented the emergence of high-affinity glucose-6-phosphate isomerase-specific and IgG1 isotype class-switched polyclonal plasmablasts, resulting in decreased levels of anti-GPI IgG1 antibodies [132]. A distinct response to A2AAR stimulation with CGS21680 has been observed in macrophages: it reduced matrix metalloproteinase (MMP) 8 expression in healthy macrophages, but it was unable to decrease MMP8 expression in macrophages from patients with ankylosing spondylitis [133].
An elevated expression of A2AAR has been identified in CD11c+T-bet+ B cells [134], a specific type of B cell that plays a critical role in autoimmunity, particularly in the development of systemic lupus erythematosus. In lupus-prone mice, CGS21680 treatment depleted CD11c+T-bet+ B cells, CD138+ B cells, and pathogenic lymphocytes and reduced anti-nuclear antibodies. Additionally, it decreased kidney pathology and lymphadenopathy and improved the overall condition of the animals, even after the disease onset [135].
In an effort to minimize potential adverse reactions resulting from the widespread distribution of ARs, recent studies focused on creating and assessing a skin-targeted method for delivering an A3AR agonist. This agonist is activated by blue light, facilitated by a photo-cleavable masking group. The A3AR agonist, known as MRS7344, effectively hindered the development of psoriatic-like characteristics in an IL-23 animal model. This successful outcome illustrates the practicality of utilizing light-directed approaches for treating psoriasis [136]. In a Phase II clinical study conducted on patients with psoriasis, the A3AR agonist, piclidenoson (also known as CF101, IB-MECA), was found to be safe and showed effectiveness, leading to significant improvements in skin lesions [137].
Deficiency of adenosine deaminase 2 (DADA2) is caused by ADA2 gene mutations, resulting in an autoinflammatory systemic vasculitis, where neutrophil extracellular traps (NETs) significantly contribute to the disease. Adenosine, through the engagement of A1 and A3 ARs, was found to trigger NET formation, particularly in neutrophils from female DADA2 patients. In contrast, A2AARs activation had an opposite effect, reducing NET formation, as well as inhibiting cytokine release mediated by NF-κB activation in macrophages derived from DADA2 patients [138].

4.2. Recent Advancements in AR Modulation for Osteoarthritis

Evidence suggests that adenosine is crucial for maintaining cartilage structure and function. Its primary role in cartilage homeostasis is acting as a buffer against the inflammatory environment on chondrocytes. Numerous studies have focused on the A2AAR, identifying it as the main mediator of adenosine’s protective effects [139]. A2AAR or ecto-5′nucleotidase KO mice develop spontaneous OA, and deleting A2AAR from chondrocytes has determined an osteoarthritis (OA) phenotype with increased MMP13 and Col10a1 expression. Additionally, injecting adenosine-containing liposomal suspensions intra-articularly prevents OA development in rats [140]. In a subsequent and recent study, A2AAR stimulation reduced senescence markers in chondrocytes in vitro and in obesity-induced OA mice. A2AAR agonism enhanced the Sirt1/AMPK pathway and increased the anti-senescent p53 variant, Δ133p53α [141]. Furthermore, primary murine chondrocytes from A2AAR−/− null mice, which develop spontaneous OA, have mitochondrial dysfunctions. Treatment with the A2AAR agonist, CGS21680, improved mitochondrial stability and function in IL-1β-exposed chondrocytes and in an obesity-induced OA mouse model [142].
Forkhead box O (FoxO) transcription factors, stress-responsive mediators, are crucial for maintaining articular cartilage homeostasis. Evidence from mouse FoxO KOs shows that their absence leads to early OA and reduced cartilage autophagy. A2AAR stimulation with the agonist, CGS21680, activated FoxO1 and FoxO3, promoting increased autophagy and improved metabolic function in chondrocytes. Enhanced activation of FoxO1 and FoxO3, along with increased autophagic flux, was demonstrated in vivo after administering the liposome-associated A2AAR agonist in an obesity-induced OA mouse model [143].
A potential role for A3AR in OA was recently suggested. Orally administered A3AR agonist, CF101, in OA rats induced by anterior cruciate ligament transection surgery reduced OA cartilage damage, pain, and cartilage pyroptosis. Mechanistically, CF101 inhibited ROS production, NLRP3 inflammasome activation, and gasdermin D cleavage in rat primary chondrocytes, indicating the inhibition of pyroptosis [144].

4.3. Recent Advancements in AR Modulation for Respiratory Diseases

The role of adenosine signaling is pivotal in the lung’s response to injuries. Initially, adenosine plays a beneficial anti-inflammatory and tissue-protective role, mainly by activating the A2A and A2B ARs during acute lung injury [145]. However, in chronic respiratory diseases, it triggers the activation of A1, A2BAR, and A3 ARs, leading to a pro-inflammatory state and uncontrolled tissue remodeling [146]. One of the extensively researched A2BAR antagonists in asthma and chronic obstructive pulmonary disease (COPD) is CVT-3883. It has been demonstrated to be as efficacious as montelukast, as it effectively lowered the count of inflammatory cells in bronchoalveolar lavage fluid and inhibited the production of proinflammatory mediators originating from macrophages [147]. Novel and potent A2BAR antagonists continued to be developed, demonstrating efficacy in an in vivo model of allergic asthma [148]. Nevertheless, the results are still controversial, and more research is required.
In a recent study, the oral A1AR antagonist, PBF-680, abrogated the late asthmatic response and reduced the early allergic response, fractional exhaled nitric oxide, and blood eosinophils in mild-to-moderate atopic asthmatics [149]. In asthmatic patients, it also reduced AMP airway hyperresponsiveness [150].
Xiao et al. identified adenosine as a key regulator that suppresses the responses of group 2 Innate Lymphoid Cells (ILC2s) and alleviates allergic airway inflammation. After exposure to the protease papain, levels of adenosine in the lungs were found to be elevated, and the A2AARs were abundantly expressed in lung ILC2s. The AR agonist, NECA, significantly subdued ILC2 responses and mitigated inflammation caused by IL-33 or papain. However, in A2AAR-KO mice or when adenosine synthesis was blocked, the inflammation worsened [151]. The DNA-derived drug, Polydeoxyribonucleotide (PDRN), which acts as an A2AAR agonist, has recently been the subject of much research. In a rat acute lung injury model induced by lipopolysaccharide (LPS), PDRN effectively reduced lung tissue damage, pro-inflammatory cytokines, apoptotic factors, and MAPK/NF-κB activation [152].
The anti-inflammatory effect of A2AAR activation has been recently hypothesized as a strategy to reduce lung inflammation in coronavirus disease 2019 (COVID-19) patients [153,154]. Patients with mild and severe cases of COVID-19 exhibited reduced extracellular adenosine levels, which were linked to elevated concentrations of pro-inflammatory cytokines and were attributed to altered expression of CD39 and CD73 in the T cells of COVID-19 patients [155]. It has been also proposed that the exacerbation of inflammation by oxygenation may result from the absence of hypoxia-induced A2AAR activation. Therefore, a direct approach might involve pairing oxygen ventilation for COVID-19 patients with the administration of inhaled adenosine or A2AAR agonists [156]. In a study involving 14 COVID-19 patients who received inhaled adenosine, 13 exhibited positive outcomes and a reduction in respiratory symptoms [157].
A link between sevoflurane and A2BARs for the reduction of acute pulmonary inflammation has been described. Sevoflurane reduced LPS-induced polymorphonuclear neutrophil (PMN) infiltration and edema in WT mice. Chimeric mice expressing A2BARs exclusively on leukocytes showed decreased PMN counts after sevoflurane treatment [158]. Recent research demonstrated that A2BAR activation by adenosine is the mechanism by which cupping, a traditional Chinese alternative therapy, attenuates LPS-induced lung inflammation [159].
A selective A3AR and partial PPARγ agonist, LJ-529, significantly improved pulmonary emphysema induce by elastase, restoring pulmonary function, reducing airspace enlargement, MMP activity, and apoptosis [160]. Activation of A3ARs was recently proven beneficial in a bleomycin murine model of lung fibrosis. The A3AR selective agonist, MRS5980, attenuated bleomycin-induced lung stiffness, TGF-β levels, α-SMA deposition, and inflammatory and oxidative stress markers [161].
In a model of bronchopulmonary dysplasia induced by hyperoxia exposure, caffeine treatment reduced lung injury and enhanced alveolar development by decreasing oxidative stress and inflammatory infiltration. Mechanistically, caffeine reduced NLRP3 activity, NF-κB pathway activation, and also downregulated the expression of A2AAR protein in the lungs of mice [162].

4.4. Recent Advancements in AR Modulation for Sepsis

The anti-inflammatory and immunosuppressive effect of adenosine is evidently counterproductive in sepsis. During systemic inflammation or tissue damage, extracellular adenosine levels increase significantly. Septic shock patients exhibit a tenfold rise in plasma adenosine concentrations due to reduced ADA and ADK activity and increased CD73 activity. The immunosuppressive effects of adenosine are mainly mediated by A2AAR, but also by A2BARs, whose expression is rapidly increased by endotoxin or inflammatory mediators. On the other hand, stimulating A1 or A3 ARs during sepsis may have advantageous effects, leading to a reduction in mortality and mitigating renal and hepatic damage [163].
A recent significant publication delineated the function of adenosine and CD39hi plasmablasts as crucial factors in the induction of immunosuppression caused by sepsis. The study revealed that sepsis led to an enlargement of a B cell group expressing CD39, resulting in an increase in immunosuppressive adenosine. This adenosine, in turn, interacted with A2AARs, weakening the bactericidal action of macrophages and boosting the production of IL-10 [164]. In the same model of sepsis induced by cecal ligation and puncture, the A2AAR antagonist, ZM241385, enhanced survival by improving bacterial clearance. However, when Treg-deleted mice were treated with ZM241385, there was no improvement in sepsis survival, indicating that the effect relies on Treg activity. Inactivating A2AAR led to decreased frequencies and impaired function of Foxp3+ Tregs.
Novel insights into the role of A2B antagonism in sepsis were also gained. Inhibition of chemokine receptors CXCR4 and CXCR7 decreased platelet–neutrophil complex formation in WT mice, but such protective anti-inflammatory effects were not observed in A2BAR−/− animals [165]. In mouse models of zymosan- and fecal-induced peritonitis, sevoflurane showed protective effects in WT animals but not in mice lacking A2BARs. The presence of A2BAR expression on both hematopoietic and nonhematopoietic compartments was necessary for sevoflurane’s protective effects [166]. Table 3 shows the most recent studies evaluating the effects of AR ligands in in vivo models of inflammatory and autoimmune diseases.

5. ARs in Cancer

Immunotherapy has revolutionized cancer treatment by harnessing the immune system’s ability to target and eliminate cancer cells in a specific manner. The tumor microenvironment (TME) can experience transient or chronic intratumoral hypoxia, leading to metabolic changes and the CD39/CD73-mediated accumulation of adenosine derived from ATP. This accumulation promotes A2A/A2B AR-dependent suppression of the host’s immune defense mechanisms, including the recruitment and differentiation of Treg cells and the inhibition of effector immune cells such as T cells, NK cells, macrophages, and DCs [167]. Because of this, adenosine and its receptors represent one of the main mechanisms through which tumor cells evade immune surveillance [168]. Furthermore, adenosine does not only affect immune cell responses to cancer cells but also influences tumor angiogenesis, lymphangiogenesis, cancer-associated thrombosis, and tumor perfusion. With the intention of exploiting these mechanisms, in recent years, there has been a surge in preclinical and clinical studies delving into the potential of the adenosinergic system for cancer immunotherapy, undoubtedly making it one of the most rapidly advancing frontiers in adenosine pharmacology [169]. Many clinical trials have explored the use of monoclonal antibodies or small molecule inhibitors that target the CD39/CD73/A2AAR pathway either as standalone treatments or in combination with anti-PD-1/PD-L1 therapies [170]. Preclinical data, however, indicate a potential role for other receptor subtypes as well.

5.1. Recent Advancements in A2AAR Modulation for Cancer

Recent experimental evidence strongly corroborates the role of A2AAR inhibition in restoring T cell effector function. Using a single-cell reporter strategy, it was found that A2AAR antagonism enhances cytotoxic T cell contact stability, improves lytic granule polarization and exocytosis, and increases the delivery of sublethal perforin hits per cytotoxic T cell contact. Moreover, A2AAR antagonism restored the functionality of tumor-infiltrating cytotoxic T cells in a melanoma model, leading to a local effector phenotype characterized by prolonged dwell time and improved sublethal hit delivery [171]. Research into the signaling pathway through which adenosine undermines the immune competence of peripheral T cells and lymphocytes infiltrating tumors has revealed that A2AAR activation, through PKA activation, disrupts the TCR/mTORC1 signaling in human CD8+ T cells. This impairment subsequently hampers both metabolic efficiency and effector functionalities [172]. In a model of chronic lymphocytic leukemia, the use of the A2AAR antagonist, SCH58261, demonstrated the restoration of immune competence. This was achieved by inhibiting the accumulation and differentiation of Treg cells, reinstating effective T cell functions, and shifting monocytes toward an inflammatory (M1-like) phenotype [173]. Monocytes/macrophages constitute a vital component within tumor tissues, playing a pivotal role in tumor progression and therapeutic response. Within human hepatocellular carcinoma (HCC) tissue, macrophages exhibited heightened proliferative potential induced by adenosine derived from the tumor itself. These rapidly dividing macrophages show reduced differentiation, display immunosuppressive characteristics, and their presence is inversely linked to the prognosis of HCC patients. The study revealed that autocrine granulocyte-macrophage colony-stimulating factor upregulated A2AAR expression in macrophages, working in synergy with adenosine to stimulate their proliferation [174]. Research has demonstrated that adenosine hampers the maturation process and hinders the antigen presentation function of CD103+ DCs, a crucial stage in promoting anti-tumor immune responses. The suppressive characteristics triggered by adenosine in human DCs, which lead to immune tolerance, were effectively counteracted by using the A2AAR antagonist, AZD4635 [175]. Using a syngeneic B cell lymphoma model, conditional deletion of A2AARs in myeloid cells augmented the therapeutic efficacy of anti-CD20 mAb [176].
The detrimental role of A2AAR activation in cancer does not end with its effects on immune cells. The involvement of A2AARs in tumor-associated lymphangiogenesis was uncovered: A2AAR-deficient mice exhibited diminished lymphangiogenesis in tumors and sentinel nodes, resulting in protection against metastasis. Lack of A2AARs in both hematopoietic and nonhematopoietic cells contributes to this outcome [177].
Adenosine signaling has been reported to be linked with an unfavorable prognosis and might hold predictive value for the response to anticancer therapy. An analysis of the gene expression signature for adenosine signaling, primarily associated with A2AAR activation, establishes an adverse correlation between adenosine and overall survival, progression-free survival, as well as decreased efficacy of anti-PD1 therapy, among a cohort of patients treated with immune checkpoint inhibitors [178]. In patients with renal cell cancer (RCC), increased expression of A2AARs in the primary tumors was associated with the presence of metastatic characteristics. Moreover, patients exhibiting lower A2AAR expression demonstrated improved response to therapy and prolonged overall survival [179].
Well-known and novel A2AAR antagonists have recently been tested in both preclinical settings and clinical trials. The novel A2AAR antagonist, DZD2269, demonstrated anti-tumor effects in syngeneic mouse models, particularly when used in conjunction with immune checkpoint inhibitors, radiotherapy, or chemotherapy. Preliminary results from a phase I clinical trial revealed that DZD2269 effectively suppresses pCREB within human T cells [180]. Similarly, other innovative A2AAR antagonists, AZD4635 and CPI-444 (Ciforadenant), exhibited the ability to diminish tumor growth while intensifying the effectiveness of checkpoint inhibitors in syngeneic tumor models [175,181]. Notably, mice that were rechallenged exhibited complete inhibition of tumor growth, highlighting the induction of systemic immune memory. Ciforadenant was subsequently tested in phase I clinical trial in a cohort of patients with advanced refractory RCC. The A2AAR antagonist demonstrated efficacy in immunotherapy-naïve patients and those resistant to anti-PD-L1 treatment, especially when combined with atezolizumab. Furthermore, an adenosine-related gene signature suggests the potential to identify patients who are likely to respond positively to adenosine pathway blockade-based treatments [182]. In castration-resistant prostate cancer (CRPC), colorectal carcinoma, non-small cell lung cancer (NSCLC), or other solid tumors, the safety and antitumor activity of AZD4635 were evaluated in a phase 1a/b open-label, multicenter study. AZD4635, whether used alone or together with durvalumab, exhibited favorable tolerability and showed no significant safety issues. Moreover, it demonstrated initial indications of clinical effectiveness in patients with metastatic CRPC [183]. Likewise, the A2AAR antagonist, Taminadenant, both with and without spartalizumab, was well tolerated in individuals with advanced NSCLC, where certain patients showed signs of clinical improvement.
Novel A2AAR antagonists for cancer immunotherapy continue to be synthesized and developed [184,185], including dual-acting compounds targeting both A2AARs and other well-recognized cancer targets such as histone deacetylases [186,187] or CD73 [188]. In an attempt to mitigate the adverse effects of systemic A2AAR inhibition and facilitate tumor-specific delivery and activation of A2AAR antagonists, various innovative approaches utilizing nanotechnologies have been investigated. A photo-modulated nanoreactor was recently shown to induce oxygen-mediated reduction in A2AAR RNA expression within immune cells. By activating the nanoreactor exclusively within the TME using near-infrared radiation, this A2AAR inhibition led to substantial secretion of immune-related cytokines, enhancing anti-tumor immune responses and promoting tumor cell killing [189]. In a similar approach, on-site oxygen generation through hydrogen peroxide catalysis was harnessed to inhibit A2AAR responses. This method utilized macrophage membrane-coated mesoporous silica nanoparticles loaded with catalase, doxorubicin, and resiquimod. Oxygen-dependent A2AAR inhibition counteracted the immunosuppressive TME caused by hypoxia. This, in turn, prompted DCs to boost the immune response mediated by CD8+ T cells [190]. Furthermore, a polydopamine nanocarrier, concealed by an acid-sensitive PEG shell, was employed for delivering the A2AAR antagonist, SCH58261, to the tumor. Once it reaches the acidic TME, SCH58261 is selectively released within the tumor tissue, thereby enhancing the immune response against the tumor [191].
A prominent avenue in oncology’s advancing immunotherapy landscape involves chimeric antigen receptor (CAR) T cells [192]. Due to its immunosuppressive role, targeting A2AARs has been confirmed as a successful approach for enhancing the effectiveness of CAR T cell therapy. The A2AAR antagonist reversed the decrease in CAR T cell proliferation and cytokine response triggered by agonists. However, it was not effective in restoring the cells’ cytotoxic functionality. The genetic removal of A2AARs, achieved through either shRNA or CRISPR/Cas9, improves the in vivo effectiveness of CAR T cells, resulting in enhanced tumor eradication capabilities [193,194,195]. Recently, the dual A2A/A2B AR antagonist, AB928 (etrumadenant), boosted CAR T cell cytokine production and proliferation, enabling the effective destruction of tumor cells in vitro and enhancing CAR T cell activation in vivo [196].

5.2. Recent Advancements in A2BAR Modulation for Cancer

A2BARs are expressed in both immune and non-immune cells, and their activation has been associated with promoting cancer cell proliferation, tumor growth, the formation of metastases, tumor angiogenesis, and the suppression of immune system responses [197]. Recent studies further indicate that A2BARs represent a potential target for cancer therapeutic intervention.
While the contribution of T cell A2BARs to immunosuppression and tumor promotion was discovered to be minimal, the presence of A2BARs on myeloid cells and antigen-presenting cells indirectly hindered CD8+ T cell responses and facilitated metastasis. The inhibition of A2BARs through genetic or pharmacological means enhanced the effectiveness of adoptive T cell therapy [198]. Hypoxia is a prevalent characteristic of the TME, and there is substantial evidence indicating an increase in A2BAR expression upon the activation of HIF-1α. A recent study demonstrated the role of A2BARs in hypoxia-induced breast cancer stem cell enrichment by activating PKC-δ/STAT3 pathway [199]. In addition to hypoxia, chemotherapy was also found to increase A2BAR expression, and its activation contributed to the expression of chemotherapy-induced pluripotency factors and the enrichment of breast cancer stem cells. As a result, in in vivo models of triple-negative breast cancer, genetic or pharmacological inhibition of A2BARs resulted in a delay of tumor recurrence after discontinuation of chemotherapy [200]. Recently, highly effective A2BAR antagonists were synthesized and subjected to functional assessment in patient-derived tumor spheroid models. These novel compounds demonstrated the capability to restore T and NK cell proliferation, enhance the production of IFNγ and perforin, and promote increased infiltration of tumor-infiltrating lymphocytes [201].
Yu et al. demonstrated that A2BAR activation enhanced CD73 expression in cancer-associated fibroblasts, initiating a feedforward circuit to amplify the CD73-adenosine axis in the TME, which ultimately led to A2AAR-dependent inhibition of immune activation [202]. Elevated levels of CD73 were detected in activated CD8+ T cells in the pancreatic ductal adenocarcinoma TME. A recent study revealed that the A2BARs on CD8+ T cells played a pivotal role in adenosine-mediated immunosuppression in pancreatic ductal adenocarcinoma models [203]. The CD73-dependent generation of adenosine through tumor-derived exosomes was demonstrated to drive the polarization of macrophages into an M2-like phenotype via A2BARs, thereby facilitating the release of angiogenic factors. Consequently, targeting A2BARs could potentially serve as a strategy to counteract tumor-derived exosome-induced tumor angiogenesis [204].

5.3. Recent Advancements in A1 and A3 AR Modulation for Cancer

While A2A and A2B ARs have long been regarded as the pivotal ARs responsible for orchestrating the immune response against tumors, emerging evidence is shedding light on the significant involvement of A1ARs as well.
In various immune-deficient cancer models, the deletion of A1ARs was shown to hinder tumor growth [205,206]. Nonetheless, recent findings have unveiled that the inhibition of A1ARs in immune-competent mice actually facilitated tumor immune evasion. The underlying mechanism for this phenomenon was identified as A1AR-mediated up-regulation of PD-L1 expression through the activation of ATF3. Notably, elevated A1AR expression was detected in tumor tissues of non-responder NSCLC patients when compared to those who responded to anti-PD-1 monoclonal antibody therapy [207]. A1AR was also identified as one of the most up-regulated genes in EGFR-mutant NSCLC associated with an immune-inert phenotype [208].
Different results were observed in HCC. Hypoxia-mediate adenosine generation mediated activated A1ARs, consequently fostering the accumulation of immunosuppressive plasmacytoid DCs. Employing an A1AR antagonist effectively impeded the migration of plasmacytoid DCs, leading to the suppression of tumor growth [209].
The overexpression of A3ARs is observed across nearly all types of cancer, suggesting their potential utility as a promising tumor biomarker. Furthermore, the activation of A3ARs could potentially introduce a novel approach to personalized cancer therapy [210]. Recently, elevated expression of A3ARs has been identified in tumor-infiltrating NK cells across various tumor types, significantly exceeding the expression levels found in peripheral NK cells [211]. The A3AR agonist, namodenoson (also known as CF102 or Cl-IB-MECA), exhibited encouraging outcomes in a phase I/II clinical trial involving advanced HCC and moderate hepatic dysfunction. The compound demonstrated favorable tolerance and displayed a modest enhancement in overall survival/progression-free survival, although it was not statistically significant [212]. Table 4 reports the latest in vivo studies on adenosine receptor modulation in cancer models.

6. Conclusions

Adenosine, its four receptor subtypes, and its metabolizing enzymes continue to be subjects of intensive research aimed at addressing numerous diseases. Over the past five years, significant progress has been achieved in understanding the mechanisms underlying adenosine-mediated modulation of crucial physiological functions and their roles in disease conditions. Given adenosine’s widespread involvement in the body, substantial advancements have been made in various fields where the modulation of ARs offers potential therapeutic avenues. These fields encompass CNS disorders, cardiovascular and metabolic conditions, inflammatory-based diseases, and cancer. However, the omnipresent nature of adenosine has posed challenges in the translation of AR-interacting drugs into clinical practice. As these challenges are gradually surmounted, the modulation of ARs could emerge as a transformative factor in treating diverse diseases. This notion is reinforced by the dedicated efforts to develop novel drug candidates centered around the manipulation of the adenosine system, as substantiated by a multitude of ongoing clinical trials (Table 5). As highlighted by this review, novel avenues have emerged for harnessing the full potential of ARs. Utilizing partial agonists, allosteric modulators, and biased agonists stands out as a promising approach to finely tune the modulation on these receptors. Moreover, the development of innovative drug delivery systems, coupled with the prospect of in situ activation, offers spatial and temporal control possibilities, thus contributing to the mitigation of adverse effects. Recent studies have also unveiled new insights into the intricate mechanisms regulated by ARs, paving the way for new therapeutic approaches across various diseases. Among these, the field of cancer treatment has witnessed remarkable activity, as evidenced by the elevated number of clinical trials in this area. Nevertheless, promising outcomes have surfaced in numerous other pathologies as well, bringing hope for the future application of therapies based on adenosinergic system modulation.

Author Contributions

F.V., P.A.B. and K.V. conceived the work and wrote the manuscript. S.P., C.C., M.C., M.N., A.T., S.M. and S.G. contributed to writing and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Ministry of University and Research (MUR) National Recovery and Resilience Plan (NRRP), Project no. PE00000006 CUP H93C22000660006—MNESYS, A multiscale integrated approach to the study of the nervous system in health and disease (DN. 1553, 11 October 2022).

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. Linden, J.; Koch-Nolte, F.; Dahl, G. Purine Release, Metabolism, and Signaling in the Inflammatory Response. Annu. Rev. Immunol. 2019, 37, 325–347. [Google Scholar] [CrossRef]
  2. Burnstock, G. Purine and Purinergic Receptors. Brain Neurosci. Adv. 2018, 2, 2398212818817494. [Google Scholar] [CrossRef]
  3. 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]
  4. Vincenzi, F.; Pasquini, S.; Borea, P.A.; Varani, K. Targeting Adenosine Receptors: A Potential Pharmacological Avenue for Acute and Chronic Pain. Int. J. Mol. Sci. 2020, 21, 8710. [Google Scholar] [CrossRef]
  5. Pasquini, S.; Contri, C.; Borea, P.A.; Vincenzi, F.; Varani, K. Adenosine and Inflammation: Here, There and Everywhere. Int. J. Mol. Sci. 2021, 22, 7685. [Google Scholar] [CrossRef]
  6. Chang, C.-P.; Wu, K.-C.; Lin, C.-Y.; Chern, Y. Emerging Roles of Dysregulated Adenosine Homeostasis in Brain Disorders with a Specific Focus on Neurodegenerative Diseases. J. Biomed. Sci. 2021, 28, 70. [Google Scholar] [CrossRef] [PubMed]
  7. Pasquini, S.; Contri, C.; Cappello, M.; Borea, P.A.; Varani, K.; Vincenzi, F. Update on the Recent Development of Allosteric Modulators for Adenosine Receptors and Their Therapeutic Applications. Front. Pharmacol. 2022, 13, 1030895. [Google Scholar] [CrossRef]
  8. 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]
  9. Shaw, S.; Uniyal, A.; Gadepalli, A.; Tiwari, V.; Belinskaia, D.A.; Shestakova, N.N.; Venugopala, K.N.; Deb, P.K.; Tiwari, V. Adenosine Receptor Signalling: Probing the Potential Pathways for the Ministration of Neuropathic Pain. Eur. J. Pharmacol. 2020, 889, 173619. [Google Scholar] [CrossRef]
  10. Jacobson, K.A.; Tosh, D.K.; Jain, S.; Gao, Z.-G. Historical and Current Adenosine Receptor Agonists in Preclinical and Clinical Development. Front. Cell. Neurosci. 2019, 13, 124. [Google Scholar] [CrossRef]
  11. Jacobson, K.A.; Giancotti, L.A.; Lauro, F.; Mufti, F.; Salvemini, D. Treatment of Chronic Neuropathic Pain: Purine Receptor Modulation. Pain 2020, 161, 1425–1441. [Google Scholar] [CrossRef]
  12. Vincenzi, F.; Pasquini, S.; Battistello, E.; Merighi, S.; Gessi, S.; Borea, P.A.; Varani, K. A1 Adenosine Receptor Partial Agonists and Allosteric Modulators: Advancing Toward the Clinic? Front. Pharmacol. 2020, 11, 625134. [Google Scholar] [CrossRef]
  13. Vincenzi, F.; Targa, M.; Romagnoli, R.; Merighi, S.; Gessi, S.; Baraldi, P.G.; Borea, P.A.; Varani, K. TRR469, a Potent A(1) Adenosine Receptor Allosteric Modulator, Exhibits Anti-Nociceptive Properties in Acute and Neuropathic Pain Models in Mice. Neuropharmacology 2014, 81, 6–14. [Google Scholar] [CrossRef]
  14. Draper-Joyce, C.J.; Bhola, R.; Wang, J.; Bhattarai, A.; Nguyen, A.T.N.; Cowie-Kent, I.; O’Sullivan, K.; Chia, L.Y.; Venugopal, H.; Valant, C.; et al. Positive Allosteric Mechanisms of Adenosine A1 Receptor-Mediated Analgesia. Nature 2021, 597, 571–576. [Google Scholar] [CrossRef]
  15. Kan, H.-W.; Chang, C.-H.; Lin, C.-L.; Lee, Y.-C.; Hsieh, S.-T.; Hsieh, Y.-L. Downregulation of Adenosine and Adenosine A1 Receptor Contributes to Neuropathic Pain in Resiniferatoxin Neuropathy. Pain 2018, 159, 1580–1591. [Google Scholar] [CrossRef]
  16. Wall, M.J.; Hill, E.; Huckstepp, R.; Barkan, K.; Deganutti, G.; Leuenberger, M.; Preti, B.; Winfield, I.; Carvalho, S.; Suchankova, A.; et al. Selective Activation of Gαob by an Adenosine A1 Receptor Agonist Elicits Analgesia without Cardiorespiratory Depression. Nat. Commun. 2022, 13, 4150. [Google Scholar] [CrossRef]
  17. Luongo, L.; Salvemini, D. Targeting Metabotropic Adenosine Receptors for Neuropathic Pain: Focus on A2A. Brain. Behav. Immun. 2018, 69, 60–61. [Google Scholar] [CrossRef]
  18. Kwilasz, A.J.; Green Fulgham, S.M.; Ellis, A.; Patel, H.P.; Duran-Malle, J.C.; Favret, J.; Harvey, L.O.; Rieger, J.; Maier, S.F.; Watkins, L.R. A Single Peri-Sciatic Nerve Administration of the Adenosine 2A Receptor Agonist ATL313 Produces Long-Lasting Anti-Allodynia and Anti-Inflammatory Effects in Male Rats. Brain. Behav. Immun. 2019, 76, 116–125. [Google Scholar] [CrossRef]
  19. Betti, M.; Catarzi, D.; Varano, F.; Falsini, M.; Varani, K.; Vincenzi, F.; Pasquini, S.; di Cesare Mannelli, L.; Ghelardini, C.; Lucarini, E.; et al. Modifications on the Amino-3,5-Dicyanopyridine Core To Obtain Multifaceted Adenosine Receptor Ligands with Antineuropathic Activity. J. Med. Chem. 2019, 62, 6894–6912. [Google Scholar] [CrossRef]
  20. Falsini, M.; Catarzi, D.; Varano, F.; Ceni, C.; Dal Ben, D.; Marucci, G.; Buccioni, M.; Volpini, R.; Di Cesare Mannelli, L.; Lucarini, E.; et al. Antioxidant-Conjugated 1,2,4-Triazolo[4,3-a]Pyrazin-3-One Derivatives: Highly Potent and Selective Human A2A Adenosine Receptor Antagonists Possessing Protective Efficacy in Neuropathic Pain. J. Med. Chem. 2019, 62, 8511–8531. [Google Scholar] [CrossRef]
  21. Varano, F.; Catarzi, D.; Vincenzi, F.; Betti, M.; Falsini, M.; Ravani, A.; Borea, P.A.; Colotta, V.; Varani, K. Design, Synthesis, and Pharmacological Characterization of 2-(2-Furanyl)Thiazolo[5,4-d]Pyrimidine-5,7-Diamine Derivatives: New Highly Potent A2A Adenosine Receptor Inverse Agonists with Antinociceptive Activity. J. Med. Chem. 2016, 59, 10564–10576. [Google Scholar] [CrossRef]
  22. Coppi, E.; Cherchi, F.; Lucarini, E.; Ghelardini, C.; Pedata, F.; Jacobson, K.A.; Di Cesare Mannelli, L.; Pugliese, A.M.; Salvemini, D. Uncovering the Mechanisms of Adenosine Receptor-Mediated Pain Control: Focus on the A3 Receptor Subtype. Int. J. Mol. Sci. 2021, 22, 7952. [Google Scholar] [CrossRef]
  23. Durante, M.; Squillace, S.; Lauro, F.; Giancotti, L.A.; Coppi, E.; Cherchi, F.; Di Cesare Mannelli, L.; Ghelardini, C.; Kolar, G.; Wahlman, C.; et al. Adenosine A3 Agonists Reverse Neuropathic Pain via T Cell-Mediated Production of IL-10. J. Clin. Investig. 2021, 131, e139299. [Google Scholar] [CrossRef]
  24. Merighi, S.; Nigro, M.; Travagli, A.; Pasquini, S.; Borea, P.A.; Varani, K.; Vincenzi, F.; Gessi, S. A2A Adenosine Receptor: A Possible Therapeutic Target for Alzheimer’s Disease by Regulating NLRP3 Inflammasome Activity? Int. J. Mol. Sci. 2022, 23, 5056. [Google Scholar] [CrossRef]
  25. Launay, A.; Nebie, O.; Vijaya Shankara, J.; Lebouvier, T.; Buée, L.; Faivre, E.; Blum, D. The Role of Adenosine A2A Receptors in Alzheimer’s Disease and Tauopathies. Neuropharmacology 2023, 226, 109379. [Google Scholar] [CrossRef]
  26. Gomez-Castro, F.; Zappettini, S.; Pressey, J.C.; Silva, C.G.; Russeau, M.; Gervasi, N.; Figueiredo, M.; Montmasson, C.; Renner, M.; Canas, P.M.; et al. Convergence of Adenosine and GABA Signaling for Synapse Stabilization during Development. Science 2021, 374, eabk2055. [Google Scholar] [CrossRef]
  27. 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]
  28. 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]
  29. 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. Alzheimers Dis. 2013, 34, 509–518. [Google Scholar] [CrossRef]
  30. 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]
  31. 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]
  32. 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. JAD 2021, 80, 1105–1117. [Google Scholar] [CrossRef] [PubMed]
  33. 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 J. Neurol. 2019, 142, 3636–3654. [Google Scholar] [CrossRef]
  34. Silva, A.C.; Lemos, C.; Gonçalves, F.Q.; Pliássova, A.V.; Machado, N.J.; Silva, H.B.; Canas, P.M.; Cunha, R.A.; Lopes, J.P.; Agostinho, P. Blockade of Adenosine A2A Receptors Recovers Early Deficits of Memory and Plasticity in the Triple Transgenic Mouse Model of Alzheimer’s Disease. Neurobiol. Dis. 2018, 117, 72–81. [Google Scholar] [CrossRef]
  35. 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]
  36. Faivre, E.; Coelho, J.E.; Zornbach, K.; Malik, E.; Baqi, Y.; Schneider, M.; Cellai, L.; Carvalho, K.; Sebda, S.; Figeac, M.; et al. Beneficial Effect of a Selective Adenosine A2A Receptor Antagonist in the APPswe/PS1dE9 Mouse Model of Alzheimer’s Disease. Front. Mol. Neurosci. 2018, 11, 235. [Google Scholar] [CrossRef]
  37. Dias, L.; Madeira, D.; Dias, R.; Tomé, Â.R.; Cunha, R.A.; Agostinho, P. Aβ1-42 Peptides Blunt the Adenosine A2A Receptor-Mediated Control of the Interplay between P2X7 and P2Y1 Receptors Mediated Calcium Responses in Astrocytes. Cell. Mol. Life Sci. CMLS 2022, 79, 457. [Google Scholar] [CrossRef]
  38. 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]
  39. Merighi, S.; Travagli, A.; Nigro, M.; Pasquini, S.; Cappello, M.; Contri, C.; Varani, K.; Vincenzi, F.; Borea, P.A.; Gessi, S. Caffeine for Prevention of Alzheimer’s Disease: Is the A2A Adenosine Receptor Its Target? Biomolecules 2023, 13, 967. [Google Scholar] [CrossRef]
  40. Paiva, I.; Cellai, L.; Meriaux, C.; Poncelet, L.; Nebie, O.; Saliou, J.-M.; Lacoste, A.-S.; Papegaey, A.; Drobecq, H.; Gras, S.L.; et al. Caffeine Intake Exerts Dual Genome-Wide Effects on Hippocampal Metabolism and Learning-Dependent Transcription. J. Clin. Investig. 2022, 132, e149371. [Google Scholar] [CrossRef]
  41. Stazi, M.; Lehmann, S.; Sakib, M.S.; Pena-Centeno, T.; Büschgens, L.; Fischer, A.; Weggen, S.; Wirths, O. Long-Term Caffeine Treatment of Alzheimer Mouse Models Ameliorates Behavioural Deficits and Neuron Loss and Promotes Cellular and Molecular Markers of Neurogenesis. Cell. Mol. Life Sci. CMLS 2021, 79, 55. [Google Scholar] [CrossRef]
  42. 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. [Google Scholar] [CrossRef]
  43. 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]
  44. Ma, L.; Day-Cooney, J.; Benavides, O.J.; Muniak, M.A.; Qin, M.; Ding, J.B.; Mao, T.; Zhong, H. Locomotion Activates PKA through Dopamine and Adenosine in Striatal Neurons. Nature 2022, 611, 762–768. [Google Scholar] [CrossRef]
  45. Ohno, Y.; Okita, E.; Kawai-Uchida, M.; Fukuda, N.; Shoukei, Y.; Soshiroda, K.; Yamada, K.; Kanda, T.; Uchida, S. Anti-Parkinsonian Activity of the Adenosine A2A Receptor Antagonist/Inverse Agonist KW-6356 as Monotherapy in MPTP-Treated Common Marmosets. Eur. J. Pharmacol. 2023, 950, 175773. [Google Scholar] [CrossRef]
  46. Ohno, Y.; Okita, E.; Kawai-Uchida, M.; Shoukei, Y.; Soshiroda, K.; Kanda, T.; Uchida, S. The Adenosine A2A Receptor Antagonist/Inverse Agonist, KW-6356 Enhances the Anti-Parkinsonian Activity of L-DOPA with a Low Risk of Dyskinesia in MPTP-Treated Common Marmosets. J. Pharmacol. Sci. 2023, 152, 193–199. [Google Scholar] [CrossRef]
  47. Carmo, M.; Gonçalves, F.Q.; Canas, P.M.; Oses, J.-P.; Fernandes, F.D.; Duarte, F.V.; Palmeira, C.M.; Tomé, A.R.; Agostinho, P.; Andrade, G.M.; et al. 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]
  48. 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 J. Neurol. 2019, 142, 700–718. [Google Scholar] [CrossRef]
  49. Gonçalves, F.Q.; Matheus, F.C.; Silva, H.B.; Real, J.I.; Rial, D.; Rodrigues, R.J.; Oses, J.-P.; Silva, A.C.; Gonçalves, N.; Prediger, R.D.; et al. Increased ATP Release and Higher Impact of Adenosine A2A Receptors on Corticostriatal Plasticity in a Rat Model of Presymptomatic Parkinson’s Disease. Mol. Neurobiol. 2023, 60, 1659–1674. [Google Scholar] [CrossRef]
  50. Ren, X.; Chen, J.-F. Caffeine and Parkinson’s Disease: Multiple Benefits and Emerging Mechanisms. Front. Neurosci. 2020, 14, 602697. [Google Scholar] [CrossRef]
  51. Kim, I.Y.; O’Reilly, É.J.; Hughes, K.C.; Gao, X.; Schwarzschild, M.A.; Ascherio, A. Differences in Parkinson’s Disease Risk with Caffeine Intake and Postmenopausal Hormone Use. J. Park. Dis. 2017, 7, 677–684. [Google Scholar] [CrossRef] [PubMed]
  52. Ishibashi, K.; Miura, Y.; Wagatsuma, K.; Toyohara, J.; Ishiwata, K.; Ishii, K. Adenosine A2A Receptor Occupancy by Caffeine After Coffee Intake in Parkinson’s Disease. Mov. Disord. Off. J. Mov. Disord. Soc. 2022, 37, 853–857. [Google Scholar] [CrossRef] [PubMed]
  53. Duarte-Silva, E.; Ulrich, H.; Oliveira-Giacomelli, Á.; Hartung, H.-P.; Meuth, S.G.; Peixoto, C.A. The Adenosinergic Signaling in the Pathogenesis and Treatment of Multiple Sclerosis. Front. Immunol. 2022, 13, 946698. [Google Scholar] [CrossRef] [PubMed]
  54. Loram, L.C.; Strand, K.A.; Taylor, F.R.; Sloane, E.; Van Dam, A.-M.; Rieger, J.; Maier, S.F.; Watkins, L.R. Adenosine 2A Receptor Agonism: A Single Intrathecal Administration Attenuates Motor Paralysis in Experimental Autoimmune Encephalopathy in Rats. Brain. Behav. Immun. 2015, 46, 50–54. [Google Scholar] [CrossRef]
  55. 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]
  56. Ingwersen, J.; Wingerath, B.; Graf, J.; Lepka, K.; Hofrichter, M.; Schröter, F.; Wedekind, F.; Bauer, A.; Schrader, J.; Hartung, H.-P.; et al. Dual Roles of the Adenosine A2a Receptor in Autoimmune Neuroinflammation. J. Neuroinflamm. 2016, 13, 48. [Google Scholar] [CrossRef]
  57. 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]
  58. Zheng, W.; Feng, Y.; Zeng, Z.; Ye, M.; Wang, M.; Liu, X.; Tang, P.; Shang, H.; Sun, X.; Lin, X.; et al. Choroid Plexus-Selective Inactivation of Adenosine A2A Receptors Protects against T Cell Infiltration and Experimental Autoimmune Encephalomyelitis. J. Neuroinflamm. 2022, 19, 52. [Google Scholar] [CrossRef]
  59. Coppi, E.; Cherchi, F.; Fusco, I.; Dettori, I.; Gaviano, L.; Magni, G.; Catarzi, D.; Colotta, V.; Varano, F.; Rossi, F.; et al. Adenosine A2B Receptors Inhibit K+ Currents and Cell Differentiation in Cultured Oligodendrocyte Precursor Cells and Modulate Sphingosine-1-Phosphate Signaling Pathway. Biochem. Pharmacol. 2020, 177, 113956. [Google Scholar] [CrossRef]
  60. Liu, Y.-J.; Chen, J.; Li, X.; Zhou, X.; Hu, Y.-M.; Chu, S.-F.; Peng, Y.; Chen, N.-H. Research Progress on Adenosine in Central Nervous System Diseases. CNS Neurosci. Ther. 2019, 25, 899–910. [Google Scholar] [CrossRef]
  61. Schädlich, I.S.; Winzer, R.; Stabernack, J.; Tolosa, E.; Magnus, T.; Rissiek, B. The Role of the ATP-Adenosine Axis in Ischemic Stroke. Semin. Immunopathol. 2023, 45, 347–365. [Google Scholar] [CrossRef] [PubMed]
  62. Cen, X.-Q.; Li, P.; Wang, B.; Chen, X.; Zhao, Y.; Yang, N.; Peng, Y.; Li, C.-H.; Ning, Y.-L.; Zhou, Y.-G. Knockdown of Adenosine A2A Receptors in Hippocampal Neurons Prevents Post-TBI Fear Memory Retrieval. Exp. Neurol. 2023, 364, 114378. [Google Scholar] [CrossRef]
  63. Farr, S.A.; Cuzzocrea, S.; Esposito, E.; Campolo, M.; Niehoff, M.L.; Doyle, T.M.; Salvemini, D. Adenosine A3 Receptor as a Novel Therapeutic Target to Reduce Secondary Events and Improve Neurocognitive Functions Following Traumatic Brain Injury. J. Neuroinflamm. 2020, 17, 339. [Google Scholar] [CrossRef] [PubMed]
  64. Pedata, F.; Dettori, I.; Coppi, E.; Melani, A.; Fusco, I.; Corradetti, R.; Pugliese, A.M. Purinergic Signalling in Brain Ischemia. Neuropharmacology 2016, 104, 105–130. [Google Scholar] [CrossRef] [PubMed]
  65. Martire, A.; Lambertucci, C.; Pepponi, R.; Ferrante, A.; Benati, N.; Buccioni, M.; Dal Ben, D.; Marucci, G.; Klotz, K.-N.; Volpini, R.; et al. Neuroprotective Potential of Adenosine A1 Receptor Partial Agonists in Experimental Models of Cerebral Ischemia. J. Neurochem. 2019, 149, 211–230. [Google Scholar] [CrossRef] [PubMed]
  66. Joya, A.; Ardaya, M.; Montilla, A.; Garbizu, M.; Plaza-García, S.; Gómez-Vallejo, V.; Padro, D.; Gutiérrez, J.J.; Rios, X.; Ramos-Cabrer, P.; et al. In Vivo Multimodal Imaging of Adenosine A1 Receptors in Neuroinflammation after Experimental Stroke. Theranostics 2021, 11, 410–425. [Google Scholar] [CrossRef]
  67. Zhou, Y.; Zeng, X.; Li, G.; Yang, Q.; Xu, J.; Zhang, M.; Mao, X.; Cao, Y.; Wang, L.; Xu, Y.; et al. Inactivation of Endothelial Adenosine A2A Receptors Protects Mice from Cerebral Ischaemia-induced Brain Injury. Br. J. Pharmacol. 2019, 176, 2250–2263. [Google Scholar] [CrossRef]
  68. Coppi, E.; Gibb, A.J. Selective Block of Adenosine A2A Receptors Prevents Ischaemic-like Effects Induced by Oxygen and Glucose Deprivation in Rat Medium Spiny Neurons. Br. J. Pharmacol. 2022, 179, 4844–4856. [Google Scholar] [CrossRef]
  69. Dettori, I.; Gaviano, L.; Ugolini, F.; Lana, D.; Bulli, I.; Magni, G.; Rossi, F.; Giovannini, M.G.; Pedata, F. Protective Effect of Adenosine A2B Receptor Agonist, BAY60-6583, Against Transient Focal Brain Ischemia in Rat. Front. Pharmacol. 2021, 11, 588757. [Google Scholar] [CrossRef]
  70. Cheng, P.; Zhang, J.; Chu, Z.; Liu, W.; Lin, H.; Wu, Y.; Zhu, J. A3 Adenosine Receptor Agonist IB-MECA Reverses Chronic Cerebral Ischemia-Induced Inhibitory Avoidance Memory Deficit. Eur. J. Pharmacol. 2022, 921, 174874. [Google Scholar] [CrossRef]
  71. Liston, T.E.; Hama, A.; Boltze, J.; Poe, R.B.; Natsume, T.; Hayashi, I.; Takamatsu, H.; Korinek, W.S.; Lechleiter, J.D. Adenosine A1R/A3R (Adenosine A1 and A3 Receptor) Agonist AST-004 Reduces Brain Infarction in a Nonhuman Primate Model of Stroke. Stroke 2022, 53, 238–248. [Google Scholar] [CrossRef]
  72. Oliveros, A.; Yoo, K.H.; Rashid, M.A.; Corujo-Ramirez, A.; Hur, B.; Sung, J.; Liu, Y.; Hawse, J.R.; Choi, D.-S.; Boison, D.; et al. Adenosine A2A Receptor Blockade Prevents Cisplatin-Induced Impairments in Neurogenesis and Cognitive Function. Proc. Natl. Acad. Sci. USA 2022, 119, e2206415119. [Google Scholar] [CrossRef]
  73. Singh, A.K.; Mahalingam, R.; Squillace, S.; Jacobson, K.A.; Tosh, D.K.; Dharmaraj, S.; Farr, S.A.; Kavelaars, A.; Salvemini, D.; Heijnen, C.J. Targeting the A3 Adenosine Receptor to Prevent and Reverse Chemotherapy-Induced Neurotoxicities in Mice. Acta Neuropathol. Commun. 2022, 10, 11. [Google Scholar] [CrossRef] [PubMed]
  74. Hargus, N.J.; Jennings, C.; Perez-Reyes, E.; Bertram, E.H.; Patel, M.K. Enhanced Actions of Adenosine in Medial Entorhinal Cortex Layer II Stellate Neurons in Temporal Lobe Epilepsy Are Mediated via A(1)-Receptor Activation. Epilepsia 2012, 53, 168–176. [Google Scholar] [CrossRef] [PubMed]
  75. Lovatt, D.; Xu, Q.; Liu, W.; Takano, T.; Smith, N.A.; Schnermann, J.; Tieu, K.; Nedergaard, M. Neuronal Adenosine Release, and Not Astrocytic ATP Release, Mediates Feedback Inhibition of Excitatory Activity. Proc. Natl. Acad. Sci. USA 2012, 109, 6265–6270. [Google Scholar] [CrossRef]
  76. Beamer, E.; Kuchukulla, M.; Boison, D.; Engel, T. ATP and Adenosine—Two Players in the Control of Seizures and Epilepsy Development. Prog. Neurobiol. 2021, 204, 102105. [Google Scholar] [CrossRef]
  77. Saggu, S.; Chen, Y.; Chen, L.; Pizarro, D.; Pati, S.; Law, W.J.; McMahon, L.; Jiao, K.; Wang, Q. A Peptide Blocking the ADORA1-Neurabin Interaction Is Anticonvulsant and Inhibits Epilepsy in an Alzheimer’s Model. JCI Insight 2022, 7, e155002. [Google Scholar] [CrossRef] [PubMed]
  78. Xie, P.; Liu, S.; Huang, Q.; Xiong, Z.; Deng, J.; Tang, C.; Xu, K.; Zhang, B.; He, B.; Wang, X.; et al. Deep Brain Stimulation Suppresses Epileptic Seizures in Rats via Inhibition of Adenosine Kinase and Activation of Adenosine A1 Receptors. CNS Neurosci. Ther. 2023, 29, 2597–2607. [Google Scholar] [CrossRef] [PubMed]
  79. Shen, H.-Y.; Baer, S.B.; Gesese, R.; Cook, J.M.; Weltha, L.; Coffman, S.Q.; Wu, J.; Chen, J.-F.; Gao, M.; Ji, T. Adenosine-A2A Receptor Signaling Plays a Crucial Role in Sudden Unexpected Death in Epilepsy. Front. Pharmacol. 2022, 13, 910535. [Google Scholar] [CrossRef]
  80. Domenici, M.R.; Ferrante, A.; Martire, A.; Chiodi, V.; Pepponi, R.; Tebano, M.T.; Popoli, P. Adenosine A2A Receptor as Potential Therapeutic Target in Neuropsychiatric Disorders. Pharmacol. Res. 2019, 147, 104338. [Google Scholar] [CrossRef]
  81. Pasquini, S.; Contri, C.; Merighi, S.; Gessi, S.; Borea, P.A.; Varani, K.; Vincenzi, F. Adenosine Receptors in Neuropsychiatric Disorders: Fine Regulators of Neurotransmission and Potential Therapeutic Targets. Int. J. Mol. Sci. 2022, 23, 1219. [Google Scholar] [CrossRef]
  82. Wang, M.; Li, P.; Li, Z.; da Silva, B.S.; Zheng, W.; Xiang, Z.; He, Y.; Xu, T.; Cordeiro, C.; Deng, L.; et al. Lateral Septum Adenosine A2A Receptors Control Stress-Induced Depressive-like Behaviors via Signaling to the Hypothalamus and Habenula. Nat. Commun. 2023, 14, 1880. [Google Scholar] [CrossRef]
  83. Yu, W.; Wu, Z.; Li, X.; Ding, M.; Xu, Y.; Zhao, P. Ketamine Counteracts Sevoflurane-Induced Depressive-like Behavior and Synaptic Plasticity Impairments through the Adenosine A2A Receptor/ERK Pathway in Rats. Mol. Neurobiol. 2023. ahead-of-print. [Google Scholar] [CrossRef]
  84. Xu, Y.; Ning, Y.; Zhao, Y.; Peng, Y.; Luo, F.; Zhou, Y.; Li, P. Caffeine Functions by Inhibiting Dorsal and Ventral Hippocampal Adenosine 2A Receptors to Modulate Memory and Anxiety, Respectively. Front. Pharmacol. 2022, 13, 807330. [Google Scholar] [CrossRef] [PubMed]
  85. Florén Lind, S.; Stam, F.; Zelleroth, S.; Meurling, E.; Frick, A.; Grönbladh, A. Acute Caffeine Differently Affects Risk-Taking and the Expression of BDNF and of Adenosine and Opioid Receptors in Rats with High or Low Anxiety-like Behavior. Pharmacol. Biochem. Behav. 2023, 227–228, 173573. [Google Scholar] [CrossRef]
  86. Singer, P.; Yee, B.K. The Adenosine Hypothesis of Schizophrenia into Its Third Decade: From Neurochemical Imbalance to Early Life Etiological Risks. Front. Cell. Neurosci. 2023, 17, 1120532. [Google Scholar] [CrossRef] [PubMed]
  87. Boison, D.; Singer, P.; Shen, H.-Y.; Feldon, J.; Yee, B.K. Adenosine Hypothesis of Schizophrenia--Opportunities for Pharmacotherapy. Neuropharmacology 2012, 62, 1527–1543. [Google Scholar] [CrossRef] [PubMed]
  88. Valle-León, M.; Callado, L.F.; Aso, E.; Cajiao-Manrique, M.M.; Sahlholm, K.; López-Cano, M.; Soler, C.; Altafaj, X.; Watanabe, M.; Ferré, S.; et al. Decreased Striatal Adenosine A2A-Dopamine D2 Receptor Heteromerization in Schizophrenia. Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2021, 46, 665–672. [Google Scholar] [CrossRef]
  89. Valle-León, M.; Casajuana-Martin, N.; Del Torrent, C.L.; Argerich, J.; Gómez-Acero, L.; Sahlholm, K.; Ferré, S.; Pardo, L.; Ciruela, F. Unique Effect of Clozapine on Adenosine A2A-Dopamine D2 Receptor Heteromerization. Biomed. Pharmacother. Biomed. Pharmacother. 2023, 160, 114327. [Google Scholar] [CrossRef]
  90. Shalaby, H.N.; Zaki, H.F.; Ain-Shoka, A.A.A.; Mohammed, R.A. Adenosine A2A Receptor Blockade Ameliorates Mania Like Symptoms in Rats: Signaling to PKC-α and Akt/GSK-3β/β-Catenin. Mol. Neurobiol. 2022, 59, 6397–6410. [Google Scholar] [CrossRef]
  91. Ma, W.-X.; Yuan, P.-C.; Zhang, H.; Kong, L.-X.; Lazarus, M.; Qu, W.-M.; Wang, Y.-Q.; Huang, Z.-L. Adenosine and P1 Receptors: Key Targets in the Regulation of Sleep, Torpor, and Hibernation. Front. Pharmacol. 2023, 14, 1098976. [Google Scholar] [CrossRef] [PubMed]
  92. Peng, W.; Wu, Z.; Song, K.; Zhang, S.; Li, Y.; Xu, M. Regulation of Sleep Homeostasis Mediator Adenosine by Basal Forebrain Glutamatergic Neurons. Science 2020, 369, eabb0556. [Google Scholar] [CrossRef] [PubMed]
  93. Jagannath, A.; Varga, N.; Dallmann, R.; Rando, G.; Gosselin, P.; Ebrahimjee, F.; Taylor, L.; Mosneagu, D.; Stefaniak, J.; Walsh, S.; et al. Adenosine Integrates Light and Sleep Signalling for the Regulation of Circadian Timing in Mice. Nat. Commun. 2021, 12, 2113. [Google Scholar] [CrossRef]
  94. Kim, T.-H.; Bormate, K.J.; Custodio, R.J.P.; Cheong, J.H.; Lee, B.K.; Kim, H.J.; Jung, Y.-S. Involvement of the Adenosine A1 Receptor in the Hypnotic Effect of Rosmarinic Acid. Biomed. Pharmacother. Biomed. Pharmacother. 2022, 146, 112483. [Google Scholar] [CrossRef]
  95. Lin, Y.; Roy, K.; Ioka, S.; Otani, R.; Amezawa, M.; Ishikawa, Y.; Cherasse, Y.; Kaushik, M.K.; Klewe-Nebenius, D.; Zhou, L.; et al. Positive Allosteric Adenosine A2A Receptor Modulation Suppresses Insomnia Associated with Mania- and Schizophrenia-like Behaviors in Mice. Front. Pharmacol. 2023, 14, 1072. [Google Scholar] [CrossRef] [PubMed]
  96. Li, R.; Wang, Y.-Q.; Liu, W.-Y.; Zhang, M.-Q.; Li, L.; Cherasse, Y.; Schiffmann, S.N.; de Kerchove d’Exaerde, A.; Lazarus, M.; Qu, W.-M.; et al. Activation of Adenosine A2A Receptors in the Olfactory Tubercle Promotes Sleep in Rodents. Neuropharmacology 2020, 168, 107923. [Google Scholar] [CrossRef]
  97. Choi, I.-S.; Kim, J.-H.; Jeong, J.-Y.; Lee, M.-G.; Suk, K.; Jang, I.-S. Astrocyte-Derived Adenosine Excites Sleep-Promoting Neurons in the Ventrolateral Preoptic Nucleus: Astrocyte-Neuron Interactions in the Regulation of Sleep. Glia 2022, 70, 1864–1885. [Google Scholar] [CrossRef]
  98. Ye, S.-S.; Tang, Y.; Song, J.-T. ATP and Adenosine in the Retina and Retinal Diseases. Front. Pharmacol. 2021, 12, 654445. [Google Scholar] [CrossRef]
  99. Boia, R.; Salinas-Navarro, M.; Gallego-Ortega, A.; Galindo-Romero, C.; Aires, I.D.; Agudo-Barriuso, M.; Ambrósio, A.F.; Vidal-Sanz, M.; Santiago, A.R. Activation of Adenosine A3 Receptor Protects Retinal Ganglion Cells from Degeneration Induced by Ocular Hypertension. Cell Death Dis. 2020, 11, 401. [Google Scholar] [CrossRef]
  100. Boia, R.; Dias, P.A.N.; Galindo-Romero, C.; Ferreira, H.; Aires, I.D.; Vidal-Sanz, M.; Agudo-Barriuso, M.; Bernardes, R.; Santos, P.F.; de Sousa, H.C.; et al. Intraocular Implants Loaded with A3R Agonist Rescue Retinal Ganglion Cells from Ischemic Damage. J. Control. Release 2022, 343, 469–481. [Google Scholar] [CrossRef]
  101. Fang, G.; Zhou, Y.; Zhou, X.; Zhou, H.; Ge, Y.-Y.; Luo, S.; Chen, J.-F.; Zhang, L. The Adenosine A2A Receptor Antagonist Protects against Retinal Mitochondrial Injury in Association with an Altered Network of Competing Endogenous RNAs. Neuropharmacology 2022, 208, 108981. [Google Scholar] [CrossRef] [PubMed]
  102. Hu, S.; Li, Y.; Zhang, Y.; Shi, R.; Tang, P.; Zhang, D.; Kuang, X.; Chen, J.; Qu, J.; Gao, Y. The Adenosine A2A Receptor Antagonist KW6002 Distinctly Regulates Retinal Ganglion Cell Morphology during Postnatal Development and Neonatal Inflammation. Front. Pharmacol. 2022, 13, 1082997. [Google Scholar] [CrossRef] [PubMed]
  103. Headrick, J.P.; Ashton, K.J.; Rose’Meyer, R.B.; Peart, J.N. Cardiovascular Adenosine Receptors: Expression, Actions and Interactions. Pharmacol. Ther. 2013, 140, 92–111. [Google Scholar] [CrossRef]
  104. Sanni, O.; Terre’Blanche, G. Therapeutic Potentials of Agonist and Antagonist of Adenosine Receptors in Type 2 Diabetes. Rev. Endocr. Metab. Disord. 2021, 22, 1073–1090. [Google Scholar] [CrossRef] [PubMed]
  105. Cai, Y.; Chen, X.; Yi, B.; Li, J.; Wen, Z. Pathophysiology Roles for Adenosine 2A Receptor in Obesity and Related Diseases. Obes. Rev. Off. J. Int. Assoc. Study Obes. 2022, 23, e13490. [Google Scholar] [CrossRef]
  106. Ruan, W.; Ma, X.; Bang, I.H.; Liang, Y.; Muehlschlegel, J.D.; Tsai, K.-L.; Mills, T.W.; Yuan, X.; Eltzschig, H.K. The Hypoxia-Adenosine Link during Myocardial Ischemia-Reperfusion Injury. Biomedicines 2022, 10, 1939. [Google Scholar] [CrossRef]
  107. Xu, S.; Gu, R.; Bian, X.; Xu, X.; Xia, X.; Liu, Y.; Jia, C.; Gu, Y.; Zhang, H. Remote Conditioning by Rhythmic Compression of Limbs Ameliorated Myocardial Infarction by Downregulation of Inflammation via A2 Adenosine Receptors. Front. Cardiovasc. Med. 2022, 8, 723332. [Google Scholar] [CrossRef]
  108. Mehaffey, J.H.; Money, D.; Charles, E.J.; Schubert, S.; Piñeros, A.F.; Wu, D.; Bontha, S.V.; Hawkins, R.; Teman, N.R.; Laubach, V.E.; et al. Adenosine 2A Receptor Activation Attenuates Ischemia Reperfusion Injury During Extracorporeal Cardiopulmonary Resuscitation. Ann. Surg. 2019, 269, 1176–1183. [Google Scholar] [CrossRef]
  109. Ruan, W.; Li, J.; Choi, S.; Ma, X.; Liang, Y.; Nair, R.; Yuan, X.; Mills, T.W.; Eltzschig, H.K. Targeting Myocardial Equilibrative Nucleoside Transporter ENT1 Provides Cardioprotection by Enhancing Myeloid Adora2b Signaling. JCI Insight 2023, 8, e166011. [Google Scholar] [CrossRef]
  110. Reichert, K.P.; Castro, M.F.V.; Assmann, C.E.; Bottari, N.B.; Miron, V.V.; Cardoso, A.; Stefanello, N.; Morsch, V.M.M.; Schetinger, M.R.C. Diabetes and Hypertension: Pivotal Involvement of Purinergic Signaling. Biomed. Pharmacother. Biomed. Pharmacother. 2021, 137, 111273. [Google Scholar] [CrossRef]
  111. Yadav, V.R.; Teng, B.; Mustafa, S.J. Enhanced A1 Adenosine Receptor-Induced Vascular Contractions in Mesenteric Artery and Aorta of in L-NAME Mouse Model of Hypertension. Eur. J. Pharmacol. 2019, 842, 111–117. [Google Scholar] [CrossRef]
  112. Jackson, E.K.; Gillespie, D.G.; Mi, Z.; Cheng, D. Adenosine Receptors Influence Hypertension in Dahl Salt-Sensitive Rats: Dependence on Receptor Subtype, Salt Diet, and Sex. Hypertens. Dallas Tex 1979 2018, 72, 511–521. [Google Scholar] [CrossRef] [PubMed]
  113. Ruan, C.-C.; Kong, L.-R.; Chen, X.-H.; Ma, Y.; Pan, X.-X.; Zhang, Z.-B.; Gao, P.-J. A2A Receptor Activation Attenuates Hypertensive Cardiac Remodeling via Promoting Brown Adipose Tissue-Derived FGF21. Cell Metab. 2018, 28, 476–489.e5. [Google Scholar] [CrossRef]
  114. Ahmad, A.; Ahmad, S.; Glover, L.; Miller, S.M.; Shannon, J.M.; Guo, X.; Franklin, W.A.; Bridges, J.P.; Schaack, J.B.; Colgan, S.P.; et al. Adenosine A2A Receptor Is a Unique Angiogenic Target of HIF-2α in Pulmonary Endothelial Cells. Proc. Natl. Acad. Sci. USA 2009, 106, 10684–10689. [Google Scholar] [CrossRef]
  115. Angioni, R.; Liboni, C.; Herkenne, S.; Sánchez-Rodríguez, R.; Borile, G.; Marcuzzi, E.; Calì, B.; Muraca, M.; Viola, A. CD73+ Extracellular Vesicles Inhibit Angiogenesis through Adenosine A2B Receptor Signalling. J. Extracell. Vesicles 2020, 9, 1757900. [Google Scholar] [CrossRef] [PubMed]
  116. Wu, Z.; Xiang, Q.; Feng, L.; Wu, D.; Huang, S.; Zhang, L.; Rao, S.; Luo, J.; Xiong, W.; Deng, J.; et al. Adenosine-ADORA2A Promotes Ang-Induced Angiogenesis in Intrauterine Growth Restriction Placenta via the Stat3/Akt Pathway. Arterioscler. Thromb. Vasc. Biol. 2023, 43, e190–e209. [Google Scholar] [CrossRef]
  117. Wu, Z.; Nie, J.; Wu, D.; Huang, S.; Chen, J.; Liang, H.; Hao, X.; Feng, L.; Luo, H.; Tan, C. Dietary Adenosine Supplementation Improves Placental Angiogenesis in IUGR Piglets by Up-Regulating Adenosine A2a Receptor. Anim. Nutr. Zhongguo Xu Mu Shou Yi Xue Hui 2023, 13, 282–288. [Google Scholar] [CrossRef] [PubMed]
  118. Gnad, T.; Navarro, G.; Lahesmaa, M.; Reverte-Salisa, L.; Copperi, F.; Cordomi, A.; Naumann, J.; Hochhäuser, A.; Haufs-Brusberg, S.; Wenzel, D.; et al. Adenosine/A2B Receptor Signaling Ameliorates the Effects of Aging and Counteracts Obesity. Cell Metab. 2020, 32, 56–70.e7. [Google Scholar] [CrossRef]
  119. Kong, L.-R.; Chen, X.-H.; Sun, Q.; Zhang, K.-Y.; Xu, L.; Ding, L.; Zhou, Y.-P.; Zhang, Z.-B.; Lin, J.-R.; Gao, P.-J. Loss of C3a and C5a Receptors Promotes Adipocyte Browning and Attenuates Diet-Induced Obesity via Activating Inosine/A2aR Pathway. Cell Rep. 2023, 42, 112078. [Google Scholar] [CrossRef]
  120. Sacramento, J.F.; Martins, F.O.; Rodrigues, T.; Matafome, P.; Ribeiro, M.J.; Olea, E.; Conde, S.V. A2 Adenosine Receptors Mediate Whole-Body Insulin Sensitivity in a Prediabetes Animal Model: Primary Effects on Skeletal Muscle. Front. Endocrinol. 2020, 11, 262. [Google Scholar] [CrossRef]
  121. Antonioli, L.; Fornai, M.; Blandizzi, C.; Pacher, P.; Haskó, G. Adenosine Signaling and the Immune System: When a Lot Could Be Too Much. Immunol. Lett. 2019, 205, 9–15. [Google Scholar] [CrossRef]
  122. Boncler, M.; Bartczak, K.; Rozalski, M. Potential for Modulation of Platelet Function via Adenosine Receptors during Inflammation. Br. J. Pharmacol. 2023. ahead-of-print. [Google Scholar] [CrossRef] [PubMed]
  123. Cronstein, B.N.; Sitkovsky, M. Adenosine and Adenosine Receptors in the Pathogenesis and Treatment of Rheumatic Diseases. Nat. Rev. Rheumatol. 2017, 13, 41–51. [Google Scholar] [CrossRef]
  124. Varani, K.; Vincenzi, F.; Tosi, A.; Targa, M.; Masieri, F.F.; Ongaro, A.; De Mattei, M.; Massari, L.; Borea, P.A. Expression and Functional Role of Adenosine Receptors in Regulating Inflammatory Responses in Human Synoviocytes. Br. J. Pharmacol. 2010, 160, 101–115. [Google Scholar] [CrossRef]
  125. Varani, K.; Padovan, M.; Vincenzi, F.; Targa, M.; Trotta, F.; Govoni, M.; Borea, P.A. A2A and A3 Adenosine Receptor Expression in Rheumatoid Arthritis: Upregulation, Inverse Correlation with Disease Activity Score and Suppression of Inflammatory Cytokine and Metalloproteinase Release. Arthritis Res. Ther. 2011, 13, R197. [Google Scholar] [CrossRef] [PubMed]
  126. Vincenzi, F.; Padovan, M.; Targa, M.; Corciulo, C.; Giacuzzo, S.; Merighi, S.; Gessi, S.; Govoni, M.; Borea, P.A.; Varani, K. A(2A) Adenosine Receptors Are Differentially Modulated by Pharmacological Treatments in Rheumatoid Arthritis Patients and Their Stimulation Ameliorates Adjuvant-Induced Arthritis in Rats. PLoS ONE 2013, 8, e54195. [Google Scholar] [CrossRef]
  127. Bortoluzzi, A.; Vincenzi, F.; Govoni, M.; Padovan, M.; Ravani, A.; Borea, P.A.; Varani, K. A2A Adenosine Receptor Upregulation Correlates with Disease Activity in Patients with Systemic Lupus Erythematosus. Arthritis Res. Ther. 2016, 18, 192. [Google Scholar] [CrossRef]
  128. Ravani, A.; Vincenzi, F.; Bortoluzzi, A.; Padovan, M.; Pasquini, S.; Gessi, S.; Merighi, S.; Borea, P.A.; Govoni, M.; Varani, K. Role and Function of A2A and A₃ Adenosine Receptors in Patients with Ankylosing Spondylitis, Psoriatic Arthritis and Rheumatoid Arthritis. Int. J. Mol. Sci. 2017, 18, 697. [Google Scholar] [CrossRef] [PubMed]
  129. Cohen, S.; Barer, F.; Bar-Yehuda, S.; IJzerman, A.P.; Jacobson, K.A.; Fishman, P. A₃ Adenosine Receptor Allosteric Modulator Induces an Anti-Inflammatory Effect: In Vivo Studies and Molecular Mechanism of Action. Mediat. Inflamm. 2014, 2014, 708746. [Google Scholar] [CrossRef]
  130. Ochaion, A.; Bar-Yehuda, S.; Cohen, S.; Amital, H.; Jacobson, K.A.; Joshi, B.V.; Gao, Z.G.; Barer, F.; Patoka, R.; Del Valle, L.; et al. The A3 Adenosine Receptor Agonist CF502 Inhibits the PI3K, PKB/Akt and NF-KappaB Signaling Pathway in Synoviocytes from Rheumatoid Arthritis Patients and in Adjuvant-Induced Arthritis Rats. Biochem. Pharmacol. 2008, 76, 482–494. [Google Scholar] [CrossRef]
  131. Flögel, U.; Burghoff, S.; van Lent, P.L.E.M.; Temme, S.; Galbarz, L.; Ding, Z.; El-Tayeb, A.; Huels, S.; Bönner, F.; Borg, N.; et al. Selective Activation of Adenosine A2A Receptors on Immune Cells by a CD73-Dependent Prodrug Suppresses Joint Inflammation in Experimental Rheumatoid Arthritis. Sci. Transl. Med. 2012, 4, 146ra108. [Google Scholar] [CrossRef]
  132. Schmiel, S.E.; Kalekar, L.A.; Zhang, N.; Blankespoor, T.W.; Robinson, L.J.; Mueller, D.L. Adenosine 2a Receptor Signals Block Autoimmune Arthritis by Inhibiting Pathogenic Germinal Center T Follicular Helper Cells. Arthritis Rheumatol. 2019, 71, 773–783. [Google Scholar] [CrossRef]
  133. Sadatpour, O.; Ebrahimi, M.T.; Akhtari, M.; Ahmadzadeh, N.; Vojdanian, M.; Jamshidi, A.; Farhadi, E.; Mahmoudi, M. A2A Adenosine Receptor Agonist Reduced MMP8 Expression in Healthy M2-like Macrophages but Not in Macrophages from Ankylosing Spondylitis Patients. BMC Musculoskelet. Disord. 2022, 23, 908. [Google Scholar] [CrossRef]
  134. Winslow, G.M.; Papillion, A.M.; Kenderes, K.J.; Levack, R.C. CD11c+ T-Bet+ Memory B Cells: Immune Maintenance during Chronic Infection and Inflammation? Cell. Immunol. 2017, 321, 8–17. [Google Scholar] [CrossRef]
  135. Levack, R.C.; Newell, K.L.; Cabrera-Martinez, B.; Cox, J.; Perl, A.; Bastacky, S.I.; Winslow, G.M. Adenosine Receptor 2a Agonists Target Mouse CD11c+T-Bet+ B Cells in Infection and Autoimmunity. Nat. Commun. 2022, 13, 452. [Google Scholar] [CrossRef]
  136. López-Cano, M.; Filgaira, I.; Nolen, E.G.; Cabré, G.; Hernando, J.; Tosh, D.K.; Jacobson, K.A.; Soler, C.; Ciruela, F. Optical Control of Adenosine A3 Receptor Function in Psoriasis. Pharmacol. Res. 2021, 170, 105731. [Google Scholar] [CrossRef]
  137. Fishman, P. Drugs Targeting the A3 Adenosine Receptor: Human Clinical Study Data. Molecules 2022, 27, 3680. [Google Scholar] [CrossRef]
  138. Carmona-Rivera, C.; Khaznadar, S.S.; Shwin, K.W.; Irizarry-Caro, J.A.; O’Neil, L.J.; Liu, Y.; Jacobson, K.A.; Ombrello, A.K.; Stone, D.L.; Tsai, W.L.; et al. Deficiency of Adenosine Deaminase 2 Triggers Adenosine-Mediated NETosis and TNF Production in Patients with DADA2. Blood 2019, 134, 395–406. [Google Scholar] [CrossRef]
  139. Bekisz, J.M.; Lopez, C.D.; Corciulo, C.; Mediero, A.; Coelho, P.G.; Witek, L.; Flores, R.L.; Cronstein, B.N. The Role of Adenosine Receptor Activation in Attenuating Cartilaginous Inflammation. Inflammation 2018, 41, 1135–1141. [Google Scholar] [CrossRef]
  140. Corciulo, C.; Lendhey, M.; Wilder, T.; Schoen, H.; Cornelissen, A.S.; Chang, G.; Kennedy, O.D.; Cronstein, B.N. Endogenous Adenosine Maintains Cartilage Homeostasis and Exogenous Adenosine Inhibits Osteoarthritis Progression. Nat. Commun. 2017, 8, 15019. [Google Scholar] [CrossRef]
  141. Friedman, B.; Larranaga-Vera, A.; Castro, C.M.; Corciulo, C.; Rabbani, P.; Cronstein, B.N. Adenosine A2A Receptor Activation Reduces Chondrocyte Senescence. FASEB J. 2023, 37, e22838. [Google Scholar] [CrossRef] [PubMed]
  142. Castro, C.M.; Corciulo, C.; Solesio, M.E.; Liang, F.; Pavlov, E.V.; Cronstein, B.N. Adenosine A2A Receptor (A2AR) Stimulation Enhances Mitochondrial Metabolism and Mitigates Reactive Oxygen Species-Mediated Mitochondrial Injury. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2020, 34, 5027–5045. [Google Scholar] [CrossRef]
  143. Friedman, B.; Corciulo, C.; Castro, C.M.; Cronstein, B.N. Adenosine A2A Receptor Signaling Promotes FoxO Associated Autophagy in Chondrocytes. Sci. Rep. 2021, 11, 968. [Google Scholar] [CrossRef]
  144. Bai, H.; Zhang, Z.; Liu, L.; Wang, X.; Song, X.; Gao, L. Activation of Adenosine A3 Receptor Attenuates Progression of Osteoarthritis through Inhibiting the NLRP3/Caspase-1/GSDMD Induced Signalling. J. Cell. Mol. Med. 2022, 26, 4230–4243. [Google Scholar] [CrossRef] [PubMed]
  145. Le, T.-T.T.; Berg, N.K.; Harting, M.T.; Li, X.; Eltzschig, H.K.; Yuan, X. Purinergic Signaling in Pulmonary Inflammation. Front. Immunol. 2019, 10, 1633. [Google Scholar] [CrossRef] [PubMed]
  146. Zhou, Y.; Schneider, D.J.; Blackburn, M.R. Adenosine Signaling and the Regulation of Chronic Lung Disease. Pharmacol. Ther. 2009, 123, 105–116. [Google Scholar] [CrossRef] [PubMed]
  147. Wendell, S.G.; Fan, H.; Zhang, C. G Protein–Coupled Receptors in Asthma Therapy: Pharmacology and Drug Action. Pharmacol. Rev. 2020, 72, 1–49. [Google Scholar] [CrossRef]
  148. Basu, S.; Barawkar, D.A.; Ramdas, V.; Patel, M.; Waman, Y.; Panmand, A.; Kumar, S.; Thorat, S.; Naykodi, M.; Goswami, A.; et al. Design and Synthesis of Novel Xanthine Derivatives as Potent and Selective A2B Adenosine Receptor Antagonists for the Treatment of Chronic Inflammatory Airway Diseases. Eur. J. Med. Chem. 2017, 134, 218–229. [Google Scholar] [CrossRef]
  149. Ramos-Barbon, D.; Brienza, N.S.; Rodríguez, T.B.; Medina, É.F.M.; Saladich, I.G.; Rodríguez, M.P.; Arbos, R.M.A.; Gancedo, S.Q.; Laria, N.C.P.; Laria, J.C.P. PBF-680, an Oral A1 Adenosine Receptor Antagonist, Inhibits the Late Allergic Response (LAR) in Mild-to-Moderate Atopic Asthmatics: A Phase-IIa Trial. Eur. Respir. J. 2020, 56, 4784. [Google Scholar] [CrossRef]
  150. Ramos-Barbon, D.; Brienza, N.S.; Rodríguez, T.B.; Saladich, I.G.; Rodríguez, M.P.; Arbos, R.M.A.; Laria, N.C.P.; Laria, J.C.P. PBF-680, an Oral A1 Adenosine Receptor Antagonist, Inhibits Adenosine Monophosphate (AMP) Airway Hyperresponsiveness (AHR) in Mild-to-Moderate Asthma: A Phase-IIa Proof-of-Concept Trial. Eur. Respir. J. 2020, 56, 2279. [Google Scholar] [CrossRef]
  151. Xiao, Q.; Han, X.; Liu, G.; Zhou, D.; Zhang, L.; He, J.; Xu, H.; Zhou, P.; Yang, Q.; Chen, J.; et al. Adenosine Restrains ILC2-Driven Allergic Airway Inflammation via A2A Receptor. Mucosal Immunol. 2022, 15, 338–350. [Google Scholar] [CrossRef] [PubMed]
  152. Ko, I.-G.; Hwang, J.J.; Chang, B.S.; Kim, S.-H.; Jin, J.-J.; Hwang, L.; Kim, C.-J.; Choi, C.W. Polydeoxyribonucleotide Ameliorates Lipopolysaccharide-Induced Acute Lung Injury via Modulation of the MAPK/NF-ΚB Signaling Pathway in Rats. Int. Immunopharmacol. 2020, 83, 106444. [Google Scholar] [CrossRef]
  153. Li, X.; Berg, N.K.; Mills, T.; Zhang, K.; Eltzschig, H.K.; Yuan, X. Adenosine at the Interphase of Hypoxia and Inflammation in Lung Injury. Front. Immunol. 2020, 11, 604944. [Google Scholar] [CrossRef] [PubMed]
  154. Korb, V.G.; Schultz, I.C.; Beckenkamp, L.R.; Wink, M.R. A Systematic Review of the Role of Purinergic Signalling Pathway in the Treatment of COVID-19. Int. J. Mol. Sci. 2023, 24, 7865. [Google Scholar] [CrossRef] [PubMed]
  155. Dorneles, G.P.; Teixeira, P.C.; da Silva, I.M.; Schipper, L.L.; Santana Filho, P.C.; Rodrigues Junior, L.C.; Bonorino, C.; Peres, A.; Fonseca, S.G.; Monteiro, M.C.; et al. Alterations in CD39/CD73 Axis of T Cells Associated with COVID-19 Severity. J. Cell. Physiol. 2022, 237, 3394–3407. [Google Scholar] [CrossRef]
  156. Halpin-Veszeleiova, K.; Hatfield, S.M. Therapeutic Targeting of Hypoxia-A2-Adenosinergic Pathway in COVID-19 Patients. Physiology 2022, 37, 46–52. [Google Scholar] [CrossRef]
  157. Correale, P.; Caracciolo, M.; Bilotta, F.; Conte, M.; Cuzzola, M.; Falcone, C.; Mangano, C.; Falzea, A.C.; Iuliano, E.; Morabito, A.; et al. Therapeutic Effects of Adenosine in High Flow 21% Oxygen Aereosol in Patients with Covid19-Pneumonia. PLoS ONE 2020, 15, e0239692. [Google Scholar] [CrossRef]
  158. Ngamsri, K.-C.; Fuhr, A.; Schindler, K.; Simelitidis, M.; Hagen, M.; Zhang, Y.; Gamper-Tsigaras, J.; Konrad, F.M. Sevoflurane Dampens Acute Pulmonary Inflammation via the Adenosine Receptor A2B and Heme Oxygenase-1. Cells 2022, 11, 1094. [Google Scholar] [CrossRef]
  159. Ren, Y.; Qi, L.; Zhang, L.; Xu, J.; Ma, J.; Lv, Y.; Zhang, Y.; Wu, R. Cupping Alleviates Lung Injury through the Adenosine/A2BAR Pathway. Heliyon 2022, 8, e12141. [Google Scholar] [CrossRef]
  160. Boo, H.-J.; Park, S.J.; Noh, M.; Min, H.-Y.; Jeong, L.S.; Lee, H.-Y. LJ-529, a Partial Peroxisome Proliferator-Activated Receptor Gamma (PPARγ) Agonist and Adenosine A3 Receptor Agonist, Ameliorates Elastase-Induced Pulmonary Emphysema in Mice. Arch. Pharm. Res. 2020, 43, 540–552. [Google Scholar] [CrossRef]
  161. Sgambellone, S.; Marri, S.; Catarinicchia, S.; Pini, A.; Tosh, D.K.; Jacobson, K.A.; Masini, E.; Salvemini, D.; Lucarini, L. Adenosine A3 Receptor (A3AR) Agonist for the Treatment of Bleomycin-Induced Lung Fibrosis in Mice. Int. J. Mol. Sci. 2022, 23, 13300. [Google Scholar] [CrossRef] [PubMed]
  162. Chen, S.; Wu, Q.; Zhong, D.; Li, C.; Du, L. Caffeine Prevents Hyperoxia-Induced Lung Injury in Neonatal Mice through NLRP3 Inflammasome and NF-ΚB Pathway. Respir. Res. 2020, 21, 140. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, T.; Yu-Jing, L.; Ma, T. The Immunomodulatory Function of Adenosine in Sepsis. Front. Immunol. 2022, 13, 936547. [Google Scholar] [CrossRef] [PubMed]
  164. Nascimento, D.C.; Viacava, P.R.; Ferreira, R.G.; Damaceno, M.A.; Piñeros, A.R.; Melo, P.H.; Donate, P.B.; Toller-Kawahisa, J.E.; Zoppi, D.; Veras, F.P.; et al. Sepsis Expands a CD39+ Plasmablast Population That Promotes Immunosuppression via Adenosine-Mediated Inhibition of Macrophage Antimicrobial Activity. Immunity 2021, 54, 2024–2041. [Google Scholar] [CrossRef] [PubMed]
  165. Ngamsri, K.-C.; Putri, R.A.; Jans, C.; Schindler, K.; Fuhr, A.; Zhang, Y.; Gamper-Tsigaras, J.; Ehnert, S.; Konrad, F.M. CXCR4 and CXCR7 Inhibition Ameliorates the Formation of Platelet-Neutrophil Complexes and Neutrophil Extracellular Traps through Adora2b Signaling. Int. J. Mol. Sci. 2021, 22, 13576. [Google Scholar] [CrossRef]
  166. Ngamsri, K.-C.; Fabian, F.; Fuhr, A.; Gamper-Tsigaras, J.; Straub, A.; Fecher, D.; Steinke, M.; Walles, H.; Reutershan, J.; Konrad, F.M. Sevoflurane Exerts Protective Effects in Murine Peritonitis-Induced Sepsis via Hypoxia-Inducible Factor 1α/Adenosine A2B Receptor Signaling. Anesthesiology 2021, 135, 136–150. [Google Scholar] [CrossRef]
  167. Yegutkin, G.G.; Boison, D. ATP and Adenosine Metabolism in Cancer: Exploitation for Therapeutic Gain. Pharmacol. Rev. 2022, 74, 797–822. [Google Scholar] [CrossRef]
  168. Zahavi, D.; Hodge, J.W. Targeting Immunosuppressive Adenosine Signaling: A Review of Potential Immunotherapy Combination Strategies. Int. J. Mol. Sci. 2023, 24, 8871. [Google Scholar] [CrossRef]
  169. Kang, C.; Liu, L.; Wu, C.; Li, L.; Jia, X.; Xie, W.; Chen, S.; Wu, X.; Zheng, H.; Liu, J.; et al. The Adenosinergic Machinery in Cancer: In-Tandem Insights from Basic Mechanisms to Therapy. Front. Immunol. 2023, 14, 1111369. [Google Scholar] [CrossRef]
  170. Xia, C.; Yin, S.; To, K.K.W.; Fu, L. CD39/CD73/A2AR Pathway and Cancer Immunotherapy. Mol. Cancer 2023, 22, 44. [Google Scholar] [CrossRef]
  171. Slaats, J.; Wagena, E.; Smits, D.; Berends, A.A.; Peters, E.; Bakker, G.-J.; van Erp, M.; Weigelin, B.; Adema, G.J.; Friedl, P. Adenosine A2a Receptor Antagonism Restores Additive Cytotoxicity by Cytotoxic T Cells in Metabolically Perturbed Tumors. Cancer Immunol. Res. 2022, 10, 1462–1474. [Google Scholar] [CrossRef] [PubMed]
  172. Mastelic-Gavillet, B.; Navarro Rodrigo, B.; Décombaz, L.; Wang, H.; Ercolano, G.; Ahmed, R.; Lozano, L.E.; Ianaro, A.; Derré, L.; Valerio, M.; et al. Adenosine Mediates Functional and Metabolic Suppression of Peripheral and Tumor-Infiltrating CD8+ T Cells. J. Immunother. Cancer 2019, 7, 257. [Google Scholar] [CrossRef]
  173. Arruga, F.; Serra, S.; Vitale, N.; Guerra, G.; Papait, A.; Gyau, B.B.; Tito, F.; Efremov, D.; Vaisitti, T.; Deaglio, S. Targeting the A2A Adenosine Receptor Counteracts Immunosuppression in Vivo in a Mouse Model of Chronic Lymphocytic Leukemia. Haematologica 2020, 106, 1343–1353. [Google Scholar] [CrossRef] [PubMed]
  174. Wang, J.; Wang, Y.; Chu, Y.; Li, Z.; Yu, X.; Huang, Z.; Xu, J.; Zheng, L. Tumor-Derived Adenosine Promotes Macrophage Proliferation in Human Hepatocellular Carcinoma. J. Hepatol. 2021, 74, 627–637. [Google Scholar] [CrossRef] [PubMed]
  175. Borodovsky, A.; Barbon, C.M.; Wang, Y.; Ye, M.; Prickett, L.; Chandra, D.; Shaw, J.; Deng, N.; Sachsenmeier, K.; Clarke, J.D.; et al. Small Molecule AZD4635 Inhibitor of A2AR Signaling Rescues Immune Cell Function Including CD103+ Dendritic Cells Enhancing Anti-Tumor Immunity. J. Immunother. Cancer 2020, 8, e000417. [Google Scholar] [CrossRef]
  176. Nakamura, K.; Casey, M.; Oey, H.; Vari, F.; Stagg, J.; Gandhi, M.K.; Smyth, M.J. Targeting an Adenosine-Mediated “Don’t Eat Me Signal” Augments Anti-Lymphoma Immunity by Anti-CD20 Monoclonal Antibody. Leukemia 2020, 34, 2708–2721. [Google Scholar] [CrossRef]
  177. Allard, B.; Cousineau, I.; Allard, D.; Buisseret, L.; Pommey, S.; Chrobak, P.; Stagg, J. Adenosine A2a Receptor Promotes Lymphangiogenesis and Lymph Node Metastasis. Oncoimmunology 2019, 8, 1601481. [Google Scholar] [CrossRef]
  178. Sidders, B.; Zhang, P.; Goodwin, K.; O’Connor, G.; Russell, D.L.; Borodovsky, A.; Armenia, J.; McEwen, R.; Linghu, B.; Bendell, J.C.; et al. Adenosine Signaling Is Prognostic for Cancer Outcome and Has Predictive Utility for Immunotherapeutic Response. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2020, 26, 2176–2187. [Google Scholar] [CrossRef]
  179. Kamai, T.; Kijima, T.; Tsuzuki, T.; Nukui, A.; Abe, H.; Arai, K.; Yoshida, K.-I. Increased Expression of Adenosine 2A Receptors in Metastatic Renal Cell Carcinoma Is Associated with Poorer Response to Anti-Vascular Endothelial Growth Factor Agents and Anti-PD-1/Anti-CTLA4 Antibodies and Shorter Survival. Cancer Immunol. Immunother. 2021, 70, 2009–2021. [Google Scholar] [CrossRef]
  180. Bai, Y.; Zhang, X.; Zheng, J.; Liu, Z.; Yang, Z.; Zhang, X. Overcoming High Level Adenosine-Mediated Immunosuppression by DZD2269, a Potent and Selective A2aR Antagonist. J. Exp. Clin. Cancer Res. 2022, 41, 302. [Google Scholar] [CrossRef]
  181. 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-(L)1 and Anti–CTLA-4 in Preclinical Models. Cancer Immunol. Res. 2018, 6, 1136–1149. [Google Scholar] [CrossRef] [PubMed]
  182. Fong, L.; Hotson, A.; Powderly, J.D.; Sznol, M.; Heist, R.S.; Choueiri, T.K.; George, S.; Hughes, B.G.M.; Hellmann, M.D.; Shepard, D.R.; et al. Adenosine A2A Receptor Blockade as an Immunotherapy for Treatment-Refractory Renal Cell Cancer. Cancer Discov. 2020, 10, 40–53. [Google Scholar] [CrossRef] [PubMed]
  183. Lim, E.A.; Bendell, J.C.; Falchook, G.S.; Bauer, T.M.; Drake, C.G.; Choe, J.H.; George, D.J.; Karlix, J.L.; Ulahannan, S.; Sachsenmeier, K.F.; et al. Phase Ia/b, Open-Label, Multicenter Study of AZD4635 (an Adenosine A2A Receptor Antagonist) as Monotherapy or Combined with Durvalumab, in Patients with Solid Tumors. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2022, 28, 4871–4884. [Google Scholar] [CrossRef] [PubMed]
  184. Zhu, C.; Ze, S.; Zhou, R.; Yang, X.; Wang, H.; Chai, X.; Fang, M.; Liu, M.; Wang, Y.; Lu, W.; et al. Discovery of Pyridinone Derivatives as Potent, Selective, and Orally Bioavailable Adenosine A2A Receptor Antagonists for Cancer Immunotherapy. J. Med. Chem. 2023, 66, 4734–4754. [Google Scholar] [CrossRef]
  185. Yu, F.; Zhu, C.; Ze, S.; Wang, H.; Yang, X.; Liu, M.; Xie, Q.; Lu, W.; Wang, Y. Design, Synthesis, and Bioevaluation of 2-Aminopteridin-7(8H)-One Derivatives as Novel Potent Adenosine A2A Receptor Antagonists for Cancer Immunotherapy. J. Med. Chem. 2022, 65, 4367–4386. [Google Scholar] [CrossRef]
  186. Zhang, J.; Luo, Z.; Duan, W.; Yang, K.; Ling, L.; Yan, W.; Liu, R.; Wüthrich, K.; Jiang, H.; Xie, C.; et al. Dual-Acting Antitumor Agents Targeting the A2A Adenosine Receptor and Histone Deacetylases: Design and Synthesis of 4-(Furan-2-Yl)-1H-Pyrazolo[3,4-d]Pyrimidin-6-Amine Derivatives. Eur. J. Med. Chem. 2022, 236, 114326. [Google Scholar] [CrossRef]
  187. Yan, W.; Ling, L.; Wu, Y.; Yang, K.; Liu, R.; Zhang, J.; Zhao, S.; Zhong, G.; Zhao, S.; Jiang, H.; et al. Structure-Based Design of Dual-Acting Compounds Targeting Adenosine A2A Receptor and Histone Deacetylase as Novel Tumor Immunotherapeutic Agents. J. Med. Chem. 2021, 64, 16573–16597. [Google Scholar] [CrossRef]
  188. Varano, F.; Catarzi, D.; Vincenzi, F.; Pasquini, S.; Pelletier, J.; Lopes Rangel Fietto, J.; Espindola Gelsleichter, N.; Sarlandie, M.; Guilbaud, A.; Sévigny, J.; et al. Structural Investigation on Thiazolo[5,4-d]Pyrimidines to Obtain Dual-Acting Blockers of CD73 and Adenosine A2A Receptor as Potential Antitumor Agents. Bioorg. Med. Chem. Lett. 2020, 30, 127067. [Google Scholar] [CrossRef]
  189. Yu, W.; Sun, J.; Wang, X.; Yu, S.; Yan, M.; Wang, F.; Liu, X. Boosting Cancer Immunotherapy via the Convenient A2AR Inhibition Using a Tunable Nanocatalyst with Light-Enhanced Activity. Adv. Mater. Deerfield Beach Fla 2022, 34, e2106967. [Google Scholar] [CrossRef]
  190. Wen, X.; Xiong, X.; Yang, G.; Xiao, W.; Hou, J.; Pan, T.; Hu, Y.; Zhou, S. A Macrophage Membrane-Coated Mesoporous Silica Nanoplatform Inhibiting Adenosine A2AR via in Situ Oxygen Supply for Immunotherapy. J. Control. Release 2023, 353, 535–548. [Google Scholar] [CrossRef]
  191. Liu, Y.; Liu, Y.; Xu, D.; Zang, J.; Zheng, X.; Zhao, Y.; Li, Y.; He, R.; Ruan, S.; Dong, H.; et al. Targeting the Negative Feedback of Adenosine-A2AR Metabolic Pathway by a Tailored Nanoinhibitor for Photothermal Immunotherapy. Adv. Sci. 2022, 9, 2104182. [Google Scholar] [CrossRef]
  192. Labanieh, L.; Mackall, C.L. CAR Immune Cells: Design Principles, Resistance and the next Generation. Nature 2023, 614, 635–648. [Google Scholar] [CrossRef]
  193. Li, N.; Tang, N.; Cheng, C.; Hu, T.; Wei, X.; Han, W.; Wang, H. Improving the Anti-Solid Tumor Efficacy of CAR-T Cells by Inhibiting Adenosine Signaling Pathway. Oncoimmunology 2020, 9, 1824643. [Google Scholar] [CrossRef] [PubMed]
  194. Giuffrida, L.; Sek, K.; Henderson, M.A.; Lai, J.; Chen, A.X.Y.; Meyran, D.; Todd, K.L.; Petley, E.V.; Mardiana, S.; Mølck, C.; et al. CRISPR/Cas9 Mediated Deletion of the Adenosine A2A Receptor Enhances CAR T Cell Efficacy. Nat. Commun. 2021, 12, 3236. [Google Scholar] [CrossRef] [PubMed]
  195. Masoumi, E.; Jafarzadeh, L.; Mirzaei, H.R.; Alishah, K.; Fallah-Mehrjardi, K.; Rostamian, H.; Khakpoor-Koosheh, M.; Meshkani, R.; Noorbakhsh, F.; Hadjati, J. Genetic and Pharmacological Targeting of A2a Receptor Improves Function of Anti-Mesothelin CAR T Cells. J. Exp. Clin. Cancer Res. 2020, 39, 49. [Google Scholar] [CrossRef] [PubMed]
  196. Seifert, M.; Benmebarek, M.-R.; Briukhovetska, D.; Märkl, F.; Dörr, J.; Cadilha, B.L.; Jobst, J.; Stock, S.; Andreu-Sanz, D.; Lorenzini, T.; et al. Impact of the Selective A2AR and A2BR Dual Antagonist AB928/Etrumadenant on CAR T Cell Function. Br. J. Cancer 2022, 127, 2175–2185. [Google Scholar] [CrossRef]
  197. Vecchio, E.A.; White, P.J.; May, L.T. The Adenosine A2B G Protein-Coupled Receptor: Recent Advances and Therapeutic Implications. Pharmacol. Ther. 2019, 198, 20–33. [Google Scholar] [CrossRef]
  198. Chen, S.; Akdemir, I.; Fan, J.; Linden, J.; Zhang, B.; Cekic, C. The Expression of Adenosine A2B Receptor on Antigen Presenting Cells Suppresses CD8+ T Cell Responses and Promotes Tumor Growth. Cancer Immunol. Res. 2020, 8, 1064–1074. [Google Scholar] [CrossRef]
  199. Lan, J.; Lu, H.; Samanta, D.; Salman, S.; Lu, Y.; Semenza, G.L. Hypoxia-Inducible Factor 1-Dependent Expression of Adenosine Receptor 2B Promotes Breast Cancer Stem Cell Enrichment. Proc. Natl. Acad. Sci. USA 2018, 115, E9640–E9648. [Google Scholar] [CrossRef]
  200. Lan, J.; Wei, G.; Liu, J.; Yang, F.; Sun, R.; Lu, H. Chemotherapy-Induced Adenosine A2B Receptor Expression Mediates Epigenetic Regulation of Pluripotency Factors and Promotes Breast Cancer Stemness. Theranostics 2022, 12, 2598–2612. [Google Scholar] [CrossRef]
  201. Tay, A.H.M.; Prieto-Díaz, R.; Neo, S.; Tong, L.; Chen, X.; Carannante, V.; Önfelt, B.; Hartman, J.; Haglund, F.; Majellaro, M.; et al. A2B Adenosine Receptor Antagonists Rescue Lymphocyte Activity in Adenosine-Producing Patient-Derived Cancer Models. J. Immunother. Cancer 2022, 10, e004592. [Google Scholar] [CrossRef]
  202. Yu, M.; Guo, G.; Huang, L.; Deng, L.; Chang, C.-S.; Achyut, B.R.; Canning, M.; Xu, N.; Arbab, A.S.; Bollag, R.J.; et al. CD73 on Cancer-Associated Fibroblasts Enhanced by the A2B-Mediated Feedforward Circuit Enforces an Immune Checkpoint. Nat. Commun. 2020, 11, 515. [Google Scholar] [CrossRef]
  203. Faraoni, E.Y.; Singh, K.; Chandra, V.; Le Roux, O.; Dai, Y.; Sahin, I.; O’Brien, B.J.; Strickland, L.N.; Li, L.; Vucic, E.; et al. CD73-Dependent Adenosine Signaling through Adora2b Drives Immunosuppression in Ductal Pancreatic Cancer. Cancer Res. 2023, 83, 1111–1127. [Google Scholar] [CrossRef]
  204. Ludwig, N.; Yerneni, S.S.; Azambuja, J.H.; Gillespie, D.G.; Menshikova, E.V.; Jackson, E.K.; Whiteside, T.L. Tumor-Derived Exosomes Promote Angiogenesis via Adenosine A2B Receptor Signaling. Angiogenesis 2020, 23, 599–610. [Google Scholar] [CrossRef]
  205. Zhou, Y.; Tong, L.; Chu, X.; Deng, F.; Tang, J.; Tang, Y.; Dai, Y. The Adenosine A1 Receptor Antagonist DPCPX Inhibits Tumor Progression via the ERK/JNK Pathway in Renal Cell Carcinoma. Cell. Physiol. Biochem. 2017, 43, 733–742. [Google Scholar] [CrossRef] [PubMed]
  206. Mirza, A.; Basso, A.; Black, S.; Malkowski, M.; Kwee, L.; Patcher, J.A.; Lachowicz, J.E.; Wang, Y.; Liu, S. RNA Interference Targeting of A1 Receptor-Overexpressing Breast Carcinoma Cells Leads to Diminished Rates of Cell Proliferation and Induction of Apoptosis. Cancer Biol. Ther. 2005, 4, 1355–1360. [Google Scholar] [CrossRef] [PubMed]
  207. Liu, H.; Kuang, X.; Zhang, Y.; Ye, Y.; Li, J.; Liang, L.; Xie, Z.; Weng, L.; Guo, J.; Li, H.; et al. ADORA1 Inhibition Promotes Tumor Immune Evasion by Regulating the ATF3-PD-L1 Axis. Cancer Cell 2020, 37, 324–339.e8. [Google Scholar] [CrossRef]
  208. Le, X.; Negrao, M.V.; Reuben, A.; Federico, L.; Diao, L.; McGrail, D.; Nilsson, M.; Robichaux, J.; Munoz, I.G.; Patel, S.; et al. Characterization of the Immune Landscape of EGFR-Mutant NSCLC Identifies CD73/Adenosine Pathway as a Potential Therapeutic Target. J. Thorac. Oncol. 2021, 16, 583–600. [Google Scholar] [CrossRef] [PubMed]
  209. Pang, L.; Ng, K.T.-P.; Liu, J.; Yeung, W.-H.O.; Zhu, J.; Chiu, T.-L.S.; Liu, H.; Chen, Z.; Lo, C.-M.; Man, K. Plasmacytoid Dendritic Cells Recruited by HIF-1α/EADO/ADORA1 Signaling Induce Immunosuppression in Hepatocellular Carcinoma. Cancer Lett. 2021, 522, 80–92. [Google Scholar] [CrossRef]
  210. Merighi, S.; Battistello, E.; Giacomelli, L.; Varani, K.; Vincenzi, F.; Borea, P.A.; Gessi, S. Targeting A3 and A2A Adenosine Receptors in the Fight against Cancer. Expert Opin. Ther. Targets 2019, 23, 669–678. [Google Scholar] [CrossRef]
  211. Bi, J.; Zheng, C.; Zheng, X. Increased Expression of Adenosine A3 Receptor in Tumor-Infiltrating Natural Killer Cells. Cell. Mol. Immunol. 2021, 18, 496–497. [Google Scholar] [CrossRef] [PubMed]
  212. Stemmer, S.M.; Manojlovic, N.S.; Marinca, M.V.; Petrov, P.; Cherciu, N.; Ganea, D.; Ciuleanu, T.E.; Pusca, I.A.; Beg, M.S.; Purcell, W.T.; et al. Namodenoson in Advanced Hepatocellular Carcinoma and Child–Pugh B Cirrhosis: Randomized Placebo-Controlled Clinical Trial. Cancers 2021, 13, 187. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of adenosine metabolism and AR intracellular signaling. Extracellularly, adenosine (ADO) is primarily derived from adenosine monophosphate (AMP) via the catalytic action of ecto-5′-nucleotidase (CD73). Multiple enzymatic pathways, including nucleoside triphosphate diphosphohydrolase-1 (NTPDase1 or CD39), NTPDase2 and NTPDase3, nucleotide pyrophosphatase/phosphodiesterase-1 (NPP1), and adenylate kinase-1 (AK1), contribute to the generation of AMP from ATP. Adenosine transport across the cell membrane is facilitated by equilibrative nucleoside transporters (ENTs). Intracellularly, AMP is converted to adenosine by cytosolic 5′-nucleotidase (5′NT), and the reverse reaction is mediated by adenosine kinase (ADK). Additionally, ATP to ADP and ADP to AMP interconversions are catalyzed by AK-1 and nucleotide diphosphokinase (NDPK), respectively. Adenosine can also be generated from S-adenosylhomocysteine (SAH) through the enzymatic action of SAH hydrolase (SAHH). Its enzymatic degradation occurs via adenosine deaminase (ADA), which converts adenosine to inosine (INO). On the signaling front, A2A and A2B ARs activate adenylyl cyclase (AC) through Gs protein coupling, stimulating the conversion of ATP to cyclic AMP (cAMP). Conversely, A1 and A3 ARs inhibit AC via Gi protein coupling. The primary downstream effectors of cAMP are cAMP-dependent protein kinase (PKA) and exchange protein directly activated by cAMP (EPAC). PKA primarily regulates transcription through phosphorylation of the cAMP response element-binding protein (CREB). Protein kinase B (PKB) and mitogen-activated protein kinases (MAPK) are common substrates for PKA. Additionally, A2B and A3 ARs can activate Gq proteins, thereby stimulating phospholipase C (PLC) and subsequently leading to the formation of diacylglycerol (DAG) and inositol triphosphate (IP3). These molecules activate protein kinase C and elevate intracellular Ca2+ levels, respectively. A3ARs also regulate the phosphatidylinositol 3-kinase (PI3K)/PKB axis via Gq proteins. A1AR activation modulates ion channels, thus inhibiting Ca2+ channels and activating K+ channels.
Figure 1. Overview of adenosine metabolism and AR intracellular signaling. Extracellularly, adenosine (ADO) is primarily derived from adenosine monophosphate (AMP) via the catalytic action of ecto-5′-nucleotidase (CD73). Multiple enzymatic pathways, including nucleoside triphosphate diphosphohydrolase-1 (NTPDase1 or CD39), NTPDase2 and NTPDase3, nucleotide pyrophosphatase/phosphodiesterase-1 (NPP1), and adenylate kinase-1 (AK1), contribute to the generation of AMP from ATP. Adenosine transport across the cell membrane is facilitated by equilibrative nucleoside transporters (ENTs). Intracellularly, AMP is converted to adenosine by cytosolic 5′-nucleotidase (5′NT), and the reverse reaction is mediated by adenosine kinase (ADK). Additionally, ATP to ADP and ADP to AMP interconversions are catalyzed by AK-1 and nucleotide diphosphokinase (NDPK), respectively. Adenosine can also be generated from S-adenosylhomocysteine (SAH) through the enzymatic action of SAH hydrolase (SAHH). Its enzymatic degradation occurs via adenosine deaminase (ADA), which converts adenosine to inosine (INO). On the signaling front, A2A and A2B ARs activate adenylyl cyclase (AC) through Gs protein coupling, stimulating the conversion of ATP to cyclic AMP (cAMP). Conversely, A1 and A3 ARs inhibit AC via Gi protein coupling. The primary downstream effectors of cAMP are cAMP-dependent protein kinase (PKA) and exchange protein directly activated by cAMP (EPAC). PKA primarily regulates transcription through phosphorylation of the cAMP response element-binding protein (CREB). Protein kinase B (PKB) and mitogen-activated protein kinases (MAPK) are common substrates for PKA. Additionally, A2B and A3 ARs can activate Gq proteins, thereby stimulating phospholipase C (PLC) and subsequently leading to the formation of diacylglycerol (DAG) and inositol triphosphate (IP3). These molecules activate protein kinase C and elevate intracellular Ca2+ levels, respectively. A3ARs also regulate the phosphatidylinositol 3-kinase (PI3K)/PKB axis via Gq proteins. A1AR activation modulates ion channels, thus inhibiting Ca2+ channels and activating K+ channels.
Biomolecules 13 01387 g001
Figure 2. Therapeutic potential of adenosine receptors. The physiological effects of the receptors are in light blue, the pharmacological actions of ligands suitable for treatment are in light green, and the pathologies in which the ligands can be utilized are in pink.
Figure 2. Therapeutic potential of adenosine receptors. The physiological effects of the receptors are in light blue, the pharmacological actions of ligands suitable for treatment are in light green, and the pathologies in which the ligands can be utilized are in pink.
Biomolecules 13 01387 g002
Table 1. The latest developments concerning adenosine receptor ligands in animal models of CNS pathologies.
Table 1. The latest developments concerning adenosine receptor ligands in animal models of CNS pathologies.
Pharmacological ActionCompoundExperimental ModelSpeciesEffects
A1AR agonistsbenzyloxy-cyclopentyladenosineNeuropathic painratAnalgesia [16]
ENBACerebral ischemiaratMicroglia/macrophages proliferation reduction [66]
Dual A1/A3 AR agonistsAST-004Cerebral ischemianonhuman primateIschemic damage reduction [71]
A1AR antagonistsDPCPXEpilepsyratReversion of DBS impact on interictal epileptic discharges [78]
A1AR PAMsMIPS521Neuropathic painratAnalgesia [14]
A2AAR agonistsATL313Neuropathic painratAnti-allodynia and anti-inflammation [18]
A2AAR antagonistsKW-6002 (istradefylline)EAE modelmouseProtection against T Cell Infiltration [58]
Chemotherapy-induced cognitive impairmentmouseReversion of cisplatin-induced neurotoxicity [72]
Retinal injurymouseReduction of apoptosis of retinal ganglia cells [101]
mouseRegulation of retinal ganglion cell morphology [102]
KW-6356PDmarmosetMotor disability reversion [45]
SCH58261SUDEPmouseIncrease in theta and beta oscillations [79]
Mania-like behaviorratReduction of locomotor hyperactivity and risk-taking behavior [90]
ZM241385Craniocerebral traumamouseReduction of fear memory [62]
A2AAR PAMsA2ARPAM-1Mania-like behaviormouseInsomnia reduction [95]
A2BAR agonistsBAY60-6583Focal ischemiaratReduced brain damage [69]
A3AR agonistsCl-IB-MECAGlaucomaratPreservation of retinal ganglion cell [99]
IB-MECAChronic cerebral ischemiamouseImproved memory deficits [70]
MRS5980Traumatic brain injurymouseReduction of tissue injury and cognitive impairment [63]
Chemotherapy-induced cognitive impairmentmousePrevention of mitochondrial dysfunction and oxidative stress [73]
Neuropathic painmouseAnti-allodynia [23]
Non-selective antagonistsCaffeineADmouseNeuroprotection [41]
AD (Alzheimer’s disease), EAE (autoimmune encephalomyelitis), PAM (positive allosteric modulator), PD (Parkinson’s disease), SUDEP (sudden unexpected death in epilepsy).
Table 2. The latest developments concerning adenosine receptor ligands in animal models of cardiovascular and metabolic diseases.
Table 2. The latest developments concerning adenosine receptor ligands in animal models of cardiovascular and metabolic diseases.
Pharmacological ActionCompoundExperimental ModelSpeciesEffects
A2AAR agonistsATL1223Circulatory arrest and ECPRpigReduction of systemic ischemia-reperfusion injury and ameliorated renal, hepatic, and cardiac injury [108]
A2AAR antagonistsSCH58261High-sucrose dietratImproved insulin response [120]
A2BAR agonistsBAY 60–6583ObesitymouseIncreased whole body energy expenditure [118]
A2BAR antagonistsMRS1754High-sucrose dietratImproved insulin response [120]
ECPR (extracorporeal cardiopulmonary resuscitation).
Table 3. The latest developments concerning adenosine receptor ligands in animal models of inflammatory and autoimmune pathologies.
Table 3. The latest developments concerning adenosine receptor ligands in animal models of inflammatory and autoimmune pathologies.
Pharmacological ActionCompoundExperimental ModelSpeciesEffects
A2AAR agonistsCGS21680ArthritismouseArthritis development inhibition [132]
LESmouseDepletion of pathogenic lymphocytes and anti-nuclear antibodies reduction [135]
Obesity-induced OAmouseSenescence markers reduction and improvement of mitochondrial stability and function [141,142]
Increased autophagy and improved metabolic function in chondrocytes [143]
PDRNAcute lung injuryratDecrease in lung tissue damage, pro-inflammatory cytokines, apoptotic factors, and MAPK/NF-κB activation [152]
A2AAR antagonistsZM241385SepsismouseEnhanced survival by improving bacterial clearance [164]
A3AR agonistsCF101OAratReduction of cartilage damage and pain [144]
MRS5980Lung fibrosismouseAttenuation of lung stiffness and inflammatory/oxidative stress markers [161]
MRS7344PsoriasismouseHinder development of psoriatic-like characteristics [126]
A3AR/partial PPARγ agonistLJ-529Pulmonary EmphysemamousePulmonary function restoration, reduction of MMP activity and apoptosis [160]
Non-selective agonistsNECAAirway inflammationmouseReduction of innate lymphoid cells responses and inflammation [151]
Non-selective antagonistsCaffeineBronchopulmonary dysplasiamouseDecrease in lung injury, oxidative stress, and inflammatory infiltration [162]
LES (lupus erythematosus systemic), OA (osteoarthritis), PPARγ (peroxisome proliferator activated receptor γ).
Table 4. The latest developments concerning adenosine receptor ligands in animal models of cancer.
Table 4. The latest developments concerning adenosine receptor ligands in animal models of cancer.
Pharmacological ActionCompoundExperimental ModelSpecies985Effects
A1AR antagonistsDPCPXHCCmouseSuppression of tumor growth [209]
A2AAR antagonistsAZD4635Melanoma, colorectal carcinoma, fibrosarcomamouseDecrease in tumor volume with enhanced T cell response [175]
CPI-444 (ciforadenant)Colorectal carcinoma, kidney tumor, melanomamouseDiminution of tumor growth intensifying the effectiveness of checkpoint inhibitors [181]
DZD2269Melanoma, prostate cancer, and pancreatic cancermouseAntitumor effects particularly when used together with immune checkpoint inhibitors, radiotherapy, or chemotherapy [180]
SCH58261CLLmouseImmune competence restoration by inhibiting the accumulation and differentiation of Treg cells [173]
Breast cancermouseRegression of primary tumor and inhibition of metastasis [191]
ZM241385MelanomamouseRestoration of the functionality of tumor-infiltrating cytotoxic T cells [171]
Dual A2A/A2B AR antagonistsAB928 (etrumadenant)Colon carcinomamouseEnhancement of CAR T cell activation [196]
A2BAR antagonistsAlloxazineBreast cancermouseDelay of tumor recurrence after discontinuation of chemotherapy [200]
CLL (chronic lymphocytic leukemia), HCC (hepatocellular carcinoma).
Table 5. Clinical trials involving adenosine receptor ligands that started in the last five years.
Table 5. Clinical trials involving adenosine receptor ligands that started in the last five years.
Pharmacological ActionCompoundConditionPhaseNTC Number
A1 antagonistsPBF-680AsthmaII03774290
COPDIIa05262218
A2A agonistsRegadenosonCOVID-19I/II04606069
Myocardial ischemiaI/II04604782
High grade gliomasI03971734
Lung transplantI04521569
A2A antagonistsIstradefyllineALSI/II05377424
PDIV05885360
Cognitive impairment in PDII05333549
Apathy in PDObservational05182151
Ciforadenant
(CPI-444)
Renal cell carcinomaIb/II05501054
Multiple myelomaI04280328
TT-10Solid cancersI/II04969315
Inupadenant
(EOS100850)
NSCLCII05403385
Solid cancersI05117177
AZD4635Solid cancersI03980821
NSCLCI/II03381274
Prostate cancerII04089553
CRPCII04495179
DZD2269CRPCI04634344
Dual A2A/A2B antagonistsEtrumadenant
(AB928)
Head and neck cancersI04892875
Prostate cancerII05915442
LiposarcomaII05886634
Rectal cancerII05024097
Urothelial carcinomaII05335941
Gastrointestinal cancersI03720678
Colorectal cancerI/II04660812
CRPCIb/II04381832
CRPCII05177770
M1069Solid cancersI05198349
A2B antagonistsPBF-1129NSCLCI03274479
NSCLCI05234307
TT-702
(prodrug of TT-478)
Solid cancersI/II05272709
TT-4Solid cancersI/II04976660
A3 agonistsPiclidenoson
(CF101, IB-MECA)
Ocular hypertensionI/II04585100
COVID-19II04333472
Plaque psoriasisIII03168256
Namodenoson
(CF102, Cl-IB-MECA)
NASHII04697810
HCC/CirrhosisIII05201404
FM-101Ocular hypertensionI/II04585100
NASHII04710524
A3 antagonistsPBF-1650PsoriasisI03798236
PBF-677Ulcerative colitisII03773952
Non-selective antagonistsCaffeineHypoxic-ischemic encephalopathyI03913221
Data from clinicaltrials.gov; ALS (amyotrophic lateral sclerosis), COPD (chronic obstructive pulmonary disease), CRPC (castrate resistant prostate cancer), HCC (hepatocellular carcinoma), NASH (nonalcoholic steatohepatitis), NSCLC (non-small cell lung cancer), PD (Parkinson’s disease).
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vincenzi, F.; Pasquini, S.; Contri, C.; Cappello, M.; Nigro, M.; Travagli, A.; Merighi, S.; Gessi, S.; Borea, P.A.; Varani, K. Pharmacology of Adenosine Receptors: Recent Advancements. Biomolecules 2023, 13, 1387. https://doi.org/10.3390/biom13091387

AMA Style

Vincenzi F, Pasquini S, Contri C, Cappello M, Nigro M, Travagli A, Merighi S, Gessi S, Borea PA, Varani K. Pharmacology of Adenosine Receptors: Recent Advancements. Biomolecules. 2023; 13(9):1387. https://doi.org/10.3390/biom13091387

Chicago/Turabian Style

Vincenzi, Fabrizio, Silvia Pasquini, Chiara Contri, Martina Cappello, Manuela Nigro, Alessia Travagli, Stefania Merighi, Stefania Gessi, Pier Andrea Borea, and Katia Varani. 2023. "Pharmacology of Adenosine Receptors: Recent Advancements" Biomolecules 13, no. 9: 1387. https://doi.org/10.3390/biom13091387

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

Article Metrics

Back to TopTop