Next Article in Journal
Metabolite and Transcriptome Profiling Analysis Provides New Insights into the Distinctive Effects of Exogenous Melatonin on Flavonoids Biosynthesis in Rosa rugosa
Previous Article in Journal
Immune-Enhancing Effects of Gwakhyangjeonggi-san in RAW 264.7 Macrophage Cells through the MAPK/NF-κB Signaling Pathways
Previous Article in Special Issue
Xenon’s Sedative Effect Is Mediated by Interaction with the Cyclic Nucleotide-Binding Domain (CNBD) of HCN2 Channels Expressed by Thalamocortical Neurons of the Ventrobasal Nucleus in Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Inhibitors of NLRP3 Inflammasome Formation: A Cardioprotective Role for the Gasotransmitters Carbon Monoxide, Nitric Oxide, and Hydrogen Sulphide in Acute Myocardial Infarction

Department of Pharmacology and Toxicology and HeartOtago, School of Biomedical Sciences, University of Otago, Dunedin 9054, New Zealand
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(17), 9247; https://doi.org/10.3390/ijms25179247
Submission received: 26 July 2024 / Revised: 21 August 2024 / Accepted: 21 August 2024 / Published: 26 August 2024

Abstract

:
Myocardial ischaemia reperfusion injury (IRI) occurring from acute coronary artery disease or cardiac surgical interventions such as bypass surgery can result in myocardial dysfunction, presenting as, myocardial “stunning”, arrhythmias, infarction, and adverse cardiac remodelling, and may lead to both a systemic and a localised inflammatory response. This localised cardiac inflammatory response is regulated through the nucleotide-binding oligomerisation domain (NACHT), leucine-rich repeat (LRR)-containing protein family pyrin domain (PYD)-3 (NLRP3) inflammasome, a multimeric structure whose components are present within both cardiomyocytes and in cardiac fibroblasts. The NLRP3 inflammasome is activated via numerous danger signals produced by IRI and is central to the resultant innate immune response. Inhibition of this inherent inflammatory response has been shown to protect the myocardium and stop the occurrence of the systemic inflammatory response syndrome following the re-establishment of cardiac circulation. Therapies to prevent NLRP3 inflammasome formation in the clinic are currently lacking, and therefore, new pharmacotherapies are required. This review will highlight the role of the NLRP3 inflammasome within the myocardium during IRI and will examine the therapeutic value of inflammasome inhibition with particular attention to carbon monoxide, nitric oxide, and hydrogen sulphide as potential pharmacological inhibitors of NLRP3 inflammasome activation.

1. Introduction

Pharmacological or percutaneous revascularisation procedures are commonly applied clinical interventions used to rapidly restore coronary perfusion and salvage as much of the jeopardised myocardium as possible in acutely ischaemic hearts. In selected patients, coronary artery bypass graft (CABG) procedures can be advocated to improve coronary flow [1] to the damaged myocardium. Paradoxically, however, the reintroduction of oxygenated blood to the damaged myocardium in all these revascularisation procedures can result in ischaemia reperfusion injury (IRI) and provoke further risk of infarct development in the postoperative period. The incidence of adverse outcomes in CABG patients varies with general factors such as the duration of aortic cross-clamp time [2], patient age [3], gender, morbidity, and body weight [4]. Additionally, the increasingly common presence of underlying cardiac pathology, such as hypertrophy [5] resulting from hypertensive, diabetic or rheumatic heart diseases, or genetic aetiologies, remains a major contributing factor to poor outcomes [6]. Hence, increasing effort has been directed to the development of effective cardioprotective conditioning interventions to protect the myocardium against reperfusion injury and to improve revascularisation outcomes for high-risk patients with vulnerable hearts [7,8]. If unchecked, IRI can result in both systemic and innate localised inflammatory responses. This review will examine the current understanding of the role of the inherent localised NLRP3-regulated inflammatory response in myocardial injury and will assess the potential of the novel gaseous conditioning agents carbon monoxide (CO), nitric oxide (NO), and hydrogen sulphide (H2S) as effective inflammasome inhibitors.

2. Myocardial Ischaemia Reperfusion Injury (IRI)

The imposition of an ischaemic insult to the myocardium rapidly depletes oxygen availability to cardiomyocytes for mitochondrial β-oxidation so that cellular metabolism becomes dependent on anaerobic glycolysis [9,10]. Prolongation of the ischaemic episode, however, can lead to the myocardium becoming progressively acidotic (<pH 7) through the accumulation of lactate and H+, thereby exacerbating myocardial injury [11,12,13]. The resultant intracellular accumulation of protons in this acidotic environment activates the Na+/H+ antiporter, serving to elevate cytosolic Na+ and drive the Na+/Ca2+ exchanger to import more Ca2+ into the myocardial cells [13,14,15,16]. High cytosolic Ca2+ levels have been firmly associated with arrhythmia formation, contributing to myocardial dysfunction, morbidity and mortality after IRI [17,18]. Raised cytosolic Ca2+ levels are also sequestered by mitochondria. The resultant mitochondrial Ca2+ overload facilitates ischaemic injury via mitochondrial swelling and the activation of proteases such as calpains, which lead to cell death via activation of pro-apoptotic proteins [19,20,21,22]. Within minutes of reperfusion onset, the reactive oxygen species (ROS) formed through mitochondrial electron leak, in conjunction with elevated mitochondrial Ca2+ and intracellular pH normalisation towards physiological levels (pH 7.2), permit the opening of the non-selective mitochondrial permeable transition pore (mPTP) [23,24]. Prolonged opening of the mPTP catastrophically disrupts the mitochondrial electrochemical gradient, impairing adenosine triphosphate (ATP) production and releasing a burst of ROS and cytochrome c to initiate apoptotic and necrotic cell death processes [25,26,27,28]. These various cell death processes defined by distinct modes of pathway activation subsequentially permeate throughout the infarcted region of the myocardium, resulting in irreversible cardiac injury.

Inflammatory Responses Following Myocardial IRI

Concurrent to the cardiomyocyte apoptosis, necrosis and pyroptosis events occurring in cardiac IRI, pro-inflammatory signalling cascades are simultaneously upregulated, which leads to long-term cardiac fibrosis, ventricular hypertrophy, and stenosis, rendering the heart susceptible to failure [29,30]. Previous research has mainly explored the impact of a systemic inflammatory response, occurring on reperfusion in cardiac surgery [31]. While this complement cascade-mediated injury is significant, there have been few investigations into the inherent inflammatory response that occurs within the myocardium during ischaemic clamping and reperfusion procedures [32]. By removing the cellular debris and dead cells and initiating tissue regeneration, this sterile inflammatory response forms an essential initial recovery process that has been suggested to salvage the infarcted myocardium allowing scar tissue to form [33]. These beneficial effects are considered to outweigh the associated adverse long-term inflammatory pathology, which if unattenuated or excessive, can cause extensive cell death leading to increased fibroblast activation [34,35]. The resulting infiltrative and reparative fibrotic scaring can produce a stiffened ventricular wall adversely impacting myocardial function and viability and ultimately reducing end-organ perfusion, with significant adverse clinical outcomes [35].
This innate ischaemic myocardial inflammatory response is initiated through the formation of multimeric inflammasome structures. A number of distinct inflammasome types have been identified, although only NLRP1 and NLRP3 have been indicated to be upregulated during cardiac ischaemia and reperfusion, acting as key contributors to myocardial IRI, with NLRP3 being the most extensively studied and recognised [34,36,37,38]. Activation of NLRP3 causes necrotic cell death and releases active pro-inflammatory cytokines that can recruit innate immune cells to the injury site [39]. The subsequent systemic inflammatory response syndrome is implicated in myocardial dysfunction and end-organ injury. Additionally, caspase-1 activation via NLRP3 can lead to the release of the pro-inflammatory protein, high mobility group box protein 1 (HMGB1), into the microenvironment, which can further exacerbate the injury within the myocardium and throughout the body [40]. Inhibition of NLRP3 activation within the isolated heart during bypass surgery may therefore serve to simultaneously prevent the formation of an intrinsic myocardial inflammatory cascade and the consequent development of a systemic inflammatory response syndrome.

3. The NLRP3 Inflammasome

The NLRP3 inflammasome (also referred to as NALP3 or cryopyrin) is a large multimeric protein complex (Figure 1) containing a central nucleotide-binding oligomerisation domain (NACHT), a C-terminus leucine-rich repeat (LRR) and an N-terminal pyrin domain (PYD) [41]. The PYD domain located at the N-terminal of NLRP3 facilitates the recruitment of the apoptosis-associated speck-like protein (ASC) through PYD-PYD interactions [42]. ASC also contains a caspase activation and recruitment domain (CARD) allowing the enlistment of pro-caspase-1 and hence activation into caspase-1 [43]. The NIMA (“never in mitosis gene a”)-related serine/threonine kinase 7 (NEK7) is a scaffolding protein consisting of a catalytic domain that interacts with the LRR domain of NLRP3 [44]. This interaction is proposed to break the inactive NLRP3 cage, transforming it into an active NLRP3 inflammasome disk [45].

3.1. NLRP3 Priming

Formation and activation of the NLRP3 inflammasome requires both priming and activation steps [46,47]. While cytosolic NLRP3 is normally expressed at low levels, priming of NLRP3 is needed to increase messenger ribonucleic acid levels within the cytosol to allow for inflammasome formation [48]. Additionally, baseline levels of cytosolic NLRP3 are rendered inactive via ubiquitination at the LRR domain [49,50]. In order for this priming event to be activated, pattern recognition receptors (PRRs) must first bind pathogen- and damage-associated molecular patterns (PAMPs and DAMPs) generated during tissue damage such as cellular debris, mitochondrial ROS, and ATP released into the cytosol [51,52]. DAMPs arising during IRI have been significantly linked to the development of irreversible changes within the myocardium, including apoptosis, fibrosis, and hypertrophy [32,37,39]. PRRs involved in this process include toll-like receptors (TLRs), nucleotide-binding oligomerisation domain-like receptors (NLRs), tumour necrosis factor receptors (TNFRs), and interleukin-1 receptors (IL-1Rs) (Figure 2) [49,53,54]. Activation of PRRs will in turn activate multiple adaptor proteins downstream of each receptor subtype, which can include either myeloid differentiation factor 88 (MyD88) or tumour necrosis factor 1 death domain protein (TRADD) dependent signalling [55,56] and subsequently lead to nuclear factor (NF)-κB activation [57,58,59]. In vitro and in vivo knockout (KO) studies have suggested that the adaptor FS-7-associated surface antigen-associated death domain protein (FADD) also facilitates NF-κB activation and is indispensable for NLRP3 priming [58,60].
NLRP3 inflammasome components have been identified in both cardiomyocyte and leukocytes as well as non-cardiomyocyte populations such as fibroblasts and endothelial cells within the mammalian heart [61]. However, as none of the components of the inflammasome are constitutively expressed within healthy cardiomyocytes, translocation of NF-κB to the nucleus is responsible for inducing the expression of genes for NLRP3, ASC, and pro-caspase 1, as well as for the substrates of activated caspase-1, pro-IL-1β, pro-IL-18, and gasdermin D (GSDMD) [49,62]. The increased cytosolic levels of these proteins allow NLRP3 inflammasome formation to occur (Figure 2) and subsequently, trigger an inflammatory response. NLRP3 inflammasome activation can also be primed via post-translational modifications; activation of the interleukin receptor-associated kinases 1/4 (IRAK1/4) phosphorylates the LRR domain of NLRP3 and promotes NLRP3-ASC interactions [63]. Non-transcriptional priming involving the lysine-63 deubiquitinate (BRCC3) mediates NLRP3 deubiquitylation of free NLRP3 at the LRR domain to also promote NLRP3-ASC interactions [48,50].

3.2. NLRP3 Activation

Prior to complete NLRP3 activation, a complex combination of cellular signalling events is required to allow NLRP3 inflammasome assembly and activation. The four most important activating signals in the context of IRI are generated through ion channel dysregulation [64], ROS accumulation [65], mitochondrial rupture [52], and lysosome disruption [66].
Ischaemia depletes intracellular ATP levels, resulting in ATP-gated ion channel dysregulation, facilitating cytosolic Na+ accumulation. Consequently, Na+ accumulation results in significant K+ efflux upon reperfusion [67]. He, et al. [68] reported that upon K+ efflux, NEK7 interacts with the LRR domain of NLRP3 to facilitate inflammasome assembly, which was abolished in NEK7−/− KO cells. In addition, K+ efflux increases Ca2+-independent phospholipase A2 activation by approximately 75% in human monocytes, aiding in the cleavage of pro-IL-1β into active IL-1β [69]. While intracellular Ca2+ accumulation is considered insufficient to induce NLRP3 inflammasome activation [70], extracellular Ca2+ accumulation was established to activate the G-protein-coupled receptor class C group 6-member A to facilitate NLRP3 inflammasome activation [71].
Concurrent with ion channel dysregulation, ROS accumulation and release is a strong signal for NLRP3 activation as identified in several in vivo models of IRI [32,72,73]. ROS species can provoke mitochondrial dysfunction by stimulating the mPTP opening, resulting in mitochondrial swelling, and the release of mitochondrial ROS and oxidised mitochondrial DNA (ox-mtDNA) (Figure 2) [74]. While mitochondrial ROS can induce NLRP3 inflammasome activation, it is not the sole driver [75,76]. Instead, ox-mtDNA appears to be the most significant stimulus as it directly promotes NLRP3-ASC interactions through protein modifications on ASC, increasing the binding affinity for NLRP3 [74]. Furthermore, ox-mtDNA can also directly bind to the PYD of NLRP3 to facilitate inflammasome assembly [77]. Lastly, phagocytosis of DAMPS and other cell debris leads to lysosome destabilisation and release of the enzyme cathepsin B into the cytosol to directly act on the LRR domain of NLRP3 and aid inflammasome formation (Figure 2) [78,79,80].
The subsequent formation of the NLRP3 inflammasome activates caspase-1, which cleaves and activates inactive pro-inflammatory cytokines, pro-IL-1β, and pro-IL-18 into active IL-1β and IL-18, allowing pyroptosis, a form of programmed necrosis, to occur (Figure 2) [81,82,83]. The process of pyroptosis occurs via the cleavage and separation of the N- and C-terminals of GSDMD by caspase-1 [83]. The C-terminus acts as an auto-inhibitor of the pore-forming N-terminus; however, once GSDMD is cleaved, the N-terminal forms nanopores in the cell membrane, leading to cell swelling and the release of active IL-1β and IL-18 [84,85]. Inhibition of NF-κB was established to suppress GSDMD transcription and reduce NLRP3 inflammasome-mediated pyroptosis activity within cardiomyocytes [86]. Pyroptosis can also be facilitated by the aggregation of ASC molecules into 1–2 μm ASC dimers and the recruitment of caspase-1 [87].

3.3. Targeted NLRP3 Inhibitors

Several small molecules examined in pre-clinical trials have shown promise as direct inhibitors of the NLRP3 inflammasome signalling pathway; however, translation into clinical use will require further assessment based on safety profiles and applicability within clinical protocols. The two most well-studied NLRP3 inhibitors that have shown therapeutic promise in IRI include MCC950 and OLT1177; this review will briefly introduce these agents but will not comprehensively discuss them as they are out of the scope of this review.
MCC950 is a sulfonylurea-containing compound that potently (IC50 7.5–8 nM) binds to the ATP hydrolysis pocket of NLRP3, proximal to the Walker B motif on the NACHT domain in human monocyte-derived macrophages [88,89]. This interaction stabilises NLRP3 into an inactive conformation [90]. In a C57BL/6 mouse model of myocardial infarction (MI) involving left anterior descending artery (LAD) occlusion, 14 days of postoperative delivery of MCC950 improved ejection fraction (26.7%), reduced myocardial fibrosis (13%), and reduced myocardial NLRP3, IL-1β, IL-18, and caspase-1 whilst decreasing immune cell infiltration [91].
OLT1177 is a β-sulfonyl nitrile known to directly target and inhibit the NLRP3 inflammasome as it does not alter transcriptional regulation or activation signals involved in NLRP3 activation but decreases IL-1β and IL-18 levels as demonstrated in LPS-stimulated human monocyte-derived macrophages. [92]. In a Swiss mouse model of MI involving LAD ligation, postoperative administration of OLT1177 (6–600 mg/kg) dose-dependently reduced infarct size (between 36–62%) and preserved left ventricular fractional shortening up to 7 days post-surgery [93]. As of 2021, a Phase 1b human clinical trial conducted in patients with systolic heart failure showed that OLT1177 delivered at oral doses between 500 and 2000 mg/day for 14 days exerted no significant adverse events and improved left ventricular ejection fraction (by 5%) [94].
Recent advancements led to the development of INF4E and other chemically related non-sulfonylurea-based NLRP3 inhibitors such as INF200 [95,96]. INF4E is an acrylamide derivative considered to be a covalent inhibitor of NLRP3 and its ATPase activity. Pharmacological inhibition of the NLRP3 inflammasome with the novel potent inhibitor, INF4E, provided a cardioprotective response by activating the reperfusion injury salvage kinase (RISK) pathway and improved mitochondrial function within an ex vivo IRI rat model [97]. Pre-treatment of isolated rat hearts with INF4E (50 μM) for 20 min prior to IRI produced a 22% reduction in infarct size and improved haemodynamic function 60 min after reperfusion onset compared to controls. Penna, et al. [98] reproduced this finding, identifying RISK activation and smaller infarct size and area at risk with INF4E pre-treatment compared to the untreated group.

4. Gasotransmitters as Cardioprotectants

As described above, the initiation of an ischaemic insult and subsequent reperfusion induces not only an aggressive myocardial injury but also an intense cardiac inflammatory response, contributing to myocardial IRI. It is imperative that new, targeted cardioprotective agents capable of reducing the associated innate inflammatory response are introduced. A potential avenue currently gaining attention is the ability of the three gasotransmitters—carbon monoxide (CO), nitric oxide (NO), and hydrogen sulphide (H2S)—to reduce myocardial IRI through multiple signalling pathways, but most importantly, via a targeted response on the NLRP3 inflammasome.

4.1. Endogenous Carbon Monoxide

Carbon monoxide (CO) gas is commonly, and perhaps melodramatically, described as a silent killer, as exposure to high concentrations of CO can impair the O2-carrying capacity of the blood due to a ~220-times higher affinity of haemoglobin for CO over O2 [99]. Paradoxically, however, endogenous production of this gaseous signalling molecule occurs ubiquitously in mammalian cells, including pertinently in humans [100]. The rate-limiting enzymes, heme oxygenases (HOs), are responsible for this endogenous production [101]. Currently, there are two main human isoforms of HOs: HO-1, and HO-2. These enzymes catalyse the oxidative degradation of free heme to biliverdin, free iron, and CO [102,103]. HO-2 is constitutively expressed within endothelial cells and smooth muscle cells [104]. In contrast, HO-1 is inducible and is activated in response to periods of ischaemia and reperfusion and other cellular stresses, indicating a high capacity for the cardiovascular system to produce CO [105]. During these stresses, the expression of HO-1 within the heart is increased and can exert cytoprotective effects on tissue and the myocardium [106]. These effects are attributable to the production of biliverdin, an antioxidant, and more importantly, to the induction of cell signalling pathways mediated by CO [107,108].

4.2. Exogenous CO Delivery

While the dangers of exposure to CO are well defined, low doses of CO (<100 ppm) and molecules that release controlled amounts of CO have been demonstrated to produce therapeutic responses whilst avoiding toxic effects [109,110,111]. The initial discovery of endogenous CO production and the subsequent demonstration of the pleiotropic therapeutic effects of this gaseous molecule led to the call for pharmacological agents to be developed with the ability to systemically release controlled low concentrations of CO as potential therapeutics and as research tools. These chemical prodrugs have been classed as ‘carbon monoxide releasing molecules’ (CORMs) [112]. The first characterised CORMs consisted of a transition metal centre, such as manganese or ruthenium, surrounded by carbonyl groups, and include CORM-1 [Mn2(CO)10], CORM-2 [Ru(CO)3Cl2]2, and CORM-3 [Ru(CO)3Cl-glycinate] [112]. A non-metallo CORM-A1 [Na2H3BCO2] has also been synthesized as a slower CO-releasing molecule [113]. These CORMs have been widely studied within both in vivo and in vitro applications [113,114,115,116]. Whilst CORMs have been proposed to have a number of therapeutic indications, these compounds still possess a number of features that limit their use in clinical applications. Unlike gaseous CO, administration of 20 μM CORM-2 (48 h) produced both cytoprotective and cytotoxic effects in cardiomyocytes and in renal cell studies and this toxicity was associated with the ruthenium complex rather than CO [117]. Subsequent research confirmed the inherent toxicity of the metal CORMs by demonstrating CORM-3 toxicity at 500 μM within RAW 264.7 macrophages, illustrating differences in the specific sensitivity of cell lines to their toxic effects [118]. Effects attributed to the parent compound of some of these earlier CORMs rather than being mediated by the CO released have recently been described, which, combined with idiosyncratic CO release rates, limit the value of these CORMs in the study of CO-mediated mechanisms [119]. Identifying that these toxic and off-target responses relate to the prodrug structure, rather than to CO, was an important discovery and has emphasised the need for an improved molecule with minimal toxicity to serve as a CO prodrug [120,121]. Recently developed organic CO-releasing prodrugs such as oCOm-21 have demonstrated many of the same protective effects as CO but with few of the limitations demonstrated by earlier CORMs [111,122]. These new drug developments have been designed to release CO under various triggering conditions such as physiological pH and provide pharmacologically effective CO delivery to intracellular compartments.

4.3. Pleiotropic Effects of CO

The discovery of endogenous CO production initiated research into the physiological and therapeutic properties of this pleiotropic molecule. CO induces vasodilatory, anti-apoptotic, anti-thrombotic, anti-proliferative, and anti-inflammatory effects through p38 mitogen-activated protein kinase (p38MAPK), the phosphatidylinositol 3-kinase/protein kinase B (PI3K/Akt) and the soluble guanylate cyclase (sGC) pathway [111,123,124,125]. Investigations into the vasodilatory properties have shown that CO mediates these effects via cyclic guanosine monophosphate-dependent smooth muscle relaxation [126]. Endogenous CO generated by HO-1 was identified to suppress vascular endothelial cell apoptosis, with the mechanism being reliant on the p38MAPK pathway and NF-κB [127,128]. Furthermore, this subsequent activation of p38MAPK and NF-κB stimulates the transcription of the anti-apoptotic genes, c-IAP2 and A1, to protect vascular endothelial cells from tumour necrosis factor-alpha (TNF-α)-mediated apoptosis. Additionally, the anti-apoptotic effects of CO are not solely reliant on the p38MAPK pathway, but also on the PI3K/Akt pathway within endothelial cells [125], while anti-apoptotic studies conducted in fibroblasts show that this CO effect is mediated through sGC activation [129].
The anti-inflammatory effects of endogenous CO upregulation have also been repeatedly demonstrated [124,130,131,132]. Similarly to the anti-apoptotic response, CO can produce these beneficial anti-inflammatory effects via p38MAPK in a concentration-dependent manner within RAW 246.7 macrophages [124]. This anti-inflammatory effect of CO was associated with a modulation of the cytokine profile seen as an inhibition of LPS-induced production of TNF-α, IL-1β, and macrophage inflammatory protein 1b, whilst anti-inflammatory IL-10 production was increased. However, CO was also subsequently shown to exert concurrent anti-inflammatory effects via other mechanisms; Morse, et al. [130] found that inhalation exposure to 250 ppm CO in a mouse was capable of inhibiting the LPS-induced JNK/AP-1 pathway, while Nakahira, et al. [131] demonstrated the ability of CO to inhibit TLR signalling pathways specifically on TLR 2, 4, 5, and 9 but not on TLR3. This signal inhibition correlated with the inactivation of transcription factors such as NF-κB and interferon regulatory factor-3, which consequently, inhibited cytokine production. Additionally, Qin et al. [132] demonstrated that without the presence of nuclear factor-erythroid 2-related factor-2 (Nrf2), CO could not produce an anti-inflammatory effect in an LPS-induced inflammation mouse model. Overall, the anti-inflammatory effects of low concentrations of CO can be considered therapeutically significant and not limited to a single pathway.

4.4. Inhibitory Effects of CO on NLRP3

The ability of CO and CORMs to inhibit NLRP3 inflammasome formation through multiple pathways has been reported in different tissues and models. Early patch-clamp studies conducted by Wilkinson, et al. [133] indicated that both CO gas and CORM-2 can inhibit ATP-gated purinergic (P2X) receptors, specifically at P2X2 (Figure 3). As activation of P2X receptors results in reduced intracellular K+ (an activation signal for NLRP3), this pathway was considered to represent one possible mechanism through which CO could inhibit NLRP3 formation [51,134,135]. These studies also showed that CORM-2 was a reversible, non-competitive inhibitor at P2X4, while CO gas was not [136]. However, there is no ex vivo or in vivo evidence to support the idea that CO gas can inhibit P2X4 or P2X7 receptors in the presence of ATP, even though these receptors are regarded as the only ligand-gated ion channels regulated by CO [136,137]. Therefore, it could be assumed that the inhibition identified was due to the parent metallo carbonyl structure of the early CORMs rather than the CO molecule.
Jiang, et al. [138] elucidated an inhibitory action of CO on thioredoxin interacting protein (TXNIP) to reduce NLRP3 expression (Figure 3). While this study was performed on lung tissue, Liu, et al. [139] demonstrated that TXNIP KO in both cardiac microvascular endothelial cells and cardiomyocytes also inhibits NLRP3 activation, significantly reducing IL-1β production by ~1.7 fold and suggesting a potential, indirect, inhibitory action for CO on NLRP3. Additionally, the researchers also identified that within a mouse myocardial IRI model, infarct size, measured as a ratio of infarct area to total area at risk, was significantly reduced by ~18%, and left ventricular ejection fraction, used as a measure of cardiac function, was restored. Zhou et al. [140] showed that ROS generation from NLRP3 activators such as uric acid crystals is required to release TXNIP from oxidised thioredoxin, which allows the binding and activation of NLRP3. Interestingly, inositol-requiring enzyme 1 (IRE1) has been shown to interact with TXNIP/NLRP3 signalling, indicating a significant modulatory role in NLRP3 inflammasome activation [141]. CO was shown to suppress IRE1 phosphorylation by inducing protein kinase R-like endoplasmic reticulum kinase phosphorylation (PERK) [142,143]. This finding is further backed by Zheng et al. [144] who reported that administering CORM-3 (8 mg/kg/day, i.v.) to rats for 7 days reduces IRE1 phosphorylation and decreases NLRP3 inflammasome production and pyroptosis in a spinal cord injury model (Figure 3). Lastly, Zhang et al. [145] proposed that CO interacts with NF-κB signalling to prevent upregulation of NLRP3 and pro-IL-1β. These authors have, however, indicated that CO could also inhibit the activation stage as the addition of CORM-3 after LPS priming reduced active IL-1β and IL-18 production (Figure 3). CO, therefore, has a broad inhibitory effect on the priming stage of NLRP3, preventing NLRP3 upregulation and the subsequent inflammation. To date, no research has investigated the ability of CO to directly inhibit NLRP3 inflammasome formation within myocardial IRI.

4.5. Endogenous Nitric Oxide

NO was first identified as an endothelium-derived relaxing factor by Palmer et al. [146] and found to be endogenously synthesised in endothelial cells through the degradation of L-arginine by NO synthase (NOS) [147]. Current NOS family isoforms identified include endothelial NOS (eNOS), neuronal NOS (nNOS), and inducible NOS (iNOS) [148,149,150]. eNOS is constitutively expressed within endothelial cells and is widely accepted as the primary source of NO present in the vasculature, contributing to the regulation of vascular tone, systemic and pulmonary vascular resistance, inhibition of platelet aggregation, and leukocyte adhesion [151,152,153,154]. nNOS was preliminarily identified within the brain and autonomic nerves owing to its pivotal role in controlling cerebral functions and modulating glutamatergic/GABAergic neurotransmission [155,156]. However, nNOS has also been identified to be constitutively expressed within the vascular endothelium and can contribute towards NO production to modulate cerebral blood flow and systemic arterial pressure [157,158]. In contrast, iNOS is not constitutively expressed; instead, its expression is regulated via NF-κB signalling during pro-inflammatory and oxidative stress conditions [159]. This inducible isoform is most commonly present within circulating neutrophils and macrophages but is also upregulated within cardiomyocytes in response to pro-inflammatory cytokines [148]. Upon activation, iNOS activity produces micromolar concentrations of NO that can both induce as well as attenuate, upon elevation, pro-inflammatory responses through a biphasic regulatory effect on NF-κB [160]. In contrast, NO production via eNOS and nNOS can produce nanomolar concentrations of NO, which have proven to induce cytoprotective effects [161]. Importantly, all of these NOS isoforms have been identified within the whole heart and can contribute to NO production during myocardial IRI [162].

4.6. Exogenous NO Application

The use of gaseous NO inhalation in both therapeutic and experimental research settings is greatly hindered by delivery constraints, unlimited diffusion, and lack of compartmentalisation to target tissues, and by the extremely short (0.05–2 s) half-life of this avidly reactive radical in biological conditions [163]. These characteristics have warranted the production of pharmacological agents that have the capacity to release NO or upregulate endogenous NO production to improve site-specific delivery of active, therapeutic levels of this gaseous transmitter. These compounds include NO donors such as organic nitrates, diazeniumdiolates, S-nitrosothiols, furoxans, metal nitrosyl compounds, and nitrobenzenes [164]. For decades, these compounds have been utilised both clinically and experimentally to further assess the therapeutic capabilities of this gaseous molecule and have provided some promising results [165,166]. However, it should be noted that the free radical nature of NO can produce a variety of toxic effects that pose a barrier to the safe therapeutic use of this molecule. NO is able to readily interact with superoxide, with the potential to produce a 100-fold increase in levels of the potent oxidant, pro-inflammatory peroxynitrite for every 10-fold increase in both NO and superoxide [167] and induce cardiomyocyte injury [168,169].

4.7. Pleiotropic Effects of NO

Both the endogenous synthesis and exogenous application of NO have been shown to provide cardioprotection during myocardial IRI [160,161,167]. Similar to CO, NO has a number of valuable pleiotropic actions, including anti-inflammatory, anti-apoptotic, and, most notably, vasodilatory attributes associated with the expression of NOS within the vasculature [151,170,171]. This vasodilatory mechanism involves a NO-mediated activation of sGC leading to the formation of cGMP similar to that produced in response to CO. NO has, however, been reported to avidly bind the heme moiety on sGC to produce a 200-fold increase in cGMP compared to CO, which, potentially because of the fast dissociation of CO from heme, results in CO exerting relatively low (4-fold) increases in sGC activation [172,173].
The ability of NO to reduce apoptosis by inducing the phosphorylation, and hence activation of ERK1/2 resulting in a reduction in caspase-3 activity, was demonstrated using the NO donor S-nitroso-N-acetylpenicillamine (SNAP) at 10 µM in a mouse model of myocardial IRI [170]. This protection could be reversed by the addition of the selective ERK1/2 inhibitor, U0126, indicating that the anti-apoptotic effect induced by NO is mediated via the ERK1/2 pathway [170]. It should be noted that NO has also been reported to induce apoptosis through the activation of BAX, BAK, and caspase-9, allowing cytochrome c to be released from the mitochondria, thereby activating the intrinsic apoptotic pathway [174].
Enzymatic activity by iNOS produces high (micromolar) levels of NO, which can promote the pathogenesis of inflammatory disorders such as sepsis, rheumatoid arthritis, and inflammatory bowel disease [175,176,177]. Conversely, NO production at low (nanomolar) concentrations has been recognised to promote an anti-inflammatory effect [160]. By applying both in vitro and in vivo experimental protocols, Lee et al. [178] found that increased NO release through eNOS overexpression, NO donor administration, or endothelial–macrophage co-culture can induce macrophage polarisation. This polarisation away from the pro-inflammatory M1 towards an M2 phenotype is indicated to occur endogenously through the downstream overexpression of vasodilator-stimulated phosphoprotein within macrophages. Confirmation of the physiological significance of NO in repressing M1 activation was also observed in macrophages co-cultured with eNOS-depleted aortic endothelial cells and in vasodilator-stimulated phosphoprotein-deficient transgenic mice [178]. Niedbala et al. [179] also identified the significance of NO-induced proliferation of CD4+ CD25 T cells into regulatory T cells in inflammatory tissue, allowing these cells to produce IL-10 and IL-4. This proliferative induction of regulatory T cells by NO (200–400 nM) has been paradoxically suggested to be dependent on p53 enhancing IL-2 and OX40 synthesis. In a separate study, Niedbala et al. [180] demonstrated that NO could inhibit the expression of the aryl hydrocarbon receptor in response to environmental stimulants/pollutants, thereby suppressing the function and proliferation of human Th17 cells under autoimmune conditions.

4.8. Inhibitory Effects of NO on NLRP3

Unlike CO, treatment with either SNAP or the iNOS inhibitor L-NMMA has shown that NO molecules can directly inhibit the NLRP3 inflammasome through S-nitrosylation (Figure 4) [181,182]. S-nitrosylation is a post-translational modification mechanism where NO groups are covalently added to cysteine protein thiols, regulating their function [183]. Hernandez-Cuellar et al. [181] used a biotin-switch technique to identify that S-nitrosylation occurred both at the C-terminus of the NLRP3 protein and on caspase-1, preventing early NLRP3 inflammasome formation (Figure 4). Both caspase-1 and the C-terminus of the NLRP3 inflammasome (the LRR domain) contain cysteine residues, providing multiple functional targets for S-nitrosylation to occur [184]. Thiol sidechains contained within these active sites on caspase-1 can be S-nitrosylated, attenuating the proteolytic function of caspase-1 (Figure 4) [185]. Collective evidence by Mao et al. [182] and Mishra et al. [186] further demonstrated that S-nitrosylation is required to occur directly on both NLRP3 and caspase-1 in order to suppress the formation of IL-1β, as S-nitrosylation of caspase-1 alone is insufficient. These findings have consistently shown the ability of NO to inhibit the activating stage of the NLRP3 inflammasome within macrophages, although this mechanism is yet to be directly investigated within the myocardium. Both exogenous and endogenous-derived NO have repeatedly demonstrated cardioprotection in IRI [187,188,189] and, pertinently, to reduce myocardial inflammation [190]. These findings lead to the suggestion that the mechanism through which NO reduces inflammation during myocardial IRI is afforded by a direct inhibitory effect on NLRP3 inflammasome activation, although further research will be required to confirm this.

4.9. Endogenous Hydrogen Sulphide

Hydrogen sulphide (H2S) is the third endogenous gasotransmitter shown to have beneficial physiological properties [191,192]. The oxidation state of the sulphur atom at −2 allows H2S to become oxidised at physiological pH and react with metal centres such as iron in heme, as well as oxidised thiol products, to form persulphides. H2S is predominantly synthesised through desulphhydration of biomolecules such as L-cysteine and homocysteine via cystathionine β-synthase, 3-mercaptopyruvate sulfotransferase, and cysteine aminotransferase enzymes located within the central nervous system, and by cystathionine γ-lyase (CSE) predominately found within the cytosol of cardiovascular, hepatic, and renal tissues [193,194,195].

4.10. Exogenous H2S Application

The identification of endogenous H2S involvement in a range of cytoprotective effects has spurred research into the use of both gaseous and synthetic H2S-releasing agents in experimental and therapeutic applications. Recognition, however, of the noxious nature of H2S gas and that high (100 ppm) concentrations can produce adverse effects [196] such as mitochondrial cytochrome c inhibition [197], have tempered the use of gas inhalation as a therapeutic application. These considerations have consequently warranted the development of a range of pharmacologically effective agents that can endogenously increase localised concentrations of H2S. To date, experimental studies have mainly employed either inorganic salts such as sodium hydrosulphide (NaHS) and sodium sulphide (Na2S) or organic compounds such as the naturally occurring, fast-releasing diallyl trisulphide and the slow-releasing diallyl disulphide donors [198,199]. Classified according to the mechanism through which they increase H2S levels, these synthetic organic donors include hybrid molecules containing a H2S donating moiety, spontaneously releasing H2S donors, and compounds which serve as substrates for H2S-generating enzymes [200]. One leading slow-releasing H2S donor in the clinical setting has been the orally available SG1002, which completed phase I clinical trials for safety and tolerability in patients with heart failure (seven subjects) [201]. While phase II trials are yet to be performed with SG1002, several preclinical studies have provided strong support for the use of H2S as a cardioprotectant [202,203,204]. Unfortunately, these experimental studies have provided limited data on the long-term impact of the drug, and further research using a large animal model will be required before pursuing the use of these agents within the clinic [205].

4.11. Pleiotropic Effects of H2S

There are consistent findings that augmentation of H2S through exogenous and endogenous mechanisms provides vasorelaxant, anti-apoptotic, anti-oxidative, and anti-inflammatory properties within pathologies such as endothelial dysfunction, myocardial fibrosis, and hypertrophy, leading to the benefits shown in myocardial, hepatic, and renal IRI [203,206,207,208,209]. Zhao et al. [210] originally reported that exogenous administration of H2S-solubilised gas induced vasorelaxation with an IC50 of 125 ± 14 μM within phenylephrine-precontracted rat aortic tissues, and that this effect was also obtained through bolus intravenous delivery of H2S in vivo. This H2S-induced response was inhibited both in vivo and ex vivo by the addition of glibenclamide, demonstrating that this vasorelaxant effect is mediated by ATP-dependent K+ channels (KATP). Lee et al. [211] further indicated that the vasodilatory effects of H2S within rat aortic rings were also partly regulated by a decrease in intracellular pH via the Cl/HCO3 exchanger. There are discrepancies within the literature, however, with some reports indicating a smooth muscle vasoconstrictive effect produced by H2S. Lim et al. [212] reported that the exogenous donor, NaHS (yielding 3–30 µM H2S) induced vasoconstriction via a decrease in NO production and an attenuation of forskolin-induced cAMP accumulation. This inhibition of cAMP production by H2S has also been noted in cardiac myocytes [213,214]. Whilst these articles have identified vasoconstrictive responses to H2S, the majority of the literature largely supports a beneficial vasodilatory effect, which has also been consistently reported in vivo [215,216,217].
The anti-apoptotic effects of H2S were identified within a rat LAD ligation myocardial IRI model where pre-treatment with 3 mg/kg NaHS prevented caspase-9 activation and increased Bcl-2 expression [203]. Examination of the area of myocardium at risk after IRI found that pre-treatment with NaHS had significantly decreased cardiomyocyte apoptosis from 37 ± 4% to 17 ± 2% and caspase 9 activity from 34 ± 2% to 20 ± 2%. Abrogation of this protective effect with 5-hydroxydecanoate (5-HD), a KATP channel blocker, demonstrated that these effects were secondary to the opening of mitochondrial KATP channels. This anti-apoptotic effect was also demonstrated with much lower (0.2 and 0.4 μmol/kg) doses of NaHS in a cerebral rat IRI model where treatment increased anti-apoptotic Bcl-2 expression whilst decreasing pro-apoptotic Bax expression [218].
H2S is recognised as an important player in providing cardioprotection during IRI. Sivarajah et al. [219] found that upregulation of endogenous H2S through cystathionine-γ-lyase occurred during cardiac ischaemia and reperfusion, providing cardioprotection and limiting the extent of myocardial structural injury. The cardioprotective effects afforded by H2S were abrogated by the addition of 5-HD, supporting the importance of KATP channels in H2S signalling. Furthermore, Johansen et al. [206] showed that NaHS (0.1–10 μM) produced a concentration-dependent reduction in infarct size in ex vivo Langendorff-perfused rat hearts subjected to IRI. Comparatively, in vivo administration of 50 μg/kg of NaHS at the time of reperfusion in mice subjected to 30 min of left ventricular ischaemia significantly reduced cardiac infarct size from 47.9 ± 2.9% to 13.4 ± 1.4%, reduced cardiac inflammation, and preserved mitochondrial function 24 h post reperfusion onset [220].
The anti-inflammatory capabilities of H2S have been consistently demonstrated to act through similar pathways to those activated by CO [203,218,221,222,223]. Increased phosphorylation of both p38 MAPK and NF-κB has been identified following NaHS administration in IRI studies conducted in gastric epithelium and myocardium, resulting in attenuated neutrophil infiltration and ICAM-1 and PMN accumulation within areas of tissue at risk [203,224]. Pre-treatment of these gastric epithelial cells with 100 μM NaHS after IRI also resulted in an attenuation of JNK phosphorylation [224]. A reduction in IL-1β, IL-6, and TNF-α levels associated with an inhibition of NF-κB was identified after treatment with NaHS in a rat myocardial IRI model [225]. Furthermore, H2S has also been recognised to strongly interact with Nrf2 signalling, providing significant cardioprotective effects [226,227]. Nrf2 is the primary cellular defence against oxidative stress and is normally restricted to the cytoplasm by Keap1 [228]. H2S has been identified to interact with this association by S-sulphhydration of cysteine residues on Keap1, specifically cysteine-151, allowing Nrf2 to enter the nucleus and begin transcription of antioxidant enzymes as well as, interestingly, the CO-producing enzyme, HO-1 (Figure 5) [227,229].

4.12. Inhibitory Effects of H2S on NLRP3

The anti-inflammatory capabilities of H2S have been determined to involve interactions upon NLRP3 inflammasome signalling through a variety of mechanisms: H2S induces a post-translational modification, similar to NO, wherein an additional sulphur molecule is added to thiol groups of cysteine; this event is termed S-sulphhydration [230]. Lin et al. [231] identified this mechanism on c-Jun, the subunit of activator protein-1, at Cys269 using a biotin switch assay within macrophages. This modification enhanced binding of the activator protein-1 (AP-1) to the SIRT3 promoter, increasing transcription of SIRT3 and p62, which subsequently inhibited NLRP3 inflammasome formation (Figure 5). Whilst this event was identified within macrophages, it has been consistently recognised that macrophage NLRP3 inflammasome activation plays a role in cardiovascular disease [232,233,234]. Furthermore, H2S has also previously been shown to interact with and enhance AP-1 binding and SIRT3 expression within endothelial cells, inducing a vasoprotective effect, providing further insight into its capability to interact with NLRP3 inflammasome activation within the cardiovascular system (Figure 5) [209].
Zhao et al. [235] reported that the induction of intracerebral haemorrhage producing neurological deficit and brain oedema in a rat model temporarily decreased cystathionine-β-synthase expression and endogenous H2S levels within the striatum lasting 48 h. This attenuation of endogenous H2S was associated with a significant increase in P2X7 receptor expression and NLRP3 inflammasome component levels. Administration of either NaHS or the cystathionine β-synthase agonist S-adenosyl-L-methionine one day after intracerebral haemorrhage resulted in a significant reduction in P2X7 receptor, NLRP3, and IL-1β levels (Figure 5). Interestingly, both endogenous and exogenous H2S have also been shown to suppress hyperglycaemia-induced TXNIP expression in a type-1 diabetic cardiomyopathy rat, resulting in a decreased response in NLRP3 complex activation (Figure 5) [236,237]. This inhibition of TXNIP expression had also been previously noted following administration of the other gaseous signalling molecule, CO, in lung tissue [138]. Huang et al. [238] had shown that pre-treatment with 400 μM NaHS offered protection to H9c2 cardiomyocytes subjected to high-glucose-induced cardiotoxicity. The addition of NaHS prevented high-glucose-induced expression of the NLRP3 inflammasome via direct suppression of the TLR4/NF-κB pathway [239,240] (Figure 5). This has been further demonstrated more recently in a mouse model of LPS-induced myocardial injury, whereby 50 μmol/kg of NaHS downregulated TLR4 and NLRP3 expressions by 64% and 31%, respectively, and enhanced ventricular function by 0.19-fold [241]. Other models have been used to show that H2S can also prevent oleic acid-induced suppression of the AMPK/mTOR pathway, promoting autophagy and inhibiting the NLRP3 inflammasome [242]. As previously stated, H2S has important interactions with Nrf2 to induce cytoprotective effects, and this interaction plays a key role in inhibiting NLRP3 inflammasome priming and activation [243,244]. This interaction between Nrf2 and H2S has been shown to attenuate renal IRI [245] and while this pathway has not been investigated in myocardial IRI, it can be hypothesised that the same protective response will occur.

5. Limitations of the Current Literature

Inflammation is often viewed as a double-edged sword, destructive if hyperactive, but favourable during low activity. Evidence of the key role played by the NLRP3 inflammasome in cardiac injury during reperfusion has been demonstrated in hearts from NLRP3/ASC double KO C57BL/6 mouse models subjected to ex vivo IRI [34]. In comparison to wildtype C57BL/6 mice, cardiac infarct size was significantly reduced in the knockouts by ~17% and haemodynamic contractile function was better preserved. Most of the current literature states that the presence of the NLRP3 inflammasome during cardiac ischaemia and reperfusion is detrimental and exacerbates the degree of tissue damage; however, there are reports indicating that the formation of the inflammasome can, in fact, be cardioprotective.
Zuurbier et al. [246] identified a cardioprotective role of NLRP3 in an NLRP3/ASC KO mouse model. This study indicated that without NLRP3, IL-6 levels decreased, resulting in a loss of the cardioprotective IL-6/signal transducer and activator of transcription 3 signalling pathway and hence a loss of cardioprotective ischaemic preconditioning. This study did not, however, directly measure infarct size, but relied on lactate dehydrogenase release as a marker of tissue damage. Controversially, Sandanger et al. [247] also suggested that NLRP3 may paradoxically be cardioprotective during in vivo cardiac IRI, a finding that appears to conflict with earlier findings by the same group in an ex vivo perfused heart model [34]. By conducting in vivo myocardial IRI in an NLRP3/ASC KO mouse model, Sandanger et al. [247] reported that cardiac infarct size based upon cardiac troponin measurements increased and RISK activation was lost in the knockout animals. There are several reservations, however, with both the KO model used and the study, chief of which is that IL-1β mRNA, rather than protein expression, was used to indicate inflammasome activation. Interestingly, Hollmann and Zuurbier [248] proposed that in the absence of surgical stressors such as tissue injury, there are inadequate resident levels of NLRP3 within the myocardium to actually contribute toward myocardial IRI. In their study, Hollmann and Zuurbier [248] used an in vivo closed-chest model with the ligature left on the left anterior descending coronary artery for 10 days in situ prior to inducing IRI in order to eliminate surgical stressors as a potential factor for inducing NLRP3 expression. The inclusion of an open-chest surgical model to directly compare against the closed-chest model in the same study identified increased expression of both cardioprotective IL-6 and TNF-α and the NLRP3 inflammasome.

6. Concluding Remarks

This review provides an overview of the signalling modulatory effects of the three gasotransmitters CO, NO, and H2S, focusing on the summary of published evidence supporting the value of these molecules as cardioprotective agents. The data gathered strongly suggests that all three gasotransmitters can exert a valuable anti-inflammatory effect by downregulating NLRP3 inflammasome activation in whole animal models. How effective NLRP3 inactivation remains following pharmacological intervention with any agent remains to be determined. These promising results still await validation in larger animal model studies. Currently, research in the field is lacking, and further work is required to identify precisely how these molecules, particularly CO and H2S, interact with intermediates involved in NLRP3 inflammasome activation to induce their cardioprotective effects. We have highlighted the limitations of the current data on the role of the NLRP3 inflammasome in myocardial IRI. However, the evidence predominantly supports that NLRP3 inflammasome activation has a detrimental role, further promoting the adverse events associated with myocardial IRI. Given the pleiotropic and largely beneficial cardioprotective effects reported in the literature, strategies to develop tunable gasotransmitter-releasing pharmacological agents should be pursued to maintain the NLRP3 inflammasome in its inactive form.

Author Contributions

Review, conception, design, figures, and writing, F.M.P.; review, figures, and writing, A.R.D.; editing, conception, and final approval of the manuscript, J.C.H. and I.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded in part by a Project research grant from the Health Research Council of New Zealand (HRC 20/274).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Doenst, T.; Haverich, A.; Serruys, P.; Bonow, R.O.; Kappetein, P.; Falk, V.; Velazquez, E.; Diegeler, A.; Sigusch, H. PCI and CABG for treating stable coronary artery disease: JACC review topic of the week. J. Am. Coll. Cardiol. 2019, 73, 964–976. [Google Scholar] [CrossRef]
  2. Ruggieri, V.G.; Bounader, K.; Verhoye, J.P.; Onorati, F.; Rubino, A.S.; Gatti, G.; Tauriainen, T.; De Feo, M.; Reichart, D.; Dalén, M. Prognostic impact of prolonged cross-clamp time in coronary artery bypass grafting. Heart Lung Circ. 2018, 27, 1476–1482. [Google Scholar] [CrossRef] [PubMed]
  3. Raja, S.G. Off-pump coronary artery bypass grafting in octogenarians. J. Thorac. Dis. 2016, 8, S799. [Google Scholar] [CrossRef] [PubMed]
  4. Dobson, G.P.; Faggian, G.; Onorati, F.; Vinten-Johansen, J. Hyperkalemic cardioplegia for adult and pediatric surgery: End of an era? Front. Physiol. 2013, 4, 228. [Google Scholar] [CrossRef]
  5. Suleiman, M.; Hancock, M.; Shukla, R.; Rajakaruna, C.; Angelini, G. Cardioplegic strategies to protect the hypertrophic heart during cardiac surgery. Perfusion 2011, 26, 48–56. [Google Scholar] [CrossRef]
  6. Wilson, N. Rheumatic heart disease in indigenous populations—New Zealand experience. Heart Lung Circ. 2010, 19, 282–288. [Google Scholar] [CrossRef]
  7. Hausenloy, D.J.; Garcia-Dorado, D.; Bøtker, H.E.; Davidson, S.M.; Downey, J.; Engel, F.B.; Jennings, R.; Lecour, S.; Leor, J.; Madonna, R. Novel targets and future strategies for acute cardioprotection: Position Paper of the European Society of Cardiology Working Group on Cellular Biology of the Heart. Cardiovasc. Res. 2017, 113, 564–585. [Google Scholar] [CrossRef] [PubMed]
  8. Pagliaro, P.; Penna, C. Hypertension, hypertrophy, and reperfusion injury. J. Cardiovasc. Med. 2017, 18, 131–135. [Google Scholar] [CrossRef]
  9. Janier, M.F.; Vanoverschelde, J.-L.; Bergmann, S.R. Ischemic preconditioning stimulates anaerobic glycolysis in the isolated rabbit heart. Am. J. Physiol. Heart Circ. Physiol. 1994, 267, H1353–H1360. [Google Scholar] [CrossRef]
  10. Jennings, R.B.; Reimer, K.A.; Hill, M.L.; Mayer, S.E. Total ischemia in dog hearts, in vitro. 1. Comparison of high energy phosphate production, utilization, and depletion, and of adenine nucleotide catabolism in total ischemia in vitro vs. severe ischemia in vivo. Circ. Res. 1981, 49, 892–900. [Google Scholar] [CrossRef]
  11. Cross, H.R.; Opie, L.H.; Radda, G.K.; Clarke, K. Is a high glycogen content beneficial or detrimental to the ischemic rat heart? A controversy resolved. Circ. Res. 1996, 78, 482–491. [Google Scholar] [CrossRef]
  12. Neely, J.R.; Grotyohann, L.W. Role of glycolytic products in damage to ischemic myocardium. Dissociation of adenosine triphosphate levels and recovery of function of reperfused ischemic hearts. Circ. Res. 1984, 55, 816–824. [Google Scholar] [CrossRef] [PubMed]
  13. Tani, M.; Neely, J.R. Role of intracellular Na+ in Ca2+ overload and depressed recovery of ventricular function of reperfused ischemic rat hearts. Possible involvement of H+-Na+ and Na+-Ca2+ exchange. Circ. Res. 1989, 65, 1045–1056. [Google Scholar] [CrossRef] [PubMed]
  14. Khandoudi, N.; Bernard, M.; Cozzone, P.; Feuvray, D. Intracellular pH and role of Na+/H+ exchange during ischaemia and reperfusion of normal and diabetic rat hearts. Cardiovasc. Res. 1990, 24, 873–878. [Google Scholar] [CrossRef] [PubMed]
  15. Kim, D.; Cragoe, E., Jr.; Smith, T. Relations among sodium pump inhibition, Na-Ca and Na-H exchange activities, and Ca-H interaction in cultured chick heart cells. Circ. Res. 1987, 60, 185–193. [Google Scholar] [CrossRef] [PubMed]
  16. Murphy, E.; Perlman, M.; London, R.E.; Steenbergen, C. Amiloride delays the ischemia-induced rise in cytosolic free calcium. Circ. Res. 1991, 68, 1250–1258. [Google Scholar] [CrossRef]
  17. Bigger, J.T., Jr.; Fleiss, J.L.; Kleiger, R.; Miller, J.P.; Rolnitzky, L. The relationships among ventricular arrhythmias, left ventricular dysfunction, and mortality in the 2 years after myocardial infarction. Circulation 1984, 69, 250–258. [Google Scholar] [CrossRef]
  18. Tani, M.; Asakura, Y.; Hasegawa, H.; Shinmura, K.; Ebihara, Y.; Nakamura, Y. Effect of preconditioning on ryanodine-sensitive Ca2+ release from sarcoplasmic reticulum of rat heart. Am. J. Physiol. Heart Circ. Physiol. 1996, 271, H876–H881. [Google Scholar] [CrossRef]
  19. Correa, F.; Soto, V.; Zazueta, C. Mitochondrial permeability transition relevance for apoptotic triggering in the post-ischemic heart. Int. J. Biochem. Cell Biol. 2007, 39, 787–798. [Google Scholar] [CrossRef]
  20. Griffiths, E.J.; Halestrap, A.P. Further evidence that cyclosporin A protects mitochondria from calcium overload by inhibiting a matrix peptidyl-prolyl cis-trans isomerase. Implications for the immunosuppressive and toxic effects of cyclosporin. Biochem. J. 1991, 274, 611–614. [Google Scholar] [CrossRef]
  21. Jennings, R.; Hawkins, H.K.; Lowe, J.E.; Hill, M.L.; Klotman, S.; Reimer, K.A. Relation between high energy phosphate and lethal injury in myocardial ischemia in the dog. Am. J. Pathol. 1978, 92, 187. [Google Scholar]
  22. Singh, R.B.; Chohan, P.K.; Dhalla, N.S.; Netticadan, T. The sarcoplasmic reticulum proteins are targets for calpain action in the ischemic–reperfused heart. J. Mol. Cell. Cardiol. 2004, 37, 101–110. [Google Scholar] [CrossRef]
  23. Lindsay, D.P.; Camara, A.K.; Stowe, D.F.; Lubbe, R.; Aldakkak, M. Differential effects of buffer pH on Ca2+-induced ROS emission with inhibited mitochondrial complexes I and III. Front. Physiol. 2015, 6, 58. [Google Scholar] [CrossRef]
  24. Seidlmayer, L.K.; Juettner, V.V.; Kettlewell, S.; Pavlov, E.V.; Blatter, L.A.; Dedkova, E.N. Distinct mPTP activation mechanisms in ischaemia–reperfusion: Contributions of Ca2+, ROS, pH, and inorganic polyphosphate. Cardiovasc. Res. 2015, 106, 237–248. [Google Scholar] [CrossRef]
  25. Green, D.R.; Reed, J.C. Mitochondria and apoptosis. Science 1998, 281, 1309–1312. [Google Scholar] [CrossRef]
  26. Kim, J.-S.; Jin, Y.; Lemasters, J.J. Reactive oxygen species, but not Ca2+ overloading, trigger pH-and mitochondrial permeability transition-dependent death of adult rat myocytes after ischemia-reperfusion. Am. J. Physiol. Heart Circ. Physiol. 2006, 290, H2024–H2034. [Google Scholar] [CrossRef]
  27. Xu, T.; Ding, W.; Ao, X.; Chu, X.; Wan, Q.; Wang, Y.; Xiao, D.; Yu, W.; Li, M.; Yu, F. ARC regulates programmed necrosis and myocardial ischemia/reperfusion injury through the inhibition of mPTP opening. Redox Biol. 2019, 20, 414–426. [Google Scholar] [CrossRef]
  28. Zhu, P.; Hu, S.; Jin, Q.; Li, D.; Tian, F.; Toan, S.; Li, Y.; Zhou, H.; Chen, Y. Ripk3 promotes ER stress-induced necroptosis in cardiac IR injury: A mechanism involving calcium overload/XO/ROS/mPTP pathway. Redox Biol. 2018, 16, 157–168. [Google Scholar] [CrossRef]
  29. Becker, R.C.; Owens, A.P., III; Sadayappan, S. Tissue-level inflammation and ventricular remodeling in hypertrophic cardiomyopathy. J. Thromb. Thrombolysis 2020, 49, 177–183. [Google Scholar] [CrossRef]
  30. Rai, V.; Sharma, P.; Agrawal, S.; Agrawal, D.K. Relevance of mouse models of cardiac fibrosis and hypertrophy in cardiac research. Mol. Cell. Biochem. 2017, 424, 123–145. [Google Scholar] [CrossRef]
  31. Hirai, S. Systemic inflammatory response syndrome after cardiac surgery under cardiopulmonary bypass. Ann. Thorac. Cardiovasc. Surg. 2003, 9, 365–370. [Google Scholar]
  32. Kawaguchi, M.; Takahashi, M.; Hata, T.; Kashima, Y.; Usui, F.; Morimoto, H.; Izawa, A.; Takahashi, Y.; Masumoto, J.; Koyama, J. Inflammasome activation of cardiac fibroblasts is essential for myocardial ischemia/reperfusion injury. Circulation 2011, 123, 594–604. [Google Scholar] [CrossRef]
  33. Frangogiannis, N.G. The inflammatory response in myocardial injury, repair, and remodelling. Nat. Rev. Cardiol. 2014, 11, 255–265. [Google Scholar] [CrossRef]
  34. Sandanger, Ø.; Ranheim, T.; Vinge, L.E.; Bliksøen, M.; Alfsnes, K.; Finsen, A.V.; Dahl, C.P.; Askevold, E.T.; Florholmen, G.; Christensen, G. The NLRP3 inflammasome is up-regulated in cardiac fibroblasts and mediates myocardial ischaemia–reperfusion injury. Cardiovasc. Res. 2013, 99, 164–174. [Google Scholar] [CrossRef]
  35. Westermann, D.; Lindner, D.; Kasner, M.; Zietsch, C.; Savvatis, K.; Escher, F.; Von Schlippenbach, J.; Skurk, C.; Steendijk, P.; Riad, A. Cardiac inflammation contributes to changes in the extracellular matrix in patients with heart failure and normal ejection fraction. Circ. Heart Fail. 2011, 4, 44–52. [Google Scholar] [CrossRef]
  36. Cao, L.; Chen, Y.; Zhang, Z.; Li, Y.; Zhao, P. Endoplasmic reticulum stress-induced NLRP1 inflammasome activation contributes to myocardial ischemia/reperfusion injury. Shock 2019, 51, 511–518. [Google Scholar] [CrossRef]
  37. Mezzaroma, E.; Toldo, S.; Farkas, D.; Seropian, I.M.; Van Tassell, B.W.; Salloum, F.N.; Kannan, H.R.; Menna, A.C.; Voelkel, N.F.; Abbate, A. The inflammasome promotes adverse cardiac remodeling following acute myocardial infarction in the mouse. Proc. Natl. Acad. Sci. USA 2011, 108, 19725–19730. [Google Scholar] [CrossRef]
  38. Toldo, S.; Marchetti, C.; Mauro, A.G.; Chojnacki, J.; Mezzaroma, E.; Carbone, S.; Zhang, S.; Van Tassell, B.; Salloum, F.N.; Abbate, A. Inhibition of the NLRP3 inflammasome limits the inflammatory injury following myocardial ischemia–reperfusion in the mouse. Int. J. Cardiol. 2016, 209, 215–220. [Google Scholar] [CrossRef]
  39. Bracey, N.A.; Beck, P.L.; Muruve, D.A.; Hirota, S.A.; Guo, J.; Jabagi, H.; Wright, J.R., Jr.; MacDonald, J.A.; Lees-Miller, J.P.; Roach, D. The Nlrp3 inflammasome promotes myocardial dysfunction in structural cardiomyopathy through interleukin-1β. Exp. Physiol. 2013, 98, 462–472. [Google Scholar] [CrossRef]
  40. Willingham, S.B.; Allen, I.C.; Bergstralh, D.T.; Brickey, W.J.; Huang, M.T.-H.; Taxman, D.J.; Duncan, J.A.; Ting, J.P.-Y. NLRP3 (NALP3, Cryopyrin) facilitates in vivo caspase-1 activation, necrosis, and HMGB1 release via inflammasome-dependent and-independent pathways. J. Immunol. 2009, 183, 2008–2015. [Google Scholar] [CrossRef]
  41. Hoffman, H.M.; Mueller, J.L.; Broide, D.H.; Wanderer, A.A.; Kolodner, R.D. Mutation of a new gene encoding a putative pyrin-like protein causes familial cold autoinflammatory syndrome and Muckle–Wells syndrome. Nat. Genet. 2001, 29, 301–305. [Google Scholar] [CrossRef]
  42. Vajjhala, P.R.; Mirams, R.E.; Hill, J.M. Multiple binding sites on the pyrin domain of ASC protein allow self-association and interaction with NLRP3 protein. J. Biol. Chem. 2012, 287, 41732–41743. [Google Scholar] [CrossRef]
  43. Proell, M.; Gerlic, M.; Mace, P.D.; Reed, J.C.; Riedl, S.J. The CARD plays a critical role in ASC foci formation and inflammasome signalling. Biochem. J. 2013, 449, 613–621. [Google Scholar] [CrossRef]
  44. Shi, H.; Wang, Y.; Li, X.; Zhan, X.; Tang, M.; Fina, M.; Su, L.; Pratt, D.; Bu, C.H.; Hildebrand, S. NLRP3 activation and mitosis are mutually exclusive events coordinated by NEK7, a new inflammasome component. Nat. Immunol. 2016, 17, 250–258. [Google Scholar] [CrossRef]
  45. Xiao, L.; Magupalli, V.G.; Wu, H. Cryo-EM structures of the active NLRP3 inflammasome disc. Nature 2023, 613, 595–600. [Google Scholar] [CrossRef]
  46. Sutterwala, F.S.; Ogura, Y.; Szczepanik, M.; Lara-Tejero, M.; Lichtenberger, G.S.; Grant, E.P.; Bertin, J.; Coyle, A.J.; Galán, J.E.; Askenase, P.W. Critical role for NALP3/CIAS1/Cryopyrin in innate and adaptive immunity through its regulation of caspase-1. Immunity 2006, 24, 317–327. [Google Scholar] [CrossRef]
  47. Toldo, S.; Mezzaroma, E.; McGeough, M.D.; Peña, C.A.; Marchetti, C.; Sonnino, C.; Van Tassell, B.W.; Salloum, F.N.; Voelkel, N.F.; Hoffman, H.M. Independent roles of the priming and the triggering of the NLRP3 inflammasome in the heart. Cardiovasc. Res. 2015, 105, 203–212. [Google Scholar] [CrossRef]
  48. Juliana, C.; Fernandes-Alnemri, T.; Kang, S.; Farias, A.; Qin, F.; Alnemri, E.S. Non-transcriptional priming and deubiquitination regulate NLRP3 inflammasome activation. J. Biol. Chem. 2012, 287, 36617–36622. [Google Scholar] [CrossRef]
  49. Bauernfeind, F.G.; Horvath, G.; Stutz, A.; Alnemri, E.S.; MacDonald, K.; Speert, D.; Fernandes-Alnemri, T.; Wu, J.; Monks, B.G.; Fitzgerald, K.A. Cutting edge: NF-κB activating pattern recognition and cytokine receptors license NLRP3 inflammasome activation by regulating NLRP3 expression. J. Immunol. 2009, 183, 787–791. [Google Scholar] [CrossRef]
  50. Py, B.F.; Kim, M.-S.; Vakifahmetoglu-Norberg, H.; Yuan, J. Deubiquitination of NLRP3 by BRCC3 critically regulates inflammasome activity. Mol. Cell 2013, 49, 331–338. [Google Scholar] [CrossRef]
  51. Mariathasan, S.; Weiss, D.S.; Newton, K.; McBride, J.; O’Rourke, K.; Roose-Girma, M.; Lee, W.P.; Weinrauch, Y.; Monack, D.M.; Dixit, V.M. Cryopyrin activates the inflammasome in response to toxins and ATP. Nature 2006, 440, 228–232. [Google Scholar] [CrossRef]
  52. Zhou, R.; Yazdi, A.S.; Menu, P.; Tschopp, J. A role for mitochondria in NLRP3 inflammasome activation. Nature 2011, 469, 221–225. [Google Scholar] [CrossRef]
  53. Franchi, L.; Eigenbrod, T.; Núñez, G. Cutting edge: TNF-α mediates sensitization to ATP and silica via the NLRP3 inflammasome in the absence of microbial stimulation. J. Immunol. 2009, 183, 792–796. [Google Scholar] [CrossRef]
  54. Qiao, Y.; Wang, P.; Qi, J.; Zhang, L.; Gao, C. TLR-induced NF-κB activation regulates NLRP3 expression in murine macrophages. FEBS Lett. 2012, 586, 1022–1026. [Google Scholar] [CrossRef]
  55. Ermolaeva, M.A.; Michallet, M.-C.; Papadopoulou, N.; Utermöhlen, O.; Kranidioti, K.; Kollias, G.; Tschopp, J.; Pasparakis, M. Function of TRADD in tumor necrosis factor receptor 1 signaling and in TRIF-dependent inflammatory responses. Nat. Immunol. 2008, 9, 1037–1046. [Google Scholar] [CrossRef]
  56. Zhang, J.; Xiao, F.; Zhang, L.; Wang, X.; Lai, X.; Shen, Y.; Zhang, M.; Zhou, B.; Lang, H.; Yu, P. Alpha-lipoic acid preconditioning and ischaemic postconditioning synergistically protect rats from cerebral injury induced by ischemia and reperfusion partly via inhibition TLR4/MyD88/NF-κB signaling pathway. Cell. Physiol. Biochem. 2018, 51, 1448–1460. [Google Scholar] [CrossRef]
  57. Bauernfeind, F.; Bartok, E.; Rieger, A.; Franchi, L.; Núñez, G.; Hornung, V. Cutting edge: Reactive oxygen species inhibitors block priming, but not activation, of the NLRP3 inflammasome. J. Immunol. 2011, 187, 613–617. [Google Scholar] [CrossRef]
  58. Gurung, P.; Anand, P.K.; Malireddi, R.S.; Walle, L.V.; Van Opdenbosch, N.; Dillon, C.P.; Weinlich, R.; Green, D.R.; Lamkanfi, M.; Kanneganti, T.-D. FADD and caspase-8 mediate priming and activation of the canonical and noncanonical Nlrp3 inflammasomes. J. Immunol. 2014, 192, 1835–1846. [Google Scholar] [CrossRef]
  59. Zhang, X.; Fan, C.; Zhang, H.; Zhao, Q.; Liu, Y.; Xu, C.; Xie, Q.; Wu, X.; Yu, X.; Zhang, J. MLKL and FADD are critical for suppressing progressive lymphoproliferative disease and activating the NLRP3 inflammasome. Cell Rep. 2016, 16, 3247–3259. [Google Scholar] [CrossRef] [PubMed]
  60. Allam, R.; Lawlor, K.E.; Yu, E.C.W.; Mildenhall, A.L.; Moujalled, D.M.; Lewis, R.S.; Ke, F.; Mason, K.D.; White, M.J.; Stacey, K.J. Mitochondrial apoptosis is dispensable for NLRP 3 inflammasome activation but non-apoptotic caspase-8 is required for inflammasome priming. EMBO Rep. 2014, 15, 982–990. [Google Scholar] [CrossRef]
  61. Baudino, T.A.; Carver, W.; Giles, W.; Borg, T.K. Cardiac fibroblasts: Friend or foe? Am. J. Physiol. Heart Circ. Physiol. 2006, 291, H1015–H1026. [Google Scholar] [CrossRef]
  62. Liu, Z.; Gan, L.; Xu, Y.; Luo, D.; Ren, Q.; Wu, S.; Sun, C. Melatonin alleviates inflammasome-induced pyroptosis through inhibiting NF-κB/GSDMD signal in mice adipose tissue. J. Pineal Res. 2017, 63, e12414. [Google Scholar] [CrossRef]
  63. Fernandes-Alnemri, T.; Kang, S.; Anderson, C.; Sagara, J.; Fitzgerald, K.A.; Alnemri, E.S. Cutting edge: TLR signaling licenses IRAK1 for rapid activation of the NLRP3 inflammasome. J. Immunol. 2013, 191, 3995–3999. [Google Scholar] [CrossRef]
  64. Jha, A.; Kumar, V.; Haque, S.; Ayasolla, K.; Saha, S.; Lan, X.; Malhotra, A.; Saleem, M.A.; Skorecki, K.; Singhal, P.C. Alterations in plasma membrane ion channel structures stimulate NLRP3 inflammasome activation in APOL1 risk milieu. FEBS J. 2020, 287, 2000–2022. [Google Scholar] [CrossRef]
  65. Shen, S.; He, F.; Cheng, C.; Xu, B.; Sheng, J. Uric acid aggravates myocardial ischemia–reperfusion injury via ROS/NLRP3 pyroptosis pathway. Biomed. Pharmacother. 2021, 133, 110990. [Google Scholar] [CrossRef]
  66. Chen, Y.; Li, X.; Boini, K.M.; Pitzer, A.L.; Gulbins, E.; Zhang, Y.; Li, P.-L. Endothelial Nlrp3 inflammasome activation associated with lysosomal destabilization during coronary arteritis. Biochim. Biophys. Acta Mol. Cell Res. 2015, 1853, 396–408. [Google Scholar] [CrossRef] [PubMed]
  67. Yellon, D.M.; Hausenloy, D.J. Myocardial reperfusion injury. N. Engl. J. Med. 2007, 357, 1121–1135. [Google Scholar] [CrossRef]
  68. He, Y.; Zeng, M.Y.; Yang, D.; Motro, B.; Núñez, G. NEK7 is an essential mediator of NLRP3 activation downstream of potassium efflux. Nature 2016, 530, 354–357. [Google Scholar] [CrossRef]
  69. Walev, I.; Klein, J.; Husmann, M.; Valeva, A.; Strauch, S.; Wirtz, H.; Weichel, O.; Bhakdi, S. Potassium regulates IL-1β processing via calcium-independent phospholipase A2. J. Immunol. 2000, 164, 5120–5124. [Google Scholar] [CrossRef]
  70. Katsnelson, M.A.; Rucker, L.G.; Russo, H.M.; Dubyak, G.R. K+ efflux agonists induce NLRP3 inflammasome activation independently of Ca2+ signaling. J. Immunol. 2015, 194, 3937–3952. [Google Scholar] [CrossRef]
  71. Rossol, M.; Pierer, M.; Raulien, N.; Quandt, D.; Meusch, U.; Rothe, K.; Schubert, K.; Schöneberg, T.; Schaefer, M.; Krügel, U. Extracellular Ca2+ is a danger signal activating the NLRP3 inflammasome through G protein-coupled calcium sensing receptors. Nat. Commun. 2012, 3, 1329. [Google Scholar] [CrossRef] [PubMed]
  72. Venkatachalam, K.; Prabhu, S.D.; Reddy, V.S.; Boylston, W.H.; Valente, A.J.; Chandrasekar, B. Neutralization of interleukin-18 ameliorates ischemia/reperfusion-induced myocardial injury. J. Biol. Chem. 2009, 284, 7853–7865. [Google Scholar] [CrossRef]
  73. Ding, H.-S.; Yang, J.; Chen, P.; Yang, J.; Bo, S.-Q.; Ding, J.-W.; Yu, Q.-Q. The HMGB1–TLR4 axis contributes to myocardial ischemia/reperfusion injury via regulation of cardiomyocyte apoptosis. Gene 2013, 527, 389–393. [Google Scholar] [CrossRef]
  74. Zhong, Z.; Liang, S.; Sanchez-Lopez, E.; He, F.; Shalapour, S.; Lin, X.-j.; Wong, J.; Ding, S.; Seki, E.; Schnabl, B. New mitochondrial DNA synthesis enables NLRP3 inflammasome activation. Nature 2018, 560, 198–203. [Google Scholar] [CrossRef] [PubMed]
  75. Billingham, L.K.; Stoolman, J.S.; Vasan, K.; Rodriguez, A.E.; Poor, T.A.; Szibor, M.; Jacobs, H.T.; Reczek, C.R.; Rashidi, A.; Zhang, P. Mitochondrial electron transport chain is necessary for NLRP3 inflammasome activation. Nat. Immunol. 2022, 23, 692–704. [Google Scholar] [CrossRef]
  76. Xian, H.; Watari, K.; Sanchez-Lopez, E.; Offenberger, J.; Onyuru, J.; Sampath, H.; Ying, W.; Hoffman, H.M.; Shadel, G.S.; Karin, M. Oxidized DNA fragments exit mitochondria via mPTP-and VDAC-dependent channels to activate NLRP3 inflammasome and interferon signaling. Immunity 2022, 55, 1370–1385.e8. [Google Scholar] [CrossRef]
  77. Cabral, A.; Cabral, J.E.; Wang, A.; Zhang, Y.; Liang, H.; Nikbakht, D.; Corona, L.; Hoffman, H.M.; McNulty, R. Differential Binding of NLRP3 to non-oxidized and Ox-mtDNA mediates NLRP3 Inflammasome Activation. Commun. Biol. 2023, 6, 578. [Google Scholar]
  78. Chevriaux, A.; Pilot, T.; Derangère, V.; Simonin, H.; Martine, P.; Chalmin, F.; Ghiringhelli, F.; Rébé, C. Cathepsin B is required for NLRP3 inflammasome activation in macrophages, through NLRP3 interaction. Front. Cell Dev. Biol. 2020, 8, 167. [Google Scholar] [CrossRef]
  79. Bai, H.; Yang, B.; Yu, W.; Xiao, Y.; Yu, D.; Zhang, Q. Cathepsin B links oxidative stress to the activation of NLRP3 inflammasome. Exp. Cell Res. 2018, 362, 180–187. [Google Scholar] [CrossRef] [PubMed]
  80. Alizadeh, P.; Smit-McBride, Z.; Oltjen, S.L.; Hjelmeland, L.M. Regulation of cysteine cathepsin expression by oxidative stress in the retinal pigment epithelium/choroid of the mouse. Exp. Eye Res. 2006, 83, 679–687. [Google Scholar] [CrossRef]
  81. Jha, S.; Srivastava, S.Y.; Brickey, W.J.; Iocca, H.; Toews, A.; Morrison, J.P.; Chen, V.S.; Gris, D.; Matsushima, G.K.; Ting, J.P.-Y. The inflammasome sensor, NLRP3, regulates CNS inflammation and demyelination via caspase-1 and interleukin-18. J. Neurosci. 2010, 30, 15811–15820. [Google Scholar] [CrossRef] [PubMed]
  82. Thornberry, N.A.; Bull, H.G.; Calaycay, J.R.; Chapman, K.T.; Howard, A.D.; Kostura, M.J.; Miller, D.K.; Molineaux, S.M.; Weidner, J.R.; Aunins, J. A novel heterodimeric cysteine protease is required for interleukin-1βprocessing in monocytes. Nature 1992, 356, 768–774. [Google Scholar] [CrossRef] [PubMed]
  83. Shi, J.; Zhao, Y.; Wang, K.; Shi, X.; Wang, Y.; Huang, H.; Zhuang, Y.; Cai, T.; Wang, F.; Shao, F. Cleavage of GSDMD by inflammatory caspases determines pyroptotic cell death. Nature 2015, 526, 660–665. [Google Scholar] [CrossRef] [PubMed]
  84. Ding, J.; Wang, K.; Liu, W.; She, Y.; Sun, Q.; Shi, J.; Sun, H.; Wang, D.-C.; Shao, F. Pore-forming activity and structural autoinhibition of the gasdermin family. Nature 2016, 535, 111–116. [Google Scholar] [CrossRef]
  85. Sborgi, L.; Rühl, S.; Mulvihill, E.; Pipercevic, J.; Heilig, R.; Stahlberg, H.; Farady, C.J.; Müller, D.J.; Broz, P.; Hiller, S. GSDMD membrane pore formation constitutes the mechanism of pyroptotic cell death. EMBO J. 2016, 35, 1766–1778. [Google Scholar] [CrossRef]
  86. Lei, Q.; Yi, T.; Chen, C. NF-κB-Gasdermin D (GSDMD) axis couples oxidative stress and NACHT, LRR and PYD domains-containing protein 3 (NLRP3) inflammasome-mediated cardiomyocyte pyroptosis following myocardial infarction. Med. Sci. Monit. 2018, 24, 6044. [Google Scholar] [CrossRef]
  87. Fernandes-Alnemri, T.; Wu, J.; Yu, J.; Datta, P.; Miller, B.; Jankowski, W.; Rosenberg, S.; Zhang, J.; Alnemri, E. The pyroptosome: A supramolecular assembly of ASC dimers mediating inflammatory cell death via caspase-1 activation. Cell Death Differ. 2007, 14, 1590–1604. [Google Scholar] [CrossRef]
  88. Coll, R.C.; Robertson, A.A.; Chae, J.J.; Higgins, S.C.; Muñoz-Planillo, R.; Inserra, M.C.; Vetter, I.; Dungan, L.S.; Monks, B.G.; Stutz, A. A small-molecule inhibitor of the NLRP3 inflammasome for the treatment of inflammatory diseases. Nat. Med. 2015, 21, 248–255. [Google Scholar] [CrossRef]
  89. Coll, R.C.; Hill, J.R.; Day, C.J.; Zamoshnikova, A.; Boucher, D.; Massey, N.L.; Chitty, J.L.; Fraser, J.A.; Jennings, M.P.; Robertson, A.A. MCC950 directly targets the NLRP3 ATP-hydrolysis motif for inflammasome inhibition. Nat. Chem. Biol. 2019, 15, 556–559. [Google Scholar] [CrossRef]
  90. Tapia-Abellán, A.; Angosto-Bazarra, D.; Martínez-Banaclocha, H.; de Torre-Minguela, C.; Cerón-Carrasco, J.P.; Pérez-Sánchez, H.; Arostegui, J.I.; Pelegrin, P. MCC950 closes the active conformation of NLRP3 to an inactive state. Nat. Chem. Biol. 2019, 15, 560–564. [Google Scholar] [CrossRef]
  91. Gao, R.; Shi, H.; Chang, S.; Gao, Y.; Li, X.; Lv, C.; Yang, H.; Xiang, H.; Yang, J.; Xu, L. The selective NLRP3-inflammasome inhibitor MCC950 reduces myocardial fibrosis and improves cardiac remodeling in a mouse model of myocardial infarction. Int. Immunopharmacol. 2019, 74, 105575. [Google Scholar] [CrossRef]
  92. Marchetti, C.; Swartzwelter, B.; Gamboni, F.; Neff, C.P.; Richter, K.; Azam, T.; Carta, S.; Tengesdal, I.; Nemkov, T.; D’Alessandro, A. OLT1177, a β-sulfonyl nitrile compound, safe in humans, inhibits the NLRP3 inflammasome and reverses the metabolic cost of inflammation. Proc. Natl. Acad. Sci. USA 2018, 115, E1530–E1539. [Google Scholar] [CrossRef] [PubMed]
  93. Toldo, S.; Mauro, A.G.; Cutter, Z.; Van Tassell, B.W.; Mezzaroma, E.; Del Buono, M.G.; Prestamburgo, A.; Potere, N.; Abbate, A. The NLRP3 inflammasome inhibitor, OLT1177 (dapansutrile), reduces infarct size and preserves contractile function after ischemia reperfusion injury in the mouse. J. Cardiovasc. Pharmacol. 2019, 73, 215–222. [Google Scholar] [CrossRef] [PubMed]
  94. Wohlford, G.F.; Van Tassell, B.W.; Billingsley, H.E.; Kadariya, D.; Carbone, S.; Mihalick, V.L.; Bonaventura, A.; Vecchié, A.; Chiabrando, J.G.; Bressi, E. Phase 1b, randomized, double-blinded, dose escalation, single-center, repeat dose safety and pharmacodynamics study of the oral nlrp3 inhibitor dapansutrile in subjects with nyha ii–iii systolic heart failure. J. Cardiovasc. Pharmacol. 2021, 77, 49–60. [Google Scholar] [CrossRef]
  95. Gastaldi, S.; Rocca, C.; Gianquinto, E.; Granieri, M.C.; Boscaro, V.; Blua, F.; Rolando, B.; Marini, E.; Gallicchio, M.; De Bartolo, A. Discovery of a novel 1, 3, 4-oxadiazol-2-one-based NLRP3 inhibitor as a pharmacological agent to mitigate cardiac and metabolic complications in an experimental model of diet-induced metaflammation. Eur. J. Med. Chem. 2023, 257, 115542. [Google Scholar] [CrossRef] [PubMed]
  96. Bertinaria, M.; Gastaldi, S.; Marini, E.; Giorgis, M. Development of covalent NLRP3 inflammasome inhibitors: Chemistry and biological activity. Arch. Biochem. Biophys. 2019, 670, 116–139. [Google Scholar] [CrossRef] [PubMed]
  97. Mastrocola, R.; Penna, C.; Tullio, F.; Femminò, S.; Nigro, D.; Chiazza, F.; Serpe, L.; Collotta, D.; Alloatti, G.; Cocco, M. Pharmacological inhibition of NLRP3 inflammasome attenuates myocardial ischemia/reperfusion injury by activation of RISK and mitochondrial pathways. Oxid. Med. Cell. Longev. 2016, 2016, 5271251. [Google Scholar] [CrossRef]
  98. Penna, C.; Aragno, M.; Cento, A.S.; Femminò, S.; Russo, I.; Bello, F.D.; Chiazza, F.; Collotta, D.; Alves, G.F.; Bertinaria, M. Ticagrelor conditioning effects are not additive to cardioprotection induced by direct NLRP3 inflammasome inhibition: Role of RISK, NLRP3, and Redox Cascades. Oxid. Med. Cell. Longev. 2020, 2020, 9219825. [Google Scholar] [CrossRef]
  99. Allen, T.A.; Root, W.S. Partition of carbon monoxide and oxygen between air and whole blood of rats, dogs and men as affected by plasma pH. J. Appl. Physiol. 1957, 10, 186–190. [Google Scholar] [CrossRef]
  100. Sjöstrand, T. Endogenous formation of carbon monoxide in man under normal and pathological conditions. Scand. J. Clin. Investig. 1949, 1, 201–214. [Google Scholar] [CrossRef]
  101. Tenhunen, R.; Marver, H.S.; Schmid, R. Microsomal heme oxygenase: Characterization of the enzyme. J. Biol. Chem. 1969, 244, 6388–6394. [Google Scholar] [CrossRef] [PubMed]
  102. Maines, M.D.; Trakshel, G.; Kutty, R.K. Characterization of two constitutive forms of rat liver microsomal heme oxygenase. Only one molecular species of the enzyme is inducible. J. Biol. Chem. 1986, 261, 411–419. [Google Scholar] [CrossRef] [PubMed]
  103. Mccoubrey, W.K., Jr.; Huang, T.; Maines, M.D. Isolation and characterization of a cDNA from the rat brain that encodes hemoprotein heme oxygenase-3. Eur. J. Biochem. 1997, 247, 725–732. [Google Scholar] [CrossRef]
  104. Zakhary, R.; Gaine, S.P.; Dinerman, J.L.; Ruat, M.; Flavahan, N.A.; Snyder, S.H. Heme oxygenase 2: Endothelial and neuronal localization and role in endothelium-dependent relaxation. Proc. Natl. Acad. Sci. USA 1996, 93, 795–798. [Google Scholar] [CrossRef]
  105. Jancsó, G.; Cserepes, B.; Gasz, B.; Benko, L.; Borsiczky, B.; Ferenc, A.; Kürthy, M.; Rácz, B.; Lantos, J.; Gál, J. Expression and Protective Role of Heme Oxygenase-1 in Delayed Myocardial Preconditioning. Ann. N. Y. Acad. Sci. 2007, 1095, 251–261. [Google Scholar] [CrossRef] [PubMed]
  106. Yet, S.-F.; Tian, R.; Layne, M.D.; Wang, Z.Y.; Maemura, K.; Solovyeva, M.; Ith, B.; Melo, L.G.; Zhang, L.; Ingwall, J.S. Cardiac-specific expression of heme oxygenase-1 protects against ischemia and reperfusion injury in transgenic mice. Circ. Res. 2001, 89, 168–173. [Google Scholar] [CrossRef]
  107. Nakao, A.; Neto, J.S.; Kanno, S.; Stolz, D.B.; Kimizuka, K.; Liu, F.; Bach, F.H.; Billiar, T.R.; Choi, A.M.; Otterbein, L.E. Protection against ischemia/reperfusion injury in cardiac and renal transplantation with carbon monoxide, biliverdin and both. Am. J. Transplant. 2005, 5, 282–291. [Google Scholar] [CrossRef]
  108. Segersvärd, H.; Lakkisto, P.; Hänninen, M.; Forsten, H.; Siren, J.; Immonen, K.; Kosonen, R.; Sarparanta, M.; Laine, M.; Tikkanen, I. Carbon monoxide releasing molecule improves structural and functional cardiac recovery after myocardial injury. Eur. J. Pharmacol. 2018, 818, 57–66. [Google Scholar] [CrossRef]
  109. Goldbaum, L.; Orellano, T.; Dergal, E. Mechanism of the toxic action of carbon monoxide. Ann. Clin. Lab. Sci. 1976, 6, 372–376. [Google Scholar]
  110. Stupfel, M.; Bouley, G. Physiological and biochemical effects on rats and mice exposed to small concentrations of carbon monoxide for long periods. Ann. N. Y. Acad. Sci. 1970, 174, 342–368. [Google Scholar] [CrossRef]
  111. Kueh, J.T.B.; Stanley, N.J.; Hewitt, R.J.; Woods, L.M.; Larsen, L.; Harrison, J.C.; Rennison, D.; Brimble, M.A.; Sammut, I.A.; Larsen, D.S. Norborn-2-en-7-ones as physiologically-triggered carbon monoxide-releasing prodrugs. Chem. Sci. 2017, 8, 5454. [Google Scholar] [CrossRef] [PubMed]
  112. Motterlini, R.; Clark, J.E.; Foresti, R.; Sarathchandra, P.; Mann, B.E.; Green, C.J. Carbon monoxide-releasing molecules: Characterization of biochemical and vascular activities. Circ. Res. 2002, 90, e17–e24. [Google Scholar] [CrossRef]
  113. Motterlini, R.; Sawle, P.; Bains, S.; Hammad, J.; Alberto, R.; Foresti, R.; Green, C.J. CORM-A1: A new pharmacologically active carbon monoxide-releasing molecule. FASEB J. 2005, 19, 284–286. [Google Scholar] [CrossRef]
  114. Clark, J.E.; Naughton, P.; Shurey, S.; Green, C.J.; Johnson, T.R.; Mann, B.E.; Foresti, R.; Motterlini, R. Cardioprotective actions by a water-soluble carbon monoxide–releasing molecule. Circ. Res. 2003, 93, e2–e8. [Google Scholar] [CrossRef] [PubMed]
  115. Portal, L.; Morin, D.; Motterlini, R.; Ghaleh, B.; Pons, S. The CO-releasing molecule CORM-3 protects adult cardiomyocytes against hypoxia-reoxygenation by modulating pH restoration. Eur. J. Pharmacol. 2019, 862, 172636. [Google Scholar] [CrossRef]
  116. Soni, H.; Patel, P.; Rath, A.C.; Jain, M.; Mehta, A.A. Cardioprotective effect with carbon monoxide releasing molecule-2 (CORM-2) in isolated perfused rat heart: Role of coronary endothelium and underlying mechanism. Vasc. Pharmacol. 2010, 53, 68–76. [Google Scholar] [CrossRef] [PubMed]
  117. Winburn, I.C.; Gunatunga, K.; McKernan, R.D.; Walker, R.J.; Sammut, I.A.; Harrison, J.C. Cell damage following carbon monoxide releasing molecule exposure: Implications for therapeutic applications. Basic Clin. Pharmacol. Toxicol. 2012, 111, 31–41. [Google Scholar] [CrossRef]
  118. Desmard, M.; Davidge, K.S.; Bouvet, O.; Morin, D.; Roux, D.; Foresti, R.; Ricard, J.D.; Denamur, E.; Poole, R.K.; Montravers, P. A carbon monoxide-releasing molecule (CORM-3) exerts bactericidal activity against Pseudomonas aeruginosa and improves survival in an animal model of bacteraemia. FASEB J. 2009, 23, 1023–1031. [Google Scholar] [CrossRef]
  119. Bauer, N.; Yang, X.; Yuan, Z.; Wang, B. Reassessing CORM-A1: Redox chemistry and idiosyncratic CO-releasing characteristics of the widely used carbon monoxide donor. Chem. Sci. 2023, 14, 3215–3228. [Google Scholar] [CrossRef]
  120. Bauer, N.; Yuan, Z.; Yang, X.; Wang, B. Plight of CORMs: The unreliability of four commercially available CO-releasing molecules, CORM-2, CORM-3, CORM-A1, and CORM-401, in studying CO biology. Biochem. Pharmacol. 2023, 214, 115642. [Google Scholar] [CrossRef]
  121. Bell, N.T.; Payne, C.M.; Sammut, I.A.; Larsen, D.S. Mechanistic Studies of Carbon Monoxide Release from Norborn-2-en-7-one CORMs. Asian J. Org. Chem. 2022, 11, e202200350. [Google Scholar] [CrossRef]
  122. Payne, F.M.; Nie, S.; Diffee, G.M.; Wilkins, G.T.; Larsen, D.S.; Harrison, J.C.; Baldi, J.C.; Sammut, I.A. The carbon monoxide prodrug oCOm-21 increases Ca2+ sensitivity of the cardiac myofilament. Physiol. Rep. 2024, 12, e15974. [Google Scholar] [CrossRef] [PubMed]
  123. Kharitonov, V.G.; Sharma, V.S.; Pilz, R.B.; Magde, D.; Koesling, D. Basis of guanylate cyclase activation by carbon monoxide. Proc. Natl. Acad. Sci. USA 1995, 92, 2568–2571. [Google Scholar] [CrossRef]
  124. Otterbein, L.E.; Bach, F.H.; Alam, J.; Soares, M.; Lu, H.T.; Wysk, M.; Davis, R.J.; Flavell, R.A.; Choi, A.M. Carbon monoxide has anti-inflammatory effects involving the mitogen-activated protein kinase pathway. Nat. Med. 2000, 6, 422–428. [Google Scholar] [CrossRef]
  125. Zhang, X.; Shan, P.; Alam, J.; Fu, X.-Y.; Lee, P.J. Carbon monoxide differentially modulates STAT1 and STAT3 and inhibits apoptosis via a phosphatidylinositol 3-kinase/Akt and p38 kinase-dependent STAT3 pathway during anoxia-reoxygenation injury. J. Biol. Chem. 2005, 280, 8714–8721. [Google Scholar] [CrossRef]
  126. Sammut, I.A.; Foresti, R.; Clark, J.E.; Exon, D.J.; Vesely, M.J.; Sarathchandra, P.; Green, C.J.; Motterlini, R. Carbon monoxide is a major contributor to the regulation of vascular tone in aortas expressing high levels of haeme oxygenase-1. Br. J. Pharmacol. 1998, 125, 1437–1444. [Google Scholar] [CrossRef] [PubMed]
  127. Brouard, S.; Otterbein, L.E.; Anrather, J.; Tobiasch, E.; Bach, F.H.; Choi, A.M.; Soares, M.P. Carbon monoxide generated by heme oxygenase 1 suppresses endothelial cell apoptosis. J. Exp. Med. 2000, 192, 1015–1026. [Google Scholar] [CrossRef]
  128. Brouard, S.; Berberat, P.O.; Tobiasch, E.; Seldon, M.P.; Bach, F.H.; Soares, M.P. Heme oxygenase-1-derived carbon monoxide requires the activation of transcription factor NF-κB to protect endothelial cells from tumor necrosis factor-α-mediated apoptosis. J. Biol. Chem. 2002, 277, 17950–17961. [Google Scholar] [CrossRef]
  129. Petrache, I.; Otterbein, L.E.; Alam, J.; Wiegand, G.W.; Choi, A.M. Heme oxygenase-1 inhibits TNF-α-induced apoptosis in cultured fibroblasts. Am. J. Physiol. Lung Cell Mol. Physiol. 2000, 278, L312–L319. [Google Scholar] [CrossRef]
  130. Morse, D.; Pischke, S.E.; Zhou, Z.; Davis, R.J.; Flavell, R.A.; Loop, T.; Otterbein, S.L.; Otterbein, L.E.; Choi, A.M. Suppression of inflammatory cytokine production by carbon monoxide involves the JNK pathway and AP-1. J. Biol. Chem. 2003, 278, 36993–36998. [Google Scholar] [CrossRef]
  131. Nakahira, K.; Kim, H.P.; Geng, X.H.; Nakao, A.; Wang, X.; Murase, N.; Drain, P.F.; Wang, X.; Sasidhar, M.; Nabel, E.G. Carbon monoxide differentially inhibits TLR signaling pathways by regulating ROS-induced trafficking of TLRs to lipid rafts. J. Exp. Med. 2006, 203, 2377–2389. [Google Scholar] [CrossRef]
  132. Qin, S.; Du, R.; Yin, S.; Liu, X.; Xu, G.; Cao, W. Nrf2 is essential for the anti-inflammatory effect of carbon monoxide in LPS-induced inflammation. Inflamm. Res. 2015, 64, 537–548. [Google Scholar] [CrossRef] [PubMed]
  133. Wilkinson, W.J.; Gadeberg, H.; Harrison, A.; Allen, N.D.; Riccardi, D.; Kemp, P.J. Carbon monoxide is a rapid modulator of recombinant and native P2X2 ligand-gated ion channels. Br. J. Pharmacol. 2009, 158, 862–871. [Google Scholar] [CrossRef] [PubMed]
  134. Chen, K.; Zhang, J.; Zhang, W.; Zhang, J.; Yang, J.; Li, K.; He, Y. ATP-P2X4 signaling mediates NLRP3 inflammasome activation: A novel pathway of diabetic nephropathy. Int. J. Biochem. Cell Biol. 2013, 45, 932–943. [Google Scholar] [CrossRef] [PubMed]
  135. Karmakar, M.; Katsnelson, M.A.; Dubyak, G.R.; Pearlman, E. Neutrophil P2X 7 receptors mediate NLRP3 inflammasome-dependent IL-1β secretion in response to ATP. Nat. Commun. 2016, 7, 10555. [Google Scholar] [CrossRef] [PubMed]
  136. Wilkinson, W.J.; Kemp, P.J. The carbon monoxide donor, CORM-2, is an antagonist of ATP-gated, human P2X4 receptors. Purinergic Signal. 2011, 7, 57–64. [Google Scholar] [CrossRef]
  137. Jung, S.-S.; Moon, J.-S.; Xu, J.-F.; Ifedigbo, E.; Ryter, S.W.; Choi, A.M.; Nakahira, K. Carbon monoxide negatively regulates NLRP3 inflammasome activation in macrophages. Am. J. Physiol. Lung Cell. Mol. Physiol. 2015, 308, L1058–L1067. [Google Scholar] [CrossRef]
  138. Jiang, L.; Fei, D.; Gong, R.; Yang, W.; Yu, W.; Pan, S.; Zhao, M.; Zhao, M. CORM-2 inhibits TXNIP/NLRP3 inflammasome pathway in LPS-induced acute lung injury. Inflamm. Res. 2016, 65, 905–915. [Google Scholar] [CrossRef]
  139. Liu, Y.; Lian, K.; Zhang, L.; Wang, R.; Yi, F.; Gao, C.; Xin, C.; Zhu, D.; Li, Y.; Yan, W. TXNIP mediates NLRP3 inflammasome activation in cardiac microvascular endothelial cells as a novel mechanism in myocardial ischemia/reperfusion injury. Basic Res. Cardiol. 2014, 109, 415. [Google Scholar] [CrossRef]
  140. Zhou, R.; Tardivel, A.; Thorens, B.; Choi, I.; Tschopp, J. Thioredoxin-interacting protein links oxidative stress to inflammasome activation. Nat. Immunol. 2010, 11, 136–140. [Google Scholar] [CrossRef]
  141. Chen, D.; Dixon, B.J.; Doycheva, D.M.; Li, B.; Zhang, Y.; Hu, Q.; He, Y.; Guo, Z.; Nowrangi, D.; Flores, J. IRE1α inhibition decreased TXNIP/NLRP3 inflammasome activation through miR-17-5p after neonatal hypoxic–ischemic brain injury in rats. J. Neuroinflamm. 2018, 15, 32. [Google Scholar] [CrossRef] [PubMed]
  142. Chung, J.; Shin, D.-Y.; Zheng, M.; Joe, Y.; Pae, H.-O.; Ryter, S.W.; Chung, H.-T. Carbon monoxide, a reaction product of heme oxygenase-1, suppresses the expression of C-reactive protein by endoplasmic reticulum stress through modulation of the unfolded protein response. Mol. Immunol. 2011, 48, 1793–1799. [Google Scholar] [CrossRef]
  143. Kim, K.M.; Pae, H.-O.; Zheng, M.; Park, R.; Kim, Y.-M.; Chung, H.-T. Carbon Monoxide Induces Heme Oxygenase-1 via Activation of Protein Kinase R–Like Endoplasmic Reticulum Kinase and Inhibits Endothelial Cell Apoptosis Triggered by Endoplasmic Reticulum Stress. Circ. Res. 2007, 101, 919–927. [Google Scholar] [CrossRef]
  144. Zheng, G.; Zhan, Y.; Wang, H.; Luo, Z.; Zheng, F.; Zhou, Y.; Wu, Y.; Wang, S.; Wu, Y.; Xiang, G. Carbon monoxide releasing molecule-3 alleviates neuron death after spinal cord injury via inflammasome regulation. EBioMedicine 2019, 40, 643–654. [Google Scholar] [CrossRef]
  145. Zhang, W.; Tao, A.; Lan, T.; Cepinskas, G.; Kao, R.; Martin, C.M.; Rui, T. Carbon monoxide releasing molecule-3 improves myocardial function in mice with sepsis by inhibiting NLRP3 inflammasome activation in cardiac fibroblasts. Basic Res. Cardiol. 2017, 112, 16. [Google Scholar] [CrossRef] [PubMed]
  146. Palmer, R.M.; Ferrige, A.; Moncada, S. Nitric oxide release accounts for the biological activity of endothelium-derived relaxing factor. Nature 1987, 327, 524–526. [Google Scholar] [CrossRef] [PubMed]
  147. Palmer, R.M.; Ashton, D.; Moncada, S. Vascular endothelial cells synthesize nitric oxide from L-arginine. Nature 1988, 333, 664–666. [Google Scholar] [CrossRef]
  148. Balligand, J.-L.; Ungureanu-Longrois, D.; Simmons, W.W.; Pimental, D.; Malinski, T.A.; Kapturczak, M.; Taha, Z.; Lowenstein, C.J.; Davidoff, A.J.; Kelly, R.A. Cytokine-inducible nitric oxide synthase (iNOS) expression in cardiac myocytes. Characterization and regulation of iNOS expression and detection of iNOS activity in single cardiac myocytes in vitro. J. Biol. Chem. 1994, 269, 27580–27588. [Google Scholar] [CrossRef]
  149. Huang, P.L.; Huang, Z.; Mashimo, H.; Bloch, K.D.; Moskowitz, M.A.; Bevan, J.A.; Fishman, M.C. Hypertension in mice lacking the gene for endothelial nitric oxide synthase. Nature 1995, 377, 239–242. [Google Scholar] [CrossRef]
  150. Xu, K.Y.; Huso, D.L.; Dawson, T.M.; Bredt, D.S.; Becker, L.C. Nitric oxide synthase in cardiac sarcoplasmic reticulum. Proc. Natl. Acad. Sci. USA 1999, 96, 657–662. [Google Scholar] [CrossRef]
  151. Kelm, M.; Schrader, J. Control of coronary vascular tone by nitric oxide. Circ. Res. 1990, 66, 1561–1575. [Google Scholar] [CrossRef] [PubMed]
  152. Kubes, P.; Suzuki, M.; Granger, D. Nitric oxide: An endogenous modulator of leukocyte adhesion. Proc. Natl. Acad. Sci. USA 1991, 88, 4651–4655. [Google Scholar] [CrossRef] [PubMed]
  153. Radomski, M.; Palmer, R.; Moncada, S. Endogenous nitric oxide inhibits human platelet adhesion to vascular endothelium. Lancet 1987, 330, 1057–1058. [Google Scholar] [CrossRef]
  154. Stamler, J.S.; Loh, E.; Roddy, M.-A.; Currie, K.E.; Creager, M.A. Nitric oxide regulates basal systemic and pulmonary vascular resistance in healthy humans. Circulation 1994, 89, 2035–2040. [Google Scholar] [CrossRef]
  155. Bredt, D.S.; Hwang, P.M.; Snyder, S.H. Localization of nitric oxide synthase indicating a neural role for nitric oxide. Nature 1990, 347, 768–770. [Google Scholar] [CrossRef] [PubMed]
  156. Ishide, T.; Nauli, S.M.; Maher, T.J.; Ally, A. Cardiovascular responses and neurotransmitter changes following blockade of nNOS within the ventrolateral medulla during static muscle contraction. Brain Res. 2003, 977, 80–89. [Google Scholar] [CrossRef]
  157. Kurihara, N.; Alfie, M.E.; Sigmon, D.H.; Rhaleb, N.-E.; Shesely, E.G.; Carretero, O.A. Role of nNOS in blood pressure regulation in eNOS null mutant mice. Hypertension 1998, 32, 856–861. [Google Scholar] [CrossRef]
  158. Wang, Q.; Pelligrino, D.A.; Baughman, V.L.; Koenig, H.M.; Albrecht, R.F. The role of neuronal nitric oxide synthase in regulation of cerebral blood flow in normocapnia and hypercapnia in rats. J. Cereb. Blood Flow. Metab. 1995, 15, 774–778. [Google Scholar] [CrossRef]
  159. Griscavage, J.M.; Wilk, S.; Ignarro, L.J. Inhibitors of the proteasome pathway interfere with induction of nitric oxide synthase in macrophages by blocking activation of transcription factor NF-kappa B. Proc. Natl. Acad. Sci. USA 1996, 93, 3308–3312. [Google Scholar] [CrossRef]
  160. Connelly, L.; Palacios-Callender, M.; Ameixa, C.; Moncada, S.; Hobbs, A. Biphasic regulation of NF-κB activity underlies the pro-and anti-inflammatory actions of nitric oxide. J. Immunol. 2001, 166, 3873–3881. [Google Scholar] [CrossRef]
  161. Sato, M.; Hida, N.; Umezawa, Y. Imaging the nanomolar range of nitric oxide with an amplifier-coupled fluorescent indicator in living cells. Proc. Natl. Acad. Sci. USA 2005, 102, 14515–14520. [Google Scholar] [CrossRef] [PubMed]
  162. Hung, L.-M.; Su, M.-J.; Chen, J.-K. Resveratrol protects myocardial ischemia–reperfusion injury through both NO-dependent and NO-independent mechanisms. Free Radic. Biol. Med. 2004, 36, 774–781. [Google Scholar] [CrossRef]
  163. Czapski, G.; Goldstein, S. The role of the reactions of· NO with superoxide and oxygen in biological systems: A kinetic approach. Free Radic. Biol. Med. 1995, 19, 785–794. [Google Scholar] [CrossRef] [PubMed]
  164. Yang, Y.; Huang, Z.; Li, L.-L. Advanced nitric oxide donors: Chemical structure of NO drugs, NO nanomedicines and biomedical applications. Nanoscale 2021, 13, 444–459. [Google Scholar] [CrossRef] [PubMed]
  165. Adhikari, N.K.; Dellinger, R.P.; Lundin, S.; Payen, D.; Vallet, B.; Gerlach, H.; Park, K.J.; Mehta, S.; Slutsky, A.S.; Friedrich, J.O. Inhaled nitric oxide does not reduce mortality in patients with acute respiratory distress syndrome regardless of severity: Systematic review and meta-analysis. Crit. Care Med. 2014, 42, 404–412. [Google Scholar] [CrossRef] [PubMed]
  166. Pisarenko, O.; Studneva, I. Modulating the Bioactivity of Nitric Oxide as a Therapeutic Strategy in Cardiac Surgery. J. Surg. Res. 2021, 257, 178–188. [Google Scholar] [CrossRef]
  167. Radi, R.; Beckman, J.S.; Bush, K.M.; Freeman, B.A. Peroxynitrite oxidation of sulfhydryls.: The cytotoxic potential of superoxide and nitric oxide. J. Biol. Chem. 1991, 266, 4244–4250. [Google Scholar] [CrossRef]
  168. Ishida, H.; Ichimori, K.; Hirota, Y.; Fukahori, M.; Nakazawa, H. Peroxynitrite-induced cardiac myocyte injury. Free Radic. Biol. Med. 1996, 20, 343–350. [Google Scholar] [CrossRef]
  169. Kaur, H.; Halliwell, B. Evidence for nitric oxide-mediated oxidative damage in chronic inflammation Nitrotyrosine in serum and synovial fluid from rheumatoid patients. FEBS Lett. 1994, 350, 9–12. [Google Scholar] [CrossRef]
  170. Li, D.-Y.; Tao, L.; Liu, H.; Christopher, T.; Lopez, B.; Ma, X. Role of ERK1/2 in the anti-apoptotic and cardioprotective effects of nitric oxide after myocardial ischemia and reperfusion. Apoptosis 2006, 11, 923–930. [Google Scholar] [CrossRef]
  171. Elliott, S.N.; McKnight, W.; Cirino, G.; Wallace, J.L. A nitric oxide-releasing nonsteroidal anti-inflammatory drug accelerates gastric ulcer healing in rats. Gastroenterology 1995, 109, 524–530. [Google Scholar] [CrossRef] [PubMed]
  172. Ma, X.; Sayed, N.; Beuve, A.; Van Den Akker, F. NO and CO differentially activate soluble guanylyl cyclase via a heme pivot-bend mechanism. EMBO J. 2007, 26, 578–588. [Google Scholar] [CrossRef]
  173. Martin, E.; Berka, V.; Bogatenkova, E.; Murad, F.; Tsai, A.-L. Ligand selectivity of soluble guanylyl cyclase: Effect of the hydrogen-bonding tyrosine in the distal heme pocket on binding of oxygen, nitric oxide, and carbon monoxide. J. Biol. Chem. 2006, 281, 27836–27845. [Google Scholar] [CrossRef]
  174. Snyder, C.M.; Shroff, E.H.; Liu, J.; Chandel, N.S. Nitric oxide induces cell death by regulating anti-apoptotic BCL-2 family members. PLoS ONE 2009, 4, e7059. [Google Scholar] [CrossRef]
  175. Grabowski, P.; Wright, P.; Van’t Hof, R.; Helfrich, M.; Ohshima, H.; Ralston, S. Immunolocalization of inducible nitric oxide synthase in synovium and cartilage in rheumatoid arthritis and osteoarthritis. Br. J. Rheumatol. 1997, 36, 651–655. [Google Scholar] [CrossRef] [PubMed]
  176. Hinder, F.; Stubbe, H.D.; VAN AKEN, H.; Waurick, R.; Booke, M.; Meyer, J. Role of nitric oxide in sepsis-associated pulmonary edema. Am. J. Respir. Crit. Care Med. 1999, 159, 252–257. [Google Scholar] [CrossRef] [PubMed]
  177. Rachmilewitz, D.; Stamler, J.; Bachwich, D.; Karmeli, F.; Ackerman, Z.; Podolsky, D. Enhanced colonic nitric oxide generation and nitric oxide synthase activity in ulcerative colitis and Crohn’s disease. Gut 1995, 36, 718–723. [Google Scholar] [CrossRef] [PubMed]
  178. Lee, W.J.; Tateya, S.; Cheng, A.M.; Rizzo-DeLeon, N.; Wang, N.F.; Handa, P.; Wilson, C.L.; Clowes, A.W.; Sweet, I.R.; Bomsztyk, K. M2 macrophage polarization mediates anti-inflammatory effects of endothelial nitric oxide signaling. Diabetes 2015, 64, 2836–2846. [Google Scholar] [CrossRef]
  179. Niedbala, W.; Cai, B.; Liu, H.; Pitman, N.; Chang, L.; Liew, F.Y. Nitric oxide induces CD4+ CD25+ Foxp3− regulatory T cells from CD4+ CD25− T cells via p53, IL-2, and OX40. Proc. Natl. Acad. Sci. USA 2007, 104, 15478–15483. [Google Scholar] [CrossRef]
  180. Niedbala, W.; Alves-Filho, J.C.; Fukada, S.Y.; Vieira, S.M.; Mitani, A.; Sonego, F.; Mirchandani, A.; Nascimento, D.C.; Cunha, F.Q.; Liew, F.Y. Regulation of type 17 helper T-cell function by nitric oxide during inflammation. Proc. Natl. Acad. Sci. USA 2011, 108, 9220–9225. [Google Scholar] [CrossRef]
  181. Hernandez-Cuellar, E.; Tsuchiya, K.; Hara, H.; Fang, R.; Sakai, S.; Kawamura, I.; Akira, S.; Mitsuyama, M. Cutting edge: Nitric oxide inhibits the NLRP3 inflammasome. J. Immunol. 2012, 189, 5113–5117. [Google Scholar] [CrossRef] [PubMed]
  182. Mao, K.; Chen, S.; Chen, M.; Ma, Y.; Wang, Y.; Huang, B.; He, Z.; Zeng, Y.; Hu, Y.; Sun, S. Nitric oxide suppresses NLRP3 inflammasome activation and protects against LPS-induced septic shock. Cell Res. 2013, 23, 201–212. [Google Scholar] [CrossRef] [PubMed]
  183. Jaffrey, S.R.; Erdjument-Bromage, H.; Ferris, C.D.; Tempst, P.; Snyder, S.H. Protein S-nitrosylation: A physiological signal for neuronal nitric oxide. Nat. Cell Biol. 2001, 3, 193–197. [Google Scholar] [CrossRef] [PubMed]
  184. Juliana, C.; Fernandes-Alnemri, T.; Wu, J.; Datta, P.; Solorzano, L.; Yu, J.-W.; Meng, R.; Quong, A.A.; Latz, E.; Scott, C.P. Anti-inflammatory compounds parthenolide and Bay 11-7082 are direct inhibitors of the inflammasome. J. Biol. Chem. 2010, 285, 9792–9802. [Google Scholar] [CrossRef]
  185. Dimmeler, S.; Haendeler, J.; Nehls, M.; Zeiher, A.M. Suppression of apoptosis by nitric oxide via inhibition of interleukin-1β–converting enzyme (ice)-like and cysteine protease protein (cpp)-32–like proteases. J. Exp. Med. 1997, 185, 601–608. [Google Scholar] [CrossRef]
  186. Mishra, B.B.; Rathinam, V.A.; Martens, G.W.; Martinot, A.J.; Kornfeld, H.; Fitzgerald, K.A.; Sassetti, C.M. Nitric oxide controls the immunopathology of tuberculosis by inhibiting NLRP3 inflammasome–dependent processing of IL-1β. Nat. Immunol. 2013, 14, 52–60. [Google Scholar] [CrossRef]
  187. Asanuma, H.; Kitakaze, M.; Funaya, H.; Takashima, S.; Minamino, T.; Node, K.; Sakata, Y.; Asakura, M.; Sanada, S.; Shinozaki, Y. Nifedipine limits infarct size via NO-dependent mechanisms in dogs. Basic Res. Cardiol. 2001, 96, 497–505. [Google Scholar] [CrossRef]
  188. Kamenshchikov, N.O.; Mandel, I.A.; Podoksenov, Y.K.; Svirko, Y.S.; Lomivorotov, V.V.; Mikheev, S.L.; Kozlov, B.N.; Shipulin, V.M.; Nenakhova, A.A.; Anfinogenova, Y.J. Nitric oxide provides myocardial protection when added to the cardiopulmonary bypass circuit during cardiac surgery: Randomized trial. J. Thorac. Cardiovasc. Surg. 2019, 157, 2328–2336.e1. [Google Scholar] [CrossRef]
  189. Sun, J.; Aponte, A.M.; Menazza, S.; Gucek, M.; Steenbergen, C.; Murphy, E. Additive cardioprotection by pharmacological postconditioning with hydrogen sulfide and nitric oxide donors in mouse heart: S-sulfhydration vs. S-nitrosylation. Cardiovasc. Res. 2016, 110, 96–106. [Google Scholar] [CrossRef]
  190. Gianetti, J.; Del Sarto, P.; Bevilacqua, S.; Vassalle, C.; De Filippis, R.; Kacila, M.; Farneti, P.A.; Clerico, A.; Glauber, M.; Biagini, A. Supplemental nitric oxide and its effect on myocardial injury and function in patients undergoing cardiac surgery with extracorporeal circulation. J. Thorac. Cardiovasc. Surg. 2004, 127, 44–50. [Google Scholar] [CrossRef]
  191. Goodwin, L.R.; Francom, D.; Dieken, F.P.; Taylor, J.D.; Warenycia, M.W.; Reiffenstein, R.; Dowling, G. Determination of sulfide in brain tissue by gas dialysis/ion chromatography: Postmortem studies and two case reports. J. Anal. Toxicol. 1989, 13, 105–109. [Google Scholar] [CrossRef]
  192. Ishigami, M.; Hiraki, K.; Umemura, K.; Ogasawara, Y.; Ishii, K.; Kimura, H. A source of hydrogen sulfide and a mechanism of its release in the brain. Antioxid. Redox Signal. 2009, 11, 205–214. [Google Scholar] [CrossRef] [PubMed]
  193. Stipanuk, M.H.; Beck, P.W. Characterization of the enzymic capacity for cysteine desulphhydration in liver and kidney of the rat. Biochem. J. 1982, 206, 267–277. [Google Scholar] [CrossRef]
  194. Tripatara, P.; Patel, N.S.; Brancaleone, V.; Renshaw, D.; Rocha, J.; Sepodes, B.; Mota-Filipe, H.; Perretti, M.; Thiemermann, C. Characterisation of cystathionine gamma-lyase/hydrogen sulphide pathway in ischaemia/reperfusion injury of the mouse kidney: An in vivo study. Eur. J. Pharmacol. 2009, 606, 205–209. [Google Scholar] [CrossRef]
  195. Filipovic, M.R.; Zivanovic, J.; Alvarez, B.; Banerjee, R. Chemical biology of H2S signaling through persulfidation. Chem. Rev. 2018, 118, 1253–1337. [Google Scholar] [CrossRef]
  196. Chou, C.-H.; Ogden, J.M.; Pohl, H.R.; Scinicariello, F.; Ingerman, L.; Barber, L.; Citra, M.J. Toxicological Profile for Hydrogen Sulfide and Carbonyl Sulfide; Agency for Toxic Substances and Disease Registry: Washington, DC, USA, 2016. [Google Scholar]
  197. Hill, B.; Woon, T.; Nicholls, P.; Peterson, J.; Greenwood, C.; Thomson, A. Interactions of sulphide and other ligands with cytochrome c oxidase. An electron-paramagnetic-resonance study. Biochem. J. 1984, 224, 591–600. [Google Scholar] [CrossRef] [PubMed]
  198. Liang, D.; Wu, H.; Wong, M.W.; Huang, D. Diallyl trisulfide is a fast H2S donor, but diallyl disulfide is a slow one: The reaction pathways and intermediates of glutathione with polysulfides. Org. Lett. 2015, 17, 4196–4199. [Google Scholar] [CrossRef]
  199. Zanardo, R.C.; Brancaleone, V.; Distrutti, E.; Fiorucci, S.; Cirino, G.; Wallace, J.L.; Zanardo, R.C.; Brancaleone, V.; Distrutti, E.; Fiorucci, S. Hydrogen sulfide is an endogenous modulator of leukocyte-mediated inflammation. FASEB J. 2006, 20, 2118–2120. [Google Scholar] [CrossRef]
  200. Papapetropoulos, A.; Whiteman, M.; Cirino, G. Pharmacological tools for hydrogen sulphide research: A brief, introductory guide for beginners. Br. J. Pharmacol. 2015, 172, 1633–1637. [Google Scholar] [CrossRef]
  201. Polhemus, D.J.; Li, Z.; Pattillo, C.B.; Gojon, G., Sr.; Gojon, G., Jr.; Giordano, T.; Krum, H. A novel hydrogen sulfide prodrug, SG 1002, promotes hydrogen sulfide and nitric oxide bioavailability in heart failure patients. Cardiovasc. Ther. 2015, 33, 216–226. [Google Scholar] [CrossRef]
  202. Nguyen, K.; Chau, V.Q.; Mauro, A.G.; Durrant, D.; Toldo, S.; Abbate, A.; Das, A.; Salloum, F.N. Hydrogen sulfide therapy suppresses cofilin-2 and attenuates ischemic heart failure in a mouse model of myocardial infarction. J. Cardiovasc. Pharmacol. Ther. 2020, 25, 472–483. [Google Scholar] [CrossRef] [PubMed]
  203. Sivarajah, A.; Collino, M.; Yasin, M.; Benetti, E.; Gallicchio, M.; Mazzon, E.; Cuzzocrea, S.; Fantozzi, R.; Thiemermann, C. Anti-apoptotic and anti-inflammatory effects of hydrogen sulfide in a rat model of regional myocardial I/R. Shock 2009, 31, 267–274. [Google Scholar] [CrossRef] [PubMed]
  204. Karwi, Q.G.; Bice, J.S.; Baxter, G.F. Pre-and postconditioning the heart with hydrogen sulfide (H2S) against ischemia/reperfusion injury in vivo: A systematic review and meta-analysis. Basic. Res. Cardiol. 2018, 113, 6. [Google Scholar] [CrossRef]
  205. Ertugrul, I.A.; van Suylen, V.; Damman, K.; de Koning, M.-S.L.; van Goor, H.; Erasmus, M.E. Donor Heart Preservation with Hydrogen Sulfide: A Systematic Review and Meta-Analysis. Int. J. Mol. Sci. 2021, 22, 5737. [Google Scholar] [CrossRef]
  206. Johansen, D.; Ytrehus, K.; Baxter, G.F. Exogenous hydrogen sulfide (H2S) protects against regional myocardial ischemia–reperfusion injury. Basic Res. Cardiol. 2006, 101, 53–60. [Google Scholar] [CrossRef]
  207. Meng, G.; Zhu, J.; Xiao, Y.; Huang, Z.; Zhang, Y.; Tang, X.; Xie, L.; Chen, Y.; Shao, Y.; Ferro, A. Hydrogen sulfide donor GYY4137 protects against myocardial fibrosis. Oxid. Med. Cell. Longev. 2015, 2015, 691070. [Google Scholar] [CrossRef]
  208. Meng, G.; Xiao, Y.; Ma, Y.; Tang, X.; Xie, L.; Liu, J.; Gu, Y.; Yu, Y.; Park, C.M.; Xian, M. Hydrogen sulfide regulates Krüppel-like factor 5 transcription activity via specificity protein 1 S-sulfhydration at Cys664 to prevent myocardial hypertrophy. J. Am. Heart Assoc. 2016, 5, e004160. [Google Scholar] [CrossRef] [PubMed]
  209. Xie, L.; Feng, H.; Li, S.; Meng, G.; Liu, S.; Tang, X.; Ma, Y.; Han, Y.; Xiao, Y.; Gu, Y. SIRT3 mediates the antioxidant effect of hydrogen sulfide in endothelial cells. Antioxid. Redox Signal. 2016, 24, 329–343. [Google Scholar] [CrossRef]
  210. Zhao, W.; Zhang, J.; Lu, Y.; Wang, R. The vasorelaxant effect of H2S as a novel endogenous gaseous KATP channel opener. EMBO J. 2001, 20, 6008–6016. [Google Scholar] [CrossRef]
  211. Lee, S.W.; Cheng, Y.; Moore, P.K.; Bian, J.-S. Hydrogen sulphide regulates intracellular pH in vascular smooth muscle cells. Biochem. Biophys. Res. Commun. 2007, 358, 1142–1147. [Google Scholar] [CrossRef]
  212. Lim, J.J.; Liu, Y.-H.; Khin, E.S.W.; Bian, J.-S. Vasoconstrictive effect of hydrogen sulfide involves downregulation of cAMP in vascular smooth muscle cells. Am. J. Physiol. Cell Physiol. 2008, 295, C1261–C1270. [Google Scholar] [CrossRef]
  213. Yong, Q.C.; Pan, T.-T.; Hu, L.-F.; Bian, J.-S. Negative regulation of β-adrenergic function by hydrogen sulphide in the rat hearts. J. Mol. Cell. Cardiol. 2008, 44, 701–710. [Google Scholar] [CrossRef]
  214. Sun, Y.-G.; Cao, Y.-X.; Wang, W.-W.; Ma, S.-F.; Yao, T.; Zhu, Y.-C. Hydrogen sulphide is an inhibitor of L-type calcium channels and mechanical contraction in rat cardiomyocytes. Cardiovasc. Res. 2008, 79, 632–641. [Google Scholar] [CrossRef] [PubMed]
  215. Kuo, M.M.; Kim, D.H.; Jandu, S.; Bergman, Y.; Tan, S.; Wang, H.; Pandey, D.R.; Abraham, T.P.; Shoukas, A.A.; Berkowitz, D.E. MPST but not CSE is the primary regulator of hydrogen sulfide production and function in the coronary artery. Am. J. Physiol. Heart Circ. Physiol. 2016, 310, H71–H79. [Google Scholar] [CrossRef]
  216. White, B.J.; Smith, P.A.; Dunn, W.R. Hydrogen sulphide–mediated vasodilatation involves the release of neurotransmitters from sensory nerves in pressurized mesenteric small arteries isolated from rats. Br. J. Pharmacol. 2013, 168, 785–793. [Google Scholar] [CrossRef]
  217. Ariyaratnam, P.; Loubani, M.; Morice, A.H. Hydrogen sulphide vasodilates human pulmonary arteries: A possible role in pulmonary hypertension? Microvasc. Res. 2013, 90, 135–137. [Google Scholar] [CrossRef]
  218. Yin, J.; Tu, C.; Zhao, J.; Ou, D.; Chen, G.; Liu, Y.; Xiao, X. Exogenous hydrogen sulfide protects against global cerebral ischemia/reperfusion injury via its anti-oxidative, anti-inflammatory and anti-apoptotic effects in rats. Brain Res. 2013, 1491, 188–196. [Google Scholar] [CrossRef] [PubMed]
  219. Sivarajah, A.; McDonald, M.; Thiemermann, C. The production of hydrogen sulfide limits myocardial ischemia and reperfusion injury and contributes to the cardioprotective effects of preconditioning with endotoxin, but not ischemia in the rat. Shock 2006, 26, 154–161. [Google Scholar] [CrossRef] [PubMed]
  220. Elrod, J.W.; Calvert, J.W.; Morrison, J.; Doeller, J.E.; Kraus, D.W.; Tao, L.; Jiao, X.; Scalia, R.; Kiss, L.; Szabo, C. Hydrogen sulfide attenuates myocardial ischemia-reperfusion injury by preservation of mitochondrial function. Proc. Natl. Acad. Sci. USA 2007, 104, 15560–15565. [Google Scholar] [CrossRef]
  221. Yang, G.; Sun, X.; Wang, R. Hydrogen sulfide-induced apoptosis of human aorta smooth muscle cells via the activation of mitogen-activated protein kinases and caspase-3. FASEB J. 2004, 18, 1782–1784. [Google Scholar] [CrossRef]
  222. Yusof, M.; Kamada, K.; Kalogeris, T.; Gaskin, F.S.; Korthuis, R.J. Hydrogen sulfide triggers late-phase preconditioning in postischemic small intestine by an NO-and p38 MAPK-dependent mechanism. Am. J. Physiol. Heart Circ. Physiol. 2009, 296, H868–H876. [Google Scholar] [CrossRef]
  223. Lohninger, L.; Tomasova, L.; Praschberger, M.; Hintersteininger, M.; Erker, T.; Gmeiner, B.M.; Laggner, H. Hydrogen sulphide induces HIF-1α and Nrf2 in THP-1 macrophages. Biochimie 2015, 112, 187–195. [Google Scholar] [CrossRef] [PubMed]
  224. Guo, C.; Liang, F.; Masood, W.S.; Yan, X. Hydrogen sulfide protected gastric epithelial cell from ischemia/reperfusion injury by Keap1 s-sulfhydration, MAPK dependent anti-apoptosis and NF-κB dependent anti-inflammation pathway. Eur. J. Pharmacol. 2014, 725, 70–78. [Google Scholar] [CrossRef] [PubMed]
  225. Liu, F.; Liu, G.J.; Liu, N.; Zhang, G.; Zhang, J.X.; Li, L.F. Effect of hydrogen sulfide on inflammatory cytokines in acute myocardial ischemia injury in rats. Exp. Ther. Med. 2015, 9, 1068–1074. [Google Scholar] [CrossRef] [PubMed]
  226. Calvert, J.W.; Jha, S.; Gundewar, S.; Elrod, J.W.; Ramachandran, A.; Pattillo, C.B.; Kevil, C.G.; Lefer, D.J. Hydrogen sulfide mediates cardioprotection through Nrf2 signaling. Circ. Res. 2009, 105, 365–374. [Google Scholar] [CrossRef]
  227. Xie, L.; Gu, Y.; Wen, M.; Zhao, S.; Wang, W.; Ma, Y.; Meng, G.; Han, Y.; Wang, Y.; Liu, G. Hydrogen sulfide induces Keap1 S-sulfhydration and suppresses diabetes-accelerated atherosclerosis via Nrf2 activation. Diabetes 2016, 65, 3171–3184. [Google Scholar] [CrossRef]
  228. Itoh, K.; Wakabayashi, N.; Katoh, Y.; Ishii, T.; Igarashi, K.; Engel, J.D.; Yamamoto, M. Keap1 represses nuclear activation of antioxidant responsive elements by Nrf2 through binding to the amino-terminal Neh2 domain. Genes Dev. 1999, 13, 76–86. [Google Scholar] [CrossRef]
  229. Yang, G.; Zhao, K.; Ju, Y.; Mani, S.; Cao, Q.; Puukila, S.; Khaper, N.; Wu, L.; Wang, R. Hydrogen sulfide protects against cellular senescence via S-sulfhydration of Keap1 and activation of Nrf2. Antioxid. Redox Signal. 2013, 18, 1906–1919. [Google Scholar] [CrossRef]
  230. Zhang, D.; Macinkovic, I.; Devarie-Baez, N.O.; Pan, J.; Park, C.M.; Carroll, K.S.; Filipovic, M.R.; Xian, M. Detection of protein S-sulfhydration by a tag-switch technique. Angew. Chem. Int. Ed. 2014, 53, 575–581. [Google Scholar] [CrossRef]
  231. Lin, Z.; Altaf, N.; Li, C.; Chen, M.; Pan, L.; Wang, D.; Xie, L.; Zheng, Y.; Fu, H.; Han, Y. Hydrogen sulfide attenuates oxidative stress-induced NLRP3 inflammasome activation via S-sulfhydrating c-Jun at Cys269 in macrophages. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 2890–2900. [Google Scholar] [CrossRef]
  232. Duewell, P.; Kono, H.; Rayner, K.J.; Sirois, C.M.; Vladimer, G.; Bauernfeind, F.G.; Abela, G.S.; Franchi, L.; Nunez, G.; Schnurr, M. NLRP3 inflammasomes are required for atherogenesis and activated by cholesterol crystals. Nature 2010, 464, 1357–1361. [Google Scholar] [CrossRef]
  233. Gan, W.; Ren, J.; Li, T.; Lv, S.; Li, C.; Liu, Z.; Yang, M. The SGK1 inhibitor EMD638683, prevents Angiotensin II–induced cardiac inflammation and fibrosis by blocking NLRP3 inflammasome activation. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 1–10. [Google Scholar] [CrossRef]
  234. Sheedy, F.J.; Grebe, A.; Rayner, K.J.; Kalantari, P.; Ramkhelawon, B.; Carpenter, S.B.; Becker, C.E.; Ediriweera, H.N.; Mullick, A.E.; Golenbock, D.T. CD36 coordinates NLRP3 inflammasome activation by facilitating intracellular nucleation of soluble ligands into particulate ligands in sterile inflammation. Nat. Immunol. 2013, 14, 812–820. [Google Scholar] [CrossRef] [PubMed]
  235. Zhao, H.; Pan, P.; Yang, Y.; Ge, H.; Chen, W.; Qu, J.; Shi, J.; Cui, G.; Liu, X.; Feng, H. Endogenous hydrogen sulphide attenuates NLRP3 inflammasome-mediated neuroinflammation by suppressing the P2X7 receptor after intracerebral haemorrhage in rats. J. Neuroinflamm. 2017, 14, 163. [Google Scholar] [CrossRef]
  236. Tian, D.; Dong, J.; Jin, S.; Teng, X.; Wu, Y. Endogenous hydrogen sulfide-mediated MAPK inhibition preserves endothelial function through TXNIP signaling. Free Radic. Biol. Med. 2017, 110, 291–299. [Google Scholar] [CrossRef] [PubMed]
  237. Jia, Q.; Mehmood, S.; Liu, X.; Ma, S.; Yang, R. Hydrogen sulfide mitigates myocardial inflammation by inhibiting nucleotide-binding oligomerization domain-like receptor protein 3 inflammasome activation in diabetic rats. Exp. Biol. Med. 2020, 245, 221–230. [Google Scholar] [CrossRef] [PubMed]
  238. Huang, Z.; Zhuang, X.; Xie, C.; Hu, X.; Dong, X.; Guo, Y.; Li, S.; Liao, X. Exogenous hydrogen sulfide attenuates high glucose-induced cardiotoxicity by inhibiting NLRP3 inflammasome activation by suppressing TLR4/NF-κB pathway in H9c2 cells. Cell. Physiol. Biochem. 2016, 40, 1578–1590. [Google Scholar] [CrossRef]
  239. Mudaliar, H.; Pollock, C.; Ma, J.; Wu, H.; Chadban, S.; Panchapakesan, U. The role of TLR2 and 4-mediated inflammatory pathways in endothelial cells exposed to high glucose. PLoS ONE 2014, 9, e108844. [Google Scholar] [CrossRef]
  240. Qiu, Z.; Lei, S.; Zhao, B.; Wu, Y.; Su, W.; Liu, M.; Meng, Q.; Zhou, B.; Leng, Y.; Xia, Z.-y. NLRP3 inflammasome activation-mediated pyroptosis aggravates myocardial ischemia/reperfusion injury in diabetic rats. Oxid. Med. Cell. Longev. 2017, 2017, 9743280. [Google Scholar] [CrossRef] [PubMed]
  241. Xia, Y.; Zhang, W.; He, K.; Bai, L.; Miao, Y.; Liu, B.; Zhang, X.; Jin, S.; Wu, Y. Hydrogen sulfide alleviates lipopolysaccharide-induced myocardial injury through TLR4-NLRP3 pathway. Physiol. Res. 2023, 72, 15. [Google Scholar] [CrossRef]
  242. Wang, H.; Zhong, P.; Sun, L. Exogenous hydrogen sulfide mitigates NLRP3 inflammasome-mediated inflammation through promoting autophagy via the AMPK-mTOR pathway. Biol. Open 2019, 8, bio043653. [Google Scholar] [CrossRef] [PubMed]
  243. Liu, X.; Zhang, X.; Ding, Y.; Zhou, W.; Tao, L.; Lu, P.; Wang, Y.; Hu, R. Nuclear factor E2-related factor-2 negatively regulates NLRP3 inflammasome activity by inhibiting reactive oxygen species-induced NLRP3 priming. Antioxid. Redox Signal. 2017, 26, 28–43. [Google Scholar] [CrossRef] [PubMed]
  244. Li, J.; Teng, X.; Jin, S.; Dong, J.; Guo, Q.; Tian, D.; Wu, Y. Hydrogen sulfide improves endothelial dysfunction by inhibiting the vicious cycle of NLRP3 inflammasome and oxidative stress in spontaneously hypertensive rats. J. Hypertens. 2019, 37, 1633–1643. [Google Scholar] [CrossRef]
  245. Su, Y.; Wang, Y.; Liu, M.; Chen, H. Hydrogen sulfide attenuates renal I/R-induced activation of the inflammatory response and apoptosis via regulating Nrf2-mediated NLRP3 signaling pathway inhibition. Mol. Med. Rep. 2021, 24, 518. [Google Scholar] [CrossRef]
  246. Zuurbier, C.J.; Jong, W.M.; Eerbeek, O.; Koeman, A.; Pulskens, W.P.; Butter, L.M.; Leemans, J.C.; Hollmann, M.W. Deletion of the innate immune NLRP3 receptor abolishes cardiac ischemic preconditioning and is associated with decreased Il-6/STAT3 signaling. PLoS ONE 2012, 7, e40643. [Google Scholar] [CrossRef] [PubMed]
  247. Sandanger, Ø.; Gao, E.; Ranheim, T.; Bliksøen, M.; Kaasbøll, O.; Alfsnes, K.; Nymo, S.H.; Rashidi, A.; Ohm, I.; Attramadal, H. NLRP3 inflammasome activation during myocardial ischemia reperfusion is cardioprotective. Biochem. Biophys. Res. Commun. 2016, 469, 1012–1020. [Google Scholar] [CrossRef]
  248. Hollmann, M.W.; Zuurbier, C.J. Nlrp3 plays no role in acute cardiac infarction due to low cardiac expression. Int. J. Cardiol. 2014, 177, 41–43. [Google Scholar]
Figure 1. Structure of the NLRP3 inflammasome. Associations between NLRP3 and ASC occur via PYD-PYD interactions. NLRP3 does not have a CARD domain and requires ASC to interface with the CARD domain present on pro-caspase-1 using a CARD-CARD interaction. NEK7 interacts with the LRR domain on NLRP3. Multiple replicates of these interactions come together and assemble the NLRP3 inflammasome with pro-caspase 1 filaments branching out from the central core of polymerised ASC to cleave proteins.
Figure 1. Structure of the NLRP3 inflammasome. Associations between NLRP3 and ASC occur via PYD-PYD interactions. NLRP3 does not have a CARD domain and requires ASC to interface with the CARD domain present on pro-caspase-1 using a CARD-CARD interaction. NEK7 interacts with the LRR domain on NLRP3. Multiple replicates of these interactions come together and assemble the NLRP3 inflammasome with pro-caspase 1 filaments branching out from the central core of polymerised ASC to cleave proteins.
Ijms 25 09247 g001
Figure 2. Simplified illustration of the signalling pathway depicting priming and activation of the NLRP3 inflammasome. PAMPs and DAMPs bind to their respective PRRs (IL-1Rs, TLRs, TNFRs, NLRs) to activate NF-κB, which upregulates the genes NLRP3, GSDMD, pro-IL-1β, and pro-IL18 within the cytosol. Concurrently, BRCC3 deubiquitinates NLRP3 and IRAK1/4 phosphorylates NLRP3 to also upregulate the priming of NLRP3 to aid inflammasome formation. Activation of NLRP3 is initiated by ion channel dysregulation, ROS accumulation, mitochondrial dysfunction, and lysosome disruption. Together, these generate key signals, including reduced K+ and increased Ca2+, Na+, ATP, ox-mtDNA, and cathepsin B to activate NLRP3 inflammasome formation. Activated inflammasomes will cleave pro-IL-1β, pro-IL18, and GSDMD to active IL-1β, IL-18, and the N-terminal of GSDMD to initiate pyroptosis.
Figure 2. Simplified illustration of the signalling pathway depicting priming and activation of the NLRP3 inflammasome. PAMPs and DAMPs bind to their respective PRRs (IL-1Rs, TLRs, TNFRs, NLRs) to activate NF-κB, which upregulates the genes NLRP3, GSDMD, pro-IL-1β, and pro-IL18 within the cytosol. Concurrently, BRCC3 deubiquitinates NLRP3 and IRAK1/4 phosphorylates NLRP3 to also upregulate the priming of NLRP3 to aid inflammasome formation. Activation of NLRP3 is initiated by ion channel dysregulation, ROS accumulation, mitochondrial dysfunction, and lysosome disruption. Together, these generate key signals, including reduced K+ and increased Ca2+, Na+, ATP, ox-mtDNA, and cathepsin B to activate NLRP3 inflammasome formation. Activated inflammasomes will cleave pro-IL-1β, pro-IL18, and GSDMD to active IL-1β, IL-18, and the N-terminal of GSDMD to initiate pyroptosis.
Ijms 25 09247 g002
Figure 3. Summary of the identified modulatory effects of CO on NLRP3 formation. CO inhibits NF-κB to reduce transcription of NLRP3 and prevent NLRP3 formation. CO also inhibits P2X receptors to reduce K+ efflux. Retention of K+ is associated with reduced Ca2+-independent phospholipase A2 activation, one of the driving signals for NLRP3 activation [69]. CO inhibits TXNIP to prevent binding of NLRP3 and reduce NLRP3 formation. Lastly, CO activates PERK and inhibits IRE1 phosphorylation to reduce NLRP3 activation.
Figure 3. Summary of the identified modulatory effects of CO on NLRP3 formation. CO inhibits NF-κB to reduce transcription of NLRP3 and prevent NLRP3 formation. CO also inhibits P2X receptors to reduce K+ efflux. Retention of K+ is associated with reduced Ca2+-independent phospholipase A2 activation, one of the driving signals for NLRP3 activation [69]. CO inhibits TXNIP to prevent binding of NLRP3 and reduce NLRP3 formation. Lastly, CO activates PERK and inhibits IRE1 phosphorylation to reduce NLRP3 activation.
Ijms 25 09247 g003
Figure 4. S-nitrosylation inhibition of NLRP3 Inflammasome activation. Post-translational covalent attachment of an NO group (to a cysteine thiol) occurs both at the LRR domain of NLRP3 and on pro-caspase-1 to prevent NLRP3 inflammasome formation.
Figure 4. S-nitrosylation inhibition of NLRP3 Inflammasome activation. Post-translational covalent attachment of an NO group (to a cysteine thiol) occurs both at the LRR domain of NLRP3 and on pro-caspase-1 to prevent NLRP3 inflammasome formation.
Ijms 25 09247 g004
Figure 5. Summary of the modulatory effects of H2S on NLRP3 inflammasome formation and activation. H2S inhibits several proteins, including NF-κB to reduce NLRP3 transcription, TNFRs to reduce NF-κB activation, TXNIP to reduce binding to NLRP3, P2X receptors to reduce K+ efflux, and Keap1 to promote translocation of Nrf2 into the nucleus. H2S activates Nrf2 to upregulate antioxidant enzymes and AP1 to upregulate SIRT3 promotor and increase SIRT3 and p62 levels to reduce NLRP3 activation. Green arrows indicate the activating response of H2S.
Figure 5. Summary of the modulatory effects of H2S on NLRP3 inflammasome formation and activation. H2S inhibits several proteins, including NF-κB to reduce NLRP3 transcription, TNFRs to reduce NF-κB activation, TXNIP to reduce binding to NLRP3, P2X receptors to reduce K+ efflux, and Keap1 to promote translocation of Nrf2 into the nucleus. H2S activates Nrf2 to upregulate antioxidant enzymes and AP1 to upregulate SIRT3 promotor and increase SIRT3 and p62 levels to reduce NLRP3 activation. Green arrows indicate the activating response of H2S.
Ijms 25 09247 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Payne, F.M.; Dabb, A.R.; Harrison, J.C.; Sammut, I.A. Inhibitors of NLRP3 Inflammasome Formation: A Cardioprotective Role for the Gasotransmitters Carbon Monoxide, Nitric Oxide, and Hydrogen Sulphide in Acute Myocardial Infarction. Int. J. Mol. Sci. 2024, 25, 9247. https://doi.org/10.3390/ijms25179247

AMA Style

Payne FM, Dabb AR, Harrison JC, Sammut IA. Inhibitors of NLRP3 Inflammasome Formation: A Cardioprotective Role for the Gasotransmitters Carbon Monoxide, Nitric Oxide, and Hydrogen Sulphide in Acute Myocardial Infarction. International Journal of Molecular Sciences. 2024; 25(17):9247. https://doi.org/10.3390/ijms25179247

Chicago/Turabian Style

Payne, Fergus M., Alisha R. Dabb, Joanne C. Harrison, and Ivan A. Sammut. 2024. "Inhibitors of NLRP3 Inflammasome Formation: A Cardioprotective Role for the Gasotransmitters Carbon Monoxide, Nitric Oxide, and Hydrogen Sulphide in Acute Myocardial Infarction" International Journal of Molecular Sciences 25, no. 17: 9247. https://doi.org/10.3390/ijms25179247

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