Next Article in Journal
Tumor Necrosis Factor Alpha: Implications of Anesthesia on Cancers
Next Article in Special Issue
Modulatory Properties of Aloe secundiflora’s Methanolic Extracts on Targeted Genes in Colorectal Cancer Management
Previous Article in Journal
The Combination of ATM and Chk1 Inhibitors Induces Synthetic Lethality in Colorectal Cancer Cells
Previous Article in Special Issue
EZH2: An Accomplice of Gastric Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

An Update of G-Protein-Coupled Receptor Signaling and Its Deregulation in Gastric Carcinogenesis

1
Department of Anatomical and Cellular Pathology, State Key Laboratory of Translational Oncology, Sir Y.K. Pao Cancer Center, Prince of Wales Hospital, The Chinese University of Hong Kong, Hong Kong 999077, China
2
State Key Laboratory of Digestive Disease, Institute of Digestive Disease, The Chinese University of Hong Kong, Hong Kong 999077, China
3
CUHK-Shenzhen Research Institute, The Chinese University of Hong Kong, Shenzhen 518000, China
4
Department of Pediatrics, The Chinese University of Hong Kong, Hong Kong 999077, China
5
Department of Medicine and Therapeutics, The Chinese University of Hong Kong, Hong Kong 999077, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 3 November 2022 / Revised: 15 January 2023 / Accepted: 19 January 2023 / Published: 25 January 2023
(This article belongs to the Special Issue Advances in Molecular Mechanisms of Gastrointestinal Tumors)

Abstract

:

Simple Summary

Gastric cancer (GC) ranks as one of the most life-threatening malignancies worldwide, and over one billion new cases and 783,000 deaths were reported last year. The incidence of GC is exceptionally high in Asian countries. Multiple oncogenic signaling pathways are aberrantly activated and implicated in gastric carcinogenesis, leading to the malignant phenotype acquisition. G-protein-coupled receptor (GPCR) signaling is one of them, and the aberrant activation of GPCRs and G proteins promotes GC progression. The activated GPCRs/G proteins might serve as useful biomarkers for early diagnosis, prognostic prediction, and even clinically therapeutic targets. This review summarized the recent research progress of GPCRs and highlighted their mechanisms in tumorigenesis, especially in GC initiation and progression.

Abstract

G-protein-coupled receptors (GPCRs) belong to a cell surface receptor superfamily responding to a wide range of external signals. The binding of extracellular ligands to GPCRs activates a heterotrimeric G protein and triggers the production of numerous secondary messengers, which transduce the extracellular signals into cellular responses. GPCR signaling is crucial and imperative for maintaining normal tissue homeostasis. High-throughput sequencing analyses revealed the occurrence of the genetic aberrations of GPCRs and G proteins in multiple malignancies. The altered GPCRs/G proteins serve as valuable biomarkers for early diagnosis, prognostic prediction, and pharmacological targets. Furthermore, the dysregulation of GPCR signaling contributes to tumor initiation and development. In this review, we have summarized the research progress of GPCRs and highlighted their mechanisms in gastric cancer (GC). The aberrant activation of GPCRs promotes GC cell proliferation and metastasis, remodels the tumor microenvironment, and boosts immune escape. Through deep investigation, novel therapeutic strategies for targeting GPCR activation have been developed, and the final aim is to eliminate GPCR-driven gastric carcinogenesis.

1. Introduction

Gastric cancer (GC) is a substantial global health burden, accounting for the fifth most commonly diagnosed cancer and the third leading cause of fatal malignancies worldwide. Incidence rates are markedly increased in Eastern Asia, especially in Mongolia, Japan, and Korea, which are strongly associated with various predisposing and etiological factors, according to several migrant studies [1,2]. Most GC-related deaths occur due to late diagnosis, lymph node metastasis, and refractory after surgery. Thus, numerous efforts have been made to develop useful prognosis markers for early detection and therapeutic targets to improve clinical outcomes. Heterogeneity represents one of the biggest challenges in GC treatment owing to the histological categories and diverse molecular drivers. The well-established histological classification divides gastric carcinomas into diffuse and intestinal types [3]. The Cancer Genome Atlas (TCGA) network also reaffirmed our understanding of molecular categories by analyzing the dysregulated pathways identified in multiomics data. This study developed a robust molecular classification scheme comprising Epstein-Barr virus (EBV), microsatellite instability (MSI), chromosomal instability (CIN), and genomically stable (GS) tumors [4]. In the past two decades, trastuzumab and chemotherapy were used as the first-line treatment, and the combination of ramucirumab and paclitaxel was used in second-line treatment [5]. However, the clinical applicability remains quite limited. There is an urgent need to uncover more targetable pathways to develop more accurate diagnosis makers against nonspecific symptoms in early-stage GC and optimize existing therapy for precision medicine.
Since G protein-coupled receptors (GPCRs) were reported in cellular transformation in 1986, emerging evidence shows that these membrane-embedded receptors regulate many biological processes and are crucial targets against several human malignancies [6]. The involvement of GPCRs in GC is emerging due to the identification of genomic aberrations that lurk at different stages and subtypes of GC and promote tumor initiation and progression [7]. This review recapitulated the current knowledge related to the aberrated regulation of the GPCR pathway in GC, including the common tactic hijacked by tumor cells for their growth, metastasis, and immune evasion. Moreover, we will discuss the advances in the current treatment strategies and summarize the ongoing clinical trials that attempt to translate biological findings into clinical applications.

2. Basic Knowledge of GPCRs

GPCRs comprise over 800 members accounting for about 4% of human genes. They have various structures and signal transduction. Based on their specific characteristics, GPCR members are further classified into different subgroups and participate in various physiological processes, whereas the aberrant expression and abnormal activation of GPCRs are associated with tumor progression.

2.1. Structure and Classification of GPCRs

GPCRs have seven transmembrane α-helices (TM1-7) that connect the N-terminal extracellular domain (ECD), three extracellular and intracellular loops: ECL1-3, ICL1-3, and the C-terminus (Figure 1). They are classified into six groups based on their structural and functional similarities, whereas only four groups (A, B, C, and F) are found in vertebrates. The Rhodopsin-like class A, which has 719 members, represents humans’ most common but diverse group. Half of the class A members serve as sensor receptors primarily in smell and vision. In contrast, diffusible ligands, such as peptides, lipids, hormones, and nucleotides, can trigger the other receptors. Class B includes secretin and adhesion receptors, which have a similar sequence in 7TM but different sequences in ECD. The secretin subgroup contains receptors for polypeptide gut hormones, such as the GLP-1 receptor, glucagon receptor, and parathyroid hormone receptor. Research has focused on the adhesion receptors by defining the mechanisms of ligand binding sites and the GPCR autoproteolysis-inducing domain (GAIN)-mediated receptor activation. The metabotropic glutamate family (class C) is characterized by a large ECD, consisting of γ-aminobutyric acid B receptors (GABAB), metabotropic glutamate receptors (mGluRs), and a calcium-sensing receptor (CasR). The frizzled/taste family (class F) includes frizzled and smoothened proteins that can be activated by the lipo-glycoproteins of the Wnt and Hedgehog families [8,9,10].
Thrilling technologies, such as X-ray crystallography, cryo-electron microscopy (cryo-EM) [11,12], and nuclear magnetic resonance (NMR) spectroscopy [13], are facilitating the exploration of the structures of GPCRs and further boosting the structure-based drug design. More extensive knowledge of GPCR biology, particularly on the function-related conformational equilibria, such as allosteric coupling, biased signaling, and dynamic modulation, is required for developing new targets and minimizing the side effects. Although structural studies reveal the direct impacts of stimulation on the receptor conformation changes, the result of the GPCR signaling pathway is regulated by various factors, including cell backgrounds, receptor expression levels, and agonists’ kinetic characteristics.

2.2. Signal Transduction of GPCRs

By binding to various extracellular ligands, GPCRs change their conformation, activate G-coupling proteins, and couple with other proteins, such as β-arrestins and GPCR kinases (GRKs). However, the determinants of GPCR selective binding have not been fully understood. Fundamentally, the G-protein barcode determines the selectivity between GPCRs and G proteins, which contain variable residues on the conserved positions and can be recognized by different GPCRs [14]. G proteins have long been recognized as the primary transducers of GPCRs, and β-arrestins are suggested as essential modulators in genome-edited cells [15]. β-arrestins can identify and bind to the GRK-phosphorylated GPCRs and thereby outcompete the G proteins, functioning as scaffolds for other signalings [16,17]. Therefore, the GRKs regulate the phosphorylation of ligand-bound GPCRs and balance the G-protein-dependent and β-arrestin-dependent nodes of GPCRs [18].
When the activated GPCRs bind with the heterotrimeric G proteins, the Gα subunits dissociate from Gβγ after the exchange of GDP with GTP on the Gα proteins. This process releases the heteromeric G proteins from the GPCRs and retains the plasma’s GTP-bound Gα and Gβγ subunits [19]. One GPCR can activate multiple Gα proteins encoded by 16 genes. These genes are classified into four subfamilies based on sequence similarity: Gi, Gq, Gs, and G12/13. Generally, the Gαs and Gαi respectively promote or inhibit adenylyl cyclase, thus intervening in cyclic adenosine monophosphate (cAMP) production. The accumulating cAMP behaves as a second messenger to activate protein kinase A (PKA). Members of the Gi family also activate phospholipase (PIs) and phosphodiesterases (PDEs), ultimately modulating the opening of numerous ion channels. PLC-β is the effector of the Gαq and Gβγ subunits that elevate intracellular Ca2+ levels and activate protein kinase C (PKC) by converting phosphatidylinositol 4,5-biphosphate (PIP2) into inositol 1,4,5-triphosphate (PIP3) and diacylglycerol (DAG). Additionally, the Gβγ subunits trigger the PI3K-γ/PI3K-β to catalyze the conversion of PIP2 to PIP3, which is a direct stimulator of AKT. The Gα12/13 and Gαq family members regulate a group of Rho GEFs, which harbor an RGS homology domain and activate Rho GTPase. As shown in Figure 2A, the segregated Gα and Gβγ subunits evoke kinase cascades by activating various second messengers [20,21].
β-arrestins undergo conformational changes when recognizing the GRK-phosphorylated GPCRs. Then, they enhance the process of desensitization, internalization, and clathrin-mediated endocytosis of the activated GPCRs. As scaffold proteins, β-arrestins facilitate GPCR-stimulated signal transduction. As one of the most prominent and earliest examples, the GPCR-mediated extracellular signal-regulated kinase 1/2 (ERK1/2) activation is a β-arrestin-dependent and G protein-independent signaling [16]. The genetic ablation or inactivation of several G proteins induces a zero functional state for the G protein and abolishes the β-arrestin-mediated signaling in response to GPCR activation [22]. However, it was reported that β-arrestins are not required for ERK1/2 phosphorylation despite their crucial roles in receptor internalization [23]. Indeed, the cumulative impact of GPCR-induced ERK1/2 activation is tightly controlled by β-arrestins and G proteins [24]. Moreover, GPCRs scaffold several signaling proteins for Wnt [25], the hedgehog (Hh) [26], and Notch [27] pathways (Figure 2B,C).

2.3. Diversification of GPCR Machinery

GPCRs are sophisticated dynamic machines rather than static on-and-off switches. When they are engaged with different ligands, receptors, and regulatory partners, they may exhibit specific conformations and undergo subcellular distributions. Exploring the dynamic nature of GPCRs is vital to elucidate the mechanisms underlying allosteric modulation, biased agonism, oligomerization, and sustained and compartmentalized signaling. These mechanisms convey novel insights into drug discovery.
Allosteric ligand binding sites in GPCRs are potential new targets for modulating GPCR functions and improving drug selectivity. These modulators augment (positive allosteric modulators [PAMs]) or reduce (negative allosteric modulators [NAMs]) the affinity and efficacy of endogenous agonists [28]. The discovery of allosteric modulators has sparked interest in central nervous system (CNS) diseases, though with limited success [29]. MK-7622 is a PAM, selectively binding with the M1 muscarinic receptor in the CNS, which has been stopped because it fails to improve recognition and increases adverse effects like diarrhea [30,31].
Another ligand-receptor dynamic is biased agonism, a mechanism in which the active conformational states of the receptors are stabilized by some ligands, resulting in distinct cellular signaling profiles [32]. There are three different modes of biased signaling, including the same receptor bound with other ligands adopting distinct conformations (ligand bias), varying stoichiometric ratios of signaling effectors in distinguished cells (system bias), and GPCR stimulation within divergent intracellular compartments (location bias) [17]. However, substantial evidence on this is limited. The endogenous ligands, CCL9 and CCL21, have been considered equipotent for activating CCR7-G protein coupling and calcium mobilization. However, both ligands cause the distinct conformation of CCR7. Only CCL9 can promote robust receptor desensitization after coupling the β-arrestins and efficiently accelerating ERK1/2 phosphorylation, which CCL21 cannot achieve [33]. Besides, small molecules targeting TRV130 and PZM21 have been utilized to improve analgesia with fewer side effects because of the biased receptor μ-OR activity, potent Gαi signaling profile, and limited β-arrestin recruitment [34,35]. These two examples can partly explain how ligands trigger the biased mechanism of GPCRs. Revealing the structural features of GPCRs under multiple activation states and different cellular backgrounds may be required to understand the biased signaling.
Receptor oligomerization conveys much more diversities in the function and physiological roles of GPCRs. However, unlike the oligomer tyrosine kinase receptors and ion channels, the formation of GPCR multimers remains controversial [36,37]. It has been found that dimerization was necessary for some GPCRs, such as the class C members. The first tangible evidence for GPCR dimerization was that the gonadotrophin-releasing hormone (GnRH) antagonist-conjugated bivalent antibody played an essential role in biphasic receptor formation [38,39]. The emergence of heteromers was associated with the preferential pattern of receptors in different tissues and cell types [40]. The balance between the monomers and heteromers of GPCRs may contribute to diseases [41,42]. Unraveling the pattern of GPCR heteromers will provide pharmacotherapeutic targets to benefit disease management.
Compartmentalized signaling may partly explain why the GPCRs can activate a typical profile of secondary messengers and kinases. In addition to locating the membrane, the GPCRs might be desensitized and undergo β-arrestin-mediated endocytosis and intracellular signaling [43,44,45]. Notably, some of the mechanisms are studied in the digestive systems. Recently, PAR2 endosomal may underlie the sustained hyperexcitability of nociceptors in patients with irritable bowel syndrome (IBS). The IBS supernatants and trypsin could persistently activate PAR2 in the colonic mucosa in a clathrin-mediated, endocytosis-dependent fashion [46,47]. The inhibitors of clathrin-mediated endocytosis and targeted PAR2 antagonists suppressed PAR2 endosomal signal [48].

2.4. Dysregulated GPCR Signaling in Tumors

Based on the recent pan-cancer analysis, GPCR signaling was among the 55 pathways most significantly mutated. Mutations and the aberrant expression of GPCRs and G proteins contribute to various diseases, including neurodegenerative, reproductive, immunological, and metabolic disorders, as well as cancers and infectious diseases [49,50]. The dysregulated GPCR signaling may exert a significant tumorigenic effect, as those alterations frequently co-occur in well-characterized oncogenes, such as tyrosine and serine-threonine kinase Ras-family members [51]. In-depth omics analysis approaches, like MutSig2CV and GISTIC (Genomic Identification of Significant Targets in Cancers), have comprehensively investigated the mutations and copy number variations (CNVs) of GPCRs and G proteins in 33 TCGA patient groups. Remarkably, mutated GPCRs and G proteins have been significantly identified in GI malignancies, even though these tumors’ mutation rates are not typically high [52]. Therefore, the relevance between these mutations and biological outcomes is vastly underestimated [53,54].

2.5. GPCR Mutation and Abundant Expression

GPCRs are mutated in over 20% of all sequenced samples [55,56]. Unlike the mutated hotspots in G proteins, GPCRs exhibit diverse mutations across different cancer types. The three-dimensional structures of GPCRs and their interaction elements were evaluated to acquire a mutational landscape of GPCRs in cancers [51]. The bulk of the alterations occurs in the conserved 7TM via the visualization of the representative GPCR 3D structure, such as the ionic lock switch E/DRY arginine motif, G protein-binding sites, and the tyrosine toggle switch motif NPxxY, and ligand-binding site. GPCR mutations impair GPCR signaling by altering the basal activity, ligand binding affinity, G-protein interaction, and cell-surface expression. As with thyroid-stimulating hormone receptors like HCRT2, P2RY12, LPAR4, and GPR174, frequent mutations in the DRY motif may result in constitutive activation due to conformational changes in TM3, TM5, and TM6 [51]. Mutations of the connection between the NPxxY motif on TM7 and a conserved tyrosine in TM5 could stabilize the inactive-state conformations of the α1B- and β2-adrenoceptors, which may account for lower agonist potency in transducing the downstream IP1 and cAMP signaling, respectively [57]. Understanding the mutated structural features will shed new light on GPCR malfunctions and devise possible therapeutic strategies [58].
The large and ever-grossing body of sequencing by pan-cancer analysis suggests that the frequently mutated GPCR families are adhesion-related GPCRs, such as the glutamate metabotropic receptors (GRM1-8, class C) [59], lysophosphatidic acid (LPA) receptors (LPAR1-6), sphingosine-1-phosphate (S1P) receptors (S1PR1-5), and muscarinic receptors (CHRM1-5, class A). However, most adhesion receptors are orphan receptors with unknown ligands and physiological functions [60,61]. GPCR genetic alternations were found in melanoma by exon capture and massively parallel sequencing. GPR98 and GRM3 were two of the most frequently mutated genes, with 27.5% and 16.3% mutation rates. GRM3 mutants selectively mediate the MEK signaling that contributes to tumor growth in melanoma, acting as an indicator for patient stratification and precision medicine [62]. MutSig2CV analysis suggests the three most mutant GPCRs in colon cancer are GPR98 (21.25%), TSHR (13.90%), and BAI3 (13.62%), while CELSR1 (11.20%), EDNRB (8.14%), and GPR45 (5.09%) account for the three most frequently mutated GPCRs in GC. However, the functional roles of these mutant GPCRs in GI cancers remain unknown.
Besides mutations, GPCRs, like chemokine and histamine receptors (HRH2), exhibit significant copy number variations (CNVs) in tumors. Several broad-type GPCRs are universally overexpressed throughout the GI tract, regulating digestive and pathophysiological processes [54,63,64]. The upregulation of receptors like 5-HTRs, FFARs, HRs, PARs, EPs, and TGRs plays pivotal roles in proliferation, invasion, metastasis, and inflammation in the small intestine and colon. It has been reported that the CNVs of chemokine receptors, LPARs, and ARs contribute to the initiation and progression of hepatocellular carcinoma [65,66]. Early studies have implicated numerous viruses that harbored open reading frames and evolved to take advantage of the signaling network for replicative success by encoding GPCRs [67]. In GC, the Epstein-Barr virus (EBV/HHV-4) encodes a class A GPCR called BILF1, affecting multiple cellular pathways [68].

2.6. Widespread Mutations of G Proteins

As oncogenic drivers in multiple prevalent cancers, many G proteins are considered part of the cancer-associated gene panels routinely employed by a wide range of clinical oncology studies. MutSig2CV analysis indicates that GNAS is the most frequently mutated G protein in TCGA cohorts, concordant with the sequence results of the catalog of somatic mutations in the cancer (COSMIC) database. GNAS aberrations widely occur in tumors originating from the pituitary (28%), pancreas (12%), thyroid (5%), colon (6%), and a few other locations [69]. Previous studies revealed that the two most frequently mutated residents, Arg 201 [70,71] and Gln 227 [72], might be functionally important. The significance of these two sites has been first confirmed in pituitary tumors [73]. The disease-causing altered resident Arg 201 leads to the constitutive cAMP signaling by reducing the GTP hydrolysis of the active GTP-bound Gαs. However, the conclusion was reshaped by a recent structural study of GNAS, indicating that the stabilization of the intramolecular hydrogen bond network (H-bond network) plays a pivotal role in mutation-mediated constitutive activation [74]. These aberrations in GNAS are responsible for initiating and progressing multiple types of GI cancer, such as colon neoplasia, GC, and pancreatic adenocarcinomas (PDAs). In colon cancer, the GNAS can mediate the tumorigenesis of inflammatory factors by stimulating the Gs-Axin-β-Catenin pathway axis [75]. In the rare gastric adenocarcinoma, GNAS mutations were tightly associated with deep submucosal invasion and increased tumor size by activating the Wnt/β-catenin pathway [76]. Besides, at the early onset of invasive PDAs, frequent GNAS mutations (~41–75%) suppressed the PKA-mediated SIK and reprogrammed lipid metabolism in the precursor of PDAs [77,78].
Although GNAQ and GNA11 mutations were less studied in tumors than GNAS, these mutations were well-established in Sturge-Weber syndrome [79] and leptomeningeal melanocytosis, arising from the central nervous system (~50%) [80], and also in the blue nevi and the primary uveal melanomas (UVM)/uveal melanoma metastases (83%) [81,82]. The somatic mutations are mainly located in the residues Q209 or R183, which are essential for GTP hydrolysis and cause constitutive activation due to loss of GTPase activity. In uveal melanoma (UVM), the more common Q209 mutations were more potent in tumorigenesis assays in nude mice models [82]. Consistently, the mutant GNAQQ209L contributed to MAPK pathway activation [81] and exhibited more significant activated ERK than the GNAQR183Q [83,84]. The activated GNAS mutant can also stimulate YAP-dependent transcription through a Trio-Rho/Rac signaling circuitry instead of the canonical Hippo pathway in UVM [85]. Furthermore, GNA13 is upregulated in several solid tumors, such as GC [86], nasopharyngeal cancer [87], breast cancer [88], squamous cell cancers [89], and colorectal cancer [90]. Interestingly, both GNA13 and RhoA have shown relevance to the transformation capacity and metastatic potential in epithelial cancer and fibroblasts, but the axis appears to play a tumor-suppressive role in B-cell lymphomas [91]. Large-scale sequencing of lymphoid and hematopoietic malignancies indicated that the mutant residues could be found throughout the gene [92,93].
The cDNA library screening distinguished the functionally relevant mutations of the Gβ proteins GNB1 and GNB2. The gain-of-function alterations of these proteins can disrupt the interactions of Gα-Gβγ and constitutively stimulate the downstream signaling effectors, conferring resistance to targeted kinase inhibitors [94,95]. Recently, emerging variants in all five Gβ proteins have been reported, such as GNB2 Arg52Leu in familial cardiac arrhythmia condition, Gly77Arg in neurodevelopmental disorder, and monoallelic missense variants in developmental delay/intellectual disability (DD/ID) [96]. Emerging evidence supports that Gβ mutants also occur in various cancer types and relate to distinct cancer subtypes. GPCRs also transduce the signal through β arrestins instead of G proteins, mediating GC cell invasion, migration, and epithelial-mesenchymal transition (EMT) [24]. For example, the protein kinase AKT exerts its oncogenic function through the signaling complex GPR39/β arrestin1/Src upon obestatin stimulation [97].
Some mutations occur in oncogenic kinase alterations, such as BCR-ABL fusion protein, ETV6-ABL1, JAKV617F, and BRAFV600K, to enhance the drug resistance of the corresponding kinase inhibitors [95,98]. Nevertheless, further investigations are needed into how these alterations influence tumorigenesis in different contexts. The potential roles of Gγ proteins, the close partners of Gβ proteins, should be clarified.

3. Aberrant GPCR Signaling in GC

GPCRs play hierarchical roles in many signaling networks. Dysregulations of the GPCRs extensively exist in tumor progression, metastasis, and immune response reprogramming. In recent years, aberrant GPCR members have been emerging in GC studies. This section will outline the updated findings of the GPCR signaling pathway in GC (Table 1).

3.1. Proliferation and Apoptosis

Mounting evidence has unveiled the multilayered crosstalk between GPCRs and proliferation- and apoptosis-related signaling circuits. The representative ones involve EGFR transactivation, MAPK cascades, the PI3K-AKT-mTOR pathway, and the Hippo signaling pathway [21,135].

3.1.1. Transactivation in the EGFR and MAPK/ERK Pathway

GPCRs share many similarities with the tyrosine kinase receptors, such as EGFR and the MAPK/ERK signaling pathways [136], in regulating cell proliferation. The EGFR-mediated signaling pathway can be ligand-dependent or independent [137,138,139]. The “three membrane-passing signal (TMPS)” model is an EGFR ligand-dependent route. The activated RTKs are triggered by activated GPCRs and subsequently activate the extracellular signal-regulated kinase (ERK)/mitogen-activated protein kinase (MAPK) cascade. On the other hand, GPCR-mediated Src activation contributed to EGFR phosphorylation more directly. Both modes have been uncovered in GC. S1P could mediate the progression of GC via Gi- and matrix metalloprotease (MMP)-independent c-Met- and EGFR-transactivation [113]. However, the S1P- or LPA-induced transactivation of ERBB2 (also known as HER2) required the activation of MMP and the tyrosine kinase activity of EGFR [110]. In addition, the knockdown of the membrane-type bile acid receptor (M-BAR)/TGR5 suppressed the deoxycholate (DC)-induced phosphorylation of EGFR, and DC transactivates EGFR through M-BAR- and ADAM/HB-EGF-dependent mechanisms [124]. Infection with H. pylori boosted the expression of interleukin-8 (IL-8), which promoted cell proliferation by inducing EGFR transactivation [140]. Due to oncogenic activation and PGE2-EP4 pathway induction, the ubiquitous overexpression of the EGFR ligands and Adams have been identified in mouse gastric tumors [141]. PGE2-induced uPAR expression has also been implicated in the activation of Src, c-Jun NH2-terminal kinase (JNK), extracellular signal-regulated kinase (Erk), and p38 mitogen-activated protein kinase (p38 MAPK) [142,143]. GPCRs may also directly trigger MAPK cascades, establishing a connection between the external stimuli and their effect factors. These effectors may be further subdivided into four core categories: ERK1/2, JNK1-3, p38α-δ MAPKs, and ERK5. The LPAR2 inhibitor suppressed the proliferative and migration abilities of GC cell line SGC-7901 through the LPAR2/Gq11/p38 pathway, suggesting that LPAR2 might be a potential target for GC treatment [112]. Protease-activated receptor family (PAR1-4) also exerted pro-carcinogenic effects via the overactivated ERK1/2-MAPK pathway. For example, the reduction of EPCR impeded PAR1 activation, thus resulting in the downregulation of phosphorylated ERK1/2 and the suppression of the proliferation and migration of GC tumor cells [144].

3.1.2. Activation of the PI3K-AKT-mTOR Pathway

PI3K is stimulated by the activated RTKs or GPCRs, ultimately leading to the synthesis of PIP3 and the recruitment of oncogenic effectors such as the serine/threonine kinase AKT. The PH domain in AKT permitted its binding with PIP3, contributing to the membrane accumulation and subsequent phosphorylation at T308 and S473 by PDK1 and mTORC2 [145]. Even though over 100 AKT substrates have been discovered in different settings, the associated mechanisms for most substrates have not been fully delineated [145]. mTOR is one of the AKT subtracts that is well-established to promote biosynthetic processes for cell growth. Since PI3K/AKT/mTOR signaling has also been identified as an ideal drug target for gastric carcinoma, the regulators may have a role in improving treatment design [146]. Indeed, some GPCRs have been proven to influence the activity of AKT in GC cells, such as the leucine-rich repeat-containing receptor Lgr6, adenosine receptor A2a, and the orphan receptor GPR39. Lgr6 was identified to empower GC cell proliferation by activating the PI3K/AKT/mTOR pathway [109]. Another GPCR, adenosine receptor A2a, was engaged in PI3K/AKT-regulated proliferation and migration in GC [123]. Additionally, GPR39 provided GC cells with a growth advantage by boosting the activity of AKT in an EGFR-dependent manner [97].

3.1.3. Regulation of the Hippo Pathway

The canonical Hippo pathway kinase cascade is a critical tumor suppressor pathway, and its dysregulation has been widely implicated in organ size modulation and carcinogenesis [147]. The core components of the Hippo pathway are composed of STE20-like protein kinase 1/2 (MST1/2) and large tumor suppressor 1/2 and the major functional output Yes-associated protein 1 (YAP) and WW domain-containing transcription regulator protein 1 (WWTR1, also known as TAZ). Because there is a lacking DNA binding site in YAP/TAZ, TEF1-4 (TEAD1-4) is characterized as a bona fide transcription enhancer factor [147,148]. GPCRs have been found to control the Hippo pathway positively and negatively as a significant regulator of the intracellular pathway. The initial implication that GPCRs modulate Hippo signaling through LATS1/2 came from the study in serum starvation cells [149]. Two components, LPA and S1P, have been identified as the effective factors in serum that are responsible for YAP/TAZ activation through the recognition of the corresponding GPCRs. The LPA/S1P-mediated GPCR activation facilitates YAP/TAZ dephosphorylation via the G protein-cytoskeleton circuit. This study has laid the foundation for how YAP/TAZ senses the diffusible extracellular signals. However, several questions have also been raised after the initial discovery. Given that GPCRs constitute ~800 members and each GPCR can be coupled to diverse G proteins [19], the integrated effects on YAP/TAZ modulation remain elusive. The case can be more complicated when the dysregulation of G proteins and GPCRs is frequently determined in cancers [54]. Moreover, different G proteins stimulate the dephosphorylation of YAP/TAZ to various degrees. GPCRs can trigger YAP/TAZ activation by interacting with Gα12/13, Gαi/o, and Gαq/11 or suppress YAP/TAZ by binding with Gαs. However, which of these that Rho is involved with has not yet been identified; it is also unclear how the actin cytoskeleton regulates Lats1/2 phosphorylation. Emerging findings have revealed how specific GPCRs may fine-tune YAP/TAZ in given cellular surroundings [150]. The triggered LPA receptors have been demonstrated to play crucial roles in activating YAP/TAZ, causing tumor progression in the colon, ovarian, prostate, and breast [151,152]. The S1P-mediated S1P receptors contribute to hepatocellular carcinoma by coupling to Gα12/13 and stimulating YAP [153], connecting GPCR signaling to the Hippo pathway. Except for LPARs and S1PRs, the other GPCR-initiated signals can influence YAP/TAZ activity, including polypeptides (Angiotensin II, Thrombin, glucagon, etc.) [154,155], metabolites (purines, fatty acids, epinephrine, glutamate, etc.) [156,157], and hormonal factors (estrogen, endothelin-1, etc.) [158,159]. These signals have been widely indicated in human malignancies and are critical cell niche or microenvironment components. Recent studies pointed out that the mesenchymal niche manipulated the initiation of colorectal cancer by the rare peri-cryptal Ptgs2-expressing fibroblasts, and these fibroblasts exhibited paracrine control over tumor-initiating stem cells via the PGE2-EP4-Yap signaling axis [160]. The GPCR-Hippo crosstalk was also identified in GC stem-like cells: PAR1 stimulated the Hippo-YAP pathway and affected invasion, metastasis, and multidrug resistance [161]. As such, the GPCR regulation of YAP/TAZ has emerged as a driver, or as a potential therapeutic target, in gastric neoplasia. However, another study has found that AMOT, rather than Lats1/2, serves as the bridge between GPCR-mediated cytoskeleton changes and YAP/TAZ modulation in uveal melanoma cells, with an activated mutation at Arg183 and Gly209 in GNAQ (encoding for Gαq) and GNA11 (encoding for Gα11), respectively [85]. As a result, the findings provide novel explanations for the alternations in actin dynamics induced by GPCR signals, which are somewhat different from previous studies. Therefore, this warrants exploring the interplay between AMOT, Lats1/2, and the actin cytoskeleton in GC, as the mechanic stress is context-dependent.

3.1.4. GPCR-Signaling Integration and Crosstalk with Other Pathways in GC

Besides the above signaling circuits, other pathways have also been linked to GPCR-mediated oncogenicity in GC. These pathways involve the Notch pathway [162], hedgehog (Hh) signaling [163], and the Wnt/β-catenin pathway [164]. The Hh pathway is crucial for GC cell growth and cancer stem cell maintenance, and its activation has been highlighted in diffuse-type GC [165,166]. Smoothened (Smo, a member of class F) and Gpr161 (an orphan member) can function as positive and negative regulators in the Hh pathway, respectively [167]. The Wnt/β-catenin pathway is involved in tissue homeostasis and embryonic development. As Wnt (Wingless/Int1) stimulates the frizzled receptor (FZD, class F GPCRs), both G-protein independent and dependent signaling can be established [164,168].

3.2. GPCRs-Driven Metastasis of GC

Metastasis is how cancer cells establish ‘bench-heads’ in other organs or anatomical sites instead of the initial lesion, and it is responsible for more than 90% of cancer-related mortality [169] (Figure 3A). The most prevalent sites for GC metastasis are the liver, lung, bone, and lymph nodes [170]. Since Paget’s ‘seed and soil’ hypothesis laid the fundamental basis for metastasis, many investigators have contributed to a better understanding of the process. Several studies have identified the sequential multistep in GC metastasis: invasion into the surrounding tissue and the degradation of the basement membrane (BM), intravasation into the blood vessels or lymphatic systems, survival and translocation to distant tissues, extravasation into the foreign environment, and finally, colonization to proliferate and form a macroscopic secondary neoplasm [170,171,172]. As the complexity and relevance of metastasis have previously been widely reviewed, we will focus mainly on the roles and mechanisms of GPCRs during the invasion, BM degradation, and angiogenesis processes in GC.

3.2.1. Inducing Epithelial-Mesenchymal Transition (EMT), Migration, and Invasion

Epithelial-to-mesenchymal transition (EMT) is regarded as the initial step for invasion, featuring a loss of cell polarity and integrity and the acquisition of motile mesenchymal characteristics. The pathologic activation of the EMT program is primarily executed by transcription factors (including SNAI1/2, TWIST1/2, and ZEB1/2) and microRNAs, ultimately resulting in the accumulation of the genes associated with mesenchymal properties, such as vimentin, fibronectin, and N-cadherin [173]. Though the above-mentioned molecular mechanisms are still lacking regarding GPCR-driven EMT in GC, GPCR signaling dysregulation is still frequently connected to EMT, migration, and invasion processes via dynamically regulating the downstream effectors and downstream cascades [174].
Although chemokines are tiny polypeptides (8-14kDa), they display pleiotropic effects in cancers. The chemokine system comprises nearly 50 chemokines that bind to 20 different chemokine receptors or four atypical chemokine receptors (ACKRs) [175]. This superfamily is distinguished by a substantial degree of redundancy, inferring that these chemokines can bind to different clusters of receptors and vice versa [176]. Intrinsic genetic or epigenetic regulators governed their expression and environmental cues such as hypoxia, microbiota, and metabolic [176]. For example, epigenetic regulator histone deacetylase 1 (HDAC1) suppressed CXCL8 expression by antagonizing the active nuclear transcription factor NF-κB [177]. Hypoxia has been revealed to induce the expression of CXCR4, CXCR7, and CXCL12 in different cancer cells, with the binding sites of hypoxia-inducible factor 1 (HIF1) as the promoters of these genes [178,179,180]. In GC, the elevated CXCL8 concentration was proved to be tightly correlated with the tumor stage instead of H. pylori infection, as shown in previous studies [181,182,183]. Several clinical investigations have suggested that the upregulation of chemokines and receptors was associated with GC pathogenesis, indicating that the specific chemokines might serve as potential diagnostic and therapeutic targets [184,185].
Chemokine receptors have attracted considerable attention due to their involvement in GC metastasis. A notable correlation was found between CCR7 expression and gastric carcinoma lymph node metastasis via stepwise regression analysis [186,187]. Strikingly, about 67% of primary gastric tumors exhibited CXCR4-positive expression [188]. The high concentration of the CXCR4 ligand CXCL12 has been validated in the malignant ascitic fluids from peritoneal carcinomatosis, and elevated CXCL12 is tightly correlated with the dissemination of GC cells to distant organs [188]. CXCL12-stimulated CXCR4 enhanced NF-κB and STAT3 signaling activation and, in turn, led to its transcriptional upregulation, which formed a positive feedback loop. This loop is linked to EMT, migration, and invasion in GC [189]. In response to CXCL12, CXCR4 also conferred the GC cell EMT and metastasis process via stimulating mTOR and some well-known oncogenic kinases: EGFR, SRC, or c-MET [190,191]. The crosstalk between TGF-β1 and the NF-κB pathway was triggered by the CCL2-CCR2 axis, leading to EMT-related protein upregulation [192]. Besides, chemokine receptors also induced organ-specific metastasis. CXCR4 and CCR7 are the primary receptors guiding the metastasizing cells, including GC cells [193]. Moreover, the high levels of CCR9 in melanoma, breast, and ovarian cancer make them efficiently translocate to the highly CCL25-expressing small intestine [194,195,196].
Many other GPCRs also govern the development of GC invasion. For example, GPER1 inhibition blocked EMT in GC cells by inhibiting the PI3K/AKT pathway [197]. Similar regulation that is mediated by adenosine receptor 2 (A2aR) or GPR30 could also be observed in GC [122,123]. In addition, the MAPK cascades were activated by the formyl peptide receptor 2 (FP2R), S1P2R, muscarinic acetylcholine receptor 3 (M3R), P2Y receptors (P2YR), and γ-Aminobutyric acid receptor A (GABAA), thus contributing to the invasion and metastasis in GC [104,105,114,119,121,128]. Many other GPCRs have also been linked to GC metastasis, while the underlying mechanisms are unknown. For instance, the angiotensin II receptor type 1/2 (AT1R/AT2R) has been locally upregulated and indicated to carry a much higher risk of nodal spread [101].

3.2.2. Degrading the Barriers to Invasion

BM, a specialized extracellular matrix (ECM), plays a critical role in normal epithelium tissue architecture. BM disruption is a must for cancer cells leaving the primary location, controlled by the balance between the expression of MMPs and their tissue inhibitors (TIMPs) [198,199]. The expression of MMPs was upregulated by a histamine-H2 receptor or Thrombin-PAR1 signaling [143]. H. pylori was reported to be crucial during the invasion by upregulating cyclooxygenase-2 (COX-2) through ATF2/MAPK stimulation. The COX-2 inhibitor or EP2 receptor antagonist repressed angiogenesis and tumor invasion via the uPA system, which is a determinant factor in transforming the zymogen plasminogen into plasmin for degrading the ECM constituents [200]. Furthermore, the bacterial pathogen of H. pylori’s consistent infection manipulated a variety of extracellular proteases [201], but the exact mechanisms need further exploration. The interactions between microbial metabolites and GPCRs may provide new insights into the complicated process [202].

3.2.3. Driving Angiogenesis

Angiogenesis is the process of vessel splitting from pre-existing vessels and is essential for tumorigenesis and progression, especially for those solid tumors exceeding 1–2 mm in diameter, as it provides oxygen and nutrients [203,204]. Many GPCRs exerted pro-tumor effects by promoting tumor-associated angiogenesis, notably Thrombin receptors, S1PRs, lysophosphatidic acid receptors (LPARs), and Prostaglandin receptors (Figure 3B). PAR1 is necessary for physio-pathological angiogenesis since poor vasculature development results in animal embryos dying after PAR1 deprivation. Thrombin-mediated PARs cleavage upregulates the transcription of many proangiogenic genes, such as VEGF and its receptor VEGFR, MMP2, angiopoietin-2 (Ang-2), and others [99,100,205,206]. Moreover, endothelial differentiation gene 1 (Edg1)/S1P1R is the first reported GPCR in blood vessel formation. Furthermore, the intrauterine death of Edg1 ablation mice happened mainly due to abnormal angiogenesis [207,208]. The Gα12/Gα13-coupled receptors LPA4 and LPA6 synergistically regulate endothelial Dll4 expression through YAP/TAZ activation, which mediates sprouting angiogenesis [209]. Moreover, H. pylori-induced VEGF upregulation was activated through p38 MAPK COX2-PEG2-EP2/4 signaling [210]. Other orphan receptors are also involved in tumor angiogenesis, such as KSHV-GPCR, GPR124, ELDT1, and GPER [211].

3.3. Remodeling the Tumor Microenvironment (TME) to Promote Immune Escape

TME acts as a unique niche populated by multiple cell types (including cancer cells, immune cells, and stromal cells), ECM, and diverse secreted factors (such as exosomes and microRNAs) [212,213]. The altered TME landscape is related to tumor progression, metastasis, and therapeutic responses [214]. Recently, the sophisticated TME infiltration pattern of GC (termed as TMEscore) has been defined based on the assessment of 22 immune cell types and cancer-associated fibroblasts (CAFs), which were correlated with genomic characteristics and pathologic features [212]. The biology and function of CAFs have emerged as an area of active investigation and have been reviewed elsewhere [215,216]. The compositions of infiltrated immune cells within TME varied greatly, and one of the most important mechanisms involved the chemokines and their receptors [176].
It is noteworthy that chemokines and chemokine receptors can be ubiquitously expressed in tumor cells, immune cells, and stromal cells [217]. Alternations in chemokines and their receptors shaped the TME immune cell constitution and remodeled the immune responses, some of which are hijacked by tumor cells to avoid immune surveillance and elimination [218]. The antitumor immune responses were driven by the recruiting immune cells, mainly including dendritic cells (DCs), CD8+ T cells, natural killer (NK) cells, and M1 macrophages. GC with a high CXCR3 expression level was shown to have increased DC and T cell infiltration. The CXCR3/CXCL4 or CXCR3/CXCL4L1 axis is necessary to recruit DCs as they elicit potent antitumor functions through substantially stimulating T cells and activating the related humoral response [219,220]. Similarly, CXCR3 also plays a vital role in CD8+ T cell infiltration that directly damages the tumor cells after being differentiated into cytotoxic CD8+ T cells [221,222]. In addition, NK cells represent professional killer cells, whose accumulation in the TME is the consequence of upregulated CXCL10 and CXCL12 signaling through CCR7 or CXCR3 [223,224].
On the other hand, chemokine signaling is also involved in the formation of immune-suppressive TME, where tumors evolve to escape recognition and clearance. This process has been largely linked to the infiltration of diverse protumor immune cell populations, such as regulatory T (Treg) cells, the M2 macrophages, monocytic myeloid-derived suppressor cells (M-MDSCs), and granulocytic (or PMN-) MDSCs [176,225]. CCL22, mainly produced by tumor cells (or macrophage-mediated), causes an abundance of Treg cells in TME via interacting with the receptor CCR4 on the surface of Treg. Another receptor, CCR10, in Treg cells also facilitated their migration in response to CCL28 [226,227]. Moreover, the nonpolarized macrophages (M0) originating from the recruited monocytes can be differentiated into two main subtypes, M1 and M2 macrophages, exhibiting extremely distinct functions toward cancers. These transitions depended on a large spectrum of chemokine signals. Active monocyte recruitment required tumor-derived chemokine releases, such as CCL2, CCL3, CCL4, CCL5, CCL20, and CCL18. Additionally, the blockade of the CCL2-CCR2 circuit led to M2 macrophage accumulation, whereas CCL11 skewed macrophages toward an M2 phenotype [228,229,230,231,232,233]. MDSCs were subdivided into two major groups: polymorphonuclear MDSCs (PMN-MDSCs) and mononuclear MDSCs (M-MDSCs). CXCR2 specifically mediated the migration of PMN-MDSCs to the tumor site by binding with CXCL1/CXCL2/CXCL5, whereas the accumulation of M-MDSCs requires CCL2-CCR2 interaction. Functionally, MDSCs employed diverse mechanisms to suppress T cell functioning, mainly through releasing high levels of arginase 1 (Arg1), reactive oxygen species (ROS), and nitric oxide (NO). Further research also suggested additional mechanisms, including the upregulation of COX2 and PGE2 in these MDSCs [234,235] (Figure 3C).
Other GPCRs are also involved in the regulations of immune responses. For example, prostaglandin (PG) production can mediate inflammation through its cognate GPCR EP1-EP4 (PTGER1-4). PGs, especially the PGE2, were produced by the cyclo-oxygenases COX-1and COX-2, the inhibitors (nonsteroidal anti-inflammatory drugs (NSAIDs)) of which have been utilized to comfort pains and reduce the incidence of a broad range of cancer types. The role of PGE2 has been extensively studied for inducing inflammation by stimulating other signaling pathways, including the Toll-like receptor (TLR)/MyD88 pathway [236], Wnt, and EGFR signal [237]. PAR1-deficient mice infected with H. pylori may suffer from severe gastritis due to lacking suppressing macrophage cytokine secretion and cellular infiltration [238]. In addition, TGR5 antagonized gastric inflammation by inhibiting the transcription activity of NF-κB signaling [239].
The involvement of GPCRs in immune remodeling is summarized in Table 2.

4. Therapeutic Strategies for Targeting GPCRs in GC

Despite the improving clinical outcomes, advanced GC patients benefit little from traditional surgery or chemotherapy and suffer from painful lives [240]. Personalized medicine and targeted therapy have been introduced to clinical applications for over two decades. For instance, trastuzumab has been integrated into the treatment for HER2-expressing patients, and ramucirumab has been utilized for VEGFR2-positive GC individuals [241]. Immune checkpoint inhibitors (ICI) also have been investigated as a frontline treatment [242,243,244,245]. Meanwhile, biomarkers and novel targeted therapies have been intensely investigated for advanced GC [246]. Substantial progress has been made by deciphering the functions of GPCR members in GC progression. However, only a handful of drugs that target GPCRs have been conducted in clinical trials for GC treatment (Table 3).
In order to accelerate the GPCR-targeted drug development for GC, many groups have identified potent compounds to inhibit or enhance the activity of GPCRs. However, the structures are only available for small partial GPCRs (~50 GPCRs), and 54% of GPCRs are orphans that are under-exploited. Machine learning approaches may be employed for predicting the interaction between immersed compounds and GPCRs based on established high-quality structural models [248]. With the evolving knowledge of GPCR pathways, we will be able to identify more effective drugs in formats, tissue-specific drug delivery systems, and appropriate treatment periods. Small molecules are the most prevalent GPCR modulators, while biologics are receiving more and more attention because of their versatility and specificity [249]. Antibodies, including antibody fragments and variable antibody domains, function with great penetration traits and are attracting considerable interest in drug development. Downregulated targeted GPCRs via RNA interference (RNAi) can represent potential approaches to gene therapy [250].
More efficient drug delivery systems with enhanced solubility and stability, lower dosages, and less toxicity have been developed, such as nanomaterials, nanocarriers, nanoconjugation, and nanoencapsulation techniques [251,252]. Furthermore, several solid tumors have well-established patient-derived xenografts (PDX) and xenograft-derived organoid models. These preclinical platforms recapitulated the genotypic and phenotypic landscape, endowed with a high predictive value for high-throughput drug screening. Nevertheless, they still have limitations, such as intratumor heterogeneity, compromised immune systems, and diverse tumor environments in GC [253].

5. Summary and Future Perspectives

GPCRs govern multiple signaling pathways and regulate GC development in various aspects. The heterogeneous and complicated characteristics of GPCRs contribute to GC heterogeneity and result in the current untimely diagnosis and inefficiency of therapeutic applications. Not only can GPCRs transduce the extracellular changes to the intracellular signaling circuits, but the conformational changes of GPCRs can also continuously influence intracellular events. Aberrant GPCR activation and mutated GPCRs/G proteins can fuel cancer cell proliferation, migration, invasion, angiogenesis, and metastasis. In addition, the dysregulation of GPCRs affords advantages for immunosuppressive TME and drug resistance to malignancies.
Although much progress has been made on novel biomarker identification and molecular mechanism investigation, the current GPCR-based diagnosis and therapy in GC are far from clinically available. It is urgent that GPCR signaling-targeted therapy be developed. In future studies, several issues need to be addressed. First, as the mutation rates of GPCRs and G proteins are prominent in some cancer types, the development of small molecules that target the driver mutations is urgent. Second, because of the heterogeneity of the cancer cells and tumor microenvironment, we need to comprehensively appraise the activation of GPCR signaling and its crosstalk by using cutting-edge techniques such as scRNA-seq or scDNA-seq. Last but not least, more preclinical models based on patient-derived samples, such as organoids or xenografts, need to be developed to evaluate the efficacies and side effects of the screened drugs. With the deep investigation of the molecular mechanisms of GPCR signaling and the multicenter clinical trials, more therapeutic strategies will be delivered for targeting GPCR signaling, which will benefit GC patients.

Author Contributions

W.K. provided direction on this manuscript. H.Y. and J.-L.Z. reviewed the literature and drafted the manuscript. K.-T.L., K.-W.L., J.Y. and K.-F.T. reviewed and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study is supported by the National Natural Science Foundation of China (NSFC) (No. 82272990).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the BioRender for the Figure generation (https://biorender.com/, accessed on 14 December 2022) and the technical support from Core Utilities of Cancer Genomics and Pathobiology, The Chinese University of Hong Kong.

Conflicts of Interest

The authors declare no competing interests.

Abbreviations

A2aR, adenosine receptor 2; ACKRs, atypical chemokine receptors; ADAM17, metallopeptidase domain 17; Ang-2, angiopoietin-2; Arg1, arginase 1; AT1R/AT2R, angiotensin II receptor type 1/2; BM, basement membrane; CAFs, cancer-associated fibroblasts; cAMP, cyclic adenosine monophosphate; CHRMs, muscarinic receptors; CIN, chromosomal instability; CNS, central nervous system; CNVs, copy number variations; COX-2, cyclooxygenase-2; cryo-EM, cryo-electron microscopy; DAG, diacylglycerol; DCs, dendritic cells; DD/ID, developmental delay/intellectual disability; EBV, Epstein Barr virus; ECD, extracellular domain; ECL, extracellular loop; ECM, extracellular matrix; Edg1, endothelial differentiation gene 1; EMT, epithelial-mesenchymal transition; ERK1/2, extracellular signal-regulated kinase 1/2; FP2R, formyl peptide receptor 2; GABAB, γ-aminobutyric acid B receptors; GAIN, autoproteolysis-inducing domain; GC, GC; GISTIC, Genomic Identification of Significant Targets in Cancers; GnRH, gonadotrophin-releasing hormone; GPCRs, G protein-coupled receptors; GRKs, GPCR kinases; GRMs, glutamate metabotropic receptors; GS, genomically stable; HDAC1, histone deacetylase 1; Hh, hedgehog; HIF1, hypoxia-inducible factor 1; HRHs, histamine receptors; IBS, irritable bowel syndrome; ICI, Immune checkpoint inhibitors; ICL, intracellular loop; JNK, c-Jun NH2-terminal kinase; LPA, lysophosphatidic acid; M-BAR, membrane-type bile acid receptor; M-MDSCs, monocytic myeloid-derived suppressor cells; M3R, muscarinic acetylcholine receptor 3; MAPK, mitogen-activated protein kinase; mGluRs, metabotropic glutamate receptors; MMP, matrix metalloprotease; MSI, microsatellite instability; MST1/2, STE20-like protein kinase 1/2; NAMs, negative allosteric modulators; NK cells, natural killer cells; NMR, nuclear magnetic resonance; NO, nitric oxide; NSAIDs, nonsteroidal anti-inflammatory drugs; PAMs, positive allosteric modulators; PARs, Protease-activated receptors; PDAs, pancreatic adenocarcinomas; PDEs, phosphodiesterases; PDX, patient-derived xenografts; PIP2, phosphatidylinositol 4,5-biphosphate; PIP3, inositol 1,4,5-triphosphate; PIs, phospholipases; PKA, activate protein kinase A; PKC, protein kinase C; PMN-MDSCs, polymorphonuclear MDSCs; RNAi, RNA interference; ROS, reactive oxygen species; S1P, sphingosine-1-phosphate; TCGA, The Cancer Genome Atlas; TEAD1-4, TEF1-4; TIMPs, tissue inhibitors of metalloprotease; TME, tumor microenvironment; TMPS, three membrane-passing signal; Treg, regulatory T; UVM, uveal melanoma; WWTR1, WW domain-containing transcription regulator protein 1; XDO, xenograft-derived organoid; YAP, Yes-associated protein 1.

References

  1. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [Green Version]
  2. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2020. CA Cancer J. Clin. 2020, 70, 7–30. [Google Scholar] [CrossRef]
  3. Lauren, P. The Two Histological Main Types of Gastric Carcinoma: Diffuse and So-Called Intestinal-Type Carcinoma. An Attempt at a Histo-Clinical Classification. Acta Pathol. Microbiol. Scand. 1965, 64, 31–49. [Google Scholar] [CrossRef] [PubMed]
  4. Cancer Genome Atlas Research, N. Comprehensive molecular characterization of gastric adenocarcinoma. Nature 2014, 513, 202–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Smyth, E.C.; Moehler, M. Late-line treatment in metastatic gastric cancer: Today and tomorrow. Ther. Adv. Med. Oncol. 2019, 11, 1758835919867522. [Google Scholar] [CrossRef]
  6. Usman, S.; Khawer, M.; Rafique, S.; Naz, Z.; Saleem, K. The current status of anti-GPCR drugs against different cancers. J. Pharm. Anal. 2020, 10, 517–521. [Google Scholar] [CrossRef]
  7. Abbaszadegan, M.R.; Mojarrad, M.; Moghbeli, M. Role of extra cellular proteins in gastric cancer progression and metastasis: An update. Genes Environ. 2020, 42, 18. [Google Scholar] [CrossRef]
  8. Fredriksson, R.; Lagerstrom, M.C.; Lundin, L.G.; Schioth, H.B. The G-protein-coupled receptors in the human genome form five main families. Phylogenetic analysis, paralogon groups, and fingerprints. Mol. Pharmacol. 2003, 63, 1256–1272. [Google Scholar] [CrossRef] [Green Version]
  9. Erlandson, S.C.; McMahon, C.; Kruse, A.C. Structural Basis for G Protein-Coupled Receptor Signaling. Annu. Rev. Biophys. 2018, 47, 1–18. [Google Scholar] [CrossRef]
  10. Alexander, S.P.H.; Christopoulos, A.; Davenport, A.P.; Kelly, E.; Mathie, A.; Peters, J.A.; Veale, E.L.; Armstrong, J.F.; Faccenda, E.; Harding, S.D.; et al. THE CONCISE GUIDE TO PHARMACOLOGY 2019/20: G protein-coupled receptors. Br. J. Pharmacol. 2019, 176, S21–S141. [Google Scholar] [CrossRef] [Green Version]
  11. Zhang, Y.; Sun, B.F.; Feng, D.; Hu, H.L.; Chu, M.; Qu, Q.H.; Tarrasch, J.T.; Li, S.; Kobilka, T.S.; Kobilka, B.K.; et al. Cryo-EM structure of the activated GLP-1 receptor in complex with a G protein. Nature 2017, 546, 248–253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Liang, Y.L.; Khoshouei, M.; Radjainia, M.; Zhang, Y.; Glukhova, A.; Tarrasch, J.; Thal, D.M.; Furness, S.G.B.; Christopoulos, G.; Coudrat, T.; et al. Phase-plate cryo-EM structure of a class B GPCR-G-protein complex. Nature 2017, 546, 118–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Paek, J.; Kalocsay, M.; Staus, D.P.; Wingler, L.; Pascolutti, R.; Paulo, J.A.; Gygi, S.P.; Kruse, A.C. Multidimensional Tracking of GPCR Signaling via Peroxidase-Catalyzed Proximity Labeling. Cell 2017, 169, 338–349. [Google Scholar] [CrossRef] [Green Version]
  14. Flock, T.; Hauser, A.S.; Lund, N.; Gloriam, D.E.; Balaji, S.; Babu, M.M. Selectivity determinants of GPCR-G-protein binding. Nature 2017, 545, 317–322. [Google Scholar] [CrossRef] [Green Version]
  15. Gurevich, V.V.; Gurevich, E.V. GPCR Signaling Regulation: The Role of GRKs and Arrestins. Front. Pharmacol. 2019, 10, 125. [Google Scholar] [CrossRef] [Green Version]
  16. Gutkind, J.S.; Kostenis, E. Arrestins as rheostats of GPCR signalling. Nat. Rev. Mol. Cell Biol. 2018, 19, 615–616. [Google Scholar] [CrossRef] [PubMed]
  17. Smith, J.S.; Lefkowitz, R.J.; Rajagopal, S. Biased signalling: From simple switches to allosteric microprocessors. Nat. Rev. Drug Discov. 2018, 17, 243–260. [Google Scholar] [CrossRef]
  18. Nogues, L.; Palacios-Garcia, J.; Reglero, C.; Rivas, V.; Neves, M.; Ribas, C.; Penela, P.; Mayor, F., Jr. G protein-coupled receptor kinases (GRKs) in tumorigenesis and cancer progression: GPCR regulators and signaling hubs. Semin. Cancer Biol. 2018, 48, 78–90. [Google Scholar] [CrossRef] [PubMed]
  19. Campbell, A.P.; Smrcka, A.V. Targeting G protein-coupled receptor signalling by blocking G proteins. Nat. Rev. Drug Discov. 2018, 17, 789–803. [Google Scholar] [CrossRef]
  20. Dorsam, R.T.; Gutkind, J.S. G-protein-coupled receptors and cancer. Nat. Rev. Cancer 2007, 7, 79–94. [Google Scholar] [CrossRef] [PubMed]
  21. O’Hayre, M.; Degese, M.S.; Gutkind, J.S. Novel insights into G protein and G protein-coupled receptor signaling in cancer. Curr. Opin. Cell Biol. 2014, 27, 126–135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Grundmann, M.; Merten, N.; Malfacini, D.; Inoue, A.; Preis, P.; Simon, K.; Ruttiger, N.; Ziegler, N.; Benkel, T.; Schmitt, N.K.; et al. Lack of beta-arrestin signaling in the absence of active G proteins. Nat. Commun. 2018, 9, 341. [Google Scholar] [CrossRef] [Green Version]
  23. O’Hayre, M.; Eichel, K.; Avino, S.; Zhao, X.F.; Steffen, D.J.; Feng, X.D.; Kawakami, K.; Aoki, J.; Messer, K.; Sunahara, R.; et al. Genetic evidence that β-arrestins are dispensable for the initiation of β2-adrenergic receptor signaling to ERK. Sci. Signal. 2017, 10, eaal3395. [Google Scholar] [CrossRef] [Green Version]
  24. Luttrell, L.M.; Wang, J.; Plouffe, B.; Smith, J.S.; Yamani, L.; Kaur, S.; Jean-Charles, P.Y.; Gauthier, C.; Lee, M.H.; Pani, B.; et al. Manifold roles of β-arrestins in GPCR signaling elucidated with siRNA and CRISPR/Cas9. Sci. Signal. 2018, 11, eaat7650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Nusse, R.; Clevers, H. Wnt/β-Catenin Signaling, Disease, and Emerging Therapeutic Modalities. Cell 2017, 169, 985–999. [Google Scholar] [CrossRef] [Green Version]
  26. Lee, R.T.; Zhao, Z.; Ingham, P.W. Hedgehog signalling. Development 2016, 143, 367–372. [Google Scholar] [CrossRef] [Green Version]
  27. Bray, S.J. Notch signalling in context. Nat. Rev. Mol. Cell Biol. 2016, 17, 722–735. [Google Scholar] [CrossRef] [PubMed]
  28. May, L.T.; Leach, K.; Sexton, P.M.; Christopoulos, A. Allosteric modulation of G protein-coupled receptors. Annu. Rev. Pharmacol. Toxicol. 2007, 47, 1–51. [Google Scholar] [CrossRef]
  29. Foster, D.J.; Conn, P.J. Allosteric Modulation of GPCRs: New Insights and Potential Utility for Treatment of Schizophrenia and Other CNS Disorders. Neuron 2017, 94, 431–446. [Google Scholar] [CrossRef] [Green Version]
  30. Voss, T.; Li, J.; Cummings, J.; Farlow, M.; Assaid, C.; Froman, S.; Leibensperger, H.; Snow-Adami, L.; McMahon, K.B.; Egan, M.; et al. Randomized, controlled, proof-of-concept trial of MK-7622 in Alzheimer’s disease. Alzheimer’s Dement. 2018, 4, 173–181. [Google Scholar] [CrossRef]
  31. Uslaner, J.M.; Kuduk, S.D.; Wittmann, M.; Lange, H.S.; Fox, S.V.; Min, C.; Pajkovic, N.; Harris, D.; Cilissen, C.; Mahon, C.; et al. Preclinical to Human Translational Pharmacology of the Novel M1 Positive Allosteric Modulator MK-7622. J. Pharmacol. Exp. Ther. 2018, 365, 556–566. [Google Scholar] [CrossRef] [Green Version]
  32. Wootten, D.; Christopoulos, A.; Marti-Solano, M.; Babu, M.M.; Sexton, P.M. Mechanisms of signalling and biased agonism in G protein-coupled receptors. Nat. Rev. Mol. Cell Biol. 2018, 19, 638–653. [Google Scholar] [CrossRef]
  33. Kohout, T.A.; Nicholas, S.L.; Perry, S.J.; Reinhart, G.; Junger, S.; Struthers, R.S. Differential desensitization, receptor phosphorylation, β-arrestin recruitment, and ERK1/2 activation by the two endogenous ligands for the CC chemokine receptor 7. J. Biol. Chem. 2004, 279, 23214–23222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Soergel, D.G.; Subach, R.A.; Burnham, N.; Lark, M.W.; James, I.E.; Sadler, B.M.; Skobieranda, F.; Violin, J.D.; Webster, L.R. Biased agonism of the mu-opioid receptor by TRV130 increases analgesia and reduces on-target adverse effects versus morphine: A randomized, double-blind, placebo-controlled, crossover study in healthy volunteers. Pain 2014, 155, 1829–1835. [Google Scholar] [CrossRef] [PubMed]
  35. Singla, N.K.; Skobieranda, F.; Soergel, D.G.; Salamea, M.; Burt, D.A.; Demitrack, M.A.; Viscusi, E.R. APOLLO-2: A Randomized, Placebo and Active-Controlled Phase III Study Investigating Oliceridine (TRV130), a G Protein-Biased Ligand at the mu-Opioid Receptor, for Management of Moderate to Severe Acute Pain Following Abdominoplasty. Pain Pract. 2019, 19, 715–731. [Google Scholar] [CrossRef] [Green Version]
  36. James, J.R.; Oliveira, M.I.; Carmo, A.M.; Iaboni, A.; Davis, S.J. A rigorous experimental framework for detecting protein oligomerization using bioluminescence resonance energy transfer. Nat. Methods 2006, 3, 1001–1006. [Google Scholar] [CrossRef] [PubMed]
  37. Bouvier, M.; Heveker, N.; Jockers, R.; Marullo, S.; Milligan, G. BRET analysis of GPCR oligomerization: Newer does not mean better. Nat. Methods 2007, 4, 3–4. [Google Scholar] [CrossRef] [Green Version]
  38. Conn, P.M.; Rogers, D.C.; McNeil, R. Potency enhancement of a GnRH agonist: GnRH-receptor microaggregation stimulates gonadotropin release. Endocrinology 1982, 111, 335–337. [Google Scholar] [CrossRef]
  39. Conn, P.M.; Rogers, D.C.; Stewart, J.M.; Niedel, J.; Sheffield, T. Conversion of a gonadotropin-releasing hormone antagonist to an agonist. Nature 1982, 296, 653–655. [Google Scholar] [CrossRef]
  40. Borroto-Escuela, D.O.; Rodriguez, D.; Romero-Fernandez, W.; Kapla, J.; Jaiteh, M.; Ranganathan, A.; Lazarova, T.; Fuxe, K.; Carlsson, J. Mapping the Interface of a GPCR Dimer: A Structural Model of the A2A Adenosine and D2 Dopamine Receptor Heteromer. Front. Pharmacol. 2018, 9, 829. [Google Scholar] [CrossRef] [Green Version]
  41. Vischer, H.F.; Watts, A.O.; Nijmeijer, S.; Leurs, R. G protein-coupled receptors: Walking hand-in-hand, talking hand-in-hand? Br. J. Pharmacol. 2011, 163, 246–260. [Google Scholar] [CrossRef] [Green Version]
  42. Smith, N.J.; Milligan, G. Allostery at G protein-coupled receptor homo- and heteromers: Uncharted pharmacological landscapes. Pharmacol. Rev. 2010, 62, 701–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Hanyaloglu, A.C.; von Zastrow, M. Regulation of GPCRs by endocytic membrane trafficking and its potential implications. Annu. Rev. Pharmacol. Toxicol. 2008, 48, 537–568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Shenoy, S.K.; Lefkowitz, R.J. Seven-transmembrane receptor signaling through β-arrestin. Sci. Signal. 2005, 2005, cm10. [Google Scholar] [CrossRef]
  45. Reiter, E.; Ahn, S.; Shukla, A.K.; Lefkowitz, R.J. Molecular mechanism of β-arrestin-biased agonism at seven-transmembrane receptors. Annu. Rev. Pharmacol. Toxicol. 2012, 52, 179–197. [Google Scholar] [CrossRef] [Green Version]
  46. Rolland-Fourcade, C.; Denadai-Souza, A.; Cirillo, C.; Lopez, C.; Jaramillo, J.O.; Desormeaux, C.; Cenac, N.; Motta, J.P.; Larauche, M.; Tache, Y.; et al. Epithelial expression and function of trypsin-3 in irritable bowel syndrome. Gut 2017, 66, 1767–1778. [Google Scholar] [CrossRef] [PubMed]
  47. Du, L.; Long, Y.; Kim, J.J.; Chen, B.; Zhu, Y.; Dai, N. Protease Activated Receptor-2 Induces Immune Activation and Visceral Hypersensitivity in Post-infectious Irritable Bowel Syndrome Mice. Dig. Dis. Sci. 2019, 64, 729–739. [Google Scholar] [CrossRef]
  48. Jimenez-Vargas, N.N.; Pattison, L.A.; Zhao, P.; Lieu, T.; Latorre, R.; Jensen, D.D.; Castro, J.; Aurelio, L.; Le, G.T.; Flynn, B.; et al. Protease-activated receptor-2 in endosomes signals persistent pain of irritable bowel syndrome. Proc. Natl. Acad. Sci. USA 2018, 115, E7438–E7447. [Google Scholar] [CrossRef] [Green Version]
  49. Heng, B.C.; Aubel, D.; Fussenegger, M. An overview of the diverse roles of G-protein coupled receptors (GPCRs) in the pathophysiology of various human diseases. Biotechnol. Adv. 2013, 31, 1676–1694. [Google Scholar] [CrossRef]
  50. Spiegel, A.M.; Weinstein, L.S. Inherited diseases involving g proteins and g protein-coupled receptors. Annu. Rev. Med. 2004, 55, 27–39. [Google Scholar] [CrossRef] [Green Version]
  51. Raimondi, F.; Inoue, A.; Kadji, F.M.N.; Shuai, N.; Gonzalez, J.C.; Singh, G.; de la Vega, A.A.; Sotillo, R.; Fischer, B.; Aoki, J.; et al. Rare, functional, somatic variants in gene families linked to cancer genes: GPCR signaling as a paradigm. Oncogene 2019, 38, 6491–6506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Lawrence, M.S.; Stojanov, P.; Polak, P.; Kryukov, G.V.; Cibulskis, K.; Sivachenko, A.; Carter, S.L.; Stewart, C.; Mermel, C.H.; Roberts, S.A.; et al. Mutational heterogeneity in cancer and the search for new cancer-associated genes. Nature 2013, 499, 214–218. [Google Scholar] [CrossRef] [Green Version]
  53. Hauser, A.S.; Chavali, S.; Masuho, I.; Jahn, L.J.; Martemyanov, K.A.; Gloriam, D.E.; Babu, M.M. Pharmacogenomics of GPCR Drug Targets. Cell 2018, 172, 41–54. [Google Scholar] [CrossRef] [Green Version]
  54. Wu, V.; Yeerna, H.; Nohata, N.; Chiou, J.; Harismendy, O.; Raimondi, F.; Inoue, A.; Russell, R.B.; Tamayo, P.; Gutkind, J.S. Illuminating the Onco-GPCRome: Novel G protein-coupled receptor-driven oncocrine networks and targets for cancer immunotherapy. J. Biol. Chem. 2019, 294, 11062–11086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. O’Hayre, M.; Vazquez-Prado, J.; Kufareva, I.; Stawiski, E.W.; Handel, T.M.; Seshagiri, S.; Gutkind, J.S. The emerging mutational landscape of G proteins and G-protein-coupled receptors in cancer. Nat. Rev. Cancer 2013, 13, 412–424. [Google Scholar] [CrossRef] [Green Version]
  56. Kan, Z.Y.; Jaiswal, B.S.; Stinson, J.; Janakiraman, V.; Bhatt, D.; Stern, H.M.; Yue, P.; Haverty, P.M.; Bourgon, R.; Zheng, J.B.; et al. Diverse somatic mutation patterns and pathway alterations in human cancers. Nature 2010, 466, 869–873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ragnarsson, L.; Andersson, A.; Thomas, W.G.; Lewis, R.J. Mutations in the NPxxY motif stabilize pharmacologically distinct conformational states of the α1B- and β2-adrenoceptors. Sci. Signal. 2019, 12, eaas9485. [Google Scholar] [CrossRef]
  58. Stoy, H.; Gurevich, V.V. How genetic errors in GPCRs affect their function: Possible therapeutic strategies. Genes Dis. 2015, 2, 108–132. [Google Scholar] [CrossRef] [Green Version]
  59. Teh, J.L.; Chen, S. Glutamatergic signaling in cellular transformation. Pigment Cell Melanoma Res. 2012, 25, 331–342. [Google Scholar] [CrossRef]
  60. Vizurraga, A.; Adhikari, R.; Yeung, J.; Yu, M.; Tall, G.G. Mechanisms of adhesion G protein-coupled receptor activation. J. Biol. Chem. 2020, 295, 14065–14083. [Google Scholar] [CrossRef]
  61. Paavola, K.J.; Hall, R.A. Adhesion G protein-coupled receptors: Signaling, pharmacology, and mechanisms of activation. Mol. Pharmacol. 2012, 82, 777–783. [Google Scholar] [CrossRef] [Green Version]
  62. Prickett, T.D.; Wei, X.; Cardenas-Navia, I.; Teer, J.K.; Lin, J.C.; Walia, V.; Gartner, J.; Jiang, J.; Cherukuri, P.F.; Molinolo, A.; et al. Exon capture analysis of G protein-coupled receptors identifies activating mutations in GRM3 in melanoma. Nat. Genet. 2011, 43, 1119–1126. [Google Scholar] [CrossRef] [PubMed]
  63. Mermel, C.H.; Schumacher, S.E.; Hill, B.; Meyerson, M.L.; Beroukhim, R.; Getz, G. GISTIC2.0 facilitates sensitive and confident localization of the targets of focal somatic copy-number alteration in human cancers. Genome Biol. 2011, 12, R41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Sriram, K.; Moyung, K.; Corriden, R.; Carter, H.; Insel, P.A. GPCRs show widespread differential mRNA expression and frequent mutation and copy number variation in solid tumors. PLoS Biol. 2019, 17, e3000434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Canals, M.; Poole, D.P.; Veldhuis, N.A.; Schmidt, B.L.; Bunnett, N.W. G-Protein-Coupled Receptors Are Dynamic Regulators of Digestion and Targets for Digestive Diseases. Gastroenterology 2019, 156, 1600–1616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Gottesman-Katz, L.; Latorre, R.; Vanner, S.; Schmidt, B.L.; Bunnett, N.W. Targeting G protein-coupled receptors for the treatment of chronic pain in the digestive system. Gut 2020, 70, 970–981. [Google Scholar] [CrossRef] [PubMed]
  67. van Senten, J.R.; Fan, T.S.; Siderius, M.; Smit, M.J. Viral G protein-coupled receptors as modulators of cancer hallmarks. Pharmacol. Res. 2020, 156, 104804. [Google Scholar] [CrossRef]
  68. Knerr, J.M.; Kledal, T.N.; Rosenkilde, M.M. Molecular Properties and Therapeutic Targeting of the EBV-Encoded Receptor BILF1. Cancers 2021, 13, 4079. [Google Scholar] [CrossRef]
  69. Wood, L.D.; Parsons, D.W.; Jones, S.; Lin, J.; Sjoblom, T.; Leary, R.J.; Shen, D.; Boca, S.M.; Barber, T.; Ptak, J.; et al. The genomic landscapes of human breast and colorectal cancers. Science 2007, 318, 1108–1113. [Google Scholar] [CrossRef] [Green Version]
  70. Cassel, D.; Selinger, Z. Mechanism of Adenylate-Cyclase Activation by Cholera Toxin: Inhibition of GTP Hydrolysis at Regulatory Site. Proc. Natl. Acad. Sci. USA 1977, 74, 3307–3311. [Google Scholar] [CrossRef] [Green Version]
  71. Vandop, C.; Tsubokawa, M.; Bourne, H.R.; Ramachandran, J. Amino-Acid-Sequence of Retinal Transducin at the Site Adp-Ribosylated by Cholera-Toxin. J. Biol. Chem. 1984, 259, 696–698. [Google Scholar] [CrossRef]
  72. Masters, S.B.; Stroud, R.M.; Bourne, H.R. Family of G protein alpha chains: Amphipathic analysis and predicted structure of functional domains. Protein Eng. 1986, 1, 47–54. [Google Scholar] [CrossRef] [PubMed]
  73. Landis, C.A.; Masters, S.B.; Spada, A.; Pace, A.M.; Bourne, H.R.; Vallar, L. GTPase inhibiting mutations activate the α chain of Gs and stimulate adenylyl cyclase in human pituitary tumours. Nature 1989, 340, 692–696. [Google Scholar] [CrossRef] [PubMed]
  74. Hu, Q.; Shokat, K.M. Disease-Causing Mutations in the G Protein Gαs Subvert the Roles of GDP and GTP. Cell 2018, 173, 1254–1264. [Google Scholar] [CrossRef] [Green Version]
  75. Castellone, M.D.; Teramoto, H.; Williams, B.O.; Druey, K.M.; Gutkind, J.S. Prostaglandin E2 promotes colon cancer cell growth through a Gs-axin-β-catenin signaling axis. Science 2005, 310, 1504–1510. [Google Scholar] [CrossRef]
  76. Ikuta, K.; Seno, H.; Chiba, T. Molecular changes leading to gastric cancer: A suggestion from rare-type gastric tumors with GNAS mutations. Gastroenterology 2014, 146, 1417–1418. [Google Scholar] [CrossRef]
  77. Hollstein, P.E.; Shaw, R.J. GNAS shifts metabolism in pancreatic cancer. Nat Cell Biol 2018, 20, 740–741. [Google Scholar] [CrossRef]
  78. Patra, K.C.; Kato, Y.; Mizukami, Y.; Widholz, S.; Boukhali, M.; Revenco, I.; Grossman, E.A.; Ji, F.; Sadreyev, R.I.; Liss, A.S.; et al. Mutant GNAS drives pancreatic tumourigenesis by inducing PKA-mediated SIK suppression and reprogramming lipid metabolism. Nat. Cell Biol. 2018, 20, 811–822. [Google Scholar] [CrossRef]
  79. Wu, Y.; Peng, C.; Huang, L.; Xu, L.; Ding, X.; Liu, Y.; Zeng, C.; Sun, H.; Guo, W. Somatic GNAQ R183Q mutation is located within the sclera and episclera in patients with Sturge-Weber syndrome. Br. J. Ophthalmol. 2022, 106, 1006–1011. [Google Scholar] [CrossRef]
  80. Kusters-Vandevelde, H.V.; Klaasen, A.; Kusters, B.; Groenen, P.J.; van Engen-van Grunsven, I.A.; van Dijk, M.R.; Reifenberger, G.; Wesseling, P.; Blokx, W.A. Activating mutations of the GNAQ gene: A frequent event in primary melanocytic neoplasms of the central nervous system. Acta Neuropathol. 2010, 119, 317–323. [Google Scholar] [CrossRef] [Green Version]
  81. Van Raamsdonk, C.D.; Bezrookove, V.; Green, G.; Bauer, J.; Gaugler, L.; O’Brien, J.M.; Simpson, E.M.; Barsh, G.S.; Bastian, B.C. Frequent somatic mutations of GNAQ in uveal melanoma and blue naevi. Nature 2009, 457, 599–602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Van Raamsdonk, C.D.; Griewank, K.G.; Crosby, M.B.; Garrido, M.C.; Vemula, S.; Wiesner, T.; Obenauf, A.C.; Wackernagel, W.; Green, G.; Bouvier, N.; et al. Mutations in GNA11 in Uveal Melanoma. N. Engl. J. Med. 2010, 363, 2191–2199. [Google Scholar] [CrossRef] [Green Version]
  83. Shirley, M.D.; Tang, H.; Gallione, C.J.; Baugher, J.D.; Frelin, L.P.; Cohen, B.; North, P.E.; Marchuk, D.A.; Comi, A.M.; Pevsner, J. Sturge-Weber syndrome and port-wine stains caused by somatic mutation in GNAQ. N. Engl. J. Med. 2013, 368, 1971–1979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Hubbard, K.B.; Hepler, J.R. Cell signalling diversity of the Gqalpha family of heterotrimeric G proteins. Cell. Signal. 2006, 18, 135–150. [Google Scholar] [CrossRef] [PubMed]
  85. Feng, X.D.; Degese, M.S.; Iglesias-Bartolome, R.; Vaque, J.P.; Molinolo, A.A.; Rodrigues, M.; Zaidi, M.R.; Ksander, B.R.; Merlino, G.; Sodhi, A.; et al. Hippo-Independent Activation of YAP by the GNAQ Uveal Melanoma Oncogene through a Trio-Regulated Rho GTPase Signaling Circuitry. Cancer Cell 2014, 25, 831–845. [Google Scholar] [CrossRef] [Green Version]
  86. Zhang, J.X.; Yun, M.; Xu, Y.; Chen, J.W.; Weng, H.W.; Zheng, Z.S.; Chen, C.; Xie, D.; Ye, S. GNA13 as a prognostic factor and mediator of gastric cancer progression. Oncotarget 2016, 7, 4414–4427. [Google Scholar] [CrossRef] [Green Version]
  87. Liu, S.C.; Jen, Y.M.; Jiang, S.S.; Chang, J.L.; Hsiung, C.A.; Wang, C.H.; Juang, J.L. Gα12-mediated pathway promotes invasiveness of nasopharyngeal carcinoma by modulating actin cytoskeleton reorganization. Cancer Res. 2009, 69, 6122–6130. [Google Scholar] [CrossRef] [Green Version]
  88. Yagi, H.; Tan, W.; Dillenburg-Pilla, P.; Armando, S.; Amornphimoltham, P.; Simaan, M.; Weigert, R.; Molinolo, A.A.; Bouvier, M.; Gutkind, J.S. A synthetic biology approach reveals a CXCR4-G13-Rho signaling axis driving transendothelial migration of metastatic breast cancer cells. Sci. Signal. 2011, 4, ra60. [Google Scholar] [CrossRef] [Green Version]
  89. Rasheed, S.A.K.; Leong, H.S.; Lakshmanan, M.; Raju, A.; Dadlani, D.; Chong, F.T.; Shannon, N.B.; Rajarethinam, R.; Skanthakumar, T.; Tan, E.Y.; et al. GNA13 expression promotes drug resistance and tumor-initiating phenotypes in squamous cell cancers. Oncogene 2018, 37, 1340–1353. [Google Scholar] [CrossRef] [Green Version]
  90. Zhang, Z.; Tan, X.; Luo, J.; Cui, B.; Lei, S.; Si, Z.; Shen, L.; Yao, H. GNA13 promotes tumor growth and angiogenesis by upregulating CXC chemokines via the NF-kappaB signaling pathway in colorectal cancer cells. Cancer Med. 2018, 7, 5611–5620. [Google Scholar] [CrossRef]
  91. O’Hayre, M.; Inoue, A.; Kufareva, I.; Wang, Z.; Mikelis, C.M.; Drummond, R.A.; Avino, S.; Finkel, K.; Kalim, K.W.; DiPasquale, G.; et al. Inactivating mutations in GNA13 and RHOA in Burkitt’s lymphoma and diffuse large B-cell lymphoma: A tumor suppressor function for the Gα13/RhoA axis in B cells. Oncogene 2016, 35, 3771–3780. [Google Scholar] [CrossRef]
  92. Lohr, J.G.; Stojanov, P.; Lawrence, M.S.; Auclair, D.; Chapuy, B.; Sougnez, C.; Cruz-Gordillo, P.; Knoechel, B.; Asmann, Y.W.; Slager, S.L.; et al. Discovery and prioritization of somatic mutations in diffuse large B-cell lymphoma (DLBCL) by whole-exome sequencing. Proc. Natl. Acad. Sci. USA 2012, 109, 3879–3884. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Love, C.; Sun, Z.; Jima, D.; Li, G.; Zhang, J.; Miles, R.; Richards, K.L.; Dunphy, C.H.; Choi, W.W.; Srivastava, G.; et al. The genetic landscape of mutations in Burkitt lymphoma. Nat. Genet. 2012, 44, 1321–1325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Ford, C.E.; Skiba, N.P.; Bae, H.S.; Daaka, Y.H.; Reuveny, E.; Shekter, L.R.; Rosal, R.; Weng, G.Z.; Yang, C.S.; Iyengar, R.; et al. Molecular basis for interactions of G protein βγ subunits with effectors. Science 1998, 280, 1271–1274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Yoda, A.; Adelmant, G.; Tamburini, J.; Chapuy, B.; Shindoh, N.; Yoda, Y.; Weigert, O.; Kopp, N.; Wu, S.C.; Kim, S.S.; et al. Mutations in G protein β subunits promote transformation and kinase inhibitor resistance. Nat. Med. 2015, 21, 71–75. [Google Scholar] [CrossRef] [Green Version]
  96. Tan, N.B.; Pagnamenta, A.T.; Ferla, M.P.; Gadian, J.; Chung, B.H.; Chan, M.C.; Fung, J.L.; Cook, E.; Guter, S.; Boschann, F.; et al. Recurrent de novo missense variants in GNB2 can cause syndromic intellectual disability. J. Med. Genet. 2022, 59, 511–516. [Google Scholar] [CrossRef]
  97. Alvarez, C.J.P.; Lodeiro, M.; Theodoropoulou, M.; Camina, J.P.; Casanueva, F.F.; Pazos, Y. Obestatin stimulates Akt signalling in gastric cancer cells through beta-arrestin-mediated epidermal growth factor receptor transactivation. Endocr. Relat. Cancer 2009, 16, 599–611. [Google Scholar] [CrossRef] [Green Version]
  98. Zimmermannova, O.; Doktorova, E.; Stuchly, J.; Kanderova, V.; Kuzilkova, D.; Strnad, H.; Starkova, J.; Alberich-Jorda, M.; Falkenburg, J.H.F.; Trka, J.; et al. An activating mutation of GNB1 is associated with resistance to tyrosine kinase inhibitors in ETV6-ABL1-positive leukemia. Oncogene 2017, 36, 5985–5994. [Google Scholar] [CrossRef]
  99. Liu, Y.G.; Teng, Y.S.; Shan, Z.G.; Cheng, P.; Hao, C.J.; Lv, Y.P.; Mao, F.Y.; Yang, S.M.; Chen, W.; Zhao, Y.L.; et al. Arrestin domain containing 3 promotes Helicobacter pylori-associated gastritis by regulating protease-activated receptor 1. JCI Insight 2020, 5, e135849. [Google Scholar] [CrossRef]
  100. Zhang, C.; Gao, G.R.; Lv, C.G.; Zhang, B.L.; Zhang, Z.L.; Zhang, X.F. Protease-activated receptor-2 induces expression of vascular endothelial growth factor and cyclooxygenase-2 via the mitogen-activated protein kinase pathway in gastric cancer cells. Oncol. Rep. 2012, 28, 1917–1923. [Google Scholar] [CrossRef] [Green Version]
  101. Rocken, C.; Rohl, F.W.; Diebler, E.; Lendeckel, U.; Pross, M.; Carl-McGrath, S.; Ebert, M.P. The angiotensin II/angiotensin II receptor system correlates with nodal spread in intestinal type gastric cancer. Cancer Epidemiol. Biomark. Prev. 2007, 16, 1206–1212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Huang, W.; Wu, Y.L.; Zhong, J.; Jiang, F.X.; Tian, X.L.; Yu, L.F. Angiotensin II type 1 receptor antagonist suppress angiogenesis and growth of gastric cancer xenografts. Dig. Dis. Sci. 2008, 53, 1206–1210. [Google Scholar] [CrossRef]
  103. Shen, W.; Xi, H.; Li, C.; Bian, S.; Cheng, H.; Cui, J.; Wang, N.; Wei, B.; Huang, X.; Chen, L. Endothelin-A receptor in gastric cancer and enhanced antitumor activity of trastuzumab in combination with the endothelin-A receptor antagonist ZD4054. Ann. N. Y. Acad. Sci. 2019, 1448, 30–41. [Google Scholar] [CrossRef] [PubMed]
  104. Prevete, N.; Liotti, F.; Visciano, C.; Marone, G.; Melillo, R.M.; de Paulis, A. The formyl peptide receptor 1 exerts a tumor suppressor function in human gastric cancer by inhibiting angiogenesis. Oncogene 2015, 34, 3826–3838. [Google Scholar] [CrossRef]
  105. Hou, X.L.; Ji, C.D.; Tang, J.; Wang, Y.X.; Xiang, D.F.; Li, H.Q.; Liu, W.W.; Wang, J.X.; Yan, H.Z.; Wang, Y.; et al. FPR2 promotes invasion and metastasis of gastric cancer cells and predicts the prognosis of patients. Sci. Rep. 2017, 7, 3153. [Google Scholar] [CrossRef] [PubMed]
  106. Prevete, N.; Liotti, F.; Illiano, A.; Amoresano, A.; Pucci, P.; de Paulis, A.; Melillo, R.M. Formyl peptide receptor 1 suppresses gastric cancer angiogenesis and growth by exploiting inflammation resolution pathways. Oncoimmunology 2017, 6, e1293213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Ogasa, M.; Miyazaki, Y.; Hiraoka, S.; Kitamura, S.; Nagasawa, Y.; Kishida, O.; Miyazaki, T.; Kiyohara, T.; Shinomura, Y.; Matsuzawa, Y. Gastrin activates nuclear factor κB (NFκB) through a protein kinase C dependent pathway involving NFκB inducing kinase, inhibitor κB (IκB) kinase, and tumour necrosis factor receptor associated factor 6 (TRAF6) in MKN-28 cells transfected with gastrin receptor. Gut 2003, 52, 813–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. de Lau, W.; Barker, N.; Low, T.Y.; Koo, B.K.; Li, V.S.W.; Teunissen, H.; Kujala, P.; Haegebarth, A.; Peters, P.J.; van de Wetering, M.; et al. Lgr5 homologues associate with Wnt receptors and mediate R-spondin signalling. Nature 2011, 476, 293–297. [Google Scholar] [CrossRef]
  109. Ke, J.; Ma, P.; Chen, J.P.; Qin, J.; Qian, H.X. LGR6 promotes the progression of gastric cancer through PI3K/AKT/mTOR pathway. OncoTargets Ther. 2018, 11, 3025–3033. [Google Scholar] [CrossRef] [Green Version]
  110. Shida, D.; Kitayama, J.; Yamaguchi, H.; Yamashita, H.; Mori, K.; Watanabe, T.; Nagawa, H. Lysophospholipids transactivate HER2/neu (erbB-2) in human gastric cancer cells. BBRC 2005, 327, 907–914. [Google Scholar] [CrossRef]
  111. Shida, D.; Kitayama, J.; Yamaguchi, H.; Hama, K.; Aoki, J.; Arai, H.; Yamashita, H.; Mori, K.; Sako, A.; Konishi, T.; et al. Dual mode regulation of migration by lysophosphatidic acid in human gastric cancer cells. Exp. Cell Res. 2004, 301, 168–178. [Google Scholar] [CrossRef] [PubMed]
  112. Yang, D.Z.; Yang, W.H.; Zhang, Q.; Hu, Y.; Bao, L.; Damirin, A. Migration of gastric cancer cells in response to lysophosphatidic acid is mediated by LPA receptor 2. Oncol. Lett. 2013, 5, 1048–1052. [Google Scholar] [CrossRef] [Green Version]
  113. Shida, D.; Kitayama, J.; Yamaguchi, H.; Yamashita, H.; Mori, K.; Watanabe, T.; Yatomi, Y.; Nagawa, H. Sphingosine 1-phosphate transactivates c-Met as well as epidermal growth factor receptor (EGFR) in human gastric cancer cells. FEBS Lett. 2004, 577, 333–338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Yamashita, H.; Kitayama, J.; Shida, D.; Yamaguchi, H.; Mori, K.; Osada, M.; Aoki, S.; Yatomi, Y.; Takuwa, Y.; Nagawa, H. Sphingosine 1-phosphate receptor expression profile in human gastric cancer cells: Differential regulation on the migration and proliferation. J. Surg. Res. 2006, 130, 80–87. [Google Scholar] [CrossRef]
  115. Wong, C.C.; Kang, W.; Xu, J.; Qian, Y.; Luk, S.T.Y.; Chen, H.; Li, W.; Zhao, L.; Zhang, X.; Chiu, P.W.; et al. Prostaglandin E2 induces DNA hypermethylation in gastric cancer in vitro and in vivo. Theranostics 2019, 9, 6256–6268. [Google Scholar] [CrossRef] [PubMed]
  116. Oshima, H.; Oshima, M.; Inaba, K.; Taketo, M.M. Hyperplastic gastric tumors induced by activated macrophages in COX-2/mPGES-1 transgenic mice. EMBO J. 2004, 23, 1669–1678. [Google Scholar] [CrossRef] [Green Version]
  117. Lee, H.J.; Song, I.C.; Yun, H.J.; Jo, D.Y.; Kim, S. CXC chemokines and chemokine receptors in gastric cancer: From basic findings towards therapeutic targeting. World J. Gastroenterol. 2014, 20, 1681–1693. [Google Scholar] [CrossRef]
  118. Xu, M.; Wang, Y.; Xia, R.; Wei, Y.; Wei, X. Role of the CCL2-CCR2 signalling axis in cancer: Mechanisms and therapeutic targeting. Cell Prolif. 2021, 54, e13115. [Google Scholar] [CrossRef]
  119. Shah, N.; Khurana, S.; Cheng, K.R.; Raufman, J.P. Muscarinic receptors and ligands in cancer. Am. J. Physiol. Cell Physiol. 2009, 296, C221–C232. [Google Scholar] [CrossRef] [Green Version]
  120. Zhang, X.; Zhang, Y.; He, Z.; Yin, K.; Li, B.; Zhang, L.; Xu, Z. Chronic stress promotes gastric cancer progression and metastasis: An essential role for ADRB2. Cell Death Dis. 2019, 10, 788. [Google Scholar] [CrossRef] [Green Version]
  121. Bellefeuille, S.D.; Molle, C.M.; Gendron, F.P. Reviewing the role of P2Y receptors in specific gastrointestinal cancers. Purinergic Signal. 2019, 15, 451–463. [Google Scholar] [CrossRef] [PubMed]
  122. Tsuchiya, A.; Nishizaki, T. Anticancer effect of adenosine on gastric cancer via diverse signaling pathways. World J. Gastroenterol. 2015, 21, 10931–10935. [Google Scholar] [CrossRef] [PubMed]
  123. Shi, L.; Wu, Z.; Miao, J.; Du, S.; Ai, S.; Xu, E.; Feng, M.; Song, J.; Guan, W. Adenosine interaction with adenosine receptor A2a promotes gastric cancer metastasis by enhancing PI3K-AKT-mTOR signaling. Mol. Biol. Cell 2019, 30, 2527–2534. [Google Scholar] [CrossRef] [PubMed]
  124. Yasuda, H.; Hirata, S.; Inoue, K.; Mashima, H.; Ohnishi, H.; Yoshiba, M. Involvement of membrane-type bile acid receptor M-BAR/TGR5 in bile acid-induced activation of epidermal growth factor receptor and mitogen-activated protein kinases in gastric carcinoma cells. Biochem. Biophys. Res. Commun. 2007, 354, 154–159. [Google Scholar] [CrossRef]
  125. Wang, X.F.; Xu, Z.Y.; Sun, J.C.; Lv, H.; Wang, Y.P.; Ni, Y.X.; Chen, S.Q.; Hu, C.; Wang, L.J.; Chen, W.; et al. Cisplatin resistance in gastric cancer cells is involved with GPR30-mediated epithelial-mesenchymal transition. J. Cell. Mol. Med. 2020, 24, 3625–3633. [Google Scholar] [CrossRef]
  126. Gan, J.F.; Ke, X.R.; Jiang, J.L.; Dong, H.M.; Yao, Z.M.; Lin, Y.S.; Lin, W.; Wu, X.; Yan, S.M.; Zhuang, Y.X.; et al. Growth hormone-releasing hormone receptor antagonists inhibit human gastric cancer through downregulation of PAK1-STAT3/NF-κB signaling. Proc. Natl. Acad. Sci. USA 2016, 113, 14745–14750. [Google Scholar] [CrossRef] [Green Version]
  127. Xie, R.; Xu, J.Y.; Xiao, Y.F.; Wu, J.L.; Wan, H.X.; Tang, B.; Liu, J.J.; Fan, Y.H.; Wang, S.M.; Wu, Y.Y.; et al. Calcium Promotes Human Gastric Cancer via a Novel Coupling of Calcium-Sensing Receptor and TRPV4 Channel. Cancer Res. 2017, 77, 6499–6512. [Google Scholar] [CrossRef] [Green Version]
  128. Wang, K.; Zhao, X.H.; Liu, J.; Zhang, R.; Li, J.P. Nervous system and gastric cancer. BBA-Rev. Cancer 2020, 1873, 188313. [Google Scholar] [CrossRef]
  129. Ferrigno, A.; Berardo, C.; Di Pasqua, L.G.; Siciliano, V.; Richelmi, P.; Vairetti, M. Localization and role of metabotropic glutamate receptors subtype 5 in the gastrointestinal tract. World J. Gastroenterol. 2017, 23, 4500–4507. [Google Scholar] [CrossRef]
  130. Gad, A.A.; Balenga, N. The Emerging Role of Adhesion GPCRs in Cancer. ACS Pharmacol. Transl. Sci. 2020, 3, 29–42. [Google Scholar] [CrossRef]
  131. Kozielewicz, P.; Turku, A.; Schulte, G. Molecular Pharmacology of Class F Receptor Activation. Mol. Pharmacol. 2020, 97, 62–71. [Google Scholar] [CrossRef] [PubMed]
  132. Flanagan, D.J.; Barker, N.; Di Costanzo, N.S.; Mason, E.A.; Gurney, A.; Meniel, V.S.; Koushyar, S.; Austin, C.R.; Ernst, M.; Pearson, H.B.; et al. Frizzled-7 Is Required for Wnt Signaling in Gastric Tumors with and Without Apc Mutations. Cancer Res. 2019, 79, 970–981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Xin, L.; Liu, L.; Liu, C.; Zhou, L.Q.; Zhou, Q.; Yuan, Y.W.; Li, S.H.; Zhang, H.T. DNA-methylation-mediated silencing of miR-7-5p promotes gastric cancer stem cell invasion via increasing Smo and Hes1. J. Cell. Physiol. 2020, 235, 2643–2654. [Google Scholar] [CrossRef]
  134. Fares, S.; Spiess, K.; Olesen, E.T.B.; Zuo, J.M.; Jackson, S.; Kledal, T.N.; Wills, M.R.; Rosenkilde, M.M. Distinct Roles of Extracellular Domains in the Epstein-Barr Virus-Encoded BILF1 Receptor for Signaling and Major Histocompatibility Complex Class I Downregulation. mBio 2019, 10, e01707-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Perez Almeria, C.V.; Setiawan, I.M.; Siderius, M.; Smit, M.J. G protein-coupled receptors as promising targets in cancer. Curr. Opin. Endocr. Metab. Res. 2021, 16, 119–127. [Google Scholar] [CrossRef]
  136. Warren, C.M.; Landgraf, R. Signaling through ERBB receptors: Multiple layers of diversity and control. Cell. Signal. 2006, 18, 923–933. [Google Scholar] [CrossRef] [PubMed]
  137. Wetzker, R.; Bohmer, F.D. Transactivation joins multiple tracks to the ERK/MAPK cascade. Nat. Rev. Mol. Cell Biol. 2003, 4, 651–657. [Google Scholar] [CrossRef]
  138. Kose, M. GPCRs and EGFR—Cross-talk of membrane receptors in cancer. Bioorg. Med. Chem. Lett. 2017, 27, 3611–3620. [Google Scholar] [CrossRef]
  139. Kilpatrick, L.E.; Hill, S.J. Transactivation of G protein-coupled receptors (GPCRs) and receptor tyrosine kinases (RTKs): Recent insights using luminescence and fluorescence technologies. Curr. Opin. Endocr. Metab. Res. 2021, 16, 102–112. [Google Scholar] [CrossRef]
  140. Joh, T.; Kataoka, H.; Tanida, S.; Watanabe, K.; Ohshima, T.; Sasaki, M.; Nakao, H.; Ohhara, H.; Higashiyama, S.; Itoh, M. Helicobacter pylori-stimulated interleukin-8 (IL-8) promotes cell proliferation through transactivation of epidermal growth factor receptor (EGFR) by disintegrin and metalloproteinase (ADAM) activation. Dig. Dis. Sci. 2005, 50, 2081–2089. [Google Scholar] [CrossRef]
  141. Oshima, H.; Popivanova, B.K.; Oguma, K.; Kong, D.; Ishikawa, T.O.; Oshima, M. Activation of epidermal growth factor receptor signaling by the prostaglandin E2 receptor EP4 pathway during gastric tumorigenesis. Cancer Sci. 2011, 102, 713–719. [Google Scholar] [CrossRef] [PubMed]
  142. Lian, S.; Xia, Y.; Ung, T.T.; Khoi, P.N.; Yoon, H.J.; Lee, S.G.; Kim, K.K.; Jung, Y.D. Prostaglandin E2 stimulates urokinase-type plasminogen activator receptor via EP2 receptor-dependent signaling pathways in human AGS gastric cancer cells. Mol. Carcinog. 2017, 56, 664–680. [Google Scholar] [CrossRef] [PubMed]
  143. Ancha, H.R.; Kurella, R.R.; Stewart, C.A.; Damera, G.; Ceresa, B.P.; Harty, R.F. Histamine stimulation of MMP-1(collagenase-1) secretion and gene expression in gastric epithelial cells: Role of EGFR transactivation and the MAP kinase pathway. Int. J. Biochem. Cell Biol. 2007, 39, 2143–2152. [Google Scholar] [CrossRef] [PubMed]
  144. Wang, Q.; Yang, H.; Zhuo, Q.; Xu, Y.; Zhang, P. Knockdown of EPCR inhibits the proliferation and migration of human gastric cancer cells via the ERK1/2 pathway in a PAR-1-dependent manner. Oncol. Rep. 2018, 39, 1843–1852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Manning, B.D.; Toker, A. AKT/PKB Signaling: Navigating the Network. Cell 2017, 169, 381–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Matsuoka, T.; Yashiro, M. The Role of PI3K/Akt/mTOR Signaling in Gastric Carcinoma. Cancers 2014, 6, 1441–1463. [Google Scholar] [CrossRef] [Green Version]
  147. Li, F.L.; Guan, K.L. The two sides of Hippo pathway in cancer. Semin. Cancer Biol. 2021, 85, 33–42. [Google Scholar] [CrossRef]
  148. Zhao, B.; Ye, X.; Yu, J.; Li, L.; Li, W.; Li, S.; Yu, J.; Lin, J.D.; Wang, C.Y.; Chinnaiyan, A.M.; et al. TEAD mediates YAP-dependent gene induction and growth control. Genes Dev. 2008, 22, 1962–1971. [Google Scholar] [CrossRef] [Green Version]
  149. Yu, F.X.; Zhao, B.; Panupinthu, N.; Jewell, J.L.; Lian, I.; Wang, L.H.; Zhao, J.; Yuan, H.; Tumaneng, K.; Li, H.; et al. Regulation of the Hippo-YAP pathway by G-protein-coupled receptor signaling. Cell 2012, 150, 780–791. [Google Scholar] [CrossRef] [Green Version]
  150. Luo, J.; Yu, F.X. GPCR-Hippo Signaling in Cancer. Cells 2019, 8, 426. [Google Scholar] [CrossRef] [Green Version]
  151. Cai, H.; Xu, Y. The role of LPA and YAP signaling in long-term migration of human ovarian cancer cells. Cell Commun. Signal. 2013, 11, 31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Yung, Y.C.; Stoddard, N.C.; Chun, J. LPA receptor signaling: Pharmacology, physiology, and pathophysiology. J. Lipid Res. 2014, 55, 1192–1214. [Google Scholar] [CrossRef] [Green Version]
  153. Cheng, J.C.; Wang, E.Y.; Yi, Y.; Thakur, A.; Tsai, S.H.; Hoodless, P.A. S1P Stimulates Proliferation by Upregulating CTGF Expression through S1PR2-Mediated YAP Activation. Mol. Cancer Res. 2018, 16, 1543–1555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Saikawa, S.; Kaji, K.; Nishimura, N.; Seki, K.; Sato, S.; Nakanishi, K.; Kitagawa, K.; Kawaratani, H.; Kitade, M.; Moriya, K.; et al. Angiotensin receptor blockade attenuates cholangiocarcinoma cell growth by inhibiting the oncogenic activity of Yes-associated protein. Cancer Lett. 2018, 434, 120–129. [Google Scholar] [CrossRef] [PubMed]
  155. Mo, J.S.; Yu, F.X.; Gong, R.; Brown, J.H.; Guan, K.L. Regulation of the Hippo-YAP pathway by protease-activated receptors (PARs). Genes Dev. 2012, 26, 2138–2143. [Google Scholar] [CrossRef] [Green Version]
  156. Anakk, S.; Bhosale, M.; Schmidt, V.A.; Johnson, R.L.; Finegold, M.J.; Moore, D.D. Bile acids activate YAP to promote liver carcinogenesis. Cell Rep. 2013, 5, 1060–1069. [Google Scholar] [CrossRef] [Green Version]
  157. Thirunavukkarasan, M.; Wang, C.; Rao, A.; Hind, T.; Teo, Y.R.; Siddiquee, A.A.; Goghari, M.A.I.; Kumar, A.P.; Herr, D.R. Short-chain fatty acid receptors inhibit invasive phenotypes in breast cancer cells. PLoS ONE 2017, 12, e0186334. [Google Scholar] [CrossRef] [PubMed]
  158. Zhou, X.; Wang, S.; Wang, Z.; Feng, X.; Liu, P.; Lv, X.B.; Li, F.; Yu, F.X.; Sun, Y.; Yuan, H.; et al. Estrogen regulates Hippo signaling via GPER in breast cancer. J. Clin. Investig. 2015, 125, 2123–2135. [Google Scholar] [CrossRef]
  159. Wang, Z.; Liu, P.; Zhou, X.; Wang, T.; Feng, X.; Sun, Y.P.; Xiong, Y.; Yuan, H.X.; Guan, K.L. Endothelin Promotes Colorectal Tumorigenesis by Activating YAP/TAZ. Cancer Res. 2017, 77, 2413–2423. [Google Scholar] [CrossRef] [Green Version]
  160. Roulis, M.; Kaklamanos, A.; Schernthanner, M.; Bielecki, P.; Zhao, J.; Kaffe, E.; Frommelt, L.S.; Qu, R.; Knapp, M.S.; Henriques, A.; et al. Paracrine orchestration of intestinal tumorigenesis by a mesenchymal niche. Nature 2020, 580, 524–529. [Google Scholar] [CrossRef]
  161. Fujimoto, D.; Ueda, Y.; Hirono, Y.; Goi, T.; Yamaguchi, A. PAR1 participates in the ability of multidrug resistance and tumorigenesis by controlling Hippo-YAP pathway. Oncotarget 2015, 6, 34788–34799. [Google Scholar] [CrossRef]
  162. Katoh, M.; Katoh, M. Precision medicine for human cancers with Notch signaling dysregulation (Review). Int. J. Mol. Med. 2020, 45, 279–297. [Google Scholar] [CrossRef] [Green Version]
  163. Merchant, J.L. Hedgehog signalling in gut development, physiology and cancer. J. Physiol. 2012, 590, 421–432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Chiurillo, M.A. Role of the Wnt/β-catenin pathway in gastric cancer: An in-depth literature review. World J. Exp. Med. 2015, 5, 84–102. [Google Scholar] [CrossRef] [PubMed]
  165. Liang, Y.; Yang, L.; Xie, J. The Role of the Hedgehog Pathway in Chemoresistance of Gastrointestinal Cancers. Cells 2021, 10, 2030. [Google Scholar] [CrossRef]
  166. Fukaya, M.; Isohata, N.; Ohta, H.; Aoyagi, K.; Ochiya, T.; Saeki, N.; Yanagihara, K.; Nakanishi, Y.; Taniguchi, H.; Sakamoto, H.; et al. Hedgehog signal activation in gastric pit cell and in diffuse-type gastric cancer. Gastroenterology 2006, 131, 14–29. [Google Scholar] [CrossRef] [PubMed]
  167. Mukhopadhyay, S.; Rohatgi, R. G-protein-coupled receptors, Hedgehog signaling and primary cilia. Semin. Cell Dev. Biol. 2014, 33, 63–72. [Google Scholar] [CrossRef] [Green Version]
  168. Koushyar, S.; Powell, A.G.; Vincan, E.; Phesse, T.J. Targeting Wnt Signaling for the Treatment of Gastric Cancer. Int. J. Mol. Sci. 2020, 21, 3927. [Google Scholar] [CrossRef]
  169. Valastyan, S.; Weinberg, R.A. Tumor metastasis: Molecular insights and evolving paradigms. Cell 2011, 147, 275–292. [Google Scholar] [CrossRef] [Green Version]
  170. Li, W.; Ng, J.M.; Wong, C.C.; Ng, E.K.W.; Yu, J. Molecular alterations of cancer cell and tumour microenvironment in metastatic gastric cancer. Oncogene 2018, 37, 4903–4920. [Google Scholar] [CrossRef]
  171. Gupta, G.P.; Massague, J. Cancer metastasis: Building a framework. Cell 2006, 127, 679–695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Chaffer, C.L.; Weinberg, R.A. A perspective on cancer cell metastasis. Science 2011, 331, 1559–1564. [Google Scholar] [CrossRef]
  173. Brabletz, S.; Schuhwerk, H.; Brabletz, T.; Stemmler, M.P. Dynamic EMT: A multi-tool for tumor progression. EMBO J. 2021, 40, e108647. [Google Scholar] [CrossRef] [PubMed]
  174. Lombardi, L.; Tavano, F.; Morelli, F.; Latiano, T.P.; Di Sebastiano, P.; Maiello, E. Chemokine receptor CXCR4: Role in gastrointestinal cancer. Crit. Rev. Oncol. Hematol. 2013, 88, 696–705. [Google Scholar] [CrossRef]
  175. Marcuzzi, E.; Angioni, R.; Molon, B.; Cali, B. Chemokines and Chemokine Receptors: Orchestrating Tumor Metastasization. Int. J. Mol. Sci. 2018, 20, 96. [Google Scholar] [CrossRef] [Green Version]
  176. Nagarsheth, N.; Wicha, M.S.; Zou, W.P. Chemokines in the cancer microenvironment and their relevance in cancer immunotherapy. Nat. Rev. Immunol. 2017, 17, 559–572. [Google Scholar] [CrossRef] [Green Version]
  177. Ashburner, B.P.; Westerheide, S.D.; Baldwin, A.S., Jr. The p65 (RelA) subunit of NF-kappaB interacts with the histone deacetylase (HDAC) corepressors HDAC1 and HDAC2 to negatively regulate gene expression. Mol. Cell. Biol. 2001, 21, 7065–7077. [Google Scholar] [CrossRef] [Green Version]
  178. Tarnowski, M.; Grymula, K.; Reca, R.; Jankowski, K.; Maksym, R.; Tarnowska, J.; Przybylski, G.; Barr, F.G.; Kucia, M.; Ratajczak, M.Z. Regulation of expression of stromal-derived factor-1 receptors: CXCR4 and CXCR7 in human rhabdomyosarcomas. Mol. Cancer Res. 2010, 8, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Schioppa, T.; Uranchimeg, B.; Saccani, A.; Biswas, S.K.; Doni, A.; Rapisarda, A.; Bernasconi, S.; Saccani, S.; Nebuloni, M.; Vago, L.; et al. Regulation of the chemokine receptor CXCR4 by hypoxia. J. Exp. Med. 2003, 198, 1391–1402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Hitchon, C.; Wong, K.; Ma, G.; Reed, J.; Lyttle, D.; El-Gabalawy, H. Hypoxia-induced production of stromal cell-derived factor 1 (CXCL12) and vascular endothelial growth factor by synovial fibroblasts. Arthritis Rheum. 2002, 46, 2587–2597. [Google Scholar] [CrossRef]
  181. Lee, K.E.; Khoi, P.N.; Xia, Y.; Park, J.S.; Joo, Y.E.; Kim, K.K.; Choi, S.Y.; Jung, Y.D. Helicobacter pylori and interleukin-8 in gastric cancer. World J. Gastroenterol. 2013, 19, 8192–8202. [Google Scholar] [CrossRef] [PubMed]
  182. Haghazali, M.; Molaei, M.; Mashayekhi, R.; Zojaji, H.; Pourhoseingholi, M.A.; Shooshtarizadeh, T.; Mirsattari, D.; Zali, M.R. Proinflammatory cytokines and thrombomodulin in patients with peptic ulcer disease and gastric cancer, infected with Helicobacter pylori. Indian J. Pathol. Microbiol. 2011, 54, 103–106. [Google Scholar] [CrossRef] [PubMed]
  183. Jafarzadeh, A.; Nemati, M.; Jafarzadeh, S. The important role played by chemokines influence the clinical outcome of Helicobacter pylori infection. Life Sci. 2019, 231, 116688. [Google Scholar] [CrossRef]
  184. Pawluczuk, E.; Lukaszewicz-Zajac, M.; Mroczko, B. The Role of Chemokines in the Development of Gastric Cancer—Diagnostic and Therapeutic Implications. Int. J. Mol. Sci. 2020, 21, 8456. [Google Scholar] [CrossRef]
  185. Baj-Krzyworzeka, M.; Weglarczyk, K.; Baran, J.; Szczepanik, A.; Szura, M.; Siedlar, M. Elevated level of some chemokines in plasma of gastric cancer patients. Cent. Eur. J. Immunol. 2016, 41, 358–362. [Google Scholar] [CrossRef] [Green Version]
  186. Mashino, K.; Sadanaga, N.; Yamaguchi, H.; Tanaka, F.; Ohta, M.; Shibuta, K.; Inoue, H.; Mori, M. Expression of chemokine receptor CCR7 is associated with lymph node metastasis of gastric carcinoma. Cancer Res. 2002, 62, 2937–2941. [Google Scholar] [PubMed]
  187. Ying, J.E.; Xu, Q.; Zhang, G.; Liu, B.X.; Zhu, L.M. The expression of CXCL12 and CXCR4 in gastric cancer and their correlation to lymph node metastasis. Med. Oncol. 2012, 29, 1716–1722. [Google Scholar] [CrossRef] [PubMed]
  188. Yasumoto, K.; Koizumi, K.; Kawashima, A.; Saitoh, Y.; Arita, Y.; Shinohara, K.; Minami, T.; Nakayama, T.; Sakurai, H.; Takahashi, Y.; et al. Role of the CXCL12/CXCR4 axis in peritoneal carcinomatosis of gastric cancer. Cancer Res. 2006, 66, 2181–2187. [Google Scholar] [CrossRef] [Green Version]
  189. Xiang, Z.; Zhou, Z.J.; Xia, G.K.; Zhang, X.H.; Wei, Z.W.; Zhu, J.T.; Yu, J.; Chen, W.; He, Y.; Schwarz, R.E.; et al. A positive crosstalk between CXCR4 and CXCR2 promotes gastric cancer metastasis. Oncogene 2017, 36, 5122–5133. [Google Scholar] [CrossRef]
  190. Cheng, Y.; Qu, J.L.; Che, X.F.; Xu, L.; Song, N.; Ma, Y.J.; Gong, J.; Qu, X.J.; Liu, Y.P. CXCL12/SDF-1 alpha induces migration via SRC-mediated CXCR4-EGFR cross-talk in gastric cancer cells. Oncol. Lett. 2017, 14, 2103–2110. [Google Scholar] [CrossRef] [Green Version]
  191. Chen, G.; Chen, S.M.; Wang, X.; Ding, X.F.; Ding, J.; Meng, L.H. Inhibition of chemokine (CXC motif) ligand 12/chemokine (CXC motif) receptor 4 axis (CXCL12/CXCR4)-mediated cell migration by targeting mammalian target of rapamycin (mTOR) pathway in human gastric carcinoma cells. J. Biol. Chem. 2012, 287, 12132–12141. [Google Scholar] [CrossRef] [Green Version]
  192. Ma, H.Y.; Gao, L.L.; Li, S.C.; Qin, J.; Chen, L.; Liu, X.Z.; Xu, P.P.; Wang, F.; Xiao, H.L.; Zhou, S.; et al. CCR7 enhances TGF-β 1-induced epithelial-mesenchymal transition and is associated with lymph node metastasis and poor overall survival in gastric cancer. Oncotarget 2015, 6, 24348–24360. [Google Scholar] [CrossRef] [Green Version]
  193. Zlotnik, A.; Burkhardt, A.M.; Homey, B. Homeostatic chemokine receptors and organ-specific metastasis. Nat. Rev. Immunol. 2011, 11, 597–606. [Google Scholar] [CrossRef] [PubMed]
  194. Amersi, F.F.; Terando, A.M.; Goto, Y.; Scolyer, R.A.; Thompson, J.F.; Tran, A.N.; Faries, M.B.; Morton, D.L.; Hoon, D.S.B. Activation of CCR9/CCL25 in cutaneous melanoma mediates preferential metastasis to the small intestine. Clin. Cancer. Res. 2008, 14, 638–645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Johnson-Holiday, C.; Singh, R.; Johnson, E.; Singh, S.; Stockard, C.R.; Grizzle, W.E.; Lillard, J.W. CCL25 mediates migration, invasion and matrix metalloproteinase expression by breast cancer cells in a CCR9-dependent fashion. Int. J. Oncol. 2011, 38, 1279–1285. [Google Scholar]
  196. Singh, R.; Stockard, C.R.; Grizzle, W.E.; Lillard, J.W.; Singh, S. Expression and histopathological correlation of CCR9 and CCL25 in ovarian cancer. Int. J. Oncol. 2011, 39, 373–381. [Google Scholar] [CrossRef]
  197. Xu, E.; Xia, X.; Jiang, C.; Li, Z.; Yang, Z.; Zheng, C.; Wang, X.; Du, S.; Miao, J.; Wang, F.; et al. GPER1 Silencing Suppresses the Proliferation, Migration, and Invasion of Gastric Cancer Cells by Inhibiting PI3K/AKT-Mediated EMT. Front. Cell Dev. Biol. 2020, 8, 591239. [Google Scholar] [CrossRef] [PubMed]
  198. Kessenbrock, K.; Plaks, V.; Werb, Z. Matrix metalloproteinases: Regulators of the tumor microenvironment. Cell 2010, 141, 52–67. [Google Scholar] [CrossRef] [Green Version]
  199. Jackson, H.W.; Defamie, V.; Waterhouse, P.; Khokha, R. TIMPs: Versatile extracellular regulators in cancer. Nat. Rev. Cancer 2017, 17, 38–53. [Google Scholar] [CrossRef]
  200. Abdi, E.; Latifi-Navid, S.; Sarvestani, F.A.; Esmailnejad, M.H. Emerging therapeutic targets for gastric cancer from a host-Helicobacter pylori interaction perspective. Expert Opin. Ther. Targets 2021, 25, 685–699. [Google Scholar] [CrossRef]
  201. Posselt, G.; Crabtree, J.E.; Wessler, S. Proteolysis in Helicobacter pylori-Induced Gastric Cancer. Toxins 2017, 9, 134. [Google Scholar] [CrossRef]
  202. Colosimo, D.A.; Kohn, J.A.; Luo, P.M.; Piscotta, F.J.; Han, S.M.; Pickard, A.J.; Rao, A.; Cross, J.R.; Cohen, L.J.; Brady, S.F. Mapping Interactions of Microbial Metabolites with Human G-Protein-Coupled Receptors. Cell Host Microbe 2019, 26, 273–282. [Google Scholar] [CrossRef] [Green Version]
  203. Ziyad, S.; Iruela-Arispe, M.L. Molecular mechanisms of tumor angiogenesis. Genes Cancer 2011, 2, 1085–1096. [Google Scholar] [CrossRef] [Green Version]
  204. Fidler, I.J. Timeline—The pathogenesis of cancer metastasis: The ‘seed and soil’ hypothesis revisited. Nat. Rev. Cancer 2003, 3, 453–458. [Google Scholar] [CrossRef]
  205. Nierodzik, M.L.; Karpatkin, S. Thrombin induces tumor growth, metastasis, and angiogenesis: Evidence for a thrombin-regulated dormant tumor phenotype. Cancer Cell 2006, 10, 355–362. [Google Scholar] [CrossRef] [Green Version]
  206. Wojtukiewicz, M.Z.; Hempel, D.; Sierko, E.; Tucker, S.C.; Honn, K.V. Protease-activated receptors (PARs)—Biology and role in cancer invasion and metastasis. Cancer Metastasis Rev. 2015, 34, 775–796. [Google Scholar] [CrossRef] [Green Version]
  207. Liu, Y.J.; Wada, R.; Yamashita, T.; Mi, Y.D.; Deng, C.X.; Hobson, J.P.; Rosenfeldt, H.M.; Nava, V.E.; Chae, S.S.; Lee, M.J.; et al. Edg-1, the G protein-coupled receptor for sphingosine-1-phosphate, is essential for vascular maturation. J. Clin. Investig. 2000, 106, 951–961. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Cartier, A.; Leigh, T.; Liu, C.H.; Hla, T. Endothelial sphingosine 1-phosphate receptors promote vascular normalization and antitumor therapy. Proc. Natl. Acad. Sci. USA 2020, 117, 3157–3166. [Google Scholar] [CrossRef] [Green Version]
  209. Yasuda, D.; Kobayashi, D.; Akahoshi, N.; Ohto-Nakanishi, T.; Yoshioka, K.; Takuwa, Y.; Mizuno, S.; Takahashi, S.; Ishii, S. Lysophosphatidic acid-induced YAP/TAZ activation promotes developmental angiogenesis by repressing Notch ligand Dll4. J. Clin. Investig. 2019, 129, 4332–4349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Liu, N.; Wu, Q.; Wang, Y.; Sui, H.; Liu, X.; Zhou, N.; Zhou, L.; Wang, Y.; Ye, N.; Fu, X.; et al. Helicobacter pylori promotes VEGF expression via the p38 MAPKmediated COX2PGE2 pathway in MKN45 cells. Mol. Med. Rep. 2014, 10, 2123–2129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. De Francesco, E.M.; Sotgia, F.; Clarke, R.B.; Lisanti, M.P.; Maggiolini, M. G Protein-Coupled Receptors at the Crossroad between Physiologic and Pathologic Angiogenesis: Old Paradigms and Emerging Concepts. Int. J. Mol. Sci. 2017, 18, 2713. [Google Scholar] [CrossRef] [Green Version]
  212. Zeng, D.Q.; Li, M.Y.; Zhou, R.; Zhang, J.W.; Sun, H.Y.; Shi, M.; Bin, J.P.; Liao, Y.L.; Rao, J.J.; Liao, W.J. Tumor Microenvironment Characterization in Gastric Cancer Identifies Prognostic and Immunotherapeutically Relevant Gene Signatures. Cancer Immunol. Res. 2019, 7, 737–750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Rojas, A.; Araya, P.; Gonzalez, I.; Morales, E. Gastric Tumor Microenvironment. Adv. Exp. Med. Biol. 2020, 1226, 23–35. [Google Scholar] [CrossRef]
  214. Quail, D.F.; Joyce, J.A. Microenvironmental regulation of tumor progression and metastasis. Nat. Med. 2013, 19, 1423–1437. [Google Scholar] [CrossRef]
  215. Chiu, K.J.; Chiou, H.C.; Huang, C.H.; Lu, P.C.; Kuo, H.R.; Wang, J.W.; Lin, M.H. Natural Compounds Targeting Cancer-Associated Fibroblasts against Digestive System Tumor Progression: Therapeutic Insights. Biomedicines 2022, 10, 713. [Google Scholar] [CrossRef] [PubMed]
  216. Peltier, A.; Seban, R.D.; Buvat, I.; Bidard, F.C.; Mechta-Grigoriou, F. Fibroblast heterogeneity in solid tumors: From single cell analysis to whole-body imaging. Semin. Cancer Biol. 2022, 86, 262–272. [Google Scholar] [CrossRef]
  217. Gorbachev, A.V.; Fairchild, R.L. Regulation of chemokine expression in the tumor microenvironment. Crit. Rev. Immunol. 2014, 34, 103–120. [Google Scholar] [CrossRef] [PubMed]
  218. Rihawi, K.; Ricci, A.D.; Rizzo, A.; Brocchi, S.; Marasco, G.; Pastore, L.V.; Llimpe, F.L.R.; Golfieri, R.; Renzulli, M. Tumor-Associated Macrophages and Inflammatory Microenvironment in Gastric Cancer: Novel Translational Implications. Int. J. Mol. Sci. 2021, 22, 3805. [Google Scholar] [CrossRef]
  219. Chen, F.F.; Yin, S.; Niu, L.; Luo, J.; Wang, B.C.; Xu, Z.G.; Yang, G.F. Expression of the Chemokine Receptor CXCR3 Correlates with Dendritic Cell Recruitment and Prognosis in Gastric Cancer. Genet. Test. Mol. Biomark. 2018, 22, 35–42. [Google Scholar] [CrossRef]
  220. Sebrell, T.A.; Hashimi, M.; Sidar, B.; Wilkinson, R.A.; Kirpotina, L.; Quinn, M.T.; Malkoc, Z.; Taylor, P.J.; Wilking, J.N.; Bimczok, D. A Novel Gastric Spheroid Co-culture Model Reveals Chemokine-Dependent Recruitment of Human Dendritic Cells to the Gastric Epithelium. Cell. Mol. Gastroenterol. Hepatol. 2019, 8, 157–171. [Google Scholar] [CrossRef] [Green Version]
  221. Maimela, N.R.; Liu, S.; Zhang, Y. Fates of CD8+ T cells in Tumor Microenvironment. Comput. Struct. Biotechnol. J. 2019, 17, 1–13. [Google Scholar] [CrossRef] [PubMed]
  222. Slaney, C.Y.; Kershaw, M.H.; Darcy, P.K. Trafficking of T Cells into Tumors. Cancer Res. 2014, 74, 7168–7174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Ishigami, S.; Natsugoe, S.; Tokuda, K.; Nakajo, A.; Che, X.M.; Iwashige, H.; Aridome, K.; Hokita, S.; Aikou, T. Prognostic value of intratumoral natural killer cells in gastric carcinoma. Cancer 2000, 88, 577–583. [Google Scholar] [CrossRef]
  224. Bald, T.; Krummel, M.F.; Smyth, M.J.; Barry, K.C. The NK cell-cancer cycle: Advances and new challenges in NK cell-based immunotherapies. Nat. Immunol. 2020, 21, 835–847. [Google Scholar] [CrossRef]
  225. Bikfalvi, A.; Billottet, C. The CC and CXC chemokines: Major regulators of tumor progression and the tumor microenvironment. Am. J. Physiol. Cell Physiol. 2020, 318, C542–C554. [Google Scholar] [CrossRef]
  226. Mizukami, Y.; Kono, K.; Kawaguchi, Y.; Akaike, H.; Kamimura, K.; Sugai, H.; Fujii, H. CCL17 and CCL22 chemokines within tumor microenvironment are related to accumulation of Foxp3+ regulatory T cells in gastric cancer. Int. J. Cancer. 2008, 122, 2286–2293. [Google Scholar] [CrossRef]
  227. Liu, X.; Zhang, Z.Z.; Zhao, G. Recent advances in the study of regulatory T cells in gastric cancer. Int. Immunopharmacol. 2019, 73, 560–567. [Google Scholar] [CrossRef]
  228. Parisi, L.; Gini, E.; Baci, D.; Tremolati, M.; Fanuli, M.; Bassani, B.; Farronato, G.; Bruno, A.; Mortara, L. Macrophage Polarization in Chronic Inflammatory Diseases: Killers or Builders? J. Immunol. Res. 2018, 2018, 8917804. [Google Scholar] [CrossRef] [Green Version]
  229. Ruytinx, P.; Proost, P.; Van Damme, J.; Struyf, S. Chemokine-Induced Macrophage Polarization in Inflammatory Conditions. Front. Immunol. 2018, 9, 1930. [Google Scholar] [CrossRef] [Green Version]
  230. Reyes, M.E.; de La Fuente, M.; Hermoso, M.; Ili, C.G.; Brebi, P. Role of CC Chemokines Subfamily in the Platinum Drugs Resistance Promotion in Cancer. Front. Immunol. 2020, 11, 901. [Google Scholar] [CrossRef]
  231. Chen, F.F.; Yuan, J.P.; Yan, H.L.; Liu, H.; Yin, S. Chemokine Receptor CXCR3 Correlates with Decreased M2 Macrophage Infiltration and Favorable Prognosis in Gastric Cancer. BioMed Res. Int. 2019, 2019, 6832867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Gambardella, V.; Castillo, J.; Tarazona, N.; Gimeno-Valiente, F.; Martinez-Ciarpaglini, C.; Cabeza-Segura, M.; Rosello, S.; Roda, D.; Huerta, M.; Cervantes, A.; et al. The role of tumor-associated macrophages in gastric cancer development and their potential as a therapeutic target. Cancer Treat. Rev. 2020, 86, 102015. [Google Scholar] [CrossRef] [Green Version]
  233. Boutilier, A.J.; Elsawa, S.F. Macrophage Polarization States in the Tumor Microenvironment. Int. J. Mol. Sci. 2021, 22, 6995. [Google Scholar] [CrossRef]
  234. Condamine, T.; Ramachandran, I.; Youn, J.I.; Gabrilovich, D.I. Regulation of Tumor Metastasis by Myeloid-Derived Suppressor Cells. Annu. Rev. Med. 2015, 66, 97–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Kramer, E.D.; Abrams, S.I. Granulocytic Myeloid-Derived Suppressor Cells as Negative Regulators of Anticancer Immunity. Front. Immunol. 2020, 11, 1963. [Google Scholar] [CrossRef]
  236. Echizen, K.; Hirose, O.; Maeda, Y.; Oshima, M. Inflammation in gastric cancer: Interplay of the COX-2/prostaglandin E2 and Toll-like receptor/MyD88 pathways. Cancer Sci. 2016, 107, 391–397. [Google Scholar] [CrossRef] [Green Version]
  237. Oshima, H.; Oshima, M. The role of PGE2-associated inflammatory responses in gastric cancer development. Semin. Immunopathol. 2013, 35, 139–150. [Google Scholar] [CrossRef]
  238. Chionh, Y.T.; Ng, G.Z.; Ong, L.; Arulmuruganar, A.; Stent, A.; Saeed, M.A.; Wee, J.L.; Sutton, P. Protease-activated receptor 1 suppresses Helicobacter pylori gastritis via the inhibition of macrophage cytokine secretion and interferon regulatory factor 5. Mucosal Immunol. 2015, 8, 68–79. [Google Scholar] [CrossRef] [Green Version]
  239. Guo, C.; Qi, H.; Yu, Y.J.; Zhang, Q.Q.; Su, J.; Yu, D.; Huang, W.D.; Chen, W.D.; Wang, Y.D. The G-Protein-Coupled Bile Acid Receptor Gpbar1 (TGR5) Inhibits Gastric Inflammation Through Antagonizing NF-kappa B Signaling Pathway. Front. Pharmacol. 2015, 6, 287. [Google Scholar] [CrossRef] [Green Version]
  240. Ajani, J.A.; D’Amico, T.A.; Bentrem, D.J.; Chao, J.; Cooke, D.; Corvera, C.; Das, P.; Enzinger, P.C.; Enzler, T.; Fanta, P.; et al. Gastric Cancer, Version 2.2022, NCCN Clinical Practice Guidelines in Oncology. J. Natl. Compr. Canc. Netw. 2022, 20, 167–192. [Google Scholar] [CrossRef]
  241. Van Cutsem, E.; Sagaert, X.; Topal, B.; Haustermans, K.; Prenen, H. Gastric cancer. Lancet 2016, 388, 2654–2664. [Google Scholar] [CrossRef] [PubMed]
  242. Shitara, K.; Van Cutsem, E.; Bang, Y.J.; Fuchs, C.; Wyrwicz, L.; Lee, K.W.; Kudaba, I.; Garrido, M.; Chung, H.C.; Lee, J.; et al. Efficacy and Safety of Pembrolizumab or Pembrolizumab Plus Chemotherapy vs Chemotherapy Alone for Patients With First-line, Advanced Gastric Cancer: The KEYNOTE-062 Phase 3 Randomized Clinical Trial. JAMA Oncol. 2020, 6, 1571–1580. [Google Scholar] [CrossRef] [PubMed]
  243. Hindson, J. Nivolumab plus chemotherapy for advanced gastric cancer and oesophageal adenocarcinoma. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 523. [Google Scholar] [CrossRef] [PubMed]
  244. Kubota, Y.; Kawazoe, A.; Sasaki, A.; Mishima, S.; Sawada, K.; Nakamura, Y.; Kotani, D.; Kuboki, Y.; Taniguchi, H.; Kojima, T.; et al. The Impact of Molecular Subtype on Efficacy of Chemotherapy and Checkpoint Inhibition in Advanced Gastric Cancer. Clin. Cancer. Res. 2020, 26, 3784–3790. [Google Scholar] [CrossRef] [Green Version]
  245. Smyth, E.C.; Gambardella, V.; Cervantes, A.; Fleitas, T. Checkpoint inhibitors for gastroesophageal cancers: Dissecting heterogeneity to better understand their role in first-line and adjuvant therapy. Ann. Oncol. 2021, 32, 590–599. [Google Scholar] [CrossRef]
  246. Chen, Y.; Wei, K.; Liu, D.; Xiang, J.; Wang, G.; Meng, X.; Peng, J. A Machine Learning Model for Predicting a Major Response to Neoadjuvant Chemotherapy in Advanced Gastric Cancer. Front. Oncol. 2021, 11, 675458. [Google Scholar] [CrossRef]
  247. Congreve, M.; de Graaf, C.; Swain, N.A.; Tate, C.G. Impact of GPCR Structures on Drug Discovery. Cell 2020, 181, 81–91. [Google Scholar] [CrossRef]
  248. Jabeen, A.; Ranganathan, S. Applications of machine learning in GPCR bioactive ligand discovery. Curr. Opin. Struct. Biol. 2019, 55, 66–76. [Google Scholar] [CrossRef]
  249. Hauser, A.S.; Attwood, M.M.; Rask-Andersen, M.; Schioth, H.B.; Gloriam, D.E. Trends in GPCR drug discovery: New agents, targets and indications. Nat. Rev. Drug Discov. 2017, 16, 829–842. [Google Scholar] [CrossRef]
  250. Ceylan, S.; Bahadori, F.; Akbas, F. Engineering of siRNA loaded PLGA Nano-Particles for highly efficient silencing of GPR87 gene as a target for pancreatic cancer treatment. Pharm. Dev. Technol. 2020, 25, 855–864. [Google Scholar] [CrossRef]
  251. Ma, X.; Xiong, Y.; Lee, L.T.O. Application of Nanoparticles for Targeting G Protein-Coupled Receptors. Int. J. Mol. Sci. 2018, 19, 2006. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Banerjee, D.; Harfouche, R.; Sengupta, S. Nanotechnology-mediated targeting of tumor angiogenesis. Vasc. Cell 2011, 3, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Corso, S.; Isella, C.; Bellomo, S.E.; Apicella, M.; Durando, S.; Migliore, C.; Ughetto, S.; D’Errico, L.; Menegon, S.; Moya-Rull, D.; et al. A Comprehensive PDX Gastric Cancer Collection Captures Cancer Cell-Intrinsic Transcriptional MSI Traits. Cancer Res. 2019, 79, 5884–5896. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The structure GPCRs and their dysregulation in GC (*, Conserved motifs). GPCRs are widely expressed in the stomach. GPCRs participate in a variety of physiological and pathological processes. The upregulated GPCRs, including PAR1, CXCR4 and P2YR, and BILF1, are identified in GC. The five key sequence motifs in the class A GPCRs represent the most frequent mutant sites, which are conserved and responsible for their structural integrity and essential function. Abbreviations: GC, gastric cancer; PAR1, Protease-activated receptor 1; CXCR4, Chemokine CXC receptor 4; P2YR, P2Y receptor.
Figure 1. The structure GPCRs and their dysregulation in GC (*, Conserved motifs). GPCRs are widely expressed in the stomach. GPCRs participate in a variety of physiological and pathological processes. The upregulated GPCRs, including PAR1, CXCR4 and P2YR, and BILF1, are identified in GC. The five key sequence motifs in the class A GPCRs represent the most frequent mutant sites, which are conserved and responsible for their structural integrity and essential function. Abbreviations: GC, gastric cancer; PAR1, Protease-activated receptor 1; CXCR4, Chemokine CXC receptor 4; P2YR, P2Y receptor.
Cancers 15 00736 g001
Figure 2. GPCR signaling and its crosstalk with other signaling pathways. (A) The most common GPCR-related signaling pathways. Agonist-stimulated GPCRs undergo a conformational change and facilitate the dissociation of Gα/Gβγ heterotrimer by replacing GDP with GTP on the Gα subunit. Subsequently, Ga and Gβγ trigger several downstream effectors, including secondary messenger systems, GEFs, Rho, and Ras GTPases, leading to a wide range of biological regulation. Besides the regulators of the G protein, signaling proteins (RGS proteins) promote the heteromeric complex reassociation and the signaling termination by accelerating intrinsic GTPase activity. Notably, agonist-activated GPCRs are also phosphorylated by GRKs and interact with β-arrestin, resulting in signaling desensitization and GPCR endocytosis. The endocytic β-arrestin-GPCR complex can be modulated by multiple factors and undergo degradation or recycling. (B) GPCR-associated crosstalk on the membrane and GPCR-EGFR crosstalk contain EGFR ligand-dependent transactivation and EGFR ligand-independent transactivation. The following pathways are the Wnt and Shh pathways. (C) The main pathways targeted by the multiple effectors in (a) consist of the following signaling pathways: Hippo pathway, MAPK pathway, Shh pathway, and Wnt pathway. Abbreviations: AC, adenylyl cyclase; AKT, protein kinase B; CREB, cAMP response element-binding protein; EGF, epidermal growth factor; EGFR, EGF receptor; ERK, extracellular signal-regulated kinase; GEF, guanine exchange factor; GLI, glioma-associated oncogenes; GPCR, G protein-coupled receptor; GRK, G protein-coupled receptor kinase; JNK, c-jun N-terminal kinase; LATS, large tumor suppressor kinase; MAPK, mitogen-activated protein kinase; mTOR, mammalian target of rapamycin; PDEs, phosphodiesterases; PI3K, phosphatidylinositol-3-kinase; PKA, Protein Kinase A; PLCβ, Phospholipase C β; ROCK, Rho-associated protein kinase; Shh, sonic hedgehog protein; SMO, Smoothened protein; SuFu, suppressor of fused; TAZ, transcriptional coactivator with PDZ-binding motif; TCF/LEF, T-cell factor/lymphoid enhancer factor; TEAD, transcriptional enhanced associate domain; YAP, yes-associated protein.
Figure 2. GPCR signaling and its crosstalk with other signaling pathways. (A) The most common GPCR-related signaling pathways. Agonist-stimulated GPCRs undergo a conformational change and facilitate the dissociation of Gα/Gβγ heterotrimer by replacing GDP with GTP on the Gα subunit. Subsequently, Ga and Gβγ trigger several downstream effectors, including secondary messenger systems, GEFs, Rho, and Ras GTPases, leading to a wide range of biological regulation. Besides the regulators of the G protein, signaling proteins (RGS proteins) promote the heteromeric complex reassociation and the signaling termination by accelerating intrinsic GTPase activity. Notably, agonist-activated GPCRs are also phosphorylated by GRKs and interact with β-arrestin, resulting in signaling desensitization and GPCR endocytosis. The endocytic β-arrestin-GPCR complex can be modulated by multiple factors and undergo degradation or recycling. (B) GPCR-associated crosstalk on the membrane and GPCR-EGFR crosstalk contain EGFR ligand-dependent transactivation and EGFR ligand-independent transactivation. The following pathways are the Wnt and Shh pathways. (C) The main pathways targeted by the multiple effectors in (a) consist of the following signaling pathways: Hippo pathway, MAPK pathway, Shh pathway, and Wnt pathway. Abbreviations: AC, adenylyl cyclase; AKT, protein kinase B; CREB, cAMP response element-binding protein; EGF, epidermal growth factor; EGFR, EGF receptor; ERK, extracellular signal-regulated kinase; GEF, guanine exchange factor; GLI, glioma-associated oncogenes; GPCR, G protein-coupled receptor; GRK, G protein-coupled receptor kinase; JNK, c-jun N-terminal kinase; LATS, large tumor suppressor kinase; MAPK, mitogen-activated protein kinase; mTOR, mammalian target of rapamycin; PDEs, phosphodiesterases; PI3K, phosphatidylinositol-3-kinase; PKA, Protein Kinase A; PLCβ, Phospholipase C β; ROCK, Rho-associated protein kinase; Shh, sonic hedgehog protein; SMO, Smoothened protein; SuFu, suppressor of fused; TAZ, transcriptional coactivator with PDZ-binding motif; TCF/LEF, T-cell factor/lymphoid enhancer factor; TEAD, transcriptional enhanced associate domain; YAP, yes-associated protein.
Cancers 15 00736 g002
Figure 3. GPCR-mediated metastasis and tumor microenvironment remodeling in GC. (A) The TME of GC consists of blood vessels, lymph vessels, immune cells, stromal cells (including fibroblast, pericytes, and adipocytes), extracellular matrix (ECM), and secreted soluble factors, such as proteins, RNAs, and small organelles. (B) GPCRs control the process of angiogenesis and metastasis. GPCR activation drives the production of stimulatory angiogenic factors like VEGF and EGF. These factors promote the development of new blood vessels by modulating the mitogenesis, migration, and sprouting of endothelial cells (ECs). Moreover, several GPCRs regulate the metastasis process by influencing ECM, degrading the status of cancer cells (EMT, migration, and invasion), and colonizing foreign sites. (C) Chemokine–chemokine receptors modulate immune responses. The chemokines are secreted by tumor cells, immune cells, and stromal cells. The interaction of chemokine and specific chemokine receptors recruits antitumor immune cells and immunosuppressive immune cells into the tumor microenvironment.
Figure 3. GPCR-mediated metastasis and tumor microenvironment remodeling in GC. (A) The TME of GC consists of blood vessels, lymph vessels, immune cells, stromal cells (including fibroblast, pericytes, and adipocytes), extracellular matrix (ECM), and secreted soluble factors, such as proteins, RNAs, and small organelles. (B) GPCRs control the process of angiogenesis and metastasis. GPCR activation drives the production of stimulatory angiogenic factors like VEGF and EGF. These factors promote the development of new blood vessels by modulating the mitogenesis, migration, and sprouting of endothelial cells (ECs). Moreover, several GPCRs regulate the metastasis process by influencing ECM, degrading the status of cancer cells (EMT, migration, and invasion), and colonizing foreign sites. (C) Chemokine–chemokine receptors modulate immune responses. The chemokines are secreted by tumor cells, immune cells, and stromal cells. The interaction of chemokine and specific chemokine receptors recruits antitumor immune cells and immunosuppressive immune cells into the tumor microenvironment.
Cancers 15 00736 g003
Table 1. The most reported GPCRs in GC.
Table 1. The most reported GPCRs in GC.
GPCRsLigandExpressionMechanismsBiological FunctionReferences
Class A Receptors
Peptide/Protein Receptors
Protease-activated receptors (PAR)Proteases, such as Thrombin, TFLLRN (synthetic PAR1-targeted peptide)PAR1/F2R: upregulation
PAR2/F2RL1: upregulation
PAR3/F2RL2: -
PAR4/F2RL3: -
H. pylori→ERK/PI3K-AKT→α-arrestin→PAR1→CXCL2
PAR2→MAPK→VEGF/COX-2
Inflammation, angiogenesis[99,100]
Angiotensin receptors (ATR)Angiotensin IIAT1R: upregulation
AT2R: upregulation
AT1R→VEGFAngiogenesis, metastasis[101,102]
Endothelin receptors (ETR)Endothelin-1ETAR: upregulation
ETBR: -
ETCR: -
ETAR→VEGF
ETAR→β arrestin/Src→EGFR
proliferation, metastasis[103]
Formyl peptide receptors (FPR)fMLF, capthespin GFP1R: -
FP2R/ALX: upregulation
FP3R: -
FP1R→ALOX5/15, SPMs (RvD1 and LXB4), SPM receptors (BLT1, ChemR23, GPR32)
FP2R→MAPK
FP1R: inhibiting angiogenesis and proliferation
FP2R: invasion and metastasis
[104,105,106]
Cholecystokinin receptors (CCKR)CCK, gastrinCCK1R: upregulation
CCK2R/GR: upregulation
gastrin/GR→PKC→IκB, NF-κBproliferation[107]
Leucine-rich repeat-containing receptors (LGRs) group BR-spondin1/2/3/4Lgr4: upregulation
Lgr5: upregulation
Lgr6
Lgr4/5/6→β catenin
Lgr6→PI3K/AKT/mTOR
proliferation, metastasis[108,109]
Lipid receptors
Lysophosphatidic acid receptors (LPAR)Lysophosphatidic acidLPA1/Edg-2: -
LPA2/Edg-4: upregulation
LPA3/Edg-7: -
LPAR2→tyrosine phosphorylation of c-Met
LPAR2→Gq11→p38
migration[110,111,112]
Sphingosine-1 phosphate receptors (S1PR)Lysophosphatidic acid: S1PS1P1R/Edg-1
S1P2R/Edg-5
S1P3R/Edg-3
: ubiquitously expressed
S1P4R/Edg-6
S1P5R/Edg-8
S1P1R→RAC-CDC42→ERK
S1P→EGFR, c-Met
S1P1R & S1P3R: promote proliferation and migration, angiogenesis
S1P2R: inhibit migration
[113,114]
Prostaglandin receptors (EPR)PGE2EP1R: -
EP2R: upregulation
EP3R: -
EP4R: -
PGE2→DNMT3B→5mC enrichment (DNA hypermethylation)
H. pylori→PGE2 upregulation→macrophage infiltration
proliferation, angiogenesis[115,116]
Chemokine receptors
Chemokine CXC receptors (CXCR)CXCL12-CXCR4/CXCR7CXCL8-CXCR1/CXCR2CXCL16-CXCR6CXCR1: upregulated
CXCR2: upregulated
CXCR3
CXCR4: upregulation
CXCR5
CXCR6: upregulation
CXCR7
CXCL12/CXCR4→PI3K/Akt/mTOR
CXCL12/CXCR4→ERK1/2
H. pylori→CXCL8→AKT/ERK/cyclin D1/EGFR/Bcl2/MMP9/MMP2
proliferation, migration, invasion, angiogenesis, metastasis[117]
Chemokine CC receptors (CCR)CCL2-CCR2CCL5-CCR5CCL19/CCL21-CCR7CCR1/3/4/5/6/8/9: -
CCR2/7: upregulation
CCR7→TGF-β1/NF-κBmigration, invasion, survival, metastasis[118]
Aminergic receptors
Muscarinic acetylcholine receptor Acetylcholine, carbachol,M1R:
M3R: upregulation
M2R/M4R/M5R: -
M1R-TRPC6→PKC
M2R/M4R→PKA→neurotransmitter release
M3R→EGFR→MAPK/ERK
M3R→Wnt pathway→YAP
proliferation, migration, invasion,[119]
β-adrenergic receptor (β-AR)isoproterenolβ1-adrenergic receptor: -
β2-adrenergic receptor: upregulation
β3-adrenergic receptor: -
ADRB2→NF-κB/AP-1/CREB/STAT3/ERK/JNK/MAPK→VEGF/MMP2/MMP7/MMP9proliferation, invasion, metastasis[120]
Nucleotide receptors
P2Y receptors (P2YR)ATPP2Y4: upregulation
P2Y6: downregulation
P2Y1/2/11-14: -
P2Y6→β catenin→ c-Myc
P2Y2→Gaq→p38-MAPK/ERK/JNK
proliferation[121]
Adenosine receptors (AR)adenosineA1/A3: -
A2aR: upregulation
A2bR: upregulation
A2aR→PI3K-AKT-mTOR
A2aR→PKA/PKC
proliferation, metastasis[122,123]
Steroid receptors
Membrane-type bile acid receptor (M-BAR/TGR5)Deoxyolate, bile acidsTGR5: upregulationTGR5→EGFR/MAPKproliferation[124]
Orphan receptors
GPR30G1 GPR30: upregulationGPR30→cAMP/Ca2+
GPR30→EGFR→PI3K/AKT/ERK
invasion, metastasis[125]
GPR39ObestatinGPR39: -GPR39→EGFR/MMP→AKT
GPR39/β-arrestin/Src→EGFR→AKT
proliferation[97]
Class B receptors
Hormone receptors
Growth hormone-releasing hormone (GHRH) receptor (GHRHR)GHRHGHRHR: upregulationGHRHR→PAK1→STAT3/NF-κBproliferation, inflammation[126]
Class C receptors
Ion receptors
Calcium-sensing receptor (CaSR)calcium ionsCaSR: upregulationCaSR→Ca2+/TRPV4/β-Cateninproliferation, migration, invasion[127]
Amino Acid receptors
γ-Aminobutyric acid (GABA) receptorGABAGABAA: upregulation
GABAB: -
GABAA→ERK1/2proliferation, invasion[128]
Metabotropic glutamate receptors (mGluRs)GlutamatemGluR5: upregulation
mGluR1/5 (group I): -
mGluR2/3 (group II): -
mGluR4/6/7/8 (group III): -
mGluR5→ERK1/2proliferation[129]
Adhesion receptors
ADGRE5 (CD97)CD55, α5β1 integrin, CD90ADGRE5: upregulationADGRE5→MAPKproliferation, metastasis[130]
Class F receptors
Fizzled receptorsWNT, lipoglycoproteinsFZD2/6/7: upregulation
FZD1/3/4/5/8/9/10: -
FZDs→Wingless/Int-1 (WNT)proliferation[131,132]
Smoothened receptors (SMO)cholesterol, sterolSmo: upregulationSMO→HH proliferation, invasion[131,133]
Viral receptors
EBV-encoded vGPCRmetal ion (Zn2+)BILF1: upregulationBILF1→MHC class 1proliferation, immune evasion[68,134]
Table 2. Immune cell infiltration induced by chemokine and related receptors.
Table 2. Immune cell infiltration induced by chemokine and related receptors.
Cell TypeReceptorsChemokinesMechanisms Underlying RecruitmentEffects on Tumor Cells after RecruitmentReferences
Anti-tumoral immune cells
Dendritic cellCXCR3, CXCR6CXCL4, CXCL1, CXCL16, CXCL17, CCL20IFN-γ-induced chemokines production, H. pylori involvementThe most potent professional antigen-presenting cells, activation of cellular immunity, and T cell-dependent humoral immunity[219,220]
CD8 T cellCXCR3CXCL9, CXCL10, CXCL11CAFs-mediated IL6 secretion, tumor cell chemokines secretion, adhesion molecules (ICAM-1, VCAM-1)Differentiated into cytotoxic CD8+ T cells to destroy tumor cells or memory CD8+ T cells to recirculate in the blood[221,222]
NK cellCXCR1, CXCR2, CXCR4, CX3CR1, CCR5, CCR10CXCL10, CXCL12, CCL21, CX3CL1, CCL5, CCL27Chemokine signaling regulated by HLA-G and CD47; stromal barriersCytokine production and cytotoxicity on tumor cells through STAT3; regulating DC maturation; modulating T cell activity[223,224]
M1 macrophageCCR2, CCR5CCL2, CCL5Disrupting NF-κB signaling or interacting with TNF-α;High capacity to present antigens; proinflammatory cytokines (IL-1β, IL-1α, IL-12, TNF-α, and GFAP) production; stimulation of type-I T cell responses[232,233]
Tumor-promoting immune cells
TregCCR4, CCR10CCL17, CCL22Stimulation of JAK-STAT3 signaling pathway; remodeling of gastric microbiota by H.pylori.; stimulation of DCs due to H.pylori. infectionSuppressing CD4+ T cells, CD8+ cells, antigen-presenting cell (APC), monocytes, and macrophages; inhibitory cytokines like IL10, IL35, and TGF-β; inducing apoptosis by perforin/ granzyme production[226,227]
M2 macrophageCCR2, CCR5, CXCR3, CXCR4CCL2, CCL5, CXCL9, CXCR12STAT3 activation; PI3K/AKT/mTOR signaling pathwayGrowth factors (FGF, VEGF, and IL-6) production; secreting matrix-degrading enzymes and cytokines[230,231]
Monocytic MDSCCCR2, CXCR2, CXCR4CCL2, CXCL5, CXCL12IL-6 production, JAK-STAT3 signaling,High amounts of NO, Arg1, and immune-suppressive cytokines; suppression of nonspecific T cell responses[234]
Granulocytic (or PMN-) MDSCCXCR1, CXCR2CXCL8, CXCL1, CXCL12, CXCL5, CXCL6HGF/TGF-β/MCP-1 production, JAK-STAT3 signaling, IRF-8, NF-κB pathway, hypoxiaLarge amounts of O2−, H2O2, and PNT (ROS) production; blocking T cell proliferation; depleting entry of CD8+ T cells to tumors[234,235]
Table 3. Drugs and antibodies against GPCRs in GC clinical trials [6,135,247].
Table 3. Drugs and antibodies against GPCRs in GC clinical trials [6,135,247].
Drug NameTargeted GPCRsTypes of DrugsTested Cancer TypesStatusNCT
MogamulizumabCCR4mAbCataneous/Peripheral T-cell lymphoma; Adult T-cell lymphomaPhase I: completeNCT02946671
VismodegibSMOsmall moleculeBasal-cell carcinoma; Head and neck cancerPhase II: complete
Phase II: complete
Phase II: recruiting
NCT03052478
NCT00982592
NCT02465060
SonidegibSMOsmall moleculeBasal-cell carcinomaPhase I: recruiting
Phase I: complete
NCT04007744
NCT01576666
Lutathera (Lutetium Lu 177 dotatate)SSTRpeptideGastroenteropancreatic neuroendocrine tumors (GEP-NETs)Most on recruitingNCT04949282
NCT04727723
NCT04609592
NCT04524442
NCT02736500
NCT02489604
NCT04614766
NCT01860742
LanreotideSSTRpeptideAdvanced prostate cancerPhase III: recruitingNCT04852679
NCT03043664
NCT03017690
NCT02730104
NCT02736448
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yan, H.; Zhang, J.-L.; Leung, K.-T.; Lo, K.-W.; Yu, J.; To, K.-F.; Kang, W. An Update of G-Protein-Coupled Receptor Signaling and Its Deregulation in Gastric Carcinogenesis. Cancers 2023, 15, 736. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers15030736

AMA Style

Yan H, Zhang J-L, Leung K-T, Lo K-W, Yu J, To K-F, Kang W. An Update of G-Protein-Coupled Receptor Signaling and Its Deregulation in Gastric Carcinogenesis. Cancers. 2023; 15(3):736. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers15030736

Chicago/Turabian Style

Yan, Huan, Jing-Ling Zhang, Kam-Tong Leung, Kwok-Wai Lo, Jun Yu, Ka-Fai To, and Wei Kang. 2023. "An Update of G-Protein-Coupled Receptor Signaling and Its Deregulation in Gastric Carcinogenesis" Cancers 15, no. 3: 736. https://0-doi-org.brum.beds.ac.uk/10.3390/cancers15030736

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