Next Article in Journal
Esketamine Inhibits Cocaine-Seeking Behaviour Subsequent to Various Abstinence Conditions in Rats
Previous Article in Journal
Mitochondrial Calcium Overload Plays a Causal Role in Oxidative Stress in the Failing Heart
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of G Protein-Coupled Estrogen Receptor (GPER) in Vascular Pathology and Physiology

1
Xi’an Medical University, Xi’an 710068, China
2
Department of Cardiovascular Surgery, Xijing Hospital, Fourth Military Medical University, Xi’an 710032, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomolecules 2023, 13(9), 1410; https://doi.org/10.3390/biom13091410
Submission received: 1 August 2023 / Revised: 12 September 2023 / Accepted: 15 September 2023 / Published: 19 September 2023
(This article belongs to the Section Biomacromolecules: Proteins)

Abstract

:
Objective: Estrogen is indispensable in health and disease and mainly functions through its receptors. The protection of the cardiovascular system by estrogen and its receptors has been recognized for decades. Numerous studies with a focus on estrogen and its receptor system have been conducted to elucidate the underlying mechanism. Although nuclear estrogen receptors, including estrogen receptor-α and estrogen receptor-β, have been shown to be classical receptors that mediate genomic effects, studies now show that GPER mainly mediates rapid signaling events as well as transcriptional regulation via binding to estrogen as a membrane receptor. With the discovery of selective synthetic ligands for GPER and the utilization of GPER knockout mice, significant progress has been made in understanding the function of GPER. In this review, the tissue and cellular localizations, endogenous and exogenous ligands, and signaling pathways of GPER are systematically summarized in diverse physiological and diseased conditions. This article further emphasizes the role of GPER in vascular pathology and physiology, focusing on the latest research progress and evidence of GPER as a promising therapeutic target in hypertension, pulmonary hypertension, and atherosclerosis. Thus, selective regulation of GPER by its agonists and antagonists have the potential to be used in clinical practice for treating such diseases.

1. Introduction

Cardiovascular disease (CVD) is currently the primary cause of death worldwide. In 2020, an estimated 19.05 million people died from CVD globally and 207.1 of 100,000 people died from hypertension and stroke, the prevalence of which was higher than that of cancer and chronic diseases [1]. Interestingly, epidemiological and clinical studies demonstrate clear sex differences for cardiovascular risk, prevalence, prognosis, pathophysiological manifestation, and response to treatment for CVD [2,3]. The incidence of cardiovascular disease is lower in premenopausal women than in age-matched men but sharply increases in postmenopausal women [4,5,6], implying that estrogen provides a protective effect and that the increased cardiovascular risk in women after menopause is related to the loss of the protective effect of estrogen [7,8]. The protection by estrogen was originally thought to be attributed to the classical estrogen receptors ERα and ERβ. However, the inhibition of these receptors hardly abolishes the protective effect of estrogen, indicating that a novel, unidentified receptor that was yet to be discovered was the actual mediator. This was later characterized as the G protein-coupled estrogen membrane receptor (GPER) [9,10].
GPER is present throughout the cardiovascular system, implying a critical role in maintaining cardiovascular physiological function. GPER activation by G1 restored glucose homeostasis and reduced body weight in multiple preclinical male and female models associated with obesity, diabetes, and metabolic abnormalities [11]. In blood vessels, GPER has important regulatory effects on the pathophysiological function of endothelial cells and smooth muscle cells [12,13]. Importantly, G1 treatment may regulate the renin–angiotensin system to reduce blood pressure in ovariectomized mRen2.Lewis rats [14]. These data collectively demonstrate that GPER plays a significant role in vascular-related diseases, including hypertension and atherosclerosis.

2. Tissue, Cellular Localization, Ligands, and Classical Signaling Pathways of GPER

GPER was initially identified as an orphan receptor with no known ligand, and belongs to the superfamily of 7-transmembrane G protein-coupled receptors (GPCRs). It was later named GPR30 based on the sequential numbering scheme for orphan receptors [15]. GPR30 was first shown to act as a functional estrogen receptor to mediate the fast activation of extracellular signal-regulated kinase (ERK) in the year 2000 [16]. Subsequently, experimental data from both groups showed that GPR30 directly bound to estrogen and activated its downstream intracellular signaling cascades [17,18]. Therefore, the International Union of Basic and Clinical Pharmacology officially named GPR30 as G protein-coupled estrogen receptor (GPER) [19].

2.1. Tissue and Cellular Localization of GPER

GPER is widely distributed in various tissues of the human body. GPER first attracted considerable attention in reproduction as an estrogen receptor that functioned in the breast [17,20], ovary [21], endometrium [22,23], testis [24], and prostate [25,26], as well as the placenta [27,28]. GPER is also found in other systems and organs, including the brain, lung, liver, heart, pancreatic islets, adipose tissue, vasculature, muscle, and skeleton, as well as in immune cells [29,30]. In the vascular system, GPER is distributed in both endothelial and smooth muscle cells of human arteries and veins [31,32,33].
The cellular localization of GPER has been a subject of much debate since its discovery. Most GPCRs are lipophilic, localized to the plasma membrane [34], and the signal acts by transferring from the outside to the inside of the cell across the plasma membrane. As a member of GPCRs, GPER was also initially thought to be located in the plasma membrane [35]. The immunohistochemical results obtained by Thomas et al. in breast cancer cells lacking ERα and ERβ were the first to confirm the plasma membrane localization of GPER [17]. Immunohistochemical analysis showed that GPER localized to the surface of CA2 cells [36]. Confocal microscopy using the agglutinin canavalin A as a marker of the plasma membrane also confirmed the expression of GPR30 on the cell surface [37]. Using RT-PCR and Western blot, it was found that GPER and calcin-1 were co-located on the cell membrane of human myometrium [38]. GPER was co-located with giant proteins on the membrane of proximal renal cortex tubules [39]. GPER was also detected in the dendritic spines of hippocampal neurons [40]. However, more studies have shown the intracellular localization of GPER. The intracellular localization of GPER was first discovered by Revankar et al. and was found to be localized in the inner membrane of the endoplasmic reticulum and Golgi apparatus of monkey kidney fibroblasts [18]. In the paraventricular and supraoptic nuclei, specific markers of intracellular organelles and immunoelectron microscopy revealed that GPR30 was mainly located in the Golgi apparatus of neurons, but could not be detected on the cell surface [41]. In transiently transfected cells, as well as in cells endogenously expressing GPR30, the GPER was shown to be localized to the endoplasmic reticulum [42]. In human umbilical vein endothelial cells (ECs) and vascular smooth muscle cells (VSMCs), GPER was shown to be expressed primarily on the endoplasmic reticulum and Golgi membrane [32]. Indirect evidence supporting the intracellular localization and functionality of GPER is that only cell-permeable E2 derivatives or intracellularly injected ligands resulted in rapid GPER activation [43], and intracellularly injected ligands produced faster and deeper calcium responses compared to extracellular ligands [44]. In some other cells, GPER was found expressing on the plasma membrane and inner membrane simultaneously. For example, almost all neurons in the trigeminal ganglion expressed GPER on the cell surface and in the cytoplasm [45]. Immunofluorescent results in human glioblastoma cells indicated that GPER was in the plasma membrane, cytoplasm, and nucleus [46]. In breast cancer cells, GPER has also been found to be localized in the nucleus, which is caused by the transport of GPER from the plasma membrane to the nucleus [47,48]. In addition, a prominent sex difference in the subcellular localization of GPER was detected in cardiomyocytes. In both wild-type and RenTgMK (a well-described model of angiotensin II-mediated chronic hypertension and cardiac hypertrophy [49,50,51]) female mice, the level of membrane localization of GPER protein was higher than in males of the same group [52]. Obviously, the localization of GPER is complex, which may be due to different species, sexes, and different tissues and cell types. It is worth noting that some studies have confirmed that GPER is activated within cells and then transported to the plasma membrane to play a role [53], which is regulated by co-expressed receptor activity-modifying protein 3 [52]. The transmembrane transport of GPER exerts physiological effects in a way that induces dynamic changes in the cellular localization of GPER, which undoubtedly makes the localization of GPER more complicated; thus, understanding the cellular localization of GPER is of importance for elucidating its function.

2.2. Endogenous Ligands of GPER

2.2.1. Estrogen

Estrogen is a steroid hormone which is secreted mainly from the ovaries, and four natural estrogens in females are biosynthesized: estrone (E1), 17β-estradiol (E2), 17α-estradiol (E2), and estriol (E3) [17,18,54]. Among them, 17β-estradiol is the most biologically active and common form of estrogen, and is primarily produced by the transformation of precursor testosterone in the ovaries of premenopausal women [15]. Estrogen is required for the development of female secondary sexual characteristics and affects the structure and function of the female reproductive system [55]. Estrogen was the first discovered ligand of GPER, and mediates pleiotropic functions by activating GPER in different tissues [56]. In the pathogenesis of diverse cardiovascular diseases, estrogen exerts important protective effects by activating GPER, which include the dilation of blood vessels, the prevention of atherosclerosis, a reduction in blood pressure, the suppression of myocardial hypertrophy, and the inhibition of myocardial ischemic injury [56]. Conversely, the absence of GPER exacerbates these pathologic reactions. The earliest experiments using GPER knockout mice showed that the loss of GPER increases blood pressure, eliminates E2-stimulated insulin release, and affects glucose metabolism, thereby leading to visceral obesity [57,58]. Hypertension due to GPER deficiency is associated with increased endothelial prostanoid-mediated vasoconstriction and increased contraction mediated by the thrombolin prostaglandin (TP) receptor [59]. In addition, GPER knockout mice showed an early increase in atherosclerotic progression [32]. In the heart, GPER-deficient mice exhibited impaired left ventricular cardiac function, left ventricular enlargement, and impaired systolic and diastolic function [60], or reduced ejection fraction and brachyaxis shortening, especially in older male mice [61]. Wang et al. showed that cardiomyocyte-specific GPER deletion caused left ventricular dysfunction and adverse remodeling in mice. Further specific spectral analysis showed that mitochondria-related genes were enriched in females of GPER KO mice, while genes related to the inflammatory response were enriched in male GPER KO mice [62]. This indicated that cardiac remodeling caused by GPER loss is associated with oxidative stress and NOD-like receptor thermal protein domain associated protein 3 (NLRP3)-mediated inflammatory activation [63,64]. GPER thus plays a protective function in the cardiovascular system.
GPER binds to estradiol with high affinity and was thought to mediate rapid non-genomic effects distinct from those of ERα and ERβ [16,18,65,66]. Unlike most GPCRs, GPER couples to both a Gαs protein and a pertussis toxin (PTX)-sensitive Gαi/o protein, which activate intracellular signaling cascades using cAMP and Ca2+ as second messengers [16,17,65,66]. GPER is involved in the activation of multiple downstream cellular signaling cascades, as summarized in Figure 1. The first demonstrated signaling pathway activated by GPER involves the release of a tyrosine kinase (Src) to cleave matrix metalloproteinases (MMPs) which mediated heparin-binding epidermal growth factor (HB-EGF) activation. The subsequent transactivation of the epidermal growth factor receptor (EGFR) can further activate the ERK1/2 signaling pathway [67]. Through this mechanism, GPER mediates vasodilation by promoting NO production [68], induces smooth muscle cell proliferation [69], and promotes angiogenesis by upregulating angiotensin [70], and protects the myocardium from ischemia/reperfusion injury by reducing mitochondrial dysfunction and mitochondrial autophagy [71,72]. In addition, relevant studies in lung and breast cancer have shown that the activation of the Src/ERK1/2 pathway by GPER involves the activation of mitogen-activated protein kinase kinases (MEK) [73,74,75,76], which is consistent with our general understanding of the mitogen-activated protein kinase (MAPK) signaling pathway. Another classical pathway is phosphatidylinositol 3-kinase (PI3-K) and serine/threonine kinase (Akt), also known as protein kinase B (PKB), which appears upstream of ERK or as an independent signaling pathway and plays a role in cell proliferation [77], apoptosis [78,79], and cell migration [80,81]. It should be noted that ERα also activates PI3K, which does not involve EGFR transactivation [18]. GPER was also found to activate the AMP-activated protein kinase (AMPK) signaling pathway, which enhances adenylyl cyclase (AC) activity to activate cAMP-activated protein kinase A (PKA). The PKA pathway has been shown to induce vasodilation in coronary vessels through GPER-mediated myosin light chain (MLC) phosphorylation and the ras homolog gene family, member A/ras homolog gene family (RhoA/Rho) kinase pathway [82,83]. Notably, the PKA pathway has been reported to have an inhibitory effect on ERK1/2 [65,84]. In addition, GPER has also been reported to inhibit inflammation [85,86,87], tumor cell migration, and tumor angiogenesis through reducing phosphorylation, nuclear localization, and the transcriptional activity of NF-κB, and TNF-α- or lipopolysaccharide-induced IL-6 secretion [88,89,90,91].

2.2.2. Aldosterone

As a vasoactive steroid, aldosterone has recently attracted much interest, not only due to its great importance in the regulation of salt and water balance, but also because of its key role in cardiovascular metabolism and function. The classical effects of aldosterone are mediated by the activation of mineralocorticoid receptors (MRs) and subsequent transcriptional regulatory mechanisms [92]. However, aldosterone showed some rapid regulatory effects that cannot be explained by MRs [93], which appeared to involve a different receptor [94]. Gros et al. first described the role of GPER as a receptor involved in aldosterone-mediated signal pathways [95,96], indicating that GPER is an authentic aldosterone receptor.
Despite accumulating evidence for aldosterone’s actions through GPER, there remains significant speculation that aldosterone directly binds to GPER [97,98,99]. However, a recent work has demonstrated a direct interaction of aldosterone with GPER through the data showing that aldosterone competitively binds GPER with other GPER ligands [100]. Meanwhile, in a reperfusion injury study of the heart, using a polymer analogue of aldosterone to selectively activate the non-genomic receptors of aldosterone, it was found that the GPER antagonist G-36 but not spironolactone, a MR blocker, could block the action of aldosterone [101]. These data indicate that GPER, rather than MRs, was the receptor for the non-genomic action of aldosterone.
The mechanism of the downstream signal pathways via the interaction between aldosterone and GPER are not well understood. Recent studies have revealed that aldosterone trans-activates EGFR via GPER in adult rat ventricular myocytes, followed by reactive oxygen species (ROS) production and PI3K/AKT-dependent pathway activation, ultimately resulting in the stimulation of cardiac sodium/bicarbonate cotransporter (NBC) [102,103]. In addition, aldosterone also increases the cardiac vagal tone via GPER [104]. In the kidney, aldosterone has been reported to enhance the connecting tubule–glomerular feedback (CTGF) and stimulate oxygen production through the cAMP/PKA/PKC pathway in the connecting tubule (CNT) [105]. GPER in the distal tricyclic tubule (DCT) rapidly leads to intracellular cAMP and inositol phosphate accumulation, causing the activation of various downstream pathways including the MAPK/ERK, PI3K/AKT, and cAMP/PKA pathways, and the acute activation of thiazide-sensitive NaCl cotransporter (NCC) activity, thus regulating renal NaCl excretion [106].
GPER has also been shown to regulate aldosterone synthesis. Aldosterone increases the expression of cyp11b2 mRNA through the activation of GPER, which is the last rate-limiting enzyme for aldosterone synthesis and up-regulates aldosterone secretion. This aldosterone-induced aldosterone production may, under certain circumstances, such as volumetric hypotension, physiologically increase aldosterone production to maintain blood pressure [107]. In contrast to GPER, ERβ was stated to inhibit aldosterone synthase expression and aldosterone synthesis through the finding that the selective blockade of ERβ markedly increased aldosterone expression, which was mimicked by the GPER-selective agonist G-1 and abolished by co-treatment with either the GPER antagonist G-15 or a selective protein kinase A inhibitor [108]. Thus, estrogen may exert different effects on aldosterone synthesis through different receptors, but exhibits inhibitory effects in physiological situations through ERβ [108]. In premenopausal women, high estrogen levels may maintain normotension by inhibiting aldosterone synthesis through ERβ, and after menopause, the loss of estrogen can lead to an increase in blood pressure through GPER-mediated aldosterone release [109], which may suggest the cause of the high prevalence of hypertension in postmenopausal women.

2.3. Exogenous Ligand of GPER

In addition to endogenous steroid hormones, many synthetic estrogenic compounds, as well as phytoestrogenic compounds of natural origin, have also been shown to bind GPER to exert their effects. Ligand-binding studies revealed that GPER binds E2 with much weaker binding affinities at concentrations from Kd = 3 to 6 nM [17] than classical ERs from Kd = 0.1 to 1.0 nM [110], so it was difficult to clarify the role of GPER alone in in vivo experiments without the synthetic highly selective GPER agonist G1 [111]. G-1 showed a stronger affinity (7–11 nM) and higher selectivity for GPER than E2 [111,112], and did not bind ERα/β at concentrations up to 10 μM [113], which provides an extremely convenient tool to study the independent effects of GPER in vivo. To date, G-1 has been the most widely studied GPER agonist. A lot of the beneficial effects of GPER in the whole body are replicated by G-1, and unlike estrogen, G-1 can be considered a non-feminizing estrogenic compound due to its minimal effect on reproductive tissues; it thus has potential therapeutic use in both women and men [114]. Notably, the effect of G1 on GPER is concentration-dependent. G1 has other effects independent of GPER at micromolar concentration ranges. G1 (1–10 μM) inhibits the mitosis of human vascular smooth muscle cells and induces apoptosis by destroying microtubules [115]. Similarly, reduced DNA synthesis and microtubule structure disruption were observed in 2–3 μM G1-treated microvascular endothelial cells and HUVECs, which were non-receptor-dependent [116,117]. This concentration-dependent anti-proliferative effect was also observed in endometriosis stromal cells. G1 blocks microtubule assembly and induces cell cycle arrest in ESCs [118]. Therefore, in the application of G1 therapy or experimental treatment, attention should be paid to the concentration-dependent effect of G1 to prevent the influence of non-specific effects.
After the discovery of G-1, researchers subsequently synthesized the first GPER antagonist G-15 [112] and its modified analogue G-36 [113]. G-15 is a synthetic substituted dihydroquinoline with similar structure and a weak binding affinity for ERα/β similar to G-1 [119] in the absence of the ethanone moiety [112]. Two additional compounds, GPER-L1 and GPER-L2, have been reported to act as novel GPER-selective agonists with a binding affinity of approximately 100nM [120]. The discovery of selective ligands for GPER has greatly helped us to understand the physiological role of GPER. In addition, numerous synthetic estrogenic compounds have been shown to bind and/or activate GPER, including bisphenol A [121,122,123,124], nonylphenol [124,125,126], atrazine [124,127,128], kepone, methoxychlor, p,p’-DDT, o,p’-DDT, o,p’-DDE, p,p’-DDE, 2,2’5’-PCB-4-OH, etc. [17,124]. Lastly, there are some plant-derived polyphenols that have also been shown to bind to GPER, including Genistein [124,129,130,131], Daidzein [132], Equol [133], Oleuropein [134], Hydroxytyrosol [133], Resveratrol [135,136,137], Quercetin [136,138], Tectoridin [139], Zearalenone [124,140,141], and Apigenin [136,142]. The structure and function of GPER ligands are summarized in Table 1.

3. The Role of GPER in Vascular Pathology and Physiology

3.1. Blood Pressure Regulation and Hypertensive Disorders

Hypertension (HTN) is a multifactorial disease, defined as a systolic blood pressure (SBP) greater than 130 mmHg or diastolic blood pressure (DBP) higher than 80 mmHg [146,147], and is an important risk factor for cardiovascular diseases such as coronary heart disease, heart failure, and stroke. HTN is characterized by persistent elevated blood pressure levels and increased peripheral vascular resistance [148,149]. The gene of GPER is mapped to chromosome 7p22 [150], a region associated with human arterial hypertension [151], which suggests a correlation between GPER and blood pressure control. Data have demonstrated that GPER is associated with hypertension in postmenopausal women [152]. The difference in gender hypertension may be related to the impaired function of GPER, since the frequency of the P16L allele, a lower-functioning variant of GPER, is statistically almost twice as high in female hypertension patients as in male patients [153]. However, aldosterone-induced hypertension has sex differences, which is mediated by T cells and attenuated by the non-T cell GPER-dependent pathway [154]. Aortic remodeling caused by salt-sensitive hypertension was also ameliorated via GPER activation [155].
Animal experiments proved that GPER could mediate vasodilation in rats and reduce blood pressure in normotensive rats [58]. In the hypertensive rat model, G1 treatment was able to reduce blood pressure by dilating rat mesenteric arteries [156]. This rapid blood-pressure-lowering effect of GPER protects the heart from hypertensive damage [58]. The estrogen-induced endothelium-dependent vasodilation was revealed to correlate with GPER-mediated nitric oxide (NO) production [13]. As a mediator of acute estrogen-dependent vasodilation, GPER is responsible for the control of vascular tone involving NO and hypotensive activity [157]. It is an important regulator of endothelial cells to regulate the tension of resistance vessels [158]. GPER activation induces endothelial cells to produce NO, which mediates vasodilation in a variety of blood vessels, including carotid artery rings of Sprague Dawley rats [58], isolated aortic and mesenteric rings of mRen2 mice [39,156], aortic rings [96,159,160] and mesenteric arteries of rats [66], and rat uterine arteries [161,162]. Both estrogen and aldosterone can activate GPER to exert a rapid vasodilator effect [160,163], but the vasodilator effect of aldosterone is almost entirely dependent on GPER [96]. Notably, the estrogen-mediated activation of eNOS is not only dependent on GPER, but also dependent on ERα [164], but GPER is more involved in estrogen-induced endothelial vasodilation and NO production [13].
NO is produced in endothelial cells, which involves the phosphorylation of eNOS caused by Src/EGFR/PI3K–Akt and ERK pathways [13,68]. Pollyana et al. showed that the GPER-dependent relaxation of mesenteric resistance arteries is mainly mediated by the PI3K–Akt–eNOS pathway and attenuated by non-specific potassium channel block [165]. The molecular mechanism of GPER-dependent activation of the ERK1/2, PI3K-Akt, and NO signaling pathways was also supported by dibutyl phthalate (DBP)-induced endothelial cell proliferation [70]. Notably, there appear to be sex differences in the mechanistic pathways involved in eNOS phosphorylation: the PI3k–Akt–eNOS pathway was identified in males, and the MEK–ERK–eNOS pathway was validated in females [166]. In addition to this, NO production can also occur through genomic pathways. The underlying mechanisms include the activation of GPER which increases intracellular Ca2+ levels through transient receptor potential channel (TRPC) isoforms, leading to the phosphorylation of calmodulin-dependent protein kinase 2 (CaMKKβ), AMPK, and calmodulin-dependent protein kinase II (CaMKIIα), which in turn increase the expression of eNOS through the activation of histone deacetylase 5 (HDAC5) and the Kruppel-like factor 2 (KLF2) signaling pathway [68]. GPER also increased the phosphorylation of EGFR by interacting with Src, which activates eNOS through ERK5 and muscle cell enhancer 2 peptide C (MEF2C) [68]. After endothelium-dependent NO production, VSMCs are stimulated, causing downstream guanylate cyclase-mediated cGMP release and protein kinase G stimulation, followed by Ca2+ influx channel opening and myosin light chain phosphatase (MLCP) activation [66,167] (Figure 2). Yu et al. revealed that G-1 exerts a concentration-dependent relaxation effect on isolated porcine coronary arteries by increasing potassium efflux from the large-conductance calcium-activated potassium (BKCa) channel, which was not endothelium-dependent [168]. Another report showed that the activation of GPER inhibited the RhoA/Rho kinase pathway through the cAMP/PKA signaling pathway, which in turn activated MLCP to mediate vasodilation in isolated porcine coronary arteries [82]. In addition, they identified an additional GPER-dependent exchange protein directly activated by camp 1/ras-related protein 1 (Epac/Rap1) pathway parallel to PKA, which inhibited the RhoA/Rho kinase pathway [83]. Later, they demonstrated that cAMP production was dependent on EGFR transactivation and ERK1/2 phosphorylation, which mediated the relaxation of isolated porcine coronary arteries [169]. In conclusion, GPER mediates vascular relaxation through RhoA/Rho inhibition, MLCP activation, and Ca2+ channel opening via the endothelium-dependent PKG pathway and the endothelium-independent PKA and Epac/Rap1 pathways (Figure 2). Indeed, the synergistic effect of PKA and PKG on VSMC relaxation is evidenced by vasodilation being inhibited by the administration of cAMP and cGMP antagonists [170] and vasodilation being partially abolished in mesenteric arteries by the administration of AC and GC inhibitors [66]. Due to the vasodilation and protective effects of GPER on blood vessels, GPER and its related pathways would be potential therapeutic targets for the treatment of hypertension, especially in postmenopausal women.

3.2. Pulmonary Arterial Hypertension

Pulmonary arterial hypertension (PAH), defined as elevated mean pulmonary arterial pressure (PAP, typically >25 mmHg at rest, or >30 mmHg with exercise, or >40 mmHg for systolic), is characterized by the pathological elevation of pulmonary artery pressure leading to remodeling of the right ventricle and ultimately to severe right heart failure [171]. The prevalence of PAH is higher in women [172], thus implying the vital role of estrogen in the pathogenesis of PAH. However, a large amount of experimental evidence has shown that women have a lower severity of PAH than men, as well as a better prognosis. The symptoms of PAH are aggravated after ovariectomy, indicating that estrogen has an inhibitory effect on the occurrence and development of PAH. This phenomenon is called the “estrogen paradox” [173,174,175]. The protective effect of estrogen on pulmonary hypertension has been demonstrated [174,175], but as a new membrane receptor of estrogen, the effect of GPER on PAH is still under investigation. Interestingly, the activation of GPER by G1 demonstrated beneficial effects on PAH-related cardiopulmonary functional and structural abnormalities in both genders. In male rats, it was mediated by improving pulmonary endothelial nitric oxide synthesis, Ca2+ handling regulation, a decrease in inflammation in cardiomyocytes, and a decrease in collagen deposition by acting in pulmonary and cardiac fibroblasts [176]. In female rats, it was mediated by reducing right ventricular collagen deposition and hypertrophy via decreasing glycosaminoglycan content and reactive oxygen species production in female rats [177].

3.3. Inflammation, Lipid Metabolism, and Atherosclerosis

Atherosclerosis is characterized by plaque formation leading to the stenosis of the arterial wall, and is associated with a long-term state of dyslipidemia and vascular inflammation [178]. Correlations between GPER and inflammation, lipid metabolism, and atherogenesis have been demonstrated [32]. A study using B-ultrasound to investigate atherosclerosis in postmenopausal women showed a significantly lower prevalence of atherosclerosis in late menopausal women and in women receiving postmenopausal estrogen therapy compared to early menopausal women and women without postmenopausal estrogen therapy [179]. The absence of GPER in male mice may generate some pathological changes, such as insulin resistance, dyslipidemia, and inflammatory effects [180]. Decreased high-density lipoprotein (HDL) cholesterol and increased triglyceride levels were observed in GPER-deficient female mice, and aged male GPER-deficient mice exhibited dyslipidemia and inflammation [180]. In a pathological model of atherosclerosis in female mice, a loss of GPER increased total cholesterol (TC) and low-density lipoprotein (LDL) levels [32], and treatment with G-1 was effective in attenuating atherosclerosis in postmenopausal mice, which was associated with a significant reduction in vascular inflammation [32]. The effects of GPER on inflammation, lipid metabolism, and atherosclerosis have not only been validated in animal models but also in humans. In a cohort study, humans carrying a low-functioning P16L genetic variant of GPER were found to have increased plasma LDL cholesterol [181]. Meanwhile, in the human hepatic carcinoma cell line, G1 treatment led to a concentration-dependent elevation in LDL receptor expression, and the effects of G1 were attenuated by G15 as a GPER antagonist or GPER shRNA [181], suggesting that GPER plays an important role in regulating LDL receptor expression and LDL metabolism. Ghaffari et al. used total internal reflection fluorescence microscopy to quantify the transcellular transport of LDL in human coronary-artery endothelial cells obtained from multiple donors. The rate of transcytosis was higher in coronary artery endothelial cells isolated from male patients compared to premenopausal female patients [182]. Estrogen administration reduced LDL transcytosis in endothelial cells from male but not female donors, and the inhibition of ERα and ERβ had no effect on the estrogen-mediated attenuation of LDL trans-endocytosis. Instead, estrogen’s effect on LDL transcytosis was blocked by the depletion of GPER [182]. Apparently, the effects of estrogen on LDL are mediated through GPER but not ERα or ERβ.
Further evidence for the anti-inflammatory effects of GPER in blood vessels and inflammatory cells was obtained in in vitro studies. In human umbilical vein vessels, GPER activation attenuated the TNF-α-induced upregulation of proinflammatory proteins, which was reversed by G-15 [183]. However, in an estrogen-deficient environment, GPER-mediated changes in the TRPC1/ERK1/2 pathway may cause an increased inflammatory response, which is associated with the development of atherosclerosis in postmenopausal women [184]. In cultured rat aortic smooth muscle cells, the up-regulation of GPER expression and down-regulation of ERα both reduced medial vascular hypertrophy and inflammatory processes after carotid artery ligation [185]. The selective GPR30 agonist G-1 suppressed the TNF-induced upregulation of pro-inflammatory proteins, which was completely abrogated by the selective GPR30 antagonist G-15 [183]. Remarkably, estrogen treatment had no effect on the TNF-treated endothelium, suggesting that the activation of the classical ER may block the anti-inflammatory effects of GPER [183]. The selective ligands of GPER reduced the C-reactive protein (CRP)-mediated expression of inflammatory mediators in primary cultured bone-marrow-derived macrophages (BMM) and VSMCs [186]. In addition, the anti-inflammatory effects of GPER have been confirmed by several systemic disease models. In multiple sclerosis models, a loss of GPER exacerbates the inflammatory response, which is reduced by treatment with G1 [187]. In autoimmune encephalomyelitis, estradiol treatment reduced the inflammatory response in ERα knockout mice but not in GPER knockout mice [188]. In a hepatocyte mitochondrial injury model, GPER activated the AMPK–Nrf2 signaling pathway to alleviate the oxidative stress and the inflammatory response [85]. Indeed, the immunomodulatory function of GPER would be expected for the reason that GPER was cloned from a human B cell lymphoblast cell line cDNA library [15], and is expressed and functions in leukocytes [189] and macrophages [190].
It is well accepted that the maintenance of the integrity and physiological function of endothelial cells is vital for vascular function, the dysregulation of which is the initial step for the development of atherosclerosis [191]. In a rabbit model of atherosclerosis induced by a high-fat diet and compression balloon, GPER reduced arterial lipid levels including total cholesterol (TC), low-density lipoprotein (LDL), and triglycerides (TG), and increased high-density lipoprotein (HDL) through inhibiting endothelial apoptosis and endothelial dysfunction via the PI3K/Akt pathway [192]. In ovarian ApoE-deficient mice, GPER attenuated endothelial cell apoptosis, which reduced plasma lipid levels and prevented atherosclerotic plaque formation [193]. The kaempferol-induced upregulation of GPER attenuated atherosclerosis by inhibiting inflammation and apoptosis through the PI3K/AKT/Nrf2 pathway [194]. The inhibition of the proliferation and migration of VSMCs is one of the key events in the prevention of early atherosclerosis [195,196]. G-1 reduced serum-stimulated proliferation of human vascular smooth muscle cells, which was abrogated by GPER deletion [58]. In cultured rat aortic smooth muscle cells, both estradiol and G1 inhibited the cell proliferation and migration, which was blocked by the GPER antagonist G15 [185]. Estrogen prevented the apoptosis of VSMCs, thereby inhibiting the proliferation of human cerebral VSMCs by up-regulating GPER and ERs and down-regulating caspase-3, myocardin (MYOCD), and serum reaction factor (SRF) [197]. In addition, the activation of GPER was shown to inhibit vascular endothelial growth factor (VEGF) and tumor angiogenesis in triple-negative breast cancer (TNBC) [198]. Through inhibiting VSMC proliferation and migration, as well as its anti-apoptosis effects, GPER contributes to the maintenance of vascular homeostasis. Given that the activation of GPER improves lipid metabolism, regulates endothelial and vascular smooth muscle cell function, and suppresses inflammation, it is a promising therapeutic target for the treatment of atherosclerosis. The research on the role of GPER in the vascular pathology is summarized in Table 2.

4. Conclusions and Future Perspectives

As an estrogen receptor, GPER is widely distributed throughout the body, and its beneficial effects have been demonstrated in cardiovascular diseases. The binding of GPER to aldosterone has been demonstrated, but the exact mechanism remains to be elucidated. In addition, many natural and synthetic molecules are also able to activate or inhibit GPER, but further studies are needed to determine how they affect disease and health through GPER. In blood vessels, GPER is indispensable in the physiological functions of endothelial cells and smooth muscle cells. Given that the incidence of CVD is lower in premenopausal women than in age-matched men, it is likely that estrogen plays a crucial role in this disparity [6]. However, due to the complexity of the physiological action of estrogen itself, estrogen imbalance can cause many diseases [199,200]. Moreover, due to the complexity of estrogen receptors, the same estrogen acting on different estrogen receptors can have different effects, some of which are beneficial and others of which are harmful [201]. Clinically, hormone replacement therapy (HRT) is a common treatment option [202], but clinical studies have shown that postmenopausal estrogen therapy increases the risk of stroke [203,204], myocardial infarction, and deep vein thrombosis (DVT) and/or pulmonary embolism (PE) [205]. Therefore, highly selective estrogen receptor agonists and antagonists are promising candidates for drugs. Furthermore, since selective GPER agonists, unlike natural estrogens [206], do not have uterine trophic effects [112,207], postmenopausal women may also receive GPER-targeted therapy to provide vascular protection.

Author Contributions

F.X.: Writing—original draft. J.M.: Writing—original draft. X.W. (Xiaowu Wang): Writing—review and editing. X.W. (Xiaoya Wang): Writing—review and editing. W.F.: Writing—review and editing. J.S.: Writing—review and editing. Z.L.: Conceptualization, Writing—review and editing, Supervision. J.L.: Conceptualization, Writing—review and editing, Supervision, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by grants from the Shaanxi Province Science and Technology Resources Open and Shared Platform (2023-CX-PT-06), the Key R&D Program of Shaanxi Province (2022ZDLSF01-09), and the National Natural Science Foundation of China (82070264, 82241204, and 82270420).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Abbreviations

GPER: G protein-coupled estrogen receptor; CVD, cardiovascular disease; GPCRs, G protein-coupled receptors; ERK, extracellular signal-regulated kinase; ERα, estrogen receptor alpha; ERβ, estrogen receptor beta; VSMC, vascular smooth muscle cell; EC, endothelial cell; PTX, pertussis toxin; cAMP: cyclic adenosine monophosphate; Src, tyrosine kinase; KO, knockout, NLRP3, NOD-like receptor thermal protein domain associated protein 3; MMP, matrix metalloproteinases; HB-EGF, heparin-binding epidermal growth factor; EGFR, epidermal growth factor receptor; MEK, mitogen-activated protein kinase kinase; MAPK, mitogen-activated protein kinase; PI3K, phosphatidylinositol 3-kinase; Akt, serine/threonine kinase; AMPK, AMP-activated protein kinase; AC, adenylyl cyclase; PKA, protein kinase A; MRs, mineralocorticoid receptors; NO, nitric oxide; eNOS, endothelial nitric oxide synthase; ROS, reactive oxygen species; NBC, sodium/bicarbonate cotransporter; CTGF, tubule–glomerular feedback; CNT, connecting tubule; DCT, distal tricyclic tubule; NCC, thiazide-sensitive NaCl cotransporter; HTN, hypertension; SBP, systolic blood pressure; DBP, diastolic blood pressure; TRPC, transient receptor potential channel; CaMKKβ, calmodulin-dependent protein kinase 2; CaMKIIα, calmodulin-dependent protein kinase II; HDAC5, histone deacetylase 5; KLF2, Kruppel-like factor 2; MEF2C, muscle cell enhancer 2 peptide C; MLCP, myosin light chain phosphatase; BK (Ca), large-conductance calcium-activated potassium; Epac/Rap1: exchange protein directly activated by camp 1/ras-related protein 1; RhoA/Rho: ras homolog gene family, member A/ras homolog gene family; PAH, pulmonary arterial hypertension; LDL, low-density lipoprotein; HDL, high-density lipoprotein; CRP, C-reactive protein; BMM, bone marrow-derived macrophages; TC, total cholesterol; TG, triglyceride; VEGF, vascular endothelial growth factor; TNBC, triple-negative breast cancer; MYOCD, myocardin; SRF, serum reaction factor; H2O2, hydrogen peroxide; RV, right ventricle; LV, left ventricle; OVX: ovariectomy; MCT: monocrotaline; AS: atherosclerosis; ApoE: apolipoprotein E-deficient; HRT, hormone replacement therapy; DVT, deep vein thrombosis; PE, pulmonary embolism.

References

  1. Tsao, C.W.; Aday, A.W.; Almarzooq, Z.I.; Anderson, C.A.M.; Arora, P.; Avery, C.L.; Baker-Smith, C.M.; Beaton, A.Z.; Boehme, A.K.; Buxton, A.E.; et al. Heart Disease and Stroke Statistics-2023 Update: A Report From the American Heart Association. Circulation 2023, 147, e93–e621. [Google Scholar] [CrossRef] [PubMed]
  2. Colafella, K.M.M.; Denton, K.M. Sex-specific differences in hypertension and associated cardiovascular disease. Nat. Rev. Nephrol. 2018, 14, 185–201. [Google Scholar] [CrossRef]
  3. Rexrode, K.M.; Madsen, T.E.; Yu, A.Y.X.; Carcel, C.; Lichtman, J.H.; Miller, E.C. The Impact of Sex and Gender on Stroke. Circ. Res. 2022, 130, 512–528. [Google Scholar] [CrossRef]
  4. Yoon, S.S.; Carroll, M.D.; Fryar, C.D. Hypertension Prevalence and Control among Adults: United States, 2011–2014; NCHS Data Brief; US Department of Health and Human Services: Hyattsville, MD, USA, 2015; pp. 1–8.
  5. Mehta, L.S.; Beckie, T.M.; DeVon, H.A.; Grines, C.L.; Krumholz, H.M.; Johnson, M.N.; Lindley, K.J.; Vaccarino, V.; Wang, T.Y.; Watson, K.E.; et al. Acute Myocardial Infarction in Women: A Scientific Statement From the American Heart Association. Circulation 2016, 133, 916–947. [Google Scholar] [CrossRef] [PubMed]
  6. Benjamin, E.J.; Muntner, P.; Alonso, A.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Chang, A.R.; Cheng, S.; Das, S.R.; et al. Heart Disease and Stroke Statistics-2019 Update: A Report From the American Heart Association. Circulation 2019, 139, e56–e528. [Google Scholar] [CrossRef] [PubMed]
  7. Iorga, A.; Cunningham, C.M.; Moazeni, S.; Ruffenach, G.; Umar, S.; Eghbali, M. The protective role of estrogen and estrogen receptors in cardiovascular disease and the controversial use of estrogen therapy. Biol. Sex. Differ. 2017, 8, 33. [Google Scholar] [CrossRef]
  8. Farhat, M.Y.; Lavigne, M.C.; Ramwell, P.W. The vascular protective effects of estrogen. FASEB J. 1996, 10, 615–624. [Google Scholar] [CrossRef]
  9. Li, H.Y.; Bian, J.S.; Kwan, Y.W.; Wong, T.M. Enhanced responses to 17beta-estradiol in rat hearts treated with isoproterenol: Involvement of a cyclic AMP-dependent pathway. J. Pharmacol. Exp. Ther. 2000, 293, 592–598. [Google Scholar]
  10. Ullrich, N.D.; Krust, A.; Collins, P.; MacLeod, K.T. Genomic deletion of estrogen receptors ERalpha and ERbeta does not alter estrogen-mediated inhibition of Ca2+ influx and contraction in murine cardiomyocytes. Am. J. Physiol. Heart Circ. Physiol. 2008, 294, H2421–H2427. [Google Scholar] [CrossRef]
  11. Sharma, G.; Hu, C.; Staquicini, D.I.; Brigman, J.L.; Liu, M.; Mauvais-Jarvis, F.; Pasqualini, R.; Arap, W.; Arterburn, J.B.; Hathaway, H.J.; et al. Preclinical efficacy of the GPER-selective agonist G-1 in mouse models of obesity and diabetes. Sci. Transl. Med. 2020, 12, eaau5956. [Google Scholar] [CrossRef]
  12. Xu, F.; Wang, X.; Wu, N.; He, S.; Yi, W.; Xiang, S.; Zhang, P.; Xie, X.; Ying, C. Bisphenol A induces proliferative effects on both breast cancer cells and vascular endothelial cells through a shared GPER-dependent pathway in hypoxia. Environ. Pollut. 2017, 231, 1609–1620. [Google Scholar] [CrossRef]
  13. Fredette, N.C.; Meyer, M.R.; Prossnitz, E.R. Role of GPER in estrogen-dependent nitric oxide formation and vasodilation. J. Steroid Biochem. Mol. Biol. 2018, 176, 65–72. [Google Scholar] [CrossRef] [PubMed]
  14. Lindsey, S.H.; Cohen, J.A.; Brosnihan, K.B.; Gallagher, P.E.; Chappell, M.C. Chronic treatment with the G protein-coupled receptor 30 agonist G-1 decreases blood pressure in ovariectomized mRen2.Lewis rats. Endocrinology 2009, 150, 3753–3758. [Google Scholar] [CrossRef] [PubMed]
  15. Prossnitz, E.R.; Arterburn, J.B. International Union of Basic and Clinical Pharmacology. XCVII. G Protein-Coupled Estrogen Receptor and Its Pharmacologic Modulators. Pharmacol. Rev. 2015, 67, 505–540. [Google Scholar] [CrossRef] [PubMed]
  16. Filardo, E.J.; Quinn, J.A.; Bland, K.I.; Frackelton, A.R., Jr. Estrogen-induced activation of Erk-1 and Erk-2 requires the G protein-coupled receptor homolog, GPR30, and occurs via trans-activation of the epidermal growth factor receptor through release of HB-EGF. Mol. Endocrinol. 2000, 14, 1649–1660. [Google Scholar] [CrossRef]
  17. Thomas, P.; Pang, Y.; Filardo, E.J.; Dong, J. Identity of an estrogen membrane receptor coupled to a G protein in human breast cancer cells. Endocrinology 2005, 146, 624–632. [Google Scholar] [CrossRef]
  18. Revankar, C.M.; Cimino, D.F.; Sklar, L.A.; Arterburn, J.B.; Prossnitz, E.R. A transmembrane intracellular estrogen receptor mediates rapid cell signaling. Science 2005, 307, 1625–1630. [Google Scholar] [CrossRef]
  19. Alexander, S.P.; Mathie, A.; Peters, J.A. Guide to Receptors and Channels (GRAC), 5th edition. Br. J. Pharmacol. 2011, 164 (Suppl. 1), S1–S324. [Google Scholar] [CrossRef]
  20. Molina, L.; Figueroa, C.D.; Bhoola, K.D.; Ehrenfeld, P. GPER-1/GPR30 a novel estrogen receptor sited in the cell membrane: Therapeutic coupling to breast cancer. Expert. Opin. Ther. Targets 2017, 21, 755–766. [Google Scholar] [CrossRef]
  21. Xu, X.L.; Huang, Z.Y.; Yu, K.; Li, J.; Fu, X.W.; Deng, S.L. Estrogen Biosynthesis and Signal Transduction in Ovarian Disease. Front. Endocrinol. 2022, 13, 827032. [Google Scholar] [CrossRef]
  22. Plante, B.J.; Lessey, B.A.; Taylor, R.N.; Wang, W.; Bagchi, M.K.; Yuan, L.; Scotchie, J.; Fritz, M.A.; Young, S.L. G protein-coupled estrogen receptor (GPER) expression in normal and abnormal endometrium. Reprod. Sci. 2012, 19, 684–693. [Google Scholar] [CrossRef] [PubMed]
  23. Yu, K.; Huang, Z.Y.; Xu, X.L.; Li, J.; Fu, X.W.; Deng, S.L. Estrogen Receptor Function: Impact on the Human Endometrium. Front. Endocrinol. 2022, 13, 827724. [Google Scholar] [CrossRef] [PubMed]
  24. Chimento, A.; De Luca, A.; Nocito, M.C.; Avena, P.; La Padula, D.; Zavaglia, L.; Pezzi, V. Role of GPER-Mediated Signaling in Testicular Functions and Tumorigenesis. Cells 2020, 9, 2115. [Google Scholar] [CrossRef] [PubMed]
  25. Rago, V.; Romeo, F.; Giordano, F.; Ferraro, A.; Carpino, A. Identification of the G protein-coupled estrogen receptor (GPER) in human prostate: Expression site of the estrogen receptor in the benign and neoplastic gland. Andrology 2016, 4, 121–127. [Google Scholar] [CrossRef] [PubMed]
  26. Ramirez-de-Arellano, A.; Pereira-Suarez, A.L.; Rico-Fuentes, C.; Lopez-Pulido, E.I.; Villegas-Pineda, J.C.; Sierra-Diaz, E. Distribution and Effects of Estrogen Receptors in Prostate Cancer: Associated Molecular Mechanisms. Front. Endocrinol. 2021, 12, 811578. [Google Scholar] [CrossRef] [PubMed]
  27. Kolkova, Z.; Noskova, V.; Ehinger, A.; Hansson, S.; Casslen, B. G protein-coupled estrogen receptor 1 (GPER, GPR 30) in normal human endometrium and early pregnancy decidua. Mol. Hum. Reprod. 2010, 16, 743–751. [Google Scholar] [CrossRef]
  28. Harding, A.T.; Goff, M.A.; Froggatt, H.M.; Lim, J.K.; Heaton, N.S. GPER1 is required to protect fetal health from maternal inflammation. Science 2021, 371, 271–276. [Google Scholar] [CrossRef]
  29. Hugo, E.R.; Brandebourg, T.D.; Woo, J.G.; Loftus, J.; Alexander, J.W.; Ben-Jonathan, N. Bisphenol A at environmentally relevant doses inhibits adiponectin release from human adipose tissue explants and adipocytes. Environ. Health Perspect. 2008, 116, 1642–1647. [Google Scholar] [CrossRef]
  30. Luo, J.; Liu, D. Does GPER Really Function as a G Protein-Coupled Estrogen Receptor in vivo? Front. Endocrinol. 2020, 11, 148. [Google Scholar] [CrossRef]
  31. Barton, M. Position paper: The membrane estrogen receptor GPER–Clues and questions. Steroids 2012, 77, 935–942. [Google Scholar] [CrossRef]
  32. Meyer, M.R.; Fredette, N.C.; Howard, T.A.; Hu, C.; Ramesh, C.; Daniel, C.; Amann, K.; Arterburn, J.B.; Barton, M.; Prossnitz, E.R. G protein-coupled estrogen receptor protects from atherosclerosis. Sci. Rep. 2014, 4, 7564. [Google Scholar] [CrossRef] [PubMed]
  33. Haas, E.; Meyer, M.R.; Schurr, U.; Bhattacharya, I.; Minotti, R.; Nguyen, H.H.; Heigl, A.; Lachat, M.; Genoni, M.; Barton, M. Differential effects of 17beta-estradiol on function and expression of estrogen receptor alpha, estrogen receptor beta, and GPR30 in arteries and veins of patients with atherosclerosis. Hypertension 2007, 49, 1358–1363. [Google Scholar] [CrossRef] [PubMed]
  34. Kockelkoren, G.; Lauritsen, L.; Shuttle, C.G.; Kazepidou, E.; Vonkova, I.; Zhang, Y.; Breuer, A.; Kennard, C.; Brunetti, R.M.; D’Este, E.; et al. Molecular mechanism of GPCR spatial organization at the plasma membrane. Nat. Chem. Biol. 2023, 1–9. [Google Scholar] [CrossRef]
  35. Kelly, M.J.; Levin, E.R. Rapid actions of plasma membrane estrogen receptors. Trends Endocrinol. Metab. 2001, 12, 152–156. [Google Scholar] [CrossRef]
  36. Funakoshi, T.; Yanai, A.; Shinoda, K.; Kawano, M.M.; Mizukami, Y. G protein-coupled receptor 30 is an estrogen receptor in the plasma membrane. Biochem. Biophys. Res. Commun. 2006, 346, 904–910. [Google Scholar] [CrossRef] [PubMed]
  37. Filardo, E.; Quinn, J.; Pang, Y.; Graeber, C.; Shaw, S.; Dong, J.; Thomas, P. Activation of the novel estrogen receptor G protein-coupled receptor 30 (GPR30) at the plasma membrane. Endocrinology 2007, 148, 3236–3245. [Google Scholar] [CrossRef]
  38. Maiti, K.; Paul, J.W.; Read, M.; Chan, E.C.; Riley, S.C.; Nahar, P.; Smith, R. G-1-activated membrane estrogen receptors mediate increased contractility of the human myometrium. Endocrinology 2011, 152, 2448–2455. [Google Scholar] [CrossRef]
  39. Lindsey, S.H.; Yamaleyeva, L.M.; Brosnihan, K.B.; Gallagher, P.E.; Chappell, M.C. Estrogen receptor GPR30 reduces oxidative stress and proteinuria in the salt-sensitive female mRen2.Lewis rat. Hypertension 2011, 58, 665–671. [Google Scholar] [CrossRef]
  40. Akama, K.T.; Thompson, L.I.; Milner, T.A.; McEwen, B.S. Post-synaptic density-95 (PSD-95) binding capacity of G-protein-coupled receptor 30 (GPR30), an estrogen receptor that can be identified in hippocampal dendritic spines. J. Biol. Chem. 2013, 288, 6438–6450. [Google Scholar] [CrossRef]
  41. Sakamoto, H.; Matsuda, K.; Hosokawa, K.; Nishi, M.; Morris, J.F.; Prossnitz, E.R.; Kawata, M. Expression of G protein-coupled receptor-30, a G protein-coupled membrane estrogen receptor, in oxytocin neurons of the rat paraventricular and supraoptic nuclei. Endocrinology 2007, 148, 5842–5850. [Google Scholar] [CrossRef]
  42. Otto, C.; Rohde-Schulz, B.; Schwarz, G.; Fuchs, I.; Klewer, M.; Brittain, D.; Langer, G.; Bader, B.; Prelle, K.; Nubbemeyer, R.; et al. G protein-coupled receptor 30 localizes to the endoplasmic reticulum and is not activated by estradiol. Endocrinology 2008, 149, 4846–4856. [Google Scholar] [CrossRef]
  43. Revankar, C.M.; Mitchell, H.D.; Field, A.S.; Burai, R.; Corona, C.; Ramesh, C.; Sklar, L.A.; Arterburn, J.B.; Prossnitz, E.R. Synthetic estrogen derivatives demonstrate the functionality of intracellular GPR30. ACS Chem. Biol. 2007, 2, 536–544. [Google Scholar] [CrossRef]
  44. Deliu, E.; Brailoiu, G.C.; Arterburn, J.B.; Oprea, T.I.; Benamar, K.; Dun, N.J.; Brailoiu, E. Mechanisms of G protein-coupled estrogen receptor-mediated spinal nociception. J. Pain. 2012, 13, 742–754. [Google Scholar] [CrossRef]
  45. Warfvinge, K.; Krause, D.N.; Maddahi, A.; Edvinsson, J.C.A.; Edvinsson, L.; Haanes, K.A. Estrogen receptors alpha, beta and GPER in the CNS and trigeminal system–molecular and functional aspects. J. Headache Pain. 2020, 21, 131. [Google Scholar] [CrossRef]
  46. Pena-Gutierrez, K.M.; Hernandez-Ortega, K.; Bello-Alvarez, C.; Camacho-Arroyo, I. Expression and estrogen regulation of G protein-coupled estrogen receptor in human glioblastoma cells. Oncol. Lett. 2022, 24, 397. [Google Scholar] [CrossRef]
  47. Cheng, S.B.; Graeber, C.T.; Quinn, J.A.; Filardo, E.J. Retrograde transport of the transmembrane estrogen receptor, G-protein-coupled-receptor-30 (GPR30/GPER) from the plasma membrane towards the nucleus. Steroids 2011, 76, 892–896. [Google Scholar] [CrossRef]
  48. Ramirez-Lopez, I.G.; Ramirez de Arellano, A.; Jave-Suarez, L.F.; Hernandez-Silva, C.D.; Garcia-Chagollan, M.; Hernandez-Bello, J.; Lopez-Pulido, E.I.; Macias-Barragan, J.; Montoya-Buelna, M.; Munoz-Valle, J.F.; et al. Interaction between 17beta-estradiol, prolactin and human papillomavirus induce E6/E7 transcript and modulate the expression and localization of hormonal receptors. Cancer Cell Int. 2019, 19, 227. [Google Scholar] [CrossRef]
  49. Caron, K.M.; James, L.R.; Kim, H.S.; Morham, S.G.; Sequeira Lopez, M.L.; Gomez, R.A.; Reudelhuber, T.L.; Smithies, O. A genetically clamped renin transgene for the induction of hypertension. Proc. Natl. Acad. Sci. USA 2002, 99, 8248–8252. [Google Scholar] [CrossRef]
  50. Caron, K.M.; James, L.R.; Kim, H.S.; Knowles, J.; Uhlir, R.; Mao, L.; Hagaman, J.R.; Cascio, W.; Rockman, H.; Smithies, O. Cardiac hypertrophy and sudden death in mice with a genetically clamped renin transgene. Proc. Natl. Acad. Sci. USA 2004, 101, 3106–3111. [Google Scholar] [CrossRef]
  51. Caron, K.M.; James, L.R.; Lee, G.; Kim, H.S.; Smithies, O. Lifelong genetic minipumps. Physiol. Genom. 2005, 20, 203–209. [Google Scholar] [CrossRef]
  52. Lenhart, P.M.; Broselid, S.; Barrick, C.J.; Leeb-Lundberg, L.M.; Caron, K.M. G-protein-coupled receptor 30 interacts with receptor activity-modifying protein 3 and confers sex-dependent cardioprotection. J. Mol. Endocrinol. 2013, 51, 191–202. [Google Scholar] [CrossRef]
  53. Cheng, S.B.; Quinn, J.A.; Graeber, C.T.; Filardo, E.J. Down-modulation of the G-protein-coupled estrogen receptor, GPER, from the cell surface occurs via a trans-Golgi-proteasome pathway. J. Biol. Chem. 2011, 286, 22441–22455. [Google Scholar] [CrossRef]
  54. Lappano, R.; Rosano, C.; De Marco, P.; De Francesco, E.M.; Pezzi, V.; Maggiolini, M. Estriol acts as a GPR30 antagonist in estrogen receptor-negative breast cancer cells. Mol. Cell. Endocrinol. 2010, 320, 162–170. [Google Scholar] [CrossRef]
  55. Hess, R.A. Estrogen in the adult male reproductive tract: A review. Reprod. Biol. Endocrinol. 2003, 1, 52. [Google Scholar] [CrossRef]
  56. Prossnitz, E.R.; Barton, M. The G-protein-coupled estrogen receptor GPER in health and disease. Nat. Rev. Endocrinol. 2011, 7, 715–726. [Google Scholar] [CrossRef]
  57. Martensson, U.E.; Salehi, S.A.; Windahl, S.; Gomez, M.F.; Sward, K.; Daszkiewicz-Nilsson, J.; Wendt, A.; Andersson, N.; Hellstrand, P.; Grande, P.O.; et al. Deletion of the G protein-coupled receptor 30 impairs glucose tolerance, reduces bone growth, increases blood pressure, and eliminates estradiol-stimulated insulin release in female mice. Endocrinology 2009, 150, 687–698. [Google Scholar] [CrossRef]
  58. Haas, E.; Bhattacharya, I.; Brailoiu, E.; Damjanovic, M.; Brailoiu, G.C.; Gao, X.; Mueller-Guerre, L.; Marjon, N.A.; Gut, A.; Minotti, R.; et al. Regulatory role of G protein-coupled estrogen receptor for vascular function and obesity. Circ. Res. 2009, 104, 288–291. [Google Scholar] [CrossRef]
  59. Meyer, M.R.; Amann, K.; Field, A.S.; Hu, C.; Hathaway, H.J.; Kanagy, N.L.; Walker, M.K.; Barton, M.; Prossnitz, E.R. Deletion of G protein-coupled estrogen receptor increases endothelial vasoconstriction. Hypertension 2012, 59, 507–512. [Google Scholar] [CrossRef]
  60. Delbeck, M.; Golz, S.; Vonk, R.; Janssen, W.; Hucho, T.; Isensee, J.; Schafer, S.; Otto, C. Impaired left-ventricular cardiac function in male GPR30-deficient mice. Mol. Med. Rep. 2011, 4, 37–40. [Google Scholar] [CrossRef]
  61. Meoli, L.; Isensee, J.; Zazzu, V.; Nabzdyk, C.S.; Soewarto, D.; Witt, H.; Foryst-Ludwig, A.; Kintscher, U.; Noppinger, P.R. Sex- and age-dependent effects of Gpr30 genetic deletion on the metabolic and cardiovascular profiles of diet-induced obese mice. Gene 2014, 540, 210–216. [Google Scholar] [CrossRef]
  62. Wang, H.; Sun, X.; Chou, J.; Lin, M.; Ferrario, C.M.; Zapata-Sudo, G.; Groban, L. Cardiomyocyte-specific deletion of the G protein-coupled estrogen receptor (GPER) leads to left ventricular dysfunction and adverse remodeling: A sex-specific gene profiling analysis. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 1870–1882. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, H.; Sun, X.; Lin, M.S.; Ferrario, C.M.; Van Remmen, H.; Groban, L. G protein-coupled estrogen receptor (GPER) deficiency induces cardiac remodeling through oxidative stress. Transl. Res. 2018, 199, 39–51. [Google Scholar] [CrossRef]
  64. Wang, H.; Sun, X.; Hodge, H.S.; Ferrario, C.M.; Groban, L. NLRP3 inhibition improves heart function in GPER knockout mice. Biochem. Biophys. Res. Commun. 2019, 514, 998–1003. [Google Scholar] [CrossRef] [PubMed]
  65. Filardo, E.J.; Quinn, J.A.; Frackelton, A.R., Jr.; Bland, K.I. Estrogen action via the G protein-coupled receptor, GPR30: Stimulation of adenylyl cyclase and cAMP-mediated attenuation of the epidermal growth factor receptor-to-MAPK signaling axis. Mol. Endocrinol. 2002, 16, 70–84. [Google Scholar] [CrossRef]
  66. Lindsey, S.H.; Liu, L.; Chappell, M.C. Vasodilation by GPER in mesenteric arteries involves both endothelial nitric oxide and smooth muscle cAMP signaling. Steroids 2014, 81, 99–102. [Google Scholar] [CrossRef] [PubMed]
  67. Barton, M.; Prossnitz, E.R. Emerging roles of GPER in diabetes and atherosclerosis. Trends Endocrinol. Metab. 2015, 26, 185–192. [Google Scholar] [CrossRef] [PubMed]
  68. Park, J.S.; Lee, G.H.; Jin, S.W.; Pham, T.H.; Thai, T.N.; Kim, J.Y.; Kim, C.Y.; Han, E.H.; Hwang, Y.P.; Choi, C.Y.; et al. G protein-coupled estrogen receptor regulates the KLF2-dependent eNOS expression by activating of Ca2+ and EGFR signaling pathway in human endothelial cells. Biochem. Pharmacol. 2021, 192, 114721. [Google Scholar] [CrossRef] [PubMed]
  69. Gao, F.; Huang, Y.; Zhang, L.; Liu, W. Involvement of estrogen receptor and GPER in bisphenol A induced proliferation of vascular smooth muscle cells. Toxicol. Vitr. 2019, 56, 156–162. [Google Scholar] [CrossRef]
  70. Kokai, D.; Stanic, B.; Tesic, B.; Samardzija Nenadov, D.; Pogrmic-Majkic, K.; Fa Nedeljkovic, S.; Andric, N. Dibutyl phthalate promotes angiogenesis in EA.hy926 cells through estrogen receptor-dependent activation of ERK1/2, PI3K-Akt, and NO signaling pathways. Chem. Biol. Interact. 2022, 366, 110174. [Google Scholar] [CrossRef] [PubMed]
  71. Deschamps, A.M.; Murphy, E. Activation of a novel estrogen receptor, GPER, is cardioprotective in male and female rats. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, H1806–H1813. [Google Scholar] [CrossRef]
  72. Feng, Y.; Madungwe, N.B.; da Cruz Junho, C.V.; Bopassa, J.C. Activation of G protein-coupled oestrogen receptor 1 at the onset of reperfusion protects the myocardium against ischemia/reperfusion injury by reducing mitochondrial dysfunction and mitophagy. Br. J. Pharmacol. 2017, 174, 4329–4344. [Google Scholar] [CrossRef] [PubMed]
  73. Abo-Zaid, O.A.; Moawed, F.S.; Hassan, H.A.; Moustafa, E.M. Bisphenol-A/Radiation mediated inflammatory response activates EGFR/KRAS/ERK1/2 signaling pathway leads to lung carcinogenesis incidence. Int. J. Immunopathol. Pharmacol. 2022, 36, 3946320221092918. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, L.; Liu, S.; Luo, H.; Chen, C.; Zhang, X.; He, L.; Tu, G. GPR30-mediated HMGB1 upregulation in CAFs induces autophagy and tamoxifen resistance in ERalpha-positive breast cancer cells. Aging 2021, 13, 16178–16197. [Google Scholar] [CrossRef]
  75. Yin, J.; Tu, G.; Peng, M.; Zeng, H.; Wan, X.; Qiao, Y.; Qin, Y.; Liu, M.; Luo, H. GPER-regulated lncRNA-Glu promotes glutamate secretion to enhance cellular invasion and metastasis in triple-negative breast cancer. FASEB J. 2020, 34, 4557–4572. [Google Scholar] [CrossRef]
  76. Li, M.; Guo, J.; Gao, W.; Yu, J.; Han, X.; Zhang, J.; Shao, B. Bisphenol AF-induced endogenous transcription is mediated by ERalpha and ERK1/2 activation in human breast cancer cells. PLoS ONE 2014, 9, e94725. [Google Scholar] [CrossRef]
  77. Li, F.; Yu, X.; Szynkarski, C.K.; Meng, C.; Zhou, B.; Barhoumi, R.; White, R.E.; Heaps, C.L.; Stallone, J.N.; Han, G. Activation of GPER Induces Differentiation and Inhibition of Coronary Artery Smooth Muscle Cell Proliferation. PLoS ONE 2013, 8, e64771. [Google Scholar] [CrossRef]
  78. Upadhayay, S.; Gupta, R.; Singh, S.; Mundkar, M.; Singh, G.; Kumar, P. Involvement of the G-Protein-Coupled Estrogen Receptor-1 (GPER) Signaling Pathway in Neurodegenerative Disorders: A Review. Cell Mol. Neurobiol. 2022, 43, 1833–1847. [Google Scholar] [CrossRef]
  79. Wang, X.W.; Yuan, L.J.; Yang, Y.; Zhang, M.; Chen, W.F. IGF-1 inhibits MPTP/MPP+-induced autophagy on dopaminergic neurons through the IGF-1R/PI3K-Akt-mTOR pathway and GPER. Am. J. Physiol. Endocrinol. Metab. 2020, 319, E734–E743. [Google Scholar] [CrossRef]
  80. Ariyani, W.; Miyazaki, W.; Amano, I.; Hanamura, K.; Shirao, T.; Koibuchi, N. Soy Isoflavones Accelerate Glial Cell Migration via GPER-Mediated Signal Transduction Pathway. Front. Endocrinol. 2020, 11, 554941. [Google Scholar] [CrossRef] [PubMed]
  81. Zhang, X.L.; Liu, N.; Weng, S.F.; Wang, H.S. Bisphenol A Increases the Migration and Invasion of Triple-Negative Breast Cancer Cells via Oestrogen-related Receptor Gamma. Basic Clin. Pharmacol. Toxicol. 2016, 119, 389–395. [Google Scholar] [CrossRef] [PubMed]
  82. Yu, X.; Li, F.; Klussmann, E.; Stallone, J.N.; Han, G. G protein-coupled estrogen receptor 1 mediates relaxation of coronary arteries via cAMP/PKA-dependent activation of MLCP. Am. J. Physiol. Endocrinol. Metab. 2014, 307, E398–E407. [Google Scholar] [CrossRef] [PubMed]
  83. Yu, X.; Zhang, Q.; Zhao, Y.; Schwarz, B.J.; Stallone, J.N.; Heaps, C.L.; Han, G. Activation of G protein-coupled estrogen receptor 1 induces coronary artery relaxation via Epac/Rap1-mediated inhibition of RhoA/Rho kinase pathway in parallel with PKA. PLoS ONE 2017, 12, e0173085. [Google Scholar] [CrossRef]
  84. Lau, K.M.; Ma, F.M.; Xia, J.T.; Chan, Q.K.Y.; Ng, C.F.; To, K.F. Activation of GPR30 stimulates GTP-binding of Galphai1 protein to sustain activation of Erk1/2 in inhibition of prostate cancer cell growth and modulates metastatic properties. Exp. Cell Res. 2017, 350, 199–209. [Google Scholar] [CrossRef] [PubMed]
  85. Yao, Y.; Wang, H.; Yang, Y.; Jiang, Z.; Ma, H. Dehydroepiandrosterone protects against oleic acid-triggered mitochondrial dysfunction to relieve oxidative stress and inflammation via activation of the AMPK-Nrf2 axis by targeting GPR30 in hepatocytes. Mol. Immunol. 2023, 155, 110–123. [Google Scholar] [CrossRef] [PubMed]
  86. Yu, Z.; Su, G.; Zhang, L.; Liu, G.; Zhou, Y.; Fang, S.; Zhang, Q.; Wang, T.; Huang, C.; Huang, Z.; et al. Icaritin inhibits neuroinflammation in a rat cerebral ischemia model by regulating microglial polarization through the GPER-ERK-NF-kappaB signaling pathway. Mol. Med. 2022, 28, 142. [Google Scholar] [CrossRef] [PubMed]
  87. Okamoto, M.; Suzuki, T.; Mizukami, Y.; Ikeda, T. The membrane-type estrogen receptor G-protein-coupled estrogen receptor suppresses lipopolysaccharide-induced interleukin 6 via inhibition of nuclear factor-kappa B pathway in murine macrophage cells. Anim. Sci. J. 2017, 88, 1870–1879. [Google Scholar] [CrossRef]
  88. Zhu, P.; Liao, L.Y.; Zhao, T.T.; Mo, X.M.; Chen, G.G.; Liu, Z.M. GPER/ERK&AKT/NF-kappaB pathway is involved in cadmium-induced proliferation, invasion and migration of GPER-positive thyroid cancer cells. Mol. Cell. Endocrinol. 2017, 442, 68–80. [Google Scholar] [CrossRef]
  89. Liang, S.; Chen, Z.; Jiang, G.; Zhou, Y.; Liu, Q.; Su, Q.; Wei, W.; Du, J.; Wang, H. Activation of GPER suppresses migration and angiogenesis of triple negative breast cancer via inhibition of NF-kappaB/IL-6 signals. Cancer Lett. 2017, 386, 12–23. [Google Scholar] [CrossRef] [PubMed]
  90. Zhu, G.; Huang, Y.; Wu, C.; Wei, D.; Shi, Y. Activation of G-Protein-Coupled Estrogen Receptor Inhibits the Migration of Human Nonsmall Cell Lung Cancer Cells via IKK-beta/NF-kappaB Signals. DNA Cell Biol. 2016, 35, 434–442. [Google Scholar] [CrossRef]
  91. Chen, Z.J.; Wei, W.; Jiang, G.M.; Liu, H.; Wei, W.D.; Yang, X.; Wu, Y.M.; Liu, H.; Wong, C.K.; Du, J.; et al. Activation of GPER suppresses epithelial mesenchymal transition of triple negative breast cancer cells via NF-kappaB signals. Mol. Oncol. 2016, 10, 775–788. [Google Scholar] [CrossRef] [PubMed]
  92. Funder, J.W. Aldosterone and mineralocorticoid receptors in the cardiovascular system. Prog. Cardiovasc. Dis. 2010, 52, 393–400. [Google Scholar] [CrossRef] [PubMed]
  93. Feldman, R.D.; Gros, R. Unraveling the mechanisms underlying the rapid vascular effects of steroids: Sorting out the receptors and the pathways. Br. J. Pharmacol. 2011, 163, 1163–1169. [Google Scholar] [CrossRef] [PubMed]
  94. Wendler, A.; Albrecht, C.; Wehling, M. Nongenomic actions of aldosterone and progesterone revisited. Steroids 2012, 77, 1002–1006. [Google Scholar] [CrossRef] [PubMed]
  95. Gros, R.; Ding, Q.; Sklar, L.A.; Prossnitz, E.E.; Arterburn, J.B.; Chorazyczewski, J.; Feldman, R.D. GPR30 expression is required for the mineralocorticoid receptor-independent rapid vascular effects of aldosterone. Hypertension 2011, 57, 442–451. [Google Scholar] [CrossRef] [PubMed]
  96. Gros, R.; Ding, Q.; Liu, B.; Chorazyczewski, J.; Feldman, R.D. Aldosterone mediates its rapid effects in vascular endothelial cells through GPER activation. Am. J. Physiol. Cell Physiol. 2013, 304, C532–C540. [Google Scholar] [CrossRef] [PubMed]
  97. Barton, M.; Meyer, M.R. Permissive Role of GPER for Arterial Hypertension. Hypertension 2019, 73, e9–e10. [Google Scholar] [CrossRef] [PubMed]
  98. Waghulde, H.; Galla, S.; Chakraborty, S.; Joe, B. Response to Permissive Role of GPER for Arterial Hypertension. Hypertension 2019, 73, e11. [Google Scholar] [CrossRef]
  99. Wendler, A.; Wehling, M. Is GPR30 the membrane aldosterone receptor postulated 20 years ago? Hypertension 2011, 57, e16, author reply e17. [Google Scholar] [CrossRef]
  100. Ding, Q.; Chorazyczewski, J.; Gros, R.; Motulsky, H.J.; Limbird, L.E.; Feldman, R.D. Correlation of functional and radioligand binding characteristics of GPER ligands confirming aldosterone as a GPER agonist. Pharmacol. Res. Perspect. 2022, 10, e00995. [Google Scholar] [CrossRef]
  101. Ashton, A.W.; Le, T.Y.; Gomez-Sanchez, C.E.; Morel-Kopp, M.C.; McWhinney, B.; Hudson, A.; Mihailidou, A.S. Role of Nongenomic Signaling Pathways Activated by Aldosterone During Cardiac Reperfusion Injury. Mol. Endocrinol. 2015, 29, 1144–1155. [Google Scholar] [CrossRef]
  102. De Giusti, V.C.; Orlowski, A.; Ciancio, M.C.; Espejo, M.S.; Gonano, L.A.; Caldiz, C.I.; Vila Petroff, M.G.; Villa-Abrille, M.C.; Aiello, E.A. Aldosterone stimulates the cardiac sodium/bicarbonate cotransporter via activation of the g protein-coupled receptor gpr30. J. Mol. Cell. Cardiol. 2015, 89, 260–267. [Google Scholar] [CrossRef] [PubMed]
  103. Orlowski, A.; De Giusti, V.C.; Ciancio, M.C.; Espejo, M.S.; Aiello, E.A. The cardiac electrogenic sodium/bicarbonate cotransporter (NBCe1) is activated by aldosterone through the G protein-coupled receptor 30 (GPR 30). Channels 2016, 10, 428–434. [Google Scholar] [CrossRef] [PubMed]
  104. Brailoiu, G.C.; Benamar, K.; Arterburn, J.B.; Gao, E.; Rabinowitz, J.E.; Koch, W.J.; Brailoiu, E. Aldosterone increases cardiac vagal tone via G protein-coupled oestrogen receptor activation. J. Physiol. 2013, 591, 4223–4235. [Google Scholar] [CrossRef] [PubMed]
  105. Ren, Y.; D’Ambrosio, M.A.; Garvin, J.L.; Leung, P.; Kutskill, K.; Wang, H.; Peterson, E.L.; Carretero, O.A. Aldosterone sensitizes connecting tubule glomerular feedback via the aldosterone receptor GPR30. Am. J. Physiol. Ren. Physiol. 2014, 307, F427–F434. [Google Scholar] [CrossRef]
  106. Cheng, L.; Poulsen, S.B.; Wu, Q.; Esteva-Font, C.; Olesen, E.T.B.; Peng, L.; Olde, B.; Leeb-Lundberg, L.M.F.; Pisitkun, T.; Rieg, T.; et al. Rapid Aldosterone-Mediated Signaling in the DCT Increases Activity of the Thiazide-Sensitive NaCl Cotransporter. J. Am. Soc. Nephrol. 2019, 30, 1454–1470. [Google Scholar] [CrossRef] [PubMed]
  107. Caroccia, B.; Seccia, T.M.; Piazza, M.; Prisco, S.; Zanin, S.; Iacobone, M.; Lenzini, L.; Pallafacchina, G.; Domening, O.; Poglitsch, M.; et al. Aldosterone Stimulates Its Biosynthesis Via a Novel GPER-Mediated Mechanism. J. Clin. Endocrinol. Metab. 2019, 104, 6316–6324. [Google Scholar] [CrossRef] [PubMed]
  108. Caroccia, B.; Seccia, T.M.; Campos, A.G.; Gioco, F.; Kuppusamy, M.; Ceolotto, G.; Guerzoni, E.; Simonato, F.; Mareso, S.; Lenzini, L.; et al. GPER-1 and estrogen receptor-beta ligands modulate aldosterone synthesis. Endocrinology 2014, 155, 4296–4304. [Google Scholar] [CrossRef]
  109. Rossi, G.P.; Caroccia, B.; Seccia, T.M. Role of estrogen receptors in modulating aldosterone biosynthesis and blood pressure. Steroids 2019, 152, 108486. [Google Scholar] [CrossRef] [PubMed]
  110. Kuiper, G.G.; Carlsson, B.; Grandien, K.; Enmark, E.; Haggblad, J.; Nilsson, S.; Gustafsson, J.A. Comparison of the ligand binding specificity and transcript tissue distribution of estrogen receptors alpha and beta. Endocrinology 1997, 138, 863–870. [Google Scholar] [CrossRef]
  111. Bologa, C.G.; Revankar, C.M.; Young, S.M.; Edwards, B.S.; Arterburn, J.B.; Kiselyov, A.S.; Parker, M.A.; Tkachenko, S.E.; Savchuck, N.P.; Sklar, L.A.; et al. Virtual and biomolecular screening converge on a selective agonist for GPR30. Nat. Chem. Biol. 2006, 2, 207–212. [Google Scholar] [CrossRef]
  112. Dennis, M.K.; Burai, R.; Ramesh, C.; Petrie, W.K.; Alcon, S.N.; Nayak, T.K.; Bologa, C.G.; Leitao, A.; Brailoiu, E.; Deliu, E.; et al. In vivo effects of a GPR30 antagonist. Nat. Chem. Biol. 2009, 5, 421–427. [Google Scholar] [CrossRef]
  113. Dennis, M.K.; Field, A.S.; Burai, R.; Ramesh, C.; Petrie, W.K.; Bologa, C.G.; Oprea, T.I.; Yamaguchi, Y.; Hayashi, S.; Sklar, L.A.; et al. Identification of a GPER/GPR30 antagonist with improved estrogen receptor counterselectivity. J. Steroid Biochem. Mol. Biol. 2011, 127, 358–366. [Google Scholar] [CrossRef]
  114. Prossnitz, E.R. GPER modulators: Opportunity Nox on the heels of a class Akt. J. Steroid Biochem. Mol. Biol. 2018, 176, 73–81. [Google Scholar] [CrossRef]
  115. Gui, Y.; Shi, Z.; Wang, Z.; Li, J.J.; Xu, C.; Tian, R.; Song, X.; Walsh, M.P.; Li, D.; Gao, J.; et al. The GPER agonist G-1 induces mitotic arrest and apoptosis in human vascular smooth muscle cells independent of GPER. J. Cell. Physiol. 2015, 230, 885–895. [Google Scholar] [CrossRef] [PubMed]
  116. Holm, A.; Grande, P.O.; Luduena, R.F.; Olde, B.; Prasad, V.; Leeb-Lundberg, L.M.; Nilsson, B.O. The G protein-coupled oestrogen receptor 1 agonist G-1 disrupts endothelial cell microtubule structure in a receptor-independent manner. Mol. Cell. Biochem. 2012, 366, 239–249. [Google Scholar] [CrossRef] [PubMed]
  117. Holm, A.; Baldetorp, B.; Olde, B.; Leeb-Lundberg, L.M.; Nilsson, B.O. The GPER1 agonist G-1 attenuates endothelial cell proliferation by inhibiting DNA synthesis and accumulating cells in the S and G2 phases of the cell cycle. J. Vasc. Res. 2011, 48, 327–335. [Google Scholar] [CrossRef] [PubMed]
  118. Mori, T.; Ito, F.; Matsushima, H.; Takaoka, O.; Tanaka, Y.; Koshiba, A.; Kusuki, I.; Kitawaki, J. G protein-coupled estrogen receptor 1 agonist G-1 induces cell cycle arrest in the mitotic phase, leading to apoptosis in endometriosis. Fertil. Steril. 2015, 103, 1228–1235.e1. [Google Scholar] [CrossRef] [PubMed]
  119. Arterburn, J.B.; Oprea, T.I.; Prossnitz, E.R.; Edwards, B.S.; Sklar, L.A. Discovery of selective probes and antagonists for G-protein-coupled receptors FPR/FPRL1 and GPR30. Curr. Top. Med. Chem. 2009, 9, 1227–1236. [Google Scholar] [CrossRef] [PubMed]
  120. Lappano, R.; Rosano, C.; Santolla, M.F.; Pupo, M.; De Francesco, E.M.; De Marco, P.; Ponassi, M.; Spallarossa, A.; Ranise, A.; Maggiolini, M. Two novel GPER agonists induce gene expression changes and growth effects in cancer cells. Curr. Cancer Drug Targets 2012, 12, 531–542. [Google Scholar] [CrossRef] [PubMed]
  121. Babiloni-Chust, I.; Dos Santos, R.S.; Medina-Gali, R.M.; Perez-Serna, A.A.; Encinar, J.A.; Martinez-Pinna, J.; Gustafsson, J.A.; Marroqui, L.; Nadal, A. G protein-coupled estrogen receptor activation by bisphenol-A disrupts the protection from apoptosis conferred by the estrogen receptors ERalpha and ERbeta in pancreatic beta cells. Environ. Int. 2022, 164, 107250. [Google Scholar] [CrossRef] [PubMed]
  122. He, W.; Gao, Z.; Liu, S.; Tan, L.; Wu, Y.; Liu, J.; Zheng, Z.; Fan, W.; Luo, Y.; Chen, Z.; et al. G protein-coupled estrogen receptor activation by bisphenol-A disrupts lipid metabolism and induces ferroptosis in the liver. Environ. Pollut. 2023, 334, 122211. [Google Scholar] [CrossRef] [PubMed]
  123. Pupo, M.; Pisano, A.; Lappano, R.; Santolla, M.F.; De Francesco, E.M.; Abonante, S.; Rosano, C.; Maggiolini, M. Bisphenol A induces gene expression changes and proliferative effects through GPER in breast cancer cells and cancer-associated fibroblasts. Environ. Health Perspect. 2012, 120, 1177–1182. [Google Scholar] [CrossRef] [PubMed]
  124. Thomas, P.; Dong, J. Binding and activation of the seven-transmembrane estrogen receptor GPR30 by environmental estrogens: A potential novel mechanism of endocrine disruption. J. Steroid Biochem. Mol. Biol. 2006, 102, 175–179. [Google Scholar] [CrossRef] [PubMed]
  125. Wang, Y.X.; Gu, Z.W.; Hao, L.Y. The environmental hormone nonylphenol interferes with the therapeutic effects of G protein-coupled estrogen receptor specific agonist G-1 on murine allergic rhinitis. Int. Immunopharmacol. 2020, 78, 106058. [Google Scholar] [CrossRef] [PubMed]
  126. Yang, L.; Chen, S.; Chen, Z.; Sun, X.; Gao, Q.; Lei, M.; Hao, L. Exploration of interaction property between nonylphenol and G protein-coupled receptor 30 based on molecular simulation and biological experiments. Steroids 2022, 188, 109114. [Google Scholar] [CrossRef] [PubMed]
  127. Albanito, L.; Lappano, R.; Madeo, A.; Chimento, A.; Prossnitz, E.R.; Cappello, A.R.; Dolce, V.; Abonante, S.; Pezzi, V.; Maggiolini, M. Effects of atrazine on estrogen receptor alpha- and G protein-coupled receptor 30-mediated signaling and proliferation in cancer cells and cancer-associated fibroblasts. Environ. Health Perspect. 2015, 123, 493–499. [Google Scholar] [CrossRef] [PubMed]
  128. Albanito, L.; Lappano, R.; Madeo, A.; Chimento, A.; Prossnitz, E.R.; Cappello, A.R.; Dolce, V.; Abonante, S.; Pezzi, V.; Maggiolini, M. G-protein-coupled receptor 30 and estrogen receptor-alpha are involved in the proliferative effects induced by atrazine in ovarian cancer cells. Environ. Health Perspect. 2008, 116, 1648–1655. [Google Scholar] [CrossRef] [PubMed]
  129. Vasquez-Reyes, S.; Vargas-Castillo, A.; Noriega, L.G.; Velazquez-Villegas, L.A.; Perez, B.; Sanchez-Tapia, M.; Ordaz, G.; Suarez-Monroy, R.; Ulloa-Aguirre, A.; Offner, H.; et al. Genistein Stimulation of White Adipose Tissue Thermogenesis Is Partially Dependent on GPR30 in Mice. Mol. Nutr. Food Res. 2022, 66, e2100838. [Google Scholar] [CrossRef] [PubMed]
  130. Thent, Z.C.; Froemming, G.R.A.; Ismail, A.B.M.; Fuad, S.; Muid, S. Phytoestrogens by inhibiting the non-classical oestrogen receptor, overcome the adverse effect of bisphenol A on hFOB 1.19 cells. Iran. J. Basic Med. Sci. 2020, 23, 1155–1163. [Google Scholar] [CrossRef] [PubMed]
  131. Liu, E.Y.L.; Xu, M.L.; Jin, Y.; Wu, Q.; Dong, T.T.X.; Tsim, K.W.K. Genistein, a Phytoestrogen in Soybean, Induces the Expression of Acetylcholinesterase via G Protein-Coupled Receptor 30 in PC12 Cells. Front. Mol. Neurosci. 2018, 11, 59. [Google Scholar] [CrossRef]
  132. Kajta, M.; Rzemieniec, J.; Litwa, E.; Lason, W.; Lenartowicz, M.; Krzeptowski, W.; Wojtowicz, A.K. The key involvement of estrogen receptor beta and G-protein-coupled receptor 30 in the neuroprotective action of daidzein. Neuroscience 2013, 238, 345–360. [Google Scholar] [CrossRef] [PubMed]
  133. Rowlands, D.J.; Chapple, S.; Siow, R.C.; Mann, G.E. Equol-stimulated mitochondrial reactive oxygen species activate endothelial nitric oxide synthase and redox signaling in endothelial cells: Roles for F-actin and GPR30. Hypertension 2011, 57, 833–840. [Google Scholar] [CrossRef] [PubMed]
  134. Chimento, A.; Casaburi, I.; Rosano, C.; Avena, P.; De Luca, A.; Campana, C.; Martire, E.; Santolla, M.F.; Maggiolini, M.; Pezzi, V.; et al. Oleuropein and hydroxytyrosol activate GPER/ GPR30-dependent pathways leading to apoptosis of ER-negative SKBR3 breast cancer cells. Mol. Nutr. Food Res. 2014, 58, 478–489. [Google Scholar] [CrossRef] [PubMed]
  135. Lu, D.; Wang, X.; Feng, C.; Liu, D.; Liu, Y.; Liu, Y.; Li, J.; Zhang, J.; Li, N.; Deng, Y.; et al. Study of the Sensing Kinetics of G Protein-Coupled Estrogen Receptor Sensors for Common Estrogens and Estrogen Analogs. Molecules 2023, 28, 3286. [Google Scholar] [CrossRef]
  136. D’Arrigo, G.; Gianquinto, E.; Rossetti, G.; Cruciani, G.; Lorenzetti, S.; Spyrakis, F. Binding of Androgen- and Estrogen-Like Flavonoids to Their Cognate (Non)Nuclear Receptors: A Comparison by Computational Prediction. Molecules 2021, 26, 1613. [Google Scholar] [CrossRef] [PubMed]
  137. Dong, W.H.; Chen, J.C.; He, Y.L.; Xu, J.J.; Mei, Y.A. Resveratrol inhibits K(v)2.2 currents through the estrogen receptor GPR30-mediated PKC pathway. Am. J. Physiol. Cell Physiol. 2013, 305, C547–C557. [Google Scholar] [CrossRef] [PubMed]
  138. Maggiolini, M.; Vivacqua, A.; Fasanella, G.; Recchia, A.G.; Sisci, D.; Pezzi, V.; Montanaro, D.; Musti, A.M.; Picard, D.; Ando, S. The G protein-coupled receptor GPR30 mediates c-fos up-regulation by 17beta-estradiol and phytoestrogens in breast cancer cells. J. Biol. Chem. 2004, 279, 27008–27016. [Google Scholar] [CrossRef] [PubMed]
  139. Kang, K.; Lee, S.B.; Jung, S.H.; Cha, K.H.; Park, W.D.; Sohn, Y.C.; Nho, C.W. Tectoridin, a poor ligand of estrogen receptor alpha, exerts its estrogenic effects via an ERK-dependent pathway. Mol. Cells 2009, 27, 351–357. [Google Scholar] [CrossRef] [PubMed]
  140. He, J.; Wei, C.; Li, Y.; Liu, Y.; Wang, Y.; Pan, J.; Liu, J.; Wu, Y.; Cui, S. Zearalenone and alpha-zearalenol inhibit the synthesis and secretion of pig follicle stimulating hormone via the non-classical estrogen membrane receptor GPR30. Mol. Cell. Endocrinol. 2018, 461, 43–54. [Google Scholar] [CrossRef]
  141. Nakamura, U.; Kadokawa, H. The nonsteroidal mycoestrogen zearalenone and its five metabolites suppress LH secretion from the bovine anterior pituitary cells via the estradiol receptor GPR30 in vitro. Theriogenology 2015, 84, 1342–1349. [Google Scholar] [CrossRef] [PubMed]
  142. Palmieri, D.; Perego, P.; Palombo, D. Apigenin inhibits the TNFalpha-induced expression of eNOS and MMP-9 via modulating Akt signalling through oestrogen receptor engagement. Mol. Cell. Biochem. 2012, 371, 129–136. [Google Scholar] [CrossRef]
  143. Koganti, S.; Snyder, R.; Gumaste, U.; Karamyan, V.T.; Thekkumkara, T. 2-methoxyestradiol binding of GPR30 down-regulates angiotensin AT(1) receptor. Eur. J. Pharmacol. 2014, 723, 131–140. [Google Scholar] [CrossRef] [PubMed]
  144. Chourasia, T.K.; Pang, Y.; Thomas, P. The catecholestrogen, 2-hydroxyestradiol-17beta, acts as a G protein-coupled estrogen receptor 1 (GPER/GPR30) antagonist to promote the resumption of meiosis in zebrafish oocytes. Biol. Reprod. 2015, 92, 69. [Google Scholar] [CrossRef] [PubMed]
  145. Cheng, S.B.; Dong, J.; Pang, Y.; LaRocca, J.; Hixon, M.; Thomas, P.; Filardo, E.J. Anatomical location and redistribution of G protein-coupled estrogen receptor-1 during the estrus cycle in mouse kidney and specific binding to estrogens but not aldosterone. Mol. Cell. Endocrinol. 2014, 382, 950–959. [Google Scholar] [CrossRef]
  146. Iqbal, A.M.; Jamal, S.F. Essential Hypertension. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2023. [Google Scholar]
  147. Curfman, G.; Bauchner, H.; Greenland, P. Treatment and Control of Hypertension in 2020: The Need for Substantial Improvement. JAMA 2020, 324, 1166–1167. [Google Scholar] [CrossRef] [PubMed]
  148. Doggrell, S.A.; Brown, L. Rat models of hypertension, cardiac hypertrophy and failure. Cardiovasc. Res. 1998, 39, 89–105. [Google Scholar] [CrossRef] [PubMed]
  149. Tahvanainen, A.; Taurio, J.; Maki-Jouppi, J.; Koobi, P.; Mustonen, J.; Kahonen, M.; Sand, J.; Nordback, I.; Porsti, I. Increased wall tension in response to vasoconstrictors in isolated mesenteric arterial rings from patients with high blood pressure. Basic Clin. Pharmacol. Toxicol. 2006, 99, 440–449. [Google Scholar] [CrossRef]
  150. O’Dowd, B.F.; Nguyen, T.; Marchese, A.; Cheng, R.; Lynch, K.R.; Heng, H.H.; Kolakowski, L.F., Jr.; George, S.R. Discovery of three novel G-protein-coupled receptor genes. Genomics 1998, 47, 310–313. [Google Scholar] [CrossRef]
  151. Stowasser, M.; Gordon, R.D. Aldosterone excess, hypertension, and chromosome 7p22: Evidence continues to mount. Hypertension 2007, 49, 761–762. [Google Scholar] [CrossRef]
  152. Liu, S.; Ding, T.; Liu, H.; Jian, L. GPER was Associated with Hypertension in Post-Menopausal Women. Open Med. 2018, 13, 338–343. [Google Scholar] [CrossRef]
  153. Feldman, R.D.; Gros, R.; Ding, Q.; Hussain, Y.; Ban, M.R.; McIntyre, A.D.; Hegele, R.A. A common hypofunctional genetic variant of GPER is associated with increased blood pressure in women. Br. J. Clin. Pharmacol. 2014, 78, 1441–1452. [Google Scholar] [CrossRef] [PubMed]
  154. Dinh, Q.N.; Vinh, A.; Kim, H.A.; Saini, N.; Broughton, B.R.S.; Chrissobolis, S.; Diep, H.; Judkins, C.P.; Drummond, G.R.; Sobey, C.G. Aldosterone-induced hypertension is sex-dependent, mediated by T cells and sensitive to GPER activation. Cardiovasc. Res. 2021, 117, 960–970. [Google Scholar] [CrossRef] [PubMed]
  155. Liu, L.; Kashyap, S.; Murphy, B.; Hutson, D.D.; Budish, R.A.; Trimmer, E.H.; Zimmerman, M.A.; Trask, A.J.; Miller, K.S.; Chappell, M.C.; et al. GPER activation ameliorates aortic remodeling induced by salt-sensitive hypertension. Am. J. Physiol. Heart Circ. Physiol. 2016, 310, H953–H961. [Google Scholar] [CrossRef]
  156. Delgado, N.T.B.; Rouver, W.D.N.; Freitas-Lima, L.C.; Vieira-Alves, I.; Lemos, V.S.; Dos Santos, R.L. Sex Differences in the Vasodilation Mediated by G Protein-Coupled Estrogen Receptor (GPER) in Hypertensive Rats. Front. Physiol. 2021, 12, 659291. [Google Scholar] [CrossRef]
  157. Meyer, M.R.; Prossnitz, E.R.; Barton, M. GPER/GPR30 and Regulation of Vascular Tone and Blood Pressure. Immunol. Endocr. Metab. Agents Med. Chem. 2011, 11, 255–261. [Google Scholar] [CrossRef]
  158. Giles, T.D.; Sander, G.E.; Nossaman, B.D.; Kadowitz, P.J. Impaired vasodilation in the pathogenesis of hypertension: Focus on nitric oxide, endothelial-derived hyperpolarizing factors, and prostaglandins. J. Clin. Hypertens. 2012, 14, 198–205. [Google Scholar] [CrossRef] [PubMed]
  159. Li, Z.L.; Liu, J.C.; Liu, S.B.; Li, X.Q.; Yi, D.H.; Zhao, M.G. Improvement of vascular function by acute and chronic treatment with the GPR30 agonist G1 in experimental diabetes mellitus. PLoS ONE 2012, 7, e38787. [Google Scholar] [CrossRef]
  160. Jang, E.J.; Seok, Y.M.; Arterburn, J.B.; Olatunji, L.A.; Kim, I.K. GPER-1 agonist G1 induces vasorelaxation through activation of epidermal growth factor receptor-dependent signalling pathway. J. Pharm. Pharmacol. 2013, 65, 1488–1499. [Google Scholar] [CrossRef]
  161. Bai, J.; Qi, Q.R.; Li, Y.; Day, R.; Makhoul, J.; Magness, R.R.; Chen, D.B. Estrogen Receptors and Estrogen-Induced Uterine Vasodilation in Pregnancy. Int. J. Mol. Sci. 2020, 21, 4349. [Google Scholar] [CrossRef]
  162. Tropea, T.; De Francesco, E.M.; Rigiracciolo, D.; Maggiolini, M.; Wareing, M.; Osol, G.; Mandala, M. Pregnancy Augments G Protein Estrogen Receptor (GPER) Induced Vasodilation in Rat Uterine Arteries via the Nitric Oxide–cGMP Signaling Pathway. PLoS ONE 2015, 10, e0141997. [Google Scholar] [CrossRef]
  163. Batenburg, W.W.; Jansen, P.M.; van den Bogaerdt, A.J.; AH, J.D. Angiotensin II-aldosterone interaction in human coronary microarteries involves GPR30, EGFR, and endothelial NO synthase. Cardiovasc. Res. 2012, 94, 136–143. [Google Scholar] [CrossRef] [PubMed]
  164. Kim, K.H.; Young, B.D.; Bender, J.R. Endothelial estrogen receptor isoforms and cardiovascular disease. Mol. Cell. Endocrinol. 2014, 389, 65–70. [Google Scholar] [CrossRef] [PubMed]
  165. Peixoto, P.; Aires, R.D.; Lemos, V.S.; Bissoli, N.S.; Santos, R.L.D. GPER agonist dilates mesenteric arteries via PI3K-Akt-eNOS and potassium channels in both sexes. Life Sci. 2017, 183, 21–27. [Google Scholar] [CrossRef] [PubMed]
  166. Peixoto, P.; Vieira-Alves, I.; Couto, G.K.; Lemos, V.S.; Rossoni, L.V.; Bissoli, N.S.; Santos, R.L.D. Sex differences in the participation of endothelial mediators and signaling pathways involved in the vasodilator effect of a selective GPER agonist in resistance arteries of gonadectomized Wistar rats. Life Sci. 2022, 308, 120917. [Google Scholar] [CrossRef] [PubMed]
  167. Browner, N.C.; Dey, N.B.; Bloch, K.D.; Lincoln, T.M. Regulation of cGMP-dependent protein kinase expression by soluble guanylyl cyclase in vascular smooth muscle cells. J. Biol. Chem. 2004, 279, 46631–46636. [Google Scholar] [CrossRef]
  168. Yu, X.; Ma, H.; Barman, S.A.; Liu, A.T.; Sellers, M.; Stallone, J.N.; Prossnitz, E.R.; White, R.E.; Han, G. Activation of G protein-coupled estrogen receptor induces endothelium-independent relaxation of coronary artery smooth muscle. Am. J. Physiol. Endocrinol. Metab. 2011, 301, E882–E888. [Google Scholar] [CrossRef]
  169. Yu, X.; Stallone, J.N.; Heaps, C.L.; Han, G. The activation of G protein-coupled estrogen receptor induces relaxation via cAMP as well as potentiates contraction via EGFR transactivation in porcine coronary arteries. PLoS ONE 2018, 13, e0191418. [Google Scholar] [CrossRef]
  170. Keung, W.; Vanhoutte, P.M.; Man, R.Y. Acute impairment of contractile responses by 17beta-estradiol is cAMP and protein kinase G dependent in vascular smooth muscle cells of the porcine coronary arteries. Br. J. Pharmacol. 2005, 144, 71–79. [Google Scholar] [CrossRef]
  171. Nadadur, R.D.; Umar, S.; Wong, G.; Eghbali, M.; Iorga, A.; Matori, H.; Partow-Navid, R.; Eghbali, M. Reverse right ventricular structural and extracellular matrix remodeling by estrogen in severe pulmonary hypertension. J. Appl. Physiol. 2012, 113, 149–158. [Google Scholar] [CrossRef]
  172. Frost, A.E.; Badesch, D.B.; Barst, R.J.; Benza, R.L.; Elliott, C.G.; Farber, H.W.; Krichman, A.; Liou, T.G.; Raskob, G.E.; Wason, P.; et al. The changing picture of patients with pulmonary arterial hypertension in the United States: How REVEAL differs from historic and non-US Contemporary Registries. Chest 2011, 139, 128–137. [Google Scholar] [CrossRef]
  173. Resta, T.C.; Kanagy, N.L.; Walker, B.R. Estradiol-induced attenuation of pulmonary hypertension is not associated with altered eNOS expression. Am. J. Physiol. Lung Cell. Mol. Physiol. 2001, 280, L88–L97. [Google Scholar] [CrossRef] [PubMed]
  174. Umar, S.; Rabinovitch, M.; Eghbali, M. Estrogen paradox in pulmonary hypertension: Current controversies and future perspectives. Am. J. Respir. Crit. Care Med. 2012, 186, 125–131. [Google Scholar] [CrossRef] [PubMed]
  175. Sun, Y.; Sangam, S.; Guo, Q.; Wang, J.; Tang, H.; Black, S.M.; Desai, A.A. Sex Differences, Estrogen Metabolism and Signaling in the Development of Pulmonary Arterial Hypertension. Front. Cardiovasc. Med. 2021, 8, 719058. [Google Scholar] [CrossRef] [PubMed]
  176. Alencar, A.K.; Montes, G.C.; Montagnoli, T.; Silva, A.M.; Martinez, S.T.; Fraga, A.G.; Wang, H.; Groban, L.; Sudo, R.T.; Zapata-Sudo, G. Activation of GPER ameliorates experimental pulmonary hypertension in male rats. Eur. J. Pharm. Sci. 2017, 97, 208–217. [Google Scholar] [CrossRef] [PubMed]
  177. Alencar, A.K.N.; Montes, G.C.; Costa, D.G.; Mendes, L.V.P.; Silva, A.M.S.; Martinez, S.T.; Trachez, M.M.; Cunha, V.; Montagnoli, T.L.; Fraga, A.G.M.; et al. Cardioprotection Induced by Activation of GPER in Ovariectomized Rats With Pulmonary Hypertension. J. Gerontol. A Biol. Sci. Med. Sci. 2018, 73, 1158–1166. [Google Scholar] [CrossRef]
  178. Insull, W., Jr. The pathology of atherosclerosis: Plaque development and plaque responses to medical treatment. Am. J. Med. 2009, 122, S3–S14. [Google Scholar] [CrossRef]
  179. Joakimsen, O.; Bonaa, K.H.; Stensland-Bugge, E.; Jacobsen, B.K. Population-based study of age at menopause and ultrasound assessed carotid atherosclerosis: The Tromso Study. J. Clin. Epidemiol. 2000, 53, 525–530. [Google Scholar] [CrossRef]
  180. Sharma, G.; Hu, C.; Brigman, J.L.; Zhu, G.; Hathaway, H.J.; Prossnitz, E.R. GPER deficiency in male mice results in insulin resistance, dyslipidemia, and a proinflammatory state. Endocrinology 2013, 154, 4136–4145. [Google Scholar] [CrossRef]
  181. Hussain, Y.; Ding, Q.; Connelly, P.W.; Brunt, J.H.; Ban, M.R.; McIntyre, A.D.; Huff, M.W.; Gros, R.; Hegele, R.A.; Feldman, R.D. G-protein estrogen receptor as a regulator of low-density lipoprotein cholesterol metabolism: Cellular and population genetic studies. Arter. Thromb. Vasc. Biol. 2015, 35, 213–221. [Google Scholar] [CrossRef]
  182. Ghaffari, S.; Naderi Nabi, F.; Sugiyama, M.G.; Lee, W.L. Estrogen Inhibits LDL (Low-Density Lipoprotein) Transcytosis by Human Coronary Artery Endothelial Cells via GPER (G-Protein-Coupled Estrogen Receptor) and SR-BI (Scavenger Receptor Class B Type 1). Arter. Thromb. Vasc. Biol. 2018, 38, 2283–2294. [Google Scholar] [CrossRef]
  183. Chakrabarti, S.; Davidge, S.T. G-protein coupled receptor 30 (GPR30): A novel regulator of endothelial inflammation. PLoS ONE 2012, 7, e52357. [Google Scholar] [CrossRef] [PubMed]
  184. Fan, L.; Wei, W.; Ning, H.; Sun, C. G protein-coupled estrogen receptor increases vascular endothelial inflammatory response after ovariectomized mice. Wei Sheng Yan Jiu 2021, 50, 279–283. [Google Scholar] [CrossRef] [PubMed]
  185. Gros, R.; Hussain, Y.; Chorazyczewski, J.; Pickering, J.G.; Ding, Q.; Feldman, R.D. Extent of Vascular Remodeling Is Dependent on the Balance Between Estrogen Receptor alpha and G-Protein-Coupled Estrogen Receptor. Hypertension 2016, 68, 1225–1235. [Google Scholar] [CrossRef]
  186. Bowling, M.R.; Xing, D.; Kapadia, A.; Chen, Y.F.; Szalai, A.J.; Oparil, S.; Hage, F.G. Estrogen effects on vascular inflammation are age dependent: Role of estrogen receptors. Arter. Thromb. Vasc. Biol. 2014, 34, 1477–1485. [Google Scholar] [CrossRef] [PubMed]
  187. Blasko, E.; Haskell, C.A.; Leung, S.; Gualtieri, G.; Halks-Miller, M.; Mahmoudi, M.; Dennis, M.K.; Prossnitz, E.R.; Karpus, W.J.; Horuk, R. Beneficial role of the GPR30 agonist G-1 in an animal model of multiple sclerosis. J. Neuroimmunol. 2009, 214, 67–77. [Google Scholar] [CrossRef]
  188. Yates, M.A.; Li, Y.; Chlebeck, P.J.; Offner, H. GPR30, but not estrogen receptor-alpha, is crucial in the treatment of experimental autoimmune encephalomyelitis by oral ethinyl estradiol. BMC Immunol. 2010, 11, 20. [Google Scholar] [CrossRef]
  189. Tamaki, M.; Konno, Y.; Kobayashi, Y.; Takeda, M.; Itoga, M.; Moritoki, Y.; Oyamada, H.; Kayaba, H.; Chihara, J.; Ueki, S. Expression and functional roles of G-protein-coupled estrogen receptor (GPER) in human eosinophils. Immunol. Lett. 2014, 160, 72–78. [Google Scholar] [CrossRef]
  190. Rettew, J.A.; McCall, S.H.t.; Marriott, I. GPR30/GPER-1 mediates rapid decreases in TLR4 expression on murine macrophages. Mol. Cell. Endocrinol. 2010, 328, 87–92. [Google Scholar] [CrossRef]
  191. Zhou, Z.; Chen, Y.; Ni, W.; Liu, T. Upregulation of Nuclear Factor IA Suppresses Oxidized Low-Density Lipoprotein-Induced Endoplasmic Reticulum Stress and Apoptosis in Human Umbilical Vein Endothelial Cells. Med. Sci. Monit. 2019, 25, 1009–1016. [Google Scholar] [CrossRef]
  192. Yang, G.; Zhuo, J.; Lin, Y.; Zhang, M.; Liu, L.; Chen, X.; Gao, R. Ginsenoside Rb1 Prevents Dysfunction of Endothelial Cells by Suppressing Inflammatory Response and Apoptosis in the High-Fat Diet Plus Balloon Catheter-Injured Rabbit Model via the G Protein-Coupled Estrogen Receptor-Mediated Phosphatidylinositol 3-Kinases (PI3K)/Akt Pathway. Med. Sci. Monit. 2019, 25, 7407–7417. [Google Scholar] [CrossRef]
  193. Jing, Y.; Cai, D.; Chen, Q.; Xiong, Q.; Hu, T.; Yao, Y.; Lin, C.; Sun, X.; Lu, Y.; Kong, X.; et al. Liuwei Dihuang soft capsules attenuates endothelial cell apoptosis to prevent atherosclerosis through GPR30-mediated regulation in ovariectomized ApoE-deficient mice. J. Ethnopharmacol. 2017, 208, 185–198. [Google Scholar] [CrossRef]
  194. Feng, Z.; Wang, C.; Jin, Y.; Meng, Q.; Wu, J.; Sun, H. Kaempferol-induced GPER upregulation attenuates atherosclerosis via the PI3K/AKT/Nrf2 pathway. Pharm. Biol. 2021, 59, 1106–1116. [Google Scholar] [CrossRef]
  195. Bennett, M.R.; Sinha, S.; Owens, G.K. Vascular Smooth Muscle Cells in Atherosclerosis. Circ. Res. 2016, 118, 692–702. [Google Scholar] [CrossRef] [PubMed]
  196. Zhu, J.; Liu, B.; Wang, Z.; Wang, D.; Ni, H.; Zhang, L.; Wang, Y. Exosomes from nicotine-stimulated macrophages accelerate atherosclerosis through miR-21-3p/PTEN-mediated VSMC migration and proliferation. Theranostics 2019, 9, 6901–6919. [Google Scholar] [CrossRef] [PubMed]
  197. Xia, X.; Zhou, C.; He, X.; Liu, C.; Wang, G.; Sun, X. The relationship between estrogen-induced phenotypic transformation and proliferation of vascular smooth muscle and hypertensive intracerebral hemorrhage. Ann. Transl. Med. 2020, 8, 762. [Google Scholar] [CrossRef] [PubMed]
  198. Wang, C.; Li, J.; Ye, S.; Zhang, Y.; Li, P.; Wang, L.; Wang, T.H. Oestrogen Inhibits VEGF Expression And Angiogenesis In Triple-Negative Breast Cancer By Activating GPER-1. J. Cancer 2018, 9, 3802–3811. [Google Scholar] [CrossRef]
  199. Blakemore, J.; Naftolin, F. Aromatase: Contributions to Physiology and Disease in Women and Men. Physiology 2016, 31, 258–269. [Google Scholar] [CrossRef]
  200. Lee, H.R.; Kim, T.H.; Choi, K.C. Functions and physiological roles of two types of estrogen receptors, ERalpha and ERbeta, identified by estrogen receptor knockout mouse. Lab. Anim. Res. 2012, 28, 71–76. [Google Scholar] [CrossRef]
  201. Nilsson, S.; Gustafsson, J.A. Estrogen receptors: Therapies targeted to receptor subtypes. Clin. Pharmacol. Ther. 2011, 89, 44–55. [Google Scholar] [CrossRef]
  202. Shook, L.L. An update on hormone replacement therapy: Health and medicine for women: A multidisciplinary, evidence-based review of mid-life health concerns. Yale J. Biol. Med. 2011, 84, 39–42. [Google Scholar]
  203. Wassertheil-Smoller, S.; Hendrix, S.L.; Limacher, M.; Heiss, G.; Kooperberg, C.; Baird, A.; Kotchen, T.; Curb, J.D.; Black, H.; Rossouw, J.E.; et al. Effect of estrogen plus progestin on stroke in postmenopausal women: The Women’s Health Initiative: A randomized trial. JAMA 2003, 289, 2673–2684. [Google Scholar] [CrossRef] [PubMed]
  204. Gillum, L.A.; Mamidipudi, S.K.; Johnston, S.C. Ischemic stroke risk with oral contraceptives: A meta-analysis. JAMA 2000, 284, 72–78. [Google Scholar] [CrossRef] [PubMed]
  205. Venous thromboembolic disease and combined oral contraceptives: Results of international multicentre case-control study. World Health Organization Collaborative Study of Cardiovascular Disease and Steroid Hormone Contraception. Lancet 1995, 346, 1575–1582. [CrossRef]
  206. Barton, M. Cholesterol and atherosclerosis: Modulation by oestrogen. Curr. Opin. Lipidol. 2013, 24, 214–220. [Google Scholar] [CrossRef]
  207. Prossnitz, E.R.; Barton, M. Estrogen biology: New insights into GPER function and clinical opportunities. Mol. Cell. Endocrinol. 2014, 389, 71–83. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Classical signaling pathways activated by GPER. GPER binds to its endogenous ligands (estrogen and aldosterone), leading to multiple downstream cascades, including ERK, PI3K/AKT, AMPK/PKA, and NF-kB pathways. Activation of Src via GPER results in the release of matrix metalloproteinase (MMP) mediation of heparin-binding epidermal growth factor (HB-EGF) cleavage, followed by transactivation of the EGF receptor (EGFR), which in turn causes the activation of extracellular signal-regulated kinases (ERK) and the activation of phosphatidylinositol 3-kinase (PI3K) and serine/threonine kinase (PI3K/AKT). GPER stimulates the AMPK signaling pathway, which increases adenylyl cyclase (AC) activity to stimulate the cAMP-activated downstream protein kinase A (PKA) signal pathway that inhibits ERK. GPER also reduces the phosphorylation, nuclear localization, and transcriptional activity of the NF-κB pathway.
Figure 1. Classical signaling pathways activated by GPER. GPER binds to its endogenous ligands (estrogen and aldosterone), leading to multiple downstream cascades, including ERK, PI3K/AKT, AMPK/PKA, and NF-kB pathways. Activation of Src via GPER results in the release of matrix metalloproteinase (MMP) mediation of heparin-binding epidermal growth factor (HB-EGF) cleavage, followed by transactivation of the EGF receptor (EGFR), which in turn causes the activation of extracellular signal-regulated kinases (ERK) and the activation of phosphatidylinositol 3-kinase (PI3K) and serine/threonine kinase (PI3K/AKT). GPER stimulates the AMPK signaling pathway, which increases adenylyl cyclase (AC) activity to stimulate the cAMP-activated downstream protein kinase A (PKA) signal pathway that inhibits ERK. GPER also reduces the phosphorylation, nuclear localization, and transcriptional activity of the NF-κB pathway.
Biomolecules 13 01410 g001
Figure 2. GPER-mediated relaxation of VSMCs. Activation of GPER increases intracellular Ca2+ levels through transient receptor potential channel (TRPC) isoforms, leading to the phosphorylation of calmodulin-dependent protein kinase 2 (CaMKKβ), AMP-activated protein kinase (AMPK), and calmodulin-dependent protein kinase II (CaMKIIα), which in turn elevates the expression of eNOS through the activation of histone deacetylase 5 (HDAC5) and the Kruppel-like factor 2 (KLF2) signaling pathway. GPER also enhances the phosphorylation of EGFR by interacting with various proteins around the membrane, such as Src, which further induces the expression of eNOS through ERK5 and muscle cell enhancer 2 peptide C (MEF2C). GPER phosphorylates eNOS through the non-genomic Akt/ERK pathway, which in turn promotes NO production. NO produced in ECs stimulates sGC in VSMCs to produce cGMP, which in turn activates PKG. PKG cooperates with Akt/ERK-dependent PKA and Epac/Rap1 pathways to inhibit RhoA/Rho, thereby causing MLCP activation and Ca2+ channel opening, ultimately leading to VSMC relaxation. EC, endothelial cell; VSMC, vascular smooth muscle cell.
Figure 2. GPER-mediated relaxation of VSMCs. Activation of GPER increases intracellular Ca2+ levels through transient receptor potential channel (TRPC) isoforms, leading to the phosphorylation of calmodulin-dependent protein kinase 2 (CaMKKβ), AMP-activated protein kinase (AMPK), and calmodulin-dependent protein kinase II (CaMKIIα), which in turn elevates the expression of eNOS through the activation of histone deacetylase 5 (HDAC5) and the Kruppel-like factor 2 (KLF2) signaling pathway. GPER also enhances the phosphorylation of EGFR by interacting with various proteins around the membrane, such as Src, which further induces the expression of eNOS through ERK5 and muscle cell enhancer 2 peptide C (MEF2C). GPER phosphorylates eNOS through the non-genomic Akt/ERK pathway, which in turn promotes NO production. NO produced in ECs stimulates sGC in VSMCs to produce cGMP, which in turn activates PKG. PKG cooperates with Akt/ERK-dependent PKA and Epac/Rap1 pathways to inhibit RhoA/Rho, thereby causing MLCP activation and Ca2+ channel opening, ultimately leading to VSMC relaxation. EC, endothelial cell; VSMC, vascular smooth muscle cell.
Biomolecules 13 01410 g002
Table 1. The structure and function of GPER ligands.
Table 1. The structure and function of GPER ligands.
CompoundStructureAgonist/ AntagonistAffinity/Efficacy
Kd/IC50/EC50
Reference(s)
Endogenous Ligand
 Estrogen
  Estrone (E1)Biomolecules 13 01410 i001Unknown>>10 μM[17]
  Estradiol (E2)
   17β-Estradiol Biomolecules 13 01410 i002Agonist3–6 nM[17,18]
   17α-Estradiol Biomolecules 13 01410 i003Unknown>10 μM[17]
   2-Methoxy-E2 Biomolecules 13 01410 i004Agonist10 nM[143]
   2-Hydroxy-E2 Biomolecules 13 01410 i005Antagonist0.1–1 μM[144]
  Estriol (E3) Biomolecules 13 01410 i006Antagonist>1 μM[17,54]
 AldosteroneBiomolecules 13 01410 i007AgonistNone detected[95,104,145]
Synthetic Ligands
 G1Biomolecules 13 01410 i008Agonist7–11 nM[111,112]
 G15Biomolecules 13 01410 i009Antagonist20 nM[112,119]
 G36Biomolecules 13 01410 i010Antagonist~ 20 nM[113]
 GPER-L1Biomolecules 13 01410 i011Agonist100 nM[120]
 GPER-L2Biomolecules 13 01410 i012Agonist100 nM[120]
 Bisphenol ABiomolecules 13 01410 i013Agonist0.6 μM[121,122,123,124]
 NonylphenolBiomolecules 13 01410 i014Agonist0.8 μM[124,125,126]
 AtrazineBiomolecules 13 01410 i015Agonist>10 μM[124,127,128]
 KeponeBiomolecules 13 01410 i016Agonist1.4 μM[124]
 MethoxychlorBiomolecules 13 01410 i017Unknown~10 μM[124]
 p,p′-DDTBiomolecules 13 01410 i018Unknown2.8 μM[124]
 o,p′-DDTBiomolecules 13 01410 i019Agonist7.1 μM[17,124]
 p,p′-DDEBiomolecules 13 01410 i020Unknown~10 μM[124]
 2,2′5′-PCB-4-OHBiomolecules 13 01410 i021Unknown3.8 μM[124]
Phytoestrogens
 GenisteinBiomolecules 13 01410 i022Agonist133 nM[124,129,130,131]
 DaidzeinBiomolecules 13 01410 i023Agonist<1 nM[132]
 EquolBiomolecules 13 01410 i024Agonist100 nM[133]
 OleuropeinBiomolecules 13 01410 i025Agonist200 nM[134]
 HydroxytyrosolBiomolecules 13 01410 i026Agonist100 nM[133]
 ResveratrolBiomolecules 13 01410 i027Agonist300 nM[135,136,137]
 QuercetinBiomolecules 13 01410 i028Agonist1 μM[136,138]
 TectoridinBiomolecules 13 01410 i029Agonist10 μM[139]
 ApigeninBiomolecules 13 01410 i030Agonist20–50 μM[136,142]
 ZearalenoneBiomolecules 13 01410 i031Unknown0.8 μM[124,140,141]
Table 2. The role of GPER in vascular pathology.
Table 2. The role of GPER in vascular pathology.
ModelInterventionOutcomeReference
Hypertension
Human endothelial cellsGPER agonistGPER↑→eNOS↑→Src/EGFR/PI3K/ERK↑→NO↑→Vasodilation↑→HTN↓[13]
Human endothelial cellsGPER antagonistGPER↓→NO↓→Vasodilation↓→HTN↑[13]
Human endothelial EA.hy2 cellsGPER agonistGPER↑→CaMKKβ, AMPK, CaMKIIα phosphorylation↑→HDAC5, KLF2 →eNOS↑→NO↑→Vasodilation↑[68]
Pig coronary arteryGPER agonistGPER↑→cAMP/PKA↑→RhoA↓→MLCP↑→Vasodilation↑[82]
Pig coronary arteryGPER agonistGPER↑→Epac/Rap1↑→RhoA/Rho↓→Vasodilation↑[83]
Rat aortic smooth muscle cellsGPER mutationERK↓, Apoptosis↓→ HTN↑[153]
Aldosterone-induced hypertension in miceGPER KO or antagonistGPER↓→HTN↑[154]
mRen2 female hypertensive ratGPER agonistGPER↑→Oxidative stress↓→Aortic remodeling↓[155]
Mesenteric artery of ratGPER agonistGPER↑→NO↑, H2O2↑→Vasodilation↑[156]
Mesenteric artery of ratGPER agonistGPER↑→PI3K/Akt/eNOS↑→NO↑→Vasodilation↑[165]
Mesenteric artery of male ratGPER agonistGPER↑→PI3K/Akt/eNOS↑→NO↑→Vasodilation↑[166]
Mesenteric artery of female ratGPER agonistGPER↑→MEK/ERK/eNOS↑→NO↑→Vasodilation↑[166]
Human or pig coronary smooth muscle cellsGPER agonistGPER↑→BK (Ca)↑→Potassium efflux↑→Vasodilation↑[168]
Pulmonary arterial hypertension
Male rats with MCT-induced PHGPER agonistGPER↑→eNOS↑, Collagen deposition in pulmonary and cardiac fibroblasts↓, Ca2+↑, Inflammation in cardiomyocytes↓→Pulmonary flow↑, RV hypertrophy↑, LV dysfunction↑[176]
OVX female rats with MCT-induced PHGPER agonistGPER↑→Pulmonary artery dysfunction↓, RV overload↓, RV dilation↓, Wall hypertrophy↓, Collagen deposition↓, Normalizes LV dysfunction[177]
Atherosclerosis
Mice fed high-fat dietGPER KO or OVXGPER↓→TC↑, LDL↑, Inflammation↑, NO bioactivity↓→AS↑[32]
Mice fed high-fat dietGPER agonistGPER↑→Inflammation↓→AS↓[32]
Male and female miceGPER deletionGPER↓→ endothelium-dependent vasoconstriction↑, visceral obesity↑, LDL↑, inflammation↑[67]
Male miceGPER KOGPER KO→Insulin resistance, Dyslipidemia, Inflammatory effects↑[180]
Female miceGPER KOGPER KO→HDL↓, TG↑[180]
Human coronary artery ECsGPER knockdownGPER↑→EGFR↑→endothelial scavenger receptor class B type I↓→LDL transcytosis↓[182]
Female miceGPER agonistCa↑→GPER↓→TRPC1/ERK1/2↓→AS↑[184]
Rat aortic SMCs
GPER agonist,
ERα knockdown
Hypertrophy and inflammation of medial vessels after carotid ligation↓[185]
Rabbit atherosclerotic ECsGPER agonistGPER↑→PI3K/Akt↑→Inflammation↓, Apoptosis↓→Dysfunction of ECs↓→TC↓, TG↓, LDL↓, HDL↑→AS↓[192]
ApoE miceGPER agonistGPER↑→Apoptosis of ECs↓→Plasma lipid↓→AS↓[193]
ApoE miceGPER agonistGPER↑→PI3K/Akt/Nrf2↑→Inflammation↓, Apoptosis↓→AS↓[194]
GPER: G protein-coupled estrogen receptor; eNOS: endothelial nitric oxide synthase; Src: tyrosine kinase; EGFR: epidermal growth factor receptor; PI3K: phosphatidylinositol 3-kinase; ERK: extracellular signal-regulated kinase; NO: nitric oxide; HTN: hypertension; CaMKKβ: calmodulin-dependent protein kinase 2; AMPK: AMP-activated protein kinases; CaMKIIα: calmodulin-dependent protein kinase II; HDAC5: histone deacetylase 5; KLF2: Kruppel-like factor 2; cAMP: cyclic adenosine monophosphate; PKA: protein kinase A; MLCP: myosin light chain phosphatase; Epac/Rap1: exchange protein directly activated by camp 1/ras-related protein 1; RhoA/Rho: ras homolog gene family, member A/ras homolog gene family; KO: knockout; H2O2: hydrogen peroxide; Akt: serine/threonine kinase; MEK: mitogen-activated protein kinase kinase; BK (Ca): large-conductance calcium-activated potassium; RV: right ventricle; LV: left ventricle; OVX: ovariectomy; MCT: monocrotaline; TC: total cholesterol; ECs: endothelial cells; LDL: low-density lipoprotein; AS: atherosclerosis; HDL: high-density lipoprotein; TG: triglyceride; TRPC1: transient receptor potential channel 1; SMCs: smooth muscle cells; Nrf2: nuclear factor-erythroid 2-related factor 2; ApoE: apolipoprotein E-deficient; ↑: increased; ↓: decreased; →: results in.
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

Xu, F.; Ma, J.; Wang, X.; Wang, X.; Fang, W.; Sun, J.; Li, Z.; Liu, J. The Role of G Protein-Coupled Estrogen Receptor (GPER) in Vascular Pathology and Physiology. Biomolecules 2023, 13, 1410. https://doi.org/10.3390/biom13091410

AMA Style

Xu F, Ma J, Wang X, Wang X, Fang W, Sun J, Li Z, Liu J. The Role of G Protein-Coupled Estrogen Receptor (GPER) in Vascular Pathology and Physiology. Biomolecules. 2023; 13(9):1410. https://doi.org/10.3390/biom13091410

Chicago/Turabian Style

Xu, Fujie, Jipeng Ma, Xiaowu Wang, Xiaoya Wang, Weiyi Fang, Jingwei Sun, Zilin Li, and Jincheng Liu. 2023. "The Role of G Protein-Coupled Estrogen Receptor (GPER) in Vascular Pathology and Physiology" Biomolecules 13, no. 9: 1410. https://doi.org/10.3390/biom13091410

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