Next Article in Journal
Pathogenic Mechanism of Autoimmune Diabetes Mellitus in Humans: Potential Role of Streptozotocin-Induced Selective Autoimmunity against Human Islet β-Cells
Next Article in Special Issue
Neuroprotective Effects of TRPM7 Deletion in Parvalbumin GABAergic vs. Glutamatergic Neurons following Ischemia
Previous Article in Journal
Pioneer Role of Extracellular Vesicles as Modulators of Cancer Initiation in Progression, Drug Therapy, and Vaccine Prospects
Previous Article in Special Issue
The TRPA1 Channel Amplifies the Oxidative Stress Signal in Melanoma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Systemic Review of the Integral Role of TRPM2 in Ischemic Stroke: From Upstream Risk Factors to Ultimate Neuronal Death

1
Department of Cell Biology, Calhoun Cardiology Center, University of Connecticut School of Medicine (UConnHealth), Farmington, CT 06030, USA
2
Department of Ecology and Evolutionary Biology, University of Connecticut, Storrs, CT 06269, USA
*
Author to whom correspondence should be addressed.
Cells 2022, 11(3), 491; https://doi.org/10.3390/cells11030491
Submission received: 18 December 2021 / Revised: 26 January 2022 / Accepted: 29 January 2022 / Published: 31 January 2022
(This article belongs to the Special Issue TRP Channels in Oxidative Stress Signalling)

Abstract

:
Ischemic stroke causes a heavy health burden worldwide, with over 10 million new cases every year. Despite the high prevalence and mortality rate of ischemic stroke, the underlying molecular mechanisms for the common etiological factors of ischemic stroke and ischemic stroke itself remain unclear, which results in insufficient preventive strategies and ineffective treatments for this devastating disease. In this review, we demonstrate that transient receptor potential cation channel, subfamily M, member 2 (TRPM2), a non-selective ion channel activated by oxidative stress, is actively involved in all the important steps in the etiology and pathology of ischemic stroke. TRPM2 could be a promising target in screening more effective prophylactic strategies and therapeutic medications for ischemic stroke.

1. Introduction

1.1. Stroke

Globally, stroke is a common cause of death and mortality only secondary to myocardial infarction, and is the third-most-common cause of disability [1]. There are two different types of strokes, ischemic stroke and hemorrhagic stroke. Ischemic stroke is much more common than hemorrhagic stroke, accounting for over 80% of all the cases [1,2]. The pathological feature of ischemic stroke is occlusion of supplying arteries and subsequent tissue necrosis [3]. The occlusion of brain arteries can result from two causes, emboli formed outside the brain (extracranial) or thrombus formed in situ (intracranial) [2]. Emboli from atherosclerotic plaque on carotid arteries or cardiac thrombus from atria during atrial fibrillation is the most common cause [3,4]. In situ thrombosis resulting from local atherosclerosis is less common but still an important cause of ischemic stroke [3]. Small-vessel diseases caused by hypertension and diabetes can cause small subcortical infarcts [2,4].
Following occlusion, all the cells “downstream” to the occlusion site are subjected to ischemic conditions, including endothelial cells, glial cells, and neurons. Endothelial hyperpermeability leads to vasogenic edema and immune-cell infiltration, which significantly aggregate brain damage and increase the mortality of patients [5,6,7]. The role of glial cells in ischemic stroke is controversial [8,9]. In some studies, activation of reactive astrocytes and microglia increases brain injury after ischemic insult [10,11]. In other cases, astrocytes and microglia were also proved to be neuro-protective and critical in tissue repair after ischemic stroke [12,13,14]. However, the underlying molecular mechanisms regulating endothelial permeability and glial activation after ischemic stroke remains obscure.
The ultimate pathological change in ischemic stroke is neuronal death, which is initiated in neurons themselves and aggravated by glial cell activation and immuno–inflammatory responses [15]. Neurons are the most ischemia-susceptible cells in human body—irreversible cellular injury and cell death happen within minutes following complete oxygen deletion [15]. For decades, glutamate excitotoxicity mediated by N-methyl-D-aspartate (NMDA) receptor has been thought to be the culprit of neuron death during ischemic stroke [16,17]. However, the failure of targeting NMDA receptors in the treatment of ischemic stroke is shifting researchers’ attention to exploring other molecular mechanisms contributing to neuron death [16], among which TRPM2 emerges as a pivotal molecule in aggregating neuron death (Figure 1).

1.2. TRPM2

Transient receptor potential melastatin 2 (TRPM2) is a Ca2+-permeable and non-selective ion channel [18,19,20] belonging to the TRP channel family [21,22,23]. This 172 kDa membrane protein is located in an ~90 kb area on chromosome 21q22.3 in humans [24]. Compared with other members of the TRPM family, TRPM2 shares the characteristic N-terminal homology regions (MHR1–MHR4) and C-terminal coiled-coil domains (CTD), but is featured by its unique C-terminal NUDT9-H domain [25]. The NUDT9-H domain is homologous to NUDT9, a highly conserved adenosine diphosphate ribose (ADPR) pyrophosphatase [25,26]. Therefore, TRPM2 can be referred as a chanzyme. The enzyme function of NUDT9-H varies depending on the species [27,28]. Although invertebrate NUDT9-H is an active ADPRase, vertebrate NUDT9-H does not have the ADPRase activity. However, the NUDT9-H domain is found to be critical for the surface expression of TRPM2 and the gating of TRPM2 by ADPR [29,30]. Besides the cell membrane, TRPM2 is also suggested to be localized in the membrane of lysosomes [31,32].
Upon activation, TRPM2 currents display a characteristic linear I–V relationship, with a single channel conductance about 75–78 pS [33]. TRPM2 is gated by ADPR and Ca2+, and can be inhibited by N-(p-amylcinnamoyl) anthranilic acid (ACA) [34] and 2-aminoethoxydiphenyl borate (2-APB) [34]. Recent atomic structural analysis has revealed a Ca2+ binding site [35,36,37] which plays a major role in channel gating for several TRPM channels including TRPM2, TRPM4, and TRPM8 [38]. Activation of TRPM2 also requires ADPR binding to the channels in the presence of Ca2+. It was well known that ADPR binds to the C-terminal NUDT9-H domain of TRPM2 [39]. Interestingly, recent TRPM2 structures reveal that in zebrafish (Danio rerio) TRPM2 (drTRPM2), besides the NUDT9-H domain, there is also another ADPR binding site at the N terminal MHR1/2 domain [40]. However, this newly discovered binding site in drTRPM2 is not involved in the ADPR gating in human TRPM2 [40], although the antagonist 8-Br-cADPR binds to the MHR1/2 domain. Like many other TRP channels, TRPM2 is a temperature sensor [41]. TRPM2 is activated by heat, and the temperature threshold of TRPM2 is 47.2 ± 0.2 °C, but this threshold can be significantly reduced to a physiologically reachable level (36.3 ± 0.6 °C) by oxidative stress [42,43]. Moreover, oxidative stress also indirectly activates TRPM2 by increasing the production of ADPR and Ca2+ [44]. Therefore, TRPM2 is regarded as a cellular sensor for oxidative stress [45]. The local temperature in the affected tissue during inflammation usually increases [46], and oxidative-stress-mediated Ca2+ signaling is critical for the elicitation of inflammatory responses in immune cells [47]. The combined sensing of heat and oxidative stress confers TRPM2 with a crucial function in regulating inflammatory responses (Figure 2 and Figure 3).
TRPM2 is ubiquitously expressed in almost all tissues and cell types [48], and TRPM2-mediated Ca2+ signaling is involved in various important cellular functions, including cytokine/hormone secretion [31,49], cytoskeletal rearrangement [32], cell migration [50], regulation of reactive oxygen species (ROS) production [51], autophagy [52], inflammasome activation [49], and cell death [53]. Therefore, TRPM2 is closely related to many human diseases, such as myocardial infarction [54], ischemic stroke [55,56,57], Alzheimer’s disease [58,59], cardiomyopathy [60], atrial fibrillation [61], hypertension [62], atherosclerosis [63], inflammatory lung injury [51,64], diabetes [65], ischemic kidney disease [66], and many cancers [67]. Besides ischemic stroke itself, almost all the above diseases are the upstream etiological factors for ischemic stroke [68], highlighting the critical and comprehensive role of TRPM2 in the development and progression of this devastating disease.

2. Trpm2 in Diseases Increasing Risk for Ischemic Stroke

The TOAST classification denotes the causes of ischemic stroke into five subtypes: large-artery atherosclerosis, cardio-embolism, small-vessel occlusion, other determined etiology, and undetermined etiology, which is widely accepted by clinicians [69]. Among all the etiological factors of ischemic stroke, the most common ones are atrial fibrillation, hypertension, atherosclerosis, diabetes, and thrombosis [1,2]. Well control of these diseases is important for the prevention of ischemic stroke. However, the molecular mechanisms involved in the development of these diseases are still not completely understood. Here, we give a concise review of the role of TRPM2 in the development and/or progression of these etiological factors of ischemic stroke (Figure 1).

2.1. Atrial Fibrillation

Atrial fibrillation (AF) is a common disease in the elderly, and atrium-derived thrombus caused by AF is one of the most common causes of ischemic stroke [70]. Atrial remodeling is the cellular mechanism promoting the development and maintenance of AF [71], in which atrial fibroblasts play a predominant role by increasing the production of extracellular matrix proteins, thereby causing atrial fibrosis [72].
The principal risk factor for AF is aging [73]. Aging is usually closely related to chronic systemic inflammation, which is referred as inflammaging [74]. Chronic inflammatory response is an critical driving force in the development of atrial remodeling by enhancing the fibrotic activity of atrial fibroblasts [75]. TRPM2 is an important regulator of inflammation. Previously, TRPM2 was found to be associated with age-associated inflammatory responses in the brain, and deletion of TRPM2 protected mice against age-associated cognitive function caused by inflammation [76]. Moreover, loss of glutathione, a physiological antioxidant, during neuron senescence facilitates TRPM2 activation [77]. Aging-related chronic inflammation is also critical in the development of many cardiovascular diseases [74], and previously we found that the expression of TRPM2 was significantly increased in atrial fibroblasts isolated from patients with AF compared with that from non-AF patients [78]. Therefore, TRPM2 might also play an important role in the aging-related chronic inflammation in the atria, thereby promoting the development of atrial remodeling and AF.
Oxidative stress promotes AF development by impairing the contractility of atrial myocytes [79,80] and accelerating atrial fibrosis [81]. In a recent paper we showed that TRPM2 activation markedly and rapidly promoted the production of ROS in mitochondria of macrophages [63]. TRPM2 itself is a cellular sensor for oxidative stress [45]. Therefore, the increased production of ROS will in turn further promote the activation of TRPM2, thereby forming a feed-forward vicious cycle [63]. Moreover, Ca2+ signaling is critical for the activation of fibroblasts during atrial fibrosis [82,83], and TRPM2-mediated Ca2+ influx under oxidative stress has been shown to be required for many cellular functions [31,49,50,52]. Considering the high expression of TRPM2 in atrial tissue after AF [78], there is a high possibility that TRPM2 also contributes to the progression of AF by magnifying the oxidative stress response and Ca2+ signaling in atrial myocytes and fibroblasts.
Intracellular Ca2+ is not only critical for regulating the mechanical and electrical activity of healthy atrial muscle, but it also plays an important role in the triggering of AF [84]. AF can be triggered by afterdepolarizations [84]. There are two types of afterdepolarizations, early afterdepolarizations (EAD) and delayed afterdepolarizations (DAD). Both EAD and DAD can cause abnormal electrical activities in atrial muscle, and Ca2+ overload during EAD in atrial muscle was shown to trigger AF [85]. The molecular mechanisms of EAD remain mysterious. Some studies suggest that EAD might result from the Ca2+ influx from L-type Ca2+ channels or spontaneous Ca2+ release from the endoplasmic reticulum (ER) [86]. Considering the important role and active involvement of aging and oxidative stress in the development of AF and TRPM2 in these two conditions, TRPM2-mediated Ca2+ influx or Ca2+ release from lysosome might also contribute to the Ca2+ overload during EAD and the triggering of AF.

2.2. TRPM2 in Hypertension

Hypertension, perhaps the most prevalent disease in humans, is an independent risk factor for ischemic stroke [87]. Traditionally, hypertension was thought to cause stroke mainly in the elderly and young males [88,89]. Recently, a small increase in blood pressure even as mild as 10 mm Hg was shown to be associated with a 38% increased risk of stroke in females [90]. The characteristic pathological change of hypertension is arteriosclerosis, which increases the peripheral resistance and blood pressure [91]. The progressive hardening of arteriole walls is called arterial remodeling [92], in which endothelial dysfunction and smooth-muscle proliferation play a central role [91].
Oxidative stress has long been known to promote the development and progression of hypertension by inducing endothelial dysfunction [93], smooth muscle hypertrophy [94], and vascular remodeling [95], whereas antioxidants were shown to have a protective effect against hypertension by preventing the vascular dysfunction [96]. However, the underlying molecular mechanisms remain unknown [97]. In a recent study, the Ca2+ dysregulation and hyperactivity of vascular smooth muscle cells during hypertension was found to be dependent on TRPM2-mediated Ca2+ influx, which is activated by angiotensin-II-mediated increase of ROS production [62]. Similarly, endothelial dysfunction mediated by ROS-activated TRPM2 was also found to accelerate the development of Alzheimer’s disease [58] and aggregate inflammatory lung injury [64]. As a cellular sensor for oxidative stress, TRPM2 might mediate the detrimental effects of ROS on endothelial cells and smooth muscle cells in the development of hypertension.
Intracellular Ca2+ is a critical regulator of endothelial function [98] and smooth muscle contractility [99]. Ca2+ signaling regulates the expression of endothelial nitric oxide synthase [98,100], whereas blockade of Ca2+ channels enhanced the production of nitric oxide, a potent vasodilator, in endothelial cells [101,102]. Moreover, increase of intracellular Ca2+ in smooth muscle cells enhanced muscle tone and peripheral resistance [103]. In platelets isolated from hypertensive patients, cellular Ca2+ concentration was much higher than that from patients with normal blood pressure [104]. There are two sources of cytosolic free Ca2+—one is the Ca2+ influx from the extracellular environment, and another is the Ca2+ release from intracellular organelles, such as the endoplasmic reticulum and lysosomes. The multiple roles of TRPM2-mediated Ca2+ influx have been well documented in many cell types including endothelial cells [64]. However, there are a limited number of studies reporting the function of TRPM2 in lysosomes. The lysosome is the most important organelle for autophagy and TRPM2 has long been shown to be associated with autophagy [105]. Lysosomal TRPM2-mediated Ca2+ release was shown to be responsible for the pancreatic β-cell death under oxidative stress [31]. Recently TRPM2-mediated Ca2+ release from lysosomes was found to promote autophagic degradation in vascular smooth muscle cells, thereby causing cell death [106], and knockout of TRPM2 attenuated hypertension in spontaneously hypertensive rats by reconstituting autophagy in endothelial cells and vascular smooth muscle cells [107]. In summary, TRPM2-mediated Ca2+ signaling aggregates the dysfunction of endothelial cells and vascular smooth muscle cells in the development and progression of hypertension.

2.3. TRPM2 in Atherosclerosis

Atherosclerosis is a dangerous risk factor for ischemic stroke. Atherosclerotic plaque on the aortic arch [108] and carotid artery [109] significantly increases the risk of ischemic stroke. Moreover, break of atherosclerotic plaque in cerebral arteries directly leads to the formation of in situ thrombus [110], which is usually more difficult to evaluate and predict due to the small plaque size and deep position in the skull [110]. The central pathological feature of atherosclerosis is foam cell formation [111]. Uptake of too much cholesterol transforms infiltrated macrophages into highly inflammatory foam cells, which are the culprit in the development and progression of atherosclerotic plaque [112]. Foam-cell formation includes two critical processes—one is macrophage infiltration, the another is phagocytosis of cholesterol included in oxidized low-density-lipoprotein (oxLDL) [112].
The first step of macrophage infiltration during atherosclerosis is macrophage chemotaxis toward the lesion site, which is caused by chemokines. Previously, TRPM2 was found to be critical for the chemotaxis of neutrophils by formyl-methionyl-leucyl-phenylalanine (fMLP) [113]. fMLP is also a well-established chemokine for macrophages, suggesting the potential role of TRPM2 in regulating macrophage chemotaxis. Indeed, in a recent study, we found that the in vitro macrophage migration induced by macrophage chemotaxis protein 1 (MCP1) was inhibited by deleting TRPM2 or inhibiting the activation of TRPM2 [63]. MCP1 secreted by endothelial cells in response to subendothelial deposition of oxLDL is one of the initial driving forces of macrophage infiltration during atherosclerosis [112]. Similar to our findings, hydrogen peroxide (H2O2) was shown to attract neutrophil migration both in vivo and in intro, which was also abolished by TRPM2 knockout or TRPM2 inhibition [114]. H2O2 is an important molecular signal generated during inflammation [115]. H2O2 gradient produced by wounded tissue is required for the rapid leukocyte recruitment after injury [116]. Like the chemotactic effect on neutrophils, H2O2 might also be an important chemokine for macrophages, and the recruitment of macrophages mediated by H2O2 might depend on TRPM2 activation.
The second step of macrophage infiltration into the vessel wall is attachment and penetration of the endothelium, which is called extravasation [112]. Leukocyte extravasation is mediated by several surface-adhesion molecules. TRPM2 was found to be required for the transendothelial migration of neutrophils induced by endotoxin, in which TRPM2-mediated Ca2+ influx promotes the phosphorylation of VE-cadherin and degradation of tight junctions between endothelial cells [50]. Similarly, specific deletion of TRPM2 in immune cells markedly decreased immune-cell invasion into the brain after ischemic stroke [55]. Moreover, in mice fed with high-fat diet, Trpm2 deletion caused reduced macrophage infiltration and attenuated inflammation in adipose tissue compared with wild type mice [117]. In our study, we also found that TRPM2-mediated Ca2+ influx is also needed for the in vitro transendothelial migration of macrophages induced by MCP1, and Trpm2 deletion reduces the macrophage burden in atherosclerotic plaques in vivo [63]. These studies highlight the crucial role of TRPM2 in promoting macrophage infiltration during atherogenesis.
After infiltrating into the vessel wall, macrophages engulf oxLDL and become highly proinflammatory foam cells [112]. Foam cells promote the development and progression of atherosclerosis by secreting pro-inflammatory cytokines, chemokines, and tissue-degrading enzymes, which cause profound inflammatory responses and lead to lesion expansion [111]. Activation of the NFκB signaling pathway is required for the activation of macrophages, and TRPM2-mediated Ca2+ signaling has been shown to be indispensable for NFκB signaling activation in macrophages during inflammation, suggesting the potential tole of TRPM2 in transforming macrophages into foam cells [118]. CD36 is the most important receptor for oxLDL uptake in macrophages, and activation of the CD36 downstream signaling pathways is required for macrophage activation and subsequent foam-cell formation [119,120,121,122,123,124]. Previously TRPM2 was shown to mediate the activation of macrophages during infection [125,126] and inflammation [49,127] or when temperature increases [42]. We found that TRPM2 is required for CD36 activation and oxLDL uptake in macrophages, and activation of CD36 by oxLDL further promotes the activation of TRPM2, thereby forming a feed-forward viscous cycle [63]. NLRP3 inflammasome activation by engulfed cholesterol is required for macrophage activation during atherosclerosis [128]. We also found that NLRP3 inflammasome activation by oxLDL is dependent on TRPM2-mediated Ca2+ signaling [63]. All these studies suggest TRPM2 is likely to play a crucial role in the transformation of infiltrated native macrophages into pro-inflammatory foam cells.

2.4. TRPM2 in Diabetes

Hyperglycemia causes a series of pathological changes in the vessel walls, including endothelial dysfunction, basal membrane thickening, interstitial fibrosis, and vessel stiffness, which markedly increase the risk of developing ischemic stroke [129]. Moreover, the mortality of ischemic stroke is higher and clinical outcomes are poorer in patients with diabetes [130]. Well-controlled hyperglycemia significantly decreases the risk of ischemic stroke, decreases mortality, and improves clinical outcomes [131]. The prominent pathological feature of diabetes is insufficient insulin secretion (type 1 diabetes mellitus, T1DM) or unresponsiveness of peripheral tissues to insulin (type 2 diabetes mellitus, T2DM). Many studies have demonstrated that TRPM2 is implicated in the development of both T1DM and T2DM.
T1DM is featured by gradual loss of pancreatic β cells by chronic inflammation in pancreatic islets, thereby resulting in lack of insulin secretion [132]. Ca2+ influx is required for insulin secretion [133], which is further amplified by intracellular Ca2+ release [134]. TRPM2 has a high expression level in pancreatic β cells and activation of TRPM2-mediated Ca2+ influx by ADPR is required for insulin secretion, which can be further enhanced when temperature increases [135]. Knockout of TRPM2 leads to decreased serum insulin levels, increased glucose levels in plasma, and higher insulin sensitivity of peripheral tissues [65,117]. However, excessive Ca2+ influx also leads to cell death under pathological conditions. Both TRPM2-mediated Ca2+ influx and lysosomal Ca2+ release were found to be involved in β-cell death caused by H2O2 [31,136]. Autoimmune inflammatory responses play a central role in the destruction of β cell during T1DM [132]. During chronic inflammation, infiltrated self-reactive T cells and macrophages can produce substantial amount of H2O2, and oxidative stress is also an important mechanism for β-cell dysfunction during T1DM [132,137,138,139]. Moreover, infiltrated immune cells also secrete cytokines such as tumor necrosis factor-α (TNF-α) [140], and TRPM2-mediated Ca2+ influx has been shown to mediated the cytotoxic effect of TNF-α [20,31]. Therefore, TRPM2 is a key molecule in promoting β-cell death in the development of T1DM.
The distinguishing feature of T2DM is insulin resistance [141]. Peripheral tissues, especially adipose tissues, lose their ability to respond to insulin, which results in decreased uptake of glucose in the blood [142]. Macrophage-mediated chronic inflammation is a major mechanism for the reduced insulin resistance in adipose tissue [143,144,145,146,147], and reduced macrophage load in adipose tissue is associated with better blood-glucose control in human patients [148]. In mice fed with a high-fat diet, Trpm2 deletion alleviated inflammation and macrophage infiltration and improved insulin resistance in adipose tissue compared with wild-type mice [117]. As mentioned earlier, TRPM2-mediated Ca2+ influx is required for macrophage infiltration and activation during inflammation [42,49,127] or under different pathological conditions [50,63,125,126]. Therefore, TRPM2 activation might also be required for macrophage infiltration into adipose tissue and the subsequent activation. Targeting TRPM2 can be a promising strategy to inhibit adipose tissue inflammation and improve insulin sensitivity during T2DM.

2.5. TRPM2 in Thrombosis

Extracranial thrombus (from the left atrium) and intracranial in situ thrombosis are important causes of ischemic stroke [1,2]. Platelet aggregation plays a central role in thrombus formation [147,149,150], during which increase of intracellular Ca2+ is required [151,152]. Oxidative stress promotes the recruitment and activation of platelets, but the underlying molecular mechanisms remain unclear [153]. TRPM2 is expressed in megakaryocytes and platelets, but the function of TRPM2 in megakaryocytes and platelets remains mysterious [154,155]. The important roles of TRPM2-mediated Ca2+ signaling in many cellular functions have been well documented, and TRPM2 might also be involved in platelet activation during thrombosis.
Thrombospondin-1 (TSP1) is a potent platelet activator by inhibiting intracellular nitric oxide signaling, and TSP1-CD36 ligation increases thrombus stability [153,156]. We found that TSP1-mediated macrophage activation depends on TRPM2 activation under oxidative stress [63]. Recently, we also found that TRPM2-mediated Ca2+ influx is required for TSP1 mediated endothelial dysfunction during ischemic stroke (unpublished). Platelets themselves have an abundant storage of TSP1 in intracellular α-granules, and during platelet activation TSP1 is rapidly released [156]. Moreover, after ischemic stroke, the synthesis and release of TSP1 in the brain was dramatically increased even within hours [157]. Therefore, it is very likely that TRPM2 can promote thrombus formation, and the platelet aggregation induced by TSP1 also needs the involvement of TRPM2.

3. Mechanisms by Which TRPM2 Increases Brain Injury during Ischemic Stroke

Compared with all other tissues, the brain has the highest expression of TRPM2 under physiological conditions [48]. Moreover, the expression of TRPM2 after ischemic stroke was significantly increased [57], highlighting the critical role of TPRM2 in promoting brain damage. The central pathological feature of ischemic stroke is neuronal-cell death, and all therapies for ischemic stroke are aimed at preventing or decreasing the loss of neurons [2]. The failure of directly targeting neurons in treating ischemic stroke shifts researcher’s attention to a relative new concept, the neurovascular unit (NVU) [16,158,159]. The NVU is composed of endothelial cells, pericytes, vascular smooth muscle cells, surrounding glial cells (mostly astrocytes), and related neurons [160]. After ischemic stroke, all the components of the NVU are subjected to hypoxic condition and dysfunction of non-neuron cells significantly aggregates neuron death [2]. Studies have shown that TRPM2 activation in each component of the NVU can aggregate brain damage during ischemic stroke (Figure 4, as summarized in Table 1).

3.1. TRPM2 and Cerebral Endothelial Hyperpermeability

Anatomically, cerebral endothelial cells are the core components of the NVU [161]. To fulfil the function of underlying the outer surface of the blood–brain barrier (BBB), cerebral endothelial cells (CECs) have several unique properties compared with peripheral endothelial cells, including densely distributed tight junctions and significantly higher mitochondrial content [162]. After ischemic stroke, degradation of tight junctions between CECs causes plasma extravasation and immune-cell infiltration, and the resulting cerebral edema and inflammatory damage significantly increase the mortality of patients [5,6,7,163]. However, the molecular mechanisms causing endothelial dysfunction during ischemic stroke remain unclear [164].
The roles of TRPM2 in regulating the permeability of peripheral endothelial cells have been extensively studied. In one study, transendothelial migration of neutrophils across lung-derived endothelial cell was shown to be dependent on the activation of TRPM2 in endothelial cells [50]. TRPM2-mediated Ca2+ influx dissembled adherens junctions and opened paracellular pathways, thereby allowing for neutrophil extravasation [50]. Another study showed that H2O2 treatment results in hyperpermeability in lung-derived endothelial cells by increasing ADPR production thereby activating TRPM2 and causing Ca2+ overload, and inhibition of TRPM2 produced a potent protective effect against H2O2-induced endothelial hyperpermeability [64]. Since the mitochondrion is the major organelle producing ADPR under oxidative stress, the higher mitochondria content in CECs could result in more robust ADPR production compared with lung-derived endothelial cells [162]. Thus, TRPM2 is likely to be activated under this condition. Indeed, we also found that oxygen–glucose deprivation (OGD), an in vitro mimic of ischemic stroke, induced substantial ROS production and hyperpermeability in CECs, whereas Trpm2 deletion or TRPM2 inhibition protected CECs from OGD-induced Ca2+ overload, tight-junction degradation, and endothelial hyperpermeability (under revision). Moreover, we also found that endothelial-specific deletion of TRPM2 significantly alleviated plasma extravasation and immune-cell infiltration after middle cerebral artery occlusion (MCAO), in an in vivo mimic of human ischemic stroke in mice (under revision).

3.2. TRPM2 and Immune-Cell Invasion

Normally, the densely distributed tight junctions between CECs prevent the migration of circulating immune cells into the brain [6]. During ischemic stroke, however, massive immune-cell infiltration occurs due to the degradation of tight junctions and secretion of chemokines following tissue injury, which causes profound inflammatory responses and aggregated brain damage [5,6]. As discussed earlier, TRPM2 is essential for immune cell migration in various pathological situations [42,49,50,63,117,125,126,127]. Indeed, replacement of bone marrow in wild-type mice with bone marrow from Trpm2 knockout mice significantly decreased infiltration of neutrophils and macrophages into the brain and reduced infarction size after MCAO, indicating TRPM2 activation in immune cells is required for their penetration across the BBB after ischemic stroke [55]. After entering into the brain, immune cells are activated and secrete cytotoxic cytokines, such as TNF-α and IL-1β, which further promote neuron death [6]. Considering TRPM2-mediated Ca2+ signaling is required for immune cell activation and cytokine production [42,49,50,63,117,125,126,127], targeting TRPM2 in immune cells might provide extra benefit in mitigating tissue damage following immune cell extravasation.

3.3. TRPM2 in Glial Cells

Microglia and astrocytes are the two most important glial cells involved in tissue inflammation and repair after ischemic stroke [9]. Microglia are the brain-resident close relatives of peripheral macrophages. Under physiological conditions, microglia are responsible for clearing cell debris and shaping synapses during brain development [165]. Under pathological conditions such as ischemic stroke, activated microglia phagocytize dead cells and cause inflammatory responses [166]. When activated, ROS production in microglia is increased [167]. Increased ROS is not only needed for activation of pro-inflammatory genes, but microglia also secrete excessive ROS into surrounding tissues, which causes comprehensive neuron damage [167]. As mentioned earlier, ROS activates TRPM2 by increasing the production of ADPR, and TRPM2-mediated Ca2+ signaling is required for the activation of macrophages, the peripheral cousins of microglia. In addition, the expression of TRPM2 in microglia was increased after H2O2 treatment [168]. Therefore, TRPM2 is likely to contribute to microglia activation after ischemic stroke. Indeed, H2O2 and ADPR activate TRPM2 in cultured microglia [169], and TRPM2 is required for TNF-α production in microglia [170,171]. Deletion of TRPM2 also significantly alleviated brain damage in a chronic cerebral hypoperfusion mouse model by inhibiting microglial activation [172]. Moreover, CD36 activation in microglia promoted ROS production and aggregated brain injury after ischemic stroke [173]. Recently, we also found that activation of CD36 in macrophages depends on TRPM2-mediated Ca2+ influx [63]. Thus, TRPM2 may also be required for CD36 activation in microglia.
Astrocytes are the major housekeeping cells in the brain. Their functions include recycling released neuro-transmitters in synapses, regulating water homeostasis in the brain, secreting neurotrophic factors supporting neuron development, and forming the backbone of the BBB [174]. After ischemic stroke, resting astrocytes are activated by tissue damage and become reactive astrocytes [175]. Proliferation of reactive astrocytes leads to the formation of a glial scar, which is called reactive astrogliosis [138]. Moreover, reactive astrocytes were found to enhance immune response and aggregate brain damage [176]. However, molecular mechanisms regulating astrocyte activation are still not clear. Similarly to activated microglia, ROS production is also increased in reactive astrocytes [177,178,179], and inhibiting TRPM2 suppressed astrocyte activation and reduced the secretion of pro-inflammatory cytokines by astrocytes in response to depletion of glutathione, a ROS neutralizer [171]. The results of above studies strongly suggest the critical role of TRPM2-mediated Ca2+ signaling in promoting the detrimental microglial/astrocyte activation and microglial/astrocyte-mediated tissue damage after ischemic stroke.

3.4. Neuronal TRPM2 and Neuron Death

Neurons are the most important cells for an intelligent human being, but they are also the most fragile cells when exposed to hypoxia. Neuron death is the hallmark of ischemic stroke. Molecular mechanisms underlying neuron death during ischemic stroke have been extensively studied. Among these mechanisms, two that are important are ROS production and Ca2+ overload [1]. Compared with the robust activation of immune cells by ROS and Ca2+ influx, even a moderate increase of intracellular ROS and Ca2+ can kill neurons [180]. The role of Ca2+ influx mediated by oxidative-stress-activated TRPM2 in cell death has been well documented [20,53]. H2O2 induced Ca2+ overload in cultured neurons and caused rapid neuron death, whereas knockdown [181,182], inhibition [182,183], and knockout [56] of TRPM2 produced a potent protective effect against Ca2+ overload and neuron death induced by H2O2 treatment or OGD. IN addition, infarction size was reduced in Trpm2 knockout mice compared with wild-type mice [184,185,186]. Moreover, TRPM2 was found to be related to delayed neuron death after ischemic stroke, and TRPM2 deletion prevented post-stroke cognitive dysfunction in mice [56]. Recently, we also found that selective deletion of TRPM2 in neurons protects mice from ischemic brain damage by inhibiting ROS production and Ca2+ overload [57].
Excitotoxicity caused by NMDA-receptor (NMDAR) activation plays a central role in inducing neuron death during ischemic stroke [1]. Like TRPM2, NMDAR is also a Ca2+-permeable channel and Ca2+ overload mediated by NMDAR has long been thought to be responsible for neuron death during ischemic stroke [187]. However, due to the complex nature of NMDAR, the contribution of NMDAR to neuron death depends on their composition and location. NMDAR are tetramers consisting of different subunits (GluN1, GluN2a, GluN2b, GluN2c, GluN2d, GluN3a, and GluN3b). There must be two GluN1 subunits in a NMDAR tetramer, while the other two subunits can be any combination of GluN2a-d and/or GluN3a-b [188]. The most abundant and physiologically important subunits in adults are GluN2a and GluN2b [188]. In ischemic stroke, GluN1/GluN2a promotes neuron survival, whereas GluN1/GluN2b causes neuron death [189,190]. In addition, the location of NMDAR is critical in determining the fate of neuron during ischemia. Synaptic NMDAR is crucial in maintaining the normal function of the brain, and is found to protect neuron from ischemic injury [191]. In contrast, NMDAR at extra-synaptic sites mediates long-term depression, and is critical in mediating the glutamate excitotoxicity during ischemic stroke [189]. TRPM2 was found to increase the expression of GluN2b while decreasing the expression of GluN2a, and knockout of TRPM2 reversed this change and protected mice from ischemic stroke [192]. Moreover, in our recent study, we found that TRPM2 was able to physically and functionally associate with GluN2a and GluN2b in NMDAR at extrasynaptic sites, and specific disruption of this TRPM2-NMDAR interaction using a interfering peptide produced a protective effect against ischemic stroke similar to Trpm2 deletion [57]. The above studies indicate that TRPM2 activation in neurons during ischemic stroke promotes neuron death.

4. Other TRPM Channels in Ischemic Stroke

In addition to TRPM2, TRPM4 and TRPM7 are also highly expressed in the brain, and were found to aggravate the brain injury following ischemic stroke. Similar to TRPM2, TRPM4 is activated by intracellular Ca2+ [195]. TRPM4 can also be activated by depleting intracellular ATP, suggesting its activity may play an important role in cellular damage when cells undergo energy imbalance during ischemia [195,196]. After MCAO, expression of TRPM4 in the endothelial cells was markedly upregulated, and knockdown of TRPM4 protected mice against ischemic brain injury, suggesting TRPM4 activation may promote BBB degradation after ischemic stroke [197]. Moreover, in a recent study, TRPM4 was found to interact with the NMDA receptors (NMDAR), which enhanced the excitotoxicity and increased neuronal death [198]. Disruption of TRPM4–NMDAR interaction significantly attenuated the MCAO induced brain injury [198]. Compared to TRPM4, the detrimental effects of TRPM7 in ischemic stroke have been extensively studied since 2003 [199]. The unique feature of TRPM7 among all TRPM channels is its high permeability to zinc [199], and zinc toxicity has been well documented as an important cause of neuronal death after ischemic stroke [200]. Knockdown of TRPM7 suppressed the in vivo neuronal death and prevented the post-stroke cognitive dysfunction in mice [201].

5. Perspective

Currently, the only two effective therapies for ischemic stroke are intravenous thrombolysis and endovascular thrombectomy. However, both of them have many limitations, including a tight therapeutic window, a low success-rate, and lethal complications [1,2]. For over 30 years, NMDAR-mediated glutamate excitotoxicity was thought to be the major mechanism responsible for neuronal death, but NMDAR antagonists all failed to show any protective effect in clinical trials. In recent years, TRPM2 has been shown to be involved in the pathophysiology of ischemic stroke in all the aspects, including accelerating the development of upstream risk factors outside the brain and aggregating the post-stroke pathological damage inside the brain. Therefore, TRPM2 could be a promising target for screening more effective therapies for preventing ischemic strokes and for treating patients with ischemic stroke.
There are still two major issues impeding the translation of these exciting findings from basic studies into clinical applications. Firstly, TRPM2 has long been thought to lack of specific inhibitors. However, in recent years, the complete revelation of the high-resolution atomic structure of TRPM2 by cryo-EM has paved the way for developing more specific and effective TRPM2 inhibitors. Secondly, the wide expression of TRPM2 also increases the risk of side effects caused by pharmacological inhibition of TRPM2 during treatment. Luckily, the rapid advancement of nanomedicines has led to the development of more targeted, controlled, and efficient drug delivery systems [202]. Combination of TRPM2-specific inhibitors with highly tissue/cellular specific nano-delivery vehicles could make possible the application TRPM2-inhibition therapies in treating ischemic stroke and its “upstream” diseases.

Author Contributions

P.Z. wrote the manuscript, Q.L. prepared the graphic figures, and J.F. and L.Y. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Institute of Health [R01-HL143750] and American Heart Association [19TPA34890022]. And The APC was funded by National Institute of Health.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

We choose to exclude this statement because this review did not report any data.

Acknowledgments

This work was partially supported by the National Institute of Health (R01-HL143750) and American Heart Association (19TPA34890022) to LY.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Johnson, C.O.; Nguyen, M.; Roth, G.A.; Nichols, E.; Alam, T.; Abate, D.; Abd-Allah, F.; Abdelalim, A.; Abraha, H.N.; Abu-Rmeileh, N.M.; et al. Global, regional, and national burden of stroke, 1990–2016: A systematic analysis for the Global Burden of Disease Study 2016. Lancet Neurol. 2019, 18, 439–458. [Google Scholar] [CrossRef] [Green Version]
  2. Campbell, B.C.V.; De Silva, D.A.; MacLeod, M.R.; Coutts, S.B.; Schwamm, L.H.; Davis, S.M.; Donnan, G.A. Ischaemic stroke. Nat. Rev. Dis. Prim. 2019, 5, 1–22. [Google Scholar] [CrossRef] [PubMed]
  3. Campbell, B.C.V.; Khatri, P. Stroke. Lancet 2020, 396, 129–142. [Google Scholar] [CrossRef]
  4. Allen, C.L.; Bayraktutan, U. Risk Factors for Ischaemic Stroke. Int. J. Stroke 2008, 3, 105–116. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, C.; Hawkins, K.E.; Doré, S.; Candelario-Jalil, E. Neuroinflammatory mechanisms of blood-brain barrier damage in ischemic stroke. Am. J. Physiol. Cell Physiol. 2019, 316, C135–C153. [Google Scholar] [CrossRef]
  6. Planas, A.M. Role of Immune Cells Migrating to the Ischemic Brain. Stroke 2018, 49, 2261–2267. [Google Scholar] [CrossRef]
  7. Bardutzky, J.; Schwab, S. Antiedema Therapy in Ischemic Stroke. Stroke 2007, 38, 3084–3094. [Google Scholar] [CrossRef] [Green Version]
  8. Liu, Z.; Chopp, M. Astrocytes, therapeutic targets for neuroprotection and neurorestoration in ischemic stroke. Prog. Neurobiol. 2015, 144, 103–120. [Google Scholar] [CrossRef] [Green Version]
  9. Jayaraj, R.L.; Azimullah, S.; Beiram, R.; Jalal, F.Y.; Rosenberg, G.A. Neuroinflammation: Friend and foe for ischemic stroke. J. Neuroinflamm. 2019, 16, 142. [Google Scholar] [CrossRef] [Green Version]
  10. Lin, J.H.-C.; Weigel, H.; Cotrina, M.L.; Liu, S.; Bueno, E.; Hansen, A.J.; Hansen, T.W.; Goldman, S.A.; Nedergaard, M. Gap-junction-mediated propagation and amplification of cell injury. Nat. Neurosci. 1998, 1, 494–500. [Google Scholar] [CrossRef]
  11. Hu, X.; Li, P.; Guo, Y.; Wang, H.; Leak, R.K.; Chen, S.; Gao, Y.; Chen, J. Microglia/Macrophage Polarization Dynamics Reveal Novel Mechanism of Injury Expansion After Focal Cerebral Ischemia. Stroke 2012, 43, 3063–3070. [Google Scholar] [CrossRef] [Green Version]
  12. Morizawa, Y.M.; Hirayama, Y.; Ohno, N.; Shibata, S.; Shigetomi, E.; Sui, Y.; Nabekura, J.; Sato, K.; Okajima, F.; Takebayashi, H.; et al. Reactive astrocytes function as phagocytes after brain ischemia via ABCA1-mediated pathway. Nat. Commun. 2017, 8, 28. [Google Scholar] [CrossRef] [PubMed]
  13. Neumann, J.; Gunzer, M.; Gutzeit, H.O.; Ullrich, O.; Reymann, K.G.; Dinkel, K. Microglia provide neuroprotection after ischemia. FASEB J. 2006, 20, 714–716. [Google Scholar] [CrossRef] [PubMed]
  14. Becerra-Calixto, A.; Cardona-Gómez, G.P. The Role of Astrocytes in Neuroprotection after Brain Stroke: Potential in Cell Therapy. Front. Mol. Neurosci. 2017, 10, 88. [Google Scholar] [CrossRef] [Green Version]
  15. Lipton, P. Ischemic Cell Death in Brain Neurons. Physiol. Rev. 1999, 79, 1431–1568. [Google Scholar] [CrossRef]
  16. Cheng, Y.D.; Al-Khoury, L.; Zivin, J.A. Neuroprotection for ischemic stroke: Two decades of success and failure. NeuroRX 2004, 1, 36–45. [Google Scholar] [CrossRef] [PubMed]
  17. Lai, T.W.; Zhang, S.; Wang, Y.T. Excitotoxicity and stroke: Identifying novel targets for neuroprotection. Prog. Neurobiol. 2014, 115, 157–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Sano, Y.; Inamura, K.; Miyake, A.; Mochizuki, S.; Yokoi, H.; Matsushime, H.; Furuichi, K. Immunocyte Ca2+ Influx System Mediated by LTRPC2. Science 2001, 293, 1327–1330. [Google Scholar] [CrossRef]
  19. Perraud, A.-L.; Fleig, A.; Dunn, C.A.; Bagley, L.A.; Launay, P.; Schmitz, C.; Stokes, A.; Zhu, Q.; Bessman, M.J.; Penner, R.; et al. ADP-ribose gating of the calcium-permeable LTRPC2 channel revealed by Nudix motif homology. Nature 2001, 411, 595–599. [Google Scholar] [CrossRef]
  20. Hara, Y.; Wakamori, M.; Ishii, M.; Maeno, E.; Nishida, M.; Yoshida, T.; Yamada, H.; Shimizu, S.; Mori, E.; Kudoh, J.; et al. LTRPC2 Ca2+-Permeable Channel Activated by Changes in Redox Status Confers Susceptibility to Cell Death. Mol. Cell 2002, 9, 163–173. [Google Scholar] [CrossRef]
  21. Montell, C.; Birnbaumer, L.; Flockerzi, V. The TRP Channels, a Remarkably Functional Family. Cell 2002, 108, 595–598. [Google Scholar] [CrossRef] [Green Version]
  22. Clapham, D.E. TRP channels as cellular sensors. Nature 2003, 426, 517–524. [Google Scholar] [CrossRef] [PubMed]
  23. Nilius, B.; Voets, T.; Peters, J. TRP Channels in Disease. Sci. STKE 2005, 1–9, [Epub ahead of print]. [Google Scholar] [CrossRef] [PubMed]
  24. McQuillin, A.; Bass, N.J.; Kalsi, G.; Lawrence, J.; Puri, T.; Choudhury, K.; Detera-Wadleigh, S.D.; Curtis, D.; Gurling, H.M.D. Fine mapping of a susceptibility locus for bipolar and genetically related unipolar affective disorders, to a region containing the C21ORF29 and TRPM2 genes on chromosome 21q22.3. Mol. Psychiatry 2005, 11, 134–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Huang, Y.; Winkler, P.A.; Sun, W.; Lü, W.; Du, J. Architecture of the TRPM2 channel and its activation mechanism by ADP-ribose and calcium. Nature 2018, 562, 145–149. [Google Scholar] [CrossRef] [PubMed]
  26. Perraud, A.-L.; Shen, B.; Dunn, C.A.; Rippe, K.; Smith, M.K.; Bessman, M.J.; Stoddard, B.L.; Scharenberg, A.M. NUDT9, a Member of the Nudix Hydrolase Family, Is an Evolutionarily Conserved Mitochondrial ADP-ribose Pyrophosphatase. J. Biol. Chem. 2003, 278, 1794–1801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Iordanov, I.; Tóth, B.; Szollosi, A.; Csanády, L. Enzyme activity and selectivity filter stability of ancient TRPM2 channels were simultaneously lost in early vertebrates. eLife 2019, 8, e44556. [Google Scholar] [CrossRef]
  28. Iordanov, I.; Mihályi, C.; Tóth, B.; Csanády, L. The proposed channel-enzyme transient receptor potential melastatin 2 does not possess ADP ribose hydrolase activity. eLife 2016, 5, e17600. [Google Scholar] [CrossRef]
  29. Kühn, F.J.P.; Lückhoff, A. Sites of the NUDT9-H Domain Critical for ADP-ribose Activation of the Cation Channel TRPM2. J. Biol. Chem. 2004, 279, 46431–46437. [Google Scholar] [CrossRef] [Green Version]
  30. Perraud, A.-L.; Schmitz, C.; Scharenberg, A.M. TRPM2 Ca2+ permeable cation channels: From gene to biological function. Cell Calcium 2003, 33, 519–531. [Google Scholar] [CrossRef]
  31. Lange, I.; Yamamoto, S.; Partida-Sanchez, S.; Mori, Y.; Fleig, A.; Penner, R. TRPM2 Functions as a Lysosomal Ca2+-Release Channel in beta cells. Sci. Signal. 2009, 2, ra23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Sumoza-Toledo, A.; Lange, I.; Cortado, H.; Bhagat, H.; Mori, Y.; Fleig, A.; Penner, R.; Partida-Sánchez, S. Dendritic cell maturation and chemotaxis is regulated by TRPM2-mediated lysosomal Ca2+ release. FASEB J. 2011, 25, 3529–3542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Du, J.; Xie, J.; Yue, L. Intracellular calcium activates TRPM2 and its alternative spliced isoforms. Proc. Natl. Acad. Sci. USA 2009, 106, 7239–7244. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Kraft, R.; Grimm, C.; Frenzel, H.; Harteneck, C. Inhibition of TRPM2 cation channels by N -(p -amylcinnamoyl)anthranilic acid. J. Cereb. Blood Flow Metab. 2006, 148, 264–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Autzen, H.E.; Myasnikov, A.G.; Campbell, M.G.; Asarnow, D.; Julius, D.; Cheng, Y. Structure of the human TRPM4 ion channel in a lipid nanodisc. Science 2018, 359, 228–232. [Google Scholar] [CrossRef] [Green Version]
  36. Huang, Y.; Fliegert, R.; Guse, A.H.; Lü, W.; Du, J. A structural overview of the ion channels of the TRPM family. Cell Calcium 2019, 85, 102111. [Google Scholar] [CrossRef]
  37. Yin, Y.; Le, S.C.; Hsu, A.L.; Borgnia, M.J.; Yang, H.; Lee, S.-Y. Structural basis of cooling agent and lipid sensing by the cold-activated TRPM8 channel. Science 2019, 363, eaav9334. [Google Scholar] [CrossRef]
  38. Zhang, Z.; Tóth, B.; Szollosi, A.; Chen, J.; Csanády, L. Structure of a TRPM2 channel in complex with Ca2+ explains unique gating regulation. eLife 2018, 7, e36409. [Google Scholar] [CrossRef]
  39. Csanády, L.; Törőcsik, B. Four Ca2+ Ions Activate TRPM2 Channels by Binding in Deep Crevices near the Pore but Intracellularly of the Gate. J. Gen. Physiol. 2009, 133, 189–203. [Google Scholar] [CrossRef]
  40. Wang, L.; Fu, T.-M.; Zhou, Y.; Xia, S.; Greka, A.; Wu, H. Structures and gating mechanism of human TRPM2. Science 2018, 362, 6421. [Google Scholar] [CrossRef] [Green Version]
  41. Vilar, B.; Tan, C.-H.; McNaughton, P.A. Heat detection by the TRPM2 ion channel. Nature 2020, 584, E5–E12. [Google Scholar] [CrossRef] [PubMed]
  42. Kashio, M.; Sokabe, T.; Shintaku, K.; Uematsu, T.; Fukuta, N.; Kobayashi, N.; Mori, Y.; Tominaga, M. Redox signal-mediated sensitization of transient receptor potential melastatin 2 (TRPM2) to temperature affects macrophage functions. Proc. Natl. Acad. Sci. USA 2012, 109, 6745–6750. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kashio, M.; Tominaga, M. The TRPM2 channel: A thermo-sensitive metabolic sensor. Channels 2017, 11, 426–433. [Google Scholar] [CrossRef] [PubMed]
  44. Perraud, A.-L.; Takanishi, C.L.; Shen, B.; Kang, S.; Smith, M.K.; Schmitz, C.; Knowles, H.M.; Ferraris, D.; Li, W.; Zhang, J.; et al. Accumulation of Free ADP-ribose from Mitochondria Mediates Oxidative Stress-induced Gating of TRPM2 Cation Channels. J. Biol. Chem. 2005, 280, 6138–6148. [Google Scholar] [CrossRef] [Green Version]
  45. Takahashi, N.; Kozai, D.; Kobayashi, R.; Ebert, M.; Mori, Y. Roles of TRPM2 in oxidative stress. Cell Calcium 2011, 50, 279–287. [Google Scholar] [CrossRef]
  46. Evans, S.S.; Repasky, E.A.; Fisher, D.T. Fever and the thermal regulation of immunity: The immune system feels the heat. Nat. Rev. Immunol. 2015, 15, 335–349. [Google Scholar] [CrossRef]
  47. Chen, Y.; Zhou, Z.; Min, W. Mitochondria, Oxidative Stress and Innate Immunity. Front. Physiol. 2018, 9, 1487. [Google Scholar] [CrossRef] [Green Version]
  48. Fonfria, E.; Murdock, P.R.; Cusdin, F.S.; Benham, C.D.; Kelsell, R.E.; McNulty, S. Tissue Distribution Profiles of the Human TRPM Cation Channel Family. J. Recept. Signal Transduct. 2006, 26, 159–178. [Google Scholar] [CrossRef]
  49. Zhong, Z.; Zhai, Y.; Liang, S.; Mori, Y.; Han, R.; Sutterwala, F.S.; Qiao, L. TRPM2 links oxidative stress to NLRP3 inflammasome activation. Nat. Commun. 2013, 4, 1611. [Google Scholar] [CrossRef]
  50. Mittal, M.; Nepal, S.; Tsukasaki, Y.; Hecquet, C.M.; Soni, D.; Rehman, J.; Tiruppathi, C.; Malik, A.B. Neutrophil Activation of Endothelial Cell-Expressed TRPM2 Mediates Transendothelial Neutrophil Migration and Vascular Injury. Circ. Res. 2017, 121, 1081–1091. [Google Scholar] [CrossRef]
  51. Di, A.; Gao, X.-P.; Qian, F.; Kawamura, T.; Han, J.; Hecquet, C.M.; Ye, R.D.; Vogel, S.M.; Malik, A.B. The redox-sensitive cation channel TRPM2 modulates phagocyte ROS production and inflammation. Nat. Immunol. 2011, 13, 29–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Almasi, S.; Kennedy, B.E.; El-Aghil, M.; Sterea, A.M.; Gujar, S.; Partida-Sánchez, S.; El Hiani, Y. TRPM2 channel–mediated regulation of autophagy maintains mitochondrial function and promotes gastric cancer cell survival via the JNK-signaling pathway. J. Biol. Chem. 2018, 293, 3637–3650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Zhang, W.; Chu, X.; Tong, Q.; Cheung, J.Y.; Conrad, K.; Masker, K.; Miller, B.A. A Novel TRPM2 Isoform Inhibits Calcium Influx and Susceptibility to Cell Death. J. Biol. Chem. 2003, 278, 16222–16229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Miller, B.A.; Hoffman, N.E.; Merali, S.; Zhang, X.-Q.; Wang, J.; Rajan, S.; Shanmughapriya, S.; Gao, E.; Barrero, C.; Mallilankaraman, K.; et al. TRPM2 Channels Protect against Cardiac Ischemia-Reperfusion Injury: Role of Mitochondria. J. Biol. Chem. 2014, 289, 7615–7629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Gelderblom, M.; Melzer, N.; Schattling, B.; Göb, E.; Hicking, G.; Arunachalam, P.; Bittner, S.; Ufer, F.; Herrmann, A.M.; Bernreuther, C.; et al. Transient Receptor Potential Melastatin Subfamily Member 2 Cation Channel Regulates Detrimental Immune Cell Invasion in Ischemic Stroke. Stroke A J. Cereb. Circ. 2014, 45, 3395–3402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ye, M.; Yang, W.; Ainscough, J.; Hu, X.-P.; Li, X.; Sedo, A.; Zhang, X.; Chen, Z.; Beech, D.; Sivaprasadarao, A.; et al. TRPM2 channel deficiency prevents delayed cytosolic Zn2+ accumulation and CA1 pyramidal neuronal death after transient global ischemia. Cell Death Dis. 2014, 5, e1541. [Google Scholar] [CrossRef]
  57. Zong, P.; Feng, J.; Yue, Z.; Wu, G.; Sun, B.; He, Y.; Miller, B.; Albert, S.Y.; Su, Z.; Mori, Y.; et al. Functional Coupling of TRPM2 and NMDARs exacerbates excitotoxicity in ischemic brain injury. bioRxiv 2021. [Google Scholar]
  58. Park, L.; Wang, G.; Moore, J.; Girouard, H.; Zhou, P.; Anrather, J.; Iadecola, C. The key role of transient receptor potential melastatin-2 channels in amyloid-β-induced neurovascular dysfunction. Nat. Commun. 2014, 5, 5318. [Google Scholar] [CrossRef] [Green Version]
  59. Ostapchenko, V.G.; Chen, M.; Guzman, M.S.; Xie, Y.-F.; LaVine, N.; Fan, J.; Beraldo, F.H.; Martyn, A.C.; Belrose, J.C.; Mori, Y.; et al. The Transient Receptor Potential Melastatin 2 (TRPM2) Channel Contributes to β-Amyloid Oligomer-Related Neurotoxicity and Memory Impairment. J. Neurosci. Off. J. Soc. Neurosci. 2015, 35, 15157–15169. [Google Scholar] [CrossRef] [Green Version]
  60. Hoffman, N.E.; Miller, B.A.; Wang, J.; Elrod, J.W.; Rajan, S.; Gao, E.; Song, J.; Zhang, X.-Q.; Hirschler-Laszkiewicz, I.; Shanmughapriya, S.; et al. Ca2+ entry via Trpm2 is essential for cardiac myocyte bioenergetics maintenance. Am. J. Physiol. Circ. Physiol. 2015, 308, H637–H650. [Google Scholar] [CrossRef] [Green Version]
  61. Düzen, I.V.; Yavuz, F.; Vuruskan, E.; Saracoglu, E.; Poyraz, F.; Göksülük, H.; Candemir, B.; Demiryürek, S. Leukocyte TRP channel gene expressions in patients with non-valvular atrial fibrillation. Sci. Rep. 2017, 7, 1–7. [Google Scholar] [CrossRef] [PubMed]
  62. Alves-Lopes, R.; Neves, K.B.; Anagnostopoulou, A.; Rios, F.J.; Lacchini, S.; Montezano, A.C.; Touyz, R.M. Crosstalk Between Vascular Redox and Calcium Signaling in Hypertension Involves TRPM2 (Transient Receptor Potential Melastatin 2) Cation Channel. Hypertension 2020, 75, 139–149. [Google Scholar] [CrossRef] [PubMed]
  63. Zong, P.; Feng, J.; Yue, Z.; Albert, S.Y.; Mori, Y.; Yue, L. TRPM2 deficiency protects against atherosclerosis by inhibiting TRPM2-CD36 inflammatory axis in macrophages. bioRxiv 2021. [Google Scholar]
  64. Hecquet, C.M.; Ahmmed, G.U.; Vogel, S.M.; Malik, A.B. Role of TRPM2 Channel in Mediating H2O2 -Induced Ca2+ Entry and Endothelial Hyperpermeability. Circ. Res. 2008, 102, 347–355. [Google Scholar] [CrossRef] [Green Version]
  65. Uchida, K.; Dezaki, K.; Damdindorj, B.; Inada, H.; Shiuchi, T.; Mori, Y.; Yada, T.; Minokoshi, Y.; Tominaga, M. Lack of TRPM2 Impaired Insulin Secretion and Glucose Metabolisms in Mice. Diabetes 2010, 60, 119–126. [Google Scholar] [CrossRef] [Green Version]
  66. Gao, G.; Wang, W.; Tadagavadi, R.K.; Briley, N.E.; Love, M.I.; Miller, B.A.; Reeves, W.B. TRPM2 mediates ischemic kidney injury and oxidant stress through RAC1. J. Clin. Investig. 2014, 124, 4989–5001. [Google Scholar] [CrossRef] [Green Version]
  67. Miller, B.A. TRPM2 in Cancer. Cell Calcium 2019, 80, 8–17. [Google Scholar] [CrossRef]
  68. Dardiotis, E.; Aloizou, A.-M.; Markoula, S.; Siokas, V.; Tsarouhas, K.; Tzanakakis, G.; Libra, M.; Kyritsis, A.P.; Brotis, A.; Aschner, M.; et al. Cancer-associated stroke: Pathophysiology, detection and management (Review). Int. J. Oncol. 2019, 54, 779–796. [Google Scholar] [CrossRef] [Green Version]
  69. Adams, H.P., Jr.; Bendixen, B.H.; Kappelle, L.J.; Biller, J.; Love, B.B.; Gordon, D.L.; Marsh, E.E., 3rd. Classification of subtype of acute ischemic stroke. Definitions for use in a multicenter clinical trial. TOAST. Trial of Org 10172 in Acute Stroke Treatment. Stroke 1993, 24, 35–41. [Google Scholar] [CrossRef] [Green Version]
  70. Marini, C.; De Santis, F.; Sacco, S.; Russo, T.; Olivieri, L.; Totaro, R.; Carolei, A. Contribution of Atrial Fibrillation to Incidence and Outcome of Ischemic Stroke: Results from a Population-Based Study. Stroke 2005, 36, 1115–1119. [Google Scholar] [CrossRef] [Green Version]
  71. Nattel, S.; Harada, M. Atrial Remodeling and Atrial Fibrillation: Recent advances and translational perspectives. J. Am. Coll. Cardiol. 2014, 63, 2335–2345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Nattel, S. Molecular and Cellular Mechanisms of Atrial Fibrosis in Atrial Fibrillation. JACC Clin. Electrophysiol. 2017, 3, 425–435. [Google Scholar] [CrossRef] [PubMed]
  73. Staerk, L.; Sherer, J.A.; Ko, D.; Benjamin, E.; Helm, R.H. Atrial Fibrillation: Epidemiology, Pathophysiology, and Clinical Outcomes. Circ. Res. 2017, 120, 1501–1517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Ferrucci, L.; Fabbri, E. Inflammageing: Chronic inflammation in ageing, cardiovascular disease, and frailty. Nat. Rev. Cardiol. 2018, 15, 505–522. [Google Scholar] [CrossRef]
  75. Hu, Y.-F.; Chen, Y.-J.; Lin, Y.-J.; Chen, S.-A. Inflammation and the pathogenesis of atrial fibrillation. Nat. Rev. Cardiol. 2015, 12, 230–243. [Google Scholar] [CrossRef]
  76. Kakae, M.; Miyanohara, J.; Morishima, M.; Nagayasu, K.; Mori, Y.; Shirakawa, H.; Kaneko, S.; Morsishima, M. Pathophysiological role of TRPM2 in age-related cognitive impairment in mice. Neuroscience 2019, 408, 204–213. [Google Scholar] [CrossRef]
  77. Belrose, J.C.; Xie, Y.-F.; Gierszewski, L.J.; MacDonald, J.F.; Jackson, M.F. Loss of glutathione homeostasis associated with neuronal senescence facilitates TRPM2 channel activation in cultured hippocampal pyramidal neurons. Mol. Brain 2012, 5, 11. [Google Scholar] [CrossRef] [Green Version]
  78. Yue, Z.; Du, J.; Xia, J.; Qin, X.; Zhang, Y.; He, Y.; Fusco, D.; Liang, B.; Yue, L. A Potential Role of TRPM2 Mediated Inflammation in Heart Disease. Circulation 2018, 128 (Suppl. 22). [Google Scholar]
  79. Korantzopoulos, P.; Letsas, K.; Fragakis, N.; Tse, G.; Liu, T. Oxidative stress and atrial fibrillation: An update. Free Radic. Res. 2018, 52, 1199–1209. [Google Scholar] [CrossRef]
  80. Mihm, M.J.; Yu, F.; Carnes, C.; Reiser, P.J.; McCarthy, P.M.; Van Wagoner, D.R.; Bauer, J.A. Impaired Myofibrillar Energetics and Oxidative Injury During Human Atrial Fibrillation. Circulation 2001, 104, 174–180. [Google Scholar] [CrossRef] [Green Version]
  81. Adam, O.; Lavall, D.; Theobald, K.; Hohl, M.; Grube, M.; Ameling, S.; Sussman, M.A.; Rosenkranz, S.; Kroemer, H.K.; Schäfers, H.-J.; et al. Rac1-Induced Connective Tissue Growth Factor Regulates Connexin 43 and N-Cadherin Expression in Atrial Fibrillation. J. Am. Coll. Cardiol. 2010, 55, 469–480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Feng, J.; Armillei, M.K.; Yu, A.S.; Liang, B.T.; Runnels, L.W.; Yue, L. Ca2+ Signaling in Cardiac Fibroblasts and Fibrosis-Associated Heart Diseases. J. Cardiovasc. Dev. Dis. 2019, 6, 34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Harada, M.; Luo, X.; Qi, X.Y.; Tadevosyan, A.; Maguy, A.; Ördög, B.; Ledoux, J.; Kato, T.; Naud, P.; Voigt, N.; et al. Transient Receptor Potential Canonical-3 Channel–Dependent Fibroblast Regulation in Atrial Fibrillation. Circulation 2012, 126, 2051–2064. [Google Scholar] [CrossRef]
  84. Denham, N.; Pearman, C.; Caldwell, J.L.; Madders, G.W.P.; Eisner, D.; Trafford, A.W.; Dibb, K.M. Calcium in the Pathophysiology of Atrial Fibrillation and Heart Failure. Front. Physiol. 2018, 9, 1380. [Google Scholar] [CrossRef] [Green Version]
  85. Burashnikov, A.; Antzelevitch, C. Reinduction of Atrial Fibrillation Immediately After Termination of the Arrhythmia Is Mediated by Late Phase 3 Early Afterdepolarization–Induced Triggered Activity. Circulation 2003, 107, 2355–2360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Qi, X.; Yeh, Y.-H.; Chartier, D.; Xiao, L.; Tsuji, Y.; Brundel, B.J.; Kodama, I.; Nattel, S. The Calcium/Calmodulin/Kinase System and Arrhythmogenic Afterdepolarizations in Bradycardia-Related Acquired Long-QT Syndrome. Circ. Arrhythmia Electrophysiol. 2009, 2, 295–304. [Google Scholar] [CrossRef] [Green Version]
  87. Hankey, G.J. Potential New Risk Factors for Ischemic Stroke: What is their potential? Stroke 2006, 37, 2181–2188. [Google Scholar] [CrossRef] [Green Version]
  88. 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]
  89. Howard, V.J.; Madsen, T.E.; Kleindorfer, D.O.; Judd, S.E.; Rhodes, J.D.; Soliman, E.Z.; Kissela, B.M.; Safford, M.M.; Moy, C.S.; McClure, L.A.; et al. Sex and Race Differences in the Association of Incident Ischemic Stroke With Risk Factors. JAMA Neurol. 2019, 76, 179–186. [Google Scholar] [CrossRef] [Green Version]
  90. Gorgui, J.; Gorshkov, M.; Khan, N.; Daskalopoulou, S.S. Hypertension as a Risk Factor for Ischemic Stroke in Women. Can. J. Cardiol. 2014, 30, 774–782. [Google Scholar] [CrossRef]
  91. Harrison, D.G.; Coffman, T.M.; Wilcox, C.S. Pathophysiology of Hypertension: The Mosaic Theory and Beyond. Circ. Res. 2021, 128, 847–863. [Google Scholar] [CrossRef] [PubMed]
  92. Van Varik, B.J.; Rennenberg, R.J.M.W.; Reutelingsperger, C.P.; Kroon, A.A.; De Leeuw, P.W.; Schurgers, L.J. Mechanisms of arterial remodeling: Lessons from genetic diseases. Front. Genet. 2012, 3, 290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Konukoglu, D.; Uzun, H. Endothelial Dysfunction and Hypertension. In Hypertension Basic Research to Clinical Practice; Springer: Berlin/Heidelberg, Germany, 2016; pp. 511–540. [Google Scholar] [CrossRef]
  94. Satoh, K.; Nigro, P.; Berk, B.C. Oxidative Stress and Vascular Smooth Muscle Cell Growth: A Mechanistic Linkage by Cyclophilin A. Antioxidants Redox Signal. 2010, 12, 675–682. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Fortuño, A.; José, G.S.; Moreno, M.U.; Díez, J.; Zalba, G. Oxidative stress and vascular remodelling. Exp. Physiol. 2005, 90, 457–462. [Google Scholar] [CrossRef] [PubMed]
  96. Sinha, N.; Dabla, P.K. Oxidative Stress and Antioxidants in Hypertension–A Current Review. Curr. Hypertens. Rev. 2015, 11, 132–142. [Google Scholar] [CrossRef]
  97. Griendling, K.K.; Camargo, L.L.; Rios, F.J.; Alves-Lopes, R.; Montezano, A.C.; Touyz, R.M. Oxidative Stress and Hypertension. Circ. Res. 2021, 128, 993–1020. [Google Scholar] [CrossRef]
  98. Tran, Q.-K.; Ohashi, K.; Watanabe, H. Calcium signalling in endothelial cells. Cardiovasc. Res. 2000, 48, 13–22. [Google Scholar] [CrossRef] [Green Version]
  99. Amberg, G.C.; Navedo, M.F. Calcium dynamics in vascular smooth muscle. Microcirculation 2013, 20, 281–289. [Google Scholar] [CrossRef] [Green Version]
  100. Kuchan, M.J.; Frangos, J.A. Role of calcium and calmodulin in flow-induced nitric oxide production in endothelial cells. Am. J. Physiol. Physiol. 1994, 266, C628–C636. [Google Scholar] [CrossRef]
  101. Ding, Y.; Vaziri, N.D. Calcium channel blockade enhances nitric oxide synthase expression by cultured endothelial cells. Hypertension 1998, 32, 718–723. [Google Scholar] [CrossRef] [Green Version]
  102. Batova, S.; Dewever, J.; Balligand, J.-L.; Godfraind, T.; Dessy, C.; Feron, O. The calcium channel blocker amlodipine promotes the unclamping of eNOS from caveolin in endothelial cells. Cardiovasc. Res. 2006, 71, 478–485. [Google Scholar] [CrossRef] [PubMed]
  103. Resnick, L. The cellular ionic basis of hypertension and allied clinical conditions. Prog. Cardiovasc. Dis. 1999, 42, 1–22. [Google Scholar] [CrossRef]
  104. Sonkusare, S.; Palade, P.T.; Marsh, J.D.; Telemaque, S.; Pesic, A.; Rusch, N.J. Vascular calcium channels and high blood pressure: Pathophysiology and therapeutic implications. Vasc. Pharmacol. 2006, 44, 131–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Wang, Q.; Guo, W.; Hao, B.; Shi, X.; Lu, Y.; Wong, C.W.; Ma, V.W.; Yip, T.T.; Au, J.S.; Hao, Q.; et al. Mechanistic study of TRPM2-Ca2+-CAMK2-BECN1 signaling in oxidative stress-induced autophagy inhibition. Autophagy 2016, 12, 1340–1354. [Google Scholar] [CrossRef] [Green Version]
  106. Zhao, Q.; Li, J.; Ko, W.-H.; Kwan, Y.-W.; Jiang, L.; Sun, L.; Yao, X. TRPM2 promotes autophagic degradation in vascular smooth muscle cells. Sci. Rep. 2020, 10, 1–11. [Google Scholar] [CrossRef]
  107. McCarthy, C.G.; Wenceslau, C.F.; Calmasini, F.B.; Klee, N.S.; Brands, M.W.; Joe, B.; Webb, R.C. Reconstitution of autophagy ameliorates vascular function and arterial stiffening in spontaneously hypertensive rats. Am. J. Physiol. Circ. Physiol. 2019, 317, H1013–H1027. [Google Scholar] [CrossRef]
  108. Amarenco, P.; Cohen, A.; Tzourio, C.; Bertrand, B.; Hommel, M.; Besson, G.; Chauvel, C.; Touboul, P.-J.T.; Bousser, M.-G. Atherosclerotic Disease of the Aortic Arch and the Risk of Ischemic Stroke. New Engl. J. Med. 1994, 331, 1474–1479. [Google Scholar] [CrossRef]
  109. Joh, J.H.; Cho, S. Cardiovascular risk of carotid atherosclerosis: Global consensus beyond societal guidelines. Lancet Glob. Health 2020, 8, e625–e626. [Google Scholar] [CrossRef]
  110. Banerjee, C.; Chimowitz, M.I. Stroke Caused by Atherosclerosis of the Major Intracranial Arteries. Circ. Res. 2017, 120, 502–513. [Google Scholar] [CrossRef] [Green Version]
  111. Moore, K.J.; Sheedy, F.J.; Fisher, E.A. Macrophages in atherosclerosis: A dynamic balance. Nat. Rev. Immunol. 2013, 13, 709–721. [Google Scholar] [CrossRef]
  112. Moore, K.J.; Tabas, I. Macrophages in the pathogenesis of atherosclerosis. Cell 2011, 145, 341–355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Heiner, I.; Eisfeld, J.; Warnstedt, M.U.; Radukina, N.; Jüngling, E.; Lückhoff, A. Endogenous ADP-ribose enables calcium-regulated cation currents through TRPM2 channels in neutrophil granulocytes. Biochem. J. 2006, 398, 225–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Morad, H.; Luqman, S.; Tan, C.-H.; Swann, V.; McNaughton, P.A. TRPM2 ion channels steer neutrophils towards a source of hydrogen peroxide. Sci. Rep. 2021, 11, 1–16. [Google Scholar] [CrossRef]
  115. Wittmann, C.; Chockley, P.; Singh, S.K.; Pase, L.; Lieschke, G.; Grabher, C. Hydrogen Peroxide in Inflammation: Messenger, Guide, and Assassin. Adv. Hematol. 2012, 2012, 1–6. [Google Scholar] [CrossRef] [PubMed]
  116. Niethammer, P.; Grabher, C.; Look, A.T.; Mitchison, T.J. A tissue-scale gradient of hydrogen peroxide mediates rapid wound detection in zebrafish. Nature 2009, 459, 996–999. [Google Scholar] [CrossRef]
  117. Zhang, Z.; Zhang, W.; Jung, D.Y.; Ko, H.J.; Lee, Y.; Friedline, R.H.; Lee, E.; Jun, J.; Ma, Z.; Kim, F.; et al. TRPM2 Ca2+ channel regulates energy balance and glucose metabolism. Am. J. Physiol. Metab. 2012, 302, E807–E816. [Google Scholar] [CrossRef] [Green Version]
  118. Yamamoto, S.; Shimizu, S.; Kiyonaka, S.; Takahashi, N.; Wajima, T.; Hara, Y.; Negoro, T.; Hiroi, T.; Kiuchi, Y.; Okada, T.; et al. TRPM2-mediated Ca2+ influx induces chemokine production in monocytes that aggravates inflammatory neutrophil infiltration. Nat. Med. 2008, 14, 738–747. [Google Scholar] [CrossRef]
  119. Silverstein, R.L.; Li, W.; Park, Y.M.; Rahaman, S.O. Mechanisms of cell signaling by the scavenger receptor CD36: Implications in atherosclerosis and thrombosis. Trans. Am. Clin. Clim. Assoc. 2010, 121, 206–220. [Google Scholar]
  120. Moore, K.J.; Freeman, M.W. Scavenger Receptors in Atherosclerosis: Beyond lipid uptake. Arter. Thromb. Vasc. Biol. 2006, 26, 1702–1711. [Google Scholar] [CrossRef] [Green Version]
  121. Baldrighi, M.; Mallat, Z.; Li, X. NLRP3 inflammasome pathways in atherosclerosis. Atherosclerosis 2017, 267, 127–138. [Google Scholar] [CrossRef] [Green Version]
  122. Zhao, M.; Liu, Y.; Wang, X.; New, L.; Han, J.; Brunk, U.T. Activation of the p38 MAP kinase pathway is required for foam cell formation from macrophages exposed to oxidized LDL. APMIS Acta Pathol. Microbiol. Immunol. Scand. 2002, 110, 458–468. [Google Scholar] [CrossRef] [PubMed]
  123. Ricci, R.; Sumara, G.; Sumara, I.; Rozenberg, I.; Kurrer, M.; Akhmedov, A.; Hersberger, M.; Eriksson, U.; Eberli, F.R.; Becher, B.; et al. Requirement of JNK2 for Scavenger Receptor A-Mediated Foam Cell Formation in Atherogenesis. Science 2004, 306, 1558–1561. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Rahaman, S.O.; Lennon, D.J.; Febbraio, M.; Podrez, E.A.; Hazen, S.L.; Silverstein, R.L. A CD36-dependent signaling cascade is necessary for macrophage foam cell formation. Cell Metab. 2006, 4, 211–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Beceiro, S.; Radin, J.N.; Chatuvedi, R.; Piazuelo, M.B.; Horvarth, D.J.; Cortado, H.; Gu, Y.; Dixon, B.; Gu, C.; Lange, I.; et al. TRPM2 ion channels regulate macrophage polarization and gastric inflammation during Helicobacter pylori infection. Mucosal Immunol. 2016, 10, 493–507. [Google Scholar] [CrossRef]
  126. Knowles, H.; Heizer, J.W.; Li, Y.; Chapman, K.; Ogden, C.A.; Andreasen, K.; Shapland, E.; Kucera, G.; Mogan, J.; Humann, J.; et al. Transient Receptor Potential Melastatin 2 (TRPM2) ion channel is required for innate immunity against Listeria monocytogenes. Proc. Natl. Acad. Sci. USA 2011, 108, 11578–11583. [Google Scholar] [CrossRef] [Green Version]
  127. Zou, J.; Ainscough, J.F.; Yang, W.; Sedo, A.; Yu, S.-P.; Mei, Z.-Z.; Sivaprasadarao, A.; Beech, D.J.; Jiang, L.-H. A differential role of macrophage TRPM2 channels in Ca2+ signaling and cell death in early responses to H2O2. Am. J. Physiol. Physiol. 2013, 305, C61–C69. [Google Scholar] [CrossRef]
  128. Duewell, P.; Kono, H.; Rayner, K.; Sirois, C.M.; Vladimer, G.; Bauernfeind, F.G.; Abela, G.S.; Franchi, L.; Nuñez, G.; Schnurr, M.; et al. NLRP3 inflammasomes are required for atherogenesis and activated by cholesterol crystals. Nature 2010, 464, 1357–1361, Erratum in Nature 2010, 466, 652. [Google Scholar] [CrossRef] [Green Version]
  129. Chen, R.; Ovbiagele, B.; Feng, W. Diabetes and Stroke: Epidemiology, Pathophysiology, Pharmaceuticals and Outcomes. Am. J. Med. Sci. 2016, 351, 380–386. [Google Scholar] [CrossRef] [Green Version]
  130. Kissela, B.M.; Khoury, J.; Kleindorfer, D.; Woo, D.; Schneider, A.; Alwell, K.; Miller, R.; Ewing, I.; Moomaw, C.J.; Szaflarski, J.P.; et al. Epidemiology of Ischemic Stroke in Patients With Diabetes: The greater Cincinnati/Northern Kentucky Stroke Study. Diabetes Care 2005, 28, 355–359. [Google Scholar] [CrossRef] [Green Version]
  131. Gæde, P.; Lund-Andersen, H.; Parving, H.-H.; Pedersen, O. Effect of a Multifactorial Intervention on Mortality in Type 2 Diabetes. New Engl. J. Med. 2008, 358, 580–591. [Google Scholar] [CrossRef] [Green Version]
  132. A DiMeglio, L.; Evans-Molina, C.; Oram, R. Type 1 diabetes. Lancet 2018, 391, 2449–2462. [Google Scholar] [CrossRef]
  133. Grodsky, G.M.; Bennett, L.L. Cation Requirements for Insulin Secretion in the Isolated Perfused Pancreas. Diabetes 1966, 15, 910. [Google Scholar] [CrossRef] [PubMed]
  134. Gromada, J.; Høy, M.; Renström, E.; Bokvist, K.; Eliasson, L.; Göpel, S.; Rorsman, P. CaM kinase II-dependent mobilization of secretory granules underlies acetylcholine-induced stimulation of exocytosis in mouse pancreatic B-cells. J. Physiol. 1999, 518, 745–759. [Google Scholar] [CrossRef] [PubMed]
  135. Togashi, K.; Hara, Y.; Tominaga, T.; Higashi, T.; Konishi, Y.; Mori, Y.; Tominaga, M. TRPM2 activation by cyclic ADP-ribose at body temperature is involved in insulin secretion. EMBO J. 2006, 25, 1804–1815. [Google Scholar] [CrossRef] [PubMed]
  136. Bari, M.R.; Akbar, S.; Eweida, M.; Kühn, F.J.; Gustafsson, A.J.; Lückhoff, A.; Islam, S. H2O2-induced Ca2+ influx and its inhibition by N-(p-amylcinnamoyl) anthranilic acid in the β-cells: Involvement of TRPM2 channels. J. Cell. Mol. Med. 2009, 13, 3260–3267. [Google Scholar] [CrossRef] [Green Version]
  137. Atkinson, M.A.; Eisenbarth, G.S.; Michels, A.W. Type 1 diabetes. Lancet 2014, 383, 69–82. [Google Scholar] [CrossRef] [Green Version]
  138. Espinoza-Jiménez, A.; Navarrete-Peón, A.; Terrazas, L.I. Alternatively Activated Macrophages in Types 1 and 2 Diabetes. Mediat. Inflamm. 2012, 2012, 815953. [Google Scholar] [CrossRef]
  139. Drews, G.; Krippeit-Drews, P.; Düfer, M. Oxidative stress and beta-cell dysfunction. Pflug. Arch. 2010, 460, 703–718. [Google Scholar] [CrossRef]
  140. Domínguez, C.; Ruiz, E.; Gussinye, M.; Carrascosa, A. Oxidative Stress at Onset and in Early Stages of Type 1 Diabetes in Children and Adolescents. Diabetes Care 1998, 21, 1736–1742. [Google Scholar] [CrossRef]
  141. DeFronzo, R.A.; Ferrannini, E.; Groop, L.; Henry, R.R.; Herman, W.H.; Holst, J.J.; Hu, F.B.; Kahn, C.R.; Raz, I.; Shulman, G.; et al. Type 2 diabetes mellitus. Nat. Rev. Dis. Prim. 2015, 1, 15019. [Google Scholar] [CrossRef]
  142. Galicia-Garcia, U.; Benito-Vicente, A.; Jebari, S.; Larrea-Sebal, A.; Siddiqi, H.; Uribe, K.B.; Ostolaza, H.; Martín, C. Pathophysiology of Type 2 Diabetes Mellitus. Int. J. Mol. Sci. 2020, 21, 6275. [Google Scholar] [CrossRef] [PubMed]
  143. Appari, M.; Channon, K.M.; McNeill, E. Metabolic Regulation of Adipose Tissue Macrophage Function in Obesity and Diabetes. Antioxidants Redox Signal. 2018, 29, 297–312. [Google Scholar] [CrossRef] [PubMed]
  144. Kraakman, M.; Murphy, A.; Jandeleit-Dahm, K.; Kammoun, H.L. Macrophage Polarization in Obesity and Type 2 Diabetes: Weighing Down Our Understanding of Macrophage Function. Front. Immunol. 2014, 5, 470. [Google Scholar] [CrossRef] [PubMed]
  145. Orliaguet, L.; Dalmas, E.; Drareni, K.; Venteclef, N.; Alzaid, F. Mechanisms of Macrophage Polarization in Insulin Signaling and Sensitivity. Front. Endocrinol. 2020, 11, 62. [Google Scholar] [CrossRef] [Green Version]
  146. Di Gregorio, G.B.; Yao-Borengasser, A.; Rasouli, N.; Varma, V.; Lu, T.; Miles, L.M.; Ranganathan, G.; Peterson, C.A.; McGehee, R.E.; Kern, P.A. Expression of CD68 and Macrophage Chemoattractant Protein-1 Genes in Human Adipose and Muscle Tissues: Association with cytokine expression, insulin resistance, and reduction by pioglitazone. Diabetes 2005, 54, 2305–2313. [Google Scholar] [CrossRef] [Green Version]
  147. Wellen, K.E.; Hotamisligil, G.S. Inflammation, stress, and diabetes. J. Clin. Investig. 2005, 115, 1111–1119. [Google Scholar] [CrossRef] [Green Version]
  148. Koppaka, S.; Kehlenbrink, S.; Carey, M.; Li, W.; Sanchez, E.; Lee, D.-E.; Lee, H.; Chen, J.; Carrasco, E.; Kishore, P.; et al. Reduced Adipose Tissue Macrophage Content Is Associated With Improved Insulin Sensitivity in Thiazolidinedione-Treated Diabetic Humans. Diabetes 2013, 62, 1843–1854. [Google Scholar] [CrossRef] [Green Version]
  149. Wagner, D.D.; Burger, P.C. Platelets in Inflammation and Thrombosis. Arter. Thromb. Vasc. Biol. 2003, 23, 2131–2137. [Google Scholar] [CrossRef]
  150. Koupenova, M.; Clancy, L.; Corkrey, H.A.; Freedman, J.E. Circulating Platelets as Mediators of Immunity, Inflammation, and Thrombosis. Circ. Res. 2018, 122, 337–351. [Google Scholar] [CrossRef]
  151. Nesbitt, W.; Giuliano, S.; Kulkarni, S.; Dopheide, S.M.; Harper, I.; Jackson, S.P. Intercellular calcium communication regulates platelet aggregation and thrombus growth. J. Cell Biol. 2003, 160, 1151–1161. [Google Scholar] [CrossRef] [Green Version]
  152. Varga-Szabo, D.; Braun, A.; Nieswandt, B. Calcium signaling in platelets. J. Thromb. Haemost. 2009, 7, 1057–1066. [Google Scholar] [CrossRef]
  153. Freedman, J.E. Oxidative Stress and Platelets. Arter. Thromb. Vasc. Biol. 2008, 28, s11–s16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Carter, R.N.; Tolhurst, G.; Walmsley, G.; Vizuete-Forster, M.; Miller, N.; Mahaut-Smith, M.P. Molecular and electrophysiological characterization of transient receptor potential ion channels in the primary murine megakaryocyte. J. Physiol. 2006, 576, 151–162. [Google Scholar] [CrossRef] [PubMed]
  155. Authi, K.S. TRP Channels in Platelet Function. Handb. Exp. Pharm. 2007, 179, 425–443. [Google Scholar] [CrossRef]
  156. Isenberg, J.S.; Romeo, M.J.; Yu, C.; Nghiem, K.; Monsale, J.; Rick, M.E.; Wink, D.A.; Frazier, W.A.; Roberts, D.D. Thrombospondin-1 stimulates platelet aggregation by blocking the antithrombotic activity of nitric oxide/cGMP signaling. Blood 2008, 111, 613–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Lin, T.-N.; Kim, G.-M.; Chen, J.-J.; Cheung, W.-M.; He, Y.Y.; Hsu, C.Y. Differential Regulation of Thrombospondin-1 and Thrombospondin-2 After Focal Cerebral Ischemia/Reperfusion. Stroke 2003, 34, 177–186. [Google Scholar] [CrossRef] [PubMed]
  158. Saver, J.L.; Starkman, S.; Eckstein, M.; Stratton, S.; Pratt, F.D.; Hamilton, S.; Conwit, R.; Liebeskind, D.S.; Sung, G.; Kramer, I.; et al. Prehospital Use of Magnesium Sulfate as Neuroprotection in Acute Stroke. New Engl. J. Med. 2015, 372, 528–536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Matei, N.; Camara, J.; Zhang, J.H. The Next Step in the Treatment of Stroke. Front. Neurol. 2021, 11, 582605. [Google Scholar] [CrossRef]
  160. Hawkins, B.; Davis, T. The Blood-Brain Barrier/Neurovascular Unit in Health and Disease. Pharmacol. Rev. 2005, 57, 173–185. [Google Scholar] [CrossRef]
  161. Bell, A.H.; Miller, S.L.; Castillo-Melendez, M.; Malhotra, A. The Neurovascular Unit: Effects of Brain Insults During the Perinatal Period. Front. Neurosci. 2020, 13, 1452. [Google Scholar] [CrossRef]
  162. Kniesel, U.; Wolburg, H. Tight Junctions of the Blood–Brain Barrier. Cell. Mol. Neurobiol. 2000, 20, 57–76. [Google Scholar] [CrossRef] [PubMed]
  163. Kassner, A.; Merali, Z. Assessment of Blood–Brain Barrier Disruption in Stroke. Stroke 2015, 46, 3310–3315. [Google Scholar] [CrossRef] [PubMed]
  164. Andjelkovic, A.V.; Xiang, J.; Stamatovic, S.M.; Hua, Y.; Xi, G.; Wang, M.M.; Keep, R.F. Endothelial Targets in Stroke: Translating Animal Models to Human. Arter. Thromb. Vasc. Biol. 2019, 39, 2240–2247. [Google Scholar] [CrossRef] [PubMed]
  165. Wake, H.; Fields, R.D. Physiological function of microglia. Neuron Glia Biol. 2011, 7, 1–3. [Google Scholar] [CrossRef] [Green Version]
  166. Guruswamy, R.; ElAli, A. Complex Roles of Microglial Cells in Ischemic Stroke Pathobiology: New Insights and Future Directions. Int. J. Mol. Sci. 2017, 18, 496. [Google Scholar] [CrossRef] [Green Version]
  167. Simpson, D.S.A.; Oliver, P.L. ROS Generation in Microglia: Understanding Oxidative Stress and Inflammation in Neurodegenerative Disease. Antioxidants 2020, 9, 743. [Google Scholar] [CrossRef]
  168. Fonfria, E.; Mattei, C.; Hill, K.; Brown, J.T.; Randall, A.; Benham, C.D.; Skaper, S.D.; Campbell, C.A.; Crook, B.; Murdock, P.R.; et al. TRPM2 Is Elevated in the tMCAO Stroke Model, Transcriptionally Regulated, and Functionally Expressed in C13 Microglia. J. Recept. Signal Transduct. 2006, 26, 179–198. [Google Scholar] [CrossRef]
  169. Kraft, R.; Grimm, C.; Grosse, K.; Hoffmann, A.; Sauerbruch, S.; Kettenmann, H.; Schultz, G.; Harteneck, C. Hydrogen peroxide and ADP-ribose induce TRPM2-mediated calcium influx and cation currents in microglia. Am. J. Physiol. Physiol. 2004, 286, C129–C137. [Google Scholar] [CrossRef]
  170. Mortadza, S.A.S.; Sim, J.A.; Neubrand, V.E.; Jiang, L.-H. A critical role of TRPM2 channel in Aβ42 -induced microglial activation and generation of tumor necrosis factor-α. Glia 2018, 66, 562–575. [Google Scholar] [CrossRef] [Green Version]
  171. Lee, M.; Cho, T.; Jantaratnotai, N.; Wang, Y.T.; McGeer, E.; McGeer, P.L. Depletion of GSH in glial cells induces neurotoxicity: Relevance to aging and degenerative neurological diseases. FASEB J. 2010, 24, 2533–2545. [Google Scholar] [CrossRef]
  172. Miyanohara, J.; Kakae, M.; Nagayasu, K.; Nakagawa, T.; Mori, Y.; Arai, K.; Shirakawa, H.; Kaneko, S. TRPM2 Channel Aggravates CNS Inflammation and Cognitive Impairment via Activation of Microglia in Chronic Cerebral Hypoperfusion. J. Neurosci. 2018, 38, 3520–3533. [Google Scholar] [CrossRef] [PubMed]
  173. Cho, S.; Park, E.-M.; Febbraio, M.; Anrather, J.; Park, L.; Racchumi, G.; Silverstein, R.L.; Iadecola, C. The Class B Scavenger Receptor CD36 Mediates Free Radical Production and Tissue Injury in Cerebral Ischemia. J. Neurosci. Off. J. Soc. Neurosci. 2005, 25, 2504–2512. [Google Scholar] [CrossRef] [PubMed]
  174. Takano, T.; Oberheim, N.; Cotrina, M.L.; Nedergaard, M. Astrocytes and Ischemic Injury. Stroke 2009, 40, S8–S12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Barreto, G.E.; White, R.E.; Ouyang, Y.; Xu, L.; Giffard, R.G. Astrocytes: Targets for Neuroprotection in Stroke. Central Nerv. Syst. Agents Med. Chem. 2011, 11, 164–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Roy-O’Reilly, M.; McCullough, L.D. Astrocytes fuel the fire of lymphocyte toxicity after stroke. Proc. Natl. Acad. Sci. USA 2017, 114, 425–427. [Google Scholar] [CrossRef] [Green Version]
  177. Xu, S.; Lu, J.; Shao, A.; Zhang, J.H.; Zhang, J. Glial Cells: Role of the Immune Response in Ischemic Stroke. Front. Immunol. 2020, 11, 294. [Google Scholar] [CrossRef]
  178. Yamada, T.; Kawahara, K.; Kosugi, T.; Tanaka, M. Nitric Oxide Produced During Sublethal Ischemia Is Crucial for the Preconditioning-Induced Down-Regulation of Glutamate Transporter GLT-1 in Neuron/Astrocyte Co-Cultures. Neurochem. Res. 2006, 31, 49–56. [Google Scholar] [CrossRef] [Green Version]
  179. Xu, M.-X.; Zhu, Y.-F.; Chang, H.-F.; Liang, Y. Nanoceria restrains PM2.5-induced metabolic disorder and hypothalamus inflammation by inhibition of astrocytes activation related NF-kappaB pathway in Nrf2 deficient mice. Free Radic. Biol. Med. 2016, 99, 259–272. [Google Scholar] [CrossRef]
  180. Weilinger, N.L.; Lohman, A.W.; Rakai, B.D.; Ma, E.M.M.; Bialecki, J.; Maslieieva, V.; Rilea, T.; Bandet, M.V.; Ikuta, N.T.; Scott, L.; et al. Metabotropic NMDA receptor signaling couples Src family kinases to pannexin-1 during excitotoxicity. Nat. Neurosci. 2016, 19, 432–442. [Google Scholar] [CrossRef]
  181. Kaneko, S.; Kawakami, S.; Hara, Y.; Wakamori, M.; Itoh, E.; Minami, T.; Takada, Y.; Kume, T.; Katsuki, H.; Mori, Y.; et al. A Critical Role of TRPM2 in Neuronal Cell Death by Hydrogen Peroxide. J. Pharmacol. Sci. 2006, 101, 66–76. [Google Scholar] [CrossRef] [Green Version]
  182. Verma, S.; Quillinan, N.; Yang, Y.-F.; Nakayama, S.; Cheng, J.; Kelley, M.; Herson, P. TRPM2 channel activation following in vitro ischemia contributes to male hippocampal cell death. Neurosci. Lett. 2012, 530, 41–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Olah, M.E.; Jackson, M.F.; Li, H.; Perez, Y.; Sun, H.-S.; Kiyonaka, S.; Mori, Y.; Tymianski, M.; Macdonald, J.F. Ca2+-dependent induction of TRPM2 currents in hippocampal neurons. J. Physiol. 2009, 587, 965–979. [Google Scholar] [CrossRef] [PubMed]
  184. Huang, S.; Turlova, E.; Li, F.; Bao, M.-H.; Szeto, V.; Wong, R.; Abussaud, A.; Wang, H.; Zhu, S.; Gao, X.; et al. Transient receptor potential melastatin 2 channels (TRPM2) mediate neonatal hypoxic-ischemic brain injury in mice. Exp. Neurol. 2017, 296, 32–40. [Google Scholar] [CrossRef] [PubMed]
  185. Jia, J.; Verma, S.; Nakayama, S.; Quillinan, N.; Grafe, M.R.; Hurn, P.D.; Herson, P.S. Sex Differences in Neuroprotection Provided by Inhibition of TRPM2 Channels following Experimental Stroke. J. Cereb. Blood Flow Metab. Off. J. Int. Soc. Cereb. Blood Flow Metab. 2011, 31, 2160–2168. [Google Scholar] [CrossRef] [Green Version]
  186. Shimizu, T.; A Macey, T.; Quillinan, N.; Klawitter, J.; Perraud, A.-L.L.; Traystman, R.J.; Herson, P.S. Androgen and PARP-1 Regulation of TRPM2 Channels after Ischemic Injury. J. Cereb. Blood Flow Metab. 2013, 33, 1549–1555. [Google Scholar] [CrossRef] [Green Version]
  187. Wu, Q.J.; Tymianski, M. Targeting NMDA receptors in stroke: New hope in neuroprotection. Mol. Brain 2018, 11, 1–14. [Google Scholar] [CrossRef]
  188. Sanz-Clemente, A.; Nicoll, R.A.; Roche, K.W. Diversity in NMDA Receptor Composition: Many regulators, many consequences. Neuroscientist 2012, 19, 62–75. [Google Scholar] [CrossRef] [Green Version]
  189. Hardingham, G.; Fukunaga, Y.; Bading, H. Extrasynaptic NMDARs oppose synaptic NMDARs by triggering CREB shut-off and cell death pathways. Nat. Neurosci. 2002, 5, 405–414. [Google Scholar] [CrossRef]
  190. Martel, M.-A.; Ryan, T.; Bell, K.F.; Fowler, J.; McMahon, A.; Al-Mubarak, B.; Komiyama, N.H.; Horsburgh, K.; Kind, P.; Grant, S.G.; et al. The Subtype of GluN2 C-terminal Domain Determines the Response to Excitotoxic Insults. Neuron 2012, 74, 543–556. [Google Scholar] [CrossRef] [Green Version]
  191. Hardingham, G.E.; Bading, H. Synaptic versus extrasynaptic NMDA receptor signalling: Implications for neurodegenerative disorders. Nat. Rev. Neurosci. 2010, 11, 682–696. [Google Scholar] [CrossRef] [Green Version]
  192. Alim, I.; Teves, L.; Li, R.; Mori, Y.; Tymianski, M. Modulation of NMDAR Subunit Expression by TRPM2 Channels Regulates Neuronal Vulnerability to Ischemic Cell Death. J. Neurosci. Off. J. Soc. Neurosci. 2013, 33, 17264–17277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Nakayama, S.; Vest, R.; Traystman, R.J.; Herson, P.S. Sexually Dimorphic Response of TRPM2 Inhibition Following Cardiac Arrest-Induced Global Cerebral Ischemia in Mice. J. Mol. Neurosci. 2013, 51, 92–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Shimizu, T.; Dietz, R.M.; Cruz-Torres, I.; Strnad, F.; Garske, A.K.; Moreno, M.; Venna, V.R.; Quillinan, N.; Herson, P.S. Extended therapeutic window of a novel peptide inhibitor of TRPM2 channels following focal cerebral ischemia. Exp. Neurol. 2016, 283, 151–156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Vennekens, R.; Nilius, B. Insights into TRPM4 Function, Regulation and Physiological Role. In Transient Receptor Potential (TRP) Channels; Springer: Berlin/Heidelberg, Germany, 2007; pp. 269–285. [Google Scholar] [CrossRef]
  196. Gerzanich, V.; Woo, S.K.; Vennekens, R.; Tsymbalyuk, O.; Ivanova, S.; Ivanov, A.; Geng, Z.; Chen, Z.; Nilius, B.; Flockerzi, V.; et al. De novo expression of Trpm4 initiates secondary hemorrhage in spinal cord injury. Nat. Med. 2009, 15, 185–191. [Google Scholar] [CrossRef] [PubMed]
  197. Loh, K.; Ng, G.; Yu, C.Y.; Fhu, C.K.; Yu, D.; Vennekens, R.; Nilius, B.; Soong, T.W.; Liao, P. TRPM4 inhibition promotes angiogenesis after ischemic stroke. Pflügers Arch. -Eur. J. Physiol. 2013, 466, 563–576. [Google Scholar] [CrossRef]
  198. Yan, J.; Bengtson, C.P.; Buchthal, B.; Hagenston, A.M.; Bading, H. Coupling of NMDA receptors and TRPM4 guides discovery of unconventional neuroprotectants. Science 2020, 370. [Google Scholar] [CrossRef] [PubMed]
  199. Lin, J.; Xiong, Z.-G. TRPM7 is a unique target for therapeutic intervention of stroke. Int. J. Physiol. Pathophysiol. Pharmacol. 2017, 9, 211–216. [Google Scholar]
  200. Inoue, K.; Branigan, D.; Xiong, Z.-G. Zinc-induced Neurotoxicity Mediated by Transient Receptor Potential Melastatin 7 Channels. J. Biol. Chem. 2010, 285, 7430–7439. [Google Scholar] [CrossRef] [Green Version]
  201. Sun, H.-S.; Jackson, M.F.; Martin, L.J.; Jansen, K.; Teves, L.; Cui, H.; Kiyonaka, S.; Mori, Y.; Jones, M.; Forder, J.P.; et al. Suppression of hippocampal TRPM7 protein prevents delayed neuronal death in brain ischemia. Nat. Neurosci. 2009, 12, 1300–1307. [Google Scholar] [CrossRef]
  202. Singh, A.P.; Biswas, A.; Shukla, A.W.; Maiti, P. Targeted therapy in chronic diseases using nanomaterial-based drug delivery vehicles. Signal Transduct. Target. Ther. 2019, 4, 1–21. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The comprehensive role of TRPM2 in ischemic stroke.
Figure 1. The comprehensive role of TRPM2 in ischemic stroke.
Cells 11 00491 g001
Figure 2. TRPM2 activation by oxidative stress and TRPM2-mediated Ca2+ signaling under oxidative stress boosts ROS production in mitochondria. Increased ROS production activates PARP, which produces ADPR, a potent endogenous TRPM2 activator. ADPR activates TRPM2 by binding to the NUDT9-H domain at the C terminus. TRPM2 activation leads to Ca2+ influx from the extracellular environment and Ca2+ release from lysosomes. TRPM2-mediated Ca2+ signaling is critical in regulating a series of cellular functions.
Figure 2. TRPM2 activation by oxidative stress and TRPM2-mediated Ca2+ signaling under oxidative stress boosts ROS production in mitochondria. Increased ROS production activates PARP, which produces ADPR, a potent endogenous TRPM2 activator. ADPR activates TRPM2 by binding to the NUDT9-H domain at the C terminus. TRPM2 activation leads to Ca2+ influx from the extracellular environment and Ca2+ release from lysosomes. TRPM2-mediated Ca2+ signaling is critical in regulating a series of cellular functions.
Cells 11 00491 g002
Figure 3. TRPM2 in inflammation. (A) Leukocyte extravasation during inflammation. (B) TRPM2-mediated Ca2+ influx leads to tight-junction molecule degradation (VE-cadherin and occludin) and mitochondrial dysfunction in endothelial cells. (C) TRPM2-mediated Ca2+ influx is needed for immune cell migration and activation. During inflammation, ROS production in mitochondria is increased, which activates PARP in mitochondria or in the nucleus and enhances the production of ADPR. Increased ADPR potentiates TRPM2-mediated Ca2+ influx, which further increases the production of ROS in mitochondria, leading to the formation of a feed-forward vicious cycle. ROS-, PARP-, and Ca2+ -related signaling pathways increase the expression of proinflammatory genes, such as TNF-α, IL-1β, MCP1, and MIF. Moreover, TRPM2-mediated Ca2+ influx promotes cytoskeleton rearrangement and immune cell migration.
Figure 3. TRPM2 in inflammation. (A) Leukocyte extravasation during inflammation. (B) TRPM2-mediated Ca2+ influx leads to tight-junction molecule degradation (VE-cadherin and occludin) and mitochondrial dysfunction in endothelial cells. (C) TRPM2-mediated Ca2+ influx is needed for immune cell migration and activation. During inflammation, ROS production in mitochondria is increased, which activates PARP in mitochondria or in the nucleus and enhances the production of ADPR. Increased ADPR potentiates TRPM2-mediated Ca2+ influx, which further increases the production of ROS in mitochondria, leading to the formation of a feed-forward vicious cycle. ROS-, PARP-, and Ca2+ -related signaling pathways increase the expression of proinflammatory genes, such as TNF-α, IL-1β, MCP1, and MIF. Moreover, TRPM2-mediated Ca2+ influx promotes cytoskeleton rearrangement and immune cell migration.
Cells 11 00491 g003
Figure 4. TRPM2 in ischemic stroke.
Figure 4. TRPM2 in ischemic stroke.
Cells 11 00491 g004
Table 1. Summary of studies on the role of TRPM2 in aggregating ischemic brain/neuron injury.
Table 1. Summary of studies on the role of TRPM2 in aggregating ischemic brain/neuron injury.
SpeciesModel (s)Target (s)Mechanism (s)
RatIn vitro H2O2 treatmentNeuronKnockdown of TRPM2 using siRNA inhibited H2O2−induced neuronal death [181].
MouseIn vitro OGD
In vivo tMCAO
NeuronKnockdown of TRPM2 using shRNA inhibited OGD-induced neuronal death, and reduced infarction size after MCAO [185].
MouseIn vitro OGD
In vivo BCCAO
NeuronGlobal knockout of TRPM2 inhibited increase of intracellular Zn2+ and ROS production, and attenuated neuronal death after global ischemia [56].
MouseIn vitro H2O2 treatment
In vivo tMCAO
NeuronGlobal knockout of TRPM2 inhibited increase of neuro-excitability in response to H2O2, and attenuated neuronal death and brain injury after tMCAO by promoting pro-survival signaling while inhibiting pro-apoptotic signaling [192].
MouseIn vivo neonatal hypoxic ischemic brain injury modelNeuronGlobal knockout of TRPM2 attenuated neuronal death and reduced infarct size after hypoxic–ischemic brain injury partially by regulating GSK-3β signaling [184].
MouseIn vivo CA/CPRNeuronInhibition of TRPM2 using clotrimazole reduced neuronal death in male mice, but not in female mice [193].
MouseIn vivo tMCAONeuronInhibition of TRPM2 using a peptide inhibitor reduced infarction size after MCAO [194].
MouseIn vivo tMCAOImmune cellsTransplantation of bone marrow from global TRPM2 knockout mice into wild-type mice, or inhibition of TRPM2 using ACA reduced infarction size in wild-type mice after MCAO [55].
MouseIn vivo BCASMicrogliaGlobal knockout of TRPM2 inhibited brain damage and cognitive dysfunction [172].
HumanIn vitro BSO treatmentMicroglia and astrocytesKnockdown of TRPM2 using siRNA attenuated the inflammatory responses in human microglia and astrocytes [171].
OGD, oxygen–glucose deprivation; BCCAO, bilateral common carotid artery occlusion; tMCAO, transient middle cerebral artery occlusion; CA/CPR, cardiac arrest and cardiopulmonary resuscitation; BCAS, bilateral common carotid artery stenosis; BSO, D,L-buthionine-S,R-sulfoximine.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zong, P.; Lin, Q.; Feng, J.; Yue, L. A Systemic Review of the Integral Role of TRPM2 in Ischemic Stroke: From Upstream Risk Factors to Ultimate Neuronal Death. Cells 2022, 11, 491. https://doi.org/10.3390/cells11030491

AMA Style

Zong P, Lin Q, Feng J, Yue L. A Systemic Review of the Integral Role of TRPM2 in Ischemic Stroke: From Upstream Risk Factors to Ultimate Neuronal Death. Cells. 2022; 11(3):491. https://doi.org/10.3390/cells11030491

Chicago/Turabian Style

Zong, Pengyu, Qiaoshan Lin, Jianlin Feng, and Lixia Yue. 2022. "A Systemic Review of the Integral Role of TRPM2 in Ischemic Stroke: From Upstream Risk Factors to Ultimate Neuronal Death" Cells 11, no. 3: 491. https://doi.org/10.3390/cells11030491

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