Next Article in Journal
PNPLA3 I148M Up-Regulates Hedgehog and Yap Signaling in Human Hepatic Stellate Cells
Next Article in Special Issue
A2B Adenosine Receptors: When Outsiders May Become an Attractive Target to Treat Brain Ischemia or Demyelination
Previous Article in Journal
Identification of Anthocyanins-Related Glutathione S-Transferase (GST) Genes in the Genome of Cultivated Strawberry (Fragaria × ananassa)
Previous Article in Special Issue
P2Y2 and P2X4 Receptors Mediate Ca2+ Mobilization in DH82 Canine Macrophage Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Adenosine Receptors: A Potential Pharmacological Avenue for Acute and Chronic Pain

1
Department of Translational Medicine and for Romagna, Pharmacology Section, University of Ferrara, 44121 Ferrara, Italy
2
University of Ferrara, 44121 Ferrara, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(22), 8710; https://doi.org/10.3390/ijms21228710
Submission received: 12 October 2020 / Accepted: 17 November 2020 / Published: 18 November 2020
(This article belongs to the Special Issue Purinergic Signaling in Neuroinflammation)

Abstract

:
Adenosine is a purine nucleoside, responsible for the regulation of multiple physiological and pathological cellular and tissue functions by activation of four G protein-coupled receptors (GPCR), namely A1, A2A, A2B, and A3 adenosine receptors (ARs). In recent years, extensive progress has been made to elucidate the role of adenosine in pain regulation. Most of the antinociceptive effects of adenosine are dependent upon A1AR activation located at peripheral, spinal, and supraspinal sites. The role of A2AAR and A2BAR is more controversial since their activation has both pro- and anti-nociceptive effects. A3AR agonists are emerging as promising candidates for neuropathic pain. Although their therapeutic potential has been demonstrated in diverse preclinical studies, no AR ligands have so far reached the market. To date, novel pharmacological approaches such as adenosine regulating agents and allosteric modulators have been proposed to improve efficacy and limit side effects enhancing the effect of endogenous adenosine. This review aims to provide an overview of the therapeutic potential of ligands interacting with ARs and the adenosinergic system for the treatment of acute and chronic pain.

1. Introduction

Today, although substantial progress has been made, many pathological pain conditions remain poorly understood and resist currently available treatments. There is, therefore, a need for novel molecular targets to develop new therapeutic agents with improved efficacy and tolerability. Many experimental reports have identified adenosine receptors (ARs) as potential targets for the management of acute and chronic pain.
Adenosine is a ubiquitous endogenous autacoid that mediates its physiopathological effects by interacting with four G protein-coupled receptors (GPCR), namely A1, A2A, A2B, and A3 ARs [1]. A1 and A3AR are coupled with Gi and Go members of the G protein family, through which they have an inhibitory effect on adenylyl cyclase (AC) activity, while A2AARs and A2BARs stimulate it by coupling to Gs proteins. The consequent modulation of cyclic adenosine monophosphate (cAMP) levels activates or inhibits a large variety of signaling pathways depending on the specific type of cell involved. Although there are instances in which adenosine exerts detrimental effects in various pathological conditions, it is generally considered a protective and homeostatic mediator against tissue damages and stress conditions [2,3]. In physiological and unstressed conditions, the extracellular concentrations of adenosine are maintained low as a result of the rapid metabolism and uptake [4]. However, its levels rise considerably during conditions involving increased metabolic demand, hypoxia, inflammation, and tissue injury. In particular, increased levels of extracellular adenosine were observed in pathological conditions such as epilepsy [5,6], ischemia [7,8], cancer [9,10], inflammation [11], and ultimately pain [12,13]. Although adenosine can be produced intracellularly, the main source of adenosine in pathological states is adenosine triphosphate (ATP), released by cells under stressful conditions and dephosphorylated from the combined action of two hydrolyzing enzymes termed ectonucleoside triphosphate diphosphohydrolase (CD39) and ecto-5′-nucleotidase (CD73) [1]. Regarding nociception, these elevated levels of endogenous adenosine can alter pain transmission by actions at spinal, supraspinal, and peripheral sites. The extracellular action of adenosine can then be terminated by its transformation to inosine through adenosine deaminase (ADA) and/or by intracellular uptake via nucleoside transporters [14]. Intracellularly, adenosine is phosphorylated to AMP by adenosine kinase or deaminated to inosine by ADA. Given these regulation mechanisms of adenosine concentration, potential pain management can be obtained not only with specific ligand interacting with ARs but also by manipulating endogenous tissue levels of adenosine by modulating its metabolism or transport [13] (Figure 1).
Although adenosine and its receptors represent a clear target for pharmacological treatment of various diseases and pathological states including pain, very few drugs acting on the adenosinergic system have so far reached the market. The reason behind this discrepancy may be partly due to the ubiquitous distribution of ARs in almost every cell and tissue, making it difficult to avoid unwanted side effects. In recent years, many efforts have been made to improve our understanding of the role of adenosine in nociception and identify novel strategies to exploit the therapeutic potential of the adenosinergic system such as selective ligands, partial agonists, allosteric modulators, or adenosine concentration modulating agents.
The focus of the present review is to describe the recent advances in our understanding of the role of ARs in nociception. For each receptor subtype, we will briefly summarize and discuss the preclinical experimental studies that investigated their role and mechanism of action in the modulation of acute and chronic pain.

2. ARs and Pain

2.1. A1ARs

The antinociceptive effect of adenosine has been primarily attributed to the activation of A1ARs [15] and various A1AR agonists or positive allosteric modulators have been shown to be effective in several preclinical models of pain (Table 1). The signaling pathway underlying A1ARs antinociception includes inhibition of cyclic AMP and consequently protein kinase A (PKA) activation, inhibition of Ca2+ channels, activation of K+ currents, and interactions with phospholipase C (PLC), inositol triphosphate (IP3), diacylglycerol (DAG), extracellular signal-regulated kinases (ERK), and β-arrestin pathways [3]. The prominent role of this receptor subtype in analgesic responses is due to its peculiar expression in different sites relevant to pain transmission. A1ARs are indeed located on peripheral sensory nerve endings in the spinal cord dorsal horn, and at supraspinal pain-processing structures [13,16]. Microglia represent another important localization for the antinociceptive action of A1ARs, especially for pain states involving glial activation [17]. The peripheral activation of A1ARs diminished inflammatory hypernociception caused by carrageenan intraplantar administration. Using specific inhibitors, the antinociceptive effect of the A1AR agonist CPA was shown to be dependent on the nitric oxide (NO)/cyclic guanosine monophosphate (cGMP)/protein kinase G (PKG)/KATP signaling pathway [18]. The contribution of peripheral A1ARs to antinociception was further corroborated when the selective A1AR antagonist DPCPX reversed the antinociceptive effects of locally and systemically administered acetaminophen or tramadol in the formalin test [19]. A proof of the supraspinal antinociceptive action of A1AR has been reported in a study where the A1AR agonist 2′-Me-CCPA injected into the intra-periaqueductal grey (PAG) reduced pain behavior in the plantar and formalin tests. When microinjected into the PAG, 2′-Me-CCPA decreased the ongoing activity of the pronociceptive ON cells and increased the ongoing activity of the antinociceptive OFF cell in the rostral ventromedial medulla [20]. In neuropathic pain rats, the A1AR agonist CPA reduced thermal and mechanical sensitivity, while in naïve rats it decreased hypersensitivity to heat but not to mechanical stimuli. In this study, electrophysiological experiments suggested that spinal application of CPA depressed long-term potentiation of A- and C-fiber evoked field potentials while it depressed the baseline of C-fiber but not A-fiber response. To explain this different response, authors have hypothesized that A1ARs may be more expressed at C-fiber nerve endings than at A-fiber endings. [21]. In resiniferatoxin-induced neuropathy, the downregulation of A1ARs was suggested to contribute to nociception, while the intrathecal injection of adenosine attenuated mechanical allodynia, an effect abrogated by A1AR antagonism [22].
A1ARs also seem to be involved in visceral antinociception. Centrally injected agonist CPA increased the threshold volume of colonic distension-induced abdominal withdrawal reflex in conscious rats. Besides, the use of the A1 antagonist DPCPX suggested that adenosinergic signaling via A1ARs is also involved in the central orexin-induced antinociceptive action against colonic distension [23]. In a subsequent study, the authors suggested that serotonin 5-HT1A, 5-HT2A, dopamine D1 or cannabinoid CB1 receptors, and the opioid system might specifically mediate the CPA-induced visceral antinociception [24].
The potential role of A1ARs in postoperative pain was also investigated. Intrathecal administration of the A1AR agonist R-PIA decreased nonevoked spontaneous pain behavior and increased withdrawal thresholds after plantar incision. The opening of KATP channels contributed to this antinociceptive effect [25]. In a mouse model of acute postoperative pain, ankle joint mobilization decreased hyperalgesia through the involvement of peripheral and central A1ARs [26]. In another report, intrathecal adenosine injection inhibited hyperalgesia in two neuropathic pain models but not in a postoperative pain model represented by the plantar incision. However, in this model A1AR mRNA and protein expression were decreased suggesting that the lack of antinociceptive effect of adenosine on postoperative pain was due to the decrease in A1ARs [27].
An intriguing connection has been uncovered between A1ARs and acupuncture, an invasive practice worldwide used to relieve pain. Many studies report that the antinociceptive effects of acupuncture are dependent upon A1AR activation. It was shown that extracellular adenosine concentration is increased during acupuncture in mice and A1AR expression is required for the adenosine-mediated analgesic effect of acupuncture [28]. The involvement of A1ARs in the reduction in neuropathic pain exerted by electroacupuncture was demonstrated by the intrathecal injection of the A1AR antagonist DPCPX in a chronic constriction injury (CCI) model. In this report, the effect of A1ARs was related to the inhibition of astrocyte activation. [29]. Similar results were obtained in a Complete Freund’s adjuvant (CFA)-induced inflammatory pain mouse model, corroborating the involvement of A1ARs in electroacupuncture-mediated antinociception [30]. In another study, the analgesic effect of electroacupuncture was suggested to be mediated by overexpressed A1ARs in the spinal cord [31].
Different studies suggested that A1AR activation is required for the antinociceptive action of various natural compounds. Indeed, the A1AR antagonist DPXPC blocked the effect of norisoboldine, a benzylisoquinoline alkaloid isolated from Radix Linderae that diminishes pain response, in the formalin and writhing test [32]. In addition, A1AR is necessary to the analgesic effect of paeoniflorin, the major active component extracted from Paeonia lactiflora. In a study carried out in mice, paeoniflorin increased the mechanical threshold and prolonged the thermal latency after partial sciatic nerve ligation (SCNL), an effect abolished by the A1AR antagonist CPT or the genetic deletion of A1ARs [33]. In the hot plate test, the antinociceptive effect of (–)-linalool, a natural occurring enantiomer in essential oils, was blocked by both an A1 and an A2AAR antagonist [34]. D-Fructose-1,6-bisphosphate is an intermediate in the glycolytic pathway, inhibiting hyperalgesia induced by intraplantar injection of carrageenin and its mechanism of action seems dependent on adenosine accumulation that in turns exerts antinociceptive effects by activating peripheral A1ARs [35].
Adenosine is rapidly metabolized to inosine by ADA. Interestingly, different studies have identified inosine as a putative endogenous ligand of A1ARs and demonstrated the A1-mediated antinociceptive effect of the more stable metabolite of adenosine. In particular, inosine binds to A1ARs with an affinity resembling that of adenosine and induces antinociceptive, antiallodynic, and antihyperalgesic effects. In rats, both the A1AR antagonist DPCPX and the A2AAR antagonist ZM241385 reversed the antiallodynic, and antihyperalgesic effects of inosine in models of mechanical and heat hyperalgesia induced by bradykinin and phorbol 12-myristate 13-acetate [36]. In the formalin test, inosine did not induce antinociception in A1ARs knockout (KO) mice and the A1AR antagonist DPCPX inhibited its effects [37]. In a subsequent study, DPCPX, but not the A2AAR antagonist SCH58261, abrogated the antinociceptive effect of inosine in the intraplantar glutamate test [38]. A different strategy to promote the accumulation of purines like adenosine or inosine is by using the xanthine oxidase inhibitor, allopurinol. Indeed, it has been reported that intraperitoneal administration of allopurinol increased cerebrospinal fluid concentrations of adenosine and its metabolites inducing antinociceptive effects in different pain models. The selective A1AR antagonist DPCPX, but not the selective A2AAR antagonist SCH58261, prevented allopurinol-induced anti-nociception [39,40]. Since extracellular adenosine is primarily derived from the hydrolysis of AMP, the antinociceptive effect of a soluble version of the recombinant CD73, the enzyme that converts AMP to adenosine, has been tested in different pain models. The results of this study revealed long-lasting thermal antihyperalgesic and mechanical antiallodynic effects that were dependent on A1AR activation [41]. Prostatic acid phosphatase (PAP) acts as an ectonucleotidase hydrolyzing extracellular AMP to adenosine in nociceptive dorsal root ganglia neurons [42,43]. Intrathecal injection of a secretory version of human PAP induced A1AR-dependent antinociceptive effects in inflammatory and neuropathic pain models [44,45]. Furthermore, the injection of PAP into the popliteal fossa—a common acupuncture point—reduces pain responses in mouse models that lasted up to six days after a single injection, an effect dependent upon A1AR activation [46].
Several papers in the literature proposed a link between opioid-mediated antinociception and A1ARs. In a rat with spinal cord injury (SCI), it was demonstrated a supra-additive interaction between the adenosine A1AR agonist R-PIA and morphine in the reduction in mechanical allodynia-like behavior [47]. In spinal cord neuronal nociceptive responses, the antinociceptive effects of the A1AR agonist CPA were associated with activation of κ-opioid receptors since the reversal of the CPA effect was observed with norbinaltorphimine (a selective κ-opioid receptor antagonist) but not with low doses of µ-opioid antagonist naloxone [48]. While the opioid antagonist naltrexone did not affect the antinociception induced by CPA in the formalin test, the activation of A1 or A2AAR counteracted the µ-opioid receptor increase induced by formalin in the spinal cord, confirming the interaction between adenosinergic and opioid systems [49]. In a rat model of nerve ligation injury, the intrathecal administration of morphine synergistically enhanced the antiallodynic effect of the A1AR agonist R-PIA, suggesting an interaction between µ-opioid receptors and A1ARs at the spinal level [50]. In addition, other works reported that the antiallodynic/antihyperalgesic effect of morphine is reversed in the presence of the selective A1AR antagonist DPCPX [51] or in A1ARs KO mice [52]. Beyond opioids, the involvement of A1ARs has been observed in the antinociceptive effect of non-steroidal anti-inflammatory drugs such as acetaminophen. In the formalin test, when acetaminophen was administered systemically or locally, its antinociceptive effects were reversed by the intraplantar injection of the A1AR antagonist DPCPX, suggesting a link between activation of peripheral A1ARs and acetaminophen effects [53]. The contribution of spinal A1ARs to the action of acetaminophen secondarily to the involvement of descending serotonin pathways and the release of adenosine within the spinal cord was also suggested [54]. The involvement of A1ARs was also demonstrated in the antinociceptive effects of amitriptyline [55,56], oxcarbazepine [57], levetiracetam [58], and neuropeptide S [59].
Collectively, these preclinical studies provide strong support for the therapeutic potential of A1AR agonists. However, limited clinical efficacy and relevant cardiovascular and central adverse effects have, to date, hampered the development of A1AR agonists as analgesic drugs. An alternative approach to increase selectivity and reduce the possibility of adverse effects exploiting the physiological action of endogenous adenosine is the development of A1AR-positive allosteric modulators [60,61]. These agents enhance the function of receptors activated by endogenous agonists, they are expected to have a much lower side effect potential than an exogenous orthosteric ligand, a low propensity for receptor desensitization, and a high selectivity for a given receptor subtype [62]. T62 was the first A1AR-positive allosteric modulator to be tested in animal models of pain. Intrathecal or systemic administration of T62 reduced mechanical hypersensitivity induced by spinal nerve ligation (SNL) [63,64], reversed thermal hypersensitivity in carrageenin-inflamed rats [65], and was effective for postoperative hypersensitivity following plantar incision [66]. More recently, TRR469 was characterized as one of the most potent A1AR-positive allosteric modulators so far synthesized being able to increase adenosine affinity by 33 fold [67,68,69]. TRR469 effectively inhibited nociceptive behaviors in the formalin and writhing tests, with effects comparable to morphine. Furthermore, it revealed an antiallodynic action in the streptozotocin (STZ)-induced diabetic neuropathic pain model without inducing locomotor or cataleptic side effects as the orthosteric-acting CCPA did [69].
Table 1. A1AR ligands with antinociceptive effects in preclinical models of pain.
Table 1. A1AR ligands with antinociceptive effects in preclinical models of pain.
LigandPharmacological BehaviorPain ModelSpeciesRoute of Administration
2′-Me-CCPAagonistformalin testratintra-PAG, i.p. [20]
plantar testratintra-PAG, i.p. [20]
tail flick testratintra-PAG, i.p. [20]
CCPAagonistformalin testmousei.p. [69]
writhing testmousei.p. [69]
STZ-induced mechanical allodyniamousei.p. [69]
CFA induced-mechanical allodynia and thermal hyperalgesiamouseZusanli acupoint-injection [28]
SCNL induced-mechanical allodynia and thermal hyperalgesiamouseZusanli acupoint-injection [28]
CPAagonistformalin testmousei.p. [49]; i.t. [54]
CFA-induced-mechanical allodynia and thermal hyperalgesiamousei.m. [30]; i.p. [42]; Weizhong acupoint-injection [46]
carrageenan-induced mechanical allodyniarati.pl. [18]
PGE2-induced mechanical allodyniarati.pl. [18]
SCNL-induced mechanical allodynia and thermal hyperalgesiarati.p. [21]
colonic distension-induced visceral painrats.c., i.c. [23,24]
R-PIAagonistplantar incision-induced mechanical allodyniarati.t. [25]
photochemical SCI-induced mechanical and thermal allodyniarati.t. [47]
SCNL-induced mechanical allodyniarati.t. [50]
photochemical sciatic nerve injury-induced mechanical and thermal allodyniarat, mousei.t. [52]
carrageenan-induced mechanical and thermal allodyniarat, mousei.t. [52]
T62positive allosteric modulatorSNL-induced mechanical allodyniarati.p. [63]; i.t. [63,64]; p.o. [64]
carrageenan-induced thermal hyperalgesiarati.t. [65]
plantar incision-induced mechanical allodyniarati.t. [66]
TRR469positive allosteric modulatorformalin testmousei.p. [69]
writhing testmousei.p. [69]
STZ-induced mechanical allodyniamousei.p. [69]
PAG (periaqueductal grey); i.p. (intraperitoneal); STZ (streptozotocin); CFA (Complete Freund’s adjuvant); SCNL (sciatic nerve ligation); i.t. (intrathecal); i.m. (intramuscular); i.pl. (intraplantar); s.c. (subcutaneous); i.c. (intracisternal); SCI (spinal cord injury); SNL (spinal nerve ligation); p.o. (per os).

2.2. A2AARs and Pain

The presence of A2AARS both on neurons and on glial cells is at the basis of A2AARs implications in pain [70]. The relation between A2AARs and pain has been controversial with evidence sustaining either pronociceptive and antinociceptive activity depending on the receptors’ localization and the kind of pain (Table 2) [13]. Studies supporting the pronociceptive role of A2AAR report that the selective blockade of this receptor subtype by systemic administration of SCH58261, a selective A2AAR antagonist, is able to counteract nociception; even the administration at the spinal level produced an equal effect [13,54]. These results are supported by experimental models of acute and nerve injury pain in A2AARs KO which showed a decreased algesic reaction to pain tests and even a reduction in markers of neural activity [71]. Moreover, the administration of caffeine, which is a well-known non-selective antagonist of ADA, avert the sleep deprivation due to hypersensitivity following surgical operation. A2AAR selective blockade with ZM241385 has shown to decrease surgical pain levels and the thermal hyperalgesia caused by sleep deprivation in rats. These results support the hypothesis that A2AARs are implicated in the regulation of the interplay between sleep and pain [72]. The pronociceptive effect of A2AAR stimulation was further corroborated in a study reporting that carrageenan-induced hyperalgesia was significantly reduced in A2AAR KO mice compared to wild type controls. Interestingly, the A2AAR inverse agonist ZM241385 injected into the hindpaw reduced the nociceptive behavior following carrageenan in female wild type mice, but not in males suggesting a sex difference in response to A2AAR activation in the periphery [73]. In addition, a series of inverse agonists showing two different affinity values for the A2AARs with the high affinity value in the picomolar/femtomolar range was recently synthesized [74,75] and tested for their antinociceptive properties. In particular, one of these potent inverse agonists, namely TP455, proved to be more potent than morphine in writhing and tail immersion tests in mice [74].
Furthermore, the blockade of A2AARs could provide protection in cases of neuropathic pain, which is one of the most common kinds of chronic pain, and it is found in different disorders and could lead to nerve dysfunctionalities [76]. Neuropathic pain pathophysiology is extremely intricate because it comprises central and peripheral mechanisms such as changes in ion channel expression, neurotransmitter release, and pain pathways [77]. Even oxidative stress could play an important role in the neuropathic pain origin process [78]. A body of evidence reveals that, after SCI, there are events that trigger reactive oxygen species (ROS) formation pathways such as microglia activation and glutamate release [79,80]. The injury at the sensory nerves level also involves damage to nuclear and mitochondrial DNA, and loss of antioxidant enzymes [81,82,83]. In fact, numerous studies report that the anti-oxidant or ROS scavengers administration has analgesic effects in many in vivo models of neuropathic pain. Furthermore, neuropathic pain is often a consequence of antitumoral treatments containing platinum because these drugs can provoke peripheral neuropathy and chemotherapy-induced oxidative stress is one of the important pathogenic factors damaging peripheral sensory neurons [84]. Recently, it has been proved that novel A2AAR antagonists featuring antioxidant moieties can reduce pain associated with oxaliplatin treatment in a mouse model of neuropathy reducing ROS level [85,86]. After peripheral nerve injury, A2AARs stimulation induces both activation and proliferation of microglia and astrocytes responsible for inflammation occurring in neuropathic pain, while genetic deletion of the A2AARs decreases all the behavioral and histological signs of pain [77,87]. Several studies also showed that systemic and spinal administration of the selective A2AAR antagonist SCH58261 has antinociceptive effects in different preclinical models [54,74].
Notwithstanding the coherence of the studies testifying for a pronociceptive role of A2AARs, in the literature there is evidence even for an antinociceptive role. In particular, since A2AARs are expressed in immune cells where they exert a potent anti-inflammatory action, their stimulation may be helpful in cases of inflammatory pain [3,13]. A2AARs KO animals under prolonged inflammatory conditions show an up-regulation of markers of spinal cord neural activation. In these KO mice, the loss of the antinociceptive A2AARs on immune cells exceeds the decrease in pronociceptive A2AARs on nerve terminals leading to enhanced pain signaling [88]. It is well known that the stimulation of A2AARs has anti-inflammatory effects but less is known about A2AAR agonists treatment and chronic inflammatory pain. Different studies report that A2AARs expression is up-regulated in lymphocytes of rheumatoid arthritis patients, these data should represent a basis for further investigations in this field [89,90]. The selective agonist of A2AAR CGS21680 shows the ability to slow down disease progression in an in vivo model of arthritis [91]. Even in a rat animal model, it has been demonstrated that CGS21680 treatment was very effective in decreasing clinical features in comparison to standard antirheumatic drugs such as methotrexate and etanercept [92]. The treatment with the A2AAR agonist CGS21680 was also able to inhibit the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) activation and to reduce the release of inflammatory cytokines such as tumor necrosis factor-α (TNF-α), IL-1β, and IL-6. Besides, the A2AAR stimulation leads to a decrease in metalloproteinases 1 and 3 [93]. Finally, in another mice model of monoarthritis, a new A2AAR agonist, named LASSBio-1359, showed an important analgesic effect in response to inflammatory pain. This treatment was also able to reduce inflammation by decreasing TNF-α, inducible NO synthase (iNOS) expression, and joint damage [94]. The results of the above-mentioned studies highlight a role for A2AAR agonists as a potential therapeutic tool in the management of inflammatory pain [89,93].
Different reports demonstrated an antinociceptive role of A2AAR activation in models of neuropathic pain. An acute administration of A2AAR agonists, such as ATL313 and CGS21680, leads to an analgesic effect that lasts for many weeks and reverses the mechanical allodynia and thermic hyperalgesia while decreasing the markers of microglia and astrocytes activation [95]. Interestingly, the effect of A2AARs activation was just specific for nerve injury or sensitized state suggesting a potential role of A2AAR agonists for neuropathic pain. Moreover, the blockade of A2AARs by using a receptor antagonist in the presence of an anti-IL-10 antibody reverted the effect of ATL313, suggesting that the observed effects were due to the activation of A2AARs and the simultaneous enhanced IL-10 production [95]. In a subsequent study, ATL313 induced long-lasting protection against allodynia caused by CCI, SNL, and sciatic inflammatory neuropathy (SIN), through a mechanism involving PKA and protein kinase C (PKC) [96]. In a recent study, a single intrathecal injection of the A2AAR agonists CGS21680 reversed mechanical allodynia in a rat model of SCI termed spinal neuropathic avulsion pain for at least 6weeks [97]. In the follow-up work, the peri-sciatic injection of the agonist ATL313 also demonstrated the efficacy of A2AAR activation at the site of nerve injury. These anti-allodynic effects were accompanied by a reduction in interleukin (IL)-1β and NO release, and reduced expression of iNOS and sciatic markers of monocytes/macrophages [98]. These studies revealed that the agonism toward A2AARs was able to reduce different kinds of neuropathic pain such as inflammatory neuropathic pain and traumatic ones. In all these cases the A2AARs stimulation averted and reverted the nociceptive stimuli amplification [98]. Additionally, the long time span of the analgesic effect after a single treatment suggests that A2AAR agonists could be useful for central neuropathic pain therapy. It is worth noting that in these studies, the antiallodynic effects of A2AAR agonists were associated with diminished reactive gliosis. Glial cells have a pivotal role in starting and carrying on neuropathic pain, and for this reason, many studies are directed toward the discovery of new strategies in order to defeat the pain expansion directed by glia. In recent years, A2AAR agonists have emerged as possible candidates for glial inhibition thanks to their capability to suppress inflammation in immune cells; consequently, A2AAR agonists represent a promising tool for the treatment of chronic pain of neuroinflammatory origin [99].
The activation of the A2AAR subtype also seems to be involved in the analgesic effect of neuropeptide S observed in the formalin test. Intracerebroventricular administration of this eicosapeptide reduced formalin-induced nociception during both phases 1 and phase 2 of the test, an effect counteracted by the non-selective AR antagonist caffeine or the selective A2AAR antagonist ZM241385 [59]. Besides, an interaction between A2AARs and the opioid system was reported when the antinociceptive effect exerted by the intracerebroventricular injection of Adonis, an agonist-like monoclonal antibody with high specificity for the A2AARs, was counteracted by naloxone, a non-selective opioid antagonist [100].
Table 2. A2AAR ligands with antinociceptive effects in preclinical models of pain.
Table 2. A2AAR ligands with antinociceptive effects in preclinical models of pain.
LigandPharmacological BehaviorPain ModelSpeciesRoute of Administration
ATL313agonistCCI-induced mechanical allodynia and thermal hyperalgesiarati.t. [95,96]; peri-sciatic nerve injection [98]
SNL-induced mechanical allodyniarati.t. [96]
SIN-induced mechanical allodyniarati.t. [96]
SCI-induced mechanical and thermal allodyniarati.t. [97]
CGS21680agonistformalin test (early phase)mousei.p. [49]
CFA-induced-mechanical allodynia and thermal hyperalgesiarati.p. [92]
CCI-induced mechanical allodynia and thermal hyperalgesiarati.t. [95,96]
SCI-induced mechanical and thermal allodyniarati.t. [97]
LASSBio-1359agonistformalin testmousei.p. [94]
carrageenan induced-mechanical allodynia and thermal hyperalgesiamousei.p. [94]
Adonisagonist-like monoclonal antibodyhot plate testmousei.c.v. [100]
tail flick testmousei.c.v. [100]
TP455inverse agonistwrithing testmousei.p. [74]
tail immersion testmousei.p. [74]
ZM241385antagonistwrithing testmousei.p. [74]
tail immersion testmousei.p. [74]
carrageenan induced-mechanical allodyniamouses.c. [73]
sleep deprivation-induced thermal hyperalgesiarati.c.v. [72]
plantar incision-induced mechanical allodynia and thermal hyperalgesiarati.c.v. [72]
CCI (chronic constriction injury); i.t. (intrathecal); SNL (spinal nerve ligation); SIN (sciatic inflammatory neuropathy); SCI (spinal cord injury); i.p. (intraperitoneal); CFA (Complete Freund’s adjuvant); i.c.v. (intracerebroventricular); s.c. (subcutaneous).

2.3. A2BARs and Pain

A2BARs are expressed both at the central level and in the periphery: among pain-relevant sites, they are localized on immune-inflammatory cells, where they have pro-inflammatory functions, in the spinal cord, and on astrocytes [1,101,102]. Since adenosine presents a lower affinity for A2BARs in comparison to other AR subtypes, A2BARs are more involved when adenosine concentration rises, for example in pathological conditions such as hypoxia/ischemia and inflammation [2,103].
Nonetheless, the different functions of A2BARs in various tissues and their involvement in the pathogenesis of pain are poorly known. As a consequence, more studies are needed in order to clarify their pro or anti-nociceptive actions in different types of pain conditions [11,101].
Unfortunately, studies on the relationship between pain and A2BARs are limited due to the lack of selective ligands (Table 3). One of the first studies using selective A2BAR antagonists reported an antinociceptive activity of A2BARs blockade in an acute pain model represented by the hot plate test. One of these ligands, PSB-1115, did not penetrate the blood brain barrier due to its polar sulfonate group, suggesting that peripheral A2BARs were implicated in the analgesic activity [104]. Interestingly, the efficacy of morphine was enhanced by subeffective doses of these A2BAR antagonists. In a follow-up study, the systemic administration of PSB-1115 decreased the algesic response and edema in both phases of the formalin test [105]. In the same test, the selective blockade of A2BARs by using alloxazine resulted in a dose-dependent reduction in nociceptive behavior [106]. Moreover, it has been reported that the treatment with A2BAR antagonists, MRS1754 and PSB-1115, was able to decrease pain in visceral hypersensitivity rat models [107,108]. PSB-1115 also reverted the antinociceptive effect of diphenyl diselenide, organoselenium compounds, in the hot plate test in mice [109].
A2BARs seem to be involved even in chronic pain, with evidence highlighting that these receptor subtypes stimulate the interactions between immune cells and neurons. It was reported that high extracellular adenosine levels activate A2BARs on myeloid cells, and that this leads to the activation of pain sensory neurons giving rise to hypersensitivity and chronic pain [110]. Intriguingly, the author demonstrated that A2BAR stimulation caused nociceptor hyperexcitability and promoted chronic pain via soluble IL-6 receptor trans-signaling. From these results, it is possible to deduce that the blockade of A2BARs may repress the nociceptive activity.
All these findings seem to testify for a pronociceptive role of A2BARs. However, it has been reported that even the activation of A2BARs, using a selective agonist (BAY606583), presented an analgesic effect in an accredited model of neuropathic pain, in a similar way to A2AAR agonists treatment [96]. As it is well known, both A2AARs and A2BARs lead to increased cAMP accumulation and activation of downstream pathways; they also probably have a similar spinal mechanism of action. Normally, A2BAR stimulation activates PKA and the pathway of PLC/IP3/DAG leading to changes in gene transcription, while β-arrestins are responsible for the receptor internalization mechanism [13,111].
Table 3. A2BAR ligands with antinociceptive effects in preclinical models of pain.
Table 3. A2BAR ligands with antinociceptive effects in preclinical models of pain.
LigandPharmacological BehaviorPain ModelSpeciesRoute of Administration
BAY606583agonistCCI-induced mechanical allodyniamousei.t. [96]
PSB-10antagonistformalin testmousei.p. [105]
PSB-36antagonistformalin testmousei.p. [105]
PSB-1115antagonistformalin testmousei.p. [105]
CCI (chronic constriction injury); i.t. (intrathecal); i.p. (intraperitoneal).

2.4. A3ARs and Pain

A3ARs are present at the peripheral level in many tissues including inflammatory cells; they are less expressed in the central nervous system, nonetheless their activation causes functional effects, in particular, in glial cells [112]. The possibility to exploit A3AR stimulation, using selective agonists, has been studied in different pathologies counting cancer and inflammation [112,113].
A3ARs involvement has also been investigated in relation to pain; the first pieces of evidence reported a pronociceptive role [114]. Further studies, using more selective ligands, overturned previous results showing that A3AR agonists present antinociceptive activity so, they can be useful as analgesics especially for neuropathic pain (Table 4) [113]. In fact, the systemic administration of selective A3AR agonists, such as IB-MECA, Cl-IB-MECA and MRS1898, reduced the mechanical allodynia in a model of neuropathic pain—especially IB-MECA was as efficacious as morphine. The specificity of this effect was demonstrated by blocking A3ARs with the selective antagonist MRS1523, which abrogated the analgesic effect of A3AR agonists [115]. Interestingly, the A3AR agonists have no effects in acute pain models, for instance, hot plate and tail flick tests [116]. Another A3AR selective agonist, named MRS5698, was demonstrated to be able to reduce mechanical allodynia in different models of neuropathic pain. MRS5698 had an analgesic effect in acute pain tests but its activity persisted with repeated administrations [117]. The mechanism of action of this agonist involves GABA signaling: the A3ARs activation normalizes the changes in GABA concentrations caused by nerve damages, thus restoring the GABA inhibitory effect on pain transmission [118]. Moreover, it has been noticed that A3ARs stimulation inhibits N-type calcium channel opening in isolated rat dorsal root ganglion neurons, causing a reduction in the neurotransmitter release and the neuronal excitation [119]. In another model of nerve injury that produces tactile allodynia, the daily administration of IB-MECA averted the appearance of hypersensitivities, the activation of glial cells and the altered transmission of nociceptive stimuli, resulting in an attenuation of neuropathic pain [120]. In a recent study, MRS7476, a prodrug with increased aqueous solubility compared with parent MRS5698, was found to be efficacious in reversing neuropathic pain induced by CCI [121].
Anticancer chemotherapeutic treatments often induce neuropathy as an adverse effect; the stimulation of A3ARs can help to decrease the pain in these cases. The A3AR agonist IB-MECA is able to reduce the allodynia and the hyperalgesia induced by different anticancer drugs such as paclitaxel, oxaliplatin and bortezomib without diminishing their antitumoral effectiveness; even other A3AR agonists, Cl-IB-MECA and MRS1898 present the same effects [115,116]. The pathway involved seems to imply NF-κB, ERK and p38 inhibition and the production of inflammatory cytokines. In particular, the treatment with A3AR agonists reduces the release of the pro-inflammatory cytokines TNF-α and IL-1β while increases the anti-inflammatory IL-10 [122]. Other mechanisms have been proposed to explain the antinociceptive activity of A3ARs; among these are the diminished activation of astrocytes, inhibition of cAMP, PKA, the interaction with the PLC/IP3/DAG and phosphoinositide 3-kinase (PI3K)/mitogen-activated protein kinase (MAPK)/ERK/cAMP response element-binding protein (CREB) pathways and the signaling via Gi [123]. Even in the A3AR subtype, the internalization of the receptor is mediated by β-arrestins [111]. Recently, in a model of cancer chemotherapy-induced neuropathic pain, the A3AR agonist MRS5698 attenuated pro-inflammatory IL-1β production and promoted anti-inflammatory and neuroprotective IL-10 expression by regulating the nucleotide-binding oligomerization domain-like receptor protein 3 inflammasome [124].
Besides the antinociceptive effect of A3AR agonists in cancer pain and neuropathic pain related to chemotherapy, they have also found to be potent antitumoral agents in many animal models of different forms of cancer (melanoma, prostate, colon, and hepatocellular carcinoma), where they are able to reduce tumor growth [113,125]. Their therapeutic potential has also been assessed in a model of bone cancer pain in which mammary gland tumoral cells were injected into the tibia [126]. In this model, the repeated administration of Cl-IB-MECA decreased tumor growth, mitigated the nociception and the bone degradation associated with cancer. In addition, the A3AR agonist was also effective in delaying the onset and the advancement of bone cancer with a major efficacy when the treatment with Cl-IB-MECA was done before the injection of cancer cells [126].
The involvement of A3ARs in diabetic neuropathy was also investigated. It has been demonstrated that IB-MECA treatment ameliorates mechanical hyperalgesia and thermal hypoalgesia in STZ-treated mice. Moreover, reduced expression or functionality of A3ARs promoted diabetic neuropathy development [127].
It is well established that long-lasting treatments with opioids lead to hyperalgesia and tolerance to drugs, reducing the analgesic effect of opioids in chronic pain [128,129]. In a rodent model, it has been reported that the opioid adverse effects seem to be linked to reduced A3ARs signaling. In fact, the stimulation with A3AR agonists ameliorates pain sensitivity suggesting that selective A3AR agonists may be used in addition to opioids for chronic pain management [130]. Importantly, it has been reported that the antinociceptive effects of A3AR agonists persist at least up to 2 weeks of treatment, suggesting that stimulation of A3ARs does not induce tolerance [87].
A recent study reports the effect of a new A3AR agonist, AL170, in a rat model of colitis. AL170 mitigates the colonic damage and inflammation, reducing the release of TNF-α, IL-1β, and myeloperoxidase. AL170 was demonstrated to have an efficacy comparable to that of dexamethasone, one of the most used drugs in the colitis treatment and other inflammatory bowel diseases [131]. The activation of A3ARs resulted able to decrease the infiltration of inflammatory cells and the production of inflammatory mediators thus softening visceral pain [131]. A further study revealed that the treatment with A3AR agonists is useful in another model of abdominal pain induced by colitis. In this model, Cl-IB-MECA and MRS5980 decreased visceral hypersensitivity in the postinflammatory phase as well as in the and persistence one and showed effectiveness comparable to that of linaclotide, a drug used for the treatment of irritable bowel syndrome [132,133].
The possibility to exploit A3AR agonists in rheumatic pathologies has been studied, starting from the observation that A3ARs are up-regulated in synovial tissue and peripheral blood mononuclear cells in rheumatoid arthritis patients. Treatment with A3AR agonists leads to an improvement of symptoms and clinical signs [90,93,113]. Another potential therapeutic approach for arthritis could be represented by allosteric enhancer, in order to exploit the anti-inflammatory action of A3ARs and the high adenosine concentrations typical of inflammatory pathologies. LUF6000, an A3AR allosteric modulator, was showed to reduce inflammation in models of adjuvant- and monoiodoacetate-induced arthritis [134]. Even if the analysis of the nociceptive activity was not comprised in these studies, it is reasonable to hypothesize that a decreased inflammation would be accompanied by a reduction in pain.
Table 4. A3AR ligands with antinociceptive effects in preclinical models of pain.
Table 4. A3AR ligands with antinociceptive effects in preclinical models of pain.
LigandPharmacological BehaviorPain ModelSpeciesRoute of Administration
AR170A3AR agonistcolitis-induced visceral hypersensitivityrati.p. [131]
Cl-IB-MECAA3AR agonistchemotherapy-induced mechanical allodyniamousei.p. [115]
CCI-induced mechanical allodyniamousei.p. [115]
bone cancer-induced mechanical allodyniarati.p. [126]
colitis-induced visceral hypersensitivityrati.p. [133]
IB-MECAA3AR agonistchemotherapy-induced mechanical allodyniamouse/rati.p. [115,122]
CCI-induced mechanical allodyniamousei.p. [115]; i.t. [118]
STZ-induced mechanical allodynia and thermal hyperalgesiamousei.p. [127]
opioid-induced thermal hyperalgesiaratp.o. [130]
tibial nerve injury-induced mechanical allodyniarati.p. [120]
MRS1898A3AR agonistchemotherapy-induced mechanical allodyniamousei.p. [115]
CCI-induced mechanical allodyniamousei.p. [115]
MRS5698A3AR agonistCCI-induced mechanical allodyniarats.c., i.p., i.v. [117]; i.t. [118]
spared nerve injury-induced mechanical allodyniarats.c., i.p., i.v. [117]
SNL-induced mechanical allodyniarats.c., i.p., i.v. [117]
bone cancer-induced mechanical allodyniarats.c., i.p., i.v. [117]
chemotherapy-induced mechanical allodyniarats.c., i.p., i.v. [117]; i.t. [124]
opioid-induced thermal hyperalgesiaratp.o. [130]
MRS5980A3AR agonistcolitis-induced visceral hypersensitivityrati.p. [133]
MRS7422A3AR agonistCCI-induced mechanical allodyniamousep.o. [121]
i.p. (intraperitoneal); CCI (chronic constriction injury); i.t. (intrathecal); STZ (streptozotocin); p.o. (per os); s.c. (subcutaneous); i.v. (intravenous); SNL (spinal nerve ligation).

3. Conclusions

The modulation of ARs, especially their activation, induces potent antinociceptive effects in diverse preclinical models of acute and chronic pain. Nevertheless, the efficacy of AR ligands for the pharmacological treatment of pain in humans is still ambiguous and it has also yet to be determined whether ARs modulation could be exploited to inhibit spontaneous pain. Many hopes were initially placed on A1AR agonists, but cardiovascular side effects prevented their progress in the clinic, at least with regard to their systemic administration. To get around these obstacles, alternative strategies have been proposed to continue exploiting the huge potential of adenosine and its receptors in pain management. Partial agonist and allosteric modulators of ARs have been tested in preclinical settings revealing potent antinociceptive effects with fewer side effects than conventional full agonists. Furthermore, localized activation of ARs has been proposed as a valid alternative to systemic delivery to maintain efficacy and reduce side effects, especially considering the ubiquitous localization of ARs in the human body. Exogenous ectonucleotidases as well as metabolizing enzyme inhibitors could increase the extracellular concentration of the short-living mediator adenosine, enhancing its nociceptive effects. As reviewed here, these alternative pharmacological approaches have shown promising results in preclinical models of pain and could offer a means to overcome the issues encountered so far by AR ligands in the clinic. Overall, the data summarized in this review highlight the therapeutic potential of ARs as pharmacological targets for the treatment of acute and chronic pain and the need to develop new and more effective strategies to exploit this potential.

Author Contributions

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

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

GPCRG protein coupled receptors
ARadenosine receptor
ACadenylate cyclase
cAMPcyclic adenosine monophosphate
ATPadenosine triphosphate
ADAadenosine deaminase
PKAprotein kinase A
PLCprotein lipase C
IP3inositol triphosphate
DAGdiacylglycerol
ERKextracellular signal-regulated kinase
NOnitric oxide
cGMPcyclic guanosine monophosphate
PKGprotein kinase G
PAGperiaqueductal gray
CCIchronic constriction injury
CFAcomplete Freund’s adjuvant
SCNLsciatic nerve ligation
KOknockout
PAPprostatic acid phosphatase
SCIspinal cord injury
SNLspinal nerve ligation
STZstreptozotocin
ROSreactive oxygen species
NF-κBnuclear factor kappa-light-chain-enhancer of activated B cells
TNF-αtumor necrosis factor-α
iNOSinducible nitric oxide synthase
SINsciatic inflammatory neuropathy
PKCprotein kinase C
ILinterleukin
PI3Kphosphoinositide 3-kinase
MAPKmitogen-activated protein kinase
CREBcAMP response element-binding protein

References

  1. Borea, P.A.; Gessi, S.; Merighi, S.; Vincenzi, F.; Varani, K. Pharmacology of Adenosine Receptors: The State of the Art. Physiol. Rev. 2018, 98, 1591–1625. [Google Scholar] [CrossRef] [PubMed]
  2. 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] [PubMed] [Green Version]
  3. Borea, P.A.; Gessi, S.; Merighi, S.; Varani, K. Adenosine as a Multi-Signalling Guardian Angel in Human Diseases: When, Where and How Does it Exert its Protective Effects? Trends Pharmacol. Sci. 2016, 37, 419–434. [Google Scholar] [CrossRef] [PubMed]
  4. Haskó, G.; Linden, J.; Cronstein, B.; Pacher, P. Adenosine receptors: Therapeutic aspects for inflammatory and immune diseases. Nat. Rev. Drug Discov. 2008, 7, 759–770. [Google Scholar] [CrossRef]
  5. Gompel, J.J.V.; Bower, M.R.; Worrell, G.A.; Stead, M.; Chang, S.-Y.; Goerss, S.J.; Kim, I.; Bennet, K.E.; Meyer, F.B.; Marsh, W.R.; et al. Increased cortical extracellular adenosine correlates with seizure termination. Epilepsia 2014, 55, 233–244. [Google Scholar] [CrossRef] [Green Version]
  6. During, M.J.; Spencer, D.D. Adenosine: A potential mediator of seizure arrest and postictal refractoriness. Ann. Neurol. 1992, 32, 618–624. [Google Scholar] [CrossRef]
  7. Ganesana, M.; Venton, B.J. Early changes in transient adenosine during cerebral ischemia and reperfusion injury. PLoS ONE 2018, 13, e0196932. [Google Scholar] [CrossRef]
  8. 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]
  9. 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]
  10. Gessi, S.; Merighi, S.; Sacchetto, V.; Simioni, C.; Borea, P.A. Adenosine receptors and cancer. Biochim. Biophys. Acta 2011, 1808, 1400–1412. [Google Scholar] [CrossRef] [Green Version]
  11. Antonioli, L.; Csóka, B.; Fornai, M.; Colucci, R.; Kókai, E.; Blandizzi, C.; Haskó, G. Adenosine and inflammation: What’s new on the horizon? Drug Discov. Today 2014, 19, 1051–1068. [Google Scholar] [CrossRef] [PubMed]
  12. Sawynok, J.; Liu, X.J. Adenosine in the spinal cord and periphery: Release and regulation of pain. Prog. Neurobiol. 2003, 69, 313–340. [Google Scholar] [CrossRef]
  13. Sawynok, J. Adenosine receptor targets for pain. Neuroscience 2016, 338, 1–18. [Google Scholar] [CrossRef] [PubMed]
  14. Fredholm, B.B. Adenosine—A physiological or pathophysiological agent? J. Mol. Med. 2014, 92, 201–206. [Google Scholar] [CrossRef]
  15. Varani, K.; Vincenzi, F.; Merighi, S.; Gessi, S.; Borea, P.A. Biochemical and Pharmacological Role of A1 Adenosine Receptors and Their Modulation as Novel Therapeutic Strategy. Adv. Exp. Med. Biol. 2017, 1051, 193–232. [Google Scholar] [CrossRef]
  16. Schulte, G.; Robertson, B.; Fredholm, B.B.; DeLander, G.E.; Shortland, P.; Molander, C. Distribution of antinociceptive adenosine a1 receptors in the spinal cord dorsal horn, and relationship to primary afferents and neuronal subpopulations. Neuroscience 2003, 121, 907–916. [Google Scholar] [CrossRef]
  17. Luongo, L.; Guida, F.; Imperatore, R.; Napolitano, F.; Gatta, L.; Cristino, L.; Giordano, C.; Siniscalco, D.; Di Marzo, V.; Bellini, G.; et al. The A1 adenosine receptor as a new player in microglia physiology. Glia 2014, 62, 122–132. [Google Scholar] [CrossRef]
  18. Lima, F.O.; Souza, G.R.; Verri, W.A.; Parada, C.A.; Ferreira, S.H.; Cunha, F.Q.; Cunha, T.M. Direct blockade of inflammatory hypernociception by peripheral A1 adenosine receptors: Involvement of the NO/cGMP/PKG/KATP signaling pathway. Pain 2010, 151, 506–515. [Google Scholar] [CrossRef]
  19. Sawynok, J.; Reid, A.R.; Liu, J. Spinal and peripheral adenosine A1 receptors contribute to antinociception by tramadol in the formalin test in mice. Eur. J. Pharmacol. 2013, 714, 373–378. [Google Scholar] [CrossRef]
  20. Maione, S.; de Novellis, V.; Cappellacci, L.; Palazzo, E.; Vita, D.; Luongo, L.; Stella, L.; Franchetti, P.; Marabese, I.; Rossi, F.; et al. The antinociceptive effect of 2-chloro-2′-C-methyl-N6-cyclopentyladenosine (2′-Me-CCPA), a highly selective adenosine A1 receptor agonist, in the rat. Pain 2007, 131, 281–292. [Google Scholar] [CrossRef]
  21. Gong, Q.-J.; Li, Y.-Y.; Xin, W.-J.; Wei, X.-H.; Cui, Y.; Wang, J.; Liu, Y.; Liu, C.-C.; Li, Y.-Y.; Liu, X.-G. Differential effects of adenosine A1 receptor on pain-related behavior in normal and nerve-injured rats. Brain Res. 2010, 1361, 23–30. [Google Scholar] [CrossRef] [PubMed]
  22. 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] [PubMed]
  23. Okumura, T.; Nozu, T.; Kumei, S.; Takakusaki, K.; Miyagishi, S.; Ohhira, M. Adenosine A1 receptors mediate the intracisternal injection of orexin-induced antinociceptive action against colonic distension in conscious rats. J. Neurol. Sci. 2016, 362, 106–110. [Google Scholar] [CrossRef]
  24. Okumura, T.; Nozu, T.; Ishioh, M.; Igarashi, S.; Kumei, S.; Ohhira, M. Adenosine A1 receptor agonist induces visceral antinociception via 5-HT1A, 5-HT2A, dopamine D1 or cannabinoid CB1 receptors, and the opioid system in the central nervous system. Physiol. Behav. 2020, 220, 112881. [Google Scholar] [CrossRef]
  25. Zahn, P.K.; Straub, H.; Wenk, M.; Pogatzki-Zahn, E.M. Adenosine A1 but not A2a receptor agonist reduces hyperalgesia caused by a surgical incision in rats: A pertussis toxin-sensitive G protein-dependent process. Anesthesiology 2007, 107, 797–806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Martins, D.F.; Mazzardo-Martins, L.; Cidral-Filho, F.J.; Stramosk, J.; Santos, A.R.S. Ankle joint mobilization affects postoperative pain through peripheral and central adenosine A1 receptors. Phys. Ther. 2013, 93, 401–412. [Google Scholar] [CrossRef] [Green Version]
  27. Yamaoka, G.; Horiuchi, H.; Morino, T.; Miura, H.; Ogata, T. Different analgesic effects of adenosine between postoperative and neuropathic pain. J. Orthop. Sci. 2013, 18, 130–136. [Google Scholar] [CrossRef] [Green Version]
  28. Goldman, N.; Chen, M.; Fujita, T.; Xu, Q.; Peng, W.; Liu, W.; Jensen, T.K.; Pei, Y.; Wang, F.; Han, X.; et al. Adenosine A1 receptors mediate local anti-nociceptive effects of acupuncture. Nat. Neurosci. 2010, 13, 883–888. [Google Scholar] [CrossRef] [Green Version]
  29. Zhang, M.; Dai, Q.; Liang, D.; Li, D.; Chen, S.; Chen, S.; Han, K.; Huang, L.; Wang, J. Involvement of adenosine A1 receptor in electroacupuncture-mediated inhibition of astrocyte activation during neuropathic pain. Arq. Neuro-Psiquiatr. 2018, 76, 736–742. [Google Scholar] [CrossRef] [Green Version]
  30. Liao, H.-Y.; Hsieh, C.-L.; Huang, C.-P.; Lin, Y.-W. Electroacupuncture Attenuates CFA-induced Inflammatory Pain by suppressing Nav1.8 through S100B, TRPV1, Opioid, and Adenosine Pathways in Mice. Sci. Rep. 2017, 7, 42531. [Google Scholar] [CrossRef] [Green Version]
  31. Dai, Q.-X.; Huang, L.-P.; Mo, Y.-C.; Yu, L.-N.; Du, W.-W.; Zhang, A.-Q.; Geng, W.-J.; Wang, J.-L.; Yan, M. Role of spinal adenosine A1 receptors in the analgesic effect of electroacupuncture in a rat model of neuropathic pain. J. Int. Med. Res. 2020, 48, 300060519883748. [Google Scholar] [CrossRef] [PubMed]
  32. Gao, X.; Lu, Q.; Chou, G.; Wang, Z.; Pan, R.; Xia, Y.; Hu, H.; Dai, Y. Norisoboldine attenuates inflammatory pain via the adenosine A1 receptor. Eur. J. Pain 2014, 18, 939–948. [Google Scholar] [CrossRef] [PubMed]
  33. Yin, D.; Liu, Y.-Y.; Wang, T.-X.; Hu, Z.-Z.; Qu, W.-M.; Chen, J.-F.; Cheng, N.-N.; Huang, Z.-L. Paeoniflorin exerts analgesic and hypnotic effects via adenosine A1 receptors in a mouse neuropathic pain model. Psychopharmacology 2016, 233, 281–293. [Google Scholar] [CrossRef] [PubMed]
  34. Peana, A.T.; Rubattu, P.; Piga, G.G.; Fumagalli, S.; Boatto, G.; Pippia, P.; De Montis, M.G. Involvement of adenosine A1 and A2A receptors in (-)-linalool-induced antinociception. Life Sci. 2006, 78, 2471–2474. [Google Scholar] [CrossRef] [PubMed]
  35. Valério, D.A.; Ferreira, F.I.; Cunha, T.M.; Alves-Filho, J.C.; Lima, F.O.; De Oliveira, J.R.; Ferreira, S.H.; Cunha, F.Q.; Queiroz, R.H.; Verri, W.A. Fructose-1,6-bisphosphate reduces inflammatory pain-like behaviour in mice: Role of adenosine acting on A1 receptors. Br. J. Pharmacol. 2009, 158, 558–568. [Google Scholar] [CrossRef] [Green Version]
  36. Nascimento, F.P.; Figueredo, S.M.; Marcon, R.; Martins, D.F.; Macedo, S.J.; Lima, D.A.N.; Almeida, R.C.; Ostroski, R.M.; Rodrigues, A.L.S.; Santos, A.R.S. Inosine reduces pain-related behavior in mice: Involvement of adenosine A1 and A2A receptor subtypes and protein kinase C pathways. J. Pharmacol. Exp. Ther. 2010, 334, 590–598. [Google Scholar] [CrossRef] [Green Version]
  37. Nascimento, F.P.; Macedo-Júnior, S.J.; Pamplona, F.A.; Luiz-Cerutti, M.; Córdova, M.M.; Constantino, L.; Tasca, C.I.; Dutra, R.C.; Calixto, J.B.; Reid, A.; et al. Adenosine A1 receptor-dependent antinociception induced by inosine in mice: Pharmacological, genetic and biochemical aspects. Mol. Neurobiol. 2015, 51, 1368–1378. [Google Scholar] [CrossRef]
  38. de Oliveira, E.D.; Schallenberger, C.; Böhmer, A.E.; Hansel, G.; Fagundes, A.C.; Milman, M.; Silva, M.D.P.; Oses, J.P.; Porciúncula, L.O.; Portela, L.V.; et al. Mechanisms involved in the antinociception induced by spinal administration of inosine or guanine in mice. Eur. J. Pharmacol. 2016, 772, 71–82. [Google Scholar] [CrossRef] [Green Version]
  39. Schmidt, A.P.; Böhmer, A.E.; Antunes, C.; Schallenberger, C.; Porciúncula, L.O.; Elisabetsky, E.; Lara, D.R.; Souza, D.O. Anti-nociceptive properties of the xanthine oxidase inhibitor allopurinol in mice: Role of A1 adenosine receptors. Br. J. Pharmacol. 2009, 156, 163–172. [Google Scholar] [CrossRef] [Green Version]
  40. Abad, A.S.S.; Falanji, F.; Ghanbarabadi, M.; Rad, A.; Nazemi, S.; Pejhan, A.; Amin, B. Assessment of anti-nociceptive effect of allopurinol in a neuropathic pain model. Brain Res. 2019, 1720, 146238. [Google Scholar] [CrossRef]
  41. Sowa, N.A.; Voss, M.K.; Zylka, M.J. Recombinant ecto-5’-nucleotidase (CD73) has long lasting antinociceptive effects that are dependent on adenosine A1 receptor activation. Mol. Pain 2010, 6, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Zylka, M.J.; Sowa, N.A.; Taylor-Blake, B.; Twomey, M.A.; Herrala, A.; Voikar, V.; Vihko, P. Prostatic acid phosphatase is an ectonucleotidase and suppresses pain by generating adenosine. Neuron 2008, 60, 111–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Street, S.E.; Walsh, P.L.; Sowa, N.A.; Taylor-Blake, B.; Guillot, T.S.; Vihko, P.; Wightman, R.M.; Zylka, M.J. PAP and NT5E inhibit nociceptive neurotransmission by rapidly hydrolyzing nucleotides to adenosine. Mol. Pain 2011, 7, 80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Sowa, N.A.; Street, S.E.; Vihko, P.; Zylka, M.J. Prostatic Acid Phosphatase Reduces Thermal Sensitivity and Chronic Pain Sensitization by Depleting Phosphatidylinositol 4,5-Bisphosphate. J. Neurosci. 2010, 30, 10282–10293. [Google Scholar] [CrossRef] [PubMed]
  45. Zylka, M.J. Pain-relieving prospects for adenosine receptors and ectonucleotidases. Trends Mol. Med. 2011, 17, 188–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Hurt, J.K.; Zylka, M.J. PAPupuncture has localized and long-lasting antinociceptive effects in mouse models of acute and chronic pain. Mol. Pain 2012, 8, 28. [Google Scholar] [CrossRef] [Green Version]
  47. von Heijne, M.; Hao, J.X.; Sollevi, A.; Xu, X.J.; Wiesenfeld-Hallin, Z. Marked enhancement of anti-allodynic effect by combined intrathecal administration of the adenosine A1-receptor agonist R-phenylisopropyladenosine and morphine in a rat model of central pain. Acta Anaesthesiol. Scand. 2000, 44, 665–671. [Google Scholar] [CrossRef]
  48. Ramos-Zepeda, G.A.; Herrero-Zorita, C.; Herrero, J.F. Interaction of the adenosine A1 receptor agonist N6-cyclopentyladenosine and κ-opioid receptors in rat spinal cord nociceptive reflexes. Behav. Pharm. 2014, 25, 741–749. [Google Scholar] [CrossRef]
  49. Borghi, V.; Przewlocka, B.; Labuz, D.; Maj, M.; Ilona, O.; Pavone, F. Formalin-induced pain and μ-opioid receptor density in brain and spinal cord are modulated by A1 and A2a adenosine agonists in mice. Brain Res. 2002, 956, 339–348. [Google Scholar] [CrossRef]
  50. Hwang, J.-H.; Hwang, G.-S.; Cho, S.-K.; Han, S.-M. Morphine can enhance the antiallodynic effect of intrathecal R-PIA in rats with nerve ligation injury. Anesth. Analg. 2005, 100, 461–468. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Conklin, D.R.; Li, X.; Eisenach, J.C. Intrathecal morphine reduces allodynia after peripheral nerve injury in rats via activation of a spinal A1 adenosine receptor. Anesthesiology 2005, 102, 416–420. [Google Scholar] [CrossRef] [Green Version]
  52. Wu, W.-P.; Hao, J.-X.; Halldner, L.; Lövdahl, C.; DeLander, G.E.; Wiesenfeld-Hallin, Z.; Fredholm, B.B.; Xu, X.-J. Increased nociceptive response in mice lacking the adenosine A1 receptor. Pain 2005, 113, 395–404. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, J.; Reid, A.R.; Sawynok, J. Antinociception by systemically-administered acetaminophen (paracetamol) involves spinal serotonin 5-HT7 and adenosine A1 receptors, as well as peripheral adenosine A1 receptors. Neurosci. Lett. 2013, 536, 64–68. [Google Scholar] [CrossRef]
  54. Sawynok, J.; Reid, A.R. Caffeine inhibits antinociception by acetaminophen in the formalin test by inhibiting spinal adenosine A1 receptors. Eur. J. Pharmacol. 2012, 674, 248–254. [Google Scholar] [CrossRef] [PubMed]
  55. Sawynok, J.; Reid, A.R.; Fredholm, B.B. Caffeine reverses antinociception by amitriptyline in wild type mice but not in those lacking adenosine A1 receptors. Neurosci. Lett. 2008, 440, 181–184. [Google Scholar] [CrossRef] [PubMed]
  56. Liu, J.; Reid, A.R.; Sawynok, J. Spinal serotonin 5-HT7 and adenosine A1 receptors, as well as peripheral adenosine A1 receptors, are involved in antinociception by systemically administered amitriptyline. Eur. J. Pharmacol. 2013, 698, 213–219. [Google Scholar] [CrossRef] [PubMed]
  57. Sawynok, J.; Reid, A.R.; Fredholm, B.B. Caffeine reverses antinociception by oxcarbazepine by inhibition of adenosine A1 receptors: Insights using knockout mice. Neurosci. Lett. 2010, 473, 178–181. [Google Scholar] [CrossRef] [PubMed]
  58. Stepanović-Petrović, R.M.; Micov, A.M.; Tomić, M.A.; Ugrešić, N.D. The local peripheral antihyperalgesic effect of levetiracetam and its mechanism of action in an inflammatory pain model. Anesth. Analg. 2012, 115, 1457–1466. [Google Scholar] [CrossRef] [PubMed]
  59. Victor Holanda, A.D.; Asth, L.; Santos, A.R.; Guerrini, R.; Soares-Rachetti, V.d.P.; Caló, G.; André, E.; Gavioli, E.C. Central adenosine A1 and A2A receptors mediate the antinociceptive effects of neuropeptide S in the mouse formalin test. Life Sci. 2015, 120, 8–12. [Google Scholar] [CrossRef] [Green Version]
  60. Romagnoli, R.; Baraldi, P.G.; Tabrizi, M.A.; Gessi, S.; Borea, P.A.; Merighi, S. Allosteric enhancers of A1 adenosine receptors: State of the art and new horizons for drug development. Curr. Med. Chem. 2010, 17, 3488–3502. [Google Scholar] [CrossRef]
  61. Romagnoli, R.; Baraldi, P.G.; Moorman, A.R.; Borea, P.A.; Varani, K. Current status of A1 adenosine receptor allosteric enhancers. Future Med. Chem. 2015, 7, 1247–1259. [Google Scholar] [CrossRef] [PubMed]
  62. Urwyler, S. Allosteric modulation of family C G-protein-coupled receptors: From molecular insights to therapeutic perspectives. Pharmacol. Rev. 2011, 63, 59–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Pan, H.L.; Xu, Z.; Leung, E.; Eisenach, J.C. Allosteric adenosine modulation to reduce allodynia. Anesthesiology 2001, 95, 416–420. [Google Scholar] [CrossRef] [PubMed]
  64. Li, X.; Conklin, D.; Ma, W.; Zhu, X.; Eisenach, J.C. Spinal noradrenergic activation mediates allodynia reduction from an allosteric adenosine modulator in a rat model of neuropathic pain. Pain 2002, 97, 117–125. [Google Scholar] [CrossRef]
  65. Li, X.; Conklin, D.; Pan, H.-L.; Eisenach, J.C. Allosteric adenosine receptor modulation reduces hypersensitivity following peripheral inflammation by a central mechanism. J. Pharmacol. Exp. Ther. 2003, 305, 950–955. [Google Scholar] [CrossRef] [Green Version]
  66. Obata, H.; Li, X.; Eisenach, J.C. Spinal adenosine receptor activation reduces hypersensitivity after surgery by a different mechanism than after nerve injury. Anesthesiology 2004, 100, 1258–1262. [Google Scholar] [CrossRef]
  67. Romagnoli, R.; Baraldi, P.G.; Carrion, M.D.; Cara, C.L.; Cruz-Lopez, O.; Salvador, M.K.; Preti, D.; Tabrizi, M.A.; Moorman, A.R.; Vincenzi, F.; et al. Synthesis and biological evaluation of 2-amino-3-(4-chlorobenzoyl)-4-[(4-arylpiperazin-1-yl)methyl]-5-substituted-thiophenes. effect of the 5-modification on allosteric enhancer activity at the A1 adenosine receptor. J. Med. Chem. 2012, 55, 7719–7735. [Google Scholar] [CrossRef]
  68. Vincenzi, F.; Ravani, A.; Pasquini, S.; Merighi, S.; Gessi, S.; Romagnoli, R.; Baraldi, P.G.; Borea, P.A.; Varani, K. Positive allosteric modulation of A1 adenosine receptors as a novel and promising therapeutic strategy for anxiety. Neuropharmacology 2016, 111, 283–292. [Google Scholar] [CrossRef]
  69. Vincenzi, F.; Targa, M.; Romagnoli, R.; Merighi, S.; Gessi, S.; Baraldi, P.G.; Borea, P.A.; Varani, K. TRR469, a potent A1 adenosine receptor allosteric modulator, exhibits anti-nociceptive properties in acute and neuropathic pain models in mice. Neuropharmacology 2014, 81, 6–14. [Google Scholar] [CrossRef]
  70. Gomes, C.; Ferreira, R.; George, J.; Sanches, R.; Rodrigues, D.I.; Gonçalves, N.; Cunha, R.A. Activation of microglial cells triggers a release of brain-derived neurotrophic factor (BDNF) inducing their proliferation in an adenosine A2A receptor-dependent manner: A2A receptor blockade prevents BDNF release and proliferation of microglia. J. Neuroinflammation 2013, 10, 780. [Google Scholar] [CrossRef] [Green Version]
  71. Hussey, M.J.; Clarke, G.D.; Ledent, C.; Kitchen, I.; Hourani, S.M.O. Genetic deletion of the adenosine A2A receptor in mice reduces the changes in spinal cord NMDA receptor binding and glucose uptake caused by a nociceptive stimulus. Neurosci. Lett. 2010, 479, 297–301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Hambrecht-Wiedbusch, V.S.; Gabel, M.; Liu, L.J.; Imperial, J.P.; Colmenero, A.V.; Vanini, G. Preemptive Caffeine Administration Blocks the Increase in Postoperative Pain Caused by Previous Sleep Loss in the Rat: A Potential Role for Preoptic Adenosine A2A Receptors in Sleep–Pain Interactions. Sleep 2017, 40. [Google Scholar] [CrossRef] [PubMed]
  73. Li, L.; Hao, J.X.; Fredholm, B.B.; Schulte, G.; Wiesenfeld-Hallin, Z.; Xu, X.J. Peripheral adenosine A2A receptors are involved in carrageenan-induced mechanical hyperalgesia in mice. Neuroscience 2010, 170, 923–928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. 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] [PubMed]
  75. Varano, F.; Catarzi, D.; Vincenzi, F.; Falsini, M.; Pasquini, S.; Borea, P.A.; Colotta, V.; Varani, K. Structure-activity relationship studies and pharmacological characterization of N5-heteroarylalkyl-substituted-2-(2-furanyl)thiazolo[5,4-d]pyrimidine-5,7-diamine-based derivatives as inverse agonists at human A2A adenosine receptor. Eur. J. Med. Chem. 2018, 155, 552–561. [Google Scholar] [CrossRef] [PubMed]
  76. Luongo, L.; Salvemini, D. Targeting metabotropic adenosine receptors for neuropathic pain: Focus on A2A. Brain Behav. Immun. 2018, 69, 60–61. [Google Scholar] [CrossRef]
  77. Bura, A.S.; Nadal, X.; Ledent, C.; Maldonado, R.; Valverde, O. A2A adenosine receptor regulates glia proliferation and pain after peripheral nerve injury. Pain 2008, 140, 95–103. [Google Scholar] [CrossRef]
  78. Areti, A.; Yerra, V.G.; Naidu, V.; Kumar, A. Oxidative stress and nerve damage: Role in chemotherapy induced peripheral neuropathy. Redox Biol. 2014, 2, 289–295. [Google Scholar] [CrossRef] [Green Version]
  79. Carrasco, C.; Naziroǧlu, M.; Rodríguez, A.B.; Pariente, J.A. Neuropathic Pain: Delving into the Oxidative Origin and the Possible Implication of Transient Receptor Potential Channels. Front. Physiol. 2018, 9, 95. [Google Scholar] [CrossRef]
  80. Kim, H.K.; Park, S.K.; Zhou, J.-L.; Taglialatela, G.; Chung, K.; Coggeshall, R.E.; Chung, J.M. Reactive oxygen species (ROS) play an important role in a rat model of neuropathic pain. Pain 2004, 111, 116–124. [Google Scholar] [CrossRef]
  81. Han, Y.; Smith, M.T. Pathobiology of cancer chemotherapy-induced peripheral neuropathy (CIPN). Front. Pharm. 2013, 4, 156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Kerckhove, N.; Collin, A.; Condé, S.; Chaleteix, C.; Pezet, D.; Balayssac, D. Long-Term Effects, Pathophysiological Mechanisms, and Risk Factors of Chemotherapy-Induced Peripheral Neuropathies: A Comprehensive Literature Review. Front. Pharm. 2017, 8, 86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Naik, A.K.; Tandan, S.K.; Dudhgaonkar, S.P.; Jadhav, S.H.; Kataria, M.; Prakash, V.R.; Kumar, D. Role of oxidative stress in pathophysiology of peripheral neuropathy and modulation by N-acetyl-L-cysteine in rats. Eur. J. Pain 2006, 10, 573–579. [Google Scholar] [CrossRef] [PubMed]
  84. Shim, H.S.; Bae, C.; Wang, J.; Lee, K.-H.; Hankerd, K.M.; Kim, H.K.; Chung, J.M.; La, J.-H. Peripheral and central oxidative stress in chemotherapy-induced neuropathic pain. Mol. Pain 2019, 15, 1–11. [Google Scholar] [CrossRef] [Green Version]
  85. 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]
  86. 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]
  87. 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] [PubMed]
  88. Hussey, M.J.; Clarke, G.D.; Ledent, C.; Kitchen, I.; Hourani, S.M.O. Deletion of the adenosine A2A receptor in mice enhances spinal cord neurochemical responses to an inflammatory nociceptive stimulus. Neurosci. Lett. 2012, 506, 198–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Varani, K.; Padovan, M.; Govoni, M.; Vincenzi, F.; Trotta, F.; Borea, P.A. The role of adenosine receptors in rheumatoid arthritis. Autoimmun. Rev. 2010, 10, 61–64. [Google Scholar] [CrossRef]
  90. 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] [Green Version]
  91. Mazzon, E.; Esposito, E.; Impellizzeri, D.; Di Paola, R.; Melani, A.; Bramanti, P.; Pedata, F.; Cuzzocrea, S. CGS 21680, an Agonist of the Adenosine (A2A) receptor, reduces progression of murine type II collagen-induced arthritis. J. Rheumatol. 2011, 38, 2119–2129. [Google Scholar] [CrossRef] [PubMed]
  92. 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] [PubMed] [Green Version]
  93. 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]
  94. Montes, G.C.; Hammes, N.; da Rocha, M.D.; Montagnoli, T.L.; Fraga, C.A.M.; Barreiro, E.J.; Sudo, R.T.; Zapata-Sudo, G. Treatment with Adenosine Receptor Agonist Ameliorates Pain Induced by Acute and Chronic Inflammation. J. Pharmacol. Exp. Ther. 2016, 358, 315–323. [Google Scholar] [CrossRef] [Green Version]
  95. Loram, L.C.; Harrison, J.A.; Sloane, E.M.; Hutchinson, M.R.; Sholar, P.; Taylor, F.R.; Berkelhammer, D.; Coats, B.D.; Poole, S.; Milligan, E.D.; et al. Enduring Reversal of Neuropathic Pain by a Single Intrathecal Injection of Adenosine 2A Receptor Agonists: A Novel Therapy for Neuropathic Pain. J. Neurosci. 2009, 29, 14015–14025. [Google Scholar] [CrossRef] [Green Version]
  96. Loram, L.C.; Taylor, F.R.; Strand, K.A.; Harrison, J.A.; RzasaLynn, R.; Sholar, P.; Rieger, J.; Maier, S.F.; Watkins, L.R. Intrathecal injection of adenosine 2A receptor agonists reversed neuropathic allodynia through protein kinase (PK)A/PKC signaling. Brain Behav. Immun. 2013, 33, 112–122. [Google Scholar] [CrossRef] [Green Version]
  97. Kwilasz, A.J.; Ellis, A.; Wieseler, J.; Loram, L.; Favret, J.; McFadden, A.; Springer, K.; Falci, S.; Rieger, J.; Maier, S.F.; et al. Sustained reversal of central neuropathic pain induced by a single intrathecal injection of adenosine A2A receptor agonists. Brain Behav. Immun. 2018, 69, 470–479. [Google Scholar] [CrossRef]
  98. 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]
  99. 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] [Green Version]
  100. By, Y.; Condo, J.; Durand-Gorde, J.-M.; Lejeune, P.-J.; Mallet, B.; Guieu, R.; Ruf, J. Intracerebroventricular injection of an agonist-like monoclonal antibody to adenosine A(2A) receptor has antinociceptive effects in mice. J. Neuroimmunol. 2011, 230, 178–182. [Google Scholar] [CrossRef]
  101. Feoktistov, I.; Biaggioni, I. Chapter 5—Role of Adenosine A2B Receptors in Inflammation. In Advances in Pharmacology; Jacobson, K.A., Linden, J., Eds.; Pharmacology of Purine and Pyrimidine Receptors; Academic Press: Cambridge, MA, USA, 2011; Volume 61, pp. 115–144. [Google Scholar]
  102. Popoli, P.; Pepponi, R. Potential Therapeutic Relevance of Adenosine A2B and A2A Receptors in the Central Nervous System. CNS Neurol. Disord.-Drug Targets (Former. Curr. Drug Targets CNS Neurol. Disord.) 2012, 11, 664–674. [Google Scholar] [CrossRef]
  103. Fredholm, B.B.; IJzerman, A.P.; Jacobson, K.A.; Linden, J.; Müller, C.E. International union of basic and clinical pharmacology. LXXXI. Nomenclature and classification of adenosine receptors—An update. Pharmacol. Rev. 2011, 63, 1–34. [Google Scholar] [CrossRef] [PubMed]
  104. Abo-Salem, O.M.; Hayallah, A.M.; Bilkei-Gorzo, A.; Filipek, B.; Zimmer, A.; Müller, C.E. Antinociceptive effects of novel A2B adenosine receptor antagonists. J. Pharmacol. Exp. Ther. 2004, 308, 358–366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Bilkei-Gorzo, A.; Abo-Salem, O.M.; Hayallah, A.M.; Michel, K.; Müller, C.E.; Zimmer, A. Adenosine receptor subtype-selective antagonists in inflammation and hyperalgesia. Naunyn. Schmiedebergs. Arch. Pharmacol. 2008, 377, 65–76. [Google Scholar] [CrossRef]
  106. Yoon, M.H.; Bae, H.B.; Choi, J.I.; Kim, S.J.; Chung, S.T.; Kim, C.M. Roles of Adenosine Receptor Subtypes in the Antinociceptive Effect of Intrathecal Adenosine in a Rat Formalin Test. PHA 2006, 78, 21–26. [Google Scholar] [CrossRef]
  107. Asano, T.; Tanaka, K.-I.; Tada, A.; Shimamura, H.; Tanaka, R.; Maruoka, H.; Takenaga, M.; Mizushima, T. Aminophylline suppresses stress-induced visceral hypersensitivity and defecation in irritable bowel syndrome. Sci. Rep. 2017, 7, 40214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Asano, T.; Takenaga, M. Adenosine A2B Receptors: An Optional Target for the Management of Irritable Bowel Syndrome with Diarrhea? J. Clin. Med. 2017, 6, 104. [Google Scholar] [CrossRef] [Green Version]
  109. Savegnago, L.; Jesse, C.R.; Nogueira, C.W. Caffeine and a selective adenosine A(2B) receptor antagonist but not imidazoline receptor antagonists modulate antinociception induced by diphenyl diselenide in mice. Neurosci. Lett. 2008, 436, 120–123. [Google Scholar] [CrossRef]
  110. Hu, X.; Adebiyi, M.G.; Luo, J.; Sun, K.; Le, T.-T.T.; Zhang, Y.; Wu, H.; Zhao, S.; Karmouty-Quintana, H.; Liu, H.; et al. Sustained Elevated Adenosine via ADORA2B Promotes Chronic Pain through Neuro-immune Interaction. Cell Rep. 2016, 16, 106–119. [Google Scholar] [CrossRef] [Green Version]
  111. Chen, J.-F.; Lee, C.; Chern, Y. Chapter One—Adenosine Receptor Neurobiology: Overview. In International Review of Neurobiology; Mori, A., Ed.; Adenosine Receptors in Neurology and Psychiatry; Academic Press: Cambridge, MA, USA, 2014; Volume 119, pp. 1–49. [Google Scholar]
  112. Borea, P.A.; Varani, K.; Vincenzi, F.; Baraldi, P.G.; Tabrizi, M.A.; Merighi, S.; Gessi, S. The A3 Adenosine Receptor: History and Perspectives. Pharm. Rev. 2015, 67, 74–102. [Google Scholar] [CrossRef] [Green Version]
  113. Fishman, P.; Bar-Yehuda, S.; Liang, B.T.; Jacobson, K.A. Pharmacological and therapeutic effects of A3 adenosine receptor agonists. Drug Discov. Today 2012, 17, 359–366. [Google Scholar] [CrossRef] [Green Version]
  114. Wu, W.-P.; Hao, J.-X.; Halldner-Henriksson, L.; Xu, X.-J.; Jacobson, M.A.; Wiesenfeld-Hallin, Z.; Fredholm, B.B. Decreased inflammatory pain due to reduced carrageenan-induced inflammation in mice lacking adenosine A3 receptors. Neuroscience 2002, 114, 523–527. [Google Scholar] [CrossRef]
  115. Chen, Z.; Janes, K.; Chen, C.; Doyle, T.; Bryant, L.; Tosh, D.K.; Jacobson, K.A.; Salvemini, D. Controlling murine and rat chronic pain through A3 adenosine receptor activation. FASEB J. 2012, 26, 1855–1865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Janes, K.; Symons-Liguori, A.; Jacobson, K.A.; Salvemini, D. Identification of A3 adenosine receptor agonists as novel non-narcotic analgesics. Br. J. Pharm. 2016, 173, 1253–1267. [Google Scholar] [CrossRef] [Green Version]
  117. Little, J.W.; Ford, A.; Symons-Liguori, A.M.; Chen, Z.; Janes, K.; Doyle, T.; Xie, J.; Luongo, L.; Tosh, D.K.; Maione, S.; et al. Endogenous adenosine A3 receptor activation selectively alleviates persistent pain states. Brain 2015, 138, 28–35. [Google Scholar] [CrossRef] [PubMed]
  118. Ford, A.; Castonguay, A.; Cottet, M.; Little, J.W.; Chen, Z.; Symons-Liguori, A.M.; Doyle, T.; Egan, T.M.; Vanderah, T.W.; Koninck, Y.D.; et al. Engagement of the GABA to KCC2 Signaling Pathway Contributes to the Analgesic Effects of A3AR Agonists in Neuropathic Pain. J. Neurosci. 2015, 35, 6057–6067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Coppi, E.; Cherchi, F.; Fusco, I.; Failli, P.; Vona, A.; Dettori, I.; Gaviano, L.; Lucarini, E.; Jacobson, K.A.; Tosh, D.K.; et al. Adenosine A3 receptor activation inhibits pronociceptive N-type Ca2+ currents and cell excitability in dorsal root ganglion neurons. Pain 2019, 160, 1103–1118. [Google Scholar] [CrossRef]
  120. Terayama, R.; Tabata, M.; Maruhama, K.; Iida, S. A3 adenosine receptor agonist attenuates neuropathic pain by suppressing activation of microglia and convergence of nociceptive inputs in the spinal dorsal horn. Exp. Brain Res. 2018, 236, 3203–3213. [Google Scholar] [CrossRef]
  121. Suresh, R.R.; Jain, S.; Chen, Z.; Tosh, D.K.; Ma, Y.; Podszun, M.C.; Rotman, Y.; Salvemini, D.; Jacobson, K.A. Design and in vivo activity of A3 adenosine receptor agonist prodrugs. Purinergic Signal. 2020, 16, 367–377. [Google Scholar] [CrossRef]
  122. Janes, K.; Esposito, E.; Doyle, T.; Cuzzocrea, S.; Tosh, D.K.; Jacobson, K.A.; Salvemini, D. A3 adenosine receptor agonist prevents the development of paclitaxel-induced neuropathic pain by modulating spinal glial-restricted redox-dependent signaling pathways. Pain 2014, 155, 2560–2567. [Google Scholar] [CrossRef] [Green Version]
  123. Kim, Y.; Kwon, S.Y.; Jung, H.S.; Park, Y.J.; Kim, Y.S.; In, J.H.; Choi, J.W.; Kim, J.A.; Joo, J.D. Amitriptyline inhibits the MAPK/ERK and CREB pathways and proinflammatory cytokines through A3AR activation in rat neuropathic pain models. Korean J. Anesth. 2019, 72, 60–67. [Google Scholar] [CrossRef]
  124. Wahlman, C.; Doyle, T.M.; Little, J.W.; Luongo, L.; Janes, K.; Chen, Z.; Esposito, E.; Tosh, D.K.; Cuzzocrea, S.; Jacobson, K.A.; et al. Chemotherapy-induced pain is promoted by enhanced spinal adenosine kinase levels through astrocyte-dependent mechanisms. Pain 2018, 159, 1025–1034. [Google Scholar] [CrossRef] [PubMed]
  125. Antonioli, L.; Blandizzi, C.; Pacher, P.; Haskó, G. Immunity, inflammation and cancer: A leading role for adenosine. Nat. Rev. Cancer 2013, 13, 842–857. [Google Scholar] [CrossRef] [PubMed]
  126. Varani, K.; Vincenzi, F.; Targa, M.; Paradiso, B.; Parrilli, A.; Fini, M.; Lanza, G.; Borea, P.A. The stimulation of A3 adenosine receptors reduces bone-residing breast cancer in a rat preclinical model. Eur. J. Cancer 2013, 49, 482–491. [Google Scholar] [CrossRef] [PubMed]
  127. Yan, H.; Zhang, E.; Feng, C.; Zhao, X. Role of A3 adenosine receptor in diabetic neuropathy. J. Neurosci. Res. 2016, 94, 936–946. [Google Scholar] [CrossRef]
  128. Hayhurst, C.J.; Durieux, M.E. Differential Opioid Tolerance and Opioid-induced HyperalgesiaA Clinical Reality. Anesthesiology 2016, 124, 483–488. [Google Scholar] [CrossRef]
  129. Rivat, C.; Ballantyne, J. The dark side of opioids in pain management: Basic science explains clinical observation. Pain Rep. 2016, 1, e570. [Google Scholar] [CrossRef]
  130. Doyle, T.M.; Largent-Milnes, T.M.; Chen, Z.; Staikopoulos, V.; Esposito, E.; Dalgarno, R.; Fan, C.; Tosh, D.K.; Cuzzocrea, S.; Jacobson, K.A.; et al. Chronic Morphine-Induced Changes in Signaling at the A3 Adenosine Receptor Contribute to Morphine-Induced Hyperalgesia, Tolerance, and Withdrawal. J. Pharmacol. Exp. Ther. 2020, 374, 331–341. [Google Scholar] [CrossRef]
  131. Antonioli, L.; Lucarini, E.; Lambertucci, C.; Fornai, M.; Pellegrini, C.; Benvenuti, L.; Di Cesare Mannelli, L.; Spinaci, A.; Marucci, G.; Blandizzi, C.; et al. The Anti-Inflammatory and Pain-Relieving Effects of AR170, an Adenosine A3 Receptor Agonist, in a Rat Model of Colitis. Cells 2020, 9, 1509. [Google Scholar] [CrossRef]
  132. Lucarini, E.; Coppi, E.; Micheli, L.; Parisio, C.; Vona, A.; Cherchi, F.; Pugliese, A.M.; Pedata, F.; Failli, P.; Palomino, S.; et al. Acute visceral pain relief mediated by A3AR agonists in rats: Involvement of N-type voltage-gated calcium channels. Pain 2020. [Google Scholar] [CrossRef]
  133. Grundy, L.; Harrington, A.M.; Castro, J.; Garcia-Caraballo, S.; Deiteren, A.; Maddern, J.; Rychkov, G.Y.; Ge, P.; Peters, S.; Feil, R.; et al. Chronic linaclotide treatment reduces colitis-induced neuroplasticity and reverses persistent bladder dysfunction. JCI Insight 2018, 3, e121841. [Google Scholar] [CrossRef] [PubMed]
  134. 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] [PubMed] [Green Version]
Figure 1. Adenosine (Ado) metabolism and involvement of adenosine receptors (ARs) in pain. The main source of adenosine is adenosine triphosphate (ATP) released from various cell types in response to different stimuli. ATP is dephosphorylated to adenosine diphosphate (ADP)/adenosine monophosphate (AMP) and then to adenosine by two ectonucleotidases (CD39, CD73). In nociception, the elevated levels of adenosine may alter the pain signaling. Thus, the modulation of adenosine metabolisms, increasing its levels, could represent an alternative strategy for pain management. Soluble CD73 provokes long-lasting thermal antihyperalgesic and mechanical antiallodynic effects through A1AR activation. Prostatic acid phosphatase (PAP), acting as an ectonucleotidase, induces A1AR-dependent antinociceptive effects in inflammatory and neuropathic pain models. Extracellular adenosine is rapidly metabolized to inosine by adenosine deaminase (ADA). Inosine is able to bind A1ARs, with an affinity similar to that of adenosine, inducing antinociceptive effects. Another strategy to promote the accumulation of inosine is represented by the inhibitors of the enzyme xanthine oxidase such as allopurinol. In the extracellular space, adenosine can interact with its receptors. A1ARs stimulation with adenosine, adenosine metabolites like inosine, or synthetic agonists presents analgesic effects in acute, neuropathic, visceral, postoperative, and inflammatory pain. Activation of A2AARs by endogenous adenosine or exogenous agonists results in antinociception in case of inflammatory pain. While, A2AARs blockade shows analgesic effects in neuropathic pain. Regarding A2BARs, their stimulation has antinociceptive effects in neuropathic pain and their blockade is useful for acute pain treatment. Finally, A3ARs activation gives analgesic effects in different types of pain such as neuropathic, cancer, and visceral pain.
Figure 1. Adenosine (Ado) metabolism and involvement of adenosine receptors (ARs) in pain. The main source of adenosine is adenosine triphosphate (ATP) released from various cell types in response to different stimuli. ATP is dephosphorylated to adenosine diphosphate (ADP)/adenosine monophosphate (AMP) and then to adenosine by two ectonucleotidases (CD39, CD73). In nociception, the elevated levels of adenosine may alter the pain signaling. Thus, the modulation of adenosine metabolisms, increasing its levels, could represent an alternative strategy for pain management. Soluble CD73 provokes long-lasting thermal antihyperalgesic and mechanical antiallodynic effects through A1AR activation. Prostatic acid phosphatase (PAP), acting as an ectonucleotidase, induces A1AR-dependent antinociceptive effects in inflammatory and neuropathic pain models. Extracellular adenosine is rapidly metabolized to inosine by adenosine deaminase (ADA). Inosine is able to bind A1ARs, with an affinity similar to that of adenosine, inducing antinociceptive effects. Another strategy to promote the accumulation of inosine is represented by the inhibitors of the enzyme xanthine oxidase such as allopurinol. In the extracellular space, adenosine can interact with its receptors. A1ARs stimulation with adenosine, adenosine metabolites like inosine, or synthetic agonists presents analgesic effects in acute, neuropathic, visceral, postoperative, and inflammatory pain. Activation of A2AARs by endogenous adenosine or exogenous agonists results in antinociception in case of inflammatory pain. While, A2AARs blockade shows analgesic effects in neuropathic pain. Regarding A2BARs, their stimulation has antinociceptive effects in neuropathic pain and their blockade is useful for acute pain treatment. Finally, A3ARs activation gives analgesic effects in different types of pain such as neuropathic, cancer, and visceral pain.
Ijms 21 08710 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/ijms21228710

AMA Style

Vincenzi F, Pasquini S, Borea PA, Varani K. Targeting Adenosine Receptors: A Potential Pharmacological Avenue for Acute and Chronic Pain. International Journal of Molecular Sciences. 2020; 21(22):8710. https://doi.org/10.3390/ijms21228710

Chicago/Turabian Style

Vincenzi, Fabrizio, Silvia Pasquini, Pier Andrea Borea, and Katia Varani. 2020. "Targeting Adenosine Receptors: A Potential Pharmacological Avenue for Acute and Chronic Pain" International Journal of Molecular Sciences 21, no. 22: 8710. https://doi.org/10.3390/ijms21228710

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