Open access peer-reviewed chapter

Purinergic System in Immune Response

Written By

Yerly Magnolia Useche Salvador

Submitted: 20 January 2022 Reviewed: 14 March 2022 Published: 24 May 2022

DOI: 10.5772/intechopen.104485

From the Edited Volume

Purinergic System

Edited by Margarete Dulce Bagatini

Chapter metrics overview

139 Chapter Downloads

View Full Metrics

Abstract

In mammalian cells, the purinergic signaling and inflammatory mediators regulate each other. During microbial infection, nucleotides and nucleosides from both dying host cells and pathogens may be recognized by the host receptors. These receptors include purinergic receptors such P2X, P2Y, and A2A, as well Toll-like receptors, and NOD-like receptors. The interaction with most of these receptors activates immune responses, including inflammasome activation, releasing of pro-inflammatory cytokines, reactive nitrogen and oxygen species production, apoptosis induction, and regulation of T cell responses. Conversely, activation of adenosine receptors is associated with anti-inflammatory responses. The magnitude of resultant responses may contribute not only to the host defense but also to the homeostatic clearance of pathogens, or even to the severe progression of infectious diseases. In this chapter, we discuss how the purinergic signaling activation upregulates or downregulates mechanisms in infectious diseases caused by the bacterial, parasite, and viral pathogens, including SARS-CoV-2. As a concluding remark, purinergic signaling can modulate not only infectious diseases but also cancer, metabolic, and cardiovascular diseases, constituting a strategy for the development of treatments.

Keywords

  • purinergic receptors
  • immune responses
  • inflammation
  • infectious disease

1. Introduction

The purinergic signaling modulates pathways of both neural and non-neural physiological processes, including immune responses, inflammation, pain, platelet aggregation, endothelium-mediated vasodilation, proliferation, and cell death [1]. Three main components are part of the system. Purinergic: extracellular nucleotides and nucleosides, their receptors, and the ectoenzymes responsible for regulating the levels of these molecules [2]. In addition, nucleotides, nucleosides, and uric acid resulting from the death of infected or injured cells are also recognized by other receptors better known for their role in pathogen recognition as Toll-like receptors (TLRs), and NOD-like receptors (NLRs) [3]. Other immune innate receptors are able to detect nucleic acids (RNA or DNA) from either phagocyted or circulating microbes, including retinoic acid-inducible gene I (RIG-I)-like receptors (RLRs), the C-type lectin receptors (CLRs), and cytosolic sensors [4, 5]. All immune cells are able to recognize nucleotides as a danger signal throughout either purinergic or no purinergic receptors. Immediately various immune responses can be activated, such as pro-inflammatory cytokines secretion by macrophages, quimocine production by eosinophils, maturation of dendritic cells (DCs), as well as T and B cells costimulation [6].

Extracellular nucleotides and nucleosides are released, along with many other molecules, from dead cells. Apoptosis and necrosis are the cell death mechanisms that can operate in physiological conditions. Apoptosis is activated by genetically controlled cell signals to modulate cell growth and development, as it is a programmed event. It is an ordered process that does not trigger inflammation. Conversely, necrosis is a not regulated cell death, characterized by the cell content release as a consequence of the effect of diverse environmental factors, leading to higher inflammation around [7]. Furthermore, during infections, some intracellular pathogens require cell lysis, while others have developed mechanisms to prevent cell death during their replication and dissemination outside the infected cell.

ATP is known as a damage signal, released or leaked by injured cells, or a molecular pattern associated with damage [8]. Necrotic cells may use either pannexin channels or connexin hemichannels to release intracellular ATP, and the P2X7R may be involved in this process [6]. Adenosine is a nucleoside that mediates anti-inflammatory and immunosuppressive actions, such as inhibiting the production of pro-inflammatory cytokines and lymphocyte proliferation [9]. In pathological conditions, adenosine plays a protective role acting as an endogenous regulator of innate immunity and in host defense against excessive tissue damage associated with inflammation [10]. The A1 and A2A receptors (A1R and A2AR) are activated by adenosine concentrations in the nanomolar range, while the A2B and A3 receptors (A2BR and A3R) become active only when the extracellular levels of adenosine rise in the micromolar range during periods of inflammation, hypoxia or ischemia [11] (Figure 1). Other nucleosides are recognized by several P2 receptors, which are going to explain later.

Figure 1.

Infected cells release nucleotides and nucleosides. High ATP concentration can induce both dead cell (dashed line, in dendritic cell) and several functions over immune cells as maturation, and quimiotaxis. Low ATP concentration ([ATP]) can induce (arrow) or inhibit (bar-headed line) responses depending on the immune cell. Low adenosine concentration ([ADO]) can induce ROS and NO. High adenosine concentration ([ADO]) inhibit pro-inflammatory cytokine (*Cytokines) expression. Reddish and bluish background means pro-inflammatory and anti-inflammatory context, respectively. Source: The figure 1 design is original and cell vectors were modified from freepik’s: https://www.freepik.es/vectores/personas, People Vector created by brgfx - http://www.freepik.es.

Advertisement

2. Regulation of extracellular nucleotides and nucleosides

In the context of the immune response, as mentioned, while the extracellular ATP (eATP) exhibits pro-inflammatory and stimulatory effects in the immune system, either appropriate or exacerbated responses [6], adenosine has primarily anti-inflammatory and inhibitory effects [9]. Therefore, the balance between ATP versus adenosine levels is important in modulating cellular immune responses and pathogen survival [12]. The concentrations of extracellular nucleotides and nucleosides are regulated by immune and non-immune cells through the action of enzymes anchored to the cell membrane, with their catalytic site facing the extracellular environment [13]. These enzymes, called ectonucleotidases, hydrolyze extracellular nucleotides into their respective nucleosides to control exacerbated levels of nucleotides and maintain steady-state conditions [12]. Among them, ecto-nucleoside triphosphate diphosphohydrolases (E-NTPDases, CD39, or apyrases), hydrolyzes both ATP and ADP to AMP, in the presence of divalent cations such as calcium and magnesium. Sequentially, the E-5′-nucleotidase (CD73) terminates the ectonucleotidase cascade with the hydrolysis of monophosphate nucleotides, resulting in adenosine. This in turn is hydrolyzed by the enzyme adenosine deaminase (E-ADA), transforming adenosine into inosine, its inactive metabolite [13]. In addition, the ecto-nucleotide pyrophosphatase/phosphodiesterases (E-NPPs) yield free nucleosides. The nitrogenous bases are hydrolyzed from nucleosides by the action of phosphorylases that yield ribose-1-P and free bases. If the nucleosides and/or bases are not re-utilized, the purine bases are further degraded to uric acid [3].

Two important environments for the highly reported purinergic signaling activation are the central nervous system (CNS) and the blood. The principal cell type in the brain involved in ATP degradation is the microglia through the expression of CD39 [14]. The ATP-regulation in the blood is mediated by the red blood cells, which may regulate tissue circulation and O2 delivery by releasing the vasodilator ATP in response to hypoxia. When released extracellularly, ATP is rapidly degraded to ADP in the circulation by ectonucleotidases. Moreover, ADP acting on P2Y13 receptors on red blood cells serves as a negative feedback pathway for the inhibition of ATP release [15].

Nucleosides and nucleotides are recognized by purinergic receptors P1, P2, TLRs, and NLRs. The recognition implies that nucleosides and nucleotides are temporarily held at different concentrations to activate their respective receptors. Purinergic receptors are divided into two families: P1 and P2 receptors [1]. The G-protein coupled metabotropic P1 receptors recognize exclusively extracellular adenosine and can be subdivided into A1R, A2AR, A2BR, and A3R. The P2 receptors can be subdivided into two subtypes: non-selective ion-gated channel P2X receptors (that recognize ATP) and G-coupled P2Y receptors (that recognize ATP, ADP, UTP, UDP, and UDP-glucose) [6]. Actually, it has been described seven P2X receptors (from P2X1 to P2X7) with different affinities for ATP. The P2X7 receptor has a low affinity for ATP (requiring ≥100 mM to be activated; while others can be activated at lower concentrations) [12].

The consensus about the relationship between purinergic signaling and the immune system can be summed up by the opposing effects of ATP and adenosine. The ATP contributes to triggering the inflammatory response along with molecular patterns associated with pathogens [5]. Conversely, the adenosine nucleoside mediates anti-inflammatory and immunosuppressive actions, such as inhibiting the production of pro-inflammatory cytokines and lymphocyte proliferation [9]. However, there are other nucleotides and nucleosides modulating the immune system.

Advertisement

3. Effects of purinergic signaling on the innate and adaptive immune system

The eATP released from stressed, dying or infected cells bind to P2 receptors (as P2X7R) and may lead to pathogen elimination through several mechanisms: (1) host cell death; (2) inflammasome activation and IL-1β secretion; and (3) production of reactive oxygen species (ROS) and nitric oxide (NO); promoting lysosome and phagosome fusion [16].

P2X7R activation is associated with pore formation, which depends on the concentration and duration of ATP treatment [17], as well as leads to the opening of pores that allow the passage of small molecules (< 900 Da) [18] as dinucleotides or nucleosides, increasing the extracellular concentration of these purinergic ligands. Inflammasomes are multi-protein complexes assembled in the host cell in response to infection or cellular stress, leading to non-homeostatic and lytic cell death, called pyroptosis. P2X7R was shown to activate NLRP3 inflammasomes [19]; and recently, caspase-11-induced pyroptosis was shown to require pannexin-1 channels and the P2X7R activation [20]. Pyroptosis is important because of the cytokines, chemokines, and damage-associated molecular patterns (DAMPs) which are released to the extracellular compartment, and also because intracellular pathogens are exposed to extracellular immune response, thus allowing their destruction [21]. At the same time, the inflammasomes lead to the maturation and secretion of pro-inflammatory cytokines, such as IL-1b and IL-18 [21]. IL-1β affects virtually all cells and organs of the body and is one of the most important cytokines that mediate autoimmunity, infections, and degenerative diseases [22]. This cytokine has a role in the CNS as an endogenous pyrogenic agent, and it can also induce inflammation, leukocyte recruitment, and Th17 profile immune responses [23]. In addition to eATP, uridin diphosphate (UDP) is released by the cleavage of pannexin-1 channels via caspase in apoptotic cells resulting from the vesicular stomatitis virus infection. Then, the UDP-P2Y6R signaling is able to protect both cells and mice from infection through an increase in IFN-β production, in acute neurotropic infection [24].

Purinergic receptors P1 (P1R) and P2 (P2R) can be expressed simultaneously in almost all immune cells, apparently depending on their ligand concentrations in the extracellular space [25]. Therefore, eATP-P2R interactions also activate pro-inflammatory responses in immune cells [26], as we have seen in infected or dying non-immune cells.

Neutrophils, granulocytes participating in both immune systems innate and adaptative, are the first immune cells in arrived at the inflammation site, constituting the main acute inflammatory response against pathogens by both phagocytosis and the oxidative microbicidal molecules production [25]. Neutrophils are the more affected cells by the purinergic signaling, probably because they express several purinergic receptors [26]. For instance, P2Y11 receptor (P2Y11R) is responsible for the ATP-mediated differentiation and maturation of granulocytic progenitors in the bone marrow [27]. Also, the interaction between eATP-P2Y11R mediates the inhibition of neutrophil apoptosis [28] and increases the chemotactic response of neutrophils [29]. In addition to eATP, other nucleotide released by damaged cells is the uridine 5′-diphosphoglucose (UDP-glucose), which activates the P2Y14R signaling. UDP-glucose promotes chemotaxis of freshly isolated human neutrophils through P2Y14R activation [30]. Moreover, some inflammatory diseases have been related to P2Y14R activation. During pelvic inflammatory disease, in the endometria in both women and female mice, the P2Y14R and pro-inflammatory cytokines as IL-8 are up-regulated in the epithelium [31]. Then, the design of therapies to modulate mucosal immunity may be done by targeting P2Y14R [30].

Occasionally, some nucleosides are more concentrated than the ATP in the extracellular matrix. When uridine triphosphate (UTP) is more available than ATP, P2Y2 receptors (P2Y2R) may be activated by either ATP or UTP mediating several activities in the P2Y2R expressing cells. For instance, fibrotic lung disease is related to some activities mediated by P2Y2R such as the lung fibroblast’s proliferation and migration, the recruitment of neutrophils, and IL-6 secretion in the lungs [32]. By the other way, when the eATP-P2X1R signaling is activated in neutrophils and platelets, activated neutrophils are recruited to the injury site and their adherence to vessel walls together with the platelets occurred, promoting both thrombosis and fibrinogenesis [33].

In addition to neutrophils, the eosinophils, and basophils, other granulocyte cells, which are activated during parasitic infections and allergies, are also regulated by the purinergic system. The accumulation of eosinophils during lung inflammation is triggered by UTP-P2Y2R interaction that induces the expression of VCAM-1, an adhesion molecule, which in turn, induces changes in endothelial cell shape for the opening of passageways through which eosinophils migrate [34]. Moreover, P2Y2R activation by ATP in eosinophils has been reported to induce chemotaxis in allergic lung inflammation [35]. In other circumstances, when UDP have more concentrated than UTP or ATP, the UDP-P2Y6R signaling induces IgE-dependent degranulation in human basophils [36].

During infections, dendritic cells (DCs) are responsible for presenting antigens to naive T cells and activating them, making a link between the innate and adaptive immune response [37]. The ATP-P2R interaction is involved in the migration and differentiation of DCs [38]. Specifically, the eATP-P2Y11R interaction modulates the maturation of human monocyte-derived dendritic cells (MoDCs) [39]. Moreover, P2Y2R activation by ATP promotes chemotaxis of MoDCs [35]. In fact, the eATP-P2Y11R interaction mediates the migration of DCs accordingly with the DC type, although all DC populations express P2Y11R. MoDCs down-regulate the P2Y11R expression, decreasing the inhibition of migration triggered by ATP. While either interleukin-3 receptor-positive plasmacytoid DCs or CD1c + peripheral blood DCs do not inhibit their migration by ATP. Then, the possibility of a meeting between DCs and antigens may be mediated by gradients of ATP formed in and around inflamed areas. Therefore, after vaccination, the migration of DCs charged with antigens to near lymph nodes may be increased with the inhibition of P2Y11R expression. This strategy could improve the time of response after vaccination [38]. In addition, P2Y12 receptor (P2Y12R) modulates murine DCs function by ADP, including induction of intracellular Ca2+ transportation, macropinocytosis, and T-cell stimulation [40]. However, the stimulation of the P2X7R with ATP can induce cell death; such as in murine spleen-derived DCs, which increase the permeabilization and the intracellular calcium, resulting in apoptosis [41].

As reviewed, the T-cell activation by DCs can be modulated by the purinergic system. Additionally, lymphocytes (mainly T and B cells) that are characterized by expressing antigen receptors, allowing the activation of anti-microbial responses [25], also express purinergic receptors which modulates lymphocyte proliferation, differentiation, and functioning. Immature T cells pass through the thymus for differentiation, where stromal epithelial cells are in charge of both the positive and negative selection processes, which in turn defines the T cell functional profile between CD8+ or CD4+ cells [42]. In the thymus, P2X7R and P2Y2R are expressed in several cells as murine thymic epithelial cells (TECs), leading to the release of calcium from intracellular stores and increasing the permeabilization membrane [43]; possibly leading to TECs apoptosis as well as reported in DCs [41], and ending in the alteration of both T-cell differentiation and their peripheral functioning. Similarly, the eATP-P2X7R signaling leads to the opening of a transmembrane cationic channel that allows K+ efflux and Na + and Ca2+ influx and promotes cytoplasmic membrane depolarization in the phagocytic cell of the thymic reticulum [18], leading to increase permeabilization and apoptosis, and impairing of T-cell precursors proliferation.

During immune synapse, naïve T-cells release ATP throughout pannexin-1 channels, then the eATP interaction with P2X1R and P2X4R receptors regulates T-cell activation, calcium entry, and IL-2 release [44]. Also, γδ T-cells, abundant at barrier sites such as the skin, gut, lung, and reproductive tract, are activated and upregulated tumor necrosis factor-alpha (TNF-α), and interferon-γ (IFN-γ) release through the eATP-P2X4R interaction [45]. Moreover, P2X4Rs participate in the migration process of CD4+ T-cells. This migration is mediated by the chemokine stromal-derived factor-1α (SDF-1α), which triggers the T-cell polarization by the accumulation of ATP producing mitochondria near ATP-releasing pannexin-1 channels and newly expressed P2X4Rs. This set of molecules promotes both the Ca2+ influx and sustained mitochondrial ATP production required for the pseudopod protrusion and T-cell migration [46]. Conversely, exogenous activation of P2Y11R with eATP blocks T-cell trafficking [47].

Specifically, the purinergic system may modulate inflammatory severe clinical conditions like sepsis. This condition has two phases, first is the hyperinflammatory phase, which may be later restricted by the immunosuppressive phase. However, this last phase characterized by high blood levels of regulatory T cells (Tregs) is strongly associated with mortality. Tregs proliferation is controlled by P2Y12R activation in both Tregs and platelets, and P2Y12R blockade restores the immunological homeostasis. Therefore, this strategy may guide pharmacological treatment for sepsis and increase patient survival [48].

Among innate immune cells acting mainly in chronic inflammatory responses are the mononuclear phagocytes, which are in circulation as monocytes or in several tissues as macrophages, or in specific tissues as microglial cells in the brain. These phagocytes may have either a pro-inflammatory or anti-inflammatory role depending on the type of cytokines around them and express several P2 receptors [25]. For instance, the major pathway of macrophage activation is the eATP-P2Y11R signaling, which leads to cytokine release [49]. Moreover, macrophages exposed to LPS, increase the P2YR and P2X7R activation mediated by eATP, modulating the IL-1β, TNF-α, and NO production [50]. Monocyte adhesion process is also regulated by the ATP-P2R interaction [51]. For example, activated P2Y12R induces both vascular smooth muscle inflammatory changes via MCP-1 upregulation and monocyte adhesion into the vascular wall, promoting atherosclerotic lesions [52].

Microglia is the cell responsible for the immune function of the nervous system in both physiological and pathological conditions [53]. Increased P2X7R expression and its ATP-mediated activation in microglia are observed after the LPS brain challenge, leading to increased immune response associated with NO and ROS, along with reduced neuronal viability. Inhibition of this purinergic response may be a neuroprotective strategy in brain inflammatory diseases [54]. In addition, the upregulation of P2Y6, P2Y12, P2Y13, and P2Y14 receptors in spinal microglia is associated with the development of neuropathic pain [55]. In an inflammatory context, ADP acting on P2Y12R induces extension of microglia processes thereby attracting this cell to the site of ATP/ADP leaking or release. Moreover, the ADP-P2Y12R activation in microglia induces intracellular calcium accumulation, which in turn causes the increase of CC chemokine ligand 3 (CCL3) expression in the peripheral injured site and also in the spinal cord, inducing neuropathic pain. Since the inhibition of CCL3-CCR5 signaling suppresses the development of neuropathic pain, treatments based on inhibition of CCL3 expression can be promising to control this kind of inflammatory disorder [56].

The P2Y6R is also upregulated in microglia when neurons are damaged, then the UDP-P2Y6R signaling facilitates microglial phagocytosis [57]. Consistently, the brain injury caused by ischemic accidents is increased by the inhibition of both P2Y6R expression and the microglia-phagocytic activity [58]. The UDP-P2Y6R signaling is also associated with neuropathic pain and is partially explained by the induction of CCL2 production through the MAP kinases-NF-kappaB pathway in microglia [59]. CCL2 is a recruitment factor of myeloid cells to the regions with injured neurons. In the spinal cord, CCL2 released from primary afferent neurons and reactive astrocytes could contribute to either the induction or maintenance of chronic pain [60], neurodegenerative and neuroinflammatory diseases [61].

As mentioned above, in addition to purinergic receptors, other receptors are able to recognize eATP such as the NLRs and TLRs. These receptors recognize both pathogen-associated molecular patterns (PAMPs) and damage-associated molecular patterns (DAMPs) (including eATP and uric acid). The relationship between these receptors and nucleotides or nucleosides, and the structure and functions of NLRs will be addressed immediately. TLRs will be explained through the case of gout disease.

The NLRs are cytoplasmic receptors and the structure has three domains: a common domain organization with a central conserved domain NOD (NACHT: NAIP, CIITA, HET-E, and TP-2), N-terminal effector domain, and C-terminal leucine-rich repeats (LRRs) [62]. The NAIP motif (neuronal apoptosis inhibitor protein) and the CIITA motif (MHC class II transcription activator) contain a distinct predicted nucleoside triphosphatase (NTPase) domain. In addition, NTPase domain in CIITA shows highly significant sequence similarity to CARD4 (pro-apoptotic protein). Therefore, the NACHT family includes both pro-apoptotic (e.g. CARD4) and anti-apoptotic (e.g. NAIP) predicted NTPases [63]. In consequence, all NLRs could modulate apoptotic process. However, either the possible apoptotic effect or the ATPase activity of the NATCH domain and the consequence on the concentration of nucleotides and derivatives in the cytoplasm must be the subject of study.

Most NLRs recognize various ligands activating inflammatory responses. These ligands come from different sources, including microbial pathogens (peptidoglycan, flagellin, viral RNA, fungal hyphae, etc.), host cells (ATPs, uric acid, etc.), and esteril activators (alum, silica, UV radiation, skin irritants, etc.). In addition, some NLRs respond to cytokines such as interferons. The activated NLRs show various functions that can be divided into four broad categories: inflammasome formation, signaling transduction, transcription activation, and autophagy [64]. Several NLRs have been used to identify inflammasomes, depending on the receptor that recognizes the PAMPs (for example, NLRP1, NLRP3, AIM2, NLRC4), while the other group of inflammasomes can be activated by cytosolic lipopolysaccharides (LPS) derived from gram-negative bacteria. The NLRP3 inflammasome can be activated by different stimuli, such as bacterial, viral, and fungal pathogens, pore-forming toxins, crystals, silica, and DAMPs (for example, eATP) [65]. The activation of the NLRP3 inflammasome requires two signals: (1) a PAMP, such as LPS, leading to transcription of NF-kB, upregulating genes encoding pro-inflammatory cytokines, chemokines, and proteins involved in the inflammasome platform; and (2) a DAMP, such as eATP, which induces inflammasome activation after ligation to the P2X7R. Once activated, these complexes promote activation of the protease caspase-1, which cleaves pro-IL-1β and pro-IL-18 into their active forms: IL-1β and IL-18 [12].

In addition to the inflammasome activation, eATP induces ROS production. ROS are highly reactive chemicals formed from O2 (such as peroxides, superoxides, and hydroxyl radicals) [66]. For instance, respiratory epithelial cells induce mitochondrial ROS in response to influenza infection. ROS induces the expression of type III interferon, a response associated with viral infection control [67]. Moreover, Porphyromonas gingivalis (P. gingivalis) infection of gingival epithelial cells induces assembly of the P2X4R and P2X7R to form a pore, pannexin-1, promoting ROS production triggered by ATP-P2X7R activation. Later the ROS can activate the NLRP3 inflammasome and caspase-1, resulting in bacteria death [68].

Some TLRs recognize in addition to PAMPs from intracellular or extracellular pathogens, others molecules like ATP and uric acid. TLRs recognizing uric acid are involved in several metabolic and cardiovascular diseases, including gout, chronic renal tubular damage, autosomal dominant polycystic kidney disease, and cartilage degeneration. Among them, the accumulation of uric acid crystals (monosodium urate - MSU -) in the joints causes arthritis which is the base of the metabolic disease called Gout [64]. The uric acid also could cause inflammatory events by the activation of the NLRP3 inflammasome and its activation induces the IL-1β formation, which leads to the development of gouty arthropathy [69]. Moreover, when the level of uric acid is higher than 6.8 mg/dL, MSU crystals are formed and are recognized by TLRs. These TLRs then activate the NALP3 inflammasome. MSU also triggers neutrophil activation and further produces immune mediators, which lead to a pro-inflammatory response [70]. In the mice model of Gout, it was found that MSU crystals are recognized by TLR2 and TLR4 with the participation of the TLR adapter protein myeloid differentiation factor 88 (MyD88) in bone marrow-derived macrophages. After recognition by these TLRs, MSU induced the production of pro-inflammatory cytokines as IL-1β, TNF-α, keratinocyte-derived cytokine/growth-related oncogene alpha (KDC/GROα), and transforming growth factor beta1(TGF-β1). Moreover, neutrophil influx, local induction of IL-1β, and more pro-inflammatory reaction were promoted [71].

In addition, patients with hyperuricemia (high levels of uric acid in the blood) develop vascular diseases associated with the formation of aminocarbonyl radicals from excess uric acid with the concomitant oxidative effect [72]. Additionaly, metabolic and cardiovascular diseases are associated with the activation of the renin-angiotensin system (RAS) mediated by uric acid, in adipose tissue. For instance, high blood pressure and increased expression of both TLR2/4, pro-inflammatory cytokines (TNF-α and IL-6), and RAS activation in adipocytes were found in hyperuricemic rats. These high levels of cytokines and RAS components were reverted by TLR2/4 RNA silencing [73]. Proinflammatory pathways are induced by uric acid and angiotensin II-mediated by TLR4 in renal proximal tubular cells developing chronic tubular damage [74]. Moreover, TLR-2 and TLR-4 gene expressions are associated with rapid progression in autosomal dominant polycystic kidney disease (ADPKD) patients [75]. Furthermore, in human chondrocytes, the accumulation of both calcium pyrophosphate dihydrate (CPPD) and MSU crystals was associated with increased expression of TLR2 and the NO generation triggered by TLR2 signaling, inducing inflammation and cartilage deterioration. Other TLR signaling pathways producing NO release are induced by both MSU and CPPD crystals, including the Pl3K/Akt/NF-κB signaling pathway and other mediators as MyD88, IRAK1, and TRAF6 [76].

Advertisement

4. Downregulation of immune responses by purinergic signaling

In the lymph nodes and spleen, lymphocytes are stimulated through eATP-P2X7R interaction to promote the Th1 pro-inflammatory response [77]. However, the eATP may also play an immunosuppressive role. This mechanism occurs mainly when eATP is in low (micromolar) concentrations, increasing its affinity for P2YR, located on the surface of lymphocytes. When stimulated, P2YRs promote the downregulation in the expression and release of pro-inflammatory cytokines, promoting a protective effect against excessive tissue damage [25]. Moreover, chronic exposure of DCs to low ATP doses reduces the capacity to stimulate the Th1 response, while Th2 response is favored [78], which induces the activation of T-cells with an anti-inflammatory profile. However, micromolar levels of eATP through P2Y2R induce the mechanisms of phagocytosis and increase ROS and NO production by macrophages and neutrophils [25].

The most recognized effector of anti-inflammatory responses of the purinergic system is adenosine. Extracellular adenosine is recognized by P1 receptors (A2AR and A2BR). High concentrations of adenosine activate A2AR, inhibiting the production of pro-inflammatory cytokines by macrophages [79], and also decreasing the production of ROS and NO by neutrophils, monocytes, and macrophages. However, low concentrations of adenosine (lower than micromoles) increase phagocytosis and ROS production by activation of A1R in neutrophils [25]. Also, adenosine acts on A2AR inhibiting the production of IL-12 and TNF-α in mice liver and preventing the damage by injury [10].

A1AR and A2AR are abundantly expressed at synapses in the CNS, modulating the synaptic efficacy [80]. A1AR and A2AR receptors are also expressed in the microglia and their activation promotes anti-inflammatory and migration activities, respectively [81]. In the presence of mild alterations of CNS high amounts of ATP can be release, then the activation of P2X7R induces both activation and pro-inflammatory response by microglia, leading to surrounding neuronal death [82]. Therefore, the ATP regulation in the CNS is critical; it has been suggested that CD39 expression has an essential role in cell proliferation and growth, inflammatory processes, and triggering cellular responses from ATP-induced contribute to apoptosis and host defense [83]. Moreover, in vivo studies on brain trauma and Alzheimer’s disease, neuroinflammation has been detected associated with ATP release from microglia, occurring in an uncontrolled way mainly through pannexin/connexin hemichannels [84]. While adenosine binding to A1R or A2AR during brain disorder exerts neuroprotective and immunosuppressive capacities, respectively [85].

Advertisement

5. Pathogens and purinergic signaling mediating immune response

Some obligated intracellular bacterial pathogens have diverse target organs. For instance, Mycobacterium tuberculosis (M. tuberculosis) invades lungs, kidney, spinal cord, and brain, while Chlamydia trachomatis (C. trachomatis) infects genital and ocular tissue. It has been reported that these pathogens may be controlled by eATP treatment. The eATP treatment of macrophages enhances their antimicrobial properties in a P2X7R-dependent manner. For instance, eATP-related killing of M. tuberculosis and C. trachomatis within human and murine macrophages is associated with mobilization of intracellular Ca+2 and consequently lysosomal fusion and acidification of the containing-pathogen phagosomes [86, 87]. Moreover, adenine nucleotides (AMP and ATP) and adenosine can inhibit C. trachomatis growth in epithelial cells; for instance, micromolar eATP concentrations reversibly inhibit chlamydial infection via the P2X4 receptor in epithelial cells [88]. While millimolar eATP concentrations are sufficient to inhibit chlamydial infection via P2X7 receptor in macrophages [89].

Another bacteria controlled by eATP-triggered mechanisms is P. gingivalis, an intracellular bacterium that infects gingival epithelial cells (GECs) and the oral mucosa, causing periodontitis [90]. The mechanisms described to control P. gingivalis via eATP activation are (1) P2X7R-mediated apoptosis [91]; (2) ROS production via P2X7R-NADPH oxidase signaling [92] inflammasome activation and IL-1β release [93]. Conversely, adenosine-receptors signaling downregulates the immune response. For example, A2AR stimulated by its agonist CGS21680 increases the P. gingivalis proliferation in GECs [94]. On the other hand, P. gingivalis can inhibit eATP-induced apoptosis in GECs through the secretion of the enzyme nucleoside-diphosphate-kinase (NDK), which can hydrolyze eATP [92], inhibiting the three eATP-triggered mechanisms to control the bacteria.

Intracellular protozoan parasites as Leishmania amazonensis (L. amazonensis), Toxoplasma gondii (T. gondii), and Trypanosoma cruzi (T. cruzi) may also be controlled by eATP. Murine macrophages infected with L. amazonensis and cells from established cutaneous lesions enhanced P2X7R expression and were more responsive to eATP activation, inhibiting parasite growth [95]. Also, UTP inhibits L. amazonensis infection in murine macrophages, probably by P2Y2R or P2Y4R activation, inducing morphological damage inside the parasite promoting apoptosis of macrophages, producing ROS and reactive nitrogen species (RNS), and increasing intracellular Ca+2 concentrations [96]. Nonetheless, several species of Leishmania modulate eATP and adenosine levels by directly acting on these molecules or by inducing CD39 and CD73 expression on the infected cells, influencing the immune response and contributing to parasite growth or survival [97]. First, saliva from phlebotomine sand flies, Leishmania promastigotes vector, is rich in adenosine and AMP, which levels are mediated by the enzymatic activity of apyrases and 5′-nucleotidases present in saliva [98]. Therefore, low levels of ATP decrease the activation of P2X and P2Y, inhibiting platelet aggregation [99], leading to the free spread of the parasite. Second, parasite-infected cells increase the expression of ectonucleotidases (CD39 and CD73) on their surface and, therefore, also the production of extracellular adenosine. Later, adenosine mediates the activation of A2BR, necessary for the expression of CD40 (DC activation marker). Conversely, the blockade of the A2BR inhibits the DC activation and interferes with T cell proliferation [100]. Third, the A2BR activation inhibits the production of NO and IL-12 by infected macrophages [101], allowing the parasite survive. Moreover, the increased A2BR expression involved IL-10 production by infected cells, in monocytes from patients with visceral leishmaniasis [102], inducing an anti-inflammatory response.

During the acute toxoplasmosis, in the mice brain, occurs an increase in purines (ATP, ADP, AMP, adenosine, xanthine, hypoxanthine, and uric acid), while in chronic toxoplasmosis reduction of the same purines, except the antioxidant, uric acid, occurs [103]. Specifically, the high levels of xanthine and hypoxanthine are associated with the inhibition of the enzyme xanthine oxidase, which catalyzes the production of uric acid, reported in T. gondii infected mice [104]. Moreover, in mice with toxoplasmosis, the elevated ATP levels promote increased levels of calcium inside infected cells mediated by P2X receptors, causing damage to the cells and contributing to nervous disorders and behavioral alterations [105]. Besides, T. gondii is eliminated by the activation of eATP-P2X7R signaling in infected macrophages mediated by the acidification of the parasitophorous vacuole and ROS production [106]. Additionally, UTP and UDP treatment in murine macrophages infected with T. gondii promotes 90% parasite elimination without inducing NO, ROS or apoptosis in the host cell [107]. Interestingly, UTP and UDP induced prematurely parasite egress from the host cell via P2Y2R, P2Y4R, and P2Y6R thus compromising infectivity and replication of the egressed parasites [107].

Interestingly, T. gondii does not produce adenosine, then the efficient transformation to the bradyzoite or long-lived cyst stage depends on the extracellular adenosine produced by ectonucleotidases expressed by infected cells [108]. CD73 expression promotes T. gondii differentiation and cyst formation by a mechanism dependent on adenosine generation, but independent of adenosine-receptor signaling [108]. In fact, some pathogens stimulate extracellular adenosine generation independently of the host. Staphylococcus aureus produces adenosine synthase A (AdsA), a cell wall-anchored enzyme which allows the bacteria to escape from clearance by phagocytosis and favoring the formation of organ abscesses. Moreover, bacteria from the gastrointestinal tract as Enterococcus faecalis and Streptococcus mutans possess homologs of adenosine synthase [109].

During acute T. cruzi infection in CNS, the parasite stimulates P2X7R expression in the cerebral cortex, being activated by the available eATP. As a consequence, P2X7R activation induces ATP release, from immune and non-immune cells, chiefly via pannexin hemichannels-boosting inflammation [83]. Besides, the P2X7R activation by the parasite transialidase is involved in the massive loss of immature CD4/CD8 double-positive cells, which determine the prominent thymus atrophy in acute T. cruzi infection [110]. In addition, eATP mediates mechanisms to control the parasite as astrocyte proliferation and differentiation, cytokine release, and the ROS and RNS formation. Furthermore, ATP, ADP, and AMP hydrolysis occur in infected animals, related to the enzymatic modulation in the presence of high parasitism [111]. As mentioned, adenosine has a neuroprotective role; however, E-ADA activity is augmented in infected animals, producing iosine which is later used in the purine rescue pathway of T. cruzi [83], and in other parasites such as Trypanosoma evansi [112] and Plasmodium falciparum [113].

Interestingly, some pathogens have evolved extracellular nucleotide-hydrolyzing enzymes that mimic the ectonucleotidases expressed in the host, probably inhibiting the ATP-driven immune response [114]. For instance, the surface of T. cruzi expresses an Mg2+-dependent ecto-ATPase enzyme (Mg-eATPase), which have higher activity in trypomastigotes, maybe promoting the host infection [115]. Another parasite with Mg-eATPase is L. amazonensis, whose virulent promastigotes are very efficient in hydrolyzing eATP and acquiring adenosine, which is used by the parasite [116].

The participation of the purinergic system in the immune response and pathogenesis as a consequence of SARS-CoV-2 virus infection has also been reported. SARS-CoV-2 induces the IFN response in patients, through MDA5-mediated RNA sensing with the participation of IRF3, IRF5, and NF-κB/p65 pro-inflammatory transcription factors [117]. However, Coronaviruses can evade the MDA5 recognition by forming endoplasmic reticulum-derived membrane vesicles around their RNA [118], delaying the IFN production; and in consequence, allowing higher viral replication. Viral load is highly correlated with the levels of IFNs and TNF-α, suggesting that viral load may drive high cytokine production [119]. Increased levels of TNF-α during inflammation induce ATP release via pannexin-1 channels [120]. ATP exportation out of the cell implies a deficit of intracellular ATP available for the ATP-dependent enzymes in the JAK–STAT pathway induced by IFN-I, limiting the cytokine expression and T helper cell activation [121].

At the same time, a pro-inflammatory immune response is initiated by the increase in the extracellular ATP and ADP levels in the microenvironment of immune cells activating the P2XRs and P2YRs [122]. The eATP-P2X7R signaling activation is a key process in the hyper inflammation resulting from the severe pro-inflammatory immune response against SARS-CoV-2 [123]. High levels of eATP are accompanied by the desensitization of all P1 and P2 purinergic receptors, except P2X7R, inducing more hyper inflammation [124], the worst scenario for a COVID-19 patient.

Shortly after the inflammatory explosion or simultaneously, the eATP concentration could decrease by the CD39-mediated transformation into eADP and eAMP, while adenosine quickly increases by the CD73-mediated eAMP conversion [125]. Then, immunosuppressive responses are activated by the adenosine excess in interaction with their A2AR and A2BR, including inhibition of macrophages and lymphocytes [10].

Moreover, increased eADP levels promote platelet activation and intravascular thrombosis mediated by P2YRs [126], and COVID-19 patients with pneumonia frequently developed microvascular thrombosis in their lungs [127]. In summary, the degree of involvement of purinergic receptors and their ligands in the response to SARS-CoV-2 virus infection may partially explain, the presence of asymptomatic infected people and the variation in the severity among the COVID-19 patients.

Advertisement

6. Perspectives

During inflammation, macrophages, NK cells, and some lymphocytes activities are impaired by the interaction of their adenosine receptors and the high extracellular levels of adenosine [10]. Therefore, the factors involved in extracellular adenosine production may be used in anti-inflammatory strategies, including the ectonucleotidases CD39 which degrades ATP into AMP, and the ectonucleotidase CD73 which converts AMP into adenosine. Following this rationale, several monoclonal antibodies (mAb) have been developed using CD73, CD39, and A2AR receptors as a target [128]. For instance, a humanized anti-CD39 mAb prevents the ATP-ADP conversion. Moreover, the enhancement of T cells and NK cells function was found, when CD39 was blocked by either antibodies or inhibitors such as POM-1; aside from increased T cell proliferation by the lack of suppression exerted by Treg cells [129].

Moreover, the prevention of AMP to adenosine conversion is also achieved using the mAb anti-CD73 which leads to its internalization [130]. As consequence, the adenosine low levels can not inhibit lymphocytes, therefore CD8 and macrophages activities are enhanced, while both myeloid suppressor cells and Treg lymphocytes are inhibited [128]. Lymphocyte proliferation is also promoted with the administration of an A2AR antagonist in two ways, removing checkpoints on both CD4+ FoxP3+ Tregs and CD8+ effector T cells development, and inhibiting the expression of the programmed death-1 receptor (PD-1) in draining lymph nodes [131]. Some of these drugs have been used as anti-cancer therapies, nevertheless, they have a potential action in many diseases based on immunosuppressive mechanisms.

On the other hand, antagonists of P2X7R as lidocaine can disrupt hyperinflammation, leading to the activation of anti-inflammatory responses. For instance, the clonal expansion of Tregs in lymph nodes is promoted by the P2X7Rs-mediated inhibition of the immune cells in the lymphatic system. Later, the Tregs control the hyperinflammation throughout their anti-inflammatory mechanisms [16]. Also, since eATP-P2Y11R signaling is highly activated in macrophages, P2Y11R antagonists maybe they can be used for the treatment of inflammatory diseases [40]. These strategies may constitute immunotherapy with promising results for inflammatory-based diseases, such as severe forms of various viral or bacterial infections, or even autoimmune diseases.

Advertisement

7. Conclusion

PX and PY receptors are involved in the inflammasome activation, apoptosis induction, oxidant production and activation of several immune cells, mechanisms that can control the infection of several pathogens. Conversely, adenosine is generally associated with the downregulation of inflammation. However, the effects triggered by eATP and nucleosides and their respective purinergic receptors in infected cells, depend on several aspects. These include first, the ability of the receptor expression by infected cells; second, the mechanisms to maintain the balance of nucleotide and nucleoside concentrations in the extracellular environment; and third, the survival strategies of specific pathogens.

The purinergic signaling can modulate infections by different intracellular pathogens, including viruses, bacteria, and parasites, and mediates inflammatory processes in metabolic, cardiovascular, and cancer diseases. For this reason, this knowledge field represents an important focus for future research regarding the survival and elimination of different pathogens and the maintenance of the homeostasis of the diseases related to hyper-inflammation.

Advertisement

Acknowledgments

Post-doctoral Fellowship of Research Support Foundation of the State of Rio de Janeiro–FAPERJ—Health Research Networks Program in the State of Rio de Janeiro—2019, Brazil. Processo E-26/ 202.139/2020.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Burnstock G. Purine and pyrimidine receptors. Cellular and Molecular Life Sciences. 2007;64:1471-1483
  2. 2. Yegutkin G. Nucleotide and nucleoside converting ectoenzymes: Important modulators of purinergic signalling cascade. Biochimica et Biophysica Acta. 1783;2008:673-694
  3. 3. Ishii KJ, Akira S. Potential link between the immune system and metabolism of nucleic acids. Current Opinion in Immunology. 2008;20:524-529. DOI: 10.1016/j.coi.2008.07.002
  4. 4. Medzhitov R. Recognition of microorganisms and activation of the immune response. Nature. 2007;449:819e26
  5. 5. Ishii KJ, Koyama S, Nakagawa A, Coban C, Akira S. Host innate immune receptors and beyond: Making sense of microbial infections. Cell Host & Microbe. 2008;3:352-363
  6. 6. Idzko M, Ferrari D, Eltzschig HK. Nucleotide signalling during inflammation. Nature. 2014;509:310e7
  7. 7. Fink SL, Cookson BT. Apoptosis, pyroptosis, and necrosis: Mechanistic description of dead and dying eukaryotic cells. Infection and Immunity. 2005;73:1907-1916. DOI: 10.1128/IAI.73.4.1907-1916.2005
  8. 8. Di Virgilio F. Purinergic mechanism in the immune system: A signal of danger for dendritic cells. Purinergic Signalling. 2005;1:205-209
  9. 9. Gessi S, Merighi S, Varani K, Cattabriga E, Benini A, Mirandola P, et al. Adenosine receptors in colon carcinoma tissues and colon tumoral cell lines: Focus on the a(3) adenosine subtype. Journal of Cellular Physiology. 2007;211:826-836. DOI: 10.1002/jcp.20994
  10. 10. Ohta A, Sitkovsky M. Role of G-protein-coupled adenosine receptors in downregulation of inflammation and protection from tissue damage. Nature. 2001;414:916-920. DOI: 10.1038/414916a
  11. 11. Olah ME, Stiles GL. Adenosine receptor subtypes: Characterization and therapeutic regulation. Annual Review of Pharmacology and Toxicology. 1995;35:581-606
  12. 12. Coutinho-Silva R, Ojcius DM. Role of extracellular nucleotides in the immune response against intracellular bacteria and protozoan parasites. Microbes and Infection. 2012;14:1271e7
  13. 13. Zimmermann H. Biochemistry, localization and functional roles of ectonucleotidases in the nervous system. Progress in Neurobiology. 1996;49:589-618
  14. 14. Braun N, Sévigny J, Robson SC, Enjyoji K, Guckelberger O, Hammer K, et al. Assignment of ecto-nucleoside triphosphate diphosphohydrolase-1/cd39 expression to microglia and vasculature of the brain. The European Journal of Neuroscience. 2000;12:4357-4366
  15. 15. Wang L, Olivecrona G, Gotberg M, Olsson ML, Winzell MS, Erlinge D. ADP acting on P2Y13 receptors is a negative feedback pathway for ATP release from human red blood cells. Circulation Research. 2005;96:189-196. DOI: 10.1161/01.res.0000153670.07559.e4
  16. 16. Almeida-da-Silva CLC, Morandini AC, Ulrich H, Ojcius DM, Coutinho-Silva R. Purinergic signaling during Porphyromonas gingivalis infection. Biomedical Journal. 2016;39:251-260. DOI: 10.1016/j.bj.2016.08.003
  17. 17. Morandini AC, Savio LE, Coutinho-Silva R. The role of P2X7 receptor in infectious inflammatory diseases and the influence of ectonucleotidases. Biomedical Journal. 2014;37:169e77
  18. 18. Coutinho-Silva R, Alves LA, de Carvalho AC, Savino W, Persechini PM. Characterization of P2Z purinergic receptors on phagocytic cells of the thymic reticulum in culture. Biochimica et Biophysica Acta. 1996;1280(2):217-222. DOI: 10.1016/0005-2736(95)00293-6
  19. 19. Pelegrin P. P2X7 receptor and the NLRP3 inflammasome: Partners in crime. Biochemical Pharmacology. 2021;187:114385. DOI: 10.1016/j.bcp.2020.114385
  20. 20. Yang D, He Y, Munoz-Planillo R, Liu Q , Nunez G. Caspase-11 requires the Pannexin-1 channel and the purinergic P2X7 pore to mediate pyroptosis and endotoxic shock. Immunity. 2015;43:923e32
  21. 21. Lamkanfi M, Dixit VM. Modulation of inflammasome pathways by bacterial and viral pathogens. Journal of Immunology. 2011;187:597e602
  22. 22. Garlanda C, Dinarello CA, Mantovani A. The interleukin-1 family: Back to the future. Immunity. 2013;39:1003e18
  23. 23. Dinarello CA. Anti-inflammatory agents: Present and future. Cell. 2010;140:935e50
  24. 24. Li R, Tan B, Yan Y, Ma X, Zhang N, Zhang Z, et al. Extracellular UDP and P2Y6 function as a danger signal to protect mice from vesicular stomatitis virus infection through an increase in IFN-beta production. Journal of Immunology. 2014;193:4515-4526. DOI: 10.4049/jimmunol.1301930
  25. 25. Bours MJ, Swennen EL, Di Virgilio F, Cronstein BN, Dagnelie PC. Adenosine 5′-triphosphate and adenosine as endogenous signaling molecules in immunity and inflammation. Pharmacology & Therapeutics. 2006;112:358-404. DOI: 10.1016/j.pharmthera.2005.04.013
  26. 26. Hasan D, Shono A, van Kalken CK, van der Spek PJ, Krenning EP, Kotani T. A novel definition and treatment of hyperinflammation in COVID-19 based on purinergic signalling. Purinergic Signal. 2021;10:1-47. DOI: 10.1007/s11302-021-09814-6
  27. 27. van der Weyden L, Conigrave AD, Morris MB. Signal transduction and white cell maturation via extracellular ATP and the P2Y11 receptor. Immunology and Cell Biology. 2000;78:369-374. DOI: 10.1046/j.1440-1711.2000.00918.x
  28. 28. Vaughan KR, Stokes L, Prince LR, Marriott HM, Meis S, Kassack MU, et al. Inhibition of neutrophil apoptosis by ATP is mediated by the P2Y11 receptor. Journal of Immunology. 2007;179:8544-8553
  29. 29. Alkayed F, Kashimata M, Koyama N, Hayashi T, Tamura Y, Azuma Y. P2Y11 purinoceptor mediates the ATP-enhanced chemotactic response of rat neutrophils. Journal of Pharmacological Sciences. 2012;120:288-295
  30. 30. Barrett MO, Sesma JI, Ball CB, Jayasekara PS, Jacobson KA, Lazarowski ER, et al. A selective high-affinity antagonist of the P2Y14 receptor inhibits UDP-glucose-stimulated chemotaxis of human neutrophils. Molecular Pharmacology. 2013;84:41-49. DOI: 10.1124/mol.113.085654
  31. 31. Arase T, Uchida H, Kajitani T, Ono M, Tamaki K, Oda H, et al. The UDP-glucose receptor P2RY14 triggers innate mucosal immunity in the female reproductive tract by inducing IL-8. Journal of Immunology. 2009;182:7074-7084. DOI: 10.4049/jimmunol.0900001
  32. 32. Muller T, Fay S, Vieira RP, Karmouty-Quintana H, Cicko S, Ayata K, et al. The purinergic receptor subtype P2Y2 mediates chemotaxis of neutrophils and fibroblasts in fibrotic lung disease. Oncotarget. 2017;8:35962-35972. DOI: 10.18632/oncotarget.16414
  33. 33. Oury C, Lecut C, Hego A, Wera O, Delierneux C. Purinergic control of inflammation and thrombosis: Role of P2X1 receptors. Computational and Structural Biotechnology Journal. 2015;13:106-110. DOI: 10.1016/j.csbj.2014.11.008
  34. 34. Vanderstocken G, Bondue B, Horckmans M, Di Pietrantonio L, Robaye B, Boeynaems JM, et al. P2Y2 receptor regulates VCAM-1 membrane and soluble forms and eosinophil accumulation during lung inflammation. Journal of Immunology. 2010;185:3702-3707. DOI: 10.4049/jimmunol.0903908
  35. 35. Muller T, Robaye B, Vieira RP, Ferrari D, Grimm M, Jakob T, et al. The purinergic receptor P2Y2 receptor mediates chemotaxis of dendritic cells and eosinophils in allergic lung inflammation. Allergy. 2010;65:1545-1553. DOI: 10.1111/j.1398-9995.2010.02426.x
  36. 36. Nakano M, Ito K, Yuno T, Soma N, Aburakawa S, Kasai K, et al. UDP/P2Y6 receptor signaling regulates IgE-dependent degranulation in human basophils. Allergology International. 2017;66:574-580. DOI: 10.1016/j.alit.2017.02.014
  37. 37. Banchereau J, Briere F, Caux C, Davoust J, Lebecque S, Liu YJ, et al. Immunobiology of dendritic cells. Annual Review of Immunology. 2000;18:767-811. DOI: 10.1146/annurev.immunol.18.1.767
  38. 38. Schnurr M, Toy T, Stoitzner P, Cameron P, Shin A, Beecroft T, et al. ATP gradients inhibit the migratory capacity of specific human dendritic cell types: Implications for P2Y11 receptor signaling. Blood. 2003;102:613-620. DOI: 10.1182/blood-2002-12-3745
  39. 39. Wilkin F, Duhant X, Bruyns C, Suarez-Huerta N, Boeynaems JM, Robaye B. The P2Y11 receptor mediates the ATP-induced maturation of human monocyte-derived dendritic cells. Journal of Immunology. 2001;166:7172-7177
  40. 40. Ben Addi A, Cammarata D, Conley PB, Boeynaems JM, Robaye B. Role of the P2Y12 receptor in the modulation of murine dendritic cell function by ADP. Journal of Immunology. 2010;185:5900-5906. DOI: 10.4049/jimmunol.0901799
  41. 41. Nihei OK, de Carvalho AC, Savino W, Alves LA. Pharmacologic properties of P(2Z)/P2X(7) receptor characterized in murine dendritic cells: Role on the induction of apoptosis. Blood. 2000;96:996-1005
  42. 42. Wang HX, Pan W, Zheng L, et al. Thymic Epithelial Cells Contribute to Thymopoiesis and T Cell Development. Frontiers in Immunology. 2020;10:3099. DOI: 10.3389/fimmu.2019.03099
  43. 43. Bisaggio RD, Nihei OK, Persechini PM, Savino W, Alves LA. Characterization of P2 receptors in thymic epithelial cells. Cellular and Molecular Biology. 2001;47:19-31
  44. 44. Woehrle T, Yip L, Elkhal A, Sumi Y, Chen Y, Yao Y, et al. Pannexin-1 hemichannel-mediated ATP release together with P2X1 and P2X4 receptors regulate T-cell activation at the immune synapse. Blood. 2010;116:3475-3484. DOI: 10.1182/blood-2010-04-277707
  45. 45. Manohar M, Hirsh MI, Chen Y, Woehrle T, Karande AA, Junger WG. ATP release and autocrine signaling through P2X4 receptors regulate gammadelta T cell activation. Journal of Leukocyte Biology. 2012;92:787-794. DOI: 10.1189/jlb.0312121
  46. 46. Ledderose C, Liu K, Kondo Y, Slubowski CJ, Dertnig T, Denicoló S, et al. Purinergic P2X4 receptors and mitochondrial ATP production regulate T cell migration. The Journal of Clinical Investigation. 2018;128:3583-3594. DOI: 10.1172/jci120972
  47. 47. Ledderose C, Bromberger S, Slubowski CJ, Sueyoshi K, Aytan D, Shen Y, et al. The purinergic receptor P2Y11 choreographs the polarization, mitochondrial metabolism, and migration of T lymphocytes. Science Signaling. 2020;13:651. DOI: 10.1126/scisignal.aba3300
  48. 48. Albayati S, Vemulapalli H, Tsygankov AY, Liverani E. P2Y(12) antagonism results in altered interactions between platelets and regulatory T cells during sepsis. Journal of Leukocyte Biology. 2020;110:141-153. DOI: 10.1002/jlb.3a0220-097r
  49. 49. Sakaki H, Tsukimoto M, Harada H, Moriyama Y, Kojima S. Autocrine regulation of macrophage activation via exocytosis of ATP and activation of P2Y11 receptor. PLoS One. 2013;8:e59778. DOI: 10.1371/journal.pone.0059778
  50. 50. Guerra AN, Fisette PL, Pfeiffer ZA, Quinchia-Rios BH, Prabhu U, Aga M, et al. Purinergic receptor regulation of LPS-induced signaling and pathophysiology. Journal of Endotoxin Research. 2003;9:256-263. DOI: 10.1179/096805103225001468
  51. 51. Ventura MA, Thomopoulos P. ATP and ADP activate distinct signalling pathways in human promonocyte U-937 cells differentiated with 1,25-dihydroxy-vitamin D3. Molecular Pharmacology. 1995;47:104-114
  52. 52. Satonaka H, Nagata D, Takahashi M, Kiyosue A, Myojo M, Fujita D, et al. Involvement of P2Y12 receptor in vascular smooth muscle inflammatory changes via MCP-1 upregulation and monocyte adhesion. American Journal of Physiology. Heart and Circulatory Physiology. 2015;308:H853-H861. DOI: 10.1152/ajpheart.00862.2013
  53. 53. Calovi S, Mut-Arbona P, Sperlágh B. Microglia and the purinergic Signaling system. Neuroscience. 2019;405:137-147. DOI: 10.1016/j.neuroscience.2018.12.021
  54. 54. Choi HB, Ryu JK, Kim SU, McLarnon JG. Modulation of the purinergic P2X7 receptor attenuates lipopolysaccharide-mediated microglial activation and neuronal damage in inflamed brain. The Journal of Neuroscience. 2007;27:4957-4968
  55. 55. Kobayashi K, Yamanaka H, Yanamoto F, Okubo M, Noguchi K. Multiple P2Y subtypes in spinal microglia are involved in neuropathic pain after peripheral nerve injury. Glia. 2012;60:1529e1539
  56. 56. Tozaki-Saitoh H, Miyata H, Yamashita T, Matsushita K, Tsuda M, Inoue K. P2Y12 receptors in primary microglia activate nuclear factor of activated T-cell signaling to induce C-C chemokine 3 expression. Journal of Neurochemistry. 2017;141:100-110. DOI: 10.1111/jnc.13968
  57. 57. Inoue K. UDP facilitates microglial phagocytosis through P2Y6 receptors. Cell Adhesion & Migration. 2007;1:131-132
  58. 58. Wen RX, Shen H, Huang SX, Wang LP, Li ZW, Peng P, et al. P2Y6 receptor inhibition aggravates ischemic brain injury by reducing microglial phagocytosis. CNS Neuroscience & Therapeutics. 2020;26:416-429. DOI: 10.1111/cns.13296
  59. 59. Morioka N, Tokuhara M, Harano S, Nakamura Y, Hisaoka-Nakashima K, Nakata Y. The activation of P2Y6 receptor in cultured spinal microglia induces the production of CCL2 through the MAP kinases-NF-kappaB pathway. Neuropharmacology. 2013;75:116-125. DOI: 10.1016/j.neuropharm.2013.07.017
  60. 60. Van Steenwinckel J, Reaux-Le GA, Pommier B, Mauborgne A, Dansereau MA, Kitabgi P, et al. CCL2 released from neuronal synaptic vesicles in the spinal cord is a major mediator of local inflammation and pain after peripheral nerve injury. The Journal of Neuroscience. 2011;31:5865e5875
  61. 61. Rostène W, Dansereau MA, Godefroy D, Van Steenwinckel J, Reaux-Le GA, Mélik-Parsadaniantz S, et al. Neurochemokines: A menage a trois providing new insights on the functions of chemokines in the central nervous system. Journal of Neurochemistry. 2011;118:680e694
  62. 62. Ting JP, Lovering RC, Alnemri ES, Bertin J, Boss JM, Davis BK, et al. The NLR gene family: A standard nomenclature. Immunity. 2008;28:285-287
  63. 63. Koonin EV, Aravind L. The NACHT family - a new group of predicted NTPases implicated in apoptosis and MHC transcription activation. Trends in Biochemical Sciences. 2000;25:223-224
  64. 64. Pillinger MH, Rosenthal P, Abeles AM. Hyperuricemia and gout: New insights into pathogenesis and treatment. Bulletin of the NYU Hospital for Joint Diseases. 2007;65:215-221
  65. 65. Lamkanfi M, Dixit VM. Mechanisms and functions of inflammasomes. Cell. 2014;157:1013e22
  66. 66. Hayyan M, Hashim MA, AlNashef IM. Superoxide ion: Generation and chemical implications. Chemical Reviews. 2016;116:3029-3085. DOI: 10.1021/acs.chemrev.5b00407
  67. 67. Kim HJ, Kim CH, Ryu JH, Kim MJ, Park CY, Lee JM, et al. Reactive oxygen species induce antiviral innate immune response through IFN-λ regulation in human nasal epithelial cells. American Journal of Respiratory Cell and Molecular Biology. 2013;49:855-865. DOI: 10.1165/rcmb.2013-0003OC
  68. 68. Hung SC, Choi CH, Said-Sadier N, Johnson L, Atanasova KR, Sellami H, et al. P2X4 assembles with P2X7 and pannexin-1 in gingival epithelial cells and modulates ATP-induced reactive oxygen species production and inflammasome activation. PLoS One. 2013;8:e70210
  69. 69. Martinon F, Pétrilli V, Mayor A, Tardivel A, Tschopp J. Gout-associated uric acid crystals activate the NALP3 inflammasome. Nature. 2006;440:237-241
  70. 70. Jin M, Yang F, Yang I, Yin Y, Luo JJ, Wang H, et al. Uric acid, hyperuricemia and vascular diseases. Frontiers in Bioscience. 2012;1:656-669. DOI: 10.2741/3950
  71. 71. Liu-Bryan R, Scott P, Sydlaske A, Rose DM, Terkeltaub R. Innate immunity conferred by toll-like receptors 2 and 4 and myeloid differentiation factor 88 expression is pivotal to monosodium urate monohydrate crystal-induced inflammation. Arthritis and Rheumatism. 2005;52:2936-2946. DOI: 10.1002/art.21238
  72. 72. Patterson RA, Horsley ET, Leake DS. Prooxidant and antioxidant properties of human serum ultrafiltrates toward LDL: Important role of uric acid. Journal of Lipid Research. 2003;44:512-521
  73. 73. Zhang J, Diao B, Lin X, Xu J, Tang F. TLR2 and TLR4 mediate an activation of adipose tissue renin-angiotensin system induced by uric acid. Biochimie. 2019;162:125-133. DOI: 10.1016/j.biochi.2019.04.013
  74. 74. Milanesi S, Verzola D, Cappadona F, Bonino B, Murugavel A, Pontremoli R, et al. Uric acid and angiotensin II additively promote inflammation and oxidative stress in human proximal tubule cells by activation of toll-like receptor 4. Journal of Cellular Physiology. 2019;234:10868-10876. DOI: 10.1002/jcp.27929
  75. 75. Kocyigit I, Sener EF, Taheri S, Eroglu E, Ozturk F, Unal A, et al. Toll-like receptors in the progression of autosomal dominant polycystic kidney disease. Therapeutic Apheresis and Dialysis. 2016;20:615-622. DOI: 10.1111/1744-9987.12458
  76. 76. Liu-Bryan R, Pritzker K, Firestein GS, Terkeltaub R. TLR2 signaling in chondrocytes drives calcium pyrophosphate dihydrate and monosodium urate crystal-induced nitric oxide generation. Journal of Immunology. 2005;174:5016-5023. DOI: 10.4049/jimmunol.174.8.5016
  77. 77. Grassi F. The P2X7 receptor as regulator of T cell development and function. Frontiers in Immunology. 2020;11:1179. DOI: 10.3389/fimmu.2020.01179
  78. 78. La Sala A, Sebastiani S, Ferrari D, Di Virgilio F, Idzko M, Norgauer J, et al. Dendritic cells exposed to extracellular adenosine triphosphate acquire the migratory properties of mature cells and show a reduced capacity to attract type 1 T lymphocytes. Blood. 2002;99:1715-1722
  79. 79. Kreckler LM, Gizewski E, Wan TC, Auchampach JA. Adenosine suppresses lipopolysaccharide-induced tumor necrosis factor-alpha production by murine macrophages through a protein kinase A- and exchange protein activated by cAMP-independent signaling pathway. The Journal of Pharmacology and Experimental Therapeutics. 2009;331:1051-1061. DOI: 10.1124/jpet.109.157651
  80. 80. Cunha RA. How does adenosine control neuronal dysfunction and neurodegeneration? Journal of Neurochemistry. 2016;139:1019-1055
  81. 81. Luongo L, Guida F, Imperatore R, Napolitano F, Gatta L, Cristino L, et al. The A1 adenosine receptor as a new player in microglia physiology. Glia. 2014;62:122-132. DOI: 10.1002/glia.22592
  82. 82. Savio LEB, de Andrade MP, da Silva CG, Coutinho-Silva R. The P2X7 receptor in inflammatory diseases: Angel or demon? Frontiers in Pharmacology. 2018;9:52. DOI: 10.3389/fphar.2018.00052
  83. 83. Fracasso M, Reichert K, Bottari NB, da Silva AD, Schetinger MRC, Monteiro SG, et al. Involvement of ectonucleotidases and purinergic receptor expression during acute Chagas disease in the cortex of mice treated with resveratrol and benznidazole. Purinergic Signal. 2021;17:493-502. DOI: 10.1007/s11302-021-09803-9
  84. 84. Yi C, Ezan P, Fernández P, Schmitt J, Sáez JC, Giaume C, et al. Inhibition of glial hemichannels by boldine treatment reduces neuronal suffering in a murine model of Alzheimer’s disease. Glia. 2017;65:1607-1625
  85. 85. Stone TW, Ceruti S, Abbracchio MP. Adenosine receptors and neurological disease: Neuroprotection and neurodegeneration. Handbook of Experimental Pharmacology. 2009;193:535-587. DOI: 10.1007/978-3-540-89615-9_17
  86. 86. Kusner DJ, Adams J. ATP-induced killing of virulent mycobacterium tuberculosis within human macrophages requires phospholipase D. Journal of Immunology. 2000;164:379e88
  87. 87. Kusner DJ, Barton JA. ATP stimulates human macrophages to kill intracellular virulent mycobacterium tuberculosis via calcium-dependent phagosome-lysosome fusion. Journal of Immunology. 2001;167:3308e15
  88. 88. Pettengill MA, Marques-da-Silva C, Avila ML, d'Arc dos Santos OS, Lam VW, Ollawa I, et al. Reversible inhibition of chlamydia trachomatis infection in epithelial cells due to stimulation of P2X(4) receptors. Infection and Immunity. 2012;80:4232e8
  89. 89. Coutinho-Silva R, Stahl L, Raymond MN, Jungas T, Verbeke P, Burnstock G, et al. Inhibition of chlamydial infectious activity due to P2X7R-dependent phospholipase D activation. Immunity. 2003;19:403e22
  90. 90. Almeida-da-Silva CLC, Morandini AC, Ulrich H, Ojcius DM, Coutinho-Silva R. Purinergic signaling during Porphyromonas gingivalis infection. Biomedical Journal. 2016;39:251-260. DOI: 10.1016/j.bj.2016.08.003
  91. 91. Yilmaz O, Yao L, Maeda K, Rose TM, Lewis EL, Duman M, et al. ATP scavenging by the intracellular pathogen Porphyromonas gingivalis inhibits P2X7-mediated host-cell apoptosis. Cellular Microbiology. 2008;10:863e75
  92. 92. Choi CH, Spooner R, DeGuzman J, Koutouzis T, Ojcius DM, Yilmaz O. Porphyromonas gingivalis-nucleoside-diphosphate-kinase inhibits ATP-induced reactive-oxygen-species via P2X7 receptor/NADPH-oxidase signalling and contributes to persistence. Cellular Microbiology. 2013;15:961e76
  93. 93. Johnson L, Atanasova KR, Bui PQ , Lee J, Hung SC, Yilmaz O, et al. Porphyromonas gingivalis attenuates ATP-mediated inflammasome activation and HMGB1 release through expression of a nucleoside-diphosphate kinase. Microbes and Infection. 2015;17:369e77
  94. 94. Spooner R, De Guzman J, Lee KL, Yilmaz O. Danger signal adenosine via adenosine 2a receptor stimulates growth of Porphyromonas gingivalis in primary gingival epithelial cells. Molecular Oral Microbiology. 2014;29:67e78
  95. 95. Chaves SP, Torres-Santos EC, Marques C, Figliuolo VR, Persechini PM, Coutinho-Silva R, et al. Modulation of P2X(7) purinergic receptor in macrophages by Leishmania amazonensis and its role in parasite elimination. Microbes and Infection. 2009;11:842e9
  96. 96. Marques-da-Silva C, Chaves MM, Chaves SP, Figliuolo VR, Meyer-Fernandes JR, Corte-Real S, et al. Infection with Leishmania amazonensis upregulates purinergic receptor expression and induces host-cell susceptibility to UTP-mediated apoptosis. Cellular Microbiology. 2011;13:1410e28
  97. 97. Figueiredo AB, Souza-Testasicca MC, Afonso LCC. Purinergic signaling and infection by Leishmania: A new approach to evasion of the immune response. Biomedical Journal. 2016;39:244-250. DOI: 10.1016/j.bj.2016.08.004
  98. 98. Katz O, Waitumbi JN, Zer R, Warburg A. Adenosine, AMP, and protein phosphatase activity in sandfly saliva. The American Journal of Tropical Medicine and Hygiene. 2000;62:145e50
  99. 99. Vial C, Rolf MG, Mahaut-Smith MP, Evans RJ. A study of P2X1 receptor function in murine megakaryocytes and human platelets reveals synergy with P2Y receptors. British Journal of Pharmacology. 2002;135:363e72
  100. 100. Figueiredo AB, Serafim TD, Marques-da-Silva EA, Meyer-Fernandes JR, Afonso LC. Leishmania amazonensis impairs DC function by inhibiting CD40 expression via A2B adenosine receptor activation. European Journal of Immunology. 2012;42:1203e15
  101. 101. Gomes RS, de Carvalho LC, de Souza VR, Fietto JL, Afonso LC. E-NTPDase (ecto-nucleoside triphosphate diphosphohydrolase) of Leishmania amazonensis inhibits macrophage activation. Microbes and Infection. 2015;17:295e303
  102. 102. Amit A, Kumar S, Dikhit MR, Jha PK, Singh AK, et al. Up regulation of A2B adenosine receptor on monocytes are crucially required for immune pathogenicity in Indian patients exposed to Leishmania donovani. Cytokine. 2016;79:38e44
  103. 103. Tonin AA, Da Silva AS, Casali EA, Silveira SS, Moritz CE, Camillo G, et al. Influence of infection by toxoplasma gondii on purine levels and E-ADA activity in the brain of mice experimentally infected mice. Experimental Parasitology. 2014;142:51-58. DOI: 10.1016/j.exppara.2014. 04.008
  104. 104. Gherardi A, Sarciron ME, Francoise A, Peyron F. Purine pathway enzymes in a cyst forming strain of toxoplasma gondii. Life sciences. 1999;65:1733-1738
  105. 105. Edwards FA, Gibb AJ, Colquhoun D. ATP receptor-mediated synaptic currents in the central nervous system. Nature. 1992;359:144-147
  106. 106. Correa G, da SC M, de Abreu Moreira-Souza AC, Vommaro RC, Coutinho-Silva R. Activation of the P2X(7) receptor triggers the elimination of toxoplasma gondii tachyzoites from infected macrophages. Microbes and Infection. 2010;12:497e504
  107. 107. Moreira-Souza AC, Marinho Y, Correa G, Santoro GF, Coutinho CM, Vommaro RC, et al. Pyrimidinergic receptor activation controls toxoplasma gondii infection in macrophages. PLoS One. 2015;10:e0133502
  108. 108. Mahamed DA, Mills JH, Egan EE, Denkers EY, Bynoe MS. CD73-generated adenosine facilitates toxoplasma gondii differentiation to long-lived tissue cysts in the central nervous system. Proceedings. National Academy of Sciences. United States of America. 2012;109:16312-16317
  109. 109. Thammavongsa V, Kern JW, Missiakas DM, Schneewind O. Staphylococcus aureus synthesizes adenosine to escape host immune responses. The Journal of Experimental Medicine. 2009;206:2417e27
  110. 110. Henriques-Pons A, DeMeis J, Cotta-De-Almeida V, Savino W, Araújo-Jorge TC. Fas and perforin are not required for thymus atrophy induced by Trypanosoma cruzi infection. Experimental Parasitology. 2004;107:1-4. DOI: 10.1016/j.exppara.2004.04.010
  111. 111. Santos EC, Novaes RD, Cardoso SA. Oliveira LL implication of purinergic signalling pathways in clinical management of Chagas disease. OA Biotechnology. 2013;2:27
  112. 112. Da Silva AS, Bellé LP, Bitencourt PE, Perez HA, Thomé GR, Costa MM, et al. Trypanosoma evansi: Adenosine deaminase activity in the brain of infected rats. Experimental Parasitology. 2011;127:173-177. DOI: 10.1016/j.exppara.2010.07.010
  113. 113. Ivanov A, Matsumura I. The adenosine deaminases of plasmodium vivax and plasmodium falciparum exhibit surprising differences in ligand specificity. Journal of Molecular Graphics & Modelling. 2012;35:43-48. DOI: 10.1016/j.jmgm.2012.2.004
  114. 114. Almeida-da-Silva CLC, Morandini AC, Ulrich H, Ojcius DM, Coutinho-Silva R. Purinergic signaling during Porphyromonas gingivalis infection. Biomedical Journal. 2016;39:251-260. DOI: 10.1016/j.bj.2016.08.003
  115. 115. Bisaggio DF, Peres-Sampaio CE, Meyer-Fernandes JR, Souto-Padron T. Ecto-ATPase activity on the surface of Trypanosoma cruzi and its possible role in the parasite-host cell interaction. Parasitology Research. 2003;91:273e82
  116. 116. Berredo-Pinho M, Peres-Sampaio CE, Chrispim PP, Belmont-Firpo R, Lemos AP, Martiny A, et al. A Mg-dependent ecto-ATPase in Leishmania amazonensis and its possible role in adenosine acquisition and virulence. Archives of Biochemistry and Biophysics. 2001;391:16e24
  117. 117. Yin X, Riva L, Pu Y, Martin-Sancho L, Kanamune J, Yamamoto Y, et al. MDA5 governs the innate immune response to SARS-CoV-2 in lung epithelial cells. Cell Reports. 2021;34:108628. DOI: 10.1016/j.celrep.2020.108628
  118. 118. Kindler E, Thiel V. To sense or not to sense viral RNA–essentials of coronavirus innate immune evasion. Current Opinion in Microbiology. 2014;20:69-75. DOI: 10.1016/j.mib.2014.05.005
  119. 119. Lucas C, Wong P, Klein J, Castro TBR, Silva J, Sundaram M, et al. Longitudinal analyses reveal immunological misfiring in severe COVID-19. Nature. 2020;584:463-469
  120. 120. Lohman AW, Leskov IL, Butcher JT, Johnstone SR, Stokes TA, Begandt D, et al. Pannexin 1 channels regulate leukocyte emigration through the venous endothelium during acute inflammation. Nature Communications. 2015;6:7965. DOI: 10.1038/ncomms8965
  121. 121. Seif F, Khoshmirsafa M, Aazami H, Mohsenzadegan M, Sedighi G, Bahar M. The role of JAK-STAT signaling pathway and its regulators in the fate of T helper cells. Cell Communication and Signaling: CCS. 2017;15:23. DOI: 10.1186/s12964-017-0177-y
  122. 122. Eltzschig HK, Sitkovsky MV, Robson SC. Purinergic signaling during inflammation. The New England Journal of Medicine. 2012;367:2322-2333
  123. 123. Di Virgilio F, Tang Y, Sarti AC, Rossato M. A rationale for targeting the P2X7 receptor in coronavirus disease 19 (Covid-19). British Journal of Pharmacology. 2020;177:4990-4994. DOI: 10.1111/bph.15138
  124. 124. Klaasse EC, Ijzerman AP, de Grip WJ, Beukers MW. Internalization and desensitization of adenosine receptors. Purinergic Signal. 2008;4:21-37. DOI: 10.1007/s11302-007-9086-7
  125. 125. Abdel-Magid AF. Inhibitors of CD73 may provide a treatment for Cancer and autoimmune diseases. ACS Medicinal Chemistry Letters. 2017;8:781-782. DOI: 10.1021/acsmedchemlett.7b00255
  126. 126. Nylander S, Mattsson C, Ramstrom S, Lindahl TL. Synergistic action between inhibition of P2Y12/P2Y1 and P2Y12/thrombin in ADP- and thrombin-induced human platelet activation. British Journal of Pharmacology. 2004;142:1325-1331. DOI: 10.1038/sj.bjp.0705885
  127. 127. McGonagle D, O'Donnell JS, Sharif K, Emery P, Bridgewood C. Immune mechanisms of pulmonary intravascular coagulopathy in COVID-19 pneumonia. The Lancet Rheumatology. 2020;2:e437-e445. DOI: 10.1016/s2665-9913(20)30121-1
  128. 128. Abouelkhair MA. Targeting adenosinergic pathway and adenosine A2A receptor signaling for the treatment of COVID-19: A hypothesis. Medical Hypotheses. 2020;144:110012. DOI: 10.1016/j.mehy.2020.110012
  129. 129. Vigano S, Alatzoglou D, Irving M, Ménétrier-Caux C, Caux C, Romero P. Targeting adenosine in cancer immunotherapy to enhance T-cell function. Frontiers in Immunology. 2019;10:925
  130. 130. Terp MG, Olesen KA, Arnspang EC, Lund RR, Lagerholm BC, Ditzel HJ. Anti-human CD73 monoclonal antibody inhibits metastasis formation in human breast cancer by inducing clustering and internalization of CD73 expressed on the surface of cancer cells. Journal of Immunology. 2013;191:4165-4173
  131. 131. Leone RD, Lo Y-C, Powell JD. A2aR antagonists: Next generation checkpoint blockade for cancer immunotherapy. Computational and Structural Biotechnology Journal. 2015;13:265-272

Written By

Yerly Magnolia Useche Salvador

Submitted: 20 January 2022 Reviewed: 14 March 2022 Published: 24 May 2022