Next Article in Journal
Diagnosis, Prognosis and Treatment of Canine Cutaneous and Subcutaneous Mast Cell Tumors
Next Article in Special Issue
FcεRI: A Master Regulator of Mast Cell Functions
Previous Article in Journal
Impact of β-Amyloids Induced Disruption of Ca2+ Homeostasis in a Simple Model of Neuronal Activity
Previous Article in Special Issue
The Critical Role Played by Mitochondrial MITF Serine 73 Phosphorylation in Immunologically Activated Mast Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synergistic Cytokine Production by ATP and PGE2 via P2X4 and EP3 Receptors in Mouse Bone-Marrow-Derived Mast Cells

1
Laboratory of Pharmacology, Faculty of Pharmacy, Takasaki University of Health and Welfare, Takasaki 370-0033, Gunma, Japan
2
Laboratory of Allergy and Immunology, Faculty of Pharmacy, Takasaki University of Health and Welfare, Takasaki 370-0033, Gunma, Japan
3
Department of Biomedical Engineering, Graduate School of Medicine, The University of Tokyo, Tokyo 113-0033, Japan
4
Discovery Research Laboratories, Nippon Chemiphar Co., Ltd., Misato 341-0005, Saitama, Japan
*
Author to whom correspondence should be addressed.
Cells 2022, 11(4), 616; https://doi.org/10.3390/cells11040616
Submission received: 28 December 2021 / Revised: 7 February 2022 / Accepted: 8 February 2022 / Published: 10 February 2022
(This article belongs to the Collection Mast Cells in Health and Diseases)

Abstract

:
ATP is an important intercellular messenger in the extracellular space. In mast cells (MCs), ATP stimulates the ionotropic P2X4 receptor (P2X4R), resulting in enhanced degranulation and exacerbation of acute allergic reactions. In this study, we investigate whether ATP regulates inflammatory cytokine production in MCs. Gene expression was analyzed by quantitative RT-PCR, and cytokine production was measured using ELISA. The stimulation of mouse bone-marrow-derived MCs (BMMCs) with ATP alone had little effect on cytokine secretion. However, the co-stimulation with prostaglandin (PG) E2 resulted in a marked increase in the secretion of various cytokines, such as tumor necrosis factor-α, interleukin (IL)-6, and IL-13, accompanied by an increase in their mRNA levels. The effects of ATP were inhibited by P2X4R antagonists and diminished in BMMCs derived from P2X4R-deficient mice, suggesting that P2X4R mediated the reaction. The effects of PGE2 were mimicked by an EP3 receptor (EP3R) agonist and blocked by an EP3R antagonist. The synergistic cytokine mRNA elevations induced by ATP and PGE2 were blocked by nuclear factor-κB and Ca2+-calcineurin signaling inhibitors. Altogether, these results suggest that combining P2X4R and EP3R signaling enhances acute degranulation and the subsequent cytokine secretion, exacerbating allergic inflammation.

1. Introduction

Adenosine triphosphate (ATP), an important intracellular energy source constantly supplied by cell metabolism under cellular respiration, is present in high concentrations in the cytosol. However, cells release ATP into the extracellular space as an intercellular mediator to regulate various physiological functions [1]. Extracellular ATP accumulation is recognized by various receptors called P2 receptors. Seven subtypes of ionotropic P2X receptors (P2X1-7) recognize ATP exclusively and open non-selective cation channels [2]. In contrast, eight subtypes of G protein-conjugated P2Y receptors have been identified in humans, including P2Y1, Y2, Y4, Y6, and Y11–14 receptors, and the ligands are not only ATP but also ADP and uridine nucleotides, such as UTP, UDP, and UDP-glucose [3]. Receptor activation is transmitted into cells by the G protein-dependent signaling [3]. In addition, extracellular ATP is rapidly dephosphorylated by various extracellular ATP-degrading enzymes to adenosine, acting as an agonist to another set of P1 adenosine receptors, such as Gi-coupled A1 and A3 receptors, and Gs-coupled A2A and A2B receptors [4]. Thus, ATP is involved in regulating physiological functions as a spatially and temporally diverse intercellular transmitter.
Initially, ATP was focused as a transmitter of non-cholinergic and non-adrenergic neurotransmission in the regulation of the contraction of smooth muscle organs, such as the intestine [5], bladder [6], and seminal vesicle [7]. However, tissue damage or mechanical stimulation causes a large amount of ATP release as a danger or find-me signal [8], facilitating the accumulation of evidence that ATP plays an important role in regulating immune cells. For example, ATP and UTP released from apoptotic cells stimulate P2Y and adenosine receptors on neutrophils [9], macrophages [10], and lymphocytes [11] to promote migration to the affected area. In macrophages, the stimulation of P2X7 receptors activates the inflammasome, promoting IL-1β processing and inducing IL-1β release [12]. Activating P2X4 receptors (P2X4R) in microglia plays an essential role in inducing neuropathic pain [13].
Mast cells (MCs) express various P1 and P2 purinergic receptors, and important information about the role of purinergic signaling in MC function has been reported [14]. For example, the stimulation of MCs with high concentrations of ATP induces direct degranulation via P2X7 receptors involved in the pathophysiology of inflammatory bowel disease [15] and skin hypersensitivity [16]. In addition, adenosine promotes antigen-stimulated degranulation via the A3 receptor [17]. G protein-coupled P2Y receptors are also abundantly expressed in MCs; however, their effects have been reported only for promoting antigen-induced degranulation via the stimulation of P2Y13 and P2Y14 receptors in model cell lines [18,19]. Recently, we reported that ATP promotes antigen-stimulated degranulation via ionotropic P2X4R using mouse bone-marrow-derived MCs (BMMCs) [20]. Furthermore, the P2X4R stimulating effect increased the antigen-independent degranulation reaction [21]. Co-stimulation with the EP3 receptor (EP3R) of prostaglandin (PG) E2 coupled to the Gi protein and the A3 receptor of adenosine induces a remarkable degranulation reaction. Such enhancement of antigen-dependent and antigen-independent degranulation reactions affected the degree of allergic reaction in vivo [20]. Since the magnitude of the degranulation reaction is related to the severity of the immediate allergic reaction, an increase in the P2X4R-mediated degranulation reaction is considered important for understanding allergic reactions.
MCs induce an immediate allergic reaction by releasing granule contents, such as histamine, serotonin, and various proteases, and produce various chemokines and cytokines as a delayed reaction [22]. Compared to the regulatory mechanism of the degranulation reaction, the purinergic regulation of cytokine production in MCs remains unexamined. In this study, we investigate whether extracellular ATP affects the production of cytokines by antigen and non-antigen stimulation using BMMCs.

2. Materials and Methods

2.1. Materials

ATP, α,β-methylene ATP (αβmeATP), ADP, UTP, PGE2, dexamethasone, 2,4-dinitrophenyl human serum albumin (DNP-HSA), anti-DNP IgE (clone SPE-7), cyclosporine and the GenElute Mammalian Total RNA miniprep kit were obtained from Sigma-Aldrich (Tokyo, Japan). 1,2-Bis (2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid tetrakis acetoxymethyl ester (BAPTA-AM) was obtained from Abcam (Cambridge, UK). Allophycocyanin-conjugated rat anti-mouse c-Kit antibodies (clone 2B8) were obtained from BD Pharmingen (Tokyo, Japan). Phycoerythrin-conjugated mouse anti-mouse FcεRIα antibodies (clone MAR-1) were obtained from eBioscience (San Diego, CA, USA). Recombinant mouse interleukin (IL)-3 and recombinant mouse SCF were obtained from Peprotech (London, UK). SB203580, wortmannin and AH6809 were obtained from Cayman Chemical (Ann Arbor, MI, USA). InSolution NF-kB activation inhibitor was obtained from Calbiochem (San Diego, CA, USA). U0126, anti-extracellular signal-regulated kinase (ERK)1/2, anti-phospho-ERK1/2, anti-Akt, anti-phospho-Akt (Thr308), anti-NF-κB p65 and anti-phospho-NF-κB p65 (Ser536) antibodies were obtained from Cell Signaling Technology (Danvers, MA, USA). Anti-P2X4 antibody was obtained from Alomone Labs (Jerusalem, Israel). Mouse IL-1β, IL-6, IL-13, and tumor necrosis factor (TNF)-α enzyme-linked immunosorbent assay (ELISA) kits were obtained from Thermo Fisher Scientific (Tokyo, Japan). ONO-DI-004 (EP1 agonist), ONO-AE1-259-01 (EP2 agonist), ONO-AE-248 (EP3 agonist), ONO-AE1-329 (EP4 agonist), ONO-8713 (EP1 antagonist), ONO-AE3-208 (EP3 antagonist), and ONO-AE3-240 (EP4 antagonist) were obtained from ONO Pharmaceuticals Co., Ltd. (Osaka, Japan). NP-1815-PX (P2X4 antagonist) was provided by Nippon Chemiphar Co., Ltd. (Tokyo, Japan). All other chemicals were of reagent grade or the highest quality available.

2.2. Animals

P2X4R-deficient (P2rx4/−) mice were generated by Dr. Yamamoto (University of Tokyo), as described previously [23]. C57BL/6 mice were purchased from SLC Japan (Hamamatsu, Japan). All mice were maintained in the animal facility of the Takasaki University of Health and Welfare under specific pathogen-free conditions. All experiments were performed per the Animal Research Committee of Takasaki University of Health and Welfare regulations (Approval No. 2033, 1 April 2020).

2.3. Cell Culture

BMMCs were prepared from the bone marrow of C57BL/6 mice, as described previously [24]. Briefly, bone marrow cells were collected from the femur and cultured in RPMI 1640 growth medium containing 10% fetal bovine serum, 100 units/mL penicillin, 100 μg/mL streptomycin, and 10 ng/mL mouse IL-3. After 2 weeks, recombinant SCF (10 ng/mL) was added to the growth medium and cultured for 2–3 weeks. The culture medium was changed twice per week. After these treatments, almost all (> 95%) cells displayed an MC phenotype, as indicated by CD117 (c-Kit), and FcεRI expression measured using a FACSCant II flow cytometer (BD Biosciences, Tokyo, Japan).

2.4. Measurement of Cytokine Secretion

BMMCs were sensitized with 50 ng/mL anti-DNP IgE overnight in RPMI 1640 growth medium. Cells were washed twice with RPMI medium containing 0.1% bovine serum albumin and then stimulated under various conditions for different periods at 37 °C. The reactions were terminated by centrifugation, and the supernatant and cell pellets were subjected to cytokine measurement using ELISA and RNA extraction, respectively.

2.5. Quantitative RT-PCR

Total RNA was isolated using the GenElute Mammalian Total RNA miniprep kit. First-strand cDNA was synthesized using Moloney murine leukemia virus reverse transcriptase with a 6-mer random primer. Quantitative RT-PCR was performed using an SYBR green kit (Takara bio, Tokyo Japan), as described previously [25].

2.6. Western Blotting

Cells were collected, washed twice with Krebs-Ringer-HEPES buffer (130 mM NaCl, 4.7 mM KCl, 4.0 mM NaHCO3, 1.2 mM KH2PO4, 1.2 mM MgSO4, 1.8 mM CaCl2, and 11.5 mM glucose and 10 mM HEPES (pH 7.4)), and then stimulated under various conditions at 37 °C. The reactions were terminated by adding 4x Laemmli sample buffer. The lysate was separated by 10 % sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and transferred to Immobilon-P polyvinylidene fluoride membranes. The membranes were blocked with 5% non-fat milk for 1 h at room temperature and exposed to primary antibodies overnight at 4 °C and then to secondary antibodies for 2 h at room temperature. The antibodies were diluted as follows: anti-phospho-ERK1/2 (1:1000), anti-ERK1/2 (1:1000), anti-phospho-Akt (1:1000), anti-Akt (1:1000), anti-phospho-NF-κB p65 (1:1000), anti-NF-kB p65 (1:1000), anti-P2X4 (1:2000), and horseradish peroxidase (HRP)-linked anti-rabbit IgG (1:10,000). Immunoreactive proteins were detected by enhanced chemiluminescence (GE Healthcare Bio-sciences, Tokyo, Japan) using Image Reader LAS-3000 (FUJIFILM, Tokyo, Japan). The density of bands was quantified using Multi Gauge version 3.0, and the fold change in phosphorylation was calculated from the amount of phosphoprotein relative to total proteins as a loading index.

2.7. Statistics

All experiments were repeated at least three times, yielding similar results. Data represent the mean ± standard error of the mean (SEM). Statistical analyses were performed using Student’s t-test for two-sample comparisons and one-way analysis of variance (ANOVA) with Dunnett’s two-tailed test for multiple comparisons. Statistical significance was set at p-value < 0.05.

3. Results

3.1. Effects of ATP Combined with PGE2 and Ag on Cytokine mRNA Expression and Secretion in BMMCs

First, we examined the effects of ATP on PGE2 or antigen DNP-HSA-induced mRNA expression of inflammatory cytokines, including IL-1β, IL-6, IL-13, and TNF-α, in BMMCs by quantitative RT-PCR. The BMMCs used in this study hardly expressed the inflammatory cytokine mRNA in the unstimulated condition. However, the co-stimulation of ATP (100 μM) with PGE2 (1 μM) and low doses of antigen DNP-HSA (10 ng/mL) markedly increased cytokine mRNA expression (Figure 1). The increased levels were comparable to the expression levels of glyceraldehyde-3-phosphate dehydrogenase (GAPDH), a housekeeping gene used as an internal standard. Next, we examined whether the change in mRNA reflected cytokine secretion. The stimulation of the BMMCs for 3 h with ATP alone had little effect on the production of IL-6, IL-13, and TNF-α; however, co-stimulation with antigen or PGE2 resulted in a pronounced increase in cytokine secretion. Both the rate and absolute value of the increase were more pronounced when co-stimulated with PGE2 than with antigen (Figure 2). In contrast, there was almost no IL-1β secretion, exhibiting increased mRNA levels (data not shown). Therefore, the subsequent experiments dealt with a detailed analysis of the production of IL-6, IL-13, and TNF-α by co-stimulation with ATP and PGE2.

3.2. Time-Dependent Increase in IL-6 and TNF-α Secretion in Response to ATP, PGE2, and Their Co-Stimulation

Although an increase in cytokine secretion is accompanied by de novo synthesis in response to stimuli, some cytokines, such as TNF-α, are exited as granule content and increased immediately after stimulation. Therefore, the detailed time-course of changes in cytokine secretion was investigated (Figure 3). ATP stimulation alone did not induce significant cytokine secretion until 24 h. In contrast, PGE2 stimulated IL-13 and TNFα secretion at 3 h after stimulation and IL-6 secretion at 6 h. When co-stimulated with ATP and PGE2, a marked increase in IL-6 and IL-13 secretion was observed 3 h after a 1 h time lag. In contrast, TNFα secretion was increased 1 h after co-stimulation with ATP and PGE2 and reached a peak secretion at 3 h. Based on the above results, the following cytokine secretion was examined 3 h after stimulation, whereas mRNA expression was determined 1 h after stimulation.

3.3. Effects of Different Concentrations of ATP and PGE2 on IL-6, IL-13, and TNF-α Secretion

The effects of different concentrations of ATP and PGE2 on IL-6, IL-13, and TNF-α were examined. ATP alone did not release cytokines up to a concentration of 100 μM; however, the secretion of IL-6, IL-13, and TNF-α increased slightly but constantly at 1 mM. In the presence of PGE2 (1 μM), ATP increased the secretion of IL-6, IL-13, and TNFα in a concentration-dependent manner from 1 to 100 μM, and the enhancement was decreased at 1 mM. In contrast, PGE2 elicited IL-6, IL-13, and TNFα secretion at 0.1 to 1 μM even in the absence of ATP; however, its effects were markedly increased in the presence of 100 μM ATP (Figure 4). Based on these results, 100 μM ATP and 0.1 μM PGE2 were selected as optimal concentrations for assessing the synergistic secretion of cytokines in the subsequent experiments.

3.4. Effects of Different Purinergic Receptor Agonists Combined with PGE2 on IL-6, IL-13, and TNF-α mRNA Expression

The BMMCs used in our studies expressed multiple P2 receptors stimulated by ATP, including P2X1, P2X4, P2X7, P2Y1, P2Y2, and P2Y13 [24]. To examine the receptor subtype involved in ATP enhancement of PGE2-induced cytokine secretion, the effects of various nucleotide agonists were examined. The agonists used were P2X1 agonist α,β-methylene ATP (αβmeATP), P2Y1 and P2Y13 agonist ADP, and P2Y2 agonist UTP. Among them, ADP enhanced the secretion of IL-6 and IL-13, but was much weaker than ATP. αβmeATP and UTP did not increase the effects of PGE2 (Figure 5).

3.5. Role of P2X4 Receptor in Enhancing Cytokine mRNA Expression and Secretion by Co-Stimulation of ATP and PGE2

The responses to P2 agonists shown in Figure 5 were similar to the effects on antigen- and PGE2-induced MC degranulation via P2X4R. Therefore, the effect of NP-1815-PX, a P2X4R antagonist [26], on cytokine secretion was investigated. NP-1815-PX did not affect cytokine secretion by PGE2, but significantly suppressed the augmented response induced by ATP (Figure 6). To further clarify the involvement of P2X4R, we examined the BMMCs prepared from P2rx4−/− mice. No increase in cytokine mRNA expression and cytokine secretion was observed in P2rx4−/− BMMCs after co-stimulation with ATP and PGE2 (Figure 7).

3.6. Role of EP3R in Enhancing Cytokine mRNA Expression and Secretion by Co-Stimulation of ATP and PGE2

We examined the receptors that mediate PGE2 action. The synergistic increase in cytokine secretion by the co-stimulation of PGE2 and ATP was reproduced only with the EP3R agonist ONO-AE-248, and EP1, EP2, and EP4 receptor agonists showed no effect (Figure 8A–C). Conversely, the synergistic increase in cytokine secretion by the co-stimulation of PGE2 and ATP was suppressed by the EP3R antagonist ONO-AE3-240, but not by EP1, EP2, and EP4 receptor antagonists (Figure 8D–F). Consistently, the BMMCs predominantly expressed EP3R mRNA, which was unaffected by P2X4R deficiency (Supplementary Figure S1).

3.7. Effects of ATP and PGE2 on ERK/Akt/NF-κB Signaling Pathway

The stimulation of the BMMCs with PGE2 induced the rapid phosphorylation of both ERK1/2 and Akt within 5 min, followed by a decrease to a sustained level at 10 and 20 min. In contrast, ATP had little effect on ERK1/2 and Akt phosphorylation at 5 min, but slightly increased after 10 and 20 min. The co-stimulation of ATP and PGE2 induced a rapid and sustained increase in ERK1/2 and Akt phosphorylation at 10 and 20 min (Figure 9A,C,D). These co-stimulating effects of ATP and PGE2 on ERK1/2 and Akt phosphorylation were absent in P2rx4−/− BMMCs (Figure 9B–D). Akt is known to activate nuclear factor-κB (NF-κB) [27], a transcription factor for various inflammatory cytokines. ATP or PGE2 alone only slightly increased the level of phosphorylated NF-κB p65, an activated form of NF-κB. The co-stimulation with ATP and PGE2 markedly increased the levels of phosphorylated NF-κB (Figure 9A,E). The synergistic phosphorylation of NF-κB by ATP and PGE2 was absent in P2rx4−/− BMMCs (Figure 9B,E). In another set of analysis, we examined the role of P2X4R in the synergistic increase in the phosphorylation of ERK1/2, Akt, and NF-kB by co-stimulation with ATP and PGE2. P2X4R deficiency abolished the increased phosphorylation of ERK, Akt, and NF-κB (Supplementary Figure S2).

3.8. Effects of Signal Transduction Pathway Inhibitors on the Enhanced Cytokine mRNA Expression by Co-Stimulation with ATP and PGE2

The effects of various inhibitors against ERK, p38MAPK, phosphatidylinositol 3-kinase (PI3K), calcineurin, and NF-κB were examined. The elevation of cytokine mRNA levels by the co-stimulation of ATP and PGE2 was attenuated by the ERK inhibitor U0126, PI3K inhibitor wortmannin, NF-κB activation inhibitor NFκBI, calcineurin inhibitor cyclosporine A, intracellular Ca2+ chelator BAPTA-AM and the steroidal anti-inflammatory drug dexamethasone. In contrast, the p38MAPK inhibitor SB203580 did not suppress cytokine mRNA expression (Figure 10)

4. Discussion

In a previous study, we reported that ATP augments IgE-dependent and IgE-independent MC degranulation via P2X4R activation [20,21]. MCs elicit an immediate allergic reaction by releasing chemical mediators stored in the granules; however, in the delayed phase, MCs subsequently secrete diverse inflammatory cytokines and chemokines, causing the migration and activation of other immune cells, resulting in an allergic inflammation [22]. The present study showed that extracellular ATP markedly increased the gene expression and secretion of inflammatory cytokines IL-6 and TNF-α and Th-2 cytokine IL-13 by co-stimulation with antigen and PGE2. Pharmacological analysis using agonists and antagonists revealed that the effects of ATP and PGE2 were mediated by P2X4R and EP3R, respectively, as similar to the effects on degranulation responses [20,21]. The involvement of P2X4R in the ATP-induced effect was confirmed using BMMCs prepared from P2rx4−/− mice. Regarding enhancing the degranulation reaction by ATP, the maximum response was not significantly different between the co-stimulation with antigen and with PGE2 [20,21]. However, the maximum cytokine secretion was much larger in the co-stimulation of ATP with PGE2 than that with the antigen. FcεRI activation with antigen is triggered by tyrosine phosphorylation of the immunoreceptor tyrosine-based activation motif (ITAM) in the FcεRI β subunit. It recruits and activates Lyn kinase resulting in Syk kinase activation, propagating phosphorylation-dependent signals to activate various effector proteins, including phospholipase Cγ [28]. Although FcεRI β subunit is indispensable for FcεRI signaling, the ITAM in the β subunit has an inhibitory regulatory role in MC cytokine production [29], explaining the low cytokine secretion combined with ATP and antigen. In this study, we focused on increased cytokine secretion by co-stimulation with PGE2 and ATP.
Although various ATP receptors are expressed in MCs, the receptor that mediates the ATP-induced enhancement of cytokine secretion in the presence of PGE2 was suggested to be the P2X4 receptor, as described above. No increase in cytokine secretion with ATP in the BMMCs prepared from P2rx4−/− mice and the failure of αβmeATP and UTP to stimulate cytokine secretion indicates that P2X1 and P2Y2 receptors expressed abundantly in the BMMCs [24] do not affect cytokine secretion. Among the P2 receptor agonists tested, ADP slightly but significantly enhanced the gene expression of IL-6 and IL-13. Unlike ATP, ADP did not enhance TNF-α expression; hence, it was not the result of P2X4R stimulation with low affinity but of stimulating the Gq-coupled P2Y1 receptor that recognizes ADP. This result cannot infer that Gq-mediated signals enhance cytokine production, and it is necessary to activate other G proteins, such as G12/13, in addition to Gq [30]. Further studies are needed to validate this hypothesis.
The stimulation of the P2X4 receptor, an ion channel-type receptor, activates the non-selective cation channel, leading to Ca2 + influx and depolarization of the membrane potential [31]. It is interesting to understand such a stimulating mechanism affecting the gene expression of inflammatory cytokines. The effect on cytokine production was markedly different among the P2X receptor subtypes expressed in BMMCs. For example, P2X1 receptor stimulation with αβmeATP has no effect, whereas P2X7 receptor stimulation with a high ATP concentration (~1 mM) alone slightly, but significantly stimulated IL-6, IL-13, and TNF-α secretion. However, unlike the P2X4R-mediated effect, activating the P2X7 receptor did not enhance PGE2-induced cytokine secretion, but rather inhibited P2X4R-mediated effects. In addition to ion channel function, P2X7 receptor stimulation dilate the channel pore, allowing it to pass with high molecular weight substances, affecting the MC membrane function [32]. The three ionotropic P2X1, P2X4, and P2X7 receptors are functionally expressed in human MCs [33]. The difference in the effects of these P2X receptors on cytokine secretion may be useful in elucidating the unique mechanism of P2X4R action.
Increased cytokine secretion by P2X4R stimulation combined with PGE2 was accompanied by increased mRNA levels. The time course of cytokine secretion showed an increase in gene expression before the increase in the secretion of all three cytokines examined. Cytokine secretion was not remarkable 1 h after stimulation when a marked mRNA increase was detected. IL-6, IL-13, and TNF-α secretion increased rapidly during 1–3 h after stimulation, and no continuous increase was observed thereafter, suggesting that P2X4R signaling enhances the transcription of inflammatory cytokines. NF-κB is involved in the transcription of various inflammatory cytokines [22,34]. NF-κB is present in the cytosol as a complex with IκB. When IκB is phosphorylated by IKK, ubiquitination of IκB progresses its degradation by the proteasome, and the dissociated NF-κB translocate into the nucleus to stimulate inflammatory cytokine gene expression [35,36]. Among EP3R downstream signals, ERK1/2 and PI3K-Akt activation participates in NF-κB signaling pathway [26]. PGE2 rapidly increased ERK1/2 and Akt phosphorylation, followed by a decrease to low steady levels. In addition, ATP increased phosphorylation of ERK1/2 and Akt in a time-dependent manner, but this action remained in P2rx4−/− BMMCs, suggesting effects other than P2X4R. However, our results showed that ATP did not affect PGE2-induced ERK1/2 and Akt phosphorylation in the early period (5 min), but significantly increased them at later periods (10–20 min). Importantly, this enhancement of PGE2-induced ERK1/2 and Akt phosphorylation by ATP was absent in P2rx4−/− BMMCs. Furthermore, the PI3K inhibitor WMN, the ERK inhibitor U0126, and the NF-κB inhibitor attenuated the upregulation of inflammatory cytokine expression by ATP in the presence of PGE2. These results suggest that P2X4R stimulation enhances the ERK1/2 and PI3K/Akt/NF-κB signaling pathway and increases inflammatory cytokine gene expression. Consistent with this hypothesis, the steroidal anti-inflammatory dexamethasone, which potently inhibits the NF-κB pathway [37], also inhibited inflammatory cytokine mRNA expression induced by ATP and PGE2. More direct evidence was observed with the change in phosphorylated NF-kB p65, a form of activated NF-κB. Namely, ATP or PGE2 alone induced only a slight increase in phosphorylated NF-κB p65, but their combination markedly increased the phosphorylated NF-κB p65. In addition, this effect was absent in P2rx4−/− BMMCs. The phosphorylation of NF-κB p65 by ATP and PGE2 was rapidly induced within 5 min and returned to basal levels in 20 min. These changes are consistent with the rapid induction of cytokine mRNAs. This also suggests that sustained elevations of ERK and Akt after 10 min are not responsible for NF-κB phosphorylation but may support NF-κB-dependent cytokine gene transcription.
In addition to NF-κB-dependent signaling, a recent study reported by Jordan et al. [38] demonstrated that Ca2+-calcineurin signaling is involved in cytokine secretion induced by co-stimulation with ATP and IL-33. In this study, we also showed that the intracellular Ca2+ chelating agent BAPTA-AM and the calcineurin inhibitor cyclosporine A inhibited inflammatory cytokine mRNA expression induced by ATP and PGE2. P2X4R-mediated Ca2+ influx may stimulate calcineurin-dependent NFAT signaling, resulting in increased inflammatory cytokine gene transcription. Since P2X1 and P2X7 receptors, which input signals qualitatively similar to P2X4R, did not show interplay with EP3R signaling, P2X4R may transmit unspecified and unique signals, which is considered an important issue that needs to be addressed.

5. Conclusions

The present study revealed that extracellular ATP enhanced inflammatory cytokine secretion from stimulated MCs via P2X4R activation. This enhancement was more pronounced with co-stimulation with PGE2 than with antigen. Since both PGE2 and ATP are commonly accumulated in the inflamed microenvironment, these mediators should affect MC function in various physiological and pathophysiological situations. The cross-talk between P2X4R and EP3R signals upregulates an acute degranulation response and following inflammatory cytokine secretion, suggesting that these receptors may be potential therapeutic targets for diseases caused by MCs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cells11040616/s1, Figure S1: Expression of mRNAs for EP receptor subtypes in BMMC obtained from wild type and P2X4 receptor deficient mice. Figure S2: Effects of co-stimulation of ATP and PGE2 on the phosphorylation of ERK1/2, Akt, and NF-κB.

Author Contributions

Conceptualization, K.O., K.Y. (Kazuki Yoshida), M.-a.I. and I.M.; methodology, K.O., K.Y. (Kazuki Yoshida), M.-a.I. and I.M.; formal analysis, K.O.; investigation, K.O., I.M. and T.M.; resources, K.Y. (Kimiko Yamamoto), T.I.; data curation, K.O.; writing—original draft preparation, K.O.; writing—review and editing, I.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Japan Society for the Promotion of Science: Grant-in-Aid for Scientific Research (grant numbers, 18K14925 to K.Y., 19K07328 to I.M., and 20K16010 to M.-a.I.).

Institutional Review Board Statement

The animal study protocol was approved by the Animal Research Committee of Takasaki University of Health and Welfare regulations (Approval No. 2033, 1 April 2020).

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the main text.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Burnstock, G. Pathophysiology and Therapeutic Potential of Purinergic Signaling. Pharmacol. Rev. 2006, 58, 58–86. [Google Scholar] [CrossRef]
  2. North, R.A.; Surprenant, A. Pharmacology of Cloned P2X Receptors. Annu. Rev. Pharmacol. Toxicol. 2000, 40, 563–580. [Google Scholar] [CrossRef]
  3. Jacobson, K.A.; Delicado, E.G.; Gachet, C.; Kennedy, C.; Von Kügelgen, I.; Li, B.; Miras-Portugal, M.T.; Novak, I.; Schöneberg, T.; Perez-Sen, R.; et al. Update of P2Y receptor pharmacology: IUPHAR Review 27. J. Cereb. Blood Flow Metab. 2020, 177, 2413–2433. [Google Scholar] [CrossRef] [PubMed]
  4. Matsuoka, I.; Ohkubo, S. ATP- and adenosine-mediated signaling in the central nervous system: Adenosine receptor activation by ATP through rapid and localized generation of adenosine by ecto-nucleotidases. J. Pharmacol. Sci. 2004, 94, 95–99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Pinna, C.; Knigh, G.E.; Puglisi, L.; Burnstock, G. Neurogenic and non-neurogenic responses in the urinary bladder of hiber-nating hamster. Br. J. Pharmacol. 1998, 123, 1281–1287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Hoyle, C.H.; Knight, G.E.; Burnstock, G. Suramin antagonizes responses to P2-purinoceptor agonists and purinergic nerve stimulation in the guinea-pig urinary bladder and taenia coli. J. Cereb. Blood Flow Metab. 1990, 99, 617–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Nakanish, H.; Takeda, H. The possibility that adenosine triphosphate is an excitatory transmitter in guinea-pig seminal vesicle. Jpn. J. Pharmacol. 1972, 22, 269–270. [Google Scholar] [CrossRef]
  8. Elliott, M.; Chekeni, F.B.; Trampont, P.C.; Lazarowski, E.R.; Kadl, A.; Walk, S.F.; Park, D.; Woodson, R.I.; Ostankovitch, 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] [Green Version]
  9. 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] [PubMed] [Green Version]
  10. Kronlage, M.; Song, J.; Sorokin, L.; Isfort, K.; Schwerdtle, T.; Leipziger, J.; Robaye, B.; Conley, P.B.; Kim, H.-C.; Sargin, S.; et al. Autocrine Purinergic Receptor Signaling Is Essential for Macrophage Chemotaxis. Sci. Signal. 2010, 3, ra55. [Google Scholar] [CrossRef]
  11. Yip, L.; Woehrle, T.; Corriden, R.; Hirsh, M.; Chen, Y.; Inoue, Y.; Ferrari, V.; Insel, P.A.; Junger, W.G. Autocrine regulation of T-cell activation by ATP release and P2X 7 receptors. FASEB J. 2009, 23, 1685–1693. [Google Scholar] [CrossRef] [Green Version]
  12. Karmakar, M.; Katsnelson, M.A.; Dubyak, G.R.; Pearlman, E. Neutrophil P2X7 receptors mediate NLRP3 inflam-masome-dependent IL-1beta secretion in response to ATP. Nat. Commun. 2016, 7, 10555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Tsuda, M.; Shigemoto-Mogami, Y.; Koizumi, S.; Mizokoshi, A.; Kohsaka, S.; Salter, M.W.; Inoue, K. P2X4 receptors induced in spinal microglia gate tactile allodynia after nerve injury. Nature 2003, 424, 778–783. [Google Scholar] [CrossRef]
  14. Bulanova, E.; Bulfone-Paus, S. P2 receptor-mediated signaling in mast cell biology. Purinergic Signal. 2009, 6, 3–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Kurashima, Y.; Amiya, T.; Nochi, T.; Fujisawa, K.; Haraguchi, T.; Iba, H.; Tsutsui, H.; Sato, S.; Nakajima, S.; Iijima, H.; et al. Extracellular ATP mediates mast cell-dependent intestinal inflammation through P2X7 purinoceptors. Nat. Commun. 2012, 3, 1034. [Google Scholar] [CrossRef]
  16. Weber, F.C.; Esser, P.; Müller, T.; Ganesan, J.; Pellegatti, P.; Simon, M.M.; Zeiser, R.; Idzko, M.; Jakob, T.; Martin, S.F. Lack of the purinergic receptor P2X7 results in resistance to contact hypersensitivity. J. Exp. Med. 2010, 207, 2609–2619. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Laffargue, M.; Calvez, R.; Finan, P.; Trifilieff, A.; Barbier, M.; Altruda, F.; Hirsch, E.; Wymann, M.P. Phosphoinositide 3-kinase gamma is an essential amplifier of mast cell function. Immunity 2002, 16, 441–451. [Google Scholar] [CrossRef] [Green Version]
  18. Gao, Z.-G.; Ding, Y.; Jacobson, K.A. P2Y13 receptor is responsible for ADP-mediated degranulation in RBL-2H3 rat mast cells. Pharmacol. Res. 2010, 62, 500–505. [Google Scholar] [CrossRef] [Green Version]
  19. Gao, Z.-G.; Wei, Q.; Jayasekara, M.P.S.; Jacobson, K.A. The role of P2Y14 and other P2Y receptors in degranulation of human LAD2 mast cells. Purinergic Signal. 2012, 9, 31–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Yoshida, K.; Ito, M.-A.; Sato, N.; Obayashi, K.; Yamamoto, K.; Koizumi, S.; Tanaka, S.; Furuta, K.; Matsuoka, I. Extracellular ATP Augments Antigen-Induced Murine Mast Cell Degranulation and Allergic Responses via P2X4 Receptor Activation. J. Immunol. 2020, 204, 3077–3085. [Google Scholar] [CrossRef]
  21. Yoshida, K.; Tajima, M.; Nagano, T.; Obayashi, K.; Ito, M.; Yamamoto, K.; Matsuoka, I. Ito Co-Stimulation of Purinergic P2X4 and Prostanoid EP3 Receptors Triggers Synergistic Degranulation in Murine Mast Cells. Int. J. Mol. Sci. 2019, 20, 5157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Mukai, K.; Tsai, M.; Saito, H.; Galli, S.J. Mast cells as sources of cytokines, chemokines, and growth factors. Immunol. Rev. 2018, 282, 121–150. [Google Scholar] [CrossRef] [PubMed]
  23. Yamamoto, K.; Sokabe, T.; Matsumoto, T.; Yoshimura, K.; Shibata, M.; Ohura, N.; Fukuda, T.; Sato, T.; Sekine, K.; Kato, S.; et al. Impaired flow-dependent control of vascular tone and remodeling in P2X4-deficient mice. Nat. Med. 2006, 12, 133–137. [Google Scholar] [CrossRef] [PubMed]
  24. Yoshida, K.; Ito, M.; Matsuoka, I. Divergent regulatory roles of extracellular ATP in the degranulation response of mouse bone marrow-derived mast cells. Int. Immunopharmacol. 2017, 43, 99–107. [Google Scholar] [CrossRef]
  25. Ito, M.; Matsuoka, I. Inhibition of P2Y6 receptor-mediated phospholipase C activation and Ca2+ signalling by prostaglandin E2 in J774 murine macrophages. Eur. J. Pharmacol. 2015, 749, 124–132. [Google Scholar] [CrossRef]
  26. Matsumura, Y.; Yamashita, T.; Sasaki, A.; Nakata, E.; Kohno, K.; Masuda, T.; Tozaki-Saitoh, H.; Imai, T.; Kuraishi, Y.; Tsuda, M.; et al. A novel P2X4 receptor-selective antagonist produces anti-allodynic effect in a mouse model of herpetic pain. Sci. Rep. 2016, 6, 32461. [Google Scholar] [CrossRef] [Green Version]
  27. Kane, L.P.; Mollenauer, M.N.; Xu, Z.; Turck, C.W.; Weiss, A. Akt-dependent phosphorylation specifically regulates Cot in-duction of NF-kappa B-dependent transcription. Mol. Cell Biol. 2002, 22, 5962–5974. [Google Scholar] [CrossRef] [Green Version]
  28. Gilfillan, A.M.; Rivera, J. The tyrosine kinase network regulating mast cell activation. Immunol. Rev. 2009, 228, 149–169. [Google Scholar] [CrossRef] [Green Version]
  29. Furumoto, Y.; Nunomura, S.; Terada, T.; Rivera, J.; Ra, C. The FcεRIβ immunoreceptor tyrosine-based activation motif exerts inhibitory control on MAPK and IkappaB kinase phosphorylation and mast cell cytokine production. J. Biol. Chem. 2004, 279, 49177–49187. [Google Scholar] [CrossRef] [Green Version]
  30. Ito, M.; Kojima, E.; Yanagihara, Y.; Yoshida, K.; Matsuoka, I. Differential effects of Gq protein-coupled uridine receptor stim-ulation on IL-8 production in 1321N1 human astrocytoma cells. Biol. Pharm. Bull. 2022, in press. [Google Scholar]
  31. Suurväli, J.; Boudinot, P.; Kanellopoulos, J.; Rüütel Boudinot, S. P2X4: A fast and sensitive purinergic receptor. Biomed. J. 2017, 40, 245–256. [Google Scholar] [CrossRef]
  32. Di Virgilio, F.; Dal Ben, D.; Sarti, A.C.; Giuliani, A.L.; Falzoni, S. The P2X7 Receptor in Infection and Inflammation. Immunity 2017, 47, 15–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Wareham, K.; Vial, C.; Wykes, R.; Bradding, P.; Seward, E. Functional evidence for the expression of P2X1, P2X4 and P2X7 receptors in human lung mast cells. J. Cereb. Blood Flow Metab. 2009, 157, 1215–1224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Lee, S.A.; Fitzgerald, S.M.; Huang, S.K.; Li, C.; Chi, D.S.; Milhorn, D.M.; Krishnaswamy, G. Molecular regulation of interleu-kin-13 and monocyte chemoattractant protein-1 expression in human mast cells by interleukin-1β. Am. J. Respir. Cell Mol. Biol. 2004, 31, 283–291. [Google Scholar] [CrossRef] [PubMed]
  35. Baldwin, A.S., Jr. Series introduction: The transcription factor NF-kappaB and human disease. J Clin Invest. 2001, 107, 3–6. [Google Scholar] [CrossRef] [Green Version]
  36. Karin, M.; Ben-Neriah, Y. Phosphorylation meets ubiquitination: The control of NF-[kappa]B activity. Annu. Rev. Immunol. 2000, 18, 621–663. [Google Scholar] [CrossRef] [PubMed]
  37. Ray, A.; Prefontaine, K.E. Physical association and functional antagonism between the p65 subunit of transcription factor NF-kappa B and the glucocorticoid receptor. Proc. Natl. Acad. Sci. USA 1994, 91, 752–756. [Google Scholar] [CrossRef] [Green Version]
  38. Jordan, P.M.; Andreas, N.; Groth, M.; Wegner, P.; Weber, F.; Jager, U.; Kuchler, C.; Werz, O.; Serfling, E.; Kamradt, T.; et al. ATP/IL-33-triggered hyperactivation of mast cells results in an amplified production of pro-inflammatory cyto-kines and eicosanoids. Immunology 2021, 164, 541–554. [Google Scholar] [CrossRef]
Figure 1. Effects of ATP on antigen (Ag)- and PGE2-induced inflammatory cytokine mRNA expressed in the BMMCs. The BMMCs were stimulated with a vehicle (None), DNP-HSA (Ag, 10 ng/mL), or PGE2 (1 μM) in the presence or absence of ATP (100 μM) for 1 h. mRNA expression for (A) IL-1β, (B) IL-6, (C) IL-13, and (D) TNF-α were examined by quantitative RT-PCR. Data were normalized to GAPDH mRNA levels. Values are shown as mean ± SEM (n = 3). * p < 0.05, ** p < 0.01.
Figure 1. Effects of ATP on antigen (Ag)- and PGE2-induced inflammatory cytokine mRNA expressed in the BMMCs. The BMMCs were stimulated with a vehicle (None), DNP-HSA (Ag, 10 ng/mL), or PGE2 (1 μM) in the presence or absence of ATP (100 μM) for 1 h. mRNA expression for (A) IL-1β, (B) IL-6, (C) IL-13, and (D) TNF-α were examined by quantitative RT-PCR. Data were normalized to GAPDH mRNA levels. Values are shown as mean ± SEM (n = 3). * p < 0.05, ** p < 0.01.
Cells 11 00616 g001
Figure 2. Effects of ATP on Ag- and PGE2-induced inflammatory cytokine secretion in the BMMCs. The BMMCs were stimulated with a vehicle (None), DNP-HSA (Ag, 10 ng/mL), or PGE2 (1 μM) in the presence or absence of ATP (100 μM) for 3 h. Cytokine levels of (A) IL-6, (B) IL-13, and (C) TNF-α were examined by ELISA. Values are shown as mean ± SEM (n = 3). ** p < 0.01.
Figure 2. Effects of ATP on Ag- and PGE2-induced inflammatory cytokine secretion in the BMMCs. The BMMCs were stimulated with a vehicle (None), DNP-HSA (Ag, 10 ng/mL), or PGE2 (1 μM) in the presence or absence of ATP (100 μM) for 3 h. Cytokine levels of (A) IL-6, (B) IL-13, and (C) TNF-α were examined by ELISA. Values are shown as mean ± SEM (n = 3). ** p < 0.01.
Cells 11 00616 g002
Figure 3. Effects of ATP on the time course of PGE2-induced cytokine secretion in the BMMCs. The BMMCs were stimulated with a vehicle (None), DNP-HSA (10 ng/mL) or PGE2 (1 μM) in the presence or absence of ATP (100 μM) for 1, 3, 6, 12, and 24 h. Cytokine levels of (A) IL-6, (B) IL-13, and (C) TNF-α were measured by ELISA. Values are shown as mean ± SEM from triplicate determination.
Figure 3. Effects of ATP on the time course of PGE2-induced cytokine secretion in the BMMCs. The BMMCs were stimulated with a vehicle (None), DNP-HSA (10 ng/mL) or PGE2 (1 μM) in the presence or absence of ATP (100 μM) for 1, 3, 6, 12, and 24 h. Cytokine levels of (A) IL-6, (B) IL-13, and (C) TNF-α were measured by ELISA. Values are shown as mean ± SEM from triplicate determination.
Cells 11 00616 g003
Figure 4. Effects of different concentrations of ATP or PGE2 on cytokine secretion induced by co-stimulation of ATP and PGE2 in the BMMCs. The BMMCs were stimulated with different concentrations of ATP (left) or PGE2 (right) in the presence or absence (alone) of PGE2 (1 μM) or ATP (100 μM), respectively, for 3 h. (A) IL-6, (B) IL-13, and (C) TNF-α levels in the reaction medium were measured by ELISA. Data are shown as percentage of maximum response obtained with ATP 100 μM + PGE2 (1 μM). Values are shown as mean ± SEM (n = 3).
Figure 4. Effects of different concentrations of ATP or PGE2 on cytokine secretion induced by co-stimulation of ATP and PGE2 in the BMMCs. The BMMCs were stimulated with different concentrations of ATP (left) or PGE2 (right) in the presence or absence (alone) of PGE2 (1 μM) or ATP (100 μM), respectively, for 3 h. (A) IL-6, (B) IL-13, and (C) TNF-α levels in the reaction medium were measured by ELISA. Data are shown as percentage of maximum response obtained with ATP 100 μM + PGE2 (1 μM). Values are shown as mean ± SEM (n = 3).
Cells 11 00616 g004
Figure 5. Effects of different P2 receptor agonists on PGE2-induced cytokine secretion in the BMMCs. The BMMCs were stimulated with a vehicle (None), ATP (100 μM), α,β-methylene ATP (αβmeATP, 10 μM), ADP (100 μM), or UTP (100 μM) in the presence (black column) or absence (open column) of PGE2 (0.1 μM) for 1 h. mRNA expression for (A) IL-6, (B) IL-13, and (C) TNF-α were examined by quantitative RT-PCR. Data were normalized with GAPDH mRNA levels. Values are shown as mean ± SEM (n = 3). * p < 0.05, ** p < 0.01 vs. PGE2 alone (none).
Figure 5. Effects of different P2 receptor agonists on PGE2-induced cytokine secretion in the BMMCs. The BMMCs were stimulated with a vehicle (None), ATP (100 μM), α,β-methylene ATP (αβmeATP, 10 μM), ADP (100 μM), or UTP (100 μM) in the presence (black column) or absence (open column) of PGE2 (0.1 μM) for 1 h. mRNA expression for (A) IL-6, (B) IL-13, and (C) TNF-α were examined by quantitative RT-PCR. Data were normalized with GAPDH mRNA levels. Values are shown as mean ± SEM (n = 3). * p < 0.05, ** p < 0.01 vs. PGE2 alone (none).
Cells 11 00616 g005
Figure 6. Effects of P2X4R antagonist NP-1815-PX on cytokine secretion induced by co-stimulation of ATP and PGE2 in the BMMCs. The BMMCs were preincubated vehicle (control) and NP-1815-PX (10 μM) for 10 min and then stimulated with vehicle (-) or PGE2 (0.1 μM) in the presence or absence of ATP (100 μM) for 3 h. (A) IL-6, (B) IL-13, and (C) TNF-α levels in the reaction medium were measured by ELISA. Values are shown as mean ± SEM (n = 3). ** p < 0.01.
Figure 6. Effects of P2X4R antagonist NP-1815-PX on cytokine secretion induced by co-stimulation of ATP and PGE2 in the BMMCs. The BMMCs were preincubated vehicle (control) and NP-1815-PX (10 μM) for 10 min and then stimulated with vehicle (-) or PGE2 (0.1 μM) in the presence or absence of ATP (100 μM) for 3 h. (A) IL-6, (B) IL-13, and (C) TNF-α levels in the reaction medium were measured by ELISA. Values are shown as mean ± SEM (n = 3). ** p < 0.01.
Cells 11 00616 g006
Figure 7. Lack of ATP-induced upregulation of PGE2-induced cytokine production in P2X4R-deficient (P2rx4−/−) mice. The BMMCs, prepared from wild-type (WT) and P2rx4−/− mice, were stimulated with vehicle (None) or PGE2 (0.1 μM) in the presence or absence of ATP (100 μM) for (AC) 1 h or (DF) 3 h. mRNA expression for (A) IL-6, (B) IL-13, and (C) TNF-α were examined by quantitative RT-PCR. Data were normalized with GAPDH mRNA levels. Values are shown as mean ± SEM (n = 3). ** p < 0.01. The levels of (D) IL-6, (E) IL-13, and (F) TNF-α in the reaction medium were measured by ELISA. Values are shown as mean ± SEM (n = 3). ** p < 0.01.
Figure 7. Lack of ATP-induced upregulation of PGE2-induced cytokine production in P2X4R-deficient (P2rx4−/−) mice. The BMMCs, prepared from wild-type (WT) and P2rx4−/− mice, were stimulated with vehicle (None) or PGE2 (0.1 μM) in the presence or absence of ATP (100 μM) for (AC) 1 h or (DF) 3 h. mRNA expression for (A) IL-6, (B) IL-13, and (C) TNF-α were examined by quantitative RT-PCR. Data were normalized with GAPDH mRNA levels. Values are shown as mean ± SEM (n = 3). ** p < 0.01. The levels of (D) IL-6, (E) IL-13, and (F) TNF-α in the reaction medium were measured by ELISA. Values are shown as mean ± SEM (n = 3). ** p < 0.01.
Cells 11 00616 g007
Figure 8. Involvement of EP3R in PGE2-induced enhanced cytokine secretion in the presence of ATP in BMMCs. (AC) The BMMCs were stimulated with PGE2, ONO-DI-004 (EP1 agonist), ONO-AE1-259 (EP2 agonist), ONO-AE-248 (EP3 agonist), or ONO-AE1-329 (EP4 agonist) at 0.1 μM with or without ATP (100 μM) for 3 h. The levels of (A) IL-6, (B) IL-13, and (C) TNF-α in the reaction medium were measured by ELISA. Values are shown as mean ± SEM (n = 3). (DF) The BMMCs were preincubated with a vehicle, ONO-8713 (EP1 antagonist), AH6809 (EP2 antagonist), ONO-AE3-240 (EP3 antagonist), and ONO-AE3-208 (EP4 antagonist) at 1 μM for 5 min and then stimulated with PGE2 (0.1 μM) and ATP (100 μM) for 3 h. The levels of (D) IL-6, (E) IL-13, and (F) TNF-α in the reaction medium were measured by ELISA. Values are shown as mean ± SEM (n = 3).
Figure 8. Involvement of EP3R in PGE2-induced enhanced cytokine secretion in the presence of ATP in BMMCs. (AC) The BMMCs were stimulated with PGE2, ONO-DI-004 (EP1 agonist), ONO-AE1-259 (EP2 agonist), ONO-AE-248 (EP3 agonist), or ONO-AE1-329 (EP4 agonist) at 0.1 μM with or without ATP (100 μM) for 3 h. The levels of (A) IL-6, (B) IL-13, and (C) TNF-α in the reaction medium were measured by ELISA. Values are shown as mean ± SEM (n = 3). (DF) The BMMCs were preincubated with a vehicle, ONO-8713 (EP1 antagonist), AH6809 (EP2 antagonist), ONO-AE3-240 (EP3 antagonist), and ONO-AE3-208 (EP4 antagonist) at 1 μM for 5 min and then stimulated with PGE2 (0.1 μM) and ATP (100 μM) for 3 h. The levels of (D) IL-6, (E) IL-13, and (F) TNF-α in the reaction medium were measured by ELISA. Values are shown as mean ± SEM (n = 3).
Cells 11 00616 g008
Figure 9. Effects of PGE2, ATP, and co-stimulation of PGE2 and ATP on ERK 1/2, Akt and NF-κB phosphorylation in the BMMCs. (A,B) The BMMCs prepared from WT (A) and P2rx4−/− mice (B) were stimulated with PGE2 (0.1 μM), ATP (100 μM), and PGE2 and ATP for 5, 10, and 20 min. Cell lysates were subjected to Western blot analysis for phospho-Akt and total Akt (upper), phospho-ERK 1/2 and total-ERK 1/2 (middle), or phospho-NF-κB and total-NF-κB (lower). Blots are representative of three independent experiments. (C) The densitometry value of protein bands of phospho-Akt (C), phospho-ERK 1/2 (D), and phospho-NF-κB (E) are shown as relative intensities, with the results obtained with unstimulated cells (Time 0) designated as 1. Data are shown as mean ± SEM (n = 3). The effect of co-stimulation of PGE2 and ATP on ERK 1/2, Akt, and NF-κB phosphorylation is significantly different from that of PGE2 alone. * p < 0.05, ** p < 0.01.
Figure 9. Effects of PGE2, ATP, and co-stimulation of PGE2 and ATP on ERK 1/2, Akt and NF-κB phosphorylation in the BMMCs. (A,B) The BMMCs prepared from WT (A) and P2rx4−/− mice (B) were stimulated with PGE2 (0.1 μM), ATP (100 μM), and PGE2 and ATP for 5, 10, and 20 min. Cell lysates were subjected to Western blot analysis for phospho-Akt and total Akt (upper), phospho-ERK 1/2 and total-ERK 1/2 (middle), or phospho-NF-κB and total-NF-κB (lower). Blots are representative of three independent experiments. (C) The densitometry value of protein bands of phospho-Akt (C), phospho-ERK 1/2 (D), and phospho-NF-κB (E) are shown as relative intensities, with the results obtained with unstimulated cells (Time 0) designated as 1. Data are shown as mean ± SEM (n = 3). The effect of co-stimulation of PGE2 and ATP on ERK 1/2, Akt, and NF-κB phosphorylation is significantly different from that of PGE2 alone. * p < 0.05, ** p < 0.01.
Cells 11 00616 g009
Figure 10. Effects of various inhibitors on the elevation of cytokine mRNA levels induced by the co-stimulation with ATP and PGE2. (AC) The BMMCs were preincubated with a vehicle, MEK1/2 inhibitor U0126 (U, 5 μM), p38MAPK inhibitor SB203580 (SB, 10 μM), PI3K inhibitor wortmannin (W, 0.1 μM), NF-κB activation inhibitor (κBI, 1 μM), calcineurin inhibitor cyclosporine A (CyA, 5 μM), intracellular Ca2+ chelating agent BAPTA-AM (5 μM) and steroidal anti-inflammatory dexamethasone (Dex, 0.1 μM) for 10 min, followed by stimulation with PGE2 (0.1 μM) with or without ATP (100 μM) for 1 h. mRNA expression of (A) IL-6, (B) IL-13, and (C) TNF-α was examined by quantitative RT-PCR. Data were normalized with GAPDH mRNA levels and shown as the percentage of the response to PGE2 and ATP. Values are shown as mean ± SEM (n = 6). * p < 0.05, ** p < 0.01 vs. the response to ATP and PGE2 and + p < 0.05, ++ p < 0.01 vs. the response to PGE2.
Figure 10. Effects of various inhibitors on the elevation of cytokine mRNA levels induced by the co-stimulation with ATP and PGE2. (AC) The BMMCs were preincubated with a vehicle, MEK1/2 inhibitor U0126 (U, 5 μM), p38MAPK inhibitor SB203580 (SB, 10 μM), PI3K inhibitor wortmannin (W, 0.1 μM), NF-κB activation inhibitor (κBI, 1 μM), calcineurin inhibitor cyclosporine A (CyA, 5 μM), intracellular Ca2+ chelating agent BAPTA-AM (5 μM) and steroidal anti-inflammatory dexamethasone (Dex, 0.1 μM) for 10 min, followed by stimulation with PGE2 (0.1 μM) with or without ATP (100 μM) for 1 h. mRNA expression of (A) IL-6, (B) IL-13, and (C) TNF-α was examined by quantitative RT-PCR. Data were normalized with GAPDH mRNA levels and shown as the percentage of the response to PGE2 and ATP. Values are shown as mean ± SEM (n = 6). * p < 0.05, ** p < 0.01 vs. the response to ATP and PGE2 and + p < 0.05, ++ p < 0.01 vs. the response to PGE2.
Cells 11 00616 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Obayashi, K.; Yoshida, K.; Ito, M.-a.; Mori, T.; Yamamoto, K.; Imai, T.; Matsuoka, I. Synergistic Cytokine Production by ATP and PGE2 via P2X4 and EP3 Receptors in Mouse Bone-Marrow-Derived Mast Cells. Cells 2022, 11, 616. https://doi.org/10.3390/cells11040616

AMA Style

Obayashi K, Yoshida K, Ito M-a, Mori T, Yamamoto K, Imai T, Matsuoka I. Synergistic Cytokine Production by ATP and PGE2 via P2X4 and EP3 Receptors in Mouse Bone-Marrow-Derived Mast Cells. Cells. 2022; 11(4):616. https://doi.org/10.3390/cells11040616

Chicago/Turabian Style

Obayashi, Kosuke, Kazuki Yoshida, Masa-aki Ito, Tetsuya Mori, Kimiko Yamamoto, Toshiyashu Imai, and Isao Matsuoka. 2022. "Synergistic Cytokine Production by ATP and PGE2 via P2X4 and EP3 Receptors in Mouse Bone-Marrow-Derived Mast Cells" Cells 11, no. 4: 616. https://doi.org/10.3390/cells11040616

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