Next Article in Journal
Garlic Volatile Diallyl Disulfide Induced Cucumber Resistance to Downy Mildew
Previous Article in Journal
Intrauterine L-NAME Exposure Weakens the Development of Sympathetic Innervation and Induces the Remodeling of Arterial Vessels in Two-Week-Old Rats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nuclear Receptors in Myocardial and Cerebral Ischemia—Mechanisms of Action and Therapeutic Strategies

1
Department of Pharmaceutical Sciences, University of Milan, 20133 Milan, Italy
2
Laboratory of Neuropharmacology and Epigenetics, Department of Pharmacology, Maj Institute of Pharmacology Polish Academy of Sciences, 31343 Krakow, Poland
3
Centro Cardiologico Monzino IRCCS, 20138 Milan, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(22), 12326; https://doi.org/10.3390/ijms222212326
Submission received: 26 October 2021 / Revised: 8 November 2021 / Accepted: 10 November 2021 / Published: 15 November 2021
(This article belongs to the Section Molecular Pathology, Diagnostics, and Therapeutics)

Abstract

:
Nearly 18 million people died from cardiovascular diseases in 2019, of these 85% were due to heart attack and stroke. The available therapies although efficacious, have narrow therapeutic window and long list of contraindications. Therefore, there is still an urgent need to find novel molecular targets that could protect the brain and heart against ischemia without evoking major side effects. Nuclear receptors are one of the promising targets for anti-ischemic drugs. Modulation of estrogen receptors (ERs) and peroxisome proliferator-activated receptors (PPARs) by their ligands is known to exert neuro-, and cardioprotective effects through anti-apoptotic, anti-inflammatory or anti-oxidant action. Recently, it has been shown that the expression of aryl hydrocarbon receptor (AhR) is strongly increased after brain or heart ischemia and evokes an activation of apoptosis or inflammation in injury site. We hypothesize that activation of ERs and PPARs and inhibition of AhR signaling pathways could be a promising strategy to protect the heart and the brain against ischemia. In this Review, we will discuss currently available knowledge on the mechanisms of action of ERs, PPARs and AhR in experimental models of stroke and myocardial infarction and future perspectives to use them as novel targets in cardiovascular diseases.

1. Introduction

Cardiovascular diseases (CVDs), with nearly 18 million estimated deaths globally represent the first cause of death worldwide, whereas stroke ranks second cause of death on this infamous list [1].
Heart and brain are organs with low ability for regeneration and damage of these tissues is irreversible. Myocardial infarction induces apoptosis, necrosis, autophagy and oxidative stress of cardiomyocytes [2]. Similarly, during a stroke, brain cells die mainly because of excitotoxicity, necrosis, apoptosis, autophagy and reactive oxygen species (ROS) overproduction [3]. In clinical practice, the main treatments for myocardial infarction are anti-platelet and thrombolytic therapies, angioplasty, stenting and coronary artery bypass surgery [4]. Instead, the approved treatment for stroke is the reperfusion therapy such as the recombinant tissue plasminogen activator (rt-PA) or endovascular mechanical thrombectomy (MT). In Japan, edaravone, an antioxidative radical scavenger is also used for the treatment of acute ischemic stroke [5]. However, it is needed to underline that the above mentioned treatments must be administered up to 6 h after symptoms onset for thrombolytic agents, up to 6–8 h for endovascular MT and up to 6–72 h for edaravone [5,6]. Since already existing treatments against myocardial infarction and stroke have a very narrow therapeutic window and a long list of contraindications, there is still an urgent need to find a more effective and safer therapeutic strategies to protect myocardial and brain cells against hypoxia/ischemia. Relevantly, epidemiological studies demonstrated that women are better protected against myocardial infarction and stroke than age-matched men [7,8]. However, the risk of myocardial infarction or stroke increases after menopause period [7,9]. The reason could be ascribed to the decreasing level of circulating estrogens, observed during aging, which are not only involved in the regulation of reproductive system, but also in the control of growth, differentiation and function of many tissues including heart and brain [10,11].
Estrogens exert genomic and non-genomic actions binding to one of the three subtypes of estrogen receptors (ERs), estrogen receptor alpha (ERα), estrogen receptor beta (ERβ) and G-protein coupled estrogen receptor-1 (GPER-1). Several in vitro and in vivo studies showed that estrogens acting via ERs exert cardioprotective and neuroprotective effects through anti-apoptotic, anti-inflammatory and anti-oxidant mechanisms [12,13,14,15]. It has been also shown that estrogen receptors interact with peroxisome proliferator-activated receptors (PPARs) [16] and with aryl hydrocarbon receptors (AhR) [17,18]. For example, ERα binds to PPAR response element and represses its transactivation [16]. AhR has been reported to inhibit ERs activity through the binding to the inhibitory xenobiotic response element (iXRE) presented in ERs target genes, squelching of shared coactivators or increased proteasomal degradation of ERs [18]. Therefore, in this Review, we will particularly focus on the role of ERs, AhR and PPARs in cardio-, and neuroprotection during hypoxia/ischemia in preclinical studies.

2. Targeting Estrogen Receptors as Potential Therapeutic Strategy in Myocardial Infarction and Stroke

The two classic nuclear estrogen receptors, ERα and ERβ, are encoded by the ESR1 and ESR2 genes, which are located on chromosome 6 (6q25.1) and 14 (14q23.2), respectively [19,20]. The ERβ shows a high homology to ERα in the DNA-binding domain (more than 95% amino acid identity) and in the C-terminal ligand-binding domain (~55% amino acid identity) [21]. However, due to the splicing mechanism, the two subtypes of receptors may have different isoforms. In human, ERα has at least three isoforms of particular significance, while ERβ has at least five different isoforms [22]. These receptors bind 17β-estradiol (E2) and the other physiological ligands with similar affinity in their ligand-binding domains, except for 17α-estradiol [23]. The cellular relative concentrations of ERα and ERβ significantly influence cell’s response to various ligands. Estrogen can also activate GPER-1, a plasma membrane receptor encoded by the GPER1 gene, which is located on chromosome 7 (7p22.3) [24]. ERα and ERβ are mainly localized in cytosol and nucleus but were also found at the cell membrane, whereas GPER-1 is localized only within cell membrane. ERα, ERβ and GPER-1 are widely expressed in different tissue types and regulate important physiological functions, including reproductive, cardiovascular, muscular, and the central nervous system (CNS) [24,25]. The cellular localization, the mechanisms of action and the protective effects of estrogen receptors during heart and brain ischemia are described on the Figure 1.

2.1. Cellular Localization of Estrogen Receptors in the Heart

Cells expressing ERα and ERβ are present in the neonatal and adult heart [26,27], both in the ventricles and atria [28]. Cardiomyocytes, cardiac fibroblasts [29], endothelial cells [30], vascular smooth muscle cells (VSMCs) [31] and monocytes [32] were shown to express both ERα and ERβ, but recent papers doubt the expression of ERβ in cardiomyocytes and in monocytes [33,34,35,36]. The expression and localization of ERs are modulated in a disease-dependent manner. Indeed, in patients with aortic stenosis or heart failure, the myocardial expression levels of ERα and ERβ were increased [27,37], whereas no change was observed after myocardial infarction (MI) in mice [38]. GPER-1 is strongly expressed in all the chambers of the human heart and also in the atrioventricular sinus, avriet dextra, and aorta, but is not present in the atrioventricular node and apex [39]. GPER-1 is present in cardiac myocytes [40], fibroblasts [41], mast cells [42], VSMCs [43], endothelial cells [44] and monocytes [36].

2.2. Cellular Localization of Estrogen Receptors in the Brain

Cells expressing ERα, ERβ, and GPER-1 are localized throughout the brain, from the most rostral regions of the forebrain to the cerebellum. ERα and ERβ are present in neurons, both at axon terminals, in association with synaptic vesicles, and in dendritic spines [45], in astrocytes, near the spines of pyramidal cells, and in microglia [45,46]. ERβ is more abundant in the cortex and hippocampus than ERα, where it is present in pyramidal neurons and in interneurons. ERs were also found in cerebellum and hypothalamus [47]. GPER-1 is mainly expressed in the cerebral cortex, hippocampus, hypothalamus, striatum, and substantia nigra [47], and is localized in neurons, astrocytes and oligodendrocytes [48]. Interestingly, the expression of ERs and GPER-1 were reported to change across the estrous cycle and to show sex differences [45,48].

2.3. Genomic and Non-Genomic Mechanisms of Action of Estrogen Receptors

Estrogen-dependent signaling can be classically divided into genomic and non-genomic, although it is now widely established that there is a convergence of these pathways. Genomic action of ERs includes binding of ligand to the receptor, dimerization (ERα-ERα, ERβ-ERβ or ERα-ERβ dimers), translocation of dimers from cytoplasm to the nucleus, binding to the estrogen response elements (ERE) in the promoters of target genes and regulating gene expression [49]. Several studies have shown that ERs can also influence gene expression without binding directly to ERE. Indeed, ERs can interact with activator protein 1 (AP-1) transcription factor complex such as Fos and Jun proteins [50]. ERs can also exert non-genomic actions that are too rapid to be accounted for the regulation of gene expression and protein synthesis. Non-genomic actions are mediated by membrane-associated ERs and are related to the activation of pro-survival kinases such as PI3K/Akt and MAPK/ERK, attenuation of the pro-apoptotic pathway JNK [51], and mobilization of intracellular calcium [52]. Relevantly, the activation of these signaling transduction pathways by ERs can influence the genomic action of ERs themselves. Indeed, several kinases regulate the activation of ERs in both ligand-dependent and ligand-independent manner [53]. Among these, MAPK can phosphorylate and activate either ERβ or its associated coregulators, enhancing the genomic action of ERβ [52,53]. Furthermore, depending on which amino acid residues of ERα are phosphorylated, ERα-DNA binding could be increased or inhibited, leading to altered gene transcription [53]. Taking into account the above study, the convergence of non-genomic and genomic actions at multiple levels ensure an extremely high degree of control of gene transcription by ERs.
Localization of ERα and ERβ within mitochondria and in the mitochondrial membrane provides additional actions of estrogens [53]. To date, the mechanisms by which estrogens regulate mitochondrial function are not clearly understood. It has been shown that estrogens regulate transcription of nuclear respiratory factor-1 (NRF-1), peroxisome proliferator-activated receptor-gamma coactivator 1 (PCG-1), or mitochondrial transcription factor A (TFAM) which are critical for mitochondrial biogenesis and mitochondrial electron transport chain complexes [54]. It was also demonstrated that ERs can directly interact with mitochondrial ERE (mtERE) and in turn regulate mtDNA transcription [55].
The membrane GPER-1 receptor, formerly known as the G protein-coupled orphan receptor GPR30, has been shown to induce rapid signaling cascades following estrogens binding. Once activated, GPER-1 initiates several effectors, including c-Src and adenylate cyclase, which leads to increase of cAMP level and to the activation of prosurvival MAPK, PI-3K/Akt, and CREB pathways [56]. This mechanism is observed in neurons and in cardiomyocytes [57,58,59]. Interestingly, in astrocytes the activation of GPER-1 is associated with cell death via the activation of Phospholipase C (PLC) pathway and rise in intracellular calcium levels [60].
In addition, estrogen signaling is also tightly connected to epigenetic mechanisms. Several studies showed that estrogens could either induce demethylation of DNA resulting in epigenetic upregulation of downstream targets or methylation of DNA with subsequent downregulation of target genes [52]. Interestingly, the methylation level of Esr1 decreased in female but not in male rats following middle cerebral artery occlusion (MCAO) [61]. This results were confirmed in women undergoing large-artery and cardio-embolic stroke who showed lower ESR1 methylation levels in peripheral blood compared to the controls [62].

2.4. The Role of Estrogen Receptors in Myocardial Infarction

2.4.1. ERs Modulation in Experimental Models of Myocardial Infarction

To assess the specific role of ERs in the pathophysiology of MI, several studies using ERs knock-out (KO) mouse or transgenic mouse models with ERs-overexpression have been carried out. Study performed on male and female ERα-KO mice, subjected to global myocardial ischemia/reperfusion (I/R), showed controversial results. Male ERα-KO mice subjected to global myocardial I/R, developed more severe cardiac damage, had a higher incidence of ventricular arrhythmias and showed a marked mitochondrial damage than wild-type (WT) mice, suggesting a cardioprotective role of ERα [63]. There results were not confirmed by another study, where no difference between ERα-KO male mice undergoing I/R and WT was observed. However, the female ERα-KO mice exhibited less functional recovery after I/R than WT female [64]. On the contrary, isolated Langendorff hearts from ERα-KO female mice exhibited I/R injury similar to that observed in WT females [65].
Studies in ERβ-KO mice suggest that this receptor plays a relevant role in the protection observed in the female heart. Indeed, isolated Langendorff hearts from ERβ-KO female mice exhibited significantly less functional recovery than WT female mice and were similar to WT male mice [65,66]. These ex-vivo results were corroborated by in vivo studies. At 8 weeks after permanent MI, ERβ-KO ovariectomized (OVX) female mice showed increased mortality and higher expression of pro-atrial natriuretic peptide (ANP) in left ventricle and in serum compared to WT mice [67]. In line, at two weeks after MI induced by left anterior descending (LAD) coronary artery ligation, estrogen treatment resulted in increased infarct size in ERβ-KO OVX mice respect to ERα-KO and WT OVX mice [68]. One possible mechanism of cardioprotection related to ERβ is the anti-apoptotic action. Indeed, ERβ-KO female mice undergoing MI had affected myocardial PI3K and Akt activation, which was associated with increased expression of caspase-3 and -8, as well as decreased Bcl-2 levels compared with wild-type (WT) mice. These effects were not observed in ERβ-KO male mice [66]. Luo and colleagues [69] observed that in ERα-KO, ERβ-KO and ERα/ERβ-double KO female mice the infarct volume was larger than in WT mice. Moreover, KO mice had reduced mitochondrial activity.
The role of ERs was also evaluated in inducible transgenic mice with cardiomyocyte-specific ERs-overexpression, making possible to discriminate the effect of cardiac ERα and ERβ from ERs-mediated systemic effects. In female mice ERα-overexpression improved functional myocardial adaptation, reduced collagen type-I and -III gene expression and collagen deposition, and induced phosphorylation of JNK signaling pathway 2 weeks after MI induced by permanent LAD ligation. Furthermore, ERα-overexpression enhanced angiogenesis, lymphangiogenesis and neovascularization in the peri-infarct area of female and male mice [70]. A recent study demonstrated that improved neovascularization after MI could be promoted by the activation of ERα in endothelial progenitor cells (EPCs) via enhancing their homing and angiogenic capacity [71]. Relevantly, transplantation of estrogen-stimulated EPCs preserved cardiac function after MI, but this effect disappeared in EPCs pre-conditioned with the ERα antagonist methyl-piperidino-pyrazole (MMP) [71] or derived from ERα-KO mice and, to a lower extent, from ERβ-KO mice [72].
It has been demonstrated that transgenic ERβ (Tg-ERβ) mice were more protected against MI than WT mice. At 1 week after MI induced by coronary artery ligation, Tg-ERβ mice had improved cardiac function with reduced echocardiographic end diastole and end systolic posterior wall thickness, increased ejection fraction and fractional shortening. Furthermore, cardiac collagen deposition and mRNA expression levels of collagen I, α-SMA, and TGF-β were significantly lower in Tg-ERβ than in WT mice [73]. Similarly, inducible ERβ-overexpression was associated with attenuated left ventricular (LV) dilatation, smaller increase in heart weight, improved systolic and diastolic function in MI mice of both sexes at 2 weeks from LAD ligation. These effects were associated with less reduction in diastolic Ca2+-reuptake into the sarcoplasmic reticulum [74]. Taken together, these results suggest a key role of ERs expressed in cardiomyocytes in the cardioprotective effects elicited by estrogen in both sexes.
Further insights on the specific ERs-mediated beneficial effects of estrogen have resulted from the development of selective ERα and ERβ agonists. Administration of the ERα-agonist propyl-pyrazole-triol (PPT) was cardioprotective in several models of MI. In intact or OVX rabbits subjected to cardiac I/R, treatment with 17β-estradiol (E2) or PPT decreased infarct size, release of cardiac-specific troponin-I and plasma level of C-reactive protein at the end of the 4 h reperfusion period [75]. Additionally, in isolated Langendorff hearts from adult or aged rats, PPT, improved ischemic tolerance via nongenomic ER signaling. PPT preserved PKCε levels at the nucleus and mitochondria, and enhanced the expression of PKCε anchoring protein RACK2 [76]. In sedentary OVX female rats, treatment of 2 weeks with PPT inhibited the acute I/R-induced increase of creatine kinase-(muscle–brain) (CK-MB), plasminogen activator inhibitor-1 (PAI-1) and TNF-α plasma concentrations, but not of IL-6, IL-8 and cardiac-specific troponin I. Instead, PPT failed to improve inflammatory cell infiltration, disorganization of cardiac muscle fibers and the microscopic damage score [77]. In ex-vivo hearts isolated from female OVX rats and subjected to MI by coronary ligation, PPT induced Akt and eNOS phosphorylation. These effects were abolished by co-incubation with PI3K inhibitor (LY294002) [78]. This result suggested that the ERα-activated PI3K/Akt signaling is involved in the modulation of eNOS activity. In the study with the same experimental design, PPT, but not diaryl-propio-nitrile (DPN), resulted in attenuated cofilin phosphorylation, suggesting that ERα, but not ERβ, mediated the inhibitory effect of estrogen on cofilin phosphorylation and thus on cardiac fibrosis after MI [78]. Similarly to PPT, ERA-45, another ERα-selective agonist, administered for 5 days prior to I/R in OVX rats reduced infarct size, neutrophil infiltration and oxidative stress at the end of the 2 h reperfusion period [79].
The role of ERβ was largely investigated using the ERβ agonist DPN. In isolated and Langendorff perfused hearts of OVX mice subjected to normothermic global ischemia, pre-treatment with DPN for 2 weeks induced a better functional recovery, decreased infarct size, upregulated expression of protective genes (heat shock protein 70, the antiapoptotic genes, growth arrest and DNA damage 45β, and cyclooxygenase 2) and increased the S-nitrosylation (SNO) of cardiac proteins [80,81]. The lack of these effects in ERβ-KO mice or in mice pre-treated with L-NAME, a NOS inhibitor, suggests that estrogens protect the heart through activation of ERβ and NO/SNO signaling [81]. On the contrary, in isolated and Langendorff perfused hearts of adult, aged, or aged OVX female Fischer 344 subjected to I/R, acute treatment of DPN had no effect on functional recovery [33]. Recently, it has been shown that treatment with DPN, for 14 days before and 2 days after LAD ligation in MI male mice, reduced the infarct size and the serum levels of myocardial enzymes (CK, CK-MB and lactate dehydrogenase) leading to cardiac function improvement. In parallel, DPN protected cardiomyocytes from oxidative damage (reduced the protein levels of iNOS and MDA, and increased SOD and GPX) and apoptosis (increased Bcl-2 and reduced cleaved caspase-3), exerting its cardioprotective function through the Notch1/PI3K/Akt signaling pathways [82].
Although ERs are well known to be cardioprotective, their hormone-dependent action on peripheral tissue is a strong contraindication to introduce them as a treatment of MI. Therefore; selective estrogen receptor modulators (SERMs) could be a good alternative for estrogens to contrast this disease. SERMs are compounds that act as ERs agonists or antagonist in tissue-dependent manner [83]. For example, SERMs representatives tamoxifen, raloxifene and bazedoxifene may act as ERs agonists in the cardiovascular system [84,85,86], but they antagonize ERs in breast tissue [7,87]. Rayabarapu and Patel [88] showed that tamoxifen and raloxifene significantly reduced isoproterenol-induced infarction and hypertrophy in rats. Cardioprotective effect of raloxifene was also observed by Chung and colleagues [89] who showed that long-term treatment with the SERM protects OVX rats against MI-induced arrhythmias and cardiomyocytes apoptosis via suppression of nuclear factor-kappa B (NF-κB).

2.4.2. GPER-1 Modulation in Experimental Models of Myocardial Infarction

GPER-1 was detected in human heart and its expression could be modulated under pathological conditions. In isolated and Langendorff perfused hearts of rats, hypoxia resulted in about 2.4-fold increase in Gper1 mRNA compared to basal conditions [39,59]. Furthermore, within the first 30 min of reoxygenation there was a substantial increase of Gper1 mRNA expression, reaching 10.3 fold under basal conditions [39]. The pre-treatment with G1, a selective agonist of GPER-1, significantly reduced infarct size and improved the functional recovery of the left ventricular developed pressure (LVEDP). These effects were lost when hearts were pre-treated with GPER-1 antibody [39]. Other studies showed similar results of G1 in Langendorff perfused hearts of male and female rats or male mice, and demonstrated that G1 exerts its protective effects through PI3K/Akt [58] and ERK pathway [90]. The role of ERK, but not of PI3K/Akt, in the GPER-1 mediated cardioprotection against hypoxia in Langendorff perfused hearts was confirmed using ERs-KO mice. The authors suggest that estrogens, binding to GPER-1, may initially trigger translocation of protein kinase C (PKC), which could directly or via activation of MEK1/2/ERK1/2 pathway increase phosphorylation of GSK-3β. Deactivation of GSK-3β results in the inhibition of mitochondrial permeability transition pore (mPTP) opening [91]. This last effect is highly relevant, since the opening of mPTP plays a crucial role in the mechanism of cell death after ischemia/reperfusion [92].
In addition to ex-vivo studies, the role of GPER-1 was also evaluated in in vivo studies. In OVX rats subjected to permanent MI, 4 weeks of treatment with G1 improved the long-term MI-induced remodeling, reducing cardiac hypertrophy and fibrosis through phosphorylation and activation of AKT and eNOS [78]. In OVX mice subjected to LAD ligation, G1 reduced myocardial infarcted area and cardiac fibrosis, inhibited apoptosis via stimulation of PI3K/Akt pathway and diminished inflammation via decreasing TNF-α and increasing IL-10 levels [93]. The cardiac induction of the anti-inflammatory cytokine IL-10 was also observed in OVX diabetic rats treated with E2 or tamoxifen [94]. In this study, however the effect was GPER-1 independent.
A recent work showed that acute estrogen treatment induces cardioprotective effects in male and OVX female rats subjected to cardiac I/R by GPER-1 activation. At 3 h from reperfusion, estrogen reduced the percentage of area at risk, increased mitochondrial membrane potential and Ca2+ retention capacity, and decreased the production of ROS. The estrogen-mediated cardioprotective effect was related to activation of the MEK/ERK, deactivation of GSK-3β and to the delay of mPTP opening. Moreover, estrogen reduced mitophagy via the PINK1/Parkin pathway involving LC3I, LC3II and p62 proteins. The role of GPER-1 was pointed to the lack of these effects in presence of G-15, a GPER-1-antagonist [95]. In isolated and perfused hearts subjected to I/R, G1 reduced infarct size and improved contractile recovery in both normotensive and hypertensive female rats at 2 h from reperfusion. Relevantly, these cardioprotective effects were abolished by specific inhibitors of PI3K/Akt-eNOS-MitoKATP channels and by DAPT. DAPT is an inhibitor of the γ-secretase, an enzyme required for the Notch1 cleavage and activation. The lack of protective effect of G1 in presence of DAPT was also observed in cardiac myoblasts H9c2 cells subjected to I/R. These results suggested that G1 counteracted cardiac damage through activation of PI3K/Akt/NOS/MitoKATP channel and Notch1 pathways [96].

2.5. The Role of Estrogen Receptors in Stroke

2.5.1. ERs Modulation in Experimental Models of Stroke

It is well known that estrogens exert anti-apoptotic, anti-oxidative and anti-inflammatory actions in the CNS [14,97,98]. The direct effect of E2 on microglia is well documented in many experiments in vitro. For example, E2 were able to reduce the expression of the pro-inflammatory mediators Il1b and Ccl5 and to increase the expression of the anti-inflammatory cytokine Il10 in immortalized microglial BV-2 cells undergoing hypoxia [99]. Furthermore, the pre-treatment of LPS-stimulated microglial N9 cells with E2 increased the IL-10 and decreased the TNF-α and interferon-γ release from these cells [100]. In vivo experiments using ERs-KO mice have suggested that ERα and ERβ play distinct roles in neuroprotection. The first research ruled out a role of ERα in the estrogen’s neuroprotective activity. Indeed, neurological function and ischemic volume were similar in ERα-KO and WT mice subjected to transient cerebral ischemia [101]. However, this study had some limitations, since the mice utilized were gonad-intact and thus the estradiol concentrations in ERα-KO mice were dramatically higher than in WT mice. On the contrary, in OVX mice subjected to permanent cerebral ischemia and treated with E2, deletion of ERα resulted in abolishment of neuroprotective effects, whereas in ERβ-KO mice neuroprotection was maintained [102,103]. Furthermore, the expression of ERα and ERβ was differentially modulated by ischemia and E2 treatment [103,104], leading the authors to speculate that ERα may be fundamental in the protection against cell death, while ERβ may play a role in regeneration and neurogenesis. This hypothesis is not fully shared. Indeed, the silencing of ERβ via intracerebroventricular (i.c.v.) injection of ERβ-antisense inhibited the E2-mediated hippocampal protection in OVX rats subjected to transient cerebral ischemia [105]. Estrogens can also exert protective effect on very early stages of ischemic injury. A recent study showed that estrogen or DPN o PPT pretreatment protected brain endothelial (bEnd3) cells against oxygen and glucose deprivation/reoxygenation (OGD-R) injury. DPN and PPT increased OGD-downregulated levels of occludin and claudin-5. Silencing of ERα or ERβ with the use of specific siRNAs completely reversed the effects of DPN or PPT on the outcomes of OGD-R [106]. These data strongly suggest an involvement of estrogen receptors in maintaining BBB function during the stroke.
Apart from the study carried out in ERs-KO mice or cells, the neuroprotective potential of ERs is often confirmed by the use of specific ERs agonist. In OVX rats subjected to transient global cerebral ischemia, ERα selective agonists PPT elicited a pronounced protection of CA1 pyramidal neurons in approximately 40–50% of treated ischemic rats [104]. This result was in contrast with two other studies showing a lack of neuroprotective action of PPT in OVX mice with transient global ischemia induced by bilateral carotid artery occlusion [107] and in male mice with transient global ischemia induced by cardiac arrest [108]. This discrepancy could be explained by the different dose utilized or differences in performing ischemia. A recent study demonstrated that metastasis-associated protein 1 (MTA1), which is a chromatin modifier and transcriptional regulator, could be a factor linking ERα with apoptosis. The increase of MTA1 expression in mice after transient middle cerebral artery occlusion (tMCAO) promoted interactions between ERα and anti-apoptotic Bcl-2 which in turn diminished ischemia-induced brain damage [109]. In line, endogenous estrogen in proestrus protected female rats against I/R injury by an increase of ERα-dependent Bcl-2 expression [110]. ERα could be also involved in the neuroprotection mediated by the inhibition of miR-181a. Indeed, miR-181a inhibition led to increase of Esr1 expression that in turn resulted in decrease in infarct volume and improved neurological deficit score in OVX mice subjected to tMCAO. Furthermore, it decreased death in female astrocytes cell culture subjected to glucose deprivation [111].
Not only ERα agonists but also ERβ agonists may protect the brain against ischemia. In OVX mice with transient global ischemia induced by bilateral carotid artery occlusion, ERβ agonist DPN significantly reduced ischemic damage in the caudate nucleus and in the CA1 region compared with vehicle controls [107]. Similarly, in male mice with transient global ischemia induced by cardiac arrest, DPN reduced neuronal injury in the striatum and in CA1 field [108]. The periodic DPN treatment (every 48 h) improved post-ischemic learning and memory in OVX rats subjected to transient cerebral ischemia [105]. More recent studies showed that DPN diminished I/R evoked injury in OVX mice via inhibition of microglia, astrocytes and NF-κB-mediated neuroinflammation [112,113]. Moreover, specific ERβ agonist AC-131 helped to recover the neurological function in male rats with permanent focal ischemia induced by photothrombosis [114]. In OVX rats subjected to transient cerebral ischemia, specific ERβ agonist WAY 200070-3 elicited pronounced protection of CA1 pyramidal neurons in approximately 40–50% of treated ischemic rats [104]. Interesting results were also obtained in OVX mice with transient focal brain ischemia where DPN reduced the extravasation of endogenous immunoglobulin G (IgG), vasogenic edema, and the infarct volume [115].
Despite the well-documented, beneficial action of estrogens in experimental models of stroke their clinical use is limited because of the hormone-dependent action on peripheral tissue. Therefore, a good alternative for the use of estrogens in stroke treatment could be the use of selective estrogen receptor modulators. There are already some data showing that SERMs mimic the action of estradiol after experimental ischemia avoiding hormone-dependent risks. In rats subjected to transient or permanent MCAO (pMCAO) tamoxifen significantly reduced the infarct size and protected neurons against ischemia [116,117]. In OVX rats with pMCAO, tamoxifen reduced MCAO-elevated superoxide anion production, oxidative damage and caspase-3 activation, through increasing the levels of manganese SOD (MnSOD) and attenuating the elevation of pERK1/2 [118]. However, the involvement of ERs in the mechanism of action of tamoxifen is still unclear. In rats with tMCAO, the neuroprotection by tamoxifen was maintained when co-administered with the estrogen receptor antagonist ICI 182,780, suggesting a key role of its antioxidant activity but not of estrogen receptors [119]. In contrast, tamoxifen enhanced neuronal survival in OVX rats with tMCAO by modulating ERα-36, a variant of ERα, and activating the MAPK/ERK signaling pathway [120]. Similarly to tamoxifen also raloxifene and bazedoxifene protected mouse hippocampal and neocortical cells against hypoxia via ERα, but not ERβ and GPER-1 [121,122]. Raloxifene and tamoxifen preserved spine density in the cortex of OVX rats, but only raloxifene enhanced neurogenesis after tMCAO. The Authors suggested that similarly to estrogens, the SERMs may enhance excitatory synapse formation in cortical neurons via a non-genomic ER-mediated mechanism involving up-regulation of AMPA receptor by Akt and ERK signaling pathways [123]. Another SERM representative bazedoxifene mimicked action of estradiol in male rats with tMCAO [124] and improved neurological function via ERα and ERβ, but not of GPER-1 [125]. Similar effect was observed in diabetic rats after tMCAO [126], suggesting the involvement of ERs in the neuroprotective effects of bazedoxifene.

2.5.2. GPER-1 Modulation in Experimental Models of Stroke

GPER-1 plays a relevant role in the acute neuroprotective effects of estrogen against ischemic injury. It was shown that the magnitude of neuroprotection observed in G1 treated OVX mice with tMCAO was indistinguishable from estrogen treated mice [127]. GPER-1 is involved in inhibition of neuronal deficit and inflammation after ischemic injury in OVX mice with tMCAO. Activation of this receptor reduced the infarct volume, improved the neurological deficit and alleviated neuronal injuries via inhibition of microglia activation and cytokine’s release [128,129]. In transient global ischemia in OVX rats, G1 inhibited inflammation decreasing the expression of NLRP3-ASC-caspase 1 inflammasome and IL-1β as well as NF-κB signaling. Intriguingly, G1 caused a robust elevation of endogenous anti-inflammatory factor IL1RA in neurons, likely enhancing CREB phosphorylation. Relevantly, IL1RA antisense oligodeoxynucleotide abolished the anti-inflammatory, neuroprotective, and anti-apoptotic effects of G1 [130]. The anti-inflammatory activity of GPER-1 was also corroborated by the fact that GPER-1 antagonist G-15 reversed the effect of estrogen and abolished the reduced serum level of IL-1β and TNF-α in rats with global cerebral ischemia [128].
In additions to anti-inflammatory effects, GPER-1 could mediate neuroprotection through different mechanisms. Indeed, G1 reduced the infarct volume and neurological deficit score in OVX mice with tMCAO, but these effects were abolished by simultaneous administration of PD98059, an ERK1/2 inhibitor, suggesting that ERK activation by GPER-1 is required to protect the neurons [131]. The involvement of GPER-1 in the neuroprotective effect of E2 was not confirmed in organotypic hippocampal slice cultures exposed to ischemia. Indeed, GPER-1 activation only facilitated the acute E2 mediated neuroprotection after OGD, since GPER-1 antagonist G-15 effected a slight but not significant reduction in neuroprotection elicited by E2 [132]. However, G1 could play an important role in protecting the cerebrovasculature against ischemia. In fact, in ex-vivo study, G1 significantly improved the ATP-vasodilation in both male and female rat vessels with hypoxia/reoxygenation (H/R) injury similar to pre-H/R dilations [133]. GPER-1 could also counteract BBB disruption after brain ischemia. In OVX rats with induced transient global cerebral ischemia, G1 decreased IgG extravasation and increased the levels of tight junctions (occludin and claudin-5) in the CA1 [134]. In male mice with transient global ischemia induced by cardiac arrest, G1 reduced neuronal injury in the hippocampal CA1 region and striatum via increased expression of the small conductance calcium-activated potassium channel 2 [135]. Finally, G1 reduced the neurological deficit scores and the infarct volume in WT mice with tMCAO, but these effects were completely blocked in conventional GPER-1 KO mice and only partially attenuated in astrocytic or neuronal GPER-1 KO mice. The authors suggest that astrocytic GPER-1 could play a key role in neuroprotection, via restoration of autophagic balance through the activation of p38 MAPK signaling pathway [136].
One research study showed sex-dependent protection of GPER-1 agonist. In male rats with tMCAO, G1 worsened functional outcomes and increased infarct volume, which was associated with higher level of cleaved caspase-3 in peri-infarct neurons. On the contrary, G-15 improved functional outcomes and reduced infarct volume after stroke in males. In OVX female mice, G1 reduced neurological deficit, apoptosis, and infarct volume, but had no effect in intact females [137]. The authors hypothesized that after cerebral ischemia GPER-1 may become the predominant estrogen receptor expressed in brain of male mice, whereas the ERα predominates in female mice.

3. Targeting of Peroxisome Proliferator-Activated Receptors (PPARs) as Potential Therapeutic Strategy in Myocardial Infarction and Stroke

PPARs are ligand-activated nuclear transcription factors, belonging to nuclear receptor superfamily and including PPAR-α, PPAR-β/δ and PPAR-γ subtypes [138]. They are encoded by PPARA, PPARD, PPARG, respectively. The PPARA gene is located on human chromosome 22q12.2-13.1 [139], and its expression is the highest in high-energy requiring tissues such as liver, heart, and skeletal muscle where it acts as a major regulator of the mitochondrial, peroxisomal, and microsomal fatty acid oxidation systems [140]. The PPARD gene is located on human chromosome 6p21.1-21.2 [139] and is highly expressed in adipose tissue, liver, heart, skeletal muscle, brain, kidney, colon, and vasculature [141]. Activation of PPAR-β/δ induces lipid catabolism and energy dissipation. Unlike PPAR-γ and PPAR-α, PPAR-β/δ is not easily targeted by currently available drugs because of its ubiquitous expression. Finally, the PPARG gene is located on human chromosome 3p25 [139] and is widely expressed in brown and white adipose tissues, and in the immune system, where it regulates glucose metabolism, cell differentiation, and immune and inflammatory responses [142]. The cellular localization, the mechanisms of action and the protective effects of PPARs during heart and brain ischemia are described on the Figure 2.

3.1. Cellular Localization of PPARs in the Heart

All three PPAR isoforms are found in the adult heart, with various levels of expression. PPAR-α and PPAR-δ are highly expressed at similar levels to those in other metabolically active tissues, such as the liver and skeletal muscle [143], while PPAR-γ is expressed at very low levels, approximately 2% of those in adipose tissue [144]. However, the cardiac expression of PPARs could change in pathological conditions. Indeed, PPAR-α expression and activity are diminished in cardiac tissue of pressure overload–induced hypertrophy murine model [145], in hypertrophic heart of spontaneously hypertensive stroke-prone rats [146] and in hypoxic cardiomyocytes [147] leading to a reduction in the capacity for fatty acid oxidation and increased rates of glucose utilization. Rats with diabetic cardiomyopathy presents decreased cardiac levels of PPAR-α and PPAR-β/δ [148,149], while those of PPAR-γ are increased [149]. Instead, the PPAR-α and PPAR-β/δ proteins level didn’t change after myocardial infarction in rats, while PPAR-γ showed a significant increase in the infarcted area [150].
Concerning the cellular localization of PPARs, several differences were observed among the three PPAR isotypes. In neonatal and adult rat cardiomyocytes, both PPAR-α and PPAR-β/δ were expressed comparably, while PPAR-γ is barely detectable [151]. However, PPAR-γ protein expression was strikingly increased in cardiomyocytes in the infarcted area of rat underwent LAD ligation [150]. In cardiac fibroblasts and myofibroblasts, PPAR-β/δ is higher expressed and more biologically active than PPAR-α and PPAR-γ [150,152]. Like the cardiomyocytes, the protein content of PPAR-γ was increased in connective tissue and in fibroblasts in the infarcted area of rats underwent LAD ligation [150]. Finally, all the three PPAR isotypes have been reported in the cells of vascular wall, such as smooth muscle cells, monocyte/macrophages and endothelial cells [153,154,155].

3.2. Cellular Localization of PPARs in the Brain

In adult murines, PPARs are ubiquitously expressed in all brain regions [156,157], although recent evidence has demonstrated brain region- and cell type-dependent differences in PPARs subtypes expression. As reported, mapping the PPARs isotype mRNA and protein in the adult mouse, the general order of abundance across all brain regions was PPAR-β/δ > PPAR-α ≥ PPAR-γ [157]. PPAR-α is strongly expressed in neurons, in the cell body and processes of astrocytes but weakly in microglia of the ventral tegmental area (VTA), prefrontal cortex (PFC), nucleus accumbens (NAC), or amygdala (AMY). Similarly to PPAR-α, PPAR-γ is more expressed in neurons than in astrocytes with the highest level in the NAC and the lowest in the PFC. Instead, PPAR-γ does not colocalize with microglia in the adult mouse brain. Finally, PPAR-β/δ is primarily localized in the nucleus of neurons, while it is not expressed by astrocytes in grey matter and in microglia [157]. In human brain, the pattern of PPARs expression is similar to the mouse brain [157]. In particular, PPAR-α colocalized with all cell types, while PPAR-β/δ and PPAR-γ colocalized with neurons and astrocytes, but not with microglia. Although the expression of PPAR-γ is not detectable in physiological condition in microglia, PPAR-γ may be tightly regulated and dependent on microglial functional state. Indeed, PPAR-γ expression was induced in microglia after LPS treatment or specific agonists [157]. On the contrary, PPAR-β/δ still was not expressed by microglia after lipopolysaccharides (LPS) treatment [157]. PPAR-γ and PPAR-β/δ are also expressed in oligodendrocytes, promoting survival and differentiation of precursor cells [158], while PPAR-α, PPAR-γ and PPAR-β/δ are also expressed in brain capillary endothelial cells, suggesting an involvement of the receptors in regulation of BBB [159,160,161].

3.3. Mechanisms of Action of Peroxisome Proliferator-Activated Receptors

Differently to the estrogen receptors, which create homodimers, PPARs form heterodimers with the retinoid X receptor (RXR) [162]. In the absence of ligand, PPARs/RXR heterodimer is bound to multicomponent repressors with histone deacetylase activity, such as the nuclear receptor corepressor (NCoR) and silencing mediator of retinoid and thyroid hormone receptor (SMRT), thereby inhibiting gene transcription [163]. A ligand binding triggers dissociation of corepressors from PPARs/RXR heterodimer, and recruitment of co-activators. The entire complex binds to peroxisome proliferator response elements (PPREs) located in the promoter region of target genes resulting in initiation of gene’s transcription [163].
However, it was also demonstrated that phosphorylation can modulate PPARs activity [164,165]. The protein kinase A (PKA)-induced phosphorylation of PPARs has a stimulatory effect on transcription in a ligand-independent and ligand-dependent manner [165], while MAPK and Brc kinases-induced phosphorylation deactivates PPAR-γ and reduce basal and ligand-dependent transcriptional activity [164,166]. PPARs can also inhibit gene expression in a DNA binding-independent manner, interfering with other transcription factors. First, PPARs could repress transcription in a ligand-dependent manner by competition for a limiting pool of co-activators with NF-kB and activator protein-1 (AP-1), leading attenuation of NF-kB and AP-1 target gene expression [167]. Second, PPARs could inhibit the expression of proinflammatory genes by direct interaction with NF-kB, Smad-3, AP-1, signal transducers and activators of transcription (STAT) proteins, preventing the binding to their response elements [167]. Third, PPARs can contribute to transrepression by either preventing the clearance of co-repressor complexes [168] or releasing co-repressors, which could allow co-repressor binding to NF-kB, eventually inhibiting NF-kB target gene expression [167].

3.4. The Modulation of PPARs in Experimental Models of Myocardial Infarction

PPARs are mainly involved in energy homeostasis and metabolic function, however there is increasing body of evidence on their cardioprotective potential [169,170]. Indeed, it has been shown that the PPAR-γ agonist-pioglitazone protected the mouse heart from myocardial injury by antagonizing monocyte/macrophage-mediated inflammation and promoting cardiac healing [171]. Moreover, theacrine (alkaloid derived from Chinese tea) inhibited myocardial infarction-evoked fibrosis via stimulation of PPARG and SIRT3 expression [172]. Similarly, plant-derived chrysin attenuated the MI-induced fibrosis in rats via up-regulation of PPAR-γ, down-regulation of matrix metalloproteinases-2 and -9 and inhibition of the NF-κB pathway [173]. Another natural compounds such as qiliqiangxin, apigenin or curcumin also protected cardiomyocytes against myocardial infarction via activation of PPAR-γ [174,175,176]. Shen and colleagues [177] showed that knock-out of PPAR-γ in mice myeloid cells led to cardiac hypertrophy and increased myocardial infarct size. This was correlated with induction of oxidative stress and cardiac inflammation. Interestingly vitamin D exerted anti-inflammatory and anti-oxidant effects in rat model of myocardial infarction via PPAR-γ [178].
Growing body of evidence demonstrates importance of miRNAs in the regulation of gene expression. It has been shown that PPAR-γ is a target gene of miR-130. In cardiomyocytes cell line H9C2 undergoing hypoxia, downregulation of miR-130 promoted PPAR-γ-mediated cardioprotection by inhibiting NF-κB-mediated inflammation and TGF-β1-mediated fibrosis. These results were also confirmed in vivo [179]. Zhu and colleagues [180] showed that miR-292-5p downregulation protects mice against myocardial ischemia through activation of the PPARα/PPAR-γ signaling pathway.
Platelet activation is one of the major pathophysiological mechanisms that underlie I/R injury. It has been demonstrated that humans and rodents undergoing acute myocardial infarction have lower level of PPAR-γ in platelets following by mitophagy activation and an increase in mitochondrial function. Improved mitochondrial function in turn lead to platelet aggregation and formation of microthrombus thus, the inhibition of this process may cause protective effects. Indeed, Zhou and colleagues [181] showed that melatonin inhibited platelet activation via increasing of the PPAR-γ level, blocking mitophagy, platelet hyperactivity, and cardiac I/R injury. Another therapeutic strategy in myocardial infarction is related to the replacement of dead cardiomyocytes or regeneration of the heart muscle. Although, there are studies showing positive effects of stem cells administration to the injured tissue [182,183], stem cells cannot ensure full regeneration of the heart muscle and in turn recovery of heart functions. Therefore, it is crucial to combine stem cells therapy with pharmacological treatments. Hou and colleagues [184] showed that rats simultaneously treated with mesenchymal stem cells (MSCs) and pioglitazone after MI significantly improved cardiac functions via stimulation of PPAR-γ-regulated Cx43 expression and inhibition of TGF-β1/Smad signaling pathway. This kind of strategy seems to be promising considering also that PPAR-γ is crucial for cardiomyocyte differentiation [185].
Apart from PPAR-γ, also PPAR-α seems to be involved in cardioprotection against myocardial infarction. Indeed, it has been shown that natural compound called raspberry ketone suppressed isoproterenol-induced cardiac infarct size, oxidative stress and inflammation in rats via activation of PPARα [186]. Administration of the PPAR-α agonist WY-14643 inhibited myocardial infarction and reperfusion-induced arrhythmia in rat model. PPAR-α activation protected also H9C2 cells against hypoxia-reoxygenation via increased Ucp3 expression and attenuation of ROS production [187]. However, there are also data showing that overexpression of PPAR-α in mice heart led to cardiomyocytes cell death during ischemia/reperfusion [188]. Similarly, conditional overexpression of PPAR-β/δ in cardiac endothelial cells failed the exert protection in mice with myocardial infarction [189]. Therefore, it is very important, to obtain the proper balance of PPAR-α/β/δ activation in the different cardiac cell types to observe beneficial effects on the outcome in ischemic heart disease.

3.5. The Modulation of PPARs in Experimental Models of Stroke

PPARs are highly expressed in the brain and play a crucial role in the CNS. It has been shown that PPAR-γ and its coactivator PGC-1α is engaged in cell differentiation and mitochondria biogenesis as well as in neurodegeneration and neuroinflammation [190]. PPAR-α was shown to influence metabolism of amyloid beta precursor protein (APP) and phosphorylation of Tau protein [191]. PPAR-β/δ has a role in differentiation of cells, lipid metabolism and myelination in CNS [192]. Considering that PPARs are involved in protecting the brain against neuroinflammation, neurodegeneration and oxidative stress, the use of PPARs as a target for stroke treatment has been elucidated by many researchers.
Promising results come from clinical trials on patients undergoing stroke who were treated with pioglitazone. In these patients, lowered risk of recurrent stroke and reduced number of cardiovascular deaths were observed [193,194]. In experimental study, mice lacking PPAR-γ and subjected to MCAO exhibited higher neuronal cell death than control mice. Apoptotic cell death was accompanied by an increase in caspase-3 and Bcl-2 associated X protein levels and reinforcement of endoplasmic reticulum (ER) stress [195]. Oleic acid (OA) is endogenous ligand of PPAR-γ released from the brain phospholipids after cerebral ischemia. Song and colleagues [196] showed that OA has a neuroprotective capacity in the mouse model of stroke, which might be related to its anti-inflammatory actions through PPAR-γ. Han and colleagues [197] showed that treatment with the PPAR-γ agonist rosiglitazone, improves long-term white matter integrity after cerebral ischemia, at least, in part, by promoting oligodendrogenesis and facilitating microglial polarization toward the beneficial M2 phenotype. Another study conducted in the rat model of cerebral ischemia has shown that rosiglitazone decreased ischemia-induced levels of TNF-α, IL-1β and IL-6 and it induced ischemia-downregulated IL-10 level [198]. The effects of rosiglitazone were inhibited by PPARγ antagonist GW9662. Moreover, 1, 25-dihydroxyvitamin D3, activating PPAR-γ, can maintain MCAO-disrupted BBB permeability, via PPAR-γ-dependent inhibition of neuro-inflammation [199]. Pan and colleagues [200] showed that a natural anti-oxidant from the Chinese plant Hopea hainanensis Malibatol A exerts anti-inflammatory potential in MCAO mice in a PPAR-γ-dependent manner. Another in vivo study indicated that PPARα agonist octadecylpropyl sulfamide reversed the memory and motor deficits, decreased the activation of microglia and astrocytes, and reduced neurodegeneration in mice subjected to hypoxia/ischemia [201].
The only pharmacological treatment against ischemic stroke is rt-PA administration, within 3–4.5 h from the first stroke symptoms. However, delayed rt-PA administration may lead to BBB breakdown and in turn to hemorrhagic transformation. Recently, it has been demonstrated that rosiglitazone prevented MCAO-induced BBB damage and blocked hemorrhagic transformation in delayed tPA-treated mice by activating PPAR-γ [202]. Similarly, pioglitazone prevented hemorrhagic infarction after transient focal ischemia in diabetic mice via inhibition of inflammation and oxidative stress induced by reperfusion [203].
The drug repurposing strategy identifies available drugs exerting cerebral protective effects. Among these, telmisartan, an antihypertensive agent protected rat brain against MCAO via PPAR-γ and PPAR-γ regulated factor Egr-1 [204]. Raloxifene and bazedoxifene, used in the treatment of postmenopausal osteoporosis, protected neurons against hypoxia partially via up-regulation of PPAR-γ [122]. Mifepristone used for the termination of early pregnancy exerted neuroprotective potential acting as PPAR-γ agonist and inhibiting inflammatory cytokines and metalloproteinases in rats subjected to MCAO [205]. Aleglitazar, a dual PPAR-α/γ agonist, approved for the treatment of diabetes, attenuated MCAO-induced inflammation via inhibition of microglia migration, phagocytosis, and release of cytokines [206]. Another PPAR-α/γ dual agonist, propane-2-sulfonic acid octadec-9-enyl-amide (N15), exhibited neuroprotective potential in mouse model of stroke via stimulation of PPAR-α/γ signaling and inhibiting the activation of the NF-κB, STAT3, and ERK1/2 signaling pathways [207]. Icaraside, an active flavonoid compound used for erectile dysfunction, osteoporosis, dementia and cancer treatment, administered as pre-treatment in rats model of I/R, protected against cerebral injury via up-regulation of PPAR-α and PPAR-γ and inhibition of NF-κB activation [208,209]. The 14 days pre-treatment of rats with fenofibrate and pioglitazone used for dyslipidemia and type 2 diabetes treatment respectively, improved ischemia-induced neurobehavioral dysfunction, reduced cerebral infarct volume, attenuated inflammatory and apoptotic markers and ameliorated histopathological changes in ischemia injured rats [210].
Not only pharmacological modulation of PPARs receptors may lead to neuroprotection in stroke. Li and colleagues [211] demonstrated that transcutaneous auricular vagus nerve stimulation (ta-VNS) exert a strong protective effect in rats after cerebral I/R injury. The authors observed that ta-VNS decreased neuronal injury, reduced infarct volume, and increased angiogenesis via upregulation of PPAR-γ expression in the ischemic cortex. The other group demonstrated an anti-inflammatory action of vagus nerve stimulation in cerebral ischemia/reperfusion rats via PPAR-γ upregulation [212]. Liu and colleagues [213] also showed that electrical stimulation of cerebellar fastigial nucleus protected rat brains against ischemia via PPAR-γ upregulation, attenuation of apoptosis and inflammation.
One of the promising experimental strategies against stroke-induced injury is the use of stem cells, which are able to differentiate into diverse cellular population and to replace dying cells during stroke. It has been demonstrated that in rat model of cerebral ischemia, pioglitazone was able to activate innate stem cells in the subventricular zone (SVZ) and recruitment of bone marrow (GFP+BM) stem cells with an increase in PPAR-γ and then increased the expression of Akt, Map2, and Vegf in the cortical peri-infarct area, leading to neurogenesis. Both types of proliferated stem cells migrated from the SVZ into the peri-infarct area and differentiated into mature neurons, glia, and blood vessels [214].
There are also data showing that expression of PPARs is regulated by miRNAs. In rat model of neonatal hypoxic ischemic encephalopathy (HIE), intranasal administration of the PPAR-β/δ agonist GW0742 diminished neuronal death and apoptosis via PPAR-β/δ/miR-17/TXNIP pathway [215]. Moreover, downregulation of miR-383 upregulated Pparg expression and exerted anti-inflammatory and neuroprotective effect in rat model of ischemic stroke [216]. Recently, the role of RXR receptor (PPARs heterodimerization partner) in cerebral diseases was investigated. Mice lacking RXRα in myeloid phagocytes (Mac-RXRα−/−) had a worse functional recovery and they developed brain atrophy after tMCAO [217]. Bexarotene, acting via RXR, improved neurological deficits and exerted anti-inflammatory effect partially through PPAR-γ/SIRT6/FoxO3a pathway in a rat model of subarachnoid hemorrhage [218].

4. Targeting of Aryl Hydrocarbon Receptor (AhR) as Promising Therapeutic Strategy in Myocardial Infarction and Stroke

Human aryl hydrocarbon receptor (AHR) is located on chromosome 7 (7p15) [219]. For a long time AHR was considered only as a regulator of response to environmental pollutants via induction of P450 cytochromes (CYP1A1, CYP1A2, CYP1B1) involved in detoxification. However, a high degree of conservation among the species and the phenotypic alterations observed in AhR-deficient mice suggest a strong involvement of the AhR in cell physiology [220]. Indeed, the loss of the AhR in mice resulted in malformation of the liver [221], heart [222,223,224] and mammary gland [225], in extensive immune dysfunctions [226,227], modulation of stem cells [228,229], oculomotor deficits and defective optic nerve myelin sheath [230] and neuronal deficits [231,232]. For example, lack of AhR in murine cerebellar neuron precursors led to an impairment of neurogenesis. Whereas, an activation or silencing of the AhR in the heart resulted in inhibition of cardiomyocytes differentiation [233,234,235]. It was also demonstrated that AhR is involved in the regulation of cardiomyocytes apoptosis and inflammation [236,237]. Moreover, expression of AhR increases in necrotic myocardium after myocardial infarction induced by LAD ligation [237].
Several endogenous AhR ligands as 6-formyl (3,2-b) carbazole (FICZ), 2-(10-H-indole-3-carbonyl) thiazole-4-carboxylic acid methyl ester (ITE), tryptophan metabolites including kynurenine and other gut microbial products, and leukotrienes [238,239] have been identified, but their physiological role is still under debate [240].
The cellular localization, the mechanisms of action and the protective effects of AhR inhibition during heart and brain ischemia are described on the Figure 3.

4.1. Cellular Localization of AhR in the Heart

The study with the use of AhR-KO mice implies that the myocardium could be a target of AhR signaling [241]. AhR-KO mice were characterized by cardiac hypertrophy and cardiomyopathy accompanied by diminished cardiac output [222,242]. Moreover, it has been demonstrated that porcine aorta endothelial cells, vascular smooth muscle cells and cardiac myocytes respond to AhR agonists [243,244,245]. AhR was also found in cardiac fibroblasts [246] and monocytes [247]. In the developing mouse heart (ED13.5 and ED15.5), AhR was detected mainly in the nucleus of endothelial cells lining the internal and external surfaces of the myocardium, endocardium, and epicardium. At the later phase of development (ED18.5), AhR was observed in cytoplasm of cardiac troponin T-positive cardiomyocytes [248].

4.2. Cellular Localization of AhR in the Brain

It has been shown that Ahr mRNA is present in the mouse brain at the very early developmental stage [249]. Ahr was detected in the cerebral cortex especially in innermost cortical layer on ED12.5. On ED18.5, expression was observed in the hippocampus (pyramidal cell layer of the CA1 and CA3, granule cell layer of the DG regions) and in cerebral cortex. Postnatally, the expression of Ahr was observed at 3, 7 and 14 days after birth, in the CA1 and CA3 pyramidal cell layers, DG granule cell layer of the hippocampus, in the cerebral cortex, cerebellum (the external granule cell layer on earlier days, the granule cell layer on PND 14), in the granule cell layer of the olfactory bulb and the rostral migratory stream (RMS). In the brain of 12-week-old mice, Ahr expression was observed in the hippocampal CA1 and CA3 pyramidal and DG granule cell layers, cerebral cortex, cerebellar and olfactory bulb granule cell layers, and rostral migratory system [249]. Ahr was also detected in neurons, astrocytes, microglia, oligodendrocytes [250,251,252,253,254,255,256], monocytes/macrophages [247] and cerebral endothelial cells [250]. Interestingly, higher level of AhR protein in astrocytes was detected in the brain of elderly than young people [257]. This discovery makes AhR even more attractive target for future therapies against stroke, which occurs mainly in older people.

4.3. Mechanisms of Action of Aryl Hydrocarbon Receptor

AhR is transcription factor belonging to a superfamily of basic helix-loop-helix/Per-ARNT-Sim (bHLH/PAS). In an inactive state AhR is localized in cytoplasm as part of a complex that consists of a dimer of the 90 kDa heat shock protein (HSP90), AhR-interacting protein (AIP; also known as XAP2), the co-chaperone p23 and the protein kinase SRC. Upon ligand binding, AIP dissociates from the complex and translocates to the nucleus in a transportin-dependent and importin-β-dependent manner [258]. Once in the nucleus, AhR binds to Aryl Hydrocarbon Receptor Nuclear Translocator (ARNT), and the dimer is recruited to xenobiotic response element (XRE) activating transcription of many genes involved not only in response to environment pollutants but also in development of cardiovascular and central nervous system [259,260]. Apart from binding to XRE, AhR can also bind to ERE and regulates gene expression by creating a dimers with estrogen receptors in the absence of their ligands [261]. AhR is able to inhibit ERs activity through the binding to the iXRE in the promoters of ERs target genes, squelching of shared coactivators or increased proteasomal degradation of ERs [18]. AhR controls also NF-κB and signal STAT proteins that play a crucial role in the regulation of the immune responses [224,262].

4.4. The Modulation of AhR in Experimental Models of Myocardial Infarction

It is well known that air pollution may increase a risk of development of cardiovascular diseases [263,264]. It has been demonstrated that exposure to particulate matter (PM) pollutants is linked to myocardial infarction, cardiac arrhythmias, increased blood coagulability, atherosclerosis and stroke [265]. Indeed, it has been shown that different manipulation of AhR such as AhR activation, AhR inhibition or AhR knockdown can affect cardiomyocyte differentiation via disruption of AhR-regulated genes e.g., homeobox transcription factors [228,235,266]. Vilahur and colleagues [267] showed a significant increase in AhR expression in pig myocardium after I/R injury, while post-ischemic conditioning inhibited AhR expression thus suggesting the important role of AhR signaling pathway in myocardial injury. Indeed, myocardial infarction induced AhR and AhR-regulated IL-1β and IL-6 in mice heart and a natural flavone baicalin was able to inhibit the myocardial injury and inflammation by decreasing the expression of AhR [237]. Li and colleagues [268] showed dual nature of AhR agonist beta-naphtoflavone (β-NF) in protecting H9C2 cells against OGD. From one side, β-NF reversed OGD/R-induced ROS overproduction, decreased cell death, lactate dehydrogenase release and caspase-3 activity, from another side β-NF, activating AhR, blocked the binding of ARNT to cardioprotective hypoxia-inducible factor (HIF)-1α and in turn inhibited VEGF production and stimulated induction of nitric oxide (NO).
Because of few data on the mechanisms of action of AhR in myocardial infarction, there is an urgent need to unravel the role of AhR signaling pathway during the heart injury.

4.5. The Modulation of AhR in Experimental Models of Stroke

There is an increasing body of evidence that blocking of AhR signaling pathway could be a promising strategy in stroke therapy. It has been shown that experimental stroke is followed by an increase of AhR in the murine brain. Pharmacological and genetic inhibition of AhR leads to a strong reduction of infarct volume in mice subjected to MCAO. AhR antagonist TMF increased CREB transcriptional activity, normalizing BDNF levels and decreasing apoptosis by affecting apoptosis-related genes (reduced pro-apoptotic proteins p53 and Puma and increased the anti-apoptotic Bcl-X) after MCAO [269]. Similarly, TMF decreased infarct volume, inhibited astrogliosis, microgliosis and apoptosis in rodent brains undergoing MCAO [231,270]. Furthermore, TMF treatment modulated gene and protein expression related to neurogenesis after stroke, leading an increased proliferation of neural progenitor cells at the ipsilesional neurogenic zones [231]. Tanaka and colleagues [252] showed that AhR antagonist CH223191 inhibited MCAO-increased the expression of Tnfa and edema progression, and improved the neurological severity scores in mice. The administration of 3,3’-diidnolylmethane (DIM), a selective AhR modulator, protected hippocampal neurons against hypoxia/ischemia via inhibition of AhR signaling pathway. The neuroprotective action of AhR antagonism against ischemia likely involves an inhibition of apoptosis and autophagy [271,272]. Moreover, an in vivo study confirmed that DIM protected rat pups against perinatal asphyxia via inhibition of AhR and NMDA signaling pathways [273]. The latest data showed that intracerebral hemorrhage in mouse induced the expression of AhR in microglia and neutrophils. DIM attenuated activation of microglia/macrophages and astrocytes and diminished infiltration of neutrophils into the hematoma. DIM also decreased AhR-regulated Il6 and Cxcl1 [274]. These results strongly suggest an involvement of AhR in immune cell functions during intracerebral hemorrhage. Although promising experimental evidence on the important role of AhR signaling pathway in stroke pathology was obtained, this topic is still unexplored and requires further research.

5. Conclusions

The progress in mechanical therapies (i.e., stenting or mechanical thrombectomy) and the use of thrombolytic drugs in stroke or in myocardial infarction reduced significantly the rate of mortality. However, scientists still have to look for new more efficient and safer drugs that will be able to prolong a short time-window of currently available treatments. Future therapies should focus not only on well-known ischemia-induced mechanisms, but also on new molecular targets (i.e., nuclear receptors) and compounds (i.e., SERMs, SAhRMs), which can help in reduction of infarct-induced damage and significantly improve patient’s life.

Author Contributions

J.R. Conceptualization, Writing—Original Draft Preparation, Writing—Review & Editing, Supervision, L.C. Writing—Review & Editing, P.G. Conceptualization, Writing—Original Draft Preparation, Visualization, M.M. Writing—Review & Editing, B.M. Writing—Review & Editing, L.S. Writing—Original Draft Preparation, Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AHRaryl hydrocarbon receptor
AIPAHR-interacting protein
Akt (PKB)protein kinase B
ANPpro-atrial natriuretic peptide
AP-1 activator protein 1
ARNTaryl hydrocarbon nuclear translocator
BBBblood brain barrier
CK-MBcreatine kinase-(muscle–brain)
CNScentral nervous system
CREBcAMP response element-binding protein
CVDscardiovascular diseases
CXCL1C-X-C Motif Chemokine Ligand 1
CYP1A1cytochrome P450 1A1
DIM3,3′-diindolylmethane
DPNdiaryl-propio-nitrile
E217β-estradiol
ERestrogen receptor
EREestrogen response element
ERKextracellular signal-regulated kinase
ERαestrogen receptor alpha
ERβestrogen receptor beta
GPERG-protein-coupled estrogen receptor
GSK-3βglycogen synthase kinase 3 beta
HIEhypoxic ischemic encephalopathy
HSP90heat shock protein
I/Rischemia/reperfusion
IL-6Interleukina 6
JNKc-Jun N-terminal kinase
KOknock-out
LADleft anterior descending
LPSlipopolysaccharides
LVleft ventricular
MAPKmitogen-activated protein kinases
MImyocardial infarction
miRNAmicroRNA
MMPmethyl-piperidino-pyrazole
mPTPmitochondrial permeability transition pore
MSCsmesenchymal stem cells
MTmechanical thrombectomy
MTA1metastasis-associated protein 1
mTORmammalian target of rapamycin kinase
NCoRthe nuclear receptor corepressor
NF-κBnuclear factor-kappa B
NRF-1nuclear respiratory factor-1
OAOleic acid
OVXovariectomized
PAI-1plasminogen activator inhibitor-1
PCG-1peroxisome proliferator-activated receptor-gamma coactivator 1
PI-3K3-Phosphatydylinosytol kinase
PKCεthe epsilon isoform of protein kinase C
PLCphospholipase C
pMCAOpermanent middle cerebral artery occlusion
PPARperoxisome proliferator-activated receptor
PPREsproliferator response elements
PPTpropyl-pyrazole-triol
RACK2receptor for activated C kinase
ROSreactive oxygen species
rt-PAtissue plasminogen activator
RXRretinoid X receptor
SAHRMselective aryl hydrocarbon receptor modulator
SERMselective estrogen receptor modulator
SMRTsilencing mediator of retinoid and thyroid hormone receptor
SOD1superoxide dismutase
ta-VNStranscutaneous auricular vagus nerve stimulation
TFAMmitochondrial transcription factor A
Tg-ERβtransgenic ERβ
tMCAOtransient middle cerebral artery occlusion
TNF-αTumor necrosis factor
VSMCsvascular smooth muscle cells
XRExenobiotic response element

References

  1. Virani, S.S.; Alonso, A.; Benjamin, E.J.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Chang, A.R.; Cheng, S.; Delling, F.N.; et al. Heart Disease and Stroke Statistics-2020 Update: A Report From the American Heart Association. Circulation 2020, 141, e139–e596. [Google Scholar] [CrossRef] [PubMed]
  2. Chiong, M.; Wang, Z.V.; Pedrozo, Z.; Cao, D.J.; Troncoso, R.; Ibacache, M.; Criollo, A.; Nemchenko, A.; Hill, J.A.; Lavandero, S. Cardiomyocyte death: Mechanisms and translational implications. Cell Death Dis. 2011, 2, e244. [Google Scholar] [CrossRef] [PubMed]
  3. Sekerdag, E.; Solaroglu, I.; Gursoy-Ozdemir, Y. Cell Death Mechanisms in Stroke and Novel Molecular and Cellular Treatment Options. Curr. Neuropharmacol. 2018, 16, 1396–1415. [Google Scholar] [CrossRef] [PubMed]
  4. Anderson, J.L.; Morrow, D.A. Acute Myocardial Infarction. N. Engl. J. Med. 2017, 376, 2053–2064. [Google Scholar] [CrossRef] [Green Version]
  5. Kobayashi, S.; Fukuma, S.; Ikenoue, T.; Fukuhara, S.; Kobayashi, S. Effect of Edaravone on Neurological Symptoms in Real-World Patients With Acute Ischemic Stroke. Stroke 2019, 50, 1805–1811. [Google Scholar] [CrossRef]
  6. Leng, T.; Xiong, Z.-G. Treatment for ischemic stroke: From thrombolysis to thrombectomy and remaining challenges. Brain Circ. 2019, 5, 8. [Google Scholar] [CrossRef]
  7. Regitz-Zagrosek, V.; Oertelt-Prigione, S.; Prescott, E.; Franconi, F.; Gerdts, E.; Foryst-Ludwig, A.; Maas, A.H.E.M.; Kautzky-Willer, A.; Knappe-Wegner, D.; Kintscher, U.; et al. Gender in cardiovascular diseases: Impact on clinical manifestations, management, and outcomes. Eur. Heart J. 2016, 37, 24–34. [Google Scholar] [CrossRef] [Green Version]
  8. Feigin, V.L.; Lawes, C.M.M.; Bennett, D.A.; Anderson, C.S. Stroke epidemiology: A review of population-based studies of incidence, prevalence, and case-fatality in the late 20th century. Lancet Neurol. 2003, 2, 43–53. [Google Scholar] [CrossRef]
  9. Lisabeth, L.; Bushnell, C. Menopause and stroke: An epidemiologic review. Lancet Neurol. 2012, 11, 82–91. [Google Scholar] [CrossRef] [Green Version]
  10. Babiker, F. Estrogenic hormone action in the heart: Regulatory network and function. Cardiovasc. Res. 2002, 53, 709–719. [Google Scholar] [CrossRef] [Green Version]
  11. McEwen, B.S.; Akama, K.T.; Spencer-Segal, J.L.; Milner, T.A.; Waters, E.M. Estrogen effects on the brain: Actions beyond the hypothalamus via novel mechanisms. Behav. Neurosci. 2012, 126, 4–16. [Google Scholar] [CrossRef] [Green Version]
  12. 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]
  13. Suzuki, S.; Brown, C.M.; Wise, P.M. Neuroprotective effects of estrogens following ischemic stroke. Front. Neuroendocrinol. 2009, 30, 201–211. [Google Scholar] [CrossRef] [Green Version]
  14. Vegeto, E.; Benedusi, V.; Maggi, A. Estrogen anti-inflammatory activity in brain: A therapeutic opportunity for menopause and neurodegenerative diseases. Front. Neuroendocrinol. 2008, 29, 507–519. [Google Scholar] [CrossRef] [Green Version]
  15. Kajta, M.; Beyer, C. Cellular Strategies of Estrogen-Mediated Neuroprotection During Brain Development. Endocrine 2003, 21, 3–10. [Google Scholar] [CrossRef]
  16. Bonofiglio, D.; Gabriele, S.; Aquila, S.; Catalano, S.; Gentile, M.; Middea, E.; Giordano, F.; Andò, S. Estrogen Receptor α Binds to Peroxisome Proliferator–Activated Receptor Response Element and Negatively Interferes with Peroxisome Proliferator-Activated Receptor γ Signaling in Breast Cancer Cells. Clin. Cancer Res. 2005, 11, 6139–6147. [Google Scholar] [CrossRef] [Green Version]
  17. Wormke, M.; Stoner, M.; Saville, B.; Walker, K.; Abdelrahim, M.; Burghardt, R.; Safe, S. The Aryl Hydrocarbon Receptor Mediates Degradation of Estrogen Receptor α through Activation of Proteasomes. Mol. Cell. Biol. 2003, 23, 1843–1855. [Google Scholar] [CrossRef] [Green Version]
  18. Matthews, J.; Gustafsson, J.-Å. Estrogen receptor and aryl hydrocarbon receptor signaling pathways. Nucl. Recept. Signal. 2006, 4, nrs.04016. [Google Scholar] [CrossRef]
  19. Menasce, L.P.; White, G.R.M.; Harrison, C.J.; Boyle, J.M. Localization of the Estrogen Receptor Locus (ESR) to Chromosome 6q25.1 by FISH and a Simple Post-FISH Banding Technique. Genomics 1993, 17, 263–265. [Google Scholar] [CrossRef]
  20. Mosselman, S.; Polman, J.; Dijkema, R. ERβ: Identification and characterization of a novel human estrogen receptor. FEBS Lett. 1996, 392, 49–53. [Google Scholar] [CrossRef] [Green Version]
  21. Kumar, R.; Zakharov, M.N.; Khan, S.H.; Miki, R.; Jang, H.; Toraldo, G.; Singh, R.; Bhasin, S.; Jasuja, R. The Dynamic Structure of the Estrogen Receptor. J. Amino Acids 2011, 2011, 1–7. [Google Scholar] [CrossRef] [Green Version]
  22. Kim, C.K.; Torcaso, A.; Asimes, A.; Chung, W.C.J.; Pak, T.R. Structural and functional characteristics of oestrogen receptor β splice variants: Implications for the ageing brain. J. Neuroendocrinol. 2018, 30, e12488. [Google Scholar] [CrossRef] [Green Version]
  23. Kuiper, G.G.J.M.; Carlsson, B.; Grandien, K.; Enmark, E.; Häggblad, J.; Nilsson, S.; Gustafsson, J.Å. Comparison of the ligand binding specificity and transcript tissue distribution of estrogen receptors and α and β. Endocrinology 1997, 138, 863–870. [Google Scholar] [CrossRef]
  24. Mizukami, Y. In vivo functions of GPR30/GPER-1, a membrane receptor for estrogen: From discovery to functions in vivo. Endocr. J. 2010, 57, 101–107. [Google Scholar] [CrossRef] [Green Version]
  25. Amenyogbe, E.; Chen, G.; Wang, Z.; Lu, X.; Lin, M.; Lin, A.Y. A Review on Sex Steroid Hormone Estrogen Receptors in Mammals and Fish. Int. J. Endocrinol. 2020, 2020, 1–9. [Google Scholar] [CrossRef] [Green Version]
  26. Taylor, A.H.; Al-Azzawi, F. Immunolocalisation of oestrogen receptor beta in human tissues. J. Mol. Endocrinol. 2000, 24, 145–155. [Google Scholar] [CrossRef]
  27. Mahmoodzadeh, S.; Eder, S.; Nordmeyer, J.; Ehler, E.; Huber, O.; Martus, P.; Weiske, J.; Pregla, R.; Hetzer, R.; Regitz-Zagrosek, V. Estrogen receptor alpha up-regulation and redistribution in human heart failure. FASEB J. 2006, 20, 926–934. [Google Scholar] [CrossRef] [Green Version]
  28. Lizotte, E.; Grandy, S.A.; Tremblay, A.; Allen, B.G.; Fiset, C. Cellular Physiology Biochemistry and Biochemistr y Expression, Distribution and Regulation of Sex Steroid Hormone Receptors in Mouse Heart. Cell. Physiol. Biochem. 2009, 8, 75–86. [Google Scholar] [CrossRef]
  29. Grohé, C.; Kahlert, S.; Löbbert, K.; Stimpel, M.; Karas, R.H.; Vetter, H.; Neyses, L. Cardiac myocytes and fibroblasts contain functional estrogen receptors. FEBS Lett. 1997, 416, 107–112. [Google Scholar] [CrossRef]
  30. Razandi, M.; Pedram, A.; Merchenthaler, I.; Greene, G.L.; Levin, E.R. Plasma membrane estrogen receptors exist and functions as dimers. Mol. Endocrinol. 2004, 18, 2854–2865. [Google Scholar] [CrossRef]
  31. Hodges, Y.K.; Tung, L.; Yan, X.-D.; Graham, J.D.; Horwitz, K.B.; Horwitz, L.D. Estrogen Receptors α and β. Circulation 2000, 101, 1792–1798. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Komi, J.; Lassila, O. Nonsteroidal anti-estrogens inhibit the functional differentiation of human monocyte-derived dendritic cells. Blood 2000, 95, 2875–2882. [Google Scholar] [CrossRef] [PubMed]
  33. Tomicek, N.J.; Miller-Lee, J.L.; Hunter, J.C.; Korzick, D.H. Estrogen Receptor Beta Does Not Influence Ischemic Tolerance in the Aged Female Rat Heart. Cardiovasc. Ther. 2013, 31, 32–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Pugach, E.K.; Blenck, C.L.; Dragavon, J.M.; Langer, S.J.; Leinwand, L.A. Estrogen receptor profiling and activity in cardiac myocytes. Mol. Cell. Endocrinol. 2016, 431, 62–70. [Google Scholar] [CrossRef] [Green Version]
  35. Andersson, S.; Sundberg, M.; Pristovsek, N.; Ibrahim, A.; Jonsson, P.; Katona, B.; Clausson, C.-M.; Zieba, A.; Ramström, M.; Söderberg, O.; et al. Insufficient antibody validation challenges oestrogen receptor beta research. Nat. Commun. 2017, 8, 15840. [Google Scholar] [CrossRef]
  36. Pelekanou, V.; Kampa, M.; Kiagiadaki, F.; Deli, A.; Theodoropoulos, P.; Agrogiannis, G.; Patsouris, E.; Tsapis, A.; Castanas, E.; Notas, G. Estrogen anti-inflammatory activity on human monocytes is mediated through cross-talk between estrogen receptor ERα36 and GPR30/GPER1. J. Leukoc. Biol. 2016, 99, 333–347. [Google Scholar] [CrossRef]
  37. Nordmeyer, J.; Eder, S.; Mahmoodzadeh, S.; Martus, P.; Fielitz, J.; Bass, J.; Bethke, N.; Zurbrügg, H.R.; Pregla, R.; Hetzer, R.; et al. Upregulation of myocardial estrogen receptors in human aortic stenosis. Circulation 2004, 110, 3270–3275. [Google Scholar] [CrossRef] [Green Version]
  38. Broberg, A.M.; Siddiqui, A.J.; Fischer, H.; Grinnemo, K.-H.; Wardell, E.; Andersson, A.B.; Inzunza, J.; Sylvén, C.; Gustafsson, J.Å. Estrogen receptors do not influence angiogenesis after myocardial infarction. Scand. Cardiovasc. J. 2011, 45, 215–222. [Google Scholar] [CrossRef]
  39. Patel, V.H.; Chen, J.; Ramanjaneya, M.; Karteris, E.; Zachariades, E.; Thomas, P.; Been, M.; Randeva, H.S. G-protein coupled estrogen receptor 1 expression in rat and human heart: Protective role during ischaemic stress. Int. J. Mol. Med. 2010, 26, 193–199. [Google Scholar] [CrossRef]
  40. 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 2017, 1863, 1870–1882. [Google Scholar] [CrossRef]
  41. Wang, H.; Zhao, Z.; Lin, M.; Groban, L. Activation of GPR30 inhibits cardiac fibroblast proliferation. Mol. Cell. Biochem. 2015, 405, 135–148. [Google Scholar] [CrossRef] [Green Version]
  42. Zhao, Z.; Wang, H.; Lin, M.; Groban, L. GPR30 decreases cardiac chymase/angiotensin II by inhibiting local mast cell number. Biochem. Biophys. Res. Commun. 2015, 459, 131–136. [Google Scholar] [CrossRef] [Green Version]
  43. 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. Metab. 2011, 301, E882–E888. [Google Scholar] [CrossRef] [Green Version]
  44. 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] [Green Version]
  45. Srivastava, D.P.; Waters, E.M.; Mermelstein, P.G.; Kramár, E.A.; Shors, T.J.; Liu, F. Rapid estrogen signaling in the brain: Implications for the fine-tuning of neuronal circuitry. J. Neurosci. 2011, 31, 16056–16063. [Google Scholar] [CrossRef] [Green Version]
  46. Crespo-Castrillo, A.; Arevalo, M.-A. Microglial and Astrocytic Function in Physiological and Pathological Conditions: Estrogenic Modulation. Int. J. Mol. Sci. 2020, 21, 3219. [Google Scholar] [CrossRef]
  47. Cersosimo, M.G.; Benarroch, E.E. Estrogen actions in the nervous system. Neurology 2015, 85, 263–273. [Google Scholar] [CrossRef]
  48. Llorente, R.; Marraudino, M.; Carrillo, B.; Bonaldo, B.; Simon-Areces, J.; Abellanas-Pérez, P.; Rivero-Aguilar, M.; Fernandez-Garcia, J.M.; Pinos, H.; Garcia-Segura, L.M.; et al. G Protein-Coupled Estrogen Receptor Immunoreactivity Fluctuates During the Estrous Cycle and Show Sex Differences in the Amygdala and Dorsal Hippocampus. Front. Endocrinol. 2020, 11, 1–11. [Google Scholar] [CrossRef]
  49. Fuentes, N.; Silveyra, P. Estrogen receptor signaling mechanisms. Adv. Protein Chem. Struct. Biol. 2019, 116, 135–170. [Google Scholar] [CrossRef]
  50. Björnstroöm, L.; Sjöberg, M. Mechanisms of Estrogen Receptor Signaling: Convergence of Genomic and Nongenomic Actions on Target Genes. Mol. Endocrinol. 2005, 19, 833–842. [Google Scholar] [CrossRef] [Green Version]
  51. Marino, M.; Galluzzo, P.; Ascenzi, P. Estrogen Signaling Multiple Pathways to Impact Gene Transcription. Curr. Genom. 2006, 7, 497–508. [Google Scholar] [CrossRef] [Green Version]
  52. Vrtačnik, P.; Ostanek, B.; Mencej-Bedrač, S.; Marc, J. The many faces of estrogen signaling. Biochem. Medica 2014, 24, 329–342. [Google Scholar] [CrossRef] [Green Version]
  53. Klinge, C.M. Estrogenic control of mitochondrial function and biogenesis. J. Cell. Biochem. 2008, 105, 1342–1351. [Google Scholar] [CrossRef] [Green Version]
  54. Zárate, S.; Stevnsner, T.; Gredilla, R. Role of estrogen and other sex hormones in brain aging. Neuroprotection and DNA repair. Front. Aging Neurosci. 2017, 9, 1–22. [Google Scholar] [CrossRef] [Green Version]
  55. Chen, J.Q.; Eshete, M.; Alworth, W.L.; Yager, J.D. Binding of MCF-7 cell mitochondrial proteins and recombinant human estrogen receptors alpha and beta to human mitochondrial DNA estrogen response elements. J. Cell. Biochem. 2004, 93, 358–373. [Google Scholar] [CrossRef]
  56. 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] [Green Version]
  57. Evans, P.D. Rapid signalling responses via the G protein-coupled estrogen receptor, GPER, in a hippocampal cell line. Steroids 2019, 152, 108487. [Google Scholar] [CrossRef]
  58. Deschamps, A.M.; Murphy, E. Activation of a novel estrogen receptor, GPER, is cardioprotective in male and female rats. Am. J. Physiol. Circ. Physiol. 2009, 297, H1806–H1813. [Google Scholar] [CrossRef] [Green Version]
  59. De Francesco, E.M.; Angelone, T.; Pasqua, T.; Pupo, M.; Cerra, M.C.; Maggiolini, M. GPER Mediates Cardiotropic Effects in Spontaneously Hypertensive Rat Hearts. PLoS ONE 2013, 8, e69322. [Google Scholar] [CrossRef] [Green Version]
  60. Roque, C.; Baltazar, G. G protein-coupled estrogen receptor 1 (GPER) activation triggers different signaling pathways on neurons and astrocytes. Neural Regen. Res. 2019, 14, 2069–2070. [Google Scholar] [CrossRef]
  61. Westberry, J.M.; Prewitt, A.K.; Wilson, M.E. Epigenetic regulation of the estrogen receptor alpha promoter in the cerebral cortex following ischemia in male and female rats. Neuroscience 2008, 152, 982–989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Lin, H.F.; Hsi, E.; Liao, Y.C.; Chhor, B.; Hung, J.; Juo, S.H.H.; Lin, R.T.; Chiariotti, L. Demethylation of circulating estrogen receptor alpha gene in cerebral ischemic stroke. PLoS ONE 2015, 10, e0139608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zhai, P.; Eurell, T.E.; Cotthaus, R.; Jeffery, E.H.; Bahr, J.M.; Gross, D.R. Effect of estrogen on global myocardial ischemia-reperfusion injury in female rats. Am. J. Physiol. Heart Circ. Physiol. 2000, 279, 2766–2775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Wang, M.; Crisostomo, P.; Wairiuko, G.M.; Meldrum, D.R. Estrogen receptor-α mediates acute myocardial protection in females. Am. J. Physiol. Heart Circ. Physiol. 2006, 290, 2204–2209. [Google Scholar] [CrossRef] [PubMed]
  65. Gabel, S.A.; Walker, V.R.; London, R.E.; Steenbergen, C.; Korach, K.S.; Murphy, E. Estrogen receptor beta mediates gender differences in ischemia/reperfusion injury. J. Mol. Cell. Cardiol. 2005, 38, 289–297. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, M.; Wang, Y.; Weil, B.; Abarbanell, A.; Herrmann, J.; Tan, J.; Kelly, M.; Meldrum, D.R. Estrogen receptor β mediates increased activation of PI3K/Akt signaling and improved myocardial function in female hearts following acute ischemia. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2009, 296, 972–978. [Google Scholar] [CrossRef] [Green Version]
  67. Pelzer, T.; Loza, P.A.A.; Hu, K.; Bayer, B.; Dienesch, C.; Calvillo, L.; Couse, J.F.; Korach, K.S.; Neyses, L.; Ertl, G. Increased mortality and aggravation of heart failure in estrogen receptor-β knockout mice after myocardial infarction. Circulation 2005, 111, 1492–1498. [Google Scholar] [CrossRef] [Green Version]
  68. Babiker, F.A.; Lips, D.J.; Delvaux, E.; Zandberg, P.; Janssen, B.J.A.; Prinzen, F.; van Eys, G.; Grohé, C.; Doevendans, P.A. Oestrogen modulates cardiac ischaemic remodelling through oestrogen receptor-specific mechanisms. Acta Physiol. 2007, 189, 23–31. [Google Scholar] [CrossRef]
  69. Luo, T.; Kim, J.K. The Role of Estrogen and Estrogen Receptors on Cardiomyocytes: An Overview. Can. J. Cardiol. 2016, 32, 1017–1025. [Google Scholar] [CrossRef] [Green Version]
  70. Mahmoodzadeh, S. Cardiomyocyte-specific Estrogen Receptor Alpha Increases Angiogenesis, Lymphangiogenesis and Reduces Fibrosis in the Female Mouse Heart Post-Myocardial Infarction. J. Cell Sci. Ther. 2014, 5, 153. [Google Scholar] [CrossRef] [Green Version]
  71. Yuan, Z.; Kang, L.; Wang, Z.; Chen, A.; Zhao, Q.; Li, H. 17β-estradiol promotes recovery after myocardial infarction by enhancing homing and angiogenic capacity of bone marrow-derived endothelial progenitor cells through ERα-SDF-1/CXCR4 crosstalking. Acta Biochim. Biophys. Sin. 2018, 50, 1247–1256. [Google Scholar] [CrossRef]
  72. Hamada, H.; Kim, M.K.; Iwakura, A.; Ii, M.; Thorne, T.; Qin, G.; Asai, J.; Tsutsumi, Y.; Sekiguchi, H.; Silver, M.; et al. Estrogen receptors α and β mediate contribution of bone marrow-derived endothelial progenitor cells to functional recovery after myocardial infarction. Circulation 2006, 114, 2261–2270. [Google Scholar] [CrossRef] [Green Version]
  73. Zhang, J.B.; Guo, C.L. Protective effect and mechanism of estrogen receptor β on myocardial infarction in mice. Exp. Ther. Med. 2017, 14, 1315–1320. [Google Scholar] [CrossRef] [Green Version]
  74. Schuster, I.; Mahmoodzadeh, S.; Dworatzek, E.; Jaisser, F.; Messaoudi, S.; Morano, I.; Regitz-Zagrosek, V. Cardiomyocyte-specific overexpression of oestrogen receptor β improves survival and cardiac function after myocardial infarction in female and male mice. Clin. Sci. 2016, 130, 365–376. [Google Scholar] [CrossRef] [Green Version]
  75. Booth, E.A.; Obeid, N.R.; Lucchesi, B.R. Activation of estrogen receptor-α protects the in vivo rabbit heart from ischemia-reperfusion injury. Am. J. Physiol. Circ. Physiol. 2005, 289, H2039–H2047. [Google Scholar] [CrossRef] [Green Version]
  76. Novotny, J.L.; Simpson, A.M.; Tomicek, N.J.; Lancaster, T.S.; Korzick, D.H. Rapid estrogen receptor-α activation improves ischemic tolerance in aged female rats through a novel protein kinase Cε -dependent mechanism. Endocrinology 2009, 150, 889–896. [Google Scholar] [CrossRef] [Green Version]
  77. Bulut, E.C.; Abueid, L.; Ercan, F.; Süleymanoğlu, S.; Ağırbaşlı, M.; Yeğen, B.Ç. Treatment with oestrogen-receptor agonists or oxytocin in conjunction with exercise protects against myocardial infarction in ovariectomized rats. Exp. Physiol. 2016, 101, 612–627. [Google Scholar] [CrossRef]
  78. Lee, T.; Lin, S.; Chang, N. Both GPER and membrane oestrogen receptor-α activation protect ventricular remodelling in 17β oestradiol-treated ovariectomized infarcted rats. J. Cell. Mol. Med. 2014, 18, 2454–2465. [Google Scholar] [CrossRef]
  79. Jeanes, H.L.; Tabor, C.; Black, D.; Ederveen, A.; Gray, G.A. Oestrogen-mediated cardioprotection following ischaemia and reperfusion is mimicked by an oestrogen receptor (ER)α agonist and unaffected by an ERβ antagonist. J. Endocrinol. 2008, 197, 493–501. [Google Scholar] [CrossRef]
  80. Nikolic, I.; Liu, D.; Bell, J.A.; Collins, J.; Steenbergen, C.; Murphy, E. Treatment with an estrogen receptor-beta-selective agonist is cardioprotective. J. Mol. Cell. Cardiol. 2007, 42, 769–780. [Google Scholar] [CrossRef]
  81. Lin, J.; Steenbergen, C.; Murphy, E.; Sun, J. Estrogen receptor-beta activation results in S-nitrosylation of proteins involved in cardioprotection. Circulation 2009, 120, 245–254. [Google Scholar] [CrossRef] [Green Version]
  82. Du, M.; Shan, J.; Feng, A.; Schmull, S.; Gu, J.; Xue, S. Oestrogen Receptor β Activation Protects Against Myocardial Infarction via Notch1 Signalling. Cardiovasc. Drugs Ther. 2020, 34, 165–178. [Google Scholar] [CrossRef]
  83. Martinkovich, S.; Shah, D.; Planey, S.L.; Arnott, J.A. Selective estrogen receptor modulators: Tissue specificity and clinical utility. Clin. Interv. Aging 2014, 9, 1437–1452. [Google Scholar] [CrossRef] [Green Version]
  84. Shi, W.; Ma, H.; Liu, T.; Yan, D.; Luo, P.; Zhai, M.; Tao, J.; Huo, S.; Guo, J.; Li, C.; et al. Inhibition of Interleukin-6/glycoprotein 130 signalling by Bazedoxifene ameliorates cardiac remodelling in pressure overload mice. J. Cell. Mol. Med. 2020, 24, 4748–4761. [Google Scholar] [CrossRef] [Green Version]
  85. Posa, A.; Szabó, R.; Kupai, K.; Berkó, A.M.; Veszelka, M.; Szucs, G.; Börzsei, D.; Gyöngyösi, M.; Pávó, I.; Deim, Z.; et al. Cardioprotective Effect of Selective Estrogen Receptor Modulator Raloxifene Are Mediated by Heme Oxygenase in Estrogen-Deficient Rat. Oxid. Med. Cell. Longev. 2017, 2017, 2176749. [Google Scholar] [CrossRef] [Green Version]
  86. Lamas, A.Z.; Nascimento, A.M.; Medeiros, A.R.S.; Caliman, I.F.; Dalpiaz, P.L.M.; Firmes, L.B.; Sousa, G.J.; Oliveira, P.W.C.; Andrade, T.U.; Reis, A.M.; et al. The selective estrogen receptor modulators (SERMs) raloxifene and tamoxifen improve ANP levels and decrease nuclear translocation of NF-kB in estrogen-deficient rats. Pharmacol. Rep. 2017, 69, 798–805. [Google Scholar] [CrossRef]
  87. Smith, C.L.; Santen, R.J.; Komm, B.; Mirkin, S. Breast-related effects of selective estrogen receptor modulators and tissue-selective estrogen complexes. Breast Cancer Res. 2014, 16, 1–15. [Google Scholar] [CrossRef] [Green Version]
  88. Rayabarapu, N.; Patel, B.M. Beneficial role of tamoxifen in isoproterenol-induced myocardial infarction. Can. J. Physiol. Pharmacol. 2014, 92, 849–857. [Google Scholar] [CrossRef]
  89. Chung, M.-T.; Cheng, P.-Y.; Lam, K.-K.; Chen, S.-Y.; Ting, Y.-F.; Yen, M.-H.; Lee, Y.-M. Cardioprotective effects of long-term treatment with raloxifene, a selective estrogen receptor modulator, on myocardial ischemia/reperfusion injury in ovariectomized rats. Menopause 2010, 17, 127–134. [Google Scholar] [CrossRef] [PubMed]
  90. Bopassa, J.C.; Eghbali, M.; Toro, L.; Stefani, E. A novel estrogen receptor GPER inhibits mitochondria permeability transition pore opening and protects the heart against ischemia-reperfusion injury. Am. J. Physiol. Circ. Physiol. 2010, 298, H16–H23. [Google Scholar] [CrossRef] [Green Version]
  91. Kabir, M.E.; Singh, H.; Lu, R.; Olde, B.; Leeb-Lundberg, L.M.F.; Bopassa, J.C. G protein-coupled estrogen receptor 1 mediates acute estrogen-induced cardioprotection via MEK/ERK/GSK-3β Pathway after Ischemia/Reperfusion. PLoS ONE 2015, 10, e0135988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Lim, S.Y.; Davidson, S.M.; Hausenloy, D.J.; Yellon, D.M. Preconditioning and postconditioning: The essential role of the mitochondrial permeability transition pore. Cardiovasc. Res. 2007, 75, 530–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Wang, X.; Lu, L.; Tan, Y.; Jiang, L.; Zhao, M.; Gao, E.; Yu, S.; Liu, J. GPR 30 reduces myocardial infarct area and fibrosis in female ovariectomized mice by activating the PI3K/AKT pathway. Life Sci. 2019, 226, 22–32. [Google Scholar] [CrossRef]
  94. Azizian, H.; Khaksari, M.; Asadikaram, G.; Sepehri, G.; Najafipour, H. Therapeutic effects of tamoxifen on metabolic parameters and cytokines modulation in rat model of postmenopausal diabetic cardiovascular dysfunction: Role of classic estrogen receptors. Int. Immunopharmacol. 2018, 65, 190–198. [Google Scholar] [CrossRef]
  95. 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]
  96. Rocca, C.; Femminò, S.; Aquila, G.; Granieri, M.C.; de Francesco, E.M.; Pasqua, T.; Rigiracciolo, D.C.; Fortini, F.; Cerra, M.C.; Maggiolini, M.; et al. Notch1 mediates preconditioning protection induced by GPER in normotensive and hypertensive female rat hearts. Front. Physiol. 2018, 9, 521. [Google Scholar] [CrossRef]
  97. McEwen, B.S.; Alves, S.E. Estrogen Actions in the Central Nervous System 1. Endocr. Rev. 1999, 20, 279–307. [Google Scholar] [CrossRef]
  98. Raghava, N.; Das, B.C.; Ray, S.K. Neuroprotective effects of estrogen in CNS injuries: Insights from animal models. Neurosci. Neuroecon. 2017, 6, 15–29. [Google Scholar] [CrossRef] [Green Version]
  99. Habib, P.; Dreymueller, D.; Ludwig, A.; Beyer, C.; Dang, J. Sex steroid hormone-mediated functional regulation of microglia-like BV-2 cells during hypoxia. J. Steroid Biochem. Mol. Biol. 2013, 138, 195–205. [Google Scholar] [CrossRef]
  100. Dimayuga, F.O.; Reed, J.L.; Carnero, G.A.; Wang, C.; Dimayuga, E.R.; Dimayuga, V.M.; Perger, A.; Wilson, M.E.; Keller, J.N.; Bruce-Keller, A.J. Estrogen and brain inflammation: Effects on microglial expression of MHC, costimulatory molecules and cytokines. J. Neuroimmunol. 2005, 161, 123–136. [Google Scholar] [CrossRef]
  101. Sampei, K.; Goto, S.; Alkayed, N.J.; Crain, B.J.; Korach, K.S.; Traystman, R.J.; Demas, G.E.; Nelson, R.J.; Hurn, P.D. Stroke in Estrogen Receptor-α–Deficient Mice. Stroke 2000, 31, 738–744. [Google Scholar] [CrossRef] [Green Version]
  102. Dubal, D.B. Estrogen receptor alpha, not beta, is a critical link in estradiol-mediated protection against brain injury. Proc. Natl. Acad. Sci. USA 2001, 98, 1952–1957. [Google Scholar] [CrossRef]
  103. Dubal, D.B.; Rau, S.W.; Shughrue, P.J.; Zhu, H.; Yu, J.; Cashion, A.B.; Suzuki, S.; Gerhold, L.M.; Bottner, M.B.; Dubal, S.B.; et al. Differential Modulation of Estrogen Receptors (ERs) in Ischemic Brain Injury: A Role for ERα in Estradiol-Mediated Protection against Delayed Cell Death. Endocrinology 2006, 147, 3076–3084. [Google Scholar] [CrossRef]
  104. Miller, N.R.; Jover, T.; Cohen, H.W.; Zukin, R.S.; Etgen, A.M. Estrogen can act via estrogen receptor α and β to protect hippocampal neurons against global ischemia-induced cell death. Endocrinology 2005, 146, 3070–3079. [Google Scholar] [CrossRef] [Green Version]
  105. Raval, A.P.; Borges-Garcia, R.; Javier Moreno, W.; Perez-Pinzon, M.A.; Bramlett, H. Periodic 17β-Estradiol Pretreatment Protects Rat Brain from Cerebral Ischemic Damage via Estrogen Receptor-β. PLoS ONE 2013, 8, e60716. [Google Scholar] [CrossRef] [Green Version]
  106. Shin, J.A.; Yoon, J.C.; Kim, M.; Park, E.M. Activation of classical estrogen receptor subtypes reduces tight junction disruption of brain endothelial cells under ischemia/reperfusion injury. Free Radic. Biol. Med. 2016, 92, 78–89. [Google Scholar] [CrossRef]
  107. Carswell, H.V.O.; Macrae, I.M.; Gallagher, L.; Harrop, E.; Horsburgh, K.J. Neuroprotection by a selective estrogen receptor β agonist in a mouse model of global ischemia. Am. J. Physiol. Circ. Physiol. 2004, 287, H1501–H1504. [Google Scholar] [CrossRef]
  108. Noppens, R.R.; Kofler, J.; Grafe, M.R.; Hurn, P.D.; Traystman, R.J. Estradiol after Cardiac Arrest and Cardiopulmonary Resuscitation is Neuroprotective and Mediated through Estrogen Receptor-β. J. Cereb. Blood Flow Metab. 2009, 29, 277–286. [Google Scholar] [CrossRef]
  109. Guo, J.; Zhang, T.; Yu, J.; Li, H.-Z.; Zhao, C.; Qiu, J.; Zhao, B.; Zhao, J.; Li, W.; Zhao, T.-Z. Neuroprotective effects of a chromatin modifier on ischemia/reperfusion neurons: Implication of its regulation of BCL2 transactivation by ERα signaling. Cell Tissue Res. 2016, 364, 475–488. [Google Scholar] [CrossRef]
  110. Zhang, Z.L.; Qin, P.; Liu, Y.; Zhang, L.X.; Guo, H.; Deng, Y.L.; Yizhao, L.; Hou, Y.S.; Wang, L.Y.; Miao, Y.; et al. Alleviation of ischaemia-reperfusion injury by endogenous estrogen involves maintaining Bcl-2 expression via the ERα signalling pathway. Brain Res. 2017, 1661, 15–23. [Google Scholar] [CrossRef]
  111. Stary, C.M.; Xu, L.; Li, L.; Sun, X.; Ouyang, Y.-B.; Xiong, X.; Zhao, J.; Giffard, R.G. Inhibition of miR-181a protects female mice from transient focal cerebral ischemia by targeting astrocyte estrogen receptor-α. Mol. Cell. Neurosci. 2017, 82, 118–125. [Google Scholar] [CrossRef] [PubMed]
  112. Guo, H.; Yang, J.; Liu, M.; Wang, L.; Hou, W.; Zhang, L.; Ma, Y. Selective activation of estrogen receptor β alleviates cerebral ischemia neuroinflammatory injury. Brain Res. 2020, 1726, 146536. [Google Scholar] [CrossRef] [PubMed]
  113. Ma, Y.; Guo, H.; Zhang, L.; Tao, L.; Yin, A.; Liu, Z.; Li, Y.; Dong, H.; Xiong, L.; Hou, W. Estrogen replacement therapy-induced neuroprotection against brain ischemia-reperfusion injury involves the activation of astrocytes via estrogen receptor β. Sci. Rep. 2016, 6, 21467. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Madinier, A.; Wieloch, T.; Olsson, R.; Ruscher, K. Impact of estrogen receptor beta activation on functional recovery after experimental stroke. Behav. Brain Res. 2014, 261, 282–288. [Google Scholar] [CrossRef]
  115. Shin, J.A.; Yang, S.J.; Jeong, S.I.; Park, H.J.; Choi, Y.H.; Park, E.M. Activation of estrogen receptor β reduces blood-brain barrier breakdown following ischemic injury. Neuroscience 2013, 235, 165–173. [Google Scholar] [CrossRef]
  116. Kimelberg, H.K.; Feustel, P.J.; Jin, Y.; Paquette, J.; Boulos, A.; Keller, J.; Tranmer, B.I. Acute treatment with tamoxifen reduces ischemic damage following middle cerebral artery occlusion. Neuroreport 2000, 11, 2675–2679. [Google Scholar] [CrossRef]
  117. Kimelberg, H.K.; Jin, Y.; Charniga, C.; Feustel, P.J. Neuroprotective activity of tamoxifen in permanent focal ischemia. J. Neurosurg. 2003, 99, 138–142. [Google Scholar] [CrossRef]
  118. Wakade, C.; Khan, M.M.; de Sevilla, L.M.; Zhang, Q.G.; Mahesh, V.B.; Brann, D.W. Tamoxifen neuroprotection in cerebral ischemia involves attenuation of kinase activation and superoxide production and potentiation of mitochondrial superoxide dismutase. Endocrinology 2008, 149, 367–379. [Google Scholar] [CrossRef] [Green Version]
  119. Zhang, Y.; Milatovic, D.; Aschner, M.; Feustel, P.J.; Kimelberg, H.K. Neuroprotection by tamoxifen in focal cerebral ischemia is not mediated by an agonist action at estrogen receptors but is associated with antioxidant activity. Exp. Neurol. 2007, 204, 819–827. [Google Scholar] [CrossRef] [Green Version]
  120. Zou, W.; Fang, C.; Ji, X.; Liang, X.; Liu, Y.; Han, C.; Huang, L.; Zhang, Q.; Li, H.; Zhang, Y.; et al. Estrogen receptor (ER)-aα36 is involved in estrogen- and tamoxifen-induced neuroprotective effects in ischemic stroke models. PLoS ONE 2015, 10, e0140660. [Google Scholar] [CrossRef] [Green Version]
  121. Rzemieniec, J.; Litwa, E.; Wnuk, A.; Lason, W.; Gołas, A.; Krzeptowski, W.; Kajta, M. Neuroprotective action of raloxifene against hypoxia-induced damage in mouse hippocampal cells depends on ERα but not ERβ or GPR30 signalling. J. Steroid Biochem. Mol. Biol. 2015, 146, 26–37. [Google Scholar] [CrossRef]
  122. Rzemieniec, J.; Litwa, E.; Wnuk, A.; Lason, W.; Kajta, M. Bazedoxifene and raloxifene protect neocortical neurons undergoing hypoxia via targeting ERα and PPAR-γ. Mol. Cell. Endocrinol. 2018, 461, 64–78. [Google Scholar] [CrossRef]
  123. Khan, M.M.; Wakade, C.; de Sevilla, L.; Brann, D.W. Selective estrogen receptor modulators (SERMs) enhance neurogenesis and spine density following focal cerebral ischemia. J. Steroid Biochem. Mol. Biol. 2015, 146, 38–47. [Google Scholar] [CrossRef] [Green Version]
  124. Castelló-Ruiz, M.; Torregrosa, G.; Burguete, M.C.; Miranda, F.J.; Centeno, J.M.; López-Morales, M.A.; Gasull, T.; Alborch, E. The selective estrogen receptor modulator, bazedoxifene, reduces ischemic brain damage in male rat. Neurosci. Lett. 2014, 575, 53–57. [Google Scholar] [CrossRef]
  125. Jover-Mengual, T.; Castelló-Ruiz, M.; Burguete, M.C.; Jorques, M.; López-Morales, M.A.; Aliena-Valero, A.; Jurado-Rodríguez, A.; Pérez, S.; Centeno, J.M.; Miranda, F.J.; et al. Molecular mechanisms mediating the neuroprotective role of the selective estrogen receptor modulator, bazedoxifene, in acute ischemic stroke: A comparative study with 17β-estradiol. J. Steroid Biochem. Mol. Biol. 2017, 171, 296–304. [Google Scholar] [CrossRef]
  126. Burguete, M.C.; Jover-Mengual, T.; López-Morales, M.A.; Aliena-Valero, A.; Jorques, M.; Torregrosa, G.; Alborch, E.; Castelló-Ruiz, M.; Salom, J.B. The selective oestrogen receptor modulator, bazedoxifene, mimics the neuroprotective effect of 17β-oestradiol in diabetic ischaemic stroke by modulating oestrogen receptor expression and the MAPK/ERK1/2 signalling pathway. J. Neuroendocrinol. 2019, 31, e12751. [Google Scholar] [CrossRef]
  127. Zhang, B.; Subramanian, S.; Dziennis, S.; Jia, J.; Uchida, M.; Akiyoshi, K.; Migliati, E.; Lewis, A.D.; Vandenbark, A.A.; Offner, H.; et al. Estradiol and G1 Reduce Infarct Size and Improve Immunosuppression after Experimental Stroke. J. Immunol. 2010, 184, 4087–4094. [Google Scholar] [CrossRef] [Green Version]
  128. Zhao, T.Z.; Ding, Q.; Hu, J.; He, S.M.; Shi, F.; Ma, L.T. GPER expressed on microglia mediates the anti-inflammatory effect of estradiol in ischemic stroke. Brain Behav. 2016, 6, 1–9. [Google Scholar] [CrossRef] [Green Version]
  129. Zhang, Z.; Qin, P.; Deng, Y.; Ma, Z.; Guo, H.; Guo, H.; Hou, Y.; Wang, S.; Zou, W.; Sun, Y.; et al. The novel estrogenic receptor GPR30 alleviates ischemic injury by inhibiting TLR4-mediated microglial inflammation. J. Neuroinflamm. 2018, 15, 1–13. [Google Scholar] [CrossRef] [Green Version]
  130. Bai, N.; Zhang, Q.; Zhang, W.; Liu, B.; Yang, F.; Brann, D.; Wang, R. G-protein-coupled estrogen receptor activation upregulates interleukin-1 receptor antagonist in the hippocampus after global cerebral ischemia: Implications for neuronal self-defense. J. Neuroinflamm. 2020, 17, 45. [Google Scholar] [CrossRef]
  131. Liu, S.-b.; Zhang, N.; Guo, Y.-y.; Zhao, R.; Shi, T.-y.; Feng, S.-f.; Wang, S.-q.; Yang, Q.; Li, X.-q.; Wu, Y.-m.; et al. G-Protein-Coupled Receptor 30 Mediates Rapid Neuroprotective Effects of Estrogen via Depression of NR2B-Containing NMDA Receptors. J. Neurosci. 2012, 32, 4887–4900. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Lamprecht, M.R.; Morrison, B. GPR30 activation is neither necessary nor sufficient for acute neuroprotection by 17β-estradiol after an ischemic injury in organotypic hippocampal slice cultures. Brain Res. 2014, 1563, 131–137. [Google Scholar] [CrossRef] [PubMed]
  133. Murata, T.; Dietrich, H.H.; Xiang, C.; Dacey, R.G. G Protein-Coupled Estrogen Receptor Agonist Improves Cerebral Microvascular Function After Hypoxia/Reoxygenation Injury in Male and Female Rats. Stroke 2013, 44, 779–785. [Google Scholar] [CrossRef] [PubMed]
  134. Lu, D.; Qu, Y.; Shi, F.; Feng, D.; Tao, K.; Gao, G.; He, S.; Zhao, T. Activation of G protein-coupled estrogen receptor 1 (GPER-1) ameliorates blood-brain barrier permeability after global cerebral ischemia in ovariectomized rats. Biochem. Biophys. Res. Commun. 2016, 477, 209–214. [Google Scholar] [CrossRef] [PubMed]
  135. Kosaka, Y.; Quillinan, N.; Bond, C.T.; Traystman, R.J.; Hurn, P.D.; Herson, P.S. GPER1/GPR30 Activation Improves Neuronal Survival Following Global Cerebral Ischemia Induced by Cardiac Arrest in Mice. Transl. Stroke Res. 2012, 3, 500–507. [Google Scholar] [CrossRef] [Green Version]
  136. Wang, X.S.; Yue, J.; Hu, L.N.; Tian, Z.; Zhang, K.; Yang, L.; Zhang, H.N.; Guo, Y.Y.; Feng, B.; Liu, H.Y.; et al. Activation of G protein-coupled receptor 30 protects neurons by regulating autophagy in astrocytes. Glia 2020, 68, 27–43. [Google Scholar] [CrossRef] [Green Version]
  137. Broughton, B.R.S.; Brait, V.H.; Kim, H.A.; Lee, S.; Chu, H.X.; Gardiner-Mann, C.V.; Guida, E.; Evans, M.A.; Miller, A.A.; Arumugam, T.V.; et al. Sex-Dependent Effects of G Protein–Coupled Estrogen Receptor Activity on Outcome After Ischemic Stroke. Stroke 2014, 45, 835–841. [Google Scholar] [CrossRef] [Green Version]
  138. Tyagi, S.; Gupta, P.; Saini, A.; Kaushal, C.; Sharma, S. The peroxisome proliferator-activated receptor: A family of nuclear receptors role in various diseases. J. Adv. Pharm. Technol. Res. 2011, 2, 236–240. [Google Scholar] [CrossRef]
  139. Dong, C.; Zhou, H.; Shen, C.; Yu, L.-G.; Ding, Y.; Zhang, Y.-H.; Guo, Z.-R. Role of peroxisome proliferator-activated receptors gene polymorphisms in type 2 diabetes and metabolic syndrome. World J. Diabetes 2015, 6, 654. [Google Scholar] [CrossRef] [Green Version]
  140. Lee, W.-S.; Kim, J. Peroxisome Proliferator-Activated Receptors and the Heart: Lessons from the Past and Future Directions. PPAR Res. 2015, 2015, 1–18. [Google Scholar] [CrossRef] [Green Version]
  141. Higashiyama, H.; Billin, A.N.; Okamoto, Y.; Kinoshita, M.; Asano, S. Expression profiling of Peroxisome proliferator-activated receptor-delta (PPAR-delta) in mouse tissues using tissue microarray. Histochem. Cell Biol. 2007, 127, 485–494. [Google Scholar] [CrossRef]
  142. Rosen, E.D.; Spiegelman, B.M. PPARγ: A Nuclear Regulator of Metabolism, Differentiation, and Cell Growth. J. Biol. Chem. 2001, 276, 37731–37734. [Google Scholar] [CrossRef] [Green Version]
  143. Madrazo, J.A.; Kelly, D.P. The PPAR trio: Regulators of myocardial energy metabolism in health and disease. J. Mol. Cell. Cardiol. 2008, 44, 968–975. [Google Scholar] [CrossRef]
  144. Son, N.H.; Park, T.S.; Yamashita, H.; Yokoyama, M.; Huggins, L.A.; Okajima, K.; Homma, S.; Szabolcs, M.J.; Huang, L.S.; Goldberg, I.J. Cardiomyocyte expression of PPARγ leads to cardiac dysfunction in mice. J. Clin. Investig. 2007, 117, 2791–2801. [Google Scholar] [CrossRef] [Green Version]
  145. Barger, P.M.; Brandt, J.M.; Leone, T.C.; Weinheimer, C.J.; Kelly, D.P. Deactivation of peroxisome proliferator-activated receptor-α during cardiac hypertrophic growth. J. Clin. Investig. 2000, 105, 1723–1730. [Google Scholar] [CrossRef] [Green Version]
  146. Castiglioni, L.; Pignieri, A.; Fiaschè, M.; Giudici, M.; Crestani, M.; Mitro, N.; Abbate, M.; Zoja, C.; Rottoli, D.; Foray, C.; et al. Fenofibrate attenuates cardiac and renal alterations in young salt-loaded spontaneously hypertensive stroke-prone rats through mitochondrial protection. J. Hypertens. 2018, 36, 1129–1146. [Google Scholar] [CrossRef]
  147. Huss, J.M.; Levy, F.H.; Kelly, D.P. Hypoxia inhibits the peroxisome proliferator-activated receptor α/retinoid X receptor gene regulatory pathway in cardiac myocytes: A mechanism for O2-dependent modulation of mitochondrial fatty acid oxidation. J. Biol. Chem. 2001, 276, 27605–27612. [Google Scholar] [CrossRef] [Green Version]
  148. Finck, B.N.; Lehman, J.J.; Leone, T.C.; Welch, M.J.; Bennett, M.J.; Kovacs, A.; Han, X.; Gross, R.W.; Kozak, R.; Lopaschuk, G.D.; et al. The cardiac phenotype induced by PPARα overexpression mimics that caused by diabetes mellitus. J. Clin. Investig. 2002, 109, 121–130. [Google Scholar] [CrossRef]
  149. Yu, B.C.; Chang, C.K.; Ou, H.Y.; Cheng, K.C.; Cheng, J.T. Decrease of peroxisome proliferator-activated receptor delta expression in cardiomyopathy of streptozotocin-induced diabetic rats. Cardiovasc. Res. 2008, 80, 78–87. [Google Scholar] [CrossRef] [Green Version]
  150. Fliegner, D.; Westermann, D.; Riad, A.; Schubert, C.; Becher, E.; Fielitz, J.; Tschöpe, C.; Regitz-Zagrosek, V. Up-regulation of PPARγ in myocardial infarction. Eur. J. Heart Fail. 2008, 10, 30–38. [Google Scholar] [CrossRef] [Green Version]
  151. Gilde, A.J.; van der Lee, K.A.J.M.; Willemsen, P.H.M.; Chinetti, G.; van der Leij, F.R.; van der Vusse, G.J.; Staels, B.; van Bilsen, M. Peroxisome Proliferator-Activated Receptor (PPAR) α and PPARβ/δ, but not PPARγ, Modulate the Expression of Genes Involved in Cardiac Lipid Metabolism. Circ. Res. 2003, 92, 518–524. [Google Scholar] [CrossRef] [Green Version]
  152. Teunissen, B.E.J.; Smeets, P.J.H.; Willemsen, P.H.M.; de Windt, L.J.; van der Vusse, G.J.; van Bilsen, M. Activation of PPARδ inhibits cardiac fibroblast proliferation and the transdifferentiation into myofibroblasts. Cardiovasc. Res. 2007, 75, 519–529. [Google Scholar] [CrossRef] [Green Version]
  153. Chinetti, G.; Fruchart, J.C.; Staels, B. Peroxisome proliferator-activated receptors (PPARs): Nuclear receptors with functions in the vascular wall. Z. Kardiol. 2001, 90 (Suppl. 3), 125–132. [Google Scholar] [CrossRef]
  154. Cheang, W.S.; Fang, X.; Tian, X.Y. Pleiotropic Effects of Peroxisome Proliferator-Activated Receptor γ and δ in Vascular Diseases. Circ. J. 2013, 77, 2664–2671. [Google Scholar] [CrossRef] [Green Version]
  155. Adhikary, T.; Wortmann, A.; Schumann, T.; Finkernagel, F.; Lieber, S.; Roth, K.; Toth, P.M.; Diederich, W.E.; Nist, A.; Stiewe, T.; et al. The transcriptional PPARβ/δ network in human macrophages defines a unique agonist-induced activation state. Nucleic Acids Res. 2015, 43, 5033–5051. [Google Scholar] [CrossRef] [Green Version]
  156. Cullingford, T.E.; Bhakoo, K.; Peuchen, S.; Dolphin, C.T.; Patel, R.; Clark, J.B. Distribution of mRNAs Encoding the Peroxisome Proliferator-Activated Receptor α, β, and γ and the Retinoid X Receptor α, β, and γ in Rat Central Nervous System. J. Neurochem. 2002, 70, 1366–1375. [Google Scholar] [CrossRef]
  157. Warden, A.; Truitt, J.; Merriman, M.; Ponomareva, O.; Jameson, K.; Ferguson, L.B.; Mayfield, R.D.; Harris, R.A. Localization of PPAR isotypes in the adult mouse and human brain. Sci. Rep. 2016, 6, 27618. [Google Scholar] [CrossRef]
  158. Mandrekar-Colucci, S.; Sauerbeck, A.; Popovich, P.G.; McTigue, D.M. PPAR agonists as therapeutics for CNS trauma and neurological diseases. ASN Neuro 2013, 5, e00129. [Google Scholar] [CrossRef] [Green Version]
  159. Akanuma, S.; Hori, S.; Ohtsuki, S.; Fujiyoshi, M.; Terasaki, T. Expression of nuclear receptor mRNA and liver X receptor-mediated regulation of ABC transporter A1 at rat blood–brain barrier. Neurochem. Int. 2008, 52, 669–674. [Google Scholar] [CrossRef]
  160. Mysiorek, C.; Culot, M.; Dehouck, L.; Derudas, B.; Staels, B.; Bordet, R.; Cecchelli, R.; Fenart, L.; Berezowski, V. Peroxisome-proliferator-activated receptor-alpha activation protects brain capillary endothelial cells from oxygen-glucose deprivation-induced hyperpermeability in the blood-brain barrier. Curr. Neurovasc. Res. 2009, 6, 181–193. [Google Scholar] [CrossRef]
  161. Yin, K.J.; Fan, Y.; Hamblin, M.; Zhang, J.; Zhu, T.; Li, S.; Hawse, J.R.; Subramaniam, M.; Song, C.Z.; Urrutia, R.; et al. KLF11 mediates PPARγ cerebrovascular protection in ischaemic stroke. Brain 2013, 136, 1274–1287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Berger, J.; Moller, D.E. The Mechanisms of Action of PPARs. Annu. Rev. Med. 2002, 53, 409–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Reddy, J.K.; Viswakarma, N.; Jia, Y.; Bai, L.; Vluggens, A.; Borensztajn, J.; Xu, J. Coactivators in PPAR-regulated gene expression. PPAR Res. 2010, 2010, 250126. [Google Scholar] [CrossRef] [Green Version]
  164. Adams, M.; Reginato, M.J.; Shao, D.; Lazar, M.A.; Chatterjee, V.K. Transcriptional Activation by Peroxisome Proliferator-activated Receptor γ Is Inhibited by Phosphorylation at a Consensus Mitogen-activated Protein Kinase Site. J. Biol. Chem. 1997, 272, 5128–5132. [Google Scholar] [CrossRef] [Green Version]
  165. Lazennec, G.; Canaple, L.; Saugy, D.; Wahli, W. Activation of peroxisome proliferator-activated receptors (PPARs) by their ligands and protein kinase A activators. Mol. Endocrinol. 2000, 14, 1962–1975. [Google Scholar] [CrossRef]
  166. Alexis, J.D.; Wang, N.; Che, W.; Lerner-Marmarosh, N.; Sahni, A.; Korshunov, V.A.; Zou, Y.; Ding, B.; Yan, C.; Berk, B.C.; et al. Bcr Kinase Activation by Angiotensin II Inhibits Peroxisome Proliferator-Activated Receptor γ Transcriptional Activity in Vascular Smooth Muscle Cells. Circ. Res. 2009, 104, 69–78. [Google Scholar] [CrossRef] [Green Version]
  167. Ricote, M.; Glass, C.K. PPARs and molecular mechanisms of transrepression. Biochim. Biophys. Acta 2007, 1771, 926–935. [Google Scholar] [CrossRef] [Green Version]
  168. Ghisletti, S.; Huang, W.; Ogawa, S.; Pascual, G.; Lin, M.-E.; Willson, T.M.; Rosenfeld, M.G.; Glass, C.K. Parallel SUMOylation-Dependent Pathways Mediate Gene- and Signal-Specific Transrepression by LXRs and PPARγ. Mol. Cell 2007, 25, 57–70. [Google Scholar] [CrossRef] [Green Version]
  169. Khalid, A.M.; Hafstad, A.D.; Larsen, T.S.; Severson, D.L.; Boardman, N.; Hagve, M.; Berge, R.K.; Aasum, E. Cardioprotective effect of the PPAR ligand tetradecylthioacetic acid in type 2 diabetic mice. Am. J. Physiol. Heart Circ. Physiol. 2011, 300, 2116–2122. [Google Scholar] [CrossRef] [Green Version]
  170. Chandra, M.; Miriyala, S.; Panchatcharam, M. PPAR γ and Its Role in Cardiovascular Diseases. PPAR Res. 2017, 2017, 1–10. [Google Scholar] [CrossRef] [Green Version]
  171. Tokutome, M.; Matoba, T.; Nakano, Y.; Okahara, A.; Fujiwara, M.; Koga, J.I.; Nakano, K.; Tsutsui, H.; Egashira, K. Peroxisome proliferator-activated receptor-gamma targeting nanomedicine promotes cardiac healing after acute myocardial infarction by skewing monocyte/macrophage polarization in preclinical animal models. Cardiovasc. Res. 2019, 115, 419–431. [Google Scholar] [CrossRef]
  172. Zong, X.; Cheng, K.; Yin, G.; Wu, Z.; Su, Q.; Yu, D.; Liao, P.; Hu, W.; Chen, Y. SIRT3 is a downstream target of PPAR-α implicated in high glucose-induced cardiomyocyte injury in AC16 cells. Exp. Ther. Med. 2020, 20, 1261–1268. [Google Scholar] [CrossRef]
  173. Yang, M.; Xiong, J.; Zou, Q.; Wang, D.D.; Huang, C.X. Chrysin attenuates interstitial fibrosis and improves cardiac function in a rat model of acute myocardial infarction. J. Mol. Histol. 2018, 49, 555–565. [Google Scholar] [CrossRef]
  174. Shen, S.; Jiang, H.; Bei, Y.; Zhang, J.; Zhang, H.; Zhu, H.; Zhang, C.; Yao, W.; Wei, C.; Shang, H.; et al. Qiliqiangxin Attenuates Adverse Cardiac Remodeling after Myocardial Infarction in Ovariectomized Mice via Activation of PPARγ. Cell. Physiol. Biochem. 2017, 42, 876–888. [Google Scholar] [CrossRef]
  175. Mahajan, U.B.; Chandrayan, G.; Patil, C.R.; Arya, D.S.; Suchal, K.; Agrawal, Y.O.; Ojha, S.; Goyal, S.N. The protective effect of apigenin on myocardial injury in diabetic rats mediating activation of the PPAR-γ pathway. Int. J. Mol. Sci. 2017, 18, 756. [Google Scholar] [CrossRef] [Green Version]
  176. Lv, F.H.; Yin, H.L.; He, Y.Q.; Wu, H.M.; Kong, J.; Chai, X.Y.; Zhang, S.R. Effects of curcumin on the apoptosis of cardiomyocytes and the expression of NF-κB, PPAR-γ and Bcl-2 in rats with myocardial infarction injury. Exp. Ther. Med. 2016, 12, 3877–3884. [Google Scholar] [CrossRef] [Green Version]
  177. Shen, Z.X.; Yang, Q.Z.; Li, C.; Du, L.J.; Sun, X.N.; Liu, Y.; Sun, J.Y.; Gu, H.H.; Sun, Y.M.; Wang, J.; et al. Myeloid peroxisome proliferator-activated receptor gamma deficiency aggravates myocardial infarction in mice. Atherosclerosis 2018, 274, 199–205. [Google Scholar] [CrossRef]
  178. El-Gohary, O.A.; Allam, M.M. Effect of vitamin D on isoprenaline-induced myocardial infarction in rats: Possible role of peroxisome proliferator-activated receptor-γ. Can. J. Physiol. Pharmacol. 2017, 95, 641–646. [Google Scholar] [CrossRef]
  179. Chu, X.; Wang, Y.; Pang, L.; Huang, J.; Sun, X.; Chen, X. miR-130 aggravates acute myocardial infarction-induced myocardial injury by targeting PPAR-γ. J. Cell. Biochem. 2018, 119, 7235–7244. [Google Scholar] [CrossRef]
  180. Zhu, Z.-D.; Ye, J.-Y.; Niu, H.; Ma, Y.-M.; Fu, X.-M.; Xia, Z.-H.; Zhang, X. Effects of microRNA-292-5p on myocardial ischemia-reperfusion injury through the peroxisome proliferator-activated receptor-α/-γ signaling pathway. Gene Ther. 2018, 25, 234–248. [Google Scholar] [CrossRef]
  181. Zhou, H.; Li, D.; Zhu, P.; Hu, S.; Hu, N.; Ma, S.; Zhang, Y.; Han, T.; Ren, J.; Cao, F.; et al. Melatonin suppresses platelet activation and function against cardiac ischemia/reperfusion injury via PPARγ/FUNDC1/mitophagy pathways. J. Pineal Res. 2017, 63, 1–18. [Google Scholar] [CrossRef]
  182. Mokhtari, B.; Azizi, Y.; Abookheili, A.R.; Aboutaleb, N.; Nazarinia, D.; Naderi, N. Human amniotic membrane mesenchymal stem cells-conditioned medium attenuates myocardial ischemia-reperfusion injury in rats by targeting oxidative stress. Iran. J. Basic Med. Sci. 2020, 23, 1453–1461. [Google Scholar] [CrossRef]
  183. Nasser, M.I.; Qi, X.; Zhu, S.; He, Y.; Zhao, M.; Guo, H.; Zhu, P. Current situation and future of stem cells in cardiovascular medicine. Biomed. Pharmacother. 2020, 132, 110813. [Google Scholar] [CrossRef]
  184. Hou, J.; Wang, L.; Hou, J.; Guo, T.; Xing, Y.; Zheng, S.; Zhou, C.; Huang, H.; Long, H.; Zhong, T.; et al. Peroxisome Proliferator-Activated Receptor Gamma Promotes Mesenchymal Stem Cells to Express Connexin43 via the Inhibition of TGF-β1/Smads Signaling in a Rat Model of Myocardial Infarction. Stem Cell Rev. Rep. 2015, 11, 885–899. [Google Scholar] [CrossRef]
  185. Peymani, M.; Ghaedi, K.; Irani, S.; Nasr-Esfahani, M.H. Peroxisome Proliferator-Activated Receptor γ Activity is Required for Appropriate Cardiomyocyte Differentiation. Cell J. 2016, 18, 221–228. [Google Scholar] [CrossRef]
  186. Khan, V.; Sharma, S.; Bhandari, U.; Sharma, N.; Rishi, V.; Haque, S.E. Suppression of isoproterenol-induced cardiotoxicity in rats by raspberry ketone via activation of peroxisome proliferator activated receptor-α. Eur. J. Pharmacol. 2019, 842, 157–166. [Google Scholar] [CrossRef]
  187. Song, J.W.; Kim, H.J.; Lee, H.; Kim, J.W.; Kwak, Y.L. Protective effect of peroxisome proliferator-activated receptor α activation against cardiac ischemia-reperfusion injury is related to upregulation of uncoupling protein-3. Oxid. Med. Cell. Longev. 2016, 2016, 3539649. [Google Scholar] [CrossRef] [Green Version]
  188. Duerr, G.D.; Heinemann, J.C.; Arnoldi, V.; Feisst, A.; Kley, J.; Ghanem, A.; Welz, A.; Dewald, O. Cardiomyocyte specific peroxisome proliferator-activated receptor-α overexpression leads to irreversible damage in ischemic murine heart. Life Sci. 2014, 102, 88–97. [Google Scholar] [CrossRef]
  189. Wagner, K.D.; Vukolic, A.; Baudouy, D.; Michiels, J.F.; Wagner, N. Inducible conditional vascular-specific overexpression of peroxisome proliferator-activated receptor beta/delta leads to rapid cardiac hypertrophy. PPAR Res. 2016, 2016, 7631085. [Google Scholar] [CrossRef] [Green Version]
  190. Villapol, S. Roles of Peroxisome Proliferator-Activated Receptor Gamma on Brain and Peripheral Inflammation. Cell. Mol. Neurobiol. 2018, 38, 121–132. [Google Scholar] [CrossRef]
  191. Wójtowicz, S.; Strosznajder, A.K.; Jeżyna, M.; Strosznajder, J.B. The Novel Role of PPAR Alpha in the Brain: Promising Target in Therapy of Alzheimer’s Disease and Other Neurodegenerative Disorders. Neurochem. Res. 2020, 45, 972–988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Strosznajder, A.K.; Wójtowicz, S.; Jeżyna, M.J.; Sun, G.Y.; Strosznajder, J.B. Recent Insights on the Role of PPAR-β/δ in Neuroinflammation and Neurodegeneration, and Its Potential Target for Therapy. Neuromol. Med. 2021, 23, 86–98. [Google Scholar] [CrossRef] [PubMed]
  193. Liu, J.; Wang, L.N. Peroxisome proliferator-activated receptor gamma agonists for preventing recurrent stroke and other vascular events in people with stroke or transient ischaemic attack. Cochrane Database Syst. Rev. 2019, 10, CD010693. [Google Scholar] [CrossRef]
  194. Woo, M.H.; Lee, H.S.; Kim, J. Effect of pioglitazone in acute ischemic stroke patients with diabetes mellitus: A nested case-control study. Cardiovasc. Diabetol. 2019, 18, 1–10. [Google Scholar] [CrossRef] [PubMed]
  195. Chen, Y.; Liu, S.; Chen, G. Aggravation of Cerebral Ischemia/Reperfusion Injury by Peroxisome Proliferator-Activated Receptor-Gamma Deficiency via Endoplasmic Reticulum Stress. Med. Sci. Monit. 2019, 25, 7518–7526. [Google Scholar] [CrossRef] [PubMed]
  196. Song, J.; Kim, Y.S.; Lee, D.H.; Lee, S.H.; Park, H.J.; Lee, D.; Kim, H. Neuroprotective effects of oleic acid in rodent models of cerebral ischaemia. Sci. Rep. 2019, 9, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Han, L.; Cai, W.; Mao, L.; Liu, J.; Li, P.; Leak, R.K.; Xu, Y.; Hu, X.; Chen, J. Rosiglitazone Promotes White Matter Integrity and Long-Term Functional Recovery After Focal Cerebral Ischemia. Stroke 2015, 46, 2628–2636. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Luo, Y.; He, Q.; Kuang, G.; Jiang, Q.; Yang, J. PPAR-alpha and PPAR-beta expression changes in the hippocampus of rats undergoing global cerebral ischemia/reperfusion due to PPAR-gamma status. Behav. Brain Funct. 2014, 10, 21. [Google Scholar] [CrossRef] [Green Version]
  199. Guo, T.; Wang, Y.; Guo, Y.; Wu, S.; Chen, W.; Liu, N.; Wang, Y.; Geng, D. 1, 25-D 3 protects from cerebral ischemia by maintaining bbb permeability via PPAR-γ activation. Front. Cell. Neurosci. 2018, 12, 1–9. [Google Scholar] [CrossRef]
  200. Pan, J.; Jin, J.; Ge, H.; Yin, K.; Chen, X.; Han, L.; Chen, Y.; Qian, L.; Li, X.; Xu, Y. Malibatol A regulates microglia M1/M2 polarization in experimental stroke in a PPARγ-dependent manner. J. Neuroinflamm. 2015, 12, 51. [Google Scholar] [CrossRef] [Green Version]
  201. Kossatz, E.; Silva-Peña, D.; Suárez, J.; de Fonseca, F.R.; Maldonado, R.; Robledo, P. Octadecylpropyl sulfamide reduces neurodegeneration and restores the memory deficits induced by hypoxia-ischemia in mice. Front. Pharmacol. 2018, 9, 376. [Google Scholar] [CrossRef] [Green Version]
  202. Li, Y.; Zhu, Z.Y.; Lu, B.W.; Huang, T.T.; Zhang, Y.M.; Zhou, N.Y.; Xuan, W.; Chen, Z.A.; Wen, D.X.; Yu, W.F.; et al. Rosiglitazone ameliorates tissue plasminogen activator-induced brain hemorrhage after stroke. CNS Neurosci. Ther. 2019, 25, 1343–1352. [Google Scholar] [CrossRef]
  203. Hasegawa, H.; Yatomi, K.; Mitome-Mishima, Y.; Miyamoto, N.; Tanaka, R.; Oishi, H.; Arai, H.; Hattori, N.; Urabe, T. Pioglitazone Prevents Hemorrhagic Infarction After Transient Focal Ischemia in Type 2 Diabetes. Neurosci. Res. 2021, 170, 314–321. [Google Scholar] [CrossRef]
  204. Li, Y.Y.; Guo, J.H.; Liu, Y.Q.; Dong, J.H.; Zhu, C.H. PPARγ Activation-Mediated Egr-1 Inhibition Benefits Against Brain Injury in an Experimental Ischaemic Stroke Model. J. Stroke Cerebrovasc. Dis. 2020, 29, 1–6. [Google Scholar] [CrossRef]
  205. Wu, X.J.; Sun, X.H.; Wang, S.W.; Chen, J.L.; Bi, Y.H.; Jiang, D.X. Mifepristone alleviates cerebral ischemiareperfusion injury in rats by stimulating PPAR γ. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 5688–5696. [Google Scholar] [CrossRef]
  206. Boujon, V.; Uhlemann, R.; Wegner, S.; Wright, M.B.; Laufs, U.; Endres, M.; Kronenberg, G.; Gertz, K. Dual PPARα/γ agonist aleglitazar confers stroke protection in a model of mild focal brain ischemia in mice. J. Mol. Med. 2019, 97, 1127–1138. [Google Scholar] [CrossRef] [Green Version]
  207. Li, Y.; Xu, L.; Zeng, K.; Xu, Z.; Suo, D.; Peng, L.; Ren, T.; Sun, Z.; Yang, W.; Jin, X.; et al. Propane-2-sulfonic acid octadec-9-enyl-amide, a novel PPARα/γ dual agonist, protects against ischemia-induced brain damage in mice by inhibiting inflammatory responses. Brain. Behav. Immun. 2017, 66, 289–301. [Google Scholar] [CrossRef]
  208. Deng, Y.; Xiong, D.; Yin, C.; Liu, B.; Shi, J.; Gong, Q. Icariside II protects against cerebral ischemia-reperfusion injury in rats via nuclear factor-κB inhibition and peroxisome proliferator-activated receptor up-regulation. Neurochem. Int. 2016, 96, 56–61. [Google Scholar] [CrossRef]
  209. Xiong, D.; Deng, Y.; Huang, B.; Yin, C.; Liu, B.; Shi, J.; Gong, Q. Icariin attenuates cerebral ischemia-reperfusion injury through inhibition of inflammatory response mediated by NF-κB, PPARα and PPARγ in rats. Int. Immunopharmacol. 2016, 30, 157–162. [Google Scholar] [CrossRef]
  210. Shehata, A.H.F.; Ahmed, A.S.F.; Abdelrehim, A.B.; Heeba, G.H. The impact of single and combined PPAR-α and PPAR-γ activation on the neurological outcomes following cerebral ischemia reperfusion. Life Sci. 2020, 252, 117679. [Google Scholar] [CrossRef]
  211. Li, J.; Zhang, K.; Zhang, Q.; Zhou, X.; Wen, L.; Ma, J.; Niu, L.; Li, C. PPAR- γ Mediates Ta-VNS-Induced Angiogenesis and Subsequent Functional Recovery after Experimental Stroke in Rats. Biomed. Res. Int. 2020, 2020, 8163789. [Google Scholar] [CrossRef]
  212. Jiang, Y.; Li, L.; Liu, B.; Zhang, Y.; Chen, Q.; Li, C. PPARγ upregulation induced by vagus nerve stimulation exerts anti-inflammatory effect in cerebral ischemia/reperfusion rats. Med. Sci. Monit. 2015, 21, 268–275. [Google Scholar] [CrossRef] [Green Version]
  213. Liu, B.; Zhang, Y.; Jiang, Y.; Li, L.; Li, C.; Li, J. Electrical stimulation of cerebellar fastigial nucleus protects against cerebral ischemic injury by PPARγ upregulation. Neurol. Res. 2017, 39, 23–29. [Google Scholar] [CrossRef]
  214. Kinouchi, T.; Kitazato, K.T.; Shimada, K.; Yagi, K.; Tada, Y.; Matsushita, N.; Kurashiki, Y.; Satomi, J.; Sata, M.; Nagahiro, S. Treatment with the PPARγ Agonist Pioglitazone in the Early Post-ischemia Phase Inhibits Pro-inflammatory Responses and Promotes Neurogenesis Via the Activation of Innate- and Bone Marrow-Derived Stem Cells in Rats. Transl. Stroke Res. 2018, 9, 306–316. [Google Scholar] [CrossRef] [Green Version]
  215. Gamdzyk, M.; Doycheva, D.M.; Malaguit, J.; Enkhjargal, B.; Tang, J.; Zhang, J.H. Role of PPAR-β/δ/miR-17/TXNIP pathway in neuronal apoptosis after neonatal hypoxic-ischemic injury in rats. Neuropharmacology 2018, 140, 150–161. [Google Scholar] [CrossRef]
  216. Pei, L.; Meng, S.; Yu, W.; Wang, Q.; Song, F.; Ma, L. Inhibition of MicroRNA-383 Ameliorates Injury after Focal Cerebral Ischemia via Targeting PPARγ. Cell. Physiol. Biochem. 2016, 39, 1339–1346. [Google Scholar] [CrossRef] [Green Version]
  217. Ting, S.-M.; Zhao, X.; Sun, G.; Obertas, L.; Ricote, M.; Aronowski, J. Brain Cleanup as a Potential Target for Poststroke Recovery. Stroke 2020, 51, 958–966. [Google Scholar] [CrossRef]
  218. Zuo, Y.; Huang, L.; Enkhjargal, B.; Xu, W.; Umut, O.; Travis, Z.D.; Zhang, G.; Tang, J.; Liu, F.; Zhang, J.H. Activation of retinoid X receptor by bexarotene attenuates neuroinflammation via PPARγ/SIRT6/FoxO3a pathway after subarachnoid hemorrhage in rats. J. Neuroinflamm. 2019, 16, 1–15. [Google Scholar] [CrossRef] [Green Version]
  219. Micka, J.; Milatovich, A.; Menon, A.; Grabowski, G.A.; Puga, A.; Nebert, D.W. Human Ah receptor (AHR) gene: Localization to 7p15 and suggestive correlation of polymorphism with CYP1A1 inducibility. Pharmacogenetics 1997, 7, 95–100. [Google Scholar] [CrossRef]
  220. Williams, E.G.; Mouchiroud, L.; Frochaux, M.; Pandey, A.; Andreux, P.A.; Deplancke, B.; Auwerx, J. An Evolutionarily Conserved Role for the Aryl Hydrocarbon Receptor in the Regulation of Movement. PLoS Genet. 2014, 10, e1004673. [Google Scholar] [CrossRef] [Green Version]
  221. Lahvis, G.P.; Pyzalski, R.W.; Glover, E.; Pitot, H.C.; McElwee, M.K.; Bradfield, C.A. The aryl hydrocarbon receptor is required for developmental closure of the ductus venosus in the neonatal mouse. Mol. Pharmacol. 2005, 67, 714–720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Lund, A.K.; Goens, M.B.; Kanagy, N.L.; Walker, M.K. Cardiac hypertrophy in Aryl hydrocarbon receptor null mice is correlated with elevated angiotensin II, endothelin-1, and mean arterial blood pressure. Toxicol. Appl. Pharmacol. 2003, 193, 177–187. [Google Scholar] [CrossRef] [PubMed]
  223. Lund, A.K.; Peterson, S.L.; Timmins, G.S.; Walker, M.K. Endothelin-1-Mediated Increase in Reactive Oxygen Species and NADPH Oxidase Activity in Hearts of Aryl Hydrocarbon Receptor (AhR) Null Mice. Toxicol. Sci. 2005, 88, 265–273. [Google Scholar] [CrossRef] [PubMed]
  224. Ishihara, Y.; Haarmann-Stemmann, T.; Kado, N.Y.; Vogel, C.F.A. Interleukin 33 Expression Induced by Aryl Hydrocarbon Receptor in Macrophages. Toxicol. Sci. 2019, 170, 404–414. [Google Scholar] [CrossRef]
  225. Lew, B.J.; Manickam, R.; Lawrence, B.P. Activation of the Aryl hydrocarbon receptor during pregnancy in the mouse alters mammary development through direct effects on stromal and epithelial tissues. Biol. Reprod. 2011, 84, 1094–1102. [Google Scholar] [CrossRef] [Green Version]
  226. Quintana, F.J.; Basso, A.S.; Iglesias, A.H.; Korn, T.; Farez, M.F.; Bettelli, E.; Caccamo, M.; Oukka, M.; Weiner, H.L. Control of Treg and TH17 cell differentiation by the aryl hydrocarbon receptor. Nature 2008, 453, 65–71. [Google Scholar] [CrossRef]
  227. Gutiérrez-Vázquez, C.; Quintana, F.J. Regulation of the Immune Response by the Aryl Hydrocarbon Receptor. Immunity 2018, 48, 19–33. [Google Scholar] [CrossRef] [Green Version]
  228. Wang, Q.; Hisaka, K.; Carreira, V.; Ko, C.I.; Fan, Y.; Zhang, X.; Biesiada, J.; Medvedovic, M.; Puga, A. Ah receptor activation by dioxin disrupts activin, BMP, and WNT signals during the early differentiation of mouse embryonic stem cells and inhibits cardiomyocyte functions. Toxicol. Sci. 2016, 149, 346–357. [Google Scholar] [CrossRef] [Green Version]
  229. Singh, K.P.; Bennett, J.A.; Casado, F.L.; Walrath, J.L.; Welle, S.L.; Gasiewicz, T.A. Loss of aryl hydrocarbon receptor promotes gene changes associated with premature hematopoietic stem cell exhaustion and development of a myeloproliferative disorder in aging mice. Stem Cells Dev. 2014, 23, 95–106. [Google Scholar] [CrossRef] [Green Version]
  230. Chevallier, A.; Mialot, A.; Petit, J.M.; Fernandez-Salguero, P.; Barouki, R.; Coumoul, X.; Beraneck, M. Oculomotor Deficits in Aryl Hydrocarbon Receptor Null Mouse. PLoS ONE 2013, 8, e53520. [Google Scholar] [CrossRef] [Green Version]
  231. Chen, W.C.; Chang, L.H.; Huang, S.S.; Huang, Y.J.; Chih, C.L.; Kuo, H.C.; Lee, Y.H.; Lee, I.H. Aryl hydrocarbon receptor modulates stroke-induced astrogliosis and neurogenesis in the adult mouse brain. J. Neuroinflamm. 2019, 16, 1–13. [Google Scholar] [CrossRef]
  232. Lin, C.H.; Chen, C.C.; Chou, C.M.; Wang, C.Y.; Hung, C.C.; Chen, J.Y.; Chang, H.W.; Chen, Y.C.; Yeh, G.C.; Lee, Y.H. Knockdown of the aryl hydrocarbon receptor attenuates excitotoxicity and enhances NMDA-induced BDNF expression in cortical neurons. J. Neurochem. 2009, 111, 777–789. [Google Scholar] [CrossRef]
  233. Dever, D.P.; Adham, Z.O.; Thompson, B.; Genestine, M.; Cherry, J.; Olschowka, J.A.; DiCicco-Bloom, E.; Opanashuk, L.A. Aryl hydrocarbon receptor deletion in cerebellar granule neuron precursors impairs neurogenesis. Dev. Neurobiol. 2016, 76, 533–550. [Google Scholar] [CrossRef]
  234. Zhu, C.; Chen, Y.L.; Wang, X.J.; Hu, X.S.; Yu, Z.B.; Han, S.P. ShRNA-mediated gene silencing of AHR promotes the differentiation of P19 mouse embryonic carcinoma cells into cardiomyocytes. Mol. Med. Rep. 2012, 6, 513–518. [Google Scholar] [CrossRef] [Green Version]
  235. Wang, Q.; Chen, J.; Ko, C.I.; Fan, Y.; Carreira, V.; Chen, Y.; Xia, Y.; Medvedovic, M.; Puga, A. Disruption of aryl hydrocarbon receptor homeostatic levels during embryonic stem cell differentiation alters expression of homeobox transcription factors that control cardiomyogenesis. Environ. Health Perspect. 2013, 121, 1334–1343. [Google Scholar] [CrossRef] [Green Version]
  236. Zhang, Y.; Wang, Y.; Ma, Z.; Liang, Q.; Tang, X.; Tan, H.; Xiao, C.; Gao, Y. Ginsenoside Rb1 inhibits doxorubicin-triggered H9C2 cell apoptosis via aryl hydrocarbon receptor. Biomol. Ther. 2017, 25, 202–212. [Google Scholar] [CrossRef] [Green Version]
  237. Xue, Y.; Shui, X.; Su, W.; He, Y.; Lu, X.; Zhang, Y.; Yan, G.; Huang, S.; Lei, W.; Chen, C. Baicalin inhibits inflammation and attenuates myocardial ischaemic injury by aryl hydrocarbon receptor. J. Pharm. Pharmacol. 2015, 67, 1756–1764. [Google Scholar] [CrossRef]
  238. Wu, P.Y.; Yu, I.S.; Lin, Y.C.; Chang, Y.T.; Chen, C.C.; Lin, K.H.; Tseng, T.H.; Kargren, M.; Tai, Y.L.; Shen, T.L.; et al. Activation of aryl hydrocarbon receptor by kynurenine impairs progression and metastasis of neuroblastoma. Cancer Res. 2019, 79, 5550–5562. [Google Scholar] [CrossRef] [Green Version]
  239. Thatcher, T.H.; Williams, M.A.; Pollock, S.J.; Mccarthy, C.E.; Lacy, S.H.; Phipps, R.P.; Sime, P.J. Endogenous ligands of the aryl hydrocarbon receptor regulate lung dendritic cell function. Immunology 2016, 147, 41–54. [Google Scholar] [CrossRef] [Green Version]
  240. Safe, S.; Jin, U.H.; Park, H.; Chapkin, R.S.; Jayaraman, A. Aryl hydrocarbon receptor (AHR) ligands as selective ahr modulators (SAHRMS). Int. J. Mol. Sci. 2020, 21, 6654. [Google Scholar] [CrossRef]
  241. Vasquez, A.; Atallah-Yunes, N.; Smith, F.C.; You, X.; Chase, S.E.; Silverstone, A.E.; Vikstrom, K.L. A role for the aryl hydrocarbon receptor in cardiac physiology and function as demonstrated by AhR knockout mice. Cardiovasc. Toxicol. 2003, 3, 153–163. [Google Scholar] [CrossRef]
  242. Thackaberry, E.A.; Gabaldon, D.M.; Walker, M.K.; Smith, S.M. Aryl hydrocarbon receptor null mice develop cardiac hypertrophy and increased hypoxia-inducible factor-1α in the absence of cardiac hypoxia. Cardiovasc. Toxicol. 2002, 2, 263–273. [Google Scholar] [CrossRef]
  243. Stegeman, J.J.; Hahn, M.E.; Weisbrod, R.; Woodin, B.R.; Joy, J.S.; Najibi, S.; Cohen, R.A. Induction of cytochrome P4501A1 by aryl hydrocarbon receptor agonists in porcine aorta endothelial cells in culture and cytochrome P4501A1 activity in intact cells. Mol. Pharmacol. 1995, 47, 296–306. [Google Scholar]
  244. Kerzee, J.K.; Ramos, K.S. Constitutive and inducible expression of Cyp1a1 and Cyp1b1 in vascular smooth muscle cells: Role of the Ahr bHLH/PAS transcription factor. Circ. Res. 2001, 89, 573–582. [Google Scholar] [CrossRef] [Green Version]
  245. Thum, T.; Borlak, J. Gene expression in distinct regions of the heart. Lancet 2000, 355, 979–983. [Google Scholar] [CrossRef]
  246. Lighthouse, J.K.; Burke, R.M.; Velasquez, L.S.; Dirkx, R.A.; Aiezza, A.; Moravec, C.S.; Alexis, J.D.; Rosenberg, A.; Small, E.M. Exercise promotes a cardioprotective gene program in resident cardiac fibroblasts. JCI Insight 2019, 4, e92098. [Google Scholar] [CrossRef] [Green Version]
  247. Goudot, C.; Coillard, A.; Villani, A.-C.; Gueguen, P.; Cros, A.; Sarkizova, S.; Tang-Huau, T.-L.; Bohec, M.; Baulande, S.; Hacohen, N.; et al. Aryl Hydrocarbon Receptor Controls Monocyte Differentiation into Dendritic Cells versus Macrophages. Immunity 2017, 47, 582–596.e6. [Google Scholar] [CrossRef] [Green Version]
  248. Carreira, V.S.; Fan, Y.; Wang, Q.; Zhang, X.; Kurita, H.; Ko, C.I.; Naticchioni, M.; Jiang, M.; Koch, S.; Medvedovic, M.; et al. Ah receptor signaling controls the expression of cardiac development and homeostasis genes. Toxicol. Sci. 2015, 147, 425–435. [Google Scholar] [CrossRef] [Green Version]
  249. Kimura, E.; Tohyama, C. Embryonic and postnatal expression of aryl hydrocarbon receptor mRNA in mouse brain. Front. Neuroanat. 2017, 11, 4. [Google Scholar] [CrossRef] [Green Version]
  250. Filbrandt, C.R.; Wu, Z.; Zlokovic, B.; Opanashuk, L.; Gasiewicz, T.A. Presence and functional activity of the aryl hydrocarbon receptor in isolated murine cerebral vascular endothelial cells and astrocytes. Neurotoxicology 2004, 25, 605–616. [Google Scholar] [CrossRef]
  251. Lee, Y.H.; Lin, C.H.; Hsu, P.C.; Sun, Y.Y.; Huang, Y.J.; Zhuo, J.H.; Wang, C.Y.; Gan, Y.L.; Hung, C.C.; Kuan, C.Y.; et al. Aryl hydrocarbon receptor mediates both proinflammatory and anti-inflammatory effects in lipopolysaccharide-activated microglia. Glia 2015, 63, 1138–1154. [Google Scholar] [CrossRef] [PubMed]
  252. Tanaka, M.; Fujikawa, M.; Oguro, A.; Itoh, K.; Vogel, C.F.A.; Ishihara, Y. Involvement of the Microglial Aryl Hydrocarbon Receptor in Neuroinflammation and Vasogenic Edema after Ischemic Stroke. Cells 2021, 10, 718. [Google Scholar] [CrossRef] [PubMed]
  253. Kajta, M.; Wnuk, A.; Rzemieniec, J.; Lason, W.; Mackowiak, M.; Chwastek, E.; Staniszewska, M.; Nehring, I.; Wojtowicz, A.K. Triclocarban Disrupts the Epigenetic Status of Neuronal Cells and Induces AHR/CAR-Mediated Apoptosis. Mol. Neurobiol. 2019, 56, 3113–3131. [Google Scholar] [CrossRef] [PubMed]
  254. Wnuk, A.; Rzemieniec, J.; Przepiórska, K.; Wesołowska, J.; Wójtowicz, A.K.; Kajta, M. Autophagy-related neurotoxicity is mediated via AHR and CAR in mouse neurons exposed to DDE. Sci. Total Environ. 2020, 742, 140599. [Google Scholar] [CrossRef] [PubMed]
  255. Lin, C.-H.; Juan, S.-H.; Wang, C.Y.; Sun, Y.-Y.; Chou, C.-M.; Chang, S.-F.; Hu, S.-Y.; Lee, W.-S.; Lee, Y.-H. Neuronal activity enhances aryl hydrocarbon receptor-mediated gene expression and dioxin neurotoxicity in cortical neurons. J. Neurochem. 2008, 104, 1415–1429. [Google Scholar] [CrossRef]
  256. Wu, P.-Y.; Chuang, P.-Y.; Chang, G.-D.; Chan, Y.-Y.; Tsai, T.-C.; Wang, B.-J.; Lin, K.-H.; Hsu, W.-M.; Liao, Y.-F.; Lee, H. Novel Endogenous Ligands of Aryl Hydrocarbon Receptor Mediate Neural Development and Differentiation of Neuroblastoma. ACS Chem. Neurosci. 2019, 10, 4031–4042. [Google Scholar] [CrossRef]
  257. Ramos-García, N.A.; Orozco-Ibarra, M.; Estudillo, E.; Elizondo, G.; Apo, E.G.; Macías, L.G.C.; Sosa-Ortiz, A.L.; Torres-Ramos, M.A. Aryl hydrocarbon receptor in post-mortem hippocampus and in serum from young, elder, and Alzheimer’s patients. Int. J. Mol. Sci. 2020, 21, 1983. [Google Scholar] [CrossRef] [Green Version]
  258. Rothhammer, V.; Quintana, F.J. The aryl hydrocarbon receptor: An environmental sensor integrating immune responses in health and disease. Nat. Rev. Immunol. 2019, 19, 184–197. [Google Scholar] [CrossRef]
  259. Mulero-Navarro, S.; Fernandez-Salguero, P.M. New trends in Aryl hydrocarbon receptor biology. Front. Cell Dev. Biol. 2016, 4, 45. [Google Scholar] [CrossRef] [Green Version]
  260. Juricek, L.; Coumoul, X. The aryl hydrocarbon receptor and the nervous system. Int. J. Mol. Sci. 2018, 19, 2504. [Google Scholar] [CrossRef] [Green Version]
  261. Ohtake, F.; Takeyama, K.i.; Matsumoto, T.; Kitagawa, H.; Yamamoto, Y.; Nohara, K.; Tohyama, C.; Krust, A.; Mimura, J.; Chambon, P.; et al. Modulation of oestrogen receptor signalling by association with the activated dioxin receptor. Nature 2003, 423, 545–550. [Google Scholar] [CrossRef]
  262. Kimura, A.; Naka, T.; Nohara, K.; Fujii-Kuriyama, Y.; Kishimoto, T. Aryl hydrocarbon receptor regulates Stat1 activation and participates in the development of Th17 cells. Proc. Natl. Acad. Sci. USA 2008, 105, 9721–9726. [Google Scholar] [CrossRef] [Green Version]
  263. Lee, B.J.; Kim, B.; Lee, K. Air pollution exposure and cardiovascular disease. Toxicol. Res. 2014, 30, 71–75. [Google Scholar] [CrossRef]
  264. Miller, M.R.; Newby, D.E. Air pollution and cardiovascular disease: Car sick. Cardiovasc. Res. 2020, 116, 279–294. [Google Scholar] [CrossRef]
  265. Marris, C.R.; Kompella, S.N.; Miller, M.R.; Incardona, J.P.; Brette, F.; Hancox, J.C.; Sørhus, E.; Shiels, H.A. Polyaromatic hydrocarbons in pollution: A heart-breaking matter. J. Physiol. 2020, 598, 227–247. [Google Scholar] [CrossRef]
  266. Ko, C.I.; Fan, Y.; de Gannes, M.; Wang, Q.; Xia, Y.; Puga, A. Repression of the Aryl Hydrocarbon Receptor Is Required to Maintain Mitotic Progression and Prevent Loss of Pluripotency of Embryonic Stem Cells. Stem Cells 2016, 34, 2825–2839. [Google Scholar] [CrossRef] [Green Version]
  267. Vilahur, G.; Cubedo, J.; Casani, L.; Padro, T.; Sabate-Tenas, M.; Badimon, J.J.; Badimon, L. Reperfusion-triggered stress protein response in the myocardium is blocked by post-conditioning. Systems biology pathway analysis highlights the key role of the canonical aryl-hydrocarbon receptor pathway. Eur. Heart J. 2013, 34, 2082–2093. [Google Scholar] [CrossRef] [Green Version]
  268. Li, W.; Li, Y.; Sun, R.; Zhou, S.; Li, M.; Feng, M.; Xie, Y. Dual character of flavonoids in attenuating and aggravating ischemia-reperfusion-induced myocardial injury. Exp. Ther. Med. 2017, 14, 1307–1314. [Google Scholar] [CrossRef] [Green Version]
  269. Cuartero, M.I.; Ballesteros, I.; de la Parra, J.; Harkin, A.L.; Abautret-Daly, A.; Sherwin, E.; Fernández-Salguero, P.; Corbí, Á.L.; Lizasoain, I.; Moro, M.A. L-kynurenine/aryl hydrocarbon receptor pathway mediates brain damage after experimental stroke. Circulation 2014, 130, 2040–2051. [Google Scholar] [CrossRef] [Green Version]
  270. Kwon, J.I.; Heo, H.; Ham, S.J.; Chae, Y.J.; Lee, D.W.; Kim, S.T.; Min, J.; Sung, Y.S.; Kim, K.W.; Choi, Y.; et al. Aryl hydrocarbon receptor antagonism before reperfusion attenuates cerebral ischaemia/reperfusion injury in rats. Sci. Rep. 2020, 10, 14906. [Google Scholar] [CrossRef]
  271. Rzemieniec, J.; Litwa, E.; Wnuk, A.; Lason, W.; Krzeptowski, W.; Kajta, M. Selective Aryl Hydrocarbon Receptor Modulator 3,3′-Diindolylmethane Impairs AhR and ARNT Signaling and Protects Mouse Neuronal Cells Against Hypoxia. Mol. Neurobiol. 2016, 53, 5591–5606. [Google Scholar] [CrossRef]
  272. Rzemieniec, J.; Wnuk, A.; Lasoń, W.; Bilecki, W.; Kajta, M. The neuroprotective action of 3,3′-diindolylmethane against ischemia involves an inhibition of apoptosis and autophagy that depends on HDAC and AhR/CYP1A1 but not ERα/CYP19A1 signaling. Apoptosis 2019, 24, 435–452. [Google Scholar] [CrossRef] [Green Version]
  273. Rzemieniec, J.; Bratek, E.; Wnuk, A.; Przepiórska, K.; Salińska, E.; Kajta, M. Neuroprotective effect of 3,3’-Diindolylmethane against perinatal asphyxia involves inhibition of the AhR and NMDA signaling and hypermethylation of specific genes. Apoptosis 2020, 25, 747–762. [Google Scholar] [CrossRef]
  274. Matsumoto, K.; Kinoshita, K.; Yoshimizu, A.; Kurauchi, Y.; Hisatsune, A.; Seki, T.; Katsuki, H. Laquinimod and 3,3′-diindolylemethane alleviate neuropathological events and neurological deficits in a mouse model of intracerebral hemorrhage. J. Neuroimmunol. 2020, 342, 577195. [Google Scholar] [CrossRef]
Figure 1. Schematic model representing the cellular localization, the molecular mechanisms and the effects of estrogen receptors activation after cerebral and cardiac ischemia. Genomic pathway: estrogen agonist directly induces dimerization of ERs and translocation to the nucleus. The complex ligands-ERs can (1) bind the estrogen-response-element (ERE) of gene promoters and induce target gene transcription, or (2) interacts with other transcription factors (TF) (e.g., Fos and Jun). Non-genomic pathway: estrogen agonist binds cell membrane ERs (3) or GPER-1 (4) triggering rapid cytosolic phosphorylation cascades through membrane-associated proteins. These kinases can phosphorylate and activate either ERs (5), both in a ligand-dependent and in a ligand-independent manner, or its associated co-regulators (6), enhancing or inhibiting the genomic action of ERs. Activated kinases can also affect gene transcription through phosphorylation of several TFs (7). In addition, estrogen-agonist activates ERs located in mitochondria, regulating mitochondrial function (8).
Figure 1. Schematic model representing the cellular localization, the molecular mechanisms and the effects of estrogen receptors activation after cerebral and cardiac ischemia. Genomic pathway: estrogen agonist directly induces dimerization of ERs and translocation to the nucleus. The complex ligands-ERs can (1) bind the estrogen-response-element (ERE) of gene promoters and induce target gene transcription, or (2) interacts with other transcription factors (TF) (e.g., Fos and Jun). Non-genomic pathway: estrogen agonist binds cell membrane ERs (3) or GPER-1 (4) triggering rapid cytosolic phosphorylation cascades through membrane-associated proteins. These kinases can phosphorylate and activate either ERs (5), both in a ligand-dependent and in a ligand-independent manner, or its associated co-regulators (6), enhancing or inhibiting the genomic action of ERs. Activated kinases can also affect gene transcription through phosphorylation of several TFs (7). In addition, estrogen-agonist activates ERs located in mitochondria, regulating mitochondrial function (8).
Ijms 22 12326 g001
Figure 2. Schematic model representing the cellular localization, the molecular mechanisms and the effects of PPARs activation after cerebral and cardiac ischemia. Under unliganded state, monomer or dimer of PPAR is bound to multicomponent repressors (1). Ligand-dependent transactivation: ligand binding to either PPAR or RXR causes displacement of bound repressors, recruitment of co-activators and activation of gene transcription (2). Ligand-independent repression: PPARs bind to response elements in the absence of ligand and recruit multicomponent repressors that mediate active repression (3). Ligand-dependent repression is provided by several mechanisms: competition for a limiting pool of co-activators (4), inhibition of repressors clearance (5), direct interaction with other transcription factors (e.g., p65/p50) (6,7). Finally, several kinases can phosphorylate PPARs modulating its activity (e.g., PPAR-mediated transcription is increased by PKA, while is reduced by MAPK and Brc kinase) (8). ° the expression is increased after MI; * the expression is increased after LPS treatment.
Figure 2. Schematic model representing the cellular localization, the molecular mechanisms and the effects of PPARs activation after cerebral and cardiac ischemia. Under unliganded state, monomer or dimer of PPAR is bound to multicomponent repressors (1). Ligand-dependent transactivation: ligand binding to either PPAR or RXR causes displacement of bound repressors, recruitment of co-activators and activation of gene transcription (2). Ligand-independent repression: PPARs bind to response elements in the absence of ligand and recruit multicomponent repressors that mediate active repression (3). Ligand-dependent repression is provided by several mechanisms: competition for a limiting pool of co-activators (4), inhibition of repressors clearance (5), direct interaction with other transcription factors (e.g., p65/p50) (6,7). Finally, several kinases can phosphorylate PPARs modulating its activity (e.g., PPAR-mediated transcription is increased by PKA, while is reduced by MAPK and Brc kinase) (8). ° the expression is increased after MI; * the expression is increased after LPS treatment.
Ijms 22 12326 g002
Figure 3. Schematic model representing the cellular localization, the molecular mechanisms and the effects of AhR activation after cerebral and cardiac ischemia. Under unbounded state, AhR is retained in the cytoplasm in an inactive complex. After ligand binding, AIP is released from the complex and the ligand-AhR-HSP90-p23-Scr structure translocates into the nucleus. Inside the nucleus, the ligand-AhR structure is released from the complex and heterodimerizes with the aryl hydrocarbon receptor nuclear translocator (ARNT) and interacts with the xenobiotics response element (XRE) (1), regulating the expression of several phase I and phase II metabolizing enzymes. The ligand-AhR-ARNT (2) and ligand-AhR (3) can interact with other transcription factors (e.g., NF-kB and the estrogen receptor ER), binding to their response elements (RE) and modulating the expression of their target genes. AhR signaling also includes non-genomic pathways: AhR can function as an E3 ubiquitin ligase (4), while the release of the c-Src kinase (5) results in the phosphorylation of multiple targets.
Figure 3. Schematic model representing the cellular localization, the molecular mechanisms and the effects of AhR activation after cerebral and cardiac ischemia. Under unbounded state, AhR is retained in the cytoplasm in an inactive complex. After ligand binding, AIP is released from the complex and the ligand-AhR-HSP90-p23-Scr structure translocates into the nucleus. Inside the nucleus, the ligand-AhR structure is released from the complex and heterodimerizes with the aryl hydrocarbon receptor nuclear translocator (ARNT) and interacts with the xenobiotics response element (XRE) (1), regulating the expression of several phase I and phase II metabolizing enzymes. The ligand-AhR-ARNT (2) and ligand-AhR (3) can interact with other transcription factors (e.g., NF-kB and the estrogen receptor ER), binding to their response elements (RE) and modulating the expression of their target genes. AhR signaling also includes non-genomic pathways: AhR can function as an E3 ubiquitin ligase (4), while the release of the c-Src kinase (5) results in the phosphorylation of multiple targets.
Ijms 22 12326 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rzemieniec, J.; Castiglioni, L.; Gelosa, P.; Muluhie, M.; Mercuriali, B.; Sironi, L. Nuclear Receptors in Myocardial and Cerebral Ischemia—Mechanisms of Action and Therapeutic Strategies. Int. J. Mol. Sci. 2021, 22, 12326. https://doi.org/10.3390/ijms222212326

AMA Style

Rzemieniec J, Castiglioni L, Gelosa P, Muluhie M, Mercuriali B, Sironi L. Nuclear Receptors in Myocardial and Cerebral Ischemia—Mechanisms of Action and Therapeutic Strategies. International Journal of Molecular Sciences. 2021; 22(22):12326. https://doi.org/10.3390/ijms222212326

Chicago/Turabian Style

Rzemieniec, Joanna, Laura Castiglioni, Paolo Gelosa, Majeda Muluhie, Benedetta Mercuriali, and Luigi Sironi. 2021. "Nuclear Receptors in Myocardial and Cerebral Ischemia—Mechanisms of Action and Therapeutic Strategies" International Journal of Molecular Sciences 22, no. 22: 12326. https://doi.org/10.3390/ijms222212326

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