Next Article in Journal
Coupling Cellular Automata and a Genetic Algorithm to Generate a Vibrant Urban Form—A Case Study of Wuhan, China
Previous Article in Journal
Challenges and Research Priorities for Dementia Care in Malaysia from the Perspective of Health and Allied Health Professionals
Previous Article in Special Issue
Evidence for Multiple Origins of De Novo Formed Vascular Smooth Muscle Cells in Pulmonary Hypertension: Challenging the Dominant Model of Pre-Existing Smooth Muscle Expansion
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of the Purinergic P2Y2 Receptor in Pulmonary Hypertension

Excellence Cluster Cardio-Pulmonary Institute (CPI), Universities of Giessen and Marburg Lung Center (UGMLC), German Center for Lung Research (DZL), Justus Liebig University of Giessen, Aulweg 130, 35392 Giessen, Germany
*
Author to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2021, 18(21), 11009; https://doi.org/10.3390/ijerph182111009
Submission received: 2 September 2021 / Revised: 13 October 2021 / Accepted: 15 October 2021 / Published: 20 October 2021

Abstract

:
Pulmonary arterial hypertension (PAH), group 1 pulmonary hypertension (PH), is a fatal disease that is characterized by vasoconstriction, increased pressure in the pulmonary arteries, and right heart failure. PAH can be described by abnormal vascular remodeling, hyperproliferation in the vasculature, endothelial cell dysfunction, and vascular tone dysregulation. The disease pathomechanisms, however, are as yet not fully understood at the molecular level. Purinergic receptors P2Y within the G-protein-coupled receptor family play a major role in fluid shear stress transduction, proliferation, migration, and vascular tone regulation in systemic circulation, but less is known about their contribution in PAH. Hence, studies that focus on purinergic signaling are of great importance for the identification of new therapeutic targets in PAH. Interestingly, the role of P2Y2 receptors has not yet been sufficiently studied in PAH, whereas the relevance of other P2Ys as drug targets for PAH was shown using specific agonists or antagonists. In this review, we will shed light on P2Y receptors and focus more on the P2Y2 receptor as a potential novel player in PAH and as a new therapeutic target for disease management.

1. Introduction

Pulmonary hypertension (PH) represents a heterogeneous group of clinical entities, defined as a mean pulmonary artery pressure above 20 mmHg [1]. Pulmonary arterial hypertension (PAH) is a fatal disease, which belongs to group 1 PH and is defined as an elevation of mean pulmonary arterial pressure >20 mm Hg, with pulmonary arterial wedge pressure <15 mm Hg and pulmonary vascular resistance >3 Wood units at rest [2,3]. It is characterized by persistent vasoconstriction, thickening, and progressive obstructive remodeling of the pulmonary arteries. Hyperproliferation of pulmonary arterial smooth muscle cells (PASMCs), endothelial cells (PAECs), and microvascular endothelial cells plays an important role in pulmonary vascular remodeling [4].
The elevated pulmonary artery pressure in PAH is an outcome of pathological processes that involve the purinergic receptors P2Y, as a member of the G protein–coupled receptors family (GPCRs) [5,6,7,8,9,10,11,12,13]. Crucial mediators of vasodilation, such as nitric oxide (NO), are released from endothelial cells in response to elevated blood pressure and fluid shear stress (FSS) forces [14,15,16,17,18,19,20]. Hence, the endothelial dysfunction, in this context, can be defined as a decreased secretion of vasodilators and an increased secretion of vasoconstrictors, which contribute, along with other factors, to the vasoconstriction and pulmonary resistance increase [12,21,22,23,24]. Recent studies have demonstrated a purinergic receptor P2Y2 involvement in NO secretion, which results in systemic vasodilation, and suggested the P2Y2 receptor transmitted the FSS signaling in order to keep NO secretion at physiological levels [25].
The lungs serve as a reservoir for numerous cell types, which express purinergic receptors. The purinergic receptors consist of two subfamilies: purinergic receptor 1 (P1), and purinergic receptor 2 (P2) (Figure 1). The P1 purinoreceptors are known as adenosine receptors and include four subtypes: A1R, A2AR, A2BR, and A3R [5]. The P2 purinoreceptors are categorized into two major sub-families: P2X, ligand-gated ionotropic membrane cation channels, and P2Y, metabotropic membrane bound, GPCRs (Figure 1) [26,27]. The P2Y receptors family comprises eight receptors (P2Y1, P2Y2, P2Y4, P2Y6, P2Y11, P2Y12, P2Y13, and P2Y14), with a high level of sequence similarity in transmembrane domains (especially P2Y1, P2Y2, P2Y4, and P2Y6), which does not easily allow synthetically developing specific ligands or pharmacologically discriminating one from other P2Y members, and forms an obstacle to selectively targeting one member of the P2Y family. Nevertheless, there are several structural homoplasies at the intracellular loops and the carboxy-terminus that make up the diversity and allow interactions with different G-proteins [28].
Alterations in purinergic signaling, as one of the potential mechanisms underlying PAH development, was recently reviewed by Cai et al. [5].The focus of this review article is the role of the P2Y2 receptor in PAH. The P2Y2 receptor is known to be associated with other diseases, such as dry eye syndrome [29,30], accommodative spasm [31], and cystic fibrosis [32,33]. This has stimulated many researchers to pharmacologically target P2Y2 to develop new treatment approaches for these diseases. Diquafosol tetrasodium, a uridine-5′-triphosphate (UTP) analogue and P2Y2 agonist that stimulates ophthalmic secretions has been approved as a therapeutic option for dry eye syndrome and accommodative spasm in some Asian countries [31,34]. Similarly, Denufosol (INS37217), another P2Y2 agonist, has reached phase III clinical trials to treat cystic fibrosis patients. Inhaled Denufosol restored sodium and chloride exchange and improved the airway clearance by enhancing ciliary beat frequency in patients with cystic fibrosis and normal to mildly impaired lung function [35,36,37]. Little is known, so far, about the potential role of P2Y2 receptors in PH, but this is still in the growth phase, and more information and new knowledge will, hopefully, emerge in the next few years.

2. P2Y2 Expression in Pulmonary Parenchyma, Vasculature, and Inflammatory Cells

P2Y2 has been demonstrated to be expressed on the protein level in primary rat type I alveolar epithelial cells (type I AECs) [38]. P2Y2 mRNA and protein expression were also noted, both in primary isolated rat type II AECs, and in immortalized tumor-derived AEC lines (L2, R3/1, RLE, and A549 cells) [39]. The epithelial expression of P2Y receptors in the alveolus is critically important for the surfactant secretion and the regulation of the alveolar surface liquid volume. P2Y2 activation in type II AECs in the alveolus occurs in response to stretch-induced adenosine-5′-triphosphate (ATP) release, which later, via subsequent activation of P2X4 receptors, results in surfactant release [40].
P2Y receptors were found in the pulmonary arteries of various species. Thus, mRNA expression of P2Y1, P2Y2, P2Y4, P2Y6, and P2Y12 receptors was demonstrated in rat pulmonary arteries [41,42,43]. Furthermore, mRNA expression of P2Y1, P2Y2, P2Y4, P2Y6, and P2Y12 was shown in mouse lung homogenates [44]. P2Y2 transcripts, concomitantly to the other members of the P2Y family, such as P2Y1 and P2Y4, were detected in the pulmonary arteries, peripheral lung tissues, and more specifically in PAECs of juvenile rabbits (Table 1) [45]. The P2Y receptors are broadly expressed in the endothelial cells of the pulmonary vasculature, suggesting their crucial importance for endothelial cell dysfunction, a pathophysiological process characterized by upregulation of adhesion molecules, increased chemokine secretion, leukocyte adherence, increased permeability, and vasoconstriction. The mRNA of P2Y2 was shown to be present in human microvascular endothelial cells from lung and PAECs [46,47,48]. Along with P2Y2, the expression of other P2Y receptors, such as P2Y1, P2Y6, and P2Y11 was detected in PAECs isolated from PAH patients and control human lungs [49,50,51].
There is accumulating evidence of the involvement of inflammation in the pathogenesis of PAH [56,57,58]. Pulmonary vascular lesions in PAH are associated with perivascular inflammation and inflammatory cell infiltration [56,59]. This pulmonary vascular inflammation in PAH has been shown to contribute to progressive pulmonary vascular remodeling [59]. In this regard, it is interesting to note that P2Y2 receptor expression was also revealed in the cells of the innate immune system, such as microphages and neutrophils. P2Y2 receptor transcripts were detected in human [60] and rat alveolar macrophages [61]. Furthermore, P2Y2 expression was also confirmed in human and rat neutrophils and eosinophils [62,63,64,65]. Macrophage-derived P2Y2 receptors sense the ATP and UTP released by apoptotic cells during immune responses to microbial infection. P2Y2 activation on alveolar macrophages results in an increase of intracellular Ca2+, which facilitates phagocytosis [40].

3. Deletion, Overexpression, and Disorders of P2Y2

3.1. Global P2Y2 Knockout

Compared to wild type (WT) mice, P2Y2 knockout mice (P2Y2-KO) showed no marked changes in body weight, glucose plasma concentrations, plasma osmolality, blood or urine pH, food and fluid intake, total fecal mass, or urinary fluid excretion, but P2Y2-KO mice exhibited lower plasma concentrations of aldosterone, renin, hematocrit, and K+ [66,67]. Global P2Y2 inactivation impaired the purinergic and Cl- secretory responses in the airways (trachea), gallbladder, and intestines [55,68]. A comparison of P2Y2-KO to WT mice showed the most effects in trachea, and to lesser extent in the gallbladder and intestine [55,68,69]. The phenotype of P2Y2-KO mice suggests that the biological function of P2Y2 can be compensated by P2Y4 and P2Y6 receptors in other organs [55,68]. Further insights gained from these mice suggest that epithelial P2Y2 in trachea interacts with extracellular ATP to trigger Ca2+ mobilization, and, therefore, it is necessary for clearance responses, such as mucin, water secretion, and ciliary beat frequency [33,35,70,71].
In another study, double knockout (P2Y1/P2Y2) mice displayed 1.8-fold lower survival rates in comparison to WT mice, when exposed to an intra-tracheal instillation of Pseudomonas aeruginosa [72]. Thus, P2Y2 is suggested, together with P2Y1, to play a protective immunological role and have a pro-inflammatory cytokine response in bacterial infection [72], pneumonia virus infection [73], and to regulate VCAM1 in eosinophils during lung inflammation [74]. Neutrophils from P2Y2-KO mice exhibited loss of sensing, orientation, and chemotaxis, mediated by A3 and P2Y2 receptors [65,75]. P2Y2 also serves as a sensor for the ATP released by apoptotic cells to promote phagocytic clearance, which was decreased in P2Y2-KO mice. Thus, P2Y2 receptor is considered as a sensor for the critical nucleotides released by apoptotic cells as a ‘find-me’ object for phagocytosis [76].
Compared to WT mice, P2Y2-KO mice displayed a larger infarct size and worse heart function in a myocardial infarction model, and in vivo treatment with MRS2768, a P2Y2 specific agonist, protected the heart from ischemic damage [77,78] and prevented vascular calcification [79,80]. Mice lacking P2Y2 developed salt-resistant systemic hypertension [81] and impaired ATP- and ATPyS-evoked relaxation in aorta compared to WT mice, suggesting the importance of this receptor for vascular signaling [82].

3.2. Conditional (Inducible) Knockout of P2Y2

Conditional P2Y2 knockout (P2Y2-CKO) mice were generated by crossing floxed P2Y2 with endothelial Cre-Tie2 mice to specifically target endothelial P2Y2. Endothelial P2Y2-CKO mice exhibited increased vascular tone and elevated blood pressure [25]. Furthermore, endothelial P2Y2 was shown to be essential for mechanotransduction and FSS effects [25].
Endothelial P2Y2-CKO was suggested to promote plaque stability in atherosclerosis ApoE-KO mouse model through reduced macrophage infiltration and matrix metalloproteinase-2, which increases smooth muscle cell migration [83].

3.3. P2Y2 Overexpression

Transgenic rats overexpressing P2Y2 by implementing a lentiviral vector [84] exhibit dramatically accelerated neointimal hyperplasia when subjected to femoral artery injury, which gives clues about the inflammatory role of P2Y2 [85].
Another in vivo study showed that hypoxia-inducible factor-1α (Hif1α) activates transcriptional expression of P2Y2 in human primary renal tubular cells [86]. These findings were further confirmed in another study that showed that HIF-1α upregulated P2Y2 expression at mRNA and protein levels and, thereby, prolonged the viability of human hepatocellular carcinoma (HHC), whereas treatment of HCC with a selective P2Y2 antagonist MRS2312 led to a reduction in cell viability [87].
FSS was associated with an increment of P2Y2-mRNA levels in HUVECs after 6 h of exposure. In HUVECs transiently expressing the P2Y2 Arg-Gly-Glu mutant receptors, FFS effects altered cell-alignment, actin stress-fiber formation, phosphorylation of focal adhesion kinase, cofilin-1, and wound repair and healing [88,89].
Finally, P2Y2 mRNA has been shown to be significantly upregulated 2 days post spinal cord injury in rats, suggesting a P2Y2 responsiveness to altered conditions and mechanical injury [90].

4. P2Y2 Signaling

P2Y2-mediated signaling pathways are summarized in Figure 2. P2Y2, like other GPCRs, is activated by extracellular stimuli. Conformational changes of P2Y2 receptors, therefore, induce intracellular signaling by activating heterotrimeric G-proteins, amongst other mechanisms. The activation of P2Y2 initiates a guanine nucleotide exchange factor (GEF), which leads to activation of subunit(s) of the coupled G-proteins. The activation/phosphorylation of one of the G-protein subunits (α, β, and γ) leads to its dissociation from the heterotrimeric complex, to start, in turn, a set of subsequent reactions and multiple effector proteins [3].
G-proteins comprise three main families: alpha (α), beta (β), and gamma (γ). Each family consists of subfamilies, e.g., the α subunit (16 subunits) forms four main subfamilies Gαq, Gαs, Gαi/o, and Gα12/13. Similarly, β and γ form two subfamilies, of 5 and 13 subunits, respectively. While β and γ have not been sufficiently studied regarding their interactions with P2Y2, other α subunits such as Gαq, Gαo, and Gα12, but not Gα12/13, are known to interact with P2Y2, due to their presence, distribution, and tissue specificity. Interestingly, researchers believe that G-proteins and GPCRs can elicit their actions and activate signaling pathways apart from each other; the so-called canonical and non-canonical signaling of GPCRs [102,103,104,105]. Furthermore, it has been shown in astrocytoma cells that, contrary to Gq, coupling of P2Y2 to Go and G12 requires an interaction with alpha v integrin [91,92].
Activation of Gαq leads to catalyzation of phosphatidylinositol 4,5-bisphosphate by phospholipase C-β (PLCβ) into inositol trisphosphate (IP3) and diacylglycerol (DAG). Both IP3 and DAG activate the intracellular Ca2+ channels in the endoplasmic reticulum (ER) and Ca2+ ions released into the cytosol in pulmonary arterial vasa vasorum endothelial cells and pulmonary fibroblasts [8,55]. The binding of Ca2+ to calmodulin (CaM) activates calmodulin-dependent kinase, which is known to phosphorylate endothelial nitric oxide synthase (eNOS) to generate more NO. DAG, however, can activate the protein kinase C, which is also necessary for eNOS and extracellular signal-regulated kinases 1/2 (Erk1/2) phosphorylation [106].
Furthermore, P2Y2 activates growth factor receptors or receptor tyrosine kinases, as well as non-receptor tyrosine kinases, such as Src [107]. Src is transiently associated and activated by P2Y2 via SH3-binding domains in the C-terminal tail of Gq-coupled GPCR [107]. Activation of Src, via tyrosine kinase proline-rich tyrosine kinase 2 (Pyk2), leads to phosphorylation of several growth factor receptors, including EGFR and VEGFR-2. The latter results in endothelial-leukocytes interactions and attachment of the leukocytes to the vessel wall [108].
Pyk2 plays a crucial role in the pathogenesis of PH. It is, however, important for hypoxia-induced proliferation and migration of human PASMCs. Furthermore, it is essential for reactive oxygen species (ROS) generation during hypoxic stress and the activation of Hif1α [109]. In addition, Src-phosphorylation leads to activation of serine/threonine-kinases such as ERK1/2, JNK, and p38, in addition to various transcription factors that are involved in inflammation, apoptosis, and cell differentiation (e.g., CREB, NF-kB, ELK1, c-FOS, c-JUN). Moreover, Src also activates the monomeric G-protein Rac1 and the platelet endothelial cell adhesion molecule-1 (PECAM-1), as well as phosphatidylinositol 3-kinase (PI3K), which phosphorylates protein kinase B (Akt), and which can also activate eNOS and enhance endothelial NO release [25,106].
Studies have shown P2Y2 to mediate phosphorylation of VE-cadherin via Src, Rac1, and VEGFR-2 [108,110]. The latter is essential for the endothelial regulation of angiogenesis, blood vessel permeability, and leukocyte trafficking [110].
In systemic circulation, Wang et al. demonstrated that Gq/G11-coupled to P2Y2 activates a mechanosensory complex consisting of PECAM-1, VE-cadherin, and VEGFR-2 in response to FSS. Consequently, and in response to FSS, eNOS is activated in a Ca2+-calmodulin-dependent manner, but also in a Ca2+-calmodulin-independent manner through Akt in pulmonary artery adventitial vasa vasorum endothelial cells [8]. The generation of NO in the vessel wall leads to vasodilatation in smooth muscle cells [25].
Other studies also referred to the NO-independent-generation of Ca2+-calmodulin cascades [111]. They suggested an endothelial-derived hyperpolarization response via mechanisms involving potassium channels (KCa3.1) and connexin 37 (Cx37) [112]. Furthermore, P2Y2 signaling cascades were shown in small pulmonary veins, which were found to contract strongly upon P2Y2 activation in vascular smooth muscle cells as a result of IP3/DAG activation of Ca2+ channels in ER and cytosolic Ca2+ release [6].
In general, the activation of GPCRs can result in receptor desensitization, mediated by a family of GPCR kinases (GRK1-7), which supports the binding of β-arrestins to the receptor [113,114,115,116]. β-arrestins then inhibit the interaction between GPCRs and G-proteins (desensitization), promote internalization of the GPCRs into clathrin-coated pits (endocytosis), and interact, for example, with MAPK/Src signaling in rat aortic smooth muscle cells (ASMCs) [115]. The desensitization of GPCRs does not only lead to less activation by agonists, but also to their internalization [3,117]. Furthermore, GPCR desensitization is crucial for the sufficient activation of store-operated calcium release activated calcium channels, which leads to an influx of extracellular calcium [116]. In rat mesenteric arterial smooth muscle cells, β-arrestin 2 and GRK2 mediate the P2Y2 receptor desensitization [114,115,118,119]. Apart from GPCRs and G-proteins, canonical, and noncanonical signaling, β-arrestins have been suggested to be, at least partially, responsible for the P2Y2-mediated MAPKs and ERK1/2 activation, as reported in pulmonary artery adventitial vasa vasorum endothelial cells and ASMCs [8,113,114,115].
So far, no specific signaling cascades have been described for P2Y2 in PAECs or PASMCs, but various P2Y2 signaling cascades were linked to inflammatory responses and pathogenesis of atherosclerosis and other cardiovascular diseases. Thus, studies have suggested that activation of endothelial P2Y2 is involved in oxLDL-mediated inflammasome activation and subsequent IL-1β production, through the modulation of mtROS-mtDNA and TLR9-NFkB signaling pathways in human endothelial cells [120]. Other studies, however, showed that P2Y2 in vascular smooth muscle cells activates Nox1 via Rac, which leads to ROS generation and, thus, to oxidative stress, and that P2Y2 mediates proinflammatory signaling by monocyte chemo-attractant protein 1 formation, leading to an atherogenesis response [121].
Finally, activation of Gο through P2Y2 is reported to mediate the activation of RhoA [91,108], whereas activation of Gα12 through P2Y2 is shown to activate Rac, RacGEF, and Vav2, which play a significant role in cell migration [91,110]. This suggests multiple signaling cascades and various responses that can be triggered by P2Y2, due to the broad distribution in organs and tissues [3]. Most of these signaling events have not yet been confirmed in the pulmonary vascular system and parenchyma, which prompts us towards more investigations, to clarify P2Y2 signaling in the pulmonary vasculature.

5. Physiological Function and Dysfunction of P2Y2

P2Y2 is associated with several diseases such as dry eye syndrome [29,30,122,123], accommodative spasm [123], and cystic fibrosis [32,33,35,36,124,125]. This illuminates the physiological importance of P2Y2 in different organs and biological systems, which can be summarized as follows:

5.1. Control of Vascular Tone (Vasorelaxation or Vasoconstriction)

Upregulation of vasoconstrictive mediators in PAECs and PASMCs, such as endothelin-1 and endothelin receptor A (ETA), which is accompanied by downregulation of vasodilatory mediators such as NO, prostacyclin (PGI2), endothelium-derived hyperpolarizing factor (EDHF), or prostaglandin receptors (DPs, EPs, FP, IP, or TP), is one example of the imbalance of GPCR-expression that contributes to the increased pulmonary vascular resistance in PAH patients [126]. Moreover, this imbalance in endothelial secretion systems is accompanied by disturbed G-protein expression in both PAECs and PASMCs [45,127,128]. Endothelial P2Y2 was shown to trigger vasodilatory function in response to increased FSS in systemic circulation [25]. In contrast, ATP and UTP were reported to play a vasoconstrictive role, probably via P2Y2, in intrapulmonary arteries isolated from normal and pulmonary hypertensive newborn piglets [12]. Mechanical injury, FSS, and hypoxia can also upregulate P2Y2, adding more complexity to the vascular control system [86,87,88,90,129,130,131,132]. It should be noted that ATP and ADP are non-selective P2Y2 receptor agonists, and therefore, more studies are required to investigate and clarify any vasodilatory or vasoconstrictive role of endothelial P2Y2 [133]. However, using PAECs- and PASMCs-cell specific P2Y2 knockout mice, and the application of the only available novel selective P2Y2 receptor agonist MRS2768, can provide more insights into the role of P2Y2 in pulmonary vascular regulation [134,135,136,137].
Endothelial P2Y2 coupled to Gαq/Gα11 is involved in systemic hypertension via transmission of FSS-induced phosphorylation of Akt, which, in turn, phosphorylates eNOS [25,138]. Under FSS forces, a variety of vasodilatory mediators, such as NO, PGI2, and EDHF, are released by endothelial cells in the systemic circulation [139,140,141,142,143,144]. NO produced by eNOS plays a potent role in flow-induced vasorelaxation [83,145]. Activation of P2Y2 results in decreased blood pressure through regulation of renal Na+, potassium channel, and connexin37 [112,146].

5.2. Mechanotransduction of FSS

Similar to other cell types, endothelial cells release ATP via the pannexin1 channel in response to FSS, which activates the purinergic P2Y2 receptor (Figure 3) [145,147]. Furthermore, P2Y2 is a mechanosensitive channel, located downstream of Piezo1, suggesting a complicated signaling that ends in enhanced eNOS phosphorylation and NO release [99,145]. Hence, P2Y2 controls vascular tone and systematic blood pressure via NO, which, in turn, stimulates soluble guanylyl cyclase in vascular smooth muscle cells, resulting in increased cyclic GMP concentrations and vasodilatation [111,142,148,149,150]. In addition, under FSS conditions, there is P2Y2 upregulation [88,129,130,131,132].

5.3. Involvement in Phagocytic Clearance

Injury to pulmonary vascular cells is one of the mechanisms involved in PAH initiation and progression [155]. Apoptosis is an important regulator of tissue integrity and homeostasis, removing damaged or non-functional cells. However, it causes cellular damage [155]. There is evidence suggesting that the mismatch between apoptosis and apoptotic cell clearance underlies various pathologic conditions, including PAH [155,156].
The ATP release by different cell types, which is a part of the vascular tone regulation mechanisms, has another biological function. It participates in the phagocytosis of apoptotic cells and involves inflammatory cells such as monocytes, dendritic, macrophages, eosinophils, and neutrophils in the circulation system via inflammatory P2Y2 sensing [60,76,156,157]. Thus, ATP release from different apoptotic cells in pulmonary tissues, which is sensed by mast cells, eosinophils, dendritic cells, and other inflammatory cell types, suggests a significant inflammatory clearance role for ATP, UTP, ADP, and P2Y2 receptors in asthma, chronic obstructive pulmonary disease, cystic fibrosis, and other pulmonary diseases [30,32,33,35,37,70,73,76,125,128,158,159,160,161,162].

5.4. Proliferation/Anti-Apoptosis and Vascular Remodeling

Hyper-proliferation of PASMCs and PAECs results in lumen narrowing of the pulmonary arteries and increased pulmonary arterial pressure. The molecular mechanism involves a number of GPCRs, such as AT1, ETA, ETB, 5-HT1B, 5HT2A, 5HT2B, and PAR1-3, in addition to the G-proteins Gq/G11 and G12/G13 [126]. P2Y2 mediated increased proliferation of breast cancer cells [163], human cancerous pancreatic duct epithelial cells [164], rabbit corneal endothelial cells [93], hepatocytes [165]. Furthermore, P2Y2 promoted vascular endothelial and smooth muscle cell proliferation [3,85,93,97]. The role of P2Y receptors in vascular homeostasis was established by studies on systemic vasculature. Thus, P2Y2 effects on the proliferation of PAECs and PASMCs have not yet been investigated.

5.5. Pulmonary Angiogenesis

The formation of new blood vessels in the cardio-pulmonary system from the preexisting vasculature plays a vital role in maintaining cardio-pulmonary vascularization and perfusion in various physiological and pathological conditions [105,166,167,168,169,170,171,172,173]. On the other hand, impaired angiogenesis has fatal consequences through a severe reduction or elevation of cardiac capillary density that contributes to dysfunction and heart failure [172,174,175,176]. P2Y2 has been suggested to affect endothelial cells, by promoting sprouting, angiogenesis, and vasculogenesis, and to protect the heart from ischemic damage [78,177].
There is experimental evidence supporting an important role of P2Y receptors in vasa vasorum neovascularization and pathologic vascular remodeling in PH [178]. Furthermore, in plexiform lesions from PAH patients, expression of CD39/ENTPD1, the ectonucleotidase responsible for the conversion of the nucleotides ATP and ADP to AMP, was decreased, suggesting that increased extracellular ATP level in the pulmonary vascular endothelium and within the angiomatoid proliferative lesions might contribute to excessive endothelial proliferation [44].

5.6. Blood Lung Barrier Dysfunction and Permeability of PAECs

In various experimental PH models, the initial increase in lung vascular leakage strongly correlated with the pulmonary pressure elevation [155]. P2Y2 was suggested to mediate the enhancement of adherent junctions and blood vessel permeability via ATP, Rac1, VEGFR-2, and VE-cadherin [110]. P2Y2, hence, significantly increased the trans-endothelial resistance in PAECs, which is of significant clinical relevance in terms of vascular permeability [46,48,179,180,181]. Tight junctions between endothelial cells are directly influenced, by increasing the gap formation through the actions of eNOS [179]. PAECs have not yet been sufficiently investigated concerning the stability of pulmonary endothelial tight junction formation and vascular permeability.

5.7. P2Y2 in Pulmonary Hypertension

The role of P2Y receptors in the pulmonary vasculature was investigated in several PH models. Significant upregulation of P2Y2 expression was revealed in the lungs of mice exposed to hypoxia [44]. Expression of P2Y2 along with other P2Y receptors, such as P2Y1, P2Y6, and P2Y11, was detected in PAECs isolated from PAH patients [49,50,51].
P2Y2 endogenous agonists such as ATP, ADP, and adenosine were claimed to be effective pulmonary vasodilators, possibly via P2Y2 receptors, in various animal PH models [182,183,184,185]. Likewise, ATP-MgCl2 infusion reduced mean pulmonary artery pressure and pulmonary vascular resistance in piglets with hypoxia-induced PH [186] and in children with PH secondary to congenital heart disease [187]. Furthermore, in pigs with acute hypoxia-induced PH, both P2Y1 inhibition with the selective antagonist MRS2500, and P2Y12R inhibition with cangrelor decreased pulmonary artery pressure [50].

6. Pharmacological Importance of P2Y2 as a Potential PH Therapy

The natural endogenous ligands for P2Y receptors (Table 2) are the extracellular purine and pyrimidine mononucleotides, such as ATP and UTP [68,188,189,190], and dinucleotides, such as diadenosine polyphosphates [189,190,191]. These nucleotides are normally located in the cytosol. Under certain conditions, such as cellular damage or apoptosis, these nucleotides can be released outside of the cellular compartment and, thus, bind to P2Y receptors. The latter leads to the argument that P2Y receptors and their natural ligands trigger clearance signaling [192].

6.1. P2Y2 Agonists

The natural endogenous ligands for P2Y2 receptors, as mentioned before, are ATP/UTP and other less potent oligonucleotides (Table 2) [133]. Synthetic selective P2Y2 agonists are difficult to generate due to the similarities of the sequences between P2Y receptors [193,194]. MRS2768 is the only selective P2Y2 agonist that does not show any interactions with other human P2Y receptors, such as P2Y4 or P2Y6 [133]. The best-known P2Y2 selective agonist “MRS2768” has an EC50 = 1.89 µM [133,195]. Another available drug, PSB1114, is a potent P2Y2 agonist that displays selectivity for P2Y4 and P2Y6 receptors, where the concentration is increased up to 60-fold over EC50 [194]. Although Diquafosol, also known as Diquas, is considered a selective agonist that activates P2Y2, it shows affinity to the P2Y4 receptor [193]. Diquafosol was approved for use in Japan in 2010 to treat dry eye disease [123,193]. The P2Y2 affinity order decreases from UTP = ATP > ATP-gamma-S >> 2-methylthio-ATP = ADP [196].
Evaluations of studies that involve vascular and pulmonary P2Y2 and its ligands are of pharmacological importance. Targeting the ocular P2Y2 with INS365 (Diquafosol eye drop) in a rat dry eye model promoted aqueous tear secretion in animals that underwent surgical excision of their lacrimal glands [180]. Similarly, in vivo studies have also demonstrated that P2Y2 agonists UTP and ATP increased tear secretion in rabbit [197,198] and regulated ocular mucin production in isolated rabbit and human conjunctiva [199]. Activation of P2Y2/4 receptors with INS45973, a UTP analog, resulted in lower blood pressure via distinct mechanisms, involving potassium/calcium activated channel KCa3.1 and connexin Cx37 [112].
Additionally, the P2Y2 agonists UTP and diadenosine tetraphosphate mediated in vivo arteriogenesis in a murine model of hindlimb ischemia [177]. Denufosol and other P2Y2 agonists improved lung function in cystic fibrosis patients [35,36,37,200,201,202,203]. Interestingly, very few studies have shown a protective effect of the specific P2Y2 agonist MRS2768 on hypoxic cardiomyocytes [78]. Thus, treatment with MRS2768 enhanced stability and protected the heart from ischemic damage in vitro and in vivo [78]. MRS2768 induced vasoconstriction of porcine pancreatic arteries via P2Y2, whereas ADP induced vasorelaxation [134].

6.2. P2Y2 Antagonists

A few inhibitors that target P2Y2 receptors are currently commercially available. AR-C118925XX was considered a selective and competitive P2Y2 receptor antagonist (Table 2) [204,205]. This antagonist did not display any interactions with the group of other receptors that interact with Gq at 10 µM concentrations, taking into consideration that the IC50 = 1 µM [195,206]. AR-C126313 and PSB-416 have been shown to selectively target P2Y2 [207,208].

7. Conclusions

P2Y2 is expressed, at least at mRNA levels, in all different layers of the pulmonary vasculature (adventitia, media, and intima). Deletion, overexpression, or pharmacological targeting of P2Y2 receptors leads to various biological organ and cell type specific alterations. Exposing P2Y2-expressing cells to mechanical stress, FSS, or hypoxia significantly increases the transcriptional expression of P2Y2, which, in turn, alters the physiological function. P2Y2 is known to play a crucial role in controlling systemic vascular tone, the proliferation and remodeling of vascular cells, chemotaxis sensing, airway clearance, and angiogenesis.
Despite the fact that more than 80% of newly developed medications partially or entirely target GPCRs and despite the presence of GPCR-based PAH therapies, including bosentan, macitentan, and iloprost, a sufficient GPCR-associated therapy for PAH has not yet been developed. The aim of our review was to shed light on the potential pathological involvement of the P2Y2 receptor in PAH development and the underlying mechanisms.
Investigation of P2Y2, as a mechanosensitive GPCR and a crucial player in many biological processes, in different pulmonary vascular cell-types will help identify rationales for administration of its ligands as a potentially novel pharmacological approach in PAH treatment. Studies of the application of selective P2Y2 (MRS2768) to activate mechanosensing pathways in vitro, ex vivo, and in vivo will increase our understanding of the involvement of P2Y2 in PAH pathomechanisms and offer new prospects and approaches for overcoming the disease.

Author Contributions

Conceptualization, M.S., T.N. and R.T.S.; writing—original draft preparation, M.S., T.L. and T.N.; writing—review and editing, M.S., T.N., T.L., A.S. and R.T.S.; funding acquisition, R.T.S.; supervision, M.S. and R.T.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by German Research Foundation (DFG), grant number 268555672—CRC1213 (Collaborative Research Center 1213), Projects A08 and B04.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank Julia Stockburger, Laura Klos, Christine Vroom, Gayathri Viswanathan, and Andreas Hecker for their support. Some figures in this manuscript (Figure 2 and Figure 3) were prepared using the Motifolio drawings.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Humbert, M.; Galie, N.; McLaughlin, V.V.; Rubin, L.J.; Simonneau, G. An insider view on the World Symposium on Pulmonary Hypertension. Lancet Respir. Med. 2019, 7, 484–485. [Google Scholar] [CrossRef] [Green Version]
  2. Simonneau, G.; Montani, D.; Celermajer, D.S.; Denton, C.P.; Gatzoulis, M.A.; Krowka, M.; Williams, P.G.; Souza, R. Haemodynamic definitions and updated clinical classification of pulmonary hypertension. Eur. Respir. J. 2019, 53, 1801913. [Google Scholar] [CrossRef]
  3. Erb, L.; Weisman, G.A. Coupling of P2Y receptors to G proteins and other signaling pathways. Wiley Interdiscip. Rev. Membr. Transp. Signal 2012, 1, 789–803. [Google Scholar] [CrossRef] [Green Version]
  4. Schermuly, R.T.; Ghofrani, H.A.; Wilkins, M.R.; Grimminger, F. Mechanisms of disease: Pulmonary arterial hypertension. Nat. Rev. Cardiol. 2011, 8, 443–455. [Google Scholar] [CrossRef] [PubMed]
  5. Cai, Z.; Tu, L.; Guignabert, C.; Merkus, D.; Zhou, Z. Purinergic Dysfunction in Pulmonary Arterial Hypertension. J. Am. Heart Assoc. 2020, 9, e017404. [Google Scholar] [CrossRef] [PubMed]
  6. Henriquez, M.; Fonseca, M.; Perez-Zoghbi, J.F. Purinergic receptor stimulation induces calcium oscillations and smooth muscle contraction in small pulmonary veins. J. Physiol. 2018, 596, 2491–2506. [Google Scholar] [CrossRef] [PubMed]
  7. Huang, C.; Hu, J.; Subedi, K.P.; Lin, A.H.; Paudel, O.; Ran, P.; Sham, J.S. Extracellular Adenosine Diphosphate Ribose Mobilizes Intracellular Ca2+ via Purinergic-Dependent Ca2+ Pathways in Rat Pulmonary Artery Smooth Muscle Cells. Cell. Physiol. Biochem. 2015, 37, 2043–2059. [Google Scholar] [CrossRef]
  8. Lyubchenko, T.; Woodward, H.; Veo, K.D.; Burns, N.; Nijmeh, H.; Liubchenko, G.A.; Stenmark, K.R.; Gerasimovskaya, E.V. P2Y1 and P2Y13 purinergic receptors mediate Ca2+ signaling and proliferative responses in pulmonary artery vasa vasorum endothelial cells. Am. J. Physiol. Cell Physiol. 2011, 300, C266–C275. [Google Scholar] [CrossRef] [Green Version]
  9. Giannattasio, G.; Ohta, S.; Boyce, J.R.; Xing, W.; Balestrieri, B.; Boyce, J.A. The purinergic G protein-coupled receptor 6 inhibits effector T cell activation in allergic pulmonary inflammation. J. Immunol. 2011, 187, 1486–1495. [Google Scholar] [CrossRef] [Green Version]
  10. De Proost, I.; Pintelon, I.; Wilkinson, W.J.; Goethals, S.; Brouns, I.; Van Nassauw, L.; Riccardi, D.; Timmermans, J.P.; Kemp, P.J.; Adriaensen, D. Purinergic signaling in the pulmonary neuroepithelial body microenvironment unraveled by live cell imaging. FASEB J. 2009, 23, 1153–1160. [Google Scholar] [CrossRef]
  11. Ahmad, S.; Ahmad, A.; White, C.W. Purinergic signaling and kinase activation for survival in pulmonary oxidative stress and disease. Free Radic. Biol. Med. 2006, 41, 29–40. [Google Scholar] [CrossRef]
  12. McMillan, M.R.; Burnstock, G.; Haworth, S.G. Vasoconstriction of intrapulmonary arteries to P2-receptor nucleotides in normal and pulmonary hypertensive newborn piglets. Br. J. Pharmacol. 1999, 128, 549–555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. McCormack, D.G.; Barnes, P.J.; Evans, T.W. Purinoceptors in the pulmonary circulation of the rat and their role in hypoxic vasoconstriction. Br. J. Pharmacol. 1989, 98, 367–372. [Google Scholar] [CrossRef] [Green Version]
  14. Li, H.H.; Hsu, H.H.; Chang, G.J.; Chen, I.C.; Ho, W.J.; Hsu, P.C.; Chen, W.J.; Pang, J.S.; Huang, C.C.; Lai, Y.J. Prostanoid EP4 agonist L-902,688 activates PPARgamma and attenuates pulmonary arterial hypertension. Am. J. Physiol. Lung Cell Mol. Physiol. 2018, 314, L349–L359. [Google Scholar] [CrossRef] [Green Version]
  15. Lai, Y.J.; Pullamsetti, S.S.; Dony, E.; Weissmann, N.; Butrous, G.; Banat, G.A.; Ghofrani, H.A.; Seeger, W.; Grimminger, F.; Schermuly, R.T. Role of the prostanoid EP4 receptor in iloprost-mediated vasodilatation in pulmonary hypertension. Am. J. Respir. Crit. Care Med. 2008, 178, 188–196. [Google Scholar] [CrossRef] [Green Version]
  16. Olschewski, H.; Rose, F.; Schermuly, R.; Ghofrani, H.A.; Enke, B.; Olschewski, A.; Seeger, W. Prostacyclin and its analogues in the treatment of pulmonary hypertension. Pharmacol. Ther. 2004, 102, 139–153. [Google Scholar] [CrossRef]
  17. Olschewski, H.; Rose, F.; Grunig, E.; Ghofrani, H.A.; Walmrath, D.; Schulz, R.; Schermuly, R.; Grimminger, F.; Seeger, W. Cellular pathophysiology and therapy of pulmonary hypertension. J. Lab. Clin. Med. 2001, 138, 367–377. [Google Scholar] [CrossRef] [PubMed]
  18. Glusa, E.; Adam, C. Endothelium-dependent relaxation induced by cathepsin G in porcine pulmonary arteries. Br. J. Pharmacol. 2001, 133, 422–428. [Google Scholar] [CrossRef] [PubMed]
  19. Olschewski, H.; Ghofrani, H.A.; Walmrath, D.; Schermuly, R.; Temmesfeld-Wollbruck, B.; Grimminger, F.; Seeger, W. Inhaled prostacyclin and iloprost in severe pulmonary hypertension secondary to lung fibrosis. Am. J. Respir. Crit. Care Med. 1999, 160, 600–607. [Google Scholar] [CrossRef]
  20. Walmrath, D.; Schermuly, R.; Pilch, J.; Grimminger, F.; Seeger, W. Effects of inhaled versus intravenous vasodilators in experimental pulmonary hypertension. Eur. Respir. J. 1997, 10, 1084–1092. [Google Scholar] [CrossRef] [Green Version]
  21. Strielkov, I.; Krause, N.C.; Sommer, N.; Schermuly, R.T.; Ghofrani, H.A.; Grimminger, F.; Gudermann, T.; Dietrich, A.; Weissmann, N. Hypoxic pulmonary vasoconstriction in isolated mouse pulmonary arterial vessels. Exp. Physiol. 2018, 103, 1185–1191. [Google Scholar] [CrossRef]
  22. Liu, Y.; Tian, X.Y.; Huang, Y.; Wang, N. Rosiglitazone Attenuated Endothelin-1-Induced Vasoconstriction of Pulmonary Arteries in the Rat Model of Pulmonary Arterial Hypertension via Differential Regulation of ET-1 Receptors. PPAR Res. 2014, 2014, 374075. [Google Scholar] [CrossRef]
  23. Paddenberg, R.; Tiefenbach, M.; Faulhammer, P.; Goldenberg, A.; Gries, B.; Pfeil, U.; Lips, K.S.; Piruat, J.I.; Lopez-Barneo, J.; Schermuly, R.T.; et al. Mitochondrial complex II is essential for hypoxia-induced pulmonary vasoconstriction of intra- but not of pre-acinar arteries. Cardiovasc. Res. 2012, 93, 702–710. [Google Scholar] [CrossRef]
  24. MacLean, M.R.; McCulloch, K.M.; Baird, M. Endothelin ETA- and ETB-receptor-mediated vasoconstriction in rat pulmonary arteries and arterioles. J. Cardiovasc. Pharmacol. 1994, 23, 838–845. [Google Scholar] [CrossRef]
  25. Wang, S.; Iring, A.; Strilic, B.; Albarran Juarez, J.; Kaur, H.; Troidl, K.; Tonack, S.; Burbiel, J.C.; Muller, C.E.; Fleming, I.; et al. P2Y(2) and Gq/G(1)(1) control blood pressure by mediating endothelial mechanotransduction. J. Clin. Investig. 2015, 125, 3077–3086. [Google Scholar] [CrossRef] [Green Version]
  26. Fredholm, B.B.; Abbracchio, M.P.; Burnstock, G.; Daly, J.W.; Harden, T.K.; Jacobson, K.A.; Leff, P.; Williams, M. Nomenclature and classification of purinoceptors. Pharmacol. Rev. 1994, 46, 143–156. [Google Scholar]
  27. Abbracchio, M.P.; Burnstock, G.; Boeynaems, J.M.; Barnard, E.A.; Boyer, J.L.; Kennedy, C.; Knight, G.E.; Fumagalli, M.; Gachet, C.; Jacobson, K.A.; et al. International Union of Pharmacology LVIII: Update on the P2Y G protein-coupled nucleotide receptors: From molecular mechanisms and pathophysiology to therapy. Pharmacol. Rev. 2006, 58, 281–341. [Google Scholar] [CrossRef] [PubMed]
  28. Burnstock, G. Physiology and pathophysiology of purinergic neurotransmission. Physiol. Rev. 2007, 87, 659–797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Tsubota, K. Understanding dry eye syndrome. Adv. Exp. Med. Biol. 2002, 506, 3–16. [Google Scholar] [CrossRef] [PubMed]
  30. Bielory, L. Ocular allergy and dry eye syndrome. Curr. Opin. Allergy Clin. Immunol. 2004, 4, 421–424. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, X.; Qu, Y.; He, X.; Ou, S.; Bu, J.; Jia, C.; Wang, J.; Wu, H.; Liu, Z.; Li, W. Dry Eye Management: Targeting the Ocular Surface Microenvironment. Int. J. Mol. Sci. 2017, 18, 1398. [Google Scholar] [CrossRef] [Green Version]
  32. Buscher, R.; Hoerning, A.; Patel, H.H.; Zhang, S.; Arthur, D.B.; Grasemann, H.; Ratjen, F.; Insel, P.A. P2Y2 receptor polymorphisms and haplotypes in cystic fibrosis and their impact on Ca2+ influx. Pharmacogenet. Genom. 2006, 16, 199–205. [Google Scholar] [CrossRef]
  33. Kellerman, D.; Evans, R.; Mathews, D.; Shaffer, C. Inhaled P2Y2 receptor agonists as a treatment for patients with Cystic Fibrosis lung disease. Adv. Drug Deliv. Rev. 2002, 54, 1463–1474. [Google Scholar] [CrossRef]
  34. Endo, K.I.; Sakamoto, A.; Fujisawa, K. Diquafosol tetrasodium elicits total cholesterol release from rabbit meibomian gland cells via P2Y2 purinergic receptor signalling. Sci. Rep. 2021, 11, 6989. [Google Scholar] [CrossRef]
  35. Deterding, R.; Retsch-Bogart, G.; Milgram, L.; Gibson, R.; Daines, C.; Zeitlin, P.L.; Milla, C.; Marshall, B.; Lavange, L.; Engels, J.; et al. Safety and tolerability of denufosol tetrasodium inhalation solution, a novel P2Y2 receptor agonist: Results of a phase 1/phase 2 multicenter study in mild to moderate cystic fibrosis. Pediatr. Pulmonol. 2005, 39, 339–348. [Google Scholar] [CrossRef]
  36. Deterding, R.R.; Lavange, L.M.; Engels, J.M.; Mathews, D.W.; Coquillette, S.J.; Brody, A.S.; Millard, S.P.; Ramsey, B.W.; The Cystic Fibrosis Therapeutics Development Network; The Inspire 08-103 Working Group. Phase 2 randomized safety and efficacy trial of nebulized denufosol tetrasodium in cystic fibrosis. Am. J. Respir. Crit. Care Med. 2007, 176, 362–369. [Google Scholar] [CrossRef]
  37. Accurso, F.J.; Moss, R.B.; Wilmott, R.W.; Anbar, R.D.; Schaberg, A.E.; Durham, T.A.; Ramsey, B.W.; Group, T.-I.S. Denufosol tetrasodium in patients with cystic fibrosis and normal to mildly impaired lung function. Am. J. Respir. Crit. Care Med. 2011, 183, 627–634. [Google Scholar] [CrossRef] [PubMed]
  38. Belete, H.A.; Hubmayr, R.D.; Wang, S.; Singh, R.D. The role of purinergic signaling on deformation induced injury and repair responses of alveolar epithelial cells. PLoS ONE 2011, 6, e27469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Olotu, C.; Kiefmann, M.; Ronneburg, C.; Lehmensiek, F.; Cuvenhaus, A.; Meidl, V.; Goetz, A.E.; Kiefmann, R. Analysis of purine receptor expression and functionality in alveolar epithelial cells. Purinergic Signal 2020, 16, 213–229. [Google Scholar] [CrossRef] [PubMed]
  40. Wirsching, E.; Fauler, M.; Fois, G.; Frick, M. P2 Purinergic Signaling in the Distal Lung in Health and Disease. Int. J. Mol. Sci. 2020, 21, 4973. [Google Scholar] [CrossRef]
  41. Mitchell, C.; Syed, N.I.; Tengah, A.; Gurney, A.M.; Kennedy, C. Identification of contractile P2Y1, P2Y6, and P2Y12 receptors in rat intrapulmonary artery using selective ligands. J. Pharmacol. Exp. Ther. 2012, 343, 755–762. [Google Scholar] [CrossRef] [PubMed]
  42. Gui, Y.; Walsh, M.P.; Jankowski, V.; Jankowski, J.; Zheng, X.L. Up4A stimulates endothelium-independent contraction of isolated rat pulmonary artery. Am. J. Physiol. Lung Cell Mol. Physiol. 2008, 294, L733–L738. [Google Scholar] [CrossRef]
  43. Hartley, S.A.; Kato, K.; Salter, K.J.; Kozlowski, R.Z. Functional evidence for a novel suramin-insensitive pyrimidine receptor in rat small pulmonary arteries. Circ. Res. 1998, 83, 940–946. [Google Scholar] [CrossRef] [PubMed]
  44. Visovatti, S.H.; Hyman, M.C.; Goonewardena, S.N.; Anyanwu, A.C.; Kanthi, Y.; Robichaud, P.; Wang, J.; Petrovic-Djergovic, D.; Rattan, R.; Burant, C.F.; et al. Purinergic dysregulation in pulmonary hypertension. Am. J. Physiol. Heart Circ. Physiol. 2016, 311, H286–H298. [Google Scholar] [CrossRef] [Green Version]
  45. Konduri, G.G.; Bakhutashvili, I.; Frenn, R.; Chandrasekhar, I.; Jacobs, E.R.; Khanna, A.K. P2Y purine receptor responses and expression in the pulmonary circulation of juvenile rabbits. Am. J. Physiol. Heart Circ. Physiol. 2004, 287, H157–H164. [Google Scholar] [CrossRef]
  46. Batori, R.; Kumar, S.; Bordan, Z.; Cherian-Shaw, M.; Kovacs-Kasa, A.; MacDonald, J.A.; Fulton, D.J.R.; Erdodi, F.; Verin, A.D. Differential mechanisms of adenosine- and ATPgammaS-induced microvascular endothelial barrier strengthening. J. Cell. Physiol. 2019, 234, 5863–5879. [Google Scholar] [CrossRef]
  47. Yamamoto, K.; Sokabe, T.; Ohura, N.; Nakatsuka, H.; Kamiya, A.; Ando, J. Endogenously released ATP mediates shear stress-induced Ca2+ influx into pulmonary artery endothelial cells. Am. J. Physiol. Heart Circ. Physiol. 2003, 285, H793–H803. [Google Scholar] [CrossRef] [Green Version]
  48. Zemskov, E.; Lucas, R.; Verin, A.D.; Umapathy, N.S. P2Y receptors as regulators of lung endothelial barrier integrity. J. Cardiovasc. Dis. Res. 2011, 2, 14–22. [Google Scholar] [CrossRef] [Green Version]
  49. Helenius, M.H.; Vattulainen, S.; Orcholski, M.; Aho, J.; Komulainen, A.; Taimen, P.; Wang, L.; de Jesus Perez, V.A.; Koskenvuo, J.W.; Alastalo, T.P. Suppression of endothelial CD39/ENTPD1 is associated with pulmonary vascular remodeling in pulmonary arterial hypertension. Am. J. Physiol. Lung Cell Mol. Physiol. 2015, 308, L1046–L1057. [Google Scholar] [CrossRef] [PubMed]
  50. Kylhammar, D.; Bune, L.T.; Radegran, G. P2Y(1) and P2Y(1)(2) receptors in hypoxia- and adenosine diphosphate-induced pulmonary vasoconstriction in vivo in the pig. Eur. J. Appl. Physiol. 2014, 114, 1995–2006. [Google Scholar] [CrossRef] [PubMed]
  51. Hennigs, J.K.; Luneburg, N.; Stage, A.; Schmitz, M.; Korbelin, J.; Harbaum, L.; Matuszcak, C.; Mienert, J.; Bokemeyer, C.; Boger, R.H.; et al. The P2-receptor-mediated Ca(2+) signalosome of the human pulmonary endothelium—Implications for pulmonary arterial hypertension. Purinergic Signal 2019, 15, 299–311. [Google Scholar] [CrossRef]
  52. Chen, B.C.; Lee, C.M.; Lee, Y.T.; Lin, W.W. Characterization of signaling pathways of P2Y and P2U purinoceptors in bovine pulmonary artery endothelial cells. J. Cardiovasc. Pharmacol. 1996, 28, 192–199. [Google Scholar] [CrossRef] [PubMed]
  53. Ahmad, S.; Ahmad, A.; Ghosh, M.; Leslie, C.C.; White, C.W. Extracellular ATP-mediated signaling for survival in hyperoxia-induced oxidative stress. J. Biol. Chem. 2004, 279, 16317–16325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Janssen, L.J.; Farkas, L.; Rahman, T.; Kolb, M.R. ATP stimulates Ca(2+)-waves and gene expression in cultured human pulmonary fibroblasts. Int. J. Biochem. Cell Biol. 2009, 41, 2477–2484. [Google Scholar] [CrossRef] [PubMed]
  55. Homolya, L.; Watt, W.C.; Lazarowski, E.R.; Koller, B.H.; Boucher, R.C. Nucleotide-regulated calcium signaling in lung fibroblasts and epithelial cells from normal and P2Y(2) receptor (−/−) mice. J. Biol. Chem. 1999, 274, 26454–26460. [Google Scholar] [CrossRef] [Green Version]
  56. Liang, S.; Desai, A.A.; Black, S.M.; Tang, H. Cytokines, Chemokines, and Inflammation in Pulmonary Arterial Hypertension. Adv. Exp. Med. Biol. 2021, 1303, 275–303. [Google Scholar] [CrossRef] [PubMed]
  57. Klouda, T.; Yuan, K. Inflammation in Pulmonary Arterial Hypertension. Adv. Exp. Med. Biol. 2021, 1303, 351–372. [Google Scholar] [CrossRef]
  58. Mercurio, V.; Cuomo, A.; Naranjo, M.; Hassoun, P.M. Inflammatory Mechanisms in the Pathogenesis of Pulmonary Arterial Hypertension: Recent Advances. Compr. Physiol. 2021, 11, 1805–1829. [Google Scholar] [CrossRef]
  59. Huertas, A.; Tu, L.; Humbert, M.; Guignabert, C. Chronic inflammation within the vascular wall in pulmonary arterial hypertension: More than a spectator. Cardiovasc. Res. 2020, 116, 885–893. [Google Scholar] [CrossRef]
  60. Myrtek, D.; Muller, T.; Geyer, V.; Derr, N.; Ferrari, D.; Zissel, G.; Durk, T.; Sorichter, S.; Luttmann, W.; Kuepper, M.; et al. Activation of human alveolar macrophages via P2 receptors: Coupling to intracellular Ca2+ increases and cytokine secretion. J. Immunol. 2008, 181, 2181–2188. [Google Scholar] [CrossRef] [Green Version]
  61. Bowler, J.W.; Bailey, R.J.; North, R.A.; Surprenant, A. P2X4, P2Y1 and P2Y2 receptors on rat alveolar macrophages. Br. J. Pharmacol. 2003, 140, 567–575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Mohanty, J.G.; Raible, D.G.; McDermott, L.J.; Pelleg, A.; Schulman, E.S. Effects of purine and pyrimidine nucleotides on intracellular Ca2+ in human eosinophils: Activation of purinergic P2Y receptors. J. Allergy Clin. Immunol. 2001, 107, 849–855. [Google Scholar] [CrossRef] [PubMed]
  63. Alkayed, F.; Kashimata, M.; Koyama, N.; Hayashi, T.; Tamura, Y.; Azuma, Y. P2Y11 purinoceptor mediates the ATP-enhanced chemotactic response of rat neutrophils. J. Pharmacol. Sci. 2012, 120, 288–295. [Google Scholar] [CrossRef] [Green Version]
  64. Chen, Y.; Shukla, A.; Namiki, S.; Insel, P.A.; Junger, W.G. A putative osmoreceptor system that controls neutrophil function through the release of ATP, its conversion to adenosine, and activation of A2 adenosine and P2 receptors. J. Leukoc. Biol. 2004, 76, 245–253. [Google Scholar] [CrossRef] [Green Version]
  65. Chen, Y.; Corriden, R.; Inoue, Y.; Yip, L.; Hashiguchi, N.; Zinkernagel, A.; Nizet, V.; Insel, P.A.; Junger, W.G. ATP release guides neutrophil chemotaxis via P2Y2 and A3 receptors. Science 2006, 314, 1792–1795. [Google Scholar] [CrossRef] [Green Version]
  66. Rieg, T.; Bundey, R.A.; Chen, Y.; Deschenes, G.; Junger, W.; Insel, P.A.; Vallon, V. Mice lacking P2Y2 receptors have salt-resistant hypertension and facilitated renal Na+ and water reabsorption. FASEB J. 2007, 21, 3717–3726. [Google Scholar] [CrossRef]
  67. Zhang, Y.; Sands, J.M.; Kohan, D.E.; Nelson, R.D.; Martin, C.F.; Carlson, N.G.; Kamerath, C.D.; Ge, Y.; Klein, J.D.; Kishore, B.K. Potential role of purinergic signaling in urinary concentration in inner medulla: Insights from P2Y2 receptor gene knockout mice. Am. J. Physiol. Renal Physiol. 2008, 295, F1715–F1724. [Google Scholar] [CrossRef] [Green Version]
  68. Cressman, V.L.; Lazarowski, E.; Homolya, L.; Boucher, R.C.; Koller, B.H.; Grubb, B.R. Effect of loss of P2Y(2) receptor gene expression on nucleotide regulation of murine epithelial Cl(-) transport. J. Biol. Chem. 1999, 274, 26461–26468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Dai, Q.Q.; Wang, Y.Y.; Jiang, Y.P.; Li, L.; Wang, H.J. VSNL1 Promotes Gastric Cancer Cell Proliferation and Migration by Regulating P2X3/P2Y2 Receptors and Is a Clinical Indicator of Poor Prognosis in Gastric Cancer Patients. Gastroenterol. Res. Pract. 2020, 2020, 7241942. [Google Scholar] [CrossRef]
  70. Homolya, L.; Steinberg, T.H.; Boucher, R.C. Cell to cell communication in response to mechanical stress via bilateral release of ATP and UTP in polarized epithelia. J. Cell Biol. 2000, 150, 1349–1360. [Google Scholar] [CrossRef]
  71. Winters, S.L.; Davis, C.W.; Boucher, R.C. Mechanosensitivity of mouse tracheal ciliary beat frequency: Roles for Ca2+, purinergic signaling, tonicity, and viscosity. Am. J. Physiol. Lung Cell Mol. Physiol. 2007, 292, L614–L624. [Google Scholar] [CrossRef] [Green Version]
  72. Geary, C.; Akinbi, H.; Korfhagen, T.; Fabre, J.E.; Boucher, R.; Rice, W. Increased susceptibility of purinergic receptor-deficient mice to lung infection with Pseudomonas aeruginosa. Am. J. Physiol. Lung Cell Mol. Physiol. 2005, 289, L890–L895. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Vanderstocken, G.; Van de Paar, E.; Robaye, B.; di Pietrantonio, L.; Bondue, B.; Boeynaems, J.M.; Desmecht, D.; Communi, D. Protective role of P2Y2 receptor against lung infection induced by pneumonia virus of mice. PLoS ONE 2012, 7, e50385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Vanderstocken, G.; Bondue, B.; Horckmans, M.; Di Pietrantonio, L.; Robaye, B.; Boeynaems, J.M.; Communi, D. P2Y2 receptor regulates VCAM-1 membrane and soluble forms and eosinophil accumulation during lung inflammation. J. Immunol. 2010, 185, 3702–3707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Inoue, Y.; Chen, Y.; Hirsh, M.I.; Yip, L.; Junger, W.G. A3 and P2Y2 receptors control the recruitment of neutrophils to the lungs in a mouse model of sepsis. Shock 2008, 30, 173–177. [Google Scholar] [CrossRef] [Green Version]
  76. Elliott, M.R.; Chekeni, F.B.; Trampont, P.C.; Lazarowski, E.R.; Kadl, A.; Walk, S.F.; Park, D.; Woodson, R.I.; Ostankovich, M.; Sharma, P.; et al. Nucleotides released by apoptotic cells act as a find-me signal to promote phagocytic clearance. Nature 2009, 461, 282–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Cohen, R.; Shainberg, A.; Hochhauser, E.; Cheporko, Y.; Tobar, A.; Birk, E.; Pinhas, L.; Leipziger, J.; Don, J.; Porat, E. UTP reduces infarct size and improves mice heart function after myocardial infarct via P2Y2 receptor. Biochem. Pharmacol. 2011, 82, 1126–1133. [Google Scholar] [CrossRef]
  78. Hochhauser, E.; Cohen, R.; Waldman, M.; Maksin, A.; Isak, A.; Aravot, D.; Jayasekara, P.S.; Muller, C.E.; Jacobson, K.A.; Shainberg, A. P2Y2 receptor agonist with enhanced stability protects the heart from ischemic damage in vitro and in vivo. Purinergic Signal 2013, 9, 633–642. [Google Scholar] [CrossRef] [Green Version]
  79. Qian, S.; Regan, J.N.; Shelton, M.T.; Hoggatt, A.; Mohammad, K.S.; Herring, P.B.; Seye, C.I. The P2Y2 nucleotide receptor is an inhibitor of vascular calcification. Atherosclerosis 2017, 257, 38–46. [Google Scholar] [CrossRef] [Green Version]
  80. Patel, J.J.; Zhu, D.; Opdebeeck, B.; D’Haese, P.; Millan, J.L.; Bourne, L.E.; Wheeler-Jones, C.P.D.; Arnett, T.R.; MacRae, V.E.; Orriss, I.R. Inhibition of arterial medial calcification and bone mineralization by extracellular nucleotides: The same functional effect mediated by different cellular mechanisms. J. Cell. Physiol. 2018, 233, 3230–3243. [Google Scholar] [CrossRef]
  81. Menzies, R.I.; Unwin, R.J.; Bailey, M.A. Renal P2 receptors and hypertension. Acta Physiol. (Oxf.) 2015, 213, 232–241. [Google Scholar] [CrossRef]
  82. Guns, P.J.; Van Assche, T.; Fransen, P.; Robaye, B.; Boeynaems, J.M.; Bult, H. Endothelium-dependent relaxation evoked by ATP and UTP in the aorta of P2Y2-deficient mice. Br. J. Pharmacol. 2006, 147, 569–574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Chen, X.; Qian, S.; Hoggatt, A.; Tang, H.; Hacker, T.A.; Obukhov, A.G.; Herring, P.B.; Seye, C.I. Endothelial Cell-Specific Deletion of P2Y2 Receptor Promotes Plaque Stability in Atherosclerosis-Susceptible ApoE-Null Mice. Arterioscler. Thromb. Vasc. Biol. 2017, 37, 75–83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Agca, C.; Seye, C.; Kashuba Benson, C.M.; Rikka, S.; Chan, A.W.; Weisman, G.A.; Agca, Y. Development of a novel transgenic rat overexpressing the P2Y(2) nucleotide receptor using a lentiviral vector. J. Vasc. Res. 2009, 46, 447–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Agca, Y.; Qian, S.; Agca, C.; Seye, C.I. Direct Evidence for P2Y2 Receptor Involvement in Vascular Response to Injury. J. Vasc. Res. 2016, 53, 163–171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Kraus, A.; Grampp, S.; Goppelt-Struebe, M.; Schreiber, R.; Kunzelmann, K.; Peters, D.J.; Leipziger, J.; Schley, G.; Schodel, J.; Eckardt, K.U.; et al. P2Y2R is a direct target of HIF-1alpha and mediates secretion-dependent cyst growth of renal cyst-forming epithelial cells. Purinergic Signal 2016, 12, 687–695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Tak, E.; Jun, D.Y.; Kim, S.H.; Park, G.C.; Lee, J.; Hwang, S.; Song, G.W.; Lee, S.G. Upregulation of P2Y2 nucleotide receptor in human hepatocellular carcinoma cells. J. Int. Med. Res. 2016, 44, 1234–1247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Sathanoori, R.; Rosi, F.; Gu, B.J.; Wiley, J.S.; Muller, C.E.; Olde, B.; Erlinge, D. Shear stress modulates endothelial KLF2 through activation of P2X4. Purinergic Signal 2015, 11, 139–153. [Google Scholar] [CrossRef] [Green Version]
  89. Sathanoori, R.; Bryl-Gorecka, P.; Muller, C.E.; Erb, L.; Weisman, G.A.; Olde, B.; Erlinge, D. P2Y2 receptor modulates shear stress-induced cell alignment and actin stress fibers in human umbilical vein endothelial cells. Cell. Mol. Life Sci. 2017, 74, 731–746. [Google Scholar] [CrossRef] [Green Version]
  90. Rodriguez-Zayas, A.E.; Torrado, A.I.; Miranda, J.D. P2Y2 receptor expression is altered in rats after spinal cord injury. Int. J. Dev. Neurosci. 2010, 28, 413–421. [Google Scholar] [CrossRef] [Green Version]
  91. Bagchi, S.; Liao, Z.; Gonzalez, F.A.; Chorna, N.E.; Seye, C.I.; Weisman, G.A.; Erb, L. The P2Y2 nucleotide receptor interacts with alphav integrins to activate Go and induce cell migration. J. Biol. Chem. 2005, 280, 39050–39057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Liao, Z.; Seye, C.I.; Weisman, G.A.; Erb, L. The P2Y2 nucleotide receptor requires interaction with alpha v integrins to access and activate G12. J. Cell Sci. 2007, 120, 1654–1662. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Chen, J.; Shao, C.; Lu, W.; Yan, C.; Yao, Q.; Zhu, M.; Chen, P.; Gu, P.; Fu, Y.; Fan, X. Adenosine triphosphate-induced rabbit corneal endothelial cell proliferation in vitro via the P2Y2-PI3K/Akt signaling axis. Cells Tissues Organs 2014, 199, 131–139. [Google Scholar] [CrossRef] [PubMed]
  94. El Husseini, D.; Boulanger, M.C.; Mahmut, A.; Bouchareb, R.; Laflamme, M.H.; Fournier, D.; Pibarot, P.; Bosse, Y.; Mathieu, P. P2Y2 receptor represses IL-6 expression by valve interstitial cells through Akt: Implication for calcific aortic valve disease. J. Mol. Cell. Cardiol. 2014, 72, 146–156. [Google Scholar] [CrossRef]
  95. Wang, T.; Takikawa, Y.; Watanabe, A.; Kakisaka, K.; Oikawa, K.; Miyamoto, Y.; Suzuki, K. Proliferation of mouse liver stem/progenitor cells induced by plasma from patients with acute liver failure is modulated by P2Y2 receptor-mediated JNK activation. J. Gastroenterol. 2014, 49, 1557–1566. [Google Scholar] [CrossRef]
  96. Cha, H.J.; Jung, M.S.; Ahn do, W.; Choi, J.K.; Ock, M.S.; Kim, K.S.; Yoon, J.H.; Song, E.J.; Song, K.S. Silencing of MUC8 by siRNA increases P2Y(2)-induced airway inflammation. Am. J. Physiol. Lung Cell Mol. Physiol. 2015, 308, L495–L502. [Google Scholar] [CrossRef]
  97. Eun, S.Y.; Ko, Y.S.; Park, S.W.; Chang, K.C.; Kim, H.J. IL-1beta enhances vascular smooth muscle cell proliferation and migration via P2Y2 receptor-mediated RAGE expression and HMGB1 release. Vascul. Pharmacol. 2015, 72, 108–117. [Google Scholar] [CrossRef]
  98. Byun, Y.S.; Yoo, Y.S.; Kwon, J.Y.; Joo, J.S.; Lim, S.A.; Whang, W.J.; Mok, J.W.; Choi, J.S.; Joo, C.K. Diquafosol promotes corneal epithelial healing via intracellular calcium-mediated ERK activation. Exp. Eye Res. 2016, 143, 89–97. [Google Scholar] [CrossRef] [Green Version]
  99. Albarran-Juarez, J.; Iring, A.; Wang, S.; Joseph, S.; Grimm, M.; Strilic, B.; Wettschureck, N.; Althoff, T.F.; Offermanns, S. Piezo1 and Gq/G11 promote endothelial inflammation depending on flow pattern and integrin activation. J. Exp. Med. 2018, 215, 2655–2672. [Google Scholar] [CrossRef] [Green Version]
  100. Dai, X.; Tohyama, M.; Murakami, M.; Shiraishi, K.; Liu, S.; Mori, H.; Utsunomiya, R.; Maeyama, K.; Sayama, K. House dust mite allergens induce interleukin 33 (IL-33) synthesis and release from keratinocytes via ATP-mediated extracellular signaling. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165719. [Google Scholar] [CrossRef]
  101. Zaparte, A.; Cappellari, A.R.; Brandao, C.A.; de Souza, J.B.; Borges, T.J.; Kist, L.W.; Bogo, M.R.; Zerbini, L.F.; Ribeiro Pinto, L.F.; Glaser, T.; et al. P2Y2 receptor activation promotes esophageal cancer cells proliferation via ERK1/2 pathway. Eur. J. Pharmacol. 2021, 891, 173687. [Google Scholar] [CrossRef]
  102. Uchida, Y.; Rutaganira, F.U.; Jullie, D.; Shokat, K.M.; von Zastrow, M. Endosomal Phosphatidylinositol 3-Kinase Is Essential for Canonical GPCR Signaling. Mol. Pharmacol. 2017, 91, 65–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Li, B.I.; Matteson, P.G.; Ababon, M.F.; Nato, A.Q., Jr.; Lin, Y.; Nanda, V.; Matise, T.C.; Millonig, J.H. The orphan GPCR, Gpr161, regulates the retinoic acid and canonical Wnt pathways during neurulation. Dev. Biol. 2015, 402, 17–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Ishimoto, H.; Wang, Z.; Rao, Y.; Wu, C.F.; Kitamoto, T. A novel role for ecdysone in Drosophila conditioned behavior: Linking GPCR-mediated non-canonical steroid action to cAMP signaling in the adult brain. PLoS Genet. 2013, 9, e1003843. [Google Scholar] [CrossRef] [Green Version]
  105. Favara, D.M.; Liebscher, I.; Jazayeri, A.; Nambiar, M.; Sheldon, H.; Banham, A.H.; Harris, A.L. Elevated expression of the adhesion GPCR ADGRL4/ELTD1 promotes endothelial sprouting angiogenesis without activating canonical GPCR signalling. Sci. Rep. 2021, 11, 8870. [Google Scholar] [CrossRef] [PubMed]
  106. Erb, L.; Liao, Z.; Seye, C.I.; Weisman, G.A. P2 receptors: Intracellular signaling. Pflugers Arch. 2006, 452, 552–562. [Google Scholar] [CrossRef]
  107. Liu, J.; Liao, Z.; Camden, J.; Griffin, K.D.; Garrad, R.C.; Santiago-Perez, L.I.; Gonzalez, F.A.; Seye, C.I.; Weisman, G.A.; Erb, L. Src homology 3 binding sites in the P2Y2 nucleotide receptor interact with Src and regulate activities of Src, proline-rich tyrosine kinase 2, and growth factor receptors. J. Biol. Chem. 2004, 279, 8212–8218. [Google Scholar] [CrossRef] [Green Version]
  108. Seye, C.I.; Yu, N.; Gonzalez, F.A.; Erb, L.; Weisman, G.A. The P2Y2 nucleotide receptor mediates vascular cell adhesion molecule-1 expression through interaction with VEGF receptor-2 (KDR/Flk-1). J. Biol. Chem. 2004, 279, 35679–35686. [Google Scholar] [CrossRef] [Green Version]
  109. Fukai, K.; Nakamura, A.; Hoshino, A.; Nakanishi, N.; Okawa, Y.; Ariyoshi, M.; Kaimoto, S.; Uchihashi, M.; Ono, K.; Tateishi, S.; et al. Pyk2 aggravates hypoxia-induced pulmonary hypertension by activating HIF-1alpha. Am. J. Physiol. Heart Circ. Physiol. 2015, 308, H951–H959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Liao, Z.; Cao, C.; Wang, J.; Huxley, V.H.; Baker, O.; Weisman, G.A.; Erb, L. The P2Y2 Receptor Interacts with VE-Cadherin and VEGF Receptor-2 to Regulate Rac1 Activity in Endothelial Cells. J. Biomed. Sci. Eng. 2014, 7, 1105–1121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Hein, T.W.; Kuo, L. cAMP-independent dilation of coronary arterioles to adenosine: Role of nitric oxide, G proteins, and K(ATP) channels. Circ. Res. 1999, 85, 634–642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Dominguez Rieg, J.A.; Burt, J.M.; Ruth, P.; Rieg, T. P2Y(2) receptor activation decreases blood pressure via intermediate conductance potassium channels and connexin 37. Acta Physiol. 2015, 213, 628–641. [Google Scholar] [CrossRef] [Green Version]
  113. Hoffmann, C.; Ziegler, N.; Reiner, S.; Krasel, C.; Lohse, M.J. Agonist-selective, receptor-specific interaction of human P2Y receptors with beta-arrestin-1 and -2. J. Biol. Chem. 2008, 283, 30933–30941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Morris, G.E.; Nelson, C.P.; Everitt, D.; Brighton, P.J.; Standen, N.B.; Challiss, R.A.; Willets, J.M. G protein-coupled receptor kinase 2 and arrestin2 regulate arterial smooth muscle P2Y-purinoceptor signalling. Cardiovasc. Res. 2011, 89, 193–203. [Google Scholar] [CrossRef] [Green Version]
  115. Morris, G.E.; Nelson, C.P.; Brighton, P.J.; Standen, N.B.; Challiss, R.A.; Willets, J.M. Arrestins 2 and 3 differentially regulate ETA and P2Y2 receptor-mediated cell signaling and migration in arterial smooth muscle. Am. J. Physiol. Cell Physiol. 2012, 302, C723–C734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Ng, S.W.; Bakowski, D.; Nelson, C.; Mehta, R.; Almeyda, R.; Bates, G.; Parekh, A.B. Cysteinyl leukotriene type I receptor desensitization sustains Ca2+-dependent gene expression. Nature 2012, 482, 111–115. [Google Scholar] [CrossRef] [Green Version]
  117. Shenoy, S.K.; Lefkowitz, R.J. beta-Arrestin-mediated receptor trafficking and signal transduction. Trends Pharmacol. Sci. 2011, 32, 521–533. [Google Scholar] [CrossRef] [Green Version]
  118. Li, J.; Cao, Y.X.; Liu, H.; Xu, C.B. Enhanced G-protein coupled receptors-mediated contraction and reduced endothelium-dependent relaxation in hypertension. Eur. J. Pharmacol. 2007, 557, 186–194. [Google Scholar] [CrossRef] [PubMed]
  119. Li, J.; Cao, Y.X.; Cao, L.; Liu, Y.; Xu, C.B. Heat stress alters G-protein coupled receptor-mediated function and endothelium-dependent relaxation in rat mesenteric artery. Eur. J. Pharmacol. 2008, 588, 280–285. [Google Scholar] [CrossRef]
  120. Jin, H.; Ko, Y.S.; Park, S.W.; Kim, H.J. P2Y2R activation by ATP induces oxLDL-mediated inflammasome activation through modulation of mitochondrial damage in human endothelial cells. Free Radic. Biol. Med. 2019, 136, 109–117. [Google Scholar] [CrossRef]
  121. Schuchardt, M.; Prufer, J.; Prufer, N.; Wiedon, A.; Huang, T.; Chebli, M.; Jankowski, V.; Jankowski, J.; Schafer-Korting, M.; Zidek, W.; et al. The endothelium-derived contracting factor uridine adenosine tetraphosphate induces P2Y(2)-mediated pro-inflammatory signaling by monocyte chemoattractant protein-1 formation. J. Mol. Med. 2011, 89, 799–810. [Google Scholar] [CrossRef]
  122. Nichols, K.K.; Yerxa, B.; Kellerman, D.J. Diquafosol tetrasodium: A novel dry eye therapy. Expert Opin. Investig. Drugs 2004, 13, 47–54. [Google Scholar] [CrossRef] [PubMed]
  123. Keating, G.M. Diquafosol ophthalmic solution 3%: A review of its use in dry eye. Drugs 2015, 75, 911–922. [Google Scholar] [CrossRef] [PubMed]
  124. Yerxa, B.R.; Sabater, J.R.; Davis, C.W.; Stutts, M.J.; Lang-Furr, M.; Picher, M.; Jones, A.C.; Cowlen, M.; Dougherty, R.; Boyer, J.; et al. Pharmacology of INS37217 [P(1)-(uridine 5′)-P(4)- (2′-deoxycytidine 5′)tetraphosphate, tetrasodium salt], a next-generation P2Y(2) receptor agonist for the treatment of cystic fibrosis. J. Pharmacol. Exp. Ther. 2002, 302, 871–880. [Google Scholar] [CrossRef] [PubMed]
  125. do Carmo, T.I.T.; Soares, V.E.M.; Wruck, J.; Dos Anjos, F.; de Resende, E.S.D.T.; de Oliveira Maciel, S.F.V.; Bagatini, M.D. Hyperinflammation and airway surface liquid dehydration in cystic fibrosis: Purinergic system as therapeutic target. Inflamm. Res. 2021, 70, 633–649. [Google Scholar] [CrossRef] [PubMed]
  126. Iyinikkel, J.; Murray, F. GPCRs in pulmonary arterial hypertension: Tipping the balance. Br. J. Pharmacol. 2018, 175, 3063–3079. [Google Scholar] [CrossRef]
  127. Janssens, R.; Paindavoine, P.; Parmentier, M.; Boeynaems, J.M. Human P2Y2 receptor polymorphism: Identification and pharmacological characterization of two allelic variants. Br. J. Pharmacol. 1999, 127, 709–716. [Google Scholar] [CrossRef] [Green Version]
  128. Hasan, D.; Satalin, J.; van der Zee, P.; Kollisch-Singule, M.; Blankman, P.; Shono, A.; Somhorst, P.; den Uil, C.; Meeder, H.; Kotani, T.; et al. Excessive Extracellular ATP Desensitizes P2Y2 and P2X4 ATP Receptors Provoking Surfactant Impairment Ending in Ventilation-Induced Lung Injury. Int. J. Mol. Sci. 2018, 19, 1185. [Google Scholar] [CrossRef] [Green Version]
  129. Dekker, R.J.; van Soest, S.; Fontijn, R.D.; Salamanca, S.; de Groot, P.G.; VanBavel, E.; Pannekoek, H.; Horrevoets, A.J. Prolonged fluid shear stress induces a distinct set of endothelial cell genes, most specifically lung Kruppel-like factor (KLF2). Blood 2002, 100, 1689–1698. [Google Scholar] [CrossRef]
  130. Groenendijk, B.C.; Van der Heiden, K.; Hierck, B.P.; Poelmann, R.E. The role of shear stress on ET-1, KLF2, and NOS-3 expression in the developing cardiovascular system of chicken embryos in a venous ligation model. Physiology 2007, 22, 380–389. [Google Scholar] [CrossRef]
  131. Kinderlerer, A.R.; Ali, F.; Johns, M.; Lidington, E.A.; Leung, V.; Boyle, J.J.; Hamdulay, S.S.; Evans, P.C.; Haskard, D.O.; Mason, J.C. KLF2-dependent, shear stress-induced expression of CD59: A novel cytoprotective mechanism against complement-mediated injury in the vasculature. J. Biol. Chem. 2008, 283, 14636–14644. [Google Scholar] [CrossRef] [Green Version]
  132. Doddaballapur, A.; Michalik, K.M.; Manavski, Y.; Lucas, T.; Houtkooper, R.H.; You, X.; Chen, W.; Zeiher, A.M.; Potente, M.; Dimmeler, S.; et al. Laminar shear stress inhibits endothelial cell metabolism via KLF2-mediated repression of PFKFB3. Arterioscler. Thromb. Vasc. Biol. 2015, 35, 137–145. [Google Scholar] [CrossRef] [Green Version]
  133. Ko, H.; Carter, R.L.; Cosyn, L.; Petrelli, R.; de Castro, S.; Besada, P.; Zhou, Y.; Cappellacci, L.; Franchetti, P.; Grifantini, M.; et al. Synthesis and potency of novel uracil nucleotides and derivatives as P2Y2 and P2Y6 receptor agonists. Bioorg. Med. Chem. 2008, 16, 6319–6332. [Google Scholar] [CrossRef] [Green Version]
  134. Alsaqati, M.; Chan, S.L.; Ralevic, V. Investigation of the functional expression of purine and pyrimidine receptors in porcine isolated pancreatic arteries. Purinergic Signal 2014, 10, 241–249. [Google Scholar] [CrossRef] [Green Version]
  135. Certal, M.; Vinhas, A.; Pinheiro, A.R.; Ferreirinha, F.; Barros-Barbosa, A.R.; Silva, I.; Costa, M.A.; Correia-de-Sa, P. Calcium signaling and the novel anti-proliferative effect of the UTP-sensitive P2Y11 receptor in rat cardiac myofibroblasts. Cell Calcium 2015, 58, 518–533. [Google Scholar] [CrossRef] [Green Version]
  136. Linan-Rico, A.; Ochoa-Cortes, F.; Zuleta-Alarcon, A.; Alhaj, M.; Tili, E.; Enneking, J.; Harzman, A.; Grants, I.; Bergese, S.; Christofi, F.L. UTP—Gated Signaling Pathways of 5-HT Release from BON Cells as a Model of Human Enterochromaffin Cells. Front. Pharmacol. 2017, 8, 429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Hevia, M.J.; Castro, P.; Pinto, K.; Reyna-Jeldes, M.; Rodriguez-Tirado, F.; Robles-Planells, C.; Ramirez-Rivera, S.; Madariaga, J.A.; Gutierrez, F.; Lopez, J.; et al. Differential Effects of Purinergic Signaling in Gastric Cancer-Derived Cells Through P2Y and P2X Receptors. Front. Pharmacol. 2019, 10, 612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Ando, K.; Obara, Y.; Sugama, J.; Kotani, A.; Koike, N.; Ohkubo, S.; Nakahata, N. P2Y2 receptor-Gq/11 signaling at lipid rafts is required for UTP-induced cell migration in NG 108-15 cells. J. Pharmacol. Exp. Ther. 2010, 334, 809–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Edwards, G.; Feletou, M.; Weston, A.H. Endothelium-derived hyperpolarising factors and associated pathways: A synopsis. Pflugers Arch. 2010, 459, 863–879. [Google Scholar] [CrossRef] [PubMed]
  140. Feletou, M.; Cohen, R.A.; Vanhoutte, P.M.; Verbeuren, T.J. TP receptors and oxidative stress hand in hand from endothelial dysfunction to atherosclerosis. Adv. Pharmacol. 2010, 60, 85–106. [Google Scholar] [CrossRef] [Green Version]
  141. Feletou, M.; Huang, Y.; Vanhoutte, P.M. Vasoconstrictor prostanoids. Pflugers Arch. 2010, 459, 941–950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Feletou, M.; Kohler, R.; Vanhoutte, P.M. Endothelium-derived vasoactive factors and hypertension: Possible roles in pathogenesis and as treatment targets. Curr. Hypertens. Rep. 2010, 12, 267–275. [Google Scholar] [CrossRef] [Green Version]
  143. Liang, C.F.; Au, A.L.; Leung, S.W.; Ng, K.F.; Feletou, M.; Kwan, Y.W.; Man, R.Y.; Vanhoutte, P.M. Endothelium-derived nitric oxide inhibits the relaxation of the porcine coronary artery to natriuretic peptides by desensitizing big conductance calcium-activated potassium channels of vascular smooth muscle. J. Pharmacol. Exp. Ther. 2010, 334, 223–231. [Google Scholar] [CrossRef] [Green Version]
  144. Vayssettes-Courchay, C.; Ragonnet, C.; Camenen, G.; Feletou, M.; Cordi, A.A.; Lavielle, G.; Verbeuren, T.J. Role of thromboxane TP and angiotensin AT1 receptors in lipopolysaccharide-induced arterial dysfunction in the rabbit: An in vivo study. Eur. J. Pharmacol. 2010, 634, 113–120. [Google Scholar] [CrossRef]
  145. Wang, S.; Chennupati, R.; Kaur, H.; Iring, A.; Wettschureck, N.; Offermanns, S. Endothelial cation channel PIEZO1 controls blood pressure by mediating flow-induced ATP release. J. Clin. Investig. 2016, 126, 4527–4536. [Google Scholar] [CrossRef]
  146. Rieg, T.; Gerasimova, M.; Boyer, J.L.; Insel, P.A.; Vallon, V. P2Y(2) receptor activation decreases blood pressure and increases renal Na(+) excretion. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2011, 301, R510–R518. [Google Scholar] [CrossRef] [Green Version]
  147. Velasquez, S.; Eugenin, E.A. Role of Pannexin-1 hemichannels and purinergic receptors in the pathogenesis of human diseases. Front. Physiol. 2014, 5, 96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Qi, H.; Zheng, X.; Qin, X.; Dou, D.; Xu, H.; Raj, J.U.; Gao, Y. Protein kinase G regulates the basal tension and plays a major role in nitrovasodilator-induced relaxation of porcine coronary veins. Br. J. Pharmacol. 2007, 152, 1060–1069. [Google Scholar] [CrossRef] [Green Version]
  149. Zhang, P.; Ma, Y.; Wang, Y.; Ma, X.; Huang, Y.; Li, R.A.; Wan, S.; Yao, X. Nitric oxide and protein kinase G act on TRPC1 to inhibit 11,12-EET-induced vascular relaxation. Cardiovasc. Res. 2014, 104, 138–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Amirjanians, M.; Egemnazarov, B.; Sydykov, A.; Kojonazarov, B.; Brandes, R.; Luitel, H.; Pradhan, K.; Stasch, J.P.; Redlich, G.; Weissmann, N.; et al. Chronic intratracheal application of the soluble guanylyl cyclase stimulator BAY 41-8543 ameliorates experimental pulmonary hypertension. Oncotarget 2017, 8, 29613–29624. [Google Scholar] [CrossRef]
  151. Deaglio, S.; Robson, S.C. Ectonucleotidases as regulators of purinergic signaling in thrombosis, inflammation, and immunity. Adv. Pharmacol. 2011, 61, 301–332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Evans, E.L.; Cuthbertson, K.; Endesh, N.; Rode, B.; Blythe, N.M.; Hyman, A.J.; Hall, S.J.; Gaunt, H.J.; Ludlow, M.J.; Foster, R.; et al. Yoda1 analogue (Dooku1) which antagonizes Yoda1-evoked activation of Piezo1 and aortic relaxation. Br. J. Pharmacol. 2018, 175, 1744–1759. [Google Scholar] [CrossRef]
  153. Hyman, A.J.; Tumova, S.; Beech, D.J. Piezo1 Channels in Vascular Development and the Sensing of Shear Stress. Curr. Top. Membr. 2017, 79, 37–57. [Google Scholar] [CrossRef] [PubMed]
  154. Beech, D.J. Endothelial Piezo1 channels as sensors of exercise. J. Physiol. 2018, 596, 979–984. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Rafikova, O.; Al Ghouleh, I.; Rafikov, R. Focus on Early Events: Pathogenesis of Pulmonary Arterial Hypertension Development. Antioxid. Redox Signal. 2019, 31, 933–953. [Google Scholar] [CrossRef]
  156. Yun, J.H.; Henson, P.M.; Tuder, R.M. Phagocytic clearance of apoptotic cells: Role in lung disease. Expert Rev. Respir. Med. 2008, 2, 753–765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Dolgachev, V.; Panicker, S.; Balijepalli, S.; McCandless, L.K.; Yin, Y.; Swamy, S.; Suresh, M.V.; Delano, M.J.; Hemmila, M.R.; Raghavendran, K.; et al. Electroporation-mediated delivery of FER gene enhances innate immune response and improves survival in a murine model of pneumonia. Gene Ther. 2018, 25, 359–375. [Google Scholar] [CrossRef]
  158. Bucheimer, R.E.; Linden, J. Purinergic regulation of epithelial transport. J. Physiol. 2004, 555, 311–321. [Google Scholar] [CrossRef]
  159. Griesenbach, U.; Geddes, D.M.; Alton, E.W. Advances in cystic fibrosis gene therapy. Curr. Opin. Pulm. Med. 2004, 10, 542–546. [Google Scholar] [CrossRef]
  160. Davies, J.C.; Alton, E.W. Airway gene therapy. Adv. Genet. 2005, 54, 291–314. [Google Scholar] [CrossRef]
  161. Tate, S.; Elborn, S. Progress towards gene therapy for cystic fibrosis. Expert Opin. Drug Deliv. 2005, 2, 269–280. [Google Scholar] [CrossRef]
  162. Ruan, J.; Hirai, H.; Yang, D.; Ma, L.; Hou, X.; Jiang, H.; Wei, H.; Rajagopalan, C.; Mou, H.; Wang, G.; et al. Efficient Gene Editing at Major CFTR Mutation Loci. Mol. Ther. Nucleic Acids 2019, 16, 73–81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Li, H.J.; Wang, L.Y.; Qu, H.N.; Yu, L.H.; Burnstock, G.; Ni, X.; Xu, M.; Ma, B. P2Y2 receptor-mediated modulation of estrogen-induced proliferation of breast cancer cells. Mol. Cell. Endocrinol. 2011, 338, 28–37. [Google Scholar] [CrossRef] [PubMed]
  164. Choi, J.H.; Ji, Y.G.; Lee, D.H. Uridine triphosphate increases proliferation of human cancerous pancreatic duct epithelial cells by activating P2Y2 receptor. Pancreas 2013, 42, 680–686. [Google Scholar] [CrossRef] [PubMed]
  165. Tackett, B.C.; Sun, H.; Mei, Y.; Maynard, J.P.; Cheruvu, S.; Mani, A.; Hernandez-Garcia, A.; Vigneswaran, N.; Karpen, S.J.; Thevananther, S. P2Y2 purinergic receptor activation is essential for efficient hepatocyte proliferation in response to partial hepatectomy. Am. J. Physiol. Gastrointest. Liver Physiol. 2014, 307, G1073–G1087. [Google Scholar] [CrossRef] [PubMed]
  166. Tuder, R.M.; Voelkel, N.F. Angiogenesis and pulmonary hypertension: A unique process in a unique disease. Antioxid. Redox Signal. 2002, 4, 833–843. [Google Scholar] [CrossRef]
  167. Wolf, D.; Tseng, N.; Seedorf, G.; Roe, G.; Abman, S.H.; Gien, J. Endothelin-1 decreases endothelial PPARgamma signaling and impairs angiogenesis after chronic intrauterine pulmonary hypertension. Am. J. Physiol. Lung Cell Mol. Physiol. 2014, 306, L361–L371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Ichikawa-Shindo, Y.; Sakurai, T.; Kamiyoshi, A.; Kawate, H.; Iinuma, N.; Yoshizawa, T.; Koyama, T.; Fukuchi, J.; Iimuro, S.; Moriyama, N.; et al. The GPCR modulator protein RAMP2 is essential for angiogenesis and vascular integrity. J. Clin. Investig. 2008, 118, 29–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Wyder, L.; Suply, T.; Ricoux, B.; Billy, E.; Schnell, C.; Baumgarten, B.U.; Maira, S.M.; Koelbing, C.; Ferretti, M.; Kinzel, B.; et al. Reduced pathological angiogenesis and tumor growth in mice lacking GPR4, a proton sensing receptor. Angiogenesis 2011, 14, 533–544. [Google Scholar] [CrossRef]
  170. Hanumegowda, C.; Farkas, L.; Kolb, M. Angiogenesis in pulmonary fibrosis: Too much or not enough? Chest 2012, 142, 200–207. [Google Scholar] [CrossRef] [PubMed]
  171. Yin, G.; Sheu, T.J.; Menon, P.; Pang, J.; Ho, H.C.; Shi, S.; Xie, C.; Smolock, E.; Yan, C.; Zuscik, M.J.; et al. Impaired angiogenesis during fracture healing in GPCR kinase 2 interacting protein-1 (GIT1) knock out mice. PLoS ONE 2014, 9, e89127. [Google Scholar] [CrossRef] [PubMed]
  172. Kir, D.; Schnettler, E.; Modi, S.; Ramakrishnan, S. Regulation of angiogenesis by microRNAs in cardiovascular diseases. Angiogenesis 2018, 21, 699–710. [Google Scholar] [CrossRef]
  173. Kang, H.; Hong, Z.; Zhong, M.; Klomp, J.; Bayless, K.J.; Mehta, D.; Karginov, A.V.; Hu, G.; Malik, A.B. Piezo1 mediates angiogenesis through activation of MT1-MMP signaling. Am. J. Physiol. Cell Physiol. 2019, 316, C92–C103. [Google Scholar] [CrossRef]
  174. Gogiraju, R.; Bochenek, M.L.; Schafer, K. Angiogenic Endothelial Cell Signaling in Cardiac Hypertrophy and Heart Failure. Front. Cardiovasc. Med. 2019, 6, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Patten, I.S.; Rana, S.; Shahul, S.; Rowe, G.C.; Jang, C.; Liu, L.; Hacker, M.R.; Rhee, J.S.; Mitchell, J.; Mahmood, F.; et al. Cardiac angiogenic imbalance leads to peripartum cardiomyopathy. Nature 2012, 485, 333–338. [Google Scholar] [CrossRef]
  176. Sedding, D.G.; Boyle, E.C.; Demandt, J.A.F.; Sluimer, J.C.; Dutzmann, J.; Haverich, A.; Bauersachs, J. Vasa Vasorum Angiogenesis: Key Player in the Initiation and Progression of Atherosclerosis and Potential Target for the Treatment of Cardiovascular Disease. Front. Immunol. 2018, 9, 706. [Google Scholar] [CrossRef] [Green Version]
  177. McEnaney, R.M.; Shukla, A.; Madigan, M.C.; Sachdev, U.; Tzeng, E. P2Y2 nucleotide receptor mediates arteriogenesis in a murine model of hind limb ischemia. J. Vasc. Surg. 2016, 63, 216–225. [Google Scholar] [CrossRef] [Green Version]
  178. Strassheim, D.; Verin, A.; Batori, R.; Nijmeh, H.; Burns, N.; Kovacs-Kasa, A.; Umapathy, N.S.; Kotamarthi, J.; Gokhale, Y.S.; Karoor, V.; et al. P2Y Purinergic Receptors, Endothelial Dysfunction, and Cardiovascular Diseases. Int. J. Mol. Sci. 2020, 21, 6855. [Google Scholar] [CrossRef]
  179. Hatakeyama, T.; Pappas, P.J.; Hobson, R.W., II; Boric, M.P.; Sessa, W.C.; Duran, W.N. Endothelial nitric oxide synthase regulates microvascular hyperpermeability in vivo. J. Physiol. 2006, 574, 275–281. [Google Scholar] [CrossRef]
  180. Fujihara, T.; Murakami, T.; Fujita, H.; Nakamura, M.; Nakata, K. Improvement of corneal barrier function by the P2Y(2) agonist INS365 in a rat dry eye model. Investig. Ophthalmol. Vis. Sci. 2001, 42, 96–100. [Google Scholar]
  181. Umapathy, N.S.; Fan, Z.; Zemskov, E.A.; Alieva, I.B.; Black, S.M.; Verin, A.D. Molecular mechanisms involved in adenosine-induced endothelial cell barrier enhancement. Vascul. Pharmacol. 2010, 52, 199–206. [Google Scholar] [CrossRef] [Green Version]
  182. Konduri, G.G.; Forman, K.; Mital, S. Characterization of purine receptors in fetal lamb pulmonary circulation. Pediatr. Res. 2000, 47, 114–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Konduri, G.G.; Woodard, L.L. Selective pulmonary vasodilation by low-dose infusion of adenosine triphosphate in newborn lambs. J. Pediatr. 1991, 119, 94–102. [Google Scholar] [CrossRef]
  184. Kaapa, P.; Jahnukainen, T.; Gronlund, J.; Rautanen, M.; Halkola, L.; Valimaki, I. Adenosine triphosphate treatment for meconium aspiration-induced pulmonary hypertension in pigs. Acta Physiol. Scand. 1997, 160, 283–289. [Google Scholar] [CrossRef] [PubMed]
  185. Crowley, M.R. Oxygen-induced pulmonary vasodilation is mediated by adenosine triphosphate in newborn lambs. J. Cardiovasc. Pharmacol. 1997, 30, 102–109. [Google Scholar] [CrossRef] [PubMed]
  186. Paidas, C.N.; Dudgeon, D.L.; Haller, J.A., Jr.; Clemens, M.G. Adenosine triphosphate: A potential therapy for hypoxic pulmonary hypertension. J. Pediatr. Surg. 1988, 23, 1154–1160. [Google Scholar] [CrossRef]
  187. Brook, M.M.; Fineman, J.R.; Bolinger, A.M.; Wong, A.F.; Heymann, M.A.; Soifer, S.J. Use of ATP-MgCl2 in the evaluation and treatment of children with pulmonary hypertension secondary to congenital heart defects. Circulation 1994, 90, 1287–1293. [Google Scholar] [CrossRef] [Green Version]
  188. Crooke, A.; Guzmán-Aranguez, A.; Peral, A.; Abdurrahman, M.K.A.; Pintor, J. Nucleotides in ocular secretions: Their role in ocular physiology. Pharmacol. Ther. 2008, 119, 55–73. [Google Scholar] [CrossRef]
  189. Jacobson, K.A.; Boeynaems, J.M. P2Y nucleotide receptors: Promise of therapeutic applications. Drug Discov. Today 2010, 15, 570–578. [Google Scholar] [CrossRef] [Green Version]
  190. Weisman, G.A.; Woods, L.T.; Erb, L.; Seye, C.I. P2Y receptors in the mammalian nervous system: Pharmacology, ligands and therapeutic potential. CNS Neurol. Disord. Drug Targets 2012, 11, 722–738. [Google Scholar] [CrossRef] [Green Version]
  191. Pintor, J.; Carracedo, G.; Alonso, C.M.; Bautista, A.; Peral, A. Presence of diadenosine polyphosphates in human tears. Pflügers Archiv. 2002, 443, 432–436. [Google Scholar] [CrossRef]
  192. Burnstock, G. Introductory overview of purinergic signalling. Front. Biosci. 2011, 3, 896–900. [Google Scholar] [CrossRef]
  193. Brunschweiger, A.; Muller, C.E. P2 receptors activated by uracil nucleotides—An update. Curr. Med. Chem. 2006, 13, 289–312. [Google Scholar] [CrossRef]
  194. El-Tayeb, A.; Qi, A.; Muller, C.E. Synthesis and structure-activity relationships of uracil nucleotide derivatives and analogues as agonists at human P2Y2, P2Y4, and P2Y6 receptors. J. Med. Chem. 2006, 49, 7076–7087. [Google Scholar] [CrossRef]
  195. Rafehi, M.; Burbiel, J.C.; Attah, I.Y.; Abdelrahman, A.; Muller, C.E. Synthesis, characterization, and in vitro evaluation of the selective P2Y2 receptor antagonist AR-C118925. Purinergic Signal. 2017, 13, 89–103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Troadec, J.D.; Thirion, S.; Petturiti, D.; Bohn, M.T.; Poujeol, P. ATP acting on P2Y receptors triggers calcium mobilization in primary cultures of rat neurohypophysial astrocytes (pituicytes). Pflugers Arch. 1999, 437, 745–753. [Google Scholar] [CrossRef] [PubMed]
  197. Murakami, T.; Fujihara, T.; Nakamura, M.; Nakata, K. P2Y(2) receptor stimulation increases tear fluid secretion in rabbits. Curr. Eye Res. 2000, 21, 782–787. [Google Scholar] [CrossRef] [PubMed]
  198. Murakami, T.; Fujihara, T.; Nakamura, M.; Nakata, K. P2Y(2) receptor elicits PAS-positive glycoprotein secretion from rabbit conjunctival goblet cells in vivo. J. Ocul. Pharmacol. Ther. 2003, 19, 345–352. [Google Scholar] [CrossRef] [PubMed]
  199. Jumblatt, J.E.; Jumblatt, M.M. Regulation of ocular mucin secretion by P2Y2 nucleotide receptors in rabbit and human conjunctiva. Exp. Eye Res. 1998, 67, 341–346. [Google Scholar] [CrossRef]
  200. Goss, C.H.; McKone, E.F.; Mathews, D.; Kerr, D.; Wanger, J.S.; Millard, S.P.; Cystic Fibrosis Therapeutics Development, N. Experience using centralized spirometry in the phase 2 randomized, placebo-controlled, double-blind trial of denufosol in patients with mild to moderate cystic fibrosis. J. Cyst. Fibros. 2008, 7, 147–153. [Google Scholar] [CrossRef] [Green Version]
  201. Kellerman, D.; Rossi Mospan, A.; Engels, J.; Schaberg, A.; Gorden, J.; Smiley, L. Denufosol: A review of studies with inhaled P2Y(2) agonists that led to Phase 3. Pulm. Pharmacol. Ther. 2008, 21, 600–607. [Google Scholar] [CrossRef]
  202. Ratjen, F.; Durham, T.; Navratil, T.; Schaberg, A.; Accurso, F.J.; Wainwright, C.; Barnes, M.; Moss, R.B.; Group, T.-S.I. Long term effects of denufosol tetrasodium in patients with cystic fibrosis. J. Cyst. Fibros. 2012, 11, 539–549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Moss, R.B. Pitfalls of drug development: Lessons learned from trials of denufosol in cystic fibrosis. J. Pediatr. 2013, 162, 676–680. [Google Scholar] [CrossRef]
  204. Muoboghare, M.O.; Drummond, R.M.; Kennedy, C. Characterisation of P2Y2 receptors in human vascular endothelial cells using AR-C118925XX, a competitive and selective P2Y2 antagonist. Br. J. Pharmacol. 2019, 176, 2894–2904. [Google Scholar] [CrossRef] [PubMed]
  205. Wu, K.C.; Cheng, K.S.; Wong, K.L.; Shiao, L.R.; Leung, Y.M.; Chang, L.Y. ARC 118925XX stimulates cation influx in bEND.3 endothelial cells. Fundam. Clin. Pharmacol. 2019, 33, 604–611. [Google Scholar] [CrossRef]
  206. Kemp, P.A.; Sugar, R.A.; Jackson, A.D. Nucleotide-mediated mucin secretion from differentiated human bronchial epithelial cells. Am. J. Respir. Cell Mol. Biol. 2004, 31, 446–455. [Google Scholar] [CrossRef]
  207. Jacobson, K.A.; Costanzi, S.; Ohno, M.; Joshi, B.V.; Besada, P.; Xu, B.; Tchilibon, S. Molecular recognition at purine and pyrimidine nucleotide (P2) receptors. Curr. Top. Med. Chem. 2004, 4, 805–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Hillmann, P.; Ko, G.Y.; Spinrath, A.; Raulf, A.; von Kugelgen, I.; Wolff, S.C.; Nicholas, R.A.; Kostenis, E.; Holtje, H.D.; Muller, C.E. Key determinants of nucleotide-activated G protein-coupled P2Y(2) receptor function revealed by chemical and pharmacological experiments, mutagenesis and homology modeling. J. Med. Chem. 2009, 52, 2762–2775. [Google Scholar] [CrossRef]
  209. Lazarowski, E.R.; Watt, W.C.; Stutts, M.J.; Boucher, R.C.; Harden, T.K. Pharmacological selectivity of the cloned human P2U-purinoceptor: Potent activation by diadenosine tetraphosphate. Br. J. Pharmacol. 1995, 116, 1619–1627. [Google Scholar] [CrossRef] [Green Version]
  210. Jacobson, K.A.; Jarvis, M.F.; Williams, M. Purine and pyrimidine (P2) receptors as drug targets. J. Med. Chem. 2002, 45, 4057–4093. [Google Scholar] [CrossRef]
  211. Kim, H.S.; Ravi, R.G.; Marquez, V.E.; Maddileti, S.; Wihlborg, A.K.; Erlinge, D.; Malmsjo, M.; Boyer, J.L.; Harden, T.K.; Jacobson, K.A. Methanocarba modification of uracil and adenine nucleotides: High potency of Northern ring conformation at P2Y1, P2Y2, P2Y4, and P2Y11 but not P2Y6 receptors. J. Med. Chem. 2002, 45, 208–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Rafehi, M.; Muller, C.E. Tools and drugs for uracil nucleotide-activated P2Y receptors. Pharmacol. Ther. 2018, 190, 24–80. [Google Scholar] [CrossRef]
  213. Matsumoto, T.; Tostes, R.C.; Webb, R.C. Uridine adenosine tetraphosphate-induced contraction is increased in renal but not pulmonary arteries from DOCA-salt hypertensive rats. Am. J. Physiol. Heart Circ. Physiol. 2011, 301, H409–H417. [Google Scholar] [CrossRef] [Green Version]
  214. Martin-Gil, A.; de Lara, M.J.; Crooke, A.; Santano, C.; Peral, A.; Pintor, J. Silencing of P2Y(2) receptors reduces intraocular pressure in New Zealand rabbits. Br. J. Pharmacol. 2012, 165, 1163–1172. [Google Scholar] [CrossRef] [Green Version]
  215. Degagne, E.; Degrandmaison, J.; Grbic, D.M.; Vinette, V.; Arguin, G.; Gendron, F.P. P2Y2 receptor promotes intestinal microtubule stabilization and mucosal re-epithelization in experimental colitis. J. Cell. Physiol. 2013, 228, 99–109. [Google Scholar] [CrossRef]
  216. Yu, Y.; Zhang, C. Purinergic signaling negatively regulates activity of an olfactory receptor in an odorant-dependent manner. Neuroscience 2014, 275, 89–101. [Google Scholar] [CrossRef]
  217. Bouchareb, R.; Cote, N.; Marie Chloe, B.; Le Quang, K.; El Husseini, D.; Asselin, J.; Hadji, F.; Lachance, D.; Shayhidin, E.E.; Mahmut, A.; et al. Carbonic anhydrase XII in valve interstitial cells promotes the regression of calcific aortic valve stenosis. J. Mol. Cell. Cardiol. 2015, 82, 104–115. [Google Scholar] [CrossRef]
  218. Yitzhaki, S.; Hochhauser, E.; Porat, E.; Shainberg, A. Uridine-5′-triphosphate (UTP) maintains cardiac mitochondrial function following chemical and hypoxic stress. J. Mol. Cell. Cardiol. 2007, 43, 653–662. [Google Scholar] [CrossRef] [PubMed]
  219. Castro, E.; Pintor, J.; Miras-Portugal, M.T. Ca(2+)-stores mobilization by diadenosine tetraphosphate, Ap4A, through a putative P2Y purinoceptor in adrenal chromaffin cells. Br. J. Pharmacol. 1992, 106, 833–837. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Patel, K.; Barnes, A.; Camacho, J.; Paterson, C.; Boughtflower, R.; Cousens, D.; Marshall, F. Activity of diadenosine polyphosphates at P2Y receptors stably expressed in 1321N1 cells. Eur. J. Pharmacol. 2001, 430, 203–210. [Google Scholar] [CrossRef]
  221. Ivanov, A.A.; Fricks, I.; Kendall Harden, T.; Jacobson, K.A. Molecular dynamics simulation of the P2Y14 receptor. Ligand docking and identification of a putative binding site of the distal hexose moiety. Bioorg. Med. Chem. Lett. 2007, 17, 761–766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. El-Tayeb, A.; Qi, A.; Nicholas, R.A.; Muller, C.E. Structural modifications of UMP, UDP, and UTP leading to subtype-selective agonists for P2Y2, P2Y4, and P2Y6 receptors. J. Med. Chem. 2011, 54, 2878–2890. [Google Scholar] [CrossRef]
  223. Jacobson, K.A.; Muller, C.E. Medicinal chemistry of adenosine, P2Y and P2X receptors. Neuropharmacology 2016, 104, 31–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Chen, M.; Chen, H.; Gu, Y.; Sun, P.; Sun, J.; Yu, H.; Zheng, H.; Chen, D. P2Y2 promotes fibroblasts activation and skeletal muscle fibrosis through AKT, ERK, and PKC. BMC Musculoskelet. Disord. 2021, 22, 680. [Google Scholar] [CrossRef]
  225. Lazarowski, E.R.; Watt, W.C.; Stutts, M.J.; Brown, H.A.; Boucher, R.C.; Harden, T.K. Enzymatic synthesis of UTP gamma S, a potent hydrolysis resistant agonist of P2U-purinoceptors. Br. J. Pharmacol. 1996, 117, 203–209. [Google Scholar] [CrossRef]
  226. Pendergast, W.; Yerxa, B.R.; Douglass, J.G., III; Shaver, S.R.; Dougherty, R.W.; Redick, C.C.; Sims, I.F.; Rideout, J.L. Synthesis and P2Y receptor activity of a series of uridine dinucleoside 5′-polyphosphates. Bioorg. Med. Chem. Lett. 2001, 11, 157–160. [Google Scholar] [CrossRef]
  227. Murakami, T.; Fujihara, T.; Horibe, Y.; Nakamura, M. Diquafosol elicits increases in net Cl- transport through P2Y2 receptor stimulation in rabbit conjunctiva. Ophthalmic Res. 2004, 36, 89–93. [Google Scholar] [CrossRef] [PubMed]
  228. Dominguez-Godinez, C.; Carracedo, G.; Pintor, J. Diquafosol Delivery from Silicone Hydrogel Contact Lenses: Improved Effect on Tear Secretion. J. Ocul. Pharmacol. Ther. 2018, 34, 170–176. [Google Scholar] [CrossRef]
  229. Mironova, E.; Suliman, F.; Stockand, J.D. Renal Na(+) excretion consequent to pharmacogenetic activation of Gq-DREADD in principal cells. Am. J. Physiol. Renal Physiol. 2019, 316, F758–F767. [Google Scholar] [CrossRef]
  230. Ivanov, A.A.; Ko, H.; Cosyn, L.; Maddileti, S.; Besada, P.; Fricks, I.; Costanzi, S.; Harden, T.K.; Calenbergh, S.V.; Jacobson, K.A. Molecular modeling of the human P2Y2 receptor and design of a selective agonist, 2′-amino-2′-deoxy-2-thiouridine 5′-triphosphate. J. Med. Chem. 2007, 50, 1166–1176. [Google Scholar] [CrossRef] [Green Version]
  231. Kim, J.C.; Woo, S.H. Shear stress induces a longitudinal Ca(2+) wave via autocrine activation of P2Y1 purinergic signalling in rat atrial myocytes. J. Physiol. 2015, 593, 5091–5109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Xie, R.; Xu, J.; Wen, G.; Jin, H.; Liu, X.; Yang, Y.; Ji, B.; Jiang, Y.; Song, P.; Dong, H.; et al. The P2Y2 nucleotide receptor mediates the proliferation and migration of human hepatocellular carcinoma cells induced by ATP. J. Biol. Chem. 2014, 289, 19137–19149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Purinergic receptor tree, and the classification of P2Y2: Adenosine receptors and ATP receptors (P1 and P2, respectively).
Figure 1. Purinergic receptor tree, and the classification of P2Y2: Adenosine receptors and ATP receptors (P1 and P2, respectively).
Ijerph 18 11009 g001
Figure 2. P2Y2 mediated signaling pathways that are generally expected in different cell types and tissues, including the pulmonary vasculature. Abbreviations: MRS = MRS2768; DAG = diacylglycerol; PKC = proteinkinase C; Erk1/2 = extracellular-signal regulated kinases; PLCβ = phospholipase C (β); IP3 = inositol trisphosphate; CaM = calmodulin; CaMK = Ca2+/calmodulin-dependent protein kinase; CREB = cAMP response element-binding protein; NOS = nitric oxide (NO) synthase; SH3 = SRC homology 3 domain; Gα, Gβ, and Gγ = G-protein subunits (α, β and γ); Src = proto-oncogene tyrosine-protein kinase; PECAM-1 = platelet endothelial cell adhesion molecule; Pyk2 = protein tyrosine kinase 2 beta; VEGFR-2 = vascular endothelial growth factor receptor 2; VE-cadherin = vascular endothelial cadherin; Rac1 = Ras-related C3 botulinum toxin substrate 1; PI3K = phosphoinositide 3-kinases; Akt = protein kinase B (PKB) [3,25,91,92,93,94,95,96,97,98,99,100,101].
Figure 2. P2Y2 mediated signaling pathways that are generally expected in different cell types and tissues, including the pulmonary vasculature. Abbreviations: MRS = MRS2768; DAG = diacylglycerol; PKC = proteinkinase C; Erk1/2 = extracellular-signal regulated kinases; PLCβ = phospholipase C (β); IP3 = inositol trisphosphate; CaM = calmodulin; CaMK = Ca2+/calmodulin-dependent protein kinase; CREB = cAMP response element-binding protein; NOS = nitric oxide (NO) synthase; SH3 = SRC homology 3 domain; Gα, Gβ, and Gγ = G-protein subunits (α, β and γ); Src = proto-oncogene tyrosine-protein kinase; PECAM-1 = platelet endothelial cell adhesion molecule; Pyk2 = protein tyrosine kinase 2 beta; VEGFR-2 = vascular endothelial growth factor receptor 2; VE-cadherin = vascular endothelial cadherin; Rac1 = Ras-related C3 botulinum toxin substrate 1; PI3K = phosphoinositide 3-kinases; Akt = protein kinase B (PKB) [3,25,91,92,93,94,95,96,97,98,99,100,101].
Ijerph 18 11009 g002
Figure 3. Endothelial P2Y2 interactions with Pannexin1 and Piezo1 under fluid shear stress conditions in systemic circulation. Fluid shear stress activates Piezo1, a mechanosensitive channel that mediates ATP release via pannexin1 channels. Ectonucleotidases hydrolyze the extracellular nucleotide ATP to ADP, AMP, and adenosine, which activate P2Y2 and other P2Y receptors, suggesting P2Y2 as being downstream of Piezo1 [99,145,151,152,153,154].
Figure 3. Endothelial P2Y2 interactions with Pannexin1 and Piezo1 under fluid shear stress conditions in systemic circulation. Fluid shear stress activates Piezo1, a mechanosensitive channel that mediates ATP release via pannexin1 channels. Ectonucleotidases hydrolyze the extracellular nucleotide ATP to ADP, AMP, and adenosine, which activate P2Y2 and other P2Y receptors, suggesting P2Y2 as being downstream of Piezo1 [99,145,151,152,153,154].
Ijerph 18 11009 g003
Table 1. Expression of P2Y2 receptors in the pulmonary vasculature.
Table 1. Expression of P2Y2 receptors in the pulmonary vasculature.
Cell TypeSpeciesDetection LevelRef.
PAECsHumanmRNA[47,48]
BovineAgonist/activity[52]
RabbitsmRNA and agonist/activity[45]
LMVECsHumanmRNA[48,53]
PASMCsRatAgonist/activity[43]
PFsHumanmRNA[54]
MousemRNA[55]
PAECs, pulmonary arterial endothelial cells; LMVECs, human microvascular endothelial cells from lung; PASMCs, pulmonary arterial smooth muscle cells; PFs, pulmonary fibroblasts.
Table 2. P2Y2 specific and non-specific ligands.
Table 2. P2Y2 specific and non-specific ligands.
LigandActionTargetRef.
ATPAgonistP2Y1, P2Y11, P2Y12, P2Y2, P2Y4, P2Y6, P2Y13[71,209,210,211]
UTPAgonistP2Y1, P2Y11, P2Y2, P2Y4, P2Y6[204,211,212]
2-thioUTPAgonistP2Y2, P2Y4, P2Y6[194,213,214,215,216,217]
4-thio-UTPAgonistP2Y2, P2Y4[193]
UTPγSAgonistP2Y2[209,218]
5BrUTPAgonistP2Y2, P2Y6[209]
Ap4AAgonistP2Y2, P2Y4, P2Y12, P2Y13[93,127,209,219,220]
PSB1114AgonistP2Y2, P2Y4, P2Y6[194,212,221,222,223,224]
Denufosol (INS37217)AgonistP2Y2, P2Y4, P2Y6[35,36,37,146,200,201,202,203]
Diquafosol (INS365)AgonistP2Y1, P2Y2, P2Y4, P2Y6[34,98,122,123,124,225,226,227,228]
INS45973AgonistP2Y2, P2Y4[146,229]
MRS2698AgonistP2Y2, P2Y4, P2Y6[212,230]
MRS2768AgonistP2Y2[78,133,134,135,136,137,164,229]
AR-C126313AntagonistP2Y2[207,230]
Reactive blue-2AntagonistP2Y2, P2Y4, P2Y6, P2Y11, P2Y13[127,135]
AR-C118925XXAntagonistP2Y2[34,206]
PSB-416AntagonistP2Y2[208]
SuraminAntagonistP2Y1, P2Y11, P2Y12, P2Y2, P2Y4, P2Y6, P2Y1, P2Y13[6,7,44,98,127,135,164,206,231,232]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shihan, M.; Novoyatleva, T.; Lehmeyer, T.; Sydykov, A.; Schermuly, R.T. Role of the Purinergic P2Y2 Receptor in Pulmonary Hypertension. Int. J. Environ. Res. Public Health 2021, 18, 11009. https://doi.org/10.3390/ijerph182111009

AMA Style

Shihan M, Novoyatleva T, Lehmeyer T, Sydykov A, Schermuly RT. Role of the Purinergic P2Y2 Receptor in Pulmonary Hypertension. International Journal of Environmental Research and Public Health. 2021; 18(21):11009. https://doi.org/10.3390/ijerph182111009

Chicago/Turabian Style

Shihan, Mazen, Tatyana Novoyatleva, Thilo Lehmeyer, Akylbek Sydykov, and Ralph T. Schermuly. 2021. "Role of the Purinergic P2Y2 Receptor in Pulmonary Hypertension" International Journal of Environmental Research and Public Health 18, no. 21: 11009. https://doi.org/10.3390/ijerph182111009

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