Next Article in Journal
Regulation of IDO2 by the Aryl Hydrocarbon Receptor (AhR) in Breast Cancer
Next Article in Special Issue
Multi-Drug Cocktail Therapy Improves Survival and Neurological Function after Asphyxial Cardiac Arrest in Rodents
Previous Article in Journal
Genome-Wide Analysis of MYB Transcription Factors in the Wheat Genome and Their Roles in Salt Stress Response
Previous Article in Special Issue
Newly Developed Di-Block Copolymer-Based Cell Membrane Stabilizers Protect Mouse Coronary Artery Endothelial Cells against Hypoxia/Reoxygenation Injury
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pharmacological Cardioprotection against Ischemia Reperfusion Injury—The Search for a Clinical Effective Therapy

1
Department of Anesthesiology—L.E.I.C.A., Amsterdam University Medical Centers, Location AMC, Cardiovascular Science, Meibergdreef 11, 1105 AZ Amsterdam, The Netherlands
2
Department of Anesthesiology, Kerckhoff-Clinic-Center for Heart, Lung, Vascular and Rheumatic Disease, Justus-Liebig-University Giessen, Benekestr. 2-8, 61231 Bad Nauheim, Germany
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Cells 2023, 12(10), 1432; https://doi.org/10.3390/cells12101432
Submission received: 6 April 2023 / Revised: 10 May 2023 / Accepted: 17 May 2023 / Published: 20 May 2023

Abstract

:
Pharmacological conditioning aims to protect the heart from myocardial ischemia-reperfusion injury (IRI). Despite extensive research in this area, today, a significant gap remains between experimental findings and clinical practice. This review provides an update on recent developments in pharmacological conditioning in the experimental setting and summarizes the clinical evidence of these cardioprotective strategies in the perioperative setting. We start describing the crucial cellular processes during ischemia and reperfusion that drive acute IRI through changes in critical compounds (∆GATP, Na+, Ca2+, pH, glycogen, succinate, glucose-6-phosphate, mitoHKII, acylcarnitines, BH4, and NAD+). These compounds all precipitate common end-effector mechanisms of IRI, such as reactive oxygen species (ROS) generation, Ca2+ overload, and mitochondrial permeability transition pore opening (mPTP). We further discuss novel promising interventions targeting these processes, with emphasis on cardiomyocytes and the endothelium. The limited translatability from basic research to clinical practice is likely due to the lack of comorbidities, comedications, and peri-operative treatments in preclinical animal models, employing only monotherapy/monointervention, and the use of no-flow (always in preclinical models) versus low-flow ischemia (often in humans). Future research should focus on improved matching between preclinical models and clinical reality, and on aligning multitarget therapy with optimized dosing and timing towards the human condition.

Graphical Abstract

1. Introduction

The present article describes the pivotal cellular processes occurring during ischemia and reperfusion that contribute to acute IRI and provides an up-to-date overview of novel and promising interventions that target these processes. We focus on cardiomyocytes and the endothelium. The review highlights the need to address the limited translatability of findings from basic research to clinical practice, which may be attributed to factors such as the absence of comorbidities, concurrent medications, and peri-operative treatments in preclinical animal models. In order to bridge this gap, future research should prioritize improved alignment between preclinical models and clinical reality, considering factors such as comorbidities, comedications, and peri-operative treatments.
The review focuses purposely only on novel pharmacological interventions, thus not on any “ischemic conditioning” protocol or drugs that have been tried already in several clinical trials, such as, e.g., cyclosporine A, nitrite, antioxidants, or gene therapies. It is also not the intention of this review to cover all IR mechanisms and all protective strategies reported in the literature. Such an extensive review was already recently published [1]. Additionally, we mainly focus on novel compounds for which the on-target IR mechanisms are clearly defined, such as NAD+ precursors to restore the IR-induced depletion of NAD+, or inhibitors of succinate oxidation (acidic malonate) to reduce the succinate-driven ROS production at early reperfusion. We have left out other novel promising compounds that have a much broader but more undefined mechanism(s) of protection, such as microRNAs [2,3], SGLT2i’s [4,5], and extracellular vesicles [6,7].

2. Mechanisms of Cardiac IRI

Myocardial IRI is, first of all, caused by the alteration of several specific cellular processes during the actual ischemic episode. Early restoration of flow to ischemic tissue, i.e., reperfusion, which is clinically achieved by thrombolytics, percutaneous coronary intervention, or other revascularization modalities, prevents further ischemic cardiomyocyte death, but paradoxically results in further damage of the cardiac tissue [8]. Important cellular processes during both the ischemic and reperfusion period that are known to contribute to acute cardiac IRI will be discussed and summarized in Figure 1. We will mainly focus on the first hours of cardiac IRI because, in this period, most of the damage occurs.
During ischemia, the lack of oxygen causes ΔGATP to decrease, resulting in increased acidosis and the build-up of intracellular Na+ and, consequently, Ca2+. In the mitochondria, succinate accumulates from the breakdown of TCA intermediates and glycogen. The decreased pH, together with the build-up of G6P from glycogen breakdown, detach HKII from mitochondria, facilitating mitochondrial ROS production. The Ca2+-activated NOX, together with eNOS uncoupling driven by BH4 oxidation, increase cytosolic ROS production. Mitochondrial acylcarnitines accumulate due to the activation of CPT1. At reperfusion, the efflux of H+ through reactivation of NHE results in a burst of cytosolic Na+ and Ca2+. Accumulated succinate drives mitochondrial ROS production through RET. HKII detachment from mitochondria, high mitochondrial acylcarnitine inhibition of the ETC, and Ca2+-induced NOX contribute to ROS production. Following pH normalization, the high ROS and Ca2+ induce mPTP opening, allowing NAD+ and CytC to escape from mitochondria and dissipating mitochondrial membrane potential. NAD+ is degraded by NADases into the Ca2+-mobilizing second messenger cADPR, and cytC activates apoptosis. Additionally, high Ca2+ causes hypercontracture, calpain, and caspase activation. Acytel-CoA = acetyl coenzyme A; ADP = adenosine diphosphate; ATP = adenosine triphosphate; BH4 = tetrahydrobiopterin; cADPR = cyclic ADP-ribose; Casp = caspase; Cyt C = cytochrome C; CTP1 = carnitine palmitoyltransferase 1; eNOS = endothelial nitric oxide synthase; FAD = flavin adenine dinucleotide; FADH2 = reduced flavin adenine dinucleotide; Gly = glycogen; G6P = glucose 6-phosphate; HK2 = hexokinase 2; mPTP = mitochondrial permeability transition pore; NAD+ = oxidized nicotinamide adenine dinucleotide; NADH = reduced nicotinamide adenine dinucleotide; NCX = Na+/Ca2+ exchanger; NHE = Na+/H+ exchanger; NKA = Na+/K+-ATPase; NLRP3 = nucleotide-binding domain, leucine-rich–containing family, pyrin domain–containing-3; NOX = nicotinamide adenine dinucleotide phosphate oxidase; RET = reverse electron transfer; ROS = reactive oxygen species; SERCA = sarco-endoplasmic reticulum calcium ATPase; SR = sarcoplasmic reticulum; 4-HNE = 4-hydroxynonenal; 3-NT = 3-nitrotyrosine; and α kg = alpha ketoglutarate.

2.1. Decreased ΔGATP, Acidosis, and Ion Disturbances as Initial Processes Driving IRI

Myocardial IRI starts from the onset of ischemia. During ischemia, cardiomyocyte metabolism shifts from oxidative phosphorylation to anaerobic glycolysis because of the lack of oxygen and nutrients. The increased glycolytic ATP turnover, together with the diminishing mitochondrial ATP synthesis, CO2 accumulation, and net ATP breakdown, increases the build-up of H+, resulting in intracellular acidosis [9]. Ion homeostasis is largely dictated by thermodynamic control of the active ion pumps (e.g., the sodium/potassium ATPase (NKA pump)) and exchangers (e.g., the sarcoplasmic–endoplasmic reticulum calcium-ATPase (SERCA)), i.e., determined by the amount of free energy that can be liberated during ATP hydrolysis (ΔGATP). During the first ten minutes of ischemia, a first initial small drop in ΔGATP that starts to inhibit the pumps, and that causes intracellular Na+ to increase and intracellular K+ to decrease, is observed [10]. Then, following the depletion of glycogen and halting of anaerobic glycolysis, ΔGATP starts to fall much more, driving further increases in Na+ and decreases in K+ [10]. The increase in Na+ drives the rise in intracellular Ca2+ by impairing the normal forward mode action (exchanging intracellular Ca2+ for extracellular Na+) of the sodium–calcium exchanger (NCX). Ischemic contracture starts to develop because of rising cytosolic calcium and lack of ATP [11]. Intracellular acidosis may further increase intracellular Na+ load through activation of the sodium–hydrogen exchanger (NHE), although the quick development of extracellular acidosis during ischemia will start to inhibit NHE activity [12,13]. At reperfusion, extracellular acidosis is quickly normalized, reactivating NHE to supranormal activity [12] to restore intracellular pH at the expense of further increasing intracellular Na+. The rise in intracellular Na+ further impairs or even reverses the NCX, extruding Na+, now at the expense of a large Ca2+ overload.
This burst in cytosolic Ca2+ within the first minutes of reperfusion, together with the fast normalization of intracellular pH, is one of the crucial mediators of IRI. The high calcium level facilitates (1) opening of the mitochondrial permeability transition pore (mPTP), dissipating mitochondrial potential and thereby ATP synthesis, causing swelling and rupture of mitochondria and consequently cardiac cells; (2) activation of calpains, proteolytic proteases that breakdown the cell, and activates caspases through dislodgement of these programmed cell death enzymes from the cytoskeleton, resulting in pyroptosis through an NLRP3 dependent or independent mechanism [14,15]; and (3) hypercontracture of the myofilaments and thereby mechanical rupture of cardiac tissue, also resulting in cell death [9,16].

2.2. Oxidative Stress Involvement in IRI

Part of IR-associated oxidative stress develops during the ischemic period, as demonstrated by techniques allowing for measurements of direct ROS production [17,18,19,20]. These findings are supported by end-ischemic measurements of cardiac 4-hydroxy-2-nonenal (4-HNE), a breakdown product of lipid peroxidation, and 3-nitrotyrosine, a marker of nitrated tyrosine due to peroxinitrite, both specific biomarkers of oxidative stress [21]. Ischemic ROS mostly develops during the initial period of ischemia, but not during later ischemia [21,22,23]. Interestingly, once 4-HNE is generated during ischemia, no further 4-HNE is produced during reperfusion, and reperfusion in itself cannot reduce the ischemic build-up of 4-HNE. Sources of ROS during ischemia include increasing cytochrome p450 activity, mitochondrial electron transport damage and uncoupling, NAD(P)H oxidase (NOX) activity, and uncoupled nitric oxide synthase (NOS) activity [24]. In addition, oxidative damage during ischemia can also be a result of a decreased antioxidant capacity in the ischemic heart, creating a more pro-oxidant environment. Reduced glutathione (GSH) is an important endogenous antioxidant in the heart that is rapidly oxidized during ischemia. Restoration into GSH needs ATP, which becomes scarce.
Next to ischemic ROS, a burst of ROS is also initiated during the first minutes of reperfusion [17,25,26]. A major contributor to the ROS burst during reperfusion is the oxidation of the ischemic accumulated succinate (see Section 2.3.1 below) [27]. Although two-thirds of the ischemic succinate is released from the heart upon reperfusion, the remaining one-third is oxidized during the first five minutes of reperfusion. At the end of ischemia, the heart is already in a highly reduced state, with maximal high levels of NADH and a maximally reduced ubiquinone (Q) pool. At this early stage of reperfusion, succinate is rapidly oxidized to fumarate by succinate dehydrogenase (SDH), trying to further reduce the already highly reduced Q pool. Because the Q pool is already maximally reduced, electrons are forced backward (reverse electron transport: RET), and O2 is partially reduced to O2− at complex I [27]. Other contributors to reperfusion ROS are activated NOX2/4 [28], highly charged (high ψm) mitochondria due to detachment of Hexokinase II (HKII) from mitochondria [26,29,30,31], interrupted oxidative phosphorylation due to ischemic accumulation of long-chain acylcarnitines (see below Section 2.3.3) [32,33], and uncoupled endothelial NOS (eNOS) [24,34].
ROS produced during early reperfusion primes the mPTP for the high cytosolic calcium, permanently opening this mitochondrial pore and causing cell death [35]. ROS and nitrative stress, both during ischemia and reperfusion, also oxidizes and/or nitrates several proteins, thereby hampering their function and contributing to IRI.

2.3. Involvement of Intermediate Metabolites in IRI

IRI of the heart can be considered a metabolic pathology with abruptly halted metabolism during ischemia, and is aggravated by a sudden restart of specific metabolic pathways trying to recover metabolic homeostasis at reperfusion [36]. Variations in cellular levels of metabolic intermediates during IR underlie part of the mechanisms contributing to the development of IRI. Here, we describe changes in metabolic intermediates that are now known to affect cardiac IRI.

2.3.1. Succinate

It has been known since 1978 [37] that succinate increases in the heart during ischemia. More recent research demonstrated that oxidation of ischemic succinate during the first minutes of reperfusion contributes largely to early reperfusion ROS and, therefore, IRI [27]. Originally it was suggested, by employing isolated mitochondria as a model, that ischemic succinate was generated from fumarate through mitochondrial complex II reversal [27]. However, research using a more physiological model of isolated mouse hearts demonstrated that ischemic succinate is mostly arising from remaining Krebs cycle activity during ischemia, still turning in the normal direction (so-called canonical TCA activity) [38]. The metabolic precursors of ischemic succinate are mostly (75%) intermediates of the Krebs cycle, whereas the remaining 25% is generated through complex II reversal (6%), aspartate anaplerosis (10%) via aspartate aminotransferase (AST), and glutamine anaplerosis (8%) via alanine aminotransferase (ALT) [38]. Increased glycogen breakdown and glycolysis during ischemia indirectly contribute to ischemic succinate accumulation by providing the substrates (pyruvate, NADH) for AST and ALT. Actually, the depletion of glycogen before ischemia resulted in a 40% decrease in ischemic succinate, demonstrating the importance of glycogen during ischemia for the generation of ischemic succinate [38]. Of note, the generation of succinate during ischemia can also be protective against IRI, as it provides energy in the form of GTP through substrate-phosphorylation by succinyl-coenzyme A (CoA) synthetase. Those beneficial effects were reflected by delayed contracture with increased succinate during ischemia [38]. Thus, impairing succinate generation during ischemia is a double-edged sword, whereas impairing its oxidation during early reperfusion is mostly protective against IRI.

2.3.2. Glycogen and Glucose-6-Phosphate

Several glycolytic intermediates and end-products accumulate during ischemia from the breakdown of glycogen, such as glucose-1-phosphate, glucose-6-phosphate (G6P), fructose-6-phosphate, pyruvate, and lactate [39,40]. Although for several of these intermediates, it remains to be elucidated whether they modulate IRI; elevated G6P levels are detrimental. The increase in G6P, together with acidosis during ischemia, will detach HKII from the mitochondria, thereby increasing infarct size [41,42]. The amount of HKII bound to mitochondria (mtHKII) is a strong determinant of infarct size [41], making treatments directed at maintaining mtHKII a valid target for cardioprotection [43]. The strongest cardioprotective intervention to date, ischemic preconditioning (IPC), was shown to be associated with differences in mtHKII, but not with, e.g., ischemic succinate accumulation [41,42,43,44,45]. Maintaining HKII at the mitochondria may provide protection through several mechanisms: (1) reduction in the mitochondrial membrane potential to lower mitochondrial ROS production [29], (2) activation of glucose metabolism with attenuation of mitochondrial activity [46], (3) impairment of mPTP opening [47], and (4) prevention of ischemic mitochondrial ATP hydrolysis by reversed ATPase [48], thereby also attenuating anaerobic glycolysis and acidosis during ischemia. Pretreating hearts with insulin may therefore increase cardiac IRI, because insulin will load the heart with glycogen, which, once totally depleted because of prolonged ischemia, increases G6P and acidosis, detaching HKII from mitochondria and increasing infarct size [49]. However, when total glycogen depletion is prevented, for example, due to such short ischemia that glycogen is not totally depleted during ischemia, insulin pretreatment is protective because insulin per se increases cardiac mtHKII [50] and the increased glycogen provides additional glycolytic ATP during ischemia [51]. Therefore, glycogen is also considered a double-edged sword during IRI: protective when not completely depleted, detrimental when all glycogen is broken down during ischemia [51].

2.3.3. Acylcarnitines

Older literature has reported the accumulation of acyl-CoA and acylcarnitine during ischemia [52,53]. Recent studies demonstrated especially the accumulation of long-chain acylcarnitines to contribute to cardiac IRI [32,36]. Acylcarnitines accumulate during ischemia because carnitine palmitoyltransferase 1 (CPT1, the enzyme catalyzing the transfer of acyl-coA to carnitine to form acylcarnitine) becomes highly activated due to the decrease in its natural inhibitor malonyl-CoA [54]. The enzyme-generating malonyl-CoA, acetyl-CoA carboxylase, is inhibited due to the ischemia-induced increase in phospho-AMPK. In turn, CPT1 generates large amounts of long-chain acylcarnitines that cannot be metabolized by mitochondria due to the lack of oxygen [32]. This detrimental effect of high cardiac levels of acylcarnitines on IRI is especially present in the fasted condition, when acylcarnitines can accumulate five times more than in the fed condition [55]. Mechanistically, the presence of high levels of long-chain acylcarnitines inhibits oxidative phosphorylation (OXPHOS) during early reperfusion, thereby inducing mitochondrial membrane hyperpolarization and contributing to mitochondrial ROS production [56] to increase cardiac IRI. Thus, the detrimental effects of long-chain acylcarnitines become apparent during reperfusion, when the supply of oxygen drives ROS from acylcarnitine-impaired mitochondria [33].

2.4. Cofactors Involved in IRI

Tetrahydrobiopterin (BH4): For nitric oxide (NO) synthesis, eNOS requires the redox-sensitive cofactor BH4. However, this cofactor is continuously oxidized during ischemia, such that at 30 min of ischemia, 85% of BH4 is already irreversibly degraded [34]. As a consequence, eNOS becomes uncoupled, and the enzyme starts to generate ROS (•O2−) instead of NO. It was estimated that ROS produced by uncoupled eNOS following 30 min of ischemia and reperfusion contributed to 30% of the total ROS detected [55]. Restoring BH4 directly by supplying BH4 in the liposomal formulation [34], or indirectly through, e.g., folic acid administration, protected hearts against IRI [57,58].
Nicotinamide adenine dinucleotide (NAD+): NAD+ is a critical cofactor involved in more than 500 enzymatic reactions impacting cellular metabolism, inflammation, energetics, and cell survival, with decreased levels related to a diverse area of pathologies such as aging, diabetes, metabolic diseases, heart failure, and IRI [59,60,61]. In the heart, 72% of all NAD+ is localized in the mitochondrial matrix [62]. During cardiac ischemia, about 30% of NAD+ is lost, which is largely increased to 70% during reperfusion [62]. The loss of NAD+ during reperfusion can be explained by the opening of the mitochondrial permeability transition pore (mPTP), a phenomenon mainly occurring during reperfusion and allowing NAD+ to escape the mitochondria. Outside the mitochondria, NAD+ is broken down by NADases (e.g., the glycohydrolase enzyme CD38) [62,63] to generate Ca2+-mobilizing second messenger cyclic ADPR (cADPR) to raise Ca2+ and thus contribute to IRI. Inhibiting the enzyme aldose reductase of the polyol pathway may be another means to conserve NAD+ during IR by attenuation of NAD+ use through reduced sorbitol dehydrogenase activity. The preserved NAD+ is then used for glyceraldehyde 3-phosphate dehydrogenase in glycolysis, facilitating increases in glycolytic ATP synthesis [64,65,66]. Major mechanisms explaining increased cardiac IRI with decreased NAD+ relates to the loss of function of enzymes that need NAD+ as a cofactor, such as sirtuins (deacetylation) and glycolytic enzymes. Decreased sirtuin and glycolytic activities are known to increase cardiac IRI [36,67,68].

3. Novel Pharmacological Strategies Targeting IRI in Cardiomyocytes

Drugs modulate their cardioprotection effects through different upstream routes; however, their final end-effector protective mechanisms culminate in the attenuation of Na+ and Ca2+ overload, inhibition of ROS production, and mPTP opening, precipitating reductions in cell death [69]. We now focus on novel pharmacological treatments that mainly target the mechanisms discussed above in Section 2.

3.1. NAD+ Precursors

NAD+ precursors that have been tested to prevent cardiac IRI ex vivo or in vivo are nicotinamide mononucleotide (NMN) and nicotinamide riboside (NR). One of the first studies examining NAD+ precursors in the setting of cardiac IRI [67] reported protection with 500 mg/kg NMN intraperitonally (i.p.), either once 30 min before ischemia, or four times just before and during 24 h of reperfusion. Mechanisms of protection by NMN precursors relate to the activation of sirtuins, glycolysis, and/or autophagy [67,68,70,71]. Surprisingly, few studies have examined NR as an NAD+ precursor in cardiac IRI studies, which is somewhat surprising given NR’s superiority in elevating NAD+ in humans [72]. Pre-ischemic administration (50 mg/kg i.v.) of NR in an in vivo rat model of cardiac IRI employing clinically relevant anesthesia reduced infarct size [73]. Subsequently, employing isolated mouse hearts, we also demonstrated that NR’s protection against cardiac IRI, similar to NMN [68], is mediated through glycolysis activation [74]. Thus, it seems that NAD+ precursors offer novel opportunities for acute protection against cardiac IRI. Although studies in isolated mitochondria have suggested protection through direct inhibition of mPTP by NAD+ [75], studies in intact hearts or in vivo favor activation of glycolysis as the major protective mechanism of these compounds.

3.2. Malonate

Malonate is a three-carbon dicarboxylic acid and a competitive succinate dehydrogenase (SDH) inhibitor, and as such, it inhibits succinate oxidation during early reperfusion [27]. Valls-Lacalle et al. found that disodium malonate reduced infarct size without increasing the incidence of ventricular fibrillation in pigs with left anterior descending coronary artery occlusion. No malonate was detected in distant myocardium or in plasma [76]. The authors attributed the observed cardioprotective effect of disodium malonate to intracoronary administration, which is a clinically feasible method to achieve selective delivery of cardioprotective treatments. However, other studies reported a lack of malonate protection, especially under diabetic conditions [77,78]. Although the reasons for this failure are not completely clear, it could have been partly explained by the absence of malonate in the hearts at the first minutes of reperfusion due to the experimental set-up. Recently, data showed that disodium malonate protects in an ischemia-selective way, i.e., only ischemic tissue with decreased pH and increased lactate will take up disodium malonate [79]. Accumulation of lactate and protons in ischemic tissue facilitates the protonation of malonate to its monocarboxylate form, which is exchanged for lactate through the monocarboxylate transporter 1 (MCT1) to quickly enter cardiomyocytes upon early reperfusion. Then, malonate is transported into mitochondria by the mitochondrial dicarboxylate carrier and subsequently inhibits succinate oxidation. Future research should focus on this type of malonate in preclinical models considering conditions and drugs that are present under clinical conditions (see Section 4).

3.3. NLRP3 Inflammasome Inhibitors

The innate immune system is the first line of defense against exogenous (invading pathogens) or endogenous (sterile, such as traumatic impact, metabolic, or IR) stress signals, and is constituted of pattern-recognition receptors (PPRs) that are localized either on the plasma membrane and endogenous endosomes (Toll-like receptors; TLRs) or within the cytosol (Nucleotide-binding domain and Leucine-rich repeat-containing proteins; NLRs). Although TLR inhibition certainly holds promise for reducing cardiac IRI [80,81], the development of NLR inhibitors is currently more prominent [82].
The most prominent NLR suggested to be involved in acute cardiac IRI and for which inhibitors are being developed and tested is the NLRP3 inflammasome. NLRP3 inflammasome activity is regulated through expression and then oligomerization of three components: ASC, NLRP3, and caspase-1 [83]. NLRP3 is hardly expressed in healthy hearts, explaining the lack of cardioprotection following NLRP3 gene deletion [84,85]. However, stress and inflammation lead to NLRP3 expression in the heart. Stress induction can occur (1) during extensive surgery needed for opening the chest, (2) in diseased animals (diabetic, heart failure, metabolic syndrome, and aging), (3) in the presence of pathophysiology (low blood pressure, hypoxia, ischemia, and metabolic stress), or (4) examining hearts at late (>3 h) reperfusion. Impairing NLRP3 inflammasome activity through ASC deletion decreased infarct size after two days of reperfusion [86]. However, deletion of ASC or NLRP3 had no effect on IRI when studied between 30 min and 3 h of reperfusion [84,85]. Similarly, cardioprotection by NLRP3 inhibition was only observed at 24 h but not at 3 h of reperfusion [87]. Moreover, NLRP3 inhibition only reduced infarct size when administered 1 h after reperfusion, while its cardioprotective effect was lost when treatment was delayed to 3 h of reperfusion [87]. These findings highlight that the therapeutic window of NLRP3 inflammasome inhibition is limited to the first hours of reperfusion.
In a large animal model of cardiac I/R, the NLRP3 inhibitor MCC950 administered during the first 7 days of reperfusion exerted a mild protective effect on infarct size [88]. However, cardioprotection was not observed with the novel NLRP3-inflammasome inhibitor IZD334 [89]. No clinical trials have been published reporting on the use of specific NLRP3 inflammasome inhibitors in the setting of acute IRI. Only the old and aspecific anti-inflammatory drug colchicine, which inhibits the assembly of the NLRP3 inflammasome, was recently evaluated in a clinical trial. Colchicine was administered immediately at reperfusion and during the first five days of reperfusion in first-time ST-segment-elevation myocardial infarct patients [90]. However, no cardioprotection was observed. Although interleukin-6 and high-sensitivity C-reactive protein concentrations after PCI were decreased by colchicine in another study, the myocardial injury did not differ between the treatment and placebo groups [91].
In conclusion, although preclinical research shows promise for NLRP3 inhibitors to reduce cardiac IRI in settings of diseased hearts (diabetic, aged, and metabolic syndrome), and novel NLRP3 inflammasome inhibitors are being developed, the first cardioprotective trial with these selective NLRP3 inhibitors is still eagerly awaited [91].

3.4. Caspase and Calpain Inhibitors

Caspases, or cysteine-aspartic acid proteases, are cysteine endoproteases that attack and cleave a protein only after an aspartic acid residue. Caspase-3, -6, -7, -8, and -9 are mainly involved in apoptosis, whereas caspase-1, -4, -5, and -12 (in humans) and caspase-1, -11, and -12 (in mice) are mainly involved in inflammatory pathways [92].
Although apoptosis is important for heart development, a resistance to caspase-dependent apoptotic cell death has been detected in differentiated cardiomyocytes [93]. In addition, caspase-3/caspase-7 double-knockout mice showed no myocardial protection during the acute phase of reperfusion [94]. Additionally, in terminally differentiated mouse myocardium, caspase-3, -6, and -7 are silenced [95]. Directly comparing necrosis and apoptosis in an isolated rabbit heart employing different durations of ischemia, it was observed that necrosis and infarct size (12–23% of heart tissue) was 6–8 times larger than apoptosis (2–3% of cells) [96]. Controversial data have been reported for selective inhibitors of caspase-3 or -9, with one study reporting significant reductions in infarct size [97], whereas another study reported no effects on infarct size [98]. It is possible that the selective caspase inhibitors are not so selective and also partly inhibit, e.g., calpains (see below). Those results demonstrate that pharmacological treatment targeting apoptosis-related caspases does probably not play a major role in acute cardiac IRI.
In contrast, inhibition of caspase-1 and -4 by VRT-043198 (VRT) or emricasan did reduce acute cardiac IRI in both rats and mice [15,99,100]. The degree of protection obtained with the specific caspase-1 and -4 inhibitor VRT was similar to the protection observed with the pan-caspase inhibitor emricasan, again providing evidence that apoptosis does not contribute significantly to acute cardiac IRI [15]. Interestingly, specific calpain inhibition offered similar protection as that observed with VRT, and adding VRT to calpain inhibition did not further reduce infarct size, suggesting that the activation of calpain during early reperfusion is needed for caspase-1-induced infarct size reduction [15]. Calpains are activated by cytosolic increases in Ca2+, one of the two major signaling events (besides ROS production) during the first minutes of reperfusion that dictate acute cardiac IRI (see sections above). Indeed, calpain silencing ameliorated acute cardiac IRI [101]. Although caspase-1 activation has originally been thought to be a consequence of NLRP3 inflammasome activation, recent research showed that calpains can activate caspase-1 through dislodging procaspase-1 from the actin-filament network and transformation of its active caspase-1 enzyme through autoactivation [102]. The active caspase-1 can then break down gasdermin D into its lytic N-terminal fragment to make holes in the plasma membrane to kill the cell. Indeed, gasdermin D knockout hearts were protected against acute cardiac IRI [103]. This process of cell death has been called pyroptosis, and because of its significance in acute cardiac IRI, it can be considered a high-potential target for reperfusion injury therapy.

4. Translating Preclinical Cardioprotection into the Clinical Arena: Role of Risk Factors, Comorbidities, Comedications, Peri-Operative Care, and Ischemia Duration

One of the main obstacles to successful translation from the preclinical towards the clinical condition has been the omission of patient risk factors, comorbidities, comedications, and peri-operative treatments in preclinical models of cardiac IRI. Risk factors and comorbidities (aging, diabetes, hyperglycemia, sex, metabolic syndrome, and hypertension), comedications (statins, β blockers, metformin, GLP-1 agonists, and SGLT2is), and peri-operative treatments (heparin, aspirin, P2Y12 platelet inhibitors, nitroglycerine, opioids, benzodiazepines, and propofol) can all abrogate cardioprotective interventions. Most of these factors have been recently summarized in comprehensive reviews [1,104]. Interestingly, recent work revealed a possible mechanism by which comorbidities such as prediabetes, hyperglycemia, and metabolic syndrome may abrogate cardioprotective interventions such as insulin treatment or ischemic preconditioning [105]. This mechanism entailed enhanced nitration of caveolin-3 at tyrosine 73 by peroxinitrite formed from increased NADPH oxidase-induced ROS and increased iNOS-induced NO, thereby disruption the Cav3 signalosome needed for insulin sensitivity, resistance to myocardial ischemia, and several cardioprotective signaling pathways [105]. Therefore, decreasing nitration through inhibition of iNOS or NOX activity, e.g., through the use of SGLT2 inhibitors that have been reported to reduce ROS and increase NO [106,107], may help to restore the cardiac intrinsic pathways of protection.
An additional, often neglected factor that also determines the effectiveness of certain protective interventions, but largely deviates between preclinical and clinical cardiac IRI, relates to the duration and/or severity of the ischemic insult [108]. Whereas in preclinical models duration of ischemia is most often between 25 and 50 min with 100% obstructed coronary flow, in clinical models, the ischemic insult before the start of clinical treatment commonly lasts between 150 and 250 min of still-lingering (<10%) coronary flow [109,110,111,112]. Although for both infarcts, the size amounts to approximately 50% of the area at risk, it is likely that underlying cellular mechanisms causing cell death differ between short-term no-flow and long-term low-flow conditions. In addition, the sensitivity of cardioprotective interventions depending on ischemia duration may differ. For example, reducing ischemic-induced oxidative stress and boosting eNOS activity mostly provide protection against the short duration of no-flow ischemia (<30 min), whereas ischemic preconditioning, postconditioning, mPTP inhibition, opioids, sevoflurane, and metoprolol mainly protect against no-flow ischemia of 30–60 min duration. Conversely, mPTP inhibition and postconditioning can even increase IRI when applied to combat short no-flow ischemia, whereas none of these cardioprotective interventions protected against no-flow ischemia of >60 min [108]. From this perspective, devising a multitarget protective reperfusion strategy may be optimal [113]. Once an optimal strategy has been developed, step-by-step criteria for IMproving Preclinical Assessment of Cardioprotective Therapies (‘IMPACT’) should be met to improve the likelihood of translating novel cardioprotective interventions to the clinical setting [114].

5. Volatile Anesthetics and Noble Gases for Cardioprotection against IRI

5.1. Volatile Anesthetics

Nowadays, the volatile anesthetics isoflurane, sevoflurane, or desflurane are generally used for balanced anesthesia in the clinical setting. Cardioprotective properties of isoflurane were first described by Warltier et al., who showed infarct size reduction in the dog heart in vivo in the isoflurane-treated group [115]. Subsequently, several studies demonstrated cardioprotection by isoflurane against IRI by preconditioning and/or postconditioning [116,117,118]. Experimental studies tried to identify potential comorbidities, as well as concomitant use of medications in their models. Infarct size reduction by isoflurane was shown to be abolished by aging, but protection was restored by employing the antioxidant TEMPOL. Interestingly, when autophagy and mitophagy were inhibited, the cardioprotective effect of isoflurane was blocked not only in young rats but also in old rats treated with TEMPOL [119]. The importance of mitophagy and autophagy in the context of cardioprotective interventions in myocardial infarction is not yet clear [120]. The release of ROS generated after IR is inhibited by mitophagy. Furthermore, mitophagy plays a role in conditioning strategies but decreases with age, making mitophagy an interesting target for cardioprotective interventions in the aged myocardium [121]. Isoflurane-induced preconditioning inhibits cardiomyocyte autophagy by phosphorylation of p38 MAPK accompanied by a decreased expression of nucleotide-binding oligomerization domain containing 2 (NOD2) [122]. Autophagy protects against IRI; however, exaggerated autophagy has a decisive role in reperfusion injury-mediated myocardial dysfunction [120].
In the past, research on cardioprotection focused on mitochondria as a key element in the signal transduction pathway of conditioning strategies. Regarding conditioning effects on mitochondrial function, Xu and colleagues showed, irrespective of mitochondrial NO production, mitochondrial state 2 respiration uncoupling with diminished state 3 respiration by isoflurane [123]. As a part of the mPTP, Cyclophilin D (CypD) regulates mPTP by controlling the opening dimension of the pore [124]. In wild-type and CypD knockout mice, mitochondrial state 3 respiration was blocked by isoflurane, and ADP consumption was improved [125]. Isoflurane-induced cardioprotective properties were mediated by microRNA-21 through the signaling pathway containing Protein kinase B (Akt), NOS, and mPTP [126,127]. In contrast to microRNA-21 upregulation, microRNA-23 was shown to be inhibited by isoflurane [128]. Inhibition of microRNA-23 by isoflurane led to the protection of cardiomyocytes against oxidative stress [128]. The role of microRNAs in cardio- protection seems to be diverse and further research is needed to clarify the particular impact.
Cardioprotective effects induced by sevoflurane [129,130] are mediated by different pathways, i.e., Janus kinase (JAK) and signal transducers and activators of transcription (STAT) [131,132]. Janus kinase (JAK) and signal transducers and activators of transcription (STAT) are involved in sevoflurane-induced cardioprotection. JAK/STAT belong to the SAFE pathway. Sevoflurane-induced protection against reperfusion injury influences apoptosis via the SAFE pathway, and postconditioning by sevoflurane was abolished by a selective JAK2 inhibitor [133]. Besides effects on apoptosis, sevoflurane was also shown to have inhibitory effects on autophagy [134], mediated by reducing phosphatidylinositol 3-kinase catalytic subunit type 3 (PI3KC3). Sevoflurane-induced cardioprotection also restores autophagic flux impaired by IR, and this effect was NO-dependent [135]. The protective effects were completely abrogated in the presence of the NOS inhibitor L-NAME. The same results were observed following the administration of chloroquine—a blocker of autophagic flux [135]. Besides the RISK and/or SAFE pathway, other targets involved in cardioprotection were identified, i.e., vascular endothelial growth factor receptor (VEGFR), consisting of three subtypes, namely VEGFR-1, VEGFR-2, and VEGFR-3. VEGFR-1 was shown to be increased by sevoflurane [136], leading to myocardial preconditioning and decreased inflammation [136]. Furthermore, a specific inhibitor of VEGF-1, macrophage migration inhibitory factor-1 (MIF-1), abrogated the protective effect of sevoflurane [136], similar to isoflurane; in sevoflurane-induced protection microRNAs (miRNA) are also involved. A decrease in miRNA-155 expression induced by sevoflurane was associated with an increase in sirtuin 1 (SIRT1), thereby reducing infarct size and inhibiting cardiomyocyte apoptosis [137].
Notably, many patients that might need perioperative cardioprotection suffer from concomitant diseases, i.e., diabetes mellitus. Sevoflurane preconditioning-induced infarct size reduction was not affected in diabetic mice [138], an effect mediated by an AMP-activated protein kinase (AMPK)-independent activation of pro-survival mitogen-activated protein kinase (MAPK) members, whereas in non-diabetic mice, cardioprotection was AMPK-dependent [129]. Aging blocks sevoflurane-induced preconditioning [139] via activation of nuclear transcription factor kappa B (NFkB) regulated genes [139]. Promising cardioprotective effects of volatile anesthetics in diseased myocardium should be followed up by elucidation of the different and/or impaired underlying mechanisms in future research.

5.2. Noble Gases

Volatile anesthetics share some common pathways with other gases, e.g., noble gases. Within the group of six noble gases, xenon and helium have especially gained a great deal of interest regarding their cardioprotective properties [131,140,141,142,143,144]. Those so-called “inert” gases display profound biochemical activity and reduce infarct size in a rabbit IRI model [145]. Noble gases protect the myocardium when given before, during, or even after ischemia [131,140,143]. In contrast to xenon, helium is not an anesthetic, and thus, the cardioprotective effects are not linked to a hypnotic effect of the noble gases.
Xenon has direct effects within the myocardium [131], including differential expression and phosphorylation of a variety of proteins, as well as blockade or activation of various channels. One major target of xenon-induced preconditioning of the heart is protein kinase C (PKC). The phosphorylation and subsequent translocation of the isoform PKC-Ɛ is a key mediator of xenon-induced cardioprotection [146]. Further downstream of PKC-Ɛ, the p38 MAPK and MAPK-activated protein kinase 2 (MAPKAPK-2) are regulated by xenon and connect its preconditioning properties to the cardiac cytoskeleton and actin stress fiber regulation via the small heat shock protein HSP27 [146,147]. Additionally, in vivo studies in rats using 3 × 5 min of xenon conditioning before IRI revealed that ERK (p44/42 MAPK) and the p54/46 MAPK (SAPK/JNK) are differentially regulated by xenon. Only ERK could be identified as a mediator in the cardioprotection by xenon [148]. The activation of PKC has been shown to be initiated by the regulation of mitochondrial KATP channels and PDK-1 in rats and rabbits in vivo [149].
All above-mentioned studies applied early cardioprotection protocols; however, xenon was also shown to be a potent inducer of late cardioprotection in vivo: xenon inhibited progressive adverse cardiac remodeling, contractile dysfunction and reduced the expression of β-myosin heavy chain and periostin proteins up to 28 days after a 60 min coronary artery occlusion in rats [150]. A total of 90 min of right ventricular ischemia followed by 120 min of reperfusion in a porcine model, whereby xenon was administered throughout ischemia and reperfusion, increased right ventricular afterload and myocardial contractility. On the cellular level, mRNA expression of type B natriuretic peptide (BNP) was hampered in the remote area of the left ventricle by xenon [151]. Another potential mediator of xenon-induced late preconditioning was shown in an in vivo late preconditioning model of rats using the COX-2 inhibitor NS-398: cyclooxygenase 2 (COX-2). Xenon, however, did not regulate mRNA expression of COX-2 in these animals, suggesting that xenon-induced late preconditioning is mediated most likely by an increased activity of existing COX-2 rather than a transcriptional upregulation of COX-2 mRNA [152].
Helium is already in clinical use for the treatment of respiratory diseases, and it can be easily applied to awake patients suffering from IRI. Inhalation of three times five minutes of 70% helium before IR reduced infarct size in rabbit hearts [145]. In follow-up studies, the authors administered several different inhibitors blocking a variety of signaling pathways described to be involved in volatile anesthetic and ischemic preconditioning of the heart. Phosphatidylinositol-3-kinase (PI3K), mitogen/extracellular signal-related kinase 1 (MEK-1), the 70-kDa ribosomal protein s6 kinase (p70s6kinase), ROS, KATP channels, and NO are mediators of helium-induced cardioprotection [145,153,154]. Inhibition of glycogen synthase kinase-3beta (GSK-3β) activity and activated apoptotic protein p53 degradation in a model with one, three, or five cycles of helium preconditioning reduced infarct size even more, suggesting a lowered threshold when these proteins are blocked. Most importantly, protection was completely reversed by mPTP opening with atractyloside during helium preconditioning [155]. Cyclosporine A, a selective inhibitor of mPTP, counteracts the blockade of helium conditioning in the presence of mild acidosis during the reperfusion in vivo [156]. Furthermore, morphine, in combination with helium, reduces infarct size after IR, whereby morphine alone had no effect on infarct size in an in vivo rabbit model [157]. As the effect could be inversed by blocking opioid receptors with naloxone, the involvement of these receptors has been suggested [157]. Even a low concentration of 30% helium in a singular dose 24 h before an ischemic insult was cardioprotective, and this effect was mediated most likely by COX-2 [158]. Regarding helium post-conditioning in healthy animals, 15 min of helium inhalation already protected the myocardium, but a prolonged inhalation for up to 60 min had no effect [159].
Most of these early studies on helium-induced cardioprotection did not investigate direct effects on the expression, phosphorylation, or activation of proteins. Oei et al. identified several important genes that where either up- or down-regulated by 15 min of helium postconditioning. The authors used gene expression arrays and found genes regulating necrosis, apoptosis (pro- and anti-), and autophagy either up- or down-regulated, showing a distinct pattern of gene regulation in the myocardium during helium postconditioning [160].
The central targets for helium-induced cardioprotection studied are mitochondria and the mPTP. Heinen et al. focused on mitochondrial calcium-sensitive potassium (mKCa) channels and changes in mitochondrial respiration by helium [161]. By measuring the rate of oxygen consumption in isolated mitochondria of young and old rats after helium conditioning, they showed that helium reduced the respiratory control index (state 3/state 4) only in young animals. Furthermore, the blockade of mKca channels by Iberotoxin abolished the protective effect of helium on infarct size reduction and respiratory control only in young animals [161].
In young and old rats, PKA was reported as an upstream target of mKCa channels. Interestingly, activation of mKCa channels in both young and old animals by NS1619 reduced infarct size, pointing to a pivotal role of these channels also in aged myocardium [162]. The adenylyl cyclase activator forskolin in a dosage of 300 μg/kg was only able to reduce infarct size in young animals. However, a much higher dosage of 1000 μg/kg was effective even in old rats, suggesting that the upstream regulation of mKCa channels by PKA might be a critical step for the age-dependent loss of helium-induced cardioprotection [162].
The influence of hypertension on helium conditioning was investigated in spontaneously hypertensive rats that were subjected to 25 min of ischemia followed by 120 min reperfusion using three different conditioning protocols in vivo. One group of animals received 70% helium post-conditioning (15 min after the index ischemia), one group received post-conditioning in combination with helium late preconditioning (application of helium 24 h before the IR procedure), and the third group received an additional 3 × 5 min cycle of preconditioning shortly before IR on top of the second group [163]. Only the triple intervention could effectively protect spontaneously hypertensive rats against IRI.
Diabetic disorders prevent helium-induced early preconditioning and postconditioning [163]. In pre-diabetic obese Zucker rats, no helium-induced mitochondrial uncoupling could be observed. On the contrary, in the non-diabetic Zucker lean rats, mild mitochondrial uncoupling of oxygen consumption was detected [163]. Among the molecular targets investigated in this study, helium only affected GSK-3β phosphorylation in Zucker lean animals. Other important targets, such as ERK 1/2 and Akt phosphorylation, were surprisingly not regulated in both animal types [163].
Helium affects many intracellular proteins without receptor binding in the heart. Caveolins, small proteins that are embedded in the Caveolae of the cell membrane, are supposed to be a target for Helium. As Caveolins have a scaffolding domain that has been described to bind several proteins involved in helium-induced cardiac protection (e.g., Src kinases, PI3K, eNOS, PKC isoforms, and ERK), a regulation of Caveolins by helium seems likely [164,165,166].
Two of the isoforms of Caveolin (Caveolin 1 and 3) are critically involved in helium postconditioning [167]. After helium administration, an increased level of Caveolin 3 was found in the plasma of rats. Furthermore, both isoforms were upregulated in the infarcted area of rat hearts [167]. Short (5 min) administration of helium before cardiac arrest leads to differential expression patterns of both isoforms in the myocardium and reduced cardiac apoptosis [168].
Employing isolated Langendorff-perfused mice hearts, cardioprotection by helium could not be detected [169]. The lack of any circulating blood in the Langendorff system was assumed to be one explanation for this surprising outcome. In follow-up experiments, the membrane fractions showed a decrease in Caveolin 1 and 3 expressions, whereby both isoforms were elevated in the platelet-free plasma of the mice [169]. Secretion of caveolins into the blood stream following helium inhalation is suggested.
In isolated neonatal rat cardiac fibroblasts, helium conditioning induces migration of cardiac fibroblasts and might thereby mediate cardioprotection. However, this effect was not mediated by an increased release of extracellular vesicles (EV). Helium decreased the amount of medium EV [170]. The cardioprotective pathways that have been discussed for helium in this section can be found in Figure 2.
To summarize, convincing evidence supports that the biological ‘inert’ gases xenon and helium affect several molecular pathways in the heart, which contribute to the observed cardioprotection against IRI of both gases in the experimental setting.

6. Translating Preclinical Cardioprotection by Volatile Anesthetics and Noble Gases into the Clinical Arena

Although experimental findings on volatile anesthetic-induced cardioprotection were very convincing, clinical evidence for these protective effects is lacking [171,172,173,174,175,176,177]. Anesthesia with volatile anesthetics was shown to reduce mortality in cardiac surgical patients but not in patients undergoing non-cardiac surgical interventions [178]. Likhvantsev et al. compared sevoflurane with propofol in nearly 900 patients undergoing cardiac surgery. Sevoflurane reduced one-year mortality and length of hospital stay [171]. However, in a similar patient population, Landoni et al. could not demonstrate any effect on mortality and/or hospital stay when comparing sevoflurane with propofol [172]. The same author confirmed his results later in a large multicenter trial including 5400 cardiac surgical patients showing no benefits on one-year mortality when using isoflurane, sevoflurane, or desflurane compared to propofol [172]. Furthermore, other outcomes, such as myocardial infarction, did not differ between the anesthetic regimens [173]. In non-cardiac surgery, patients with increased cardiovascular risk (n = 385), no protective effect of sevoflurane versus propofol on myocardial ischemia was observed [174]. Particularly with regard to patients undergoing cardiac surgery, where balanced anesthesia has been re-commended for years, current research shows no evidence in favor of volatile anesthetics compared to intravenous anesthetics.
Similar results were reported for noble gas-induced cardioprotection in the clinical scenario.
In healthy volunteers, employing a forearm blood flow model to investigate endothelial function, helium reduced post-ischemic endothelial dysfunction without affecting plasma levels of cytokines, adhesion molecules, or microparticles known as mediators of helium-induced organ protection [179]. In patients undergoing coronary artery bypass surgery, neither helium pre-conditioning (3 × 5 min helium inhalation before aortic cross-clamping) nor post-conditioning (helium inhalation at the start of reperfusion) reduced postoperative troponin release [180]. Even a combination of pre- and postconditioning had no cardioprotective effect.
Contrary to hypothermia alone, xenon, in addition to hypothermia, attenuated myocardial damage in patients after out-of-hospital cardiac arrest and ROSC [181]. In addition, xenon combined with hypothermia was associated with greater recovery of left ventricular systolic function in comparison with hypothermia alone, indicating some cardioprotective properties of xenon in this clinical setting [182].
A multicenter, international, and randomized clinical trial assessed the cardioprotective effects of xenon anesthesia in patients undergoing coronary artery bypass graft surgery as compared to sevoflurane- or propofol-based anesthesia [176]. In 492 patients receiving either propofol, sevoflurane, or xenon, a reduction in troponin I release was shown between the xenon and propofol and between the sevoflurane and propofol group. The difference in troponin release was small, and no conclusion can be drawn whether the observed effect is clinically relevant.
To summarize, similar to volatile anesthetics, a clinically relevant cardioprotective effect of helium or xenon is absent.
It has to be mentioned, however, that also other medical gases have been investigated for their cardioprotective effects. Amongst these, NO and also Hydrogen (H2) gas have been found effective in protecting against IRI of the heart in different animal studies [183,184,185,186,187]. In a double-blind, randomized, placebo-controlled phase II trial, the impact of inhaling 80 ppm NO as an adjunctive therapy before percutaneous coronary intervention was investigated. The effects of NO were observed for up to 4 h following reperfusion. While the inhalation of NO did not show a significant reduction in infarction size (measured by magnetic resonance) between 48 and 72 h compared to the placebo group, a trend towards improved left ventricular (LV) functional recovery was observed with NO inhalation [188]. Furthermore, experimental studies have investigated the role of hydrogen gas in preventing ischemia/reperfusion injury of the heart and other organs [187] for a recent review in this area [189]. Ohsawa et al. discovered that hydrogen gas has a selective antioxidant effect that helps to scavenge harmful ROS, such as hydroxyl radicals (OH·), while preserving the beneficial signaling functions of other ROS [190]. These features help to reduce oxidative stress and limit damage to cardiac tissues. Additionally, hydrogen gas was proven to suppress the inflammatory responses by modulating pro-inflammatory cytokines and, therefore, can help prevent further damage to the heart during ischemia/reperfusion injury [189].
The pathways per drug group that have been shown to be involved in preventing IRI of the heart have been summarized in Table 1.

7. The Endothelium as Target to Protect against IRI during Cardiac Surgery with Cardiopulmonary Bypass

One of the most common surgical interventions in which global cardiac ischemia is induced is cardiac surgery. During cardiac surgery, cardiac ischemia is induced by cross-clamping of the aorta, as most surgeons prefer a non-beating heart with a blood-free operating field. During cross-clamping of the aorta, coronary artery perfusion is ceased, and ischemia is induced. To maintain systemic blood circulation, blood is transferred to a heart-lung machine, e.g., by extracorporeal cardiopulmonary bypass (CPB).
Since the start of cardiac surgery in the 1950s, considerable research has been performed to reduce the damaging effects of aorta cross-clamping during cardiac surgery. Cardioplegic strategies have been the subject of many recent reviews [191,192], but also anesthetic agents, as discussed above, have been shown to protect the heart against IRI.
Myocardial IRI is mainly characterized as a disease of cardiomyocytes, but other cellular compartments, such as the vasculature, also play a role. Most of the non-cardiomyocyte cells consist of endothelial cells [193]. Endothelial cells constitute the inner lining of arteries, veins, and capillaries and form a barrier between vessels and tissue, regulate blood flow, and are essential for tissue delivery of oxygen and nutrients. Cardiomyocytes have received the most attention as they are more sensitive to IRI compared to endothelial cells; however, with increasing knowledge of cardiac cellular composition, it has become evident that crosstalk between cardiomyocytes and endothelial cells is crucial for cardiac function [194]. Besides myocardial IRI, also endothelial dysfunction is a well-known phenomenon in cardiac surgery patients. Patients undergoing cardiac surgery with CPB have disturbed capillary perfusion [195], resulting in hypoxia and ischemia. This is, among other causes, due to a systemic inflammatory response and activation of the endothelium, leading to increased permeability of the endothelium, fluid accumulation, and tissue edema [196], thereby hampering oxygen exchange. A healthy endothelium is important for cardiac function [197], and endothelial cells can improve cardiomyocyte survival after ischemia [198]. Moreover, pharmacological targeting of the endothelium might be an effective way to protect the heart during IRI, given its capacity to release protective factors [199].

7.1. Oxidative Stress

The systemic inflammatory response during CPB is associated with increased oxidative stress and ROS formation [200]. Used antioxidants during CPB are, among others, L-arginine and N-acetyl-cysteine (NAC). Supplementation of L-arginine, a precursor for NO synthesis, in cardioplegia solution, increased NO levels and attenuated ROS radical-mediated myocardial injury in patients undergoing CABG with CPB [201]. NAC administration reduced ROS formation and creatine kinase-MB in cardiac surgery patients on CPB [202], whereas, in another study in the same patient population, NAC decreased oxidative stress substantially; however, it did not improve cardiac troponin I level [203]. Although promising results, the cardioprotective effects of antioxidants via reducing oxidative stress need further investigation.
Administration of NO, NO donors, or drugs that enhance NO release prior to ischemia protects against myocardial IRI. Drugs that enhance NO release are statins, calcium antagonists, angiotensin-converting enzyme (ACE) inhibitors, and dexamethasone, which are not discussed as these drugs are already used in the clinical setting. Supplementation of the NO donor S-nitroso human serum albumin prevented eNOS uncoupling and improved myocardial perfusion and function in a pig CPB model [204]. However, this drug has not been studied in the clinical setting. Interestingly, the NO donor Nicorandil protected the heart via the reduction in oxidative stress in a rabbit CPB model [205] and reduced inflammation and troponin T levels in patients undergoing cardiac surgery with CPB [206,207]. In addition, nitroglycerin administration in cardioplegia solution increased NO levels and decreased troponin T levels in CABG patients [208]. In contrast, treatment with glyceryl trinitrate did not affect cardiac troponin T levels and even abrogated the cardioprotective effect of RIPC in cardiac surgery patients [209]. Taken together, NO donors seem promising in protecting the heart via the vasculature during CPB; however, it should be taken into account that most NO donors have platelet inhibitory properties and might therefore have clinical consequences in terms of postoperative bleeding [210].

7.2. Glycocalyx

The surface of the vascular endothelium is covered by a gel-like layer called the endothelial glycocalyx, which regulates vascular resistance, permeability, and leukocyte recruitment. IR severely damages the endothelial glycocalyx, which can be detected by increased circulating levels of its principal constituents, syndecan-1 and heparan sulfate [211,212]. Degradation of the glycocalyx can result in myocardial edema and microvascular obstruction in patients following myocardial infarction [213]. Glycocalyx damage is also reported following aortic cross-clamping in rats [214] and in cardiac surgery patients [215,216]. During remote ischemic conditioning, glycocalyx thickness was improved in patients after myocardial infarction [217], and anesthetic conditioning protects the glycocalyx layer. In isolated pig hearts, sevoflurane protected the post-ischemic heart against endothelial dysfunction by reducing glycocalyx degradation [218]. Sevoflurane decreased glycocalyx degradation in patients undergoing heart valve surgery with CPB compared to propofol anesthesia [219]. Thus, pharmacological conditioning might preserve the endothelial glycocalyx.
Research on the effects of pharmacological conditioning on the glycocalyx during cardiac surgery is sparse. Lidoflazine, a calcium-channel blocker, preserves the glycocalyx and protects cardiac function in patients undergoing multiple aorta-coronary bypass grafting [220]. Although it has been known for many years that reperfusion results in cardiomyocyte calcium overload, as described above [221], calcium-channel blockers have not been followed up in the clinical setting. However, from a vasculature perspective, combination therapy with a calcium channel blocker could be of interest.
More recently, matrix metalloproteinases have been shown to be commonly upregulated in cardiac surgery patients with CPB [222] and are suggested as a potential mechanism triggering glycocalyx damage. Prophylactic treatment with doxycycline reduced glycocalyx damage in cardiac surgery patients with CPB by inhibiting matrix metalloproteinases [223]; however, it did not ameliorate cardiac mechanical dysfunction following reperfusion [224].
Other pharmacological means to reduce glycocalyx degradation during cardiac surgery have been explored in the context of CPB coating and priming fluids. Interestingly, the glycocalyx layer was better preserved when the CPB circuit was coated with heparin compared to phosphorylcholine in cardiac surgery patients [225]. In contrast, glycocalyx degradation did not differ between priming the CPB circuit with human albumin or gelofusine in rats [226]. Unfortunately, the above-mentioned studies did not investigate whether the protection of the endothelial glycocalyx reduced myocardial IRI. Interestingly, treatment with recombinant syndecan-1 reduced glycocalyx damage and partly ameliorated cardiomyocyte damage in mice with myocardial IRI [227], linking endothelial glycocalyx and cardiomyocytes during myocardial IRI.

7.3. Endothelial Barrier Function

Endothelial cells form a unique semi-permeable barrier for the transfer of solutes. Endothelial barrier function is regulated by cell–cell and cell–matrix adhesion, as well as endogenous mediators. Endothelial cells are joined together by adherens junctions, tight junctions, and gap junctions, and are anchored to the basement membranes via transmembrane receptors. Several types of endothelial receptors have been associated with the regulation of endothelial barrier function, such as VEGFR, sphingosine-1-phosphate (S1P) receptor, tyrosine-protein kinase (Tie2) receptor, and protease-activated receptor (PAR), and targeting them may open new therapeutic approaches to protect against myocardial IRI [228].
The involvement of VEGFR in cardioprotection was already suggested in the experimental setting during ischemic [229,230] and anesthetic conditioning [136,231]. The relationship between VEGF-A and the heart is interdependent as VEGF-A can activate cardiomyocytes, but upon inflammation, cardiomyocytes produce VEGF-A [232]. In a CPB rat model, VEGF protein is increased, especially following cardioplegia reperfusion [233]. In contrast, circulating VEGF-A levels decreased after CPB in cardiac surgery patients [234]. Interestingly, infusion of VEGF was protective for kidneys in beagles in CPB through the improvement in renal perfusion [235], but no studies have been performed yet on the effect on the heart during cardiac surgery.
Additionally, S1P signaling is well known in IRI [236]. Targeting this signaling pathway not only protects the heart but also the vasculature. In isolated rat hearts, S1P decreased infarct size and myocardial edema [237]. Reduction in myocardial water content could not be explained by protection of the endothelial glycocalyx nor hemodynamic changes, but by the involvement of S1P in endothelial permeability. However, S1P receptor subtypes have contrasting effects. Subtype 1 (S1P1) is required in the protection of endothelial barrier function, whereas the activation of S1P2 and S1P3 receptors disrupts the endothelial barrier. Pharmacological activation of the S1P1 receptor with FTY-720 and SEW2871 preserved vascular function in rats on CPB [238] and might be an interesting target to protect both the vasculature and the heart as low circulating levels of S1P were found in cardiac surgery patients [239].
The endothelial angiopoietin/Tie2 system is one of the most important mechanisms in the regulation of endothelial barrier function and is dysregulated in cardiac surgery patients [240,241]. In short, angiopoietin-1 binds to the Tie2 receptor, thereby maintaining endothelial barrier function, whereas angiopoietin-2 binds antagonistically to Tie2, increasing endothelial permeability. Targeting Tie2 has been proposed as a promising strategy to improve vascular function, due to its key position in the regulation of endothelial barrier function. Indeed, pharmacological activation of Tie2 by vasculotide protected the vasculature in a CPB rat model [242]. The Tie2 antagonist angiopoietin-2 was highly expressed in endothelial cells at the infarct border zone after myocardial infarction in mice [243]. Moreover, endothelial-derived angiopoietin-2 was involved in vascular leakage and glycocalyx degradation, thereby worsening myocardial hypoxia [243]. Inhibition of angiopoietin-2 substantially ameliorated postischemic cardiovascular remodeling [243]. Furthermore, inhibition of angiopoietin-2 in cultured cardiomyocytes enhanced the cardioprotective effects of fibroblast growth factor 2 [244]. Taken together, these data suggest that combination therapy of angiopoietin-2 inhibition together with Tie2 activation has the potential to reduce myocardial IRI.
Another pharmacological agent with protective effects on endothelial barrier function is imatinib mesylate [245]. Imatinib is a tyrosine kinase inhibitor developed to treat Bcr/Abl-expressing leukemias, but Abl kinases are also involved in endothelial barrier regulation [246]. In a CPB model, imatinib prevented endothelial barrier dysfunction and attenuated pulmonary and renal injury [247]. Unfortunately, the heart was not studied, but in a rat model of acute myocardial infarction, imatinib reduced microvascular injury and myocardial infarct size [248]. As cardiotoxic effects are reported following treatment with another tyrosine kinase inhibitor [249], future research on the effect of several generations of tyrosine kinase inhibitors on IRI is required.
During cardiac surgery, CPB is associated with the generation of thrombin. Thrombin has adverse effects on the endothelium and on cardiomyocytes, independent of its procoagulant effects, and has therefore emerged as a possible mediator of IRI [250]. Thrombin can induce endothelial hyperpermeability via activation of the PAR1 thrombin receptor. During cardiac surgery, patients receive antifibrinolytics such as aprotinin or tranexamic acid. Aprotinin can, amongst others, prevent activation of the PAR1 receptor. In a rat CPB model, aprotinin preserved endothelial integrity and reduced edema in the kidney, but the effects on the heart were not studied [251]. In a CPB pig model, aprotinin preserved the loss of coronary adherens junctions, which resulted in the preservation of the endothelial barrier and reduced myocardial edema [252]. This was confirmed in cardiac surgery patients, were aprotinin showed cardioprotective effects [253,254]. However, contradictory results on the heart have also been reported [255]. Nowadays, the most commonly used antifibrinolytic is tranexamic acid. In a recent review and meta-analysis, the authors concluded tranexamic acid administration was associated with less myocardial injury in cardiac surgery patients [256]. These results suggest that direct thrombin inhibitors might be novel pharmacological agents to protect the heart. Indeed, one study showed cardioprotective properties during myocardial IRI [257], but future studies are needed to investigate these agents during cardiac surgery. A summary of the possible new therapeutics targeting the vasculature and thereby protecting the cardiomyocyte is given in Table 1 and Figure 3.
Although the role of the vascular endothelium as a dynamic regulator of tissue responses is increasingly recognized, the relation between vascular conditioning in the ischemic heart is underappreciated. The cross-talk between endothelial cells and cardiomyocytes provides an access point for therapeutic targeting and may be a promising approach to protect the heart against IRI.
The pathways per drug group that have been shown to be involved in preventing IRI of the endothelium are summarized in Table 1.

8. Conclusions

In conclusion, while there are promising pharmacological approaches demonstrated in experimental studies for protecting the heart against IRI, there are challenges in translating these findings into the clinical setting. Several potential confounders have been identified within the last decennia, such as co-morbidities, co-medication, peri-operative care, and ischemic duration, and future studies should focus on understanding the underlying pathways involved in cardiac ischemia-reperfusion damage in order to evaluate the effectiveness of combination therapies targeting different cell types, such as endothelial cells and damage pathways. Overall, continued research in this area is critical for improving the outcomes of patients at risk of cardiac IRI.

Author Contributions

Conceptualization: N.C.W., C.J.Z. and Q.W.; manuscript preparation: Q.W., C.J.Z., R.H., C.T., M.W.H., B.P., C.E.v.d.B. and N.C.W.; manuscript editing: Q.W., C.J.Z., R.H., C.T., M.W.H., B.P., C.E.v.d.B. and N.C.W. All authors have read and agreed to the published version of the manuscript.

Funding

Q.W. is supported by the Chinese Scholarship Council (CSC no: 202107720012) fellowship program. C.B. is supported by the Dutch Research Council (Veni 2019). The research received no specific grant from any funding agency in the public, commercial, or not-for-profit sectors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ferdinandy, P.; Andreadou, I.; Baxter, G.F.; Bøtker, H.E.; Davidson, S.M.; Dobrev, D.; Gersh, B.J.; Heusch, G.; Lecour, S.; Ruiz-Meana, M.; et al. Interaction of Cardiovascular Nonmodifiable Risk Factors, Comorbidities and Comedications with Ischemia/Reperfusion Injury and Cardioprotection by Pharmacological Treatments and Ischemic Conditioning. Pharmacol. Rev. 2022, 75, 159–216. [Google Scholar] [CrossRef]
  2. Makkos, A.; Ágg, B.; Petrovich, B.; Varga, Z.V.; Görbe, A.; Ferdinandy, P. Systematic review and network analysis of microRNAs involved in cardioprotection against myocardial ischemia/reperfusion injury and infarction: Involvement of redox signalling. Free. Radic. Biol. Med. 2021, 172, 237–251. [Google Scholar] [CrossRef] [PubMed]
  3. Varga, Z.V.; Zvara, Á.; Faragó, N.; Kocsis, G.F.; Pipicz, M.; Gaspar, R.; Bencsik, P.; Görbe, A.; Csonka, C.; Puskás, L.G.; et al. MicroRNAs associated with ischemia-reperfusion injury and cardioprotection by ischemic pre- and postconditioning: ProtectomiRs. Am. J. Physiol. Heart Circ. Physiol. 2014, 307, H216–H227. [Google Scholar] [CrossRef] [PubMed]
  4. Andreadou, I.; Bell, R.M.; Bøtker, H.E.; Zuurbier, C.J. SGLT2 inhibitors reduce infarct size in reperfused ischemic heart and improve cardiac function during ischemic episodes in preclinical models. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165770. [Google Scholar] [CrossRef]
  5. Chen, S.; Wang, Q.; Christodoulou, A.; Mylonas, N.; Bakker, D.; Nederlof, R.; Hollmann, M.W.; Weber, N.C.; Coronel, R.; Wakker, V.; et al. Sodium Glucose Cotransporter-2 Inhibitor Empagliflozin Reduces Infarct Size Independently of Sodium Glucose Cotransporter-2. Circulation 2023, 147, 276–279. [Google Scholar] [CrossRef] [PubMed]
  6. Vicencio, J.M.; Yellon, D.M.; Sivaraman, V.; Das, D.; Boi-Doku, C.; Arjun, S.; Zheng, Y.; Riquelme, J.A.; Kearney, J.; Sharma, V.; et al. Plasma Exosomes Protect the Myocardium from Ischemia-Reperfusion Injury. J. Am. Coll. Cardiol. 2015, 65, 1525–1536. [Google Scholar] [CrossRef] [PubMed]
  7. Prakash, A.; E Crespo-Avilan, G.; Hernandez-Resendiz, S.; Ong, S.-G.; Hausenloy, D.J. Extracellular vesicles-mediating and delivering cardioprotection in acute myocardial infarction and heart failure. Cond. Med. 2020, 3, 227–238. [Google Scholar]
  8. Bell, R.M.; Yellon, D.M. There is More to Life than Revascularization: Therapeutic Targeting of Myocardial Ischemia/Reperfusion Injury. Cardiovasc. Ther. 2010, 29, e67–e79. [Google Scholar] [CrossRef] [PubMed]
  9. Garcia-Dorado, D.; Théroux, P.; Duran, J.M.; Solares, J.; Alonso, J.; Sanz, E.; Munoz, R.; Elizaga, J.; Botas, J.; Fernandez-Avilés, F. Selective inhibition of the contractile apparatus. A new approach to modification of infarct size, infarct composition, and infarct geometry during coronary artery occlusion and reperfusion. Circulation 1992, 85, 1160–1174. [Google Scholar] [CrossRef]
  10. Fiolet, J.; Baartscheer, A.; Schumacher, C.; Coronel, R.; ter Welle, H. The change of the free energy of ATP hydrolysis during global ischemia and anoxia in the rat heart: Its possible role in the regulation of transsarcolemmal sodium and potassium gradients. J. Mol. Cell. Cardiol. 1984, 16, 1023–1036. [Google Scholar] [CrossRef]
  11. Steenbergen, C.; Murphy, E.; Watts, J.A.; London, R.E. Correlation between cytosolic free calcium, contracture, ATP, and irreversible ischemic injury in perfused rat heart. Circ. Res. 1990, 66, 135–146. [Google Scholar] [CrossRef]
  12. Van Borren, M.M.; Zegers, J.G.; Baartscheer, A.; Ravesloot, J.H. Contribution of NHE-1 to cell length shortening of normal and failing rabbit cardiac myocytes. J. Mol. Cell. Cardiol. 2006, 41, 706–715. [Google Scholar] [CrossRef]
  13. Karmazyn, M.; Gan, X.T.; Humphreys, R.A.; Yoshida, H.; Kusumoto, K. The myocardial Na(+)-H(+) exchange: Structure, regulation, and its role in heart disease. Circ. Res. 1999, 85, 777–786. [Google Scholar] [CrossRef] [PubMed]
  14. Miyata, N.; Morris, L.L.; Chen, Q.; Thorne, C.; Singla, A.; Zhu, W.; Winter, M.; Melton, S.D.; Li, H.; Sifuentes-Dominguez, L.; et al. Microbial Sensing by Intestinal Myeloid Cells Controls Carcinogenesis and Epithelial Differentiation. Cell Rep. 2018, 24, 2342–2355. [Google Scholar] [CrossRef]
  15. Yang, X.-M.; Cohen, M.V.; Sayner, S.; Audia, J.P.; Downey, J.M. Lethal Caspase-1/4-Dependent Injury Occurs in the First Minutes of Coronary Reperfusion and Requires Calpain Activity. Int. J. Mol. Sci. 2023, 24, 3801. [Google Scholar] [CrossRef]
  16. Tani, M.; Hasegawa, H.; Suganuma, Y.; Shinmura, K.; Kayashi, Y.; Nakamura, Y. Protection of ischemic myocardium by inhibition of contracture in isolated rat heart. Am. J. Physiol. Content 1996, 271, H2515–H2519. [Google Scholar] [CrossRef] [PubMed]
  17. Zweier, J.L.; Flaherty, J.T.; Weisfeldt, M.L. Direct measurement of free radical generation following reperfusion of ischemic myocardium. Proc. Natl. Acad. Sci. USA 1987, 84, 1404–1407. [Google Scholar] [CrossRef]
  18. Maupoil, V.; Rochette, L.; Tabard, A.; Clauser, P.; Harpey, C. Evolution of free radical formation during low-flow ischemia and reperfusion in isolated rat heart. Cardiovasc. Drugs Ther. 1990, 4, 791–795. [Google Scholar] [CrossRef] [PubMed]
  19. Timoshin, A.A.; Tskitishvili, O.V.; Serebryakova, L.I.; Kuzmin, A.I.; Medvedev, O.; Ruuge, E.K. Microdialysis study of ischemia-induced hydroxyl radicals in the canine heart. Experientia 1994, 50, 677–679. [Google Scholar] [CrossRef]
  20. Vanden Hoek, T.L.; Li, C.; Shao, Z.; Schumacker, P.T.; Becker, L.B. Significant levels of oxidants are generated by isolated cardiomyocytes during ischemia prior to reperfusion. J. Mol. Cell. Cardiol. 1997, 29, 2571–2583. [Google Scholar] [CrossRef]
  21. Eaton, P.; Li, J.-M.; Hearse, D.J.; Shattock, M.J. Formation of 4-hydroxy-2-nonenal-modified proteins in ischemic rat heart. Am. J. Physiol. Content 1999, 276, H935–H943. [Google Scholar] [CrossRef] [PubMed]
  22. Xiao, Y.; Yim, K.; Zhang, H.; Bakker, D.; Nederlof, R.; Smeitink, J.A.M.; Renkema, H.; Hollmann, M.W.; Weber, N.C.; Zuurbier, C.J. The Redox Modulating Sonlicromanol Active Metabolite KH176m and the Antioxidant MPG Protect Against Short-Duration Cardiac Ischemia-Reperfusion Injury. Cardiovasc. Drugs Ther. 2021, 35, 745–758. [Google Scholar] [CrossRef] [PubMed]
  23. Eaton, P.; Hearse, D.J.; Shattock, M.J. Lipid hydroperoxide modification of proteins during myocardial ischaemia. Cardiovasc. Res. 2001, 51, 294–303. [Google Scholar] [CrossRef]
  24. Raedschelders, K.; Ansley, D.M.; Chen, D.D. The cellular and molecular origin of reactive oxygen species generation during myocardial ischemia and reperfusion. Pharmacol. Ther. 2012, 133, 230–255. [Google Scholar] [CrossRef] [PubMed]
  25. Garlick, P.B.; Davies, M.J.; Hearse, D.J.; Slater, T.F. Direct detection of free radicals in the reperfused rat heart using electron spin resonance spectroscopy. Circ. Res. 1987, 61, 757–760. [Google Scholar] [CrossRef] [PubMed]
  26. Nederlof, R.; Gürel-Gurevin, E.; Eerbeek, O.; Xie, C.; Deijs, G.S.; Konkel, M.; Hu, J.; Weber, N.C.; Schumacher, C.A.; Baartscheer, A.; et al. Reducing mitochondrial bound hexokinase II mediates transition from non-injurious into injurious ischemia/reperfusion of the intact heart. J. Physiol. Biochem. 2016, 73, 323–333. [Google Scholar] [CrossRef]
  27. Chouchani, E.T.; Pell, V.R.; Gaude, E.; Aksentijević, D.; Sundier, S.Y.; Robb, E.L.; Logan, A.; Nadtochiy, S.M.; Ord, E.N.J.; Smith, A.C.; et al. Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature 2014, 515, 431–435. [Google Scholar] [CrossRef]
  28. Matsushima, S.; Sadoshima, J. Yin and Yang of NADPH Oxidases in Myocardial Ischemia-Reperfusion. Antioxidants 2022, 11, 1069. [Google Scholar] [CrossRef]
  29. Da-Silva, W.S.; Gómez-Puyou, A.; de Gómez-Puyou, M.T.; Moreno-Sanchez, R.; De Felice, F.G.; de Meis, L.; Oliveira, M.F.; Galina, A. Mitochondrial bound hexokinase activity as a preventive antioxidant defense: Steady-state ADP formation as a regulatory mechanism of membrane potential and reactive oxygen species generation in mitochondria. J. Biol. Chem. 2004, 279, 39846–39855. [Google Scholar] [CrossRef]
  30. Ong, S.-G.; Lee, W.H.; Theodorou, L.; Kodo, K.; Lim, S.Y.; Shukla, D.H.; Briston, T.; Kiriakidis, S.; Ashcroft, M.; Davidson, S.M.; et al. HIF-1 reduces ischaemia–reperfusion injury in the heart by targeting the mitochondrial permeability transition pore. Cardiovasc. Res. 2014, 104, 24–36. [Google Scholar] [CrossRef]
  31. Yang, M.; Sun, J.; Stowe, D.F.; Tajkhorshid, E.; Kwok, W.-M.; Camara, A.K.S. Knockout of VDAC1 in H9c2 Cells Promotes Oxidative Stress-Induced Cell Apoptosis through Decreased Mitochondrial Hexokinase II Binding and Enhanced Glycolytic Stress. Cell. Physiol. Biochem. 2020, 54, 853–874. [Google Scholar] [CrossRef] [PubMed]
  32. Liepinsh, E.; Makrecka-Kuka, M.; Volska, K.; Kuka, J.; Makarova, E.; Antone, U.; Sevostjanovs, E.; Vilskersts, R.; Strods, A.; Tars, K.; et al. Long-chain acylcarnitines determine ischaemia/reperfusion-induced damage in heart mitochondria. Biochem. J. 2016, 473, 1191–1202. [Google Scholar] [CrossRef]
  33. Dambrova, M.; Zuurbier, C.J.; Borutaite, V.; Liepinsh, E.; Makrecka-Kuka, M. Energy substrate metabolism and mitochondrial oxidative stress in cardiac ischemia/reperfusion injury. Free. Radic. Biol. Med. 2021, 165, 24–37. [Google Scholar] [CrossRef]
  34. Dumitrescu, C.; Biondi, R.; Xia, Y.; Cardounel, A.J.; Druhan, L.J.; Ambrosio, G.; Zweier, J.L. Myocardial ischemia results in tetrahydrobiopterin (BH4) oxidation with impaired endothelial function ameliorated by BH4. Proc. Natl. Acad. Sci. USA 2007, 104, 15081–15086. [Google Scholar] [CrossRef] [PubMed]
  35. Halestrap, A.P.; Richardson, A.P. The mitochondrial permeability transition: A current perspective on its identity and role in ischaemia/reperfusion injury. J. Mol. Cell. Cardiol. 2015, 78, 129–141. [Google Scholar] [CrossRef]
  36. Zuurbier, C.J.; Bertrand, L.; Beauloye, C.R.; Andreadou, I.; Ruiz-Meana, M.; Jespersen, N.R.; Kula-Alwar, D.; Prag, H.A.; Eric Botker, H.; Dambrova, M.; et al. Cardiac metabolism as a driver and therapeutic target of myocardial infarction. J. Cell. Mol. Med. 2020, 24, 5937–5954. [Google Scholar] [CrossRef]
  37. Taegtmeyer, H. Metabolic responses to cardiac hypoxia. Increased production of succinate by rabbit papillary muscles. Circ. Res. 1978, 43, 808–815. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, J.; Wang, Y.T.; Miller, J.H.; Day, M.M.; Munger, J.C.; Brookes, P.S. Accumulation of Succinate in Cardiac Ischemia Primarily Occurs via Canonical Krebs Cycle Activity. Cell Rep. 2018, 23, 2617–2628. [Google Scholar] [CrossRef]
  39. Vogt, A.M.; Poolman, M.; Ackermann, C.; Yildiz, M.; Schoels, W.; Fell, D.A.; Kübler, W. Regulation of Glycolytic Flux in Ischemic Preconditioning. J. Biol. Chem. 2002, 277, 24411–24419. [Google Scholar] [CrossRef]
  40. Jennings, R.B.; Murry, C.E.; Reimer, K.A. Energy metabolism in preconditioned and control myocardium: Effect of total ischemia. J. Mol. Cell. Cardiol. 1991, 23, 1449–1458. [Google Scholar] [CrossRef]
  41. Pasdois, P.; Parker, J.E.; Halestrap, A.P. Extent of mitochondrial hexokinase II dissociation during ischemia correlates with mitochondrial cytochrome c release, reactive oxygen species production, and infarct size on reperfusion. J. Am. Heart Assoc. 2012, 2, e005645. [Google Scholar] [CrossRef] [PubMed]
  42. Smeele, K.M.A.; Southworth, R.; Wu, R.; Xie, C.; Nederlof, R.; Warley, A.; Nelson, J.K.; Van Horssen, P.; Wijngaard, J.P.V.D.; Heikkinen, S.; et al. Disruption of Hexokinase II–Mitochondrial Binding Blocks Ischemic Preconditioning and Causes Rapid Cardiac Necrosis. Circ. Res. 2011, 108, 1165–1169. [Google Scholar] [CrossRef]
  43. Nederlof, R.; Eerbeek, O.; Hollmann, M.W.; Southworth, R.; Zuurbier, C.J. Targeting hexokinase II to mitochondria to modulate energy metabolism and reduce ischaemia-reperfusion injury in heart. Br. J. Pharmacol. 2014, 171, 2067–2079. [Google Scholar] [CrossRef] [PubMed]
  44. Andrienko, T.N.; Pasdois, P.; Pereira, G.C.; Ovens, M.J.; Halestrap, A.P. The role of succinate and ROS in reperfusion injury—A critical appraisal. J. Mol. Cell. Cardiol. 2017, 110, 1–14. [Google Scholar] [CrossRef]
  45. Gürel, E.; Smeele, K.M.; Eerbeek, O.; Koeman, A.; Demirci, C.; Hollmann, M.W.; Zuurbier, C.J. Ischemic preconditioning affects hexokinase activity and HKII in different subcellular compartments throughout cardiac ischemia-reperfusion. J. Appl. Physiol. 2009, 106, 1909–1916. [Google Scholar] [CrossRef] [PubMed]
  46. Nederlof, R.; Denis, S.; Lauzier, B.; Rosiers, C.D.; Laakso, M.; Hagen, J.; Argmann, C.; Wanders, R.; Houtkooper, R.H.; Hollmann, M.W.; et al. Acute detachment of hexokinase II from mitochondria modestly increases oxygen consumption of the intact mouse heart. Metabolism 2017, 72, 66–74. [Google Scholar] [CrossRef]
  47. Chiara, F.; Castellaro, D.; Marin, O.; Petronilli, V.; Brusilow, W.S.; Juhaszova, M.; Sollott, S.J.; Forte, M.; Bernardi, P.; Rasola, A. Hexokinase II Detachment from Mitochondria Triggers Apoptosis through the Permeability Transition Pore Independent of Voltage-Dependent Anion Channels. PLoS ONE 2008, 3, e1852. [Google Scholar] [CrossRef]
  48. Perevoshchikova, I.V.; Zorov, S.D.; Kotova, E.A.; Zorov, D.B.; Antonenko, Y.N. Hexokinase inhibits flux of fluorescently labeled ATP through mitochondrial outer membrane porin. FEBS Lett. 2010, 584, 2397–2402. [Google Scholar] [CrossRef]
  49. Uthman, L.; Nederlof, R.; Eerbeek, O.; Baartscheer, A.; Schumacher, C.; Buchholtz, N.; Hollmann, M.W.; Coronel, R.; Weber, N.C.; Zuurbier, C.J. Delayed ischaemic contracture onset by empagliflozin associates with NHE1 inhibition and is dependent on insulin in isolated mouse hearts. Cardiovasc. Res. 2019, 115, 1533–1545. [Google Scholar] [CrossRef]
  50. Zuurbier, C.J.; Eerbeek, O.; Meijer, A.J. Ischemic preconditioning, insulin, and morphine all cause hexokinase redistribution. Am. J. Physiol. Heart Circ. Physiol. 2005, 289, H496–H499. [Google Scholar] [CrossRef]
  51. 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] [PubMed]
  52. Whitmer, J.T.; Idell-Wenger, J.A.; Rovetto, M.J.; Neely, J.R. Control of fatty acid metabolism in ischemic and hypoxic hearts. J. Biol. Chem. 1978, 253, 4305–4309. [Google Scholar] [CrossRef] [PubMed]
  53. Ford, D.A.; Han, X.; Horner, C.C.; Gross, R.W. Accumulation of Unsaturated Acylcarnitine Molecular Species During Acute Myocardial Ischemia: Metabolic Compartmentalization of Products of Fatty Acyl Chain Elongation in the Acylcarnitine Pool. Biochemistry 1996, 35, 7903–7909. [Google Scholar] [CrossRef] [PubMed]
  54. van Weeghel, M.; Abdurrachim, D.; Nederlof, R.; Argmann, C.A.; Houtkooper, R.H.; Hagen, J.; Nabben, M.; Denis, S.; Ciapaite, J.; Kolwicz, S.C., Jr.; et al. Increased cardiac fatty acid oxidation in a mouse model with decreased malonyl-CoA sensitivity of CPT1B. Cardiovasc. Res. 2018, 114, 1324–1334. [Google Scholar] [CrossRef]
  55. Makrecka, M.; Kuka, J.; Volska, K.; Antone, U.; Sevostjanovs, E.; Cirule, H.; Grinberga, S.; Pugovics, O.; Dambrova, M.; Liepinsh, E. Long-chain acylcarnitine content determines the pattern of energy metabolism in cardiac mitochondria. Mol. Cell. Biochem. 2014, 395, 1–10. [Google Scholar] [CrossRef]
  56. Tominaga, H.; Katoh, H.; Odagiri, K.; Takeuchi, Y.; Kawashima, H.; Saotome, M.; Urushida, T.; Satoh, H.; Hayashi, H. Different effects of palmitoyl-L-carnitine and palmitoyl-CoA on mitochondrial function in rat ventricular myocytes. Am. J. Physiol. Heart Circ. Physiol. 2008, 295, H105–H112. [Google Scholar] [CrossRef]
  57. Moens, A.L.; Champion, H.C.; Claeys, M.J.; Tavazzi, B.; Kaminski, P.M.; Wolin, M.S.; Borgonjon, D.J.; Van Nassauw, L.; Haile, A.; Zviman, M.; et al. High-Dose Folic Acid Pretreatment Blunts Cardiac Dysfunction During Ischemia Coupled to Maintenance of High-Energy Phosphates and Reduces Postreperfusion Injury. Circulation 2008, 117, 1810–1819. [Google Scholar] [CrossRef]
  58. Zuurbier, C.J.; Heinen, A.; Koeman, A.; Stuifbergen, R.; Hakvoort, T.B.M.; Weber, N.C.; Hollmann, M.W. Cardioprotective efficacy depends critically on pharmacological dose, duration of ischaemia, health status of animals and choice of anaesthetic regimen: A case study with folic acid. J. Transl. Med. 2014, 12, 325. [Google Scholar] [CrossRef]
  59. Verdin, E. NAD⁺ in aging, metabolism, and neurodegeneration. Science 2015, 350, 1208–1213. [Google Scholar] [CrossRef]
  60. Chu, X.; Raju, R.P. Regulation of NAD(+) metabolism in aging and disease. Metabolism 2022, 126, 154923. [Google Scholar] [CrossRef]
  61. Abdellatif, M.; Sedej, S.; Kroemer, G. NAD(+) Metabolism in Cardiac Health, Aging, and Disease. Circulation 2021, 144, 1795–1817. [Google Scholar] [CrossRef]
  62. Di Lisa, F.; Menabò, R.; Canton, M.; Barile, M.; Bernardi, P. Opening of the Mitochondrial Permeability Transition Pore Causes Depletion of Mitochondrial and Cytosolic NAD+and Is a Causative Event in the Death of Myocytes in Postischemic Reperfusion of the Heart. J. Biol. Chem. 2001, 276, 2571–2575. [Google Scholar] [CrossRef] [PubMed]
  63. Hogan, K.A.; Chini, C.C.S.; Chini, E.N. The Multi-faceted Ecto-enzyme CD38: Roles in Immunomodulation, Cancer, Aging, and Metabolic Diseases. Front. Immunol. 2019, 10, 1187. [Google Scholar] [CrossRef]
  64. Tran, D.H.; Wang, Z.V. Glucose Metabolism in Cardiac Hypertrophy and Heart Failure. J. Am. Heart Assoc. 2019, 8, e012673. [Google Scholar] [CrossRef] [PubMed]
  65. Hwang, Y.C.; Sato, S.; Tsai, J.-Y.; Yan, S.; Bakr, S.; Zhang, H.; Oates, P.J.; Ramasamy, R. Aldose reductase activation is a key component of myocardial response to ischemia. FASEB J. 2001, 16, 243–245. [Google Scholar] [CrossRef] [PubMed]
  66. Hwang, Y.C.; Bakr, S.; Ellery, C.A.; Oates, P.J.; Ramasamy, R. Sorbitol dehydrogenase: A novel target for adjunctive protection of ischemic myocardium. FASEB J. 2003, 17, 2331–2333. [Google Scholar] [CrossRef] [PubMed]
  67. Yamamoto, T.; Byun, J.; Zhai, P.; Ikeda, Y.; Oka, S.; Sadoshima, J. Nicotinamide mononucleotide, an intermediate of NAD+ synthesis, protects the heart from ischemia and reperfusion. PLoS ONE 2014, 9, e98972. [Google Scholar] [CrossRef]
  68. Nadtochiy, S.M.; Wang, Y.T.; Nehrke, K.; Munger, J.; Brookes, P.S. Cardioprotection by nicotinamide mononucleotide (NMN): Involvement of glycolysis and acidic pH. J. Mol. Cell. Cardiol. 2018, 121, 155–162. [Google Scholar] [CrossRef]
  69. Soares, R.O.S.; Losada, D.M.; Jordani, M.C.; Évora, P.; Castro-E-Silva, O. Ischemia/Reperfusion Injury Revisited: An Overview of the Latest Pharmacological Strategies. Int. J. Mol. Sci. 2019, 20, 5034. [Google Scholar] [CrossRef]
  70. Zhang, Y.; Wang, B.; Fu, X.; Guan, S.; Han, W.; Zhang, J.; Gan, Q.; Fang, W.; Ying, W.; Qu, X. Exogenous NAD(+) administration significantly protects against myocardial ischemia/reperfusion injury in rat model. Am. J. Transl. Res. 2016, 8, 3342–3350. [Google Scholar]
  71. Zhang, Y.-J.; Zhang, M.; Zhao, X.; Shi, K.; Ye, M.; Tian, J.; Guan, S.; Ying, W.; Qu, X. NAD+ administration decreases microvascular damage following cardiac ischemia/reperfusion by restoring autophagic flux. Basic Res. Cardiol. 2020, 115, 57. [Google Scholar] [CrossRef] [PubMed]
  72. Trammell, S.A.J.; Schmidt, M.S.; Weidemann, B.J.; Redpath, P.; Jaksch, F.; Dellinger, R.W.; Li, Z.; Abel, E.D.; Migaud, M.E.; Brenner, C. Nicotinamide riboside is uniquely and orally bioavailable in mice and humans. Nat. Commun. 2016, 7, 12948. [Google Scholar] [CrossRef] [PubMed]
  73. Xiao, Y.; Phelp, P.; Wang, Q.; Bakker, D.; Nederlof, R.; Hollmann, M.W.; Zuurbier, C.J. Cardioprotecive Properties of Known Agents in Rat Ischemia-Reperfusion Model Under Clinically Relevant Conditions: Only the NAD Precursor Nicotinamide Riboside Reduces Infarct Size in Presence of Fentanyl, Midazolam and Cangrelor, but Not Propofol. Front. Cardiovasc. Med. 2021, 8, 712478. [Google Scholar] [CrossRef]
  74. Xiao, Y.; Phelp, P.; Bakker, D.; Nederlof, R.; Hollmann, M.W.; Weber, N.C.; Zuurbier, C.J. Only the NAD precursor nicotinamide riboside maintains cardioprotection under clinically relevant conditions, possibly through activation of glycolysis. Eur. Heart J. 2021, 42, ehab724.3248. [Google Scholar] [CrossRef]
  75. Kharechkina, E.; Nikiforova, A.; Kruglov, A. NAD(H) Regulates the Permeability Transition Pore in Mitochondria through an External Site. Int. J. Mol. Sci. 2021, 22, 8560. [Google Scholar] [CrossRef]
  76. Valls-Lacalle, L.; Barba, I.; Miró-Casas, E.; Ruiz-Meana, M.; Rodríguez-Sinovas, A.; García-Dorado, D. Selective Inhibition of Succinate Dehydrogenase in Reperfused Myocardium with Intracoronary Malonate Reduces Infarct Size. Sci. Rep. 2018, 8, 2442. [Google Scholar] [CrossRef] [PubMed]
  77. Jespersen, N.R.; Hjortbak, M.V.; Lassen, T.R.; Støttrup, N.B.; Johnsen, J.; Tonnesen, P.T.; Larsen, S.; Kimose, H.-H.; Bøtker, H.E. Cardioprotective effect of succinate dehydrogenase inhibition in rat hearts and human myocardium with and without diabetes mellitus. Sci. Rep. 2020, 10, 10344. [Google Scholar] [CrossRef]
  78. Tonnesen, P.T.; Hjortbak, M.V.; Lassen, T.R.; Seefeldt, J.M.; Bøtker, H.E.; Jespersen, N.R. Myocardial salvage by succinate dehydrogenase inhibition in ischemia–reperfusion injury depends on diabetes stage in rats. Mol. Cell. Biochem. 2021, 476, 2675–2684. [Google Scholar] [CrossRef]
  79. Prag, H.A.; Aksentijevic, D.; Dannhorn, A.; Giles, A.V.; Mulvey, J.F.; Sauchanka, O.; Du, L.; Bates, G.; Reinhold, J.; Kula-Alwar, D.; et al. Ischemia-Selective Cardioprotection by Malonate for Ischemia/Reperfusion Injury. Circ. Res. 2022, 131, 528–541. [Google Scholar] [CrossRef]
  80. Arslan, F.; Smeets, M.B.; O’Neill, L.A.; Keogh, B.; McGuirk, P.; Timmers, L.; Tersteeg, C.; Hoefer, I.E.; Doevendans, P.A.; Pasterkamp, G.; et al. Myocardial Ischemia/Reperfusion Injury Is Mediated by Leukocytic Toll-Like Receptor-2 and Reduced by Systemic Administration of a Novel Anti–Toll-Like Receptor-2 Antibody. Circulation 2010, 121, 80–90. [Google Scholar] [CrossRef]
  81. Shimamoto, A.; Chong, A.J.; Yada, M.; Shomura, S.; Takayama, H.; Fleisig, A.J.; Agnew, M.L.; Hampton, C.R.; Rothnie, C.L.; Spring, D.J.; et al. Inhibition of Toll-like Receptor 4 With Eritoran Attenuates Myocardial Ischemia-Reperfusion Injury. Circulation 2006, 114, I270–I274. [Google Scholar] [CrossRef] [PubMed]
  82. Toldo, S.; Mezzaroma, E.; Buckley, L.F.; Potere, N.; Di Nisio, M.; Biondi-Zoccai, G.; Van Tassell, B.W.; Abbate, A. Targeting the NLRP3 inflammasome in cardiovascular diseases. Pharmacol. Ther. 2021, 236, 108053. [Google Scholar] [CrossRef] [PubMed]
  83. Zuurbier, C.J.; Abbate, A.; Cabrera-Fuentes, H.A.; Cohen, M.V.; Collino, M.; De Kleijn, D.P.V.; Downey, J.M.; Pagliaro, P.; Preissner, K.T.; Takahashi, M.; et al. Innate immunity as a target for acute cardioprotection. Cardiovasc. Res. 2019, 115, 1131–1142. [Google Scholar] [CrossRef] [PubMed]
  84. 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]
  85. Jong, W.M.; Leemans, J.C.; Weber, N.C.; Juffermans, N.P.; Schultz, M.J.; 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] [CrossRef] [PubMed]
  86. Kawaguchi, M.; Takahashi, M.; Hata, T.; Kashima, Y.; Usui, F.; Morimoto, H.; Izawa, A.; Takahashi, Y.; Masumoto, J.; Koyama, J.; et al. Inflammasome Activation of Cardiac Fibroblasts Is Essential for Myocardial Ischemia/Reperfusion Injury. Circulation 2011, 123, 594–604. [Google Scholar] [CrossRef]
  87. 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]
  88. Van Hout, G.P.; Bosch, L.; Ellenbroek, G.H.; de Haan, J.J.; van Solinge, W.W.; Cooper, M.A.; Arslan, F.; de Jager, S.C.; Robertson, A.A.; Pasterkamp, G.; et al. The selective NLRP3-inflammasome inhibitor MCC950 reduces infarct size and preserves cardiac function in a pig model of myocardial infarction. Eur. Heart J. 2017, 38, 828–836. [Google Scholar] [CrossRef]
  89. Silvis, M.J.M.; Demkes, E.J.; Timmers, L.; Arslan, F.; de Jager, S.C.A.; Sluijter, J.P.G.; Mosterd, A.; de Kleijn, D.P.V.; Bosch, L.; van Hout, G.P.J. NLRP3-Inflammasome Inhibition with IZD334 Does Not Reduce Cardiac Damage in a Pig Model of Myocardial Infarction. Biomedicines 2022, 10, 3056. [Google Scholar] [CrossRef]
  90. Mewton, N.; Roubille, F.; Bresson, D.; Prieur, C.; Bouleti, C.; Bochaton, T.; Ivanes, F.; Dubreuil, O.; Biere, L.; Hayek, A.; et al. Effect of Colchicine on Myocardial Injury in Acute Myocardial Infarction. Circulation 2021, 144, 859–869. [Google Scholar] [CrossRef]
  91. Shah, B.; Pillinger, M.; Zhong, H.; Cronstein, B.; Xia, Y.; Lorin, J.D.; Smilowitz, N.R.; Feit, F.; Ratnapala, N.; Keller, N.M.; et al. Effects of Acute Colchicine Administration Prior to Percutaneous Coronary Intervention: COLCHICINE-PCI Randomized Trial. Circ. Cardiovasc. Interv. 2020, 13, e008717. [Google Scholar] [CrossRef] [PubMed]
  92. McIlwain, D.R.; Berger, T.; Mak, T.W. Caspase functions in cell death and disease. Cold Spring Harb. Perspect. Biol. 2015, 7, a008656. [Google Scholar] [CrossRef] [PubMed]
  93. Bahi, N.; Zhang, J.; Llovera, M.; Ballester, M.; Comella, J.X.; Sanchis, D. Switch from caspase-dependent to caspase-independent death during heart development: Essential role of endonuclease G in ischemia-induced DNA processing of differentiated cardiomyocytes. J. Biol. Chem. 2006, 281, 22943–22952. [Google Scholar] [CrossRef] [PubMed]
  94. Inserte, J.; Cardona, M.; Poncelas-Nozal, M.; Hernando, V.; Vilardosa, Ú.; Aluja, D.; Parra, V.M.; Sanchis, D.; Garcia-Dorado, D. Studies on the role of apoptosis after transient myocardial ischemia: Genetic deletion of the executioner caspases-3 and -7 does not limit infarct size and ventricular remodeling. Basic Res. Cardiol. 2016, 111, 18. [Google Scholar] [CrossRef]
  95. Cardona, M.; Lopez, J.A.; Serafín, A.; Rongvaux, A.; Inserte, J.; García-Dorado, D.; Flavell, R.; Llovera, M.; Cañas, X.; Vázquez, J.; et al. Executioner Caspase-3 and 7 Deficiency Reduces Myocyte Number in the Developing Mouse Heart. PLoS ONE 2015, 10, e0131411. [Google Scholar] [CrossRef] [PubMed]
  96. McCully, J.D.; Wakiyama, H.; Hsieh, Y.-J.; Jones, M.; Levitsky, S. Differential contribution of necrosis and apoptosis in myocardial ischemia-reperfusion injury. Am. J. Physiol. Heart Circ. Physiol. 2004, 286, H1923–H1935. [Google Scholar] [CrossRef]
  97. Mocanu, M.M.; Baxter, G.F.; Yellon, D.M. Caspase inhibition and limitation of myocardial infarct size: Protection against lethal reperfusion injury. Br. J. Pharmacol. 2000, 130, 197–200. [Google Scholar] [CrossRef]
  98. Kovacs, P.; Bak, I.; Szendrei, L.; Vecsernyes, M.; Varga, E.; Blasig, I.E.; Tosaki, A. Non-specific caspase inhibition reduces infarct size and improves post-ischaemic recovery in isolated ischaemic/reperfused rat hearts. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2001, 364, 501–507. [Google Scholar] [CrossRef]
  99. He, Z.; Davidson, S.M.; Yellon, D.M. The importance of clinically relevant background therapy in cardioprotective studies. Basic Res. Cardiol. 2020, 115, 69. [Google Scholar] [CrossRef]
  100. Audia, J.P.; Yang, X.-M.; Crockett, E.S.; Housley, N.; Haq, E.U.; O’donnell, K.; Cohen, M.V.; Downey, J.M.; Alvarez, D.F. Caspase-1 inhibition by VX-765 administered at reperfusion in P2Y12 receptor antagonist-treated rats provides long-term reduction in myocardial infarct size and preservation of ventricular function. Basic Res. Cardiol. 2018, 113, 32. [Google Scholar] [CrossRef]
  101. Yue, R.-C.; Lu, S.-Z.; Luo, Y.; Wang, T.; Liang, H.; Zeng, J.; Liu, J.; Hu, H.-X. Calpain silencing alleviates myocardial ischemia-reperfusion injury through the NLRP3/ASC/Caspase-1 axis in mice. Life Sci. 2019, 233, 116631. [Google Scholar] [CrossRef]
  102. Zhang, Y.; Rong, H.; Zhang, F.-X.; Wu, K.; Mu, L.; Meng, J.; Xiao, B.; Zamponi, G.W.; Shi, Y. A Membrane Potential- and Calpain-Dependent Reversal of Caspase-1 Inhibition Regulates Canonical NLRP3 Inflammasome. Cell Rep. 2018, 24, 2356–2369.e5. [Google Scholar] [CrossRef] [PubMed]
  103. Shi, H.; Gao, Y.; Dong, Z.; Yang, J.; Gao, R.; Li, X.; Zhang, S.; Ma, L.; Sun, X.; Wang, Z.; et al. GSDMD-Mediated Cardiomyocyte Pyroptosis Promotes Myocardial I/R Injury. Circ. Res. 2021, 129, 383–396. [Google Scholar] [CrossRef] [PubMed]
  104. Penna, C.; Andreadou, I.; Aragno, M.; Beauloye, C.; Bertrand, L.; Lazou, A.; Falcão-Pires, I.; Bell, R.; Zuurbier, C.J.; Pagliaro, P. Effect of hyperglycaemia and diabetes on acute myocardial ischaemia–reperfusion injury and cardioprotection by ischaemic conditioning protocols. Br. J. Pharmacol. 2020, 177, 5312–5335. [Google Scholar] [CrossRef]
  105. Meng, Z.; Zhang, Z.; Zhao, J.; Liu, C.; Yao, P.; Zhang, L.; Xie, D.; Lau, W.B.; Tsukuda, J.; Christopher, T.A.; et al. Nitrative Modification of Caveolin-3: A Novel Mechanism of Cardiac Insulin Resistance and a Potential Therapeutic Target Against Ischemic Heart Failure in Prediabetic Animals. Circulation 2023, 147, 1162–1179. [Google Scholar] [CrossRef] [PubMed]
  106. Li, X.; Preckel, B.; Hermanides, J.; Hollmann, M.W.; Zuurbier, C.J.; Weber, N.C. Amelioration of endothelial dysfunction by sodium glucose co-transporter 2 inhibitors: Pieces of the puzzle explaining their cardiovascular protection. Br. J. Pharmacol. 2022, 179, 4047–4062. [Google Scholar] [CrossRef]
  107. Chen, S.; Coronel, R.; Hollmann, M.W.; Weber, N.C.; Zuurbier, C.J. Direct cardiac effects of SGLT2 inhibitors. Cardiovasc. Diabetol. 2022, 21, 45. [Google Scholar] [CrossRef]
  108. Xiao, Y.; Wang, Q.; Zuurbier, C.J. Efficacy of cardioprotective interventions depends critically on duration of ischemia. Cond. Med. 2022, 5, 93–99. [Google Scholar]
  109. Hausenloy, D.J.; Kharbanda, R.K.; Møller, U.K.; Ramlall, M.; Aarøe, J.; Butler, R.; Bulluck, H.; Clayton, T.; Dana, A.; Dodd, M.; et al. Effect of remote ischaemic conditioning on clinical outcomes in patients with acute myocardial infarction (CONDI-2/ERIC-PPCI): A single-blind randomised controlled trial. Lancet 2019, 394, 1415–1424. [Google Scholar] [CrossRef]
  110. Park, J.; Choi, K.H.; Lee, J.M.; Kim, H.K.; Hwang, D.; Rhee, T.M.; Kim, J.; Park, T.K.; Yang, J.H.; Song, Y.B.; et al. Prognostic Implications of Door-to-Balloon Time and Onset-to-Door Time on Mortality in Patients with ST -Segment-Elevation Myocardial Infarction Treated with Primary Percutaneous Coronary Intervention. J. Am. Heart Assoc. 2019, 8, e012188. [Google Scholar] [CrossRef]
  111. Zahler, D.; Rozenfeld, K.-L.; Pasternak, Y.; Itach, T.; Lupu, L.; Banai, S.; Shacham, Y. Relation of Pain-to-Balloon Time and Mortality in Patients with ST-Segment Elevation Myocardial Infarction Undergoing Primary Percutaneous Coronary Intervention. Am. J. Cardiol. 2021, 163, 38–42. [Google Scholar] [CrossRef] [PubMed]
  112. Sabia, P.J.; Powers, E.R.; Ragosta, M.; Sarembock, I.J.; Burwell, L.R.; Kaul, S. An Association between Collateral Blood Flow and Myocardial Viability in Patients with Recent Myocardial Infarction. N. Engl. J. Med. 1992, 327, 1825–1831. [Google Scholar] [CrossRef] [PubMed]
  113. Davidson, S.M.; Ferdinandy, P.; Andreadou, I.; Bøtker, H.E.; Heusch, G.; Ibáñez, B.; Ovize, M.; Schulz, R.; Yellon, D.M.; Hausenloy, D.J.; et al. Multitarget Strategies to Reduce Myocardial Ischemia/Reperfusion Injury: JACC Review Topic of the Week. J. Am. Coll. Cardiol. 2019, 73, 89–99. [Google Scholar] [CrossRef]
  114. Lecour, S.; Andreadou, I.; Bøtker, H.E.; Davidson, S.M.; Heusch, G.; Ruiz-Meana, M.; Schulz, R.; Zuurbier, C.J.; Ferdinandy, P.; Hausenloy, D.J.; et al. Improving Preclinical Assessment of Cardioprotective Therapies (IMPACT) criteria: Guidelines of the EU-CARDIOPROTECTION COST Action. Basic Res. Cardiol. 2021, 116, 52. [Google Scholar] [CrossRef]
  115. Warltier, D.C.; Al-Wathiqui, M.H.; Kampine, J.P.; Schmeling, W.T. Recovery of Contractile Function of Stunned Myocardium in Chronically Instrumented Dogs is Enhanced by Halothane or Isoflurane. Anesthesiology 1988, 69, 552–565. [Google Scholar] [CrossRef]
  116. Weber, N.C.; Goletz, C.; Huhn, R.; Grueber, Y.; Preckel, B.; Schlack, W.; Ebel, D. Blockade of anaesthetic-induced preconditioning in the hyperglycaemic myocardium: The regulation of different mitogen-activated protein kinases. Eur. J. Pharmacol. 2008, 592, 48–54. [Google Scholar] [CrossRef] [PubMed]
  117. Ge, Z.D.; Pravdic, D.; Bienengraeber, M.; Pratt, P.F., Jr.; Auchampach, J.A.; Gross, G.J.; Kersten, J.R.; Warltier, D.C. Isoflurane postconditioning protects against reperfusion injury by preventing mitochondrial permeability transition by an endothelial nitric oxide synthase-dependent mechanism. Anesthesiology 2010, 112, 73–85. [Google Scholar] [CrossRef]
  118. Lotz, C.; Kehl, F. Volatile anesthetic-induced cardiac protection: Molecular mechanisms, clinical aspects, and interactions with nonvolatile agents. J. Cardiothorac. Vasc. Anesth. 2015, 29, 749–760. [Google Scholar] [CrossRef]
  119. Ma, L.; Zhu, J.; Gao, Q.; Rebecchi, M.J.; Wang, Q.; Liu, L. Restoring Pharmacologic Preconditioning in the Aging Heart: Role of Mitophagy/Autophagy. J. Gerontol. A Biol. Sci. Med. Sci. 2017, 72, 489–498. [Google Scholar] [CrossRef]
  120. Wu, D.; Zhang, K.; Hu, P. The Role of Autophagy in Acute Myocardial Infarction. Front. Pharmacol. 2019, 10, 551. [Google Scholar] [CrossRef]
  121. Hamacher-Brady, A.; Brady, N.R.; Logue, S.E.; Sayen, M.R.; Jinno, M.; Kirshenbaum, L.A.; Gottlieb, R.A.; Gustafsson, Å.B. Response to myocardial ischemia/reperfusion injury involves Bnip3 and autophagy. Cell Death Differ. 2007, 14, 146–157. [Google Scholar] [CrossRef] [PubMed]
  122. Yang, C.; Jiao, Y.; Yan, N.; Wu, B.; Ren, Y.; Li, H.; Sun, J.; Gao, J. NOD2 mediates isoflurane preconditioning-induced protection of myocardial injury. Neurosci. Lett. 2017, 637, 154–160. [Google Scholar] [CrossRef] [PubMed]
  123. Xu, F.; Qiao, S.; Li, H.; Deng, Y.; Wang, C.; An, J. The Effect of Mitochondrial Complex I-Linked Respiration by Isoflurane Is Independent of Mitochondrial Nitric Oxide Production. Cardiorenal Med. 2018, 8, 113–122. [Google Scholar] [CrossRef] [PubMed]
  124. Javadov, S.; Jang, S.; Parodi-Rullán, R.; Khuchua, Z.; Kuznetsov, A.V. Mitochondrial permeability transition in cardiac ischemia-reperfusion: Whether cyclophilin D is a viable target for cardioprotection? Cell. Mol. Life Sci. 2017, 74, 2795–2813. [Google Scholar] [CrossRef] [PubMed]
  125. Harisseh, R.; Chiari, P.; Villedieu, C.; Sueur, P.; Abrial, M.; Fellahi, J.-L.; Ovize, M.; Gharib, A. Cyclophilin D Modulates the Cardiac Mitochondrial Target of Isoflurane, Sevoflurane, and Desflurane. J. Cardiovasc. Pharmacol. 2017, 69, 326–334. [Google Scholar] [CrossRef]
  126. Qiao, S.; Olson, J.M.; Paterson, M.; Yan, Y.; Zaja, I.; Liu, Y.; Riess, M.L.; Kersten, J.R.; Liang, M.; Warltier, D.C.; et al. MicroRNA-21 Mediates Isoflurane-induced Cardioprotection against Ischemia–Reperfusion Injury via Akt/Nitric Oxide Synthase/Mitochondrial Permeability Transition Pore Pathway. Anesthesiology 2015, 123, 786–798. [Google Scholar] [CrossRef] [PubMed]
  127. Olson, J.M.; Yan, Y.; Bai, X.; Ge, Z.D.; Liang, M.; Kriegel, A.J.; Twaroski, D.M.; Bosnjak, Z.J. Up-regulation of microRNA-21 mediates isoflurane-induced protection of cardiomyocytes. Anesthesiology 2015, 122, 795–805. [Google Scholar] [CrossRef] [PubMed]
  128. Liu, H.J.; Liu, B. Inhibition of MicroRNA-23 Contributes to the Isoflurane-Mediated Cardioprotection against Oxidative Stress. Cardiovasc. Toxicol. 2018, 18, 450–458. [Google Scholar] [CrossRef] [PubMed]
  129. Lemoine, S.; Tritapepe, L.; Hanouz, J.L.; Puddu, P.E. The mechanisms of cardio-protective effects of desflurane and sevoflurane at the time of reperfusion: Anaesthetic post-conditioning potentially translatable to humans? Br. J. Anaesth. 2016, 116, 456–475. [Google Scholar] [CrossRef]
  130. Chen, S.; Lotz, C.; Roewer, N.; Broscheit, J.-A. Comparison of volatile anesthetic-induced preconditioning in cardiac and cerebral system: Molecular mechanisms and clinical aspects. Eur. J. Med. Res. 2018, 23, 10. [Google Scholar] [CrossRef]
  131. Torregroza, C.; Raupach, A.; Feige, K.; Weber, N.C.; Hollmann, M.W.; Huhn, R. Perioperative Cardioprotection: General Mechanisms and Pharmacological Approaches. Anesth. Analg. 2020, 131, 1765–1780. [Google Scholar] [CrossRef] [PubMed]
  132. Van Allen, N.R.; Krafft, P.R.; Leitzke, A.S.; Applegate, R.L., 2nd; Tang, J.; Zhang, J.H. The role of Volatile Anesthetics in Cardioprotection: A systematic review. Med. Gas Res. 2012, 2, 22. [Google Scholar] [CrossRef] [PubMed]
  133. Wu, J.; Yu, J.; Xie, P.; Maimaitili, Y.; Wang, J.; Yang, L.; Ma, H.; Zhang, X.; Yang, Y.; Zheng, H. Sevoflurane postconditioning protects the myocardium against ischemia/reperfusion injury via activation of the JAK2–STAT3 pathway. PeerJ 2017, 5, e3196. [Google Scholar] [CrossRef] [PubMed]
  134. Lu, Y.; Bu, M.; Yun, H. Sevoflurane prevents hypoxia/reoxygenation-induced cardiomyocyte apoptosis by inhibiting PI3KC3-mediated autophagy. Hum. Cell 2019, 32, 150–159. [Google Scholar] [CrossRef] [PubMed]
  135. Qiao, S.; Sun, Y.; Sun, B.; Wang, A.; Qiu, J.; Hong, L.; An, J.-Z.; Wang, C.; Zhang, H.-L. Sevoflurane postconditioning protects against myocardial ischemia/reperfusion injury by restoring autophagic flux via an NO-dependent mechanism. Acta Pharmacol. Sin. 2018, 40, 35–45. [Google Scholar] [CrossRef]
  136. Qian, B.; Yang, Y.; Yao, Y.; Liao, Y.; Lin, Y. Upregulation of vascular endothelial growth factor receptor-1 contributes to sevoflurane preconditioning–mediated cardioprotection. Drug Des. Dev. Ther. 2018, 12, 769–776. [Google Scholar] [CrossRef]
  137. Huang, G.; Hao, F.; Hu, X. Downregulation of microRNA-155 stimulates sevoflurane-mediated cardioprotection against myocardial ischemia/reperfusion injury by binding to SIRT1 in mice. J. Cell. Biochem. 2019, 120, 15494–15505. [Google Scholar] [CrossRef]
  138. Xie, D.; Zhao, J.; Guo, R.; Jiao, L.; Zhang, Y.; Lau, W.B.; Lopez, B.; Christopher, T.; Gao, E.; Cao, J.; et al. Sevoflurane Pre-conditioning Ameliorates Diabetic Myocardial Ischemia/Reperfusion Injury Via Differential Regulation of p38 and ERK. Sci. Rep. 2020, 10, 23. [Google Scholar] [CrossRef]
  139. Zhong, C.Y.; Qiu, H.; Chen, J.; Liu, H. Effects of volatile anesthetic preconditioning on expression of NFkB-regulated genes in aged rat myocardium. J. Biomed. Res. 2017, 33, 264–270. [Google Scholar] [CrossRef]
  140. Smit, K.F.; Weber, N.C.; Hollmann, M.W.; Preckel, B. Noble gases as cardioprotectants-translatability and mechanism. Br. J. Pharmacol. 2015, 172, 2062–2073. [Google Scholar] [CrossRef]
  141. Weber, N.C.; Preckel, B. Gaseous mediators: An updated review on the effects of helium beyond blowing up balloons. Intensive Care Med. Exp. 2019, 7, 73. [Google Scholar] [CrossRef] [PubMed]
  142. Anna, R.; Rolf, R.; Mark, C. Update of the organoprotective properties of xenon and argon: From bench to beside. Intensive Care Med. Exp. 2020, 8, 11. [Google Scholar] [CrossRef] [PubMed]
  143. Roth, S.; Torregroza, C.; Feige, K.; Preckel, B.; Hollmann, M.W.; Weber, N.C.; Huhn, R. Pharmacological Conditioning of the Heart: An Update on Experimental Developments and Clinical Implications. Int. J. Mol. Sci. 2021, 22, 2519. [Google Scholar] [CrossRef] [PubMed]
  144. De Deken, J.; Rex, S.; Monbaliu, D.; Pirenne, J.; Jochmans, I. The Efficacy of Noble Gases in the Attenuation of Ischemia Reperfusion Injury: A Systematic Review and Meta-Analyses. Crit. Care Med. 2016, 44, e886–e896. [Google Scholar] [CrossRef] [PubMed]
  145. Pagel, P.S.; Krolikowski, J.G.; Shim, Y.H.; Venkatapuram, S.; Kersten, J.R.; Weihrauch, D.; Warltier, D.C.; Pratt, P.F., Jr. Noble gases without anesthetic properties protect myocardium against infarction by activating prosurvival signaling kinases and inhibiting mitochondrial permeability transition in vivo. Anesth. Analg. 2007, 105, 562–569. [Google Scholar] [CrossRef] [PubMed]
  146. Weber, N.C.; Toma, O.; Wolter, J.I.; Obal, D.; Müllenheim, J.; Preckel, B.; Schlack, W. The noble gas xenon induces pharmacological preconditioning in the rat heart in vivo via induction of PKC-epsilon and p38 MAPK. Br. J. Pharmacol. 2005, 144, 123–132. [Google Scholar] [CrossRef]
  147. Weber, N.C.; Toma, O.; Wolter, J.I.; Wirthle, N.M.; Schlack, W.; Preckel, B. Mechanisms of xenon- and isoflurane-induced preconditioning-a potential link to the cytoskeleton via the MAPKAPK-2/HSP27 pathway. Br. J. Pharmacol. 2005, 146, 445–455. [Google Scholar] [CrossRef]
  148. Weber, N.C.; Stursberg, J.; Wirthle, N.M.; Toma, O.; Schlack, W.; Preckel, B. Xenon preconditioning differently regulates p44/42 MAPK (ERK 1/2) and p46/54 MAPK (JNK 1/2 and 3) in vivo. Br. J. Anaesth. 2006, 97, 298–306. [Google Scholar] [CrossRef]
  149. Li, Q.; Lian, C.; Zhou, R.; Li, T.; Xiang, X.; Liu, B. Pretreatment With Xenon Protected Immature Rabbit Heart From Ischaemia/Reperfusion Injury by Opening of the mitoKATP Channel. Heart Lung Circ. 2013, 22, 276–283. [Google Scholar] [CrossRef]
  150. Roehl, A.B.; Funcke, S.; Becker, M.M.; Goetzenich, A.; Bleilevens, C.; Rossaint, R.; Steendijk, P.; Hein, M. Xenon and Isoflurane Reduce Left Ventricular Remodeling after Myocardial Infarction in the Rat. Anesthesiology 2013, 118, 1385–1394. [Google Scholar] [CrossRef]
  151. Hein, M.; Roehl, A.B.; Baumert, J.H.; Bleilevens, C.; Fischer, S.; Steendijk, P.; Rossaint, R. Xenon and isoflurane improved biventricular function during right ventricular ischemia and reperfusion. Acta Anaesthesiol. Scand. 2010, 54, 470–478. [Google Scholar] [CrossRef] [PubMed]
  152. Weber, N.C.; Frässdorf, J.; Ratajczak, C.; Grueber, Y.; Schlack, W.; Hollmann, M.W.; Preckel, B. Xenon induces late cardiac preconditioning in vivo: A role for cyclooxygenase 2? Anesth. Analg. 2008, 107, 1807–1813. [Google Scholar] [CrossRef] [PubMed]
  153. Pagel, P.S.; Krolikowski, J.G.; Pratt, P.F.; Shim, Y.H.; Amour, J.; Warltier, D.C.; Weihrauch, D. The Mechanism of Helium-Induced Preconditioning: A Direct Role for Nitric Oxide in Rabbits. Anesth. Analg. 2008, 107, 762–768. [Google Scholar] [CrossRef] [PubMed]
  154. Pagel, P.S.; Krolikowski, J.G.; Pratt, P.F.; Shim, Y.H.; Amour, J.; Warltier, D.C.; Weihrauch, D. Reactive Oxygen Species and Mitochondrial Adenosine Triphosphate–Regulated Potassium Channels Mediate Helium-Induced Preconditioning against Myocardial Infarction In Vivo. J. Cardiothorac. Vasc. Anesth. 2008, 22, 554–559. [Google Scholar] [CrossRef] [PubMed]
  155. Pagel, P.S.; Krolikowski, J.G.; Pratt, P.F.; Shim, Y.H.; Amour, J.; Warltier, D.C.; Weihrauch, D. Inhibition of Glycogen Synthase Kinase or the Apoptotic Protein p53 Lowers the Threshold of Helium Cardioprotection In Vivo: The Role of Mitochondrial Permeability Transition. Anesth. Analg. 2008, 107, 769–775. [Google Scholar] [CrossRef]
  156. Pagel, P.S.; Krolikowski, J.G. Transient metabolic alkalosis during early reperfusion abolishes helium preconditioning against myocardial infarction: Restoration of cardioprotection by cyclosporin A in rabbits. Anesth. Analg. 2009, 108, 1076–1082. [Google Scholar] [CrossRef]
  157. Pagel, P.S.; Krolikowski, J.G.; Amour, J.; Warltier, D.C.; Weihrauch, D. Morphine reduces the threshold of helium preconditioning against myocardial infarction: The role of opioid receptors in rabbits. J. Cardiothorac. Vasc. Anesth. 2009, 23, 619–624. [Google Scholar] [CrossRef]
  158. Huhn, R.; Heinen, A.; Weber, N.C.; Hieber, S.; Hollmann, M.W.; Schlack, W.; Preckel, B. Helium-induced late preconditioning in the rat heart in vivo. Br. J. Anaesth. 2009, 102, 614–619. [Google Scholar] [CrossRef]
  159. Oei, G.T.M.L.; Aslami, H.; Kerindongo, R.P.; Steenstra, R.J.; Beurskens, C.J.P.; Boer, A.M.T.-D.; Juffermans, N.P.; Hollmann, M.W.; Preckel, B.; Weber, N.C. Prolonged Helium Postconditioning Protocols during Early Reperfusion Do Not Induce Cardioprotection in the Rat Heart In Vivo: Role of Inflammatory Cytokines. J. Immunol. Res. 2015, 2015, 216798. [Google Scholar] [CrossRef]
  160. Oei, G.T.; Heger, M.; van Golen, R.F.; Alles, L.K.; Flick, M.; van der Wal, A.C.; van Gulik, T.M.; Hollmann, M.W.; Preckel, B.; Weber, N.C. Reduction of cardiac cell death after helium postconditioning in rats: Transcriptional analysis of cell death and survival pathways. Mol. Med. 2015, 20, 516–526. [Google Scholar] [CrossRef]
  161. Heinen, A.; Huhn, R.; Smeele, K.M.; Zuurbier, C.J.; Schlack, W.; Preckel, B.; Weber, N.C.; Hollmann, M.W. Helium-induced preconditioning in young and old rat heart: Impact of mitochondrial Ca(2+) -sensitive potassium channel activation. Anesthesiology 2008, 109, 830–836. [Google Scholar] [CrossRef] [PubMed]
  162. Huhn, R.; Weber, N.C.; Preckel, B.; Schlack, W.; Bauer, I.; Hollmann, M.W.; Heinen, A. Age-related loss of cardiac preconditioning: Impact of protein kinase A. Exp. Gerontol. 2012, 47, 116–121. [Google Scholar] [CrossRef] [PubMed]
  163. Oei, G.T.; Huhn, R.; Heinen, A.; Hollmann, M.W.; Schlack, W.S.; Preckel, B.; Weber, N.C. Helium-induced cardioprotection of healthy and hypertensive rat myocardium in vivo. Eur. J. Pharmacol. 2012, 684, 125–131. [Google Scholar] [CrossRef] [PubMed]
  164. Foster, C.R.; Satomi, S.; Kato, Y.; Patel, H.H. The caveolar-mitochondrial interface: Regulation of cellular metabolism in physiology and pathophysiology. Biochem. Soc. Trans. 2020, 48, 165–177. [Google Scholar] [CrossRef]
  165. Schilling, J.M.; Head, B.P.; Patel, H.H. Caveolins as Regulators of Stress Adaptation. Mol. Pharmacol. 2018, 93, 277–285. [Google Scholar] [CrossRef]
  166. Ohi, M.D.; Kenworthy, A.K. Emerging Insights into the Molecular Architecture of Caveolin-1. J. Membr. Biol. 2022, 255, 375–383. [Google Scholar] [CrossRef]
  167. Flick, M.; Albrecht, M.; Oei, G.T.; Steenstra, R.; Kerindongo, R.P.; Zuurbier, C.J.; Patel, H.H.; Hollmann, M.W.; Preckel, B.; Weber, N.C. Helium postconditioning regulates expression of caveolin-1 and -3 and induces RISK pathway activation after ischaemia/reperfusion in cardiac tissue of rats. Eur. J. Pharmacol. 2016, 791, 718–725. [Google Scholar] [CrossRef]
  168. Aehling, C.; Weber, N.C.; Zuurbier, C.J.; Preckel, B.; Galmbacher, R.; Stefan, K.; Hollmann, M.W.; Popp, E.; Knapp, J. Effects of combined helium pre/post-conditioning on the brain and heart in a rat resuscitation model. Acta Anaesthesiol. Scand. 2017, 62, 63–74. [Google Scholar] [CrossRef]
  169. Weber, N.C.; Schilling, J.M.; Warmbrunn, M.V.; Dhanani, M.; Kerindongo, R.; Siamwala, J.; Song, Y.; Zemljic-Harpf, A.E.; Fannon, M.J.; Hollmann, M.W.; et al. Helium-Induced Changes in Circulating Caveolin in Mice Suggest a Novel Mechanism of Cardiac Protection. Int. J. Mol. Sci. 2019, 20, 2640. [Google Scholar] [CrossRef]
  170. Jelemenský, M.; Kovácsházi, C.; Ferenczyová, K.; Hofbauerová, M.; Kiss, B.; Pállinger, É.; Kittel, Á.; Sayour, V.N.; Görbe, A.; Pelyhe, C.; et al. Helium Conditioning Increases Cardiac Fibroblast Migration Which Effect Is Not Propagated via Soluble Factors or Extracellular Vesicles. Int. J. Mol. Sci. 2021, 22, 10504. [Google Scholar] [CrossRef]
  171. Likhvantsev, V.V.; Landoni, G.; Levikov, D.I.; Grebenchikov, O.A.; Skripkin, Y.V.; Cherpakov, R.A. Sevoflurane Versus Total Intravenous Anesthesia for Isolated Coronary Artery Bypass Surgery with Cardiopulmonary Bypass: A Randomized Trial. J. Cardiothorac. Vasc. Anesth. 2016, 30, 1221–1227. [Google Scholar] [CrossRef] [PubMed]
  172. Landoni, G.; Guarracino, F.; Cariello, C.; Franco, A.; Baldassarri, R.; Borghi, G.; Covello, R.D.; Gerli, C.; Crivellari, M.; Zangrillo, A. Volatile compared with total intravenous anaesthesia in patients undergoing high-risk cardiac surgery: A randomized multicentre study. Br. J. Anaesth. 2014, 113, 955–963. [Google Scholar] [CrossRef] [PubMed]
  173. Landoni, G.; Lomivorotov, V.V.; Nigro Neto, C.; Monaco, F.; Pasyuga, V.V.; Bradic, N.; Lembo, R.; Gazivoda, G.; Likhvantsev, V.V.; Lei, C.; et al. Volatile Anesthetics versus Total Intravenous Anesthesia for Cardiac Surgery. N. Engl. J. Med. 2019, 380, 1214–1225. [Google Scholar] [CrossRef] [PubMed]
  174. Lurati Buse, G.A.; Schumacher, P.; Seeberger, E.; Studer, W.; Schuman, R.M.; Fassl, J.; Kasper, J.; Filipovic, M.; Bolliger, D.; Seeberger, M.D. Randomized comparison of sevoflurane versus propofol to reduce perioperative myocardial ischemia in patients undergoing noncardiac surgery. Circulation 2012, 126, 2696–2704. [Google Scholar] [CrossRef]
  175. Devereaux, P.J.; Sessler, D.I.; Leslie, K.; Kurz, A.; Mrkobrada, M.; Alonso-Coello, P.; Villar, J.C.; Sigamani, A.; Biccard, B.M.; Meyhoff, C.S.; et al. Clonidine in patients undergoing noncardiac surgery. N. Engl. J. Med. 2014, 370, 1504–1513. [Google Scholar] [CrossRef] [PubMed]
  176. Hofland, J.; Ouattara, A.; Fellahi, J.L.; Gruenewald, M.; Hazebroucq, J.; Ecoffey, C.; Joseph, P.; Heringlake, M.; Steib, A.; Coburn, M.; et al. Effect of Xenon Anesthesia Compared to Sevoflurane and Total Intravenous Anesthesia for Coronary Artery Bypass Graft Surgery on Postoperative Cardiac Troponin Release: An International, Multicenter, Phase 3, Single-blinded, Randomized Noninferiority Trial. Anesthesiology 2017, 127, 918–933. [Google Scholar] [CrossRef]
  177. Mehta, R.H.; Leimberger, J.D.; van Diepen, S.; Meza, J.; Wang, A.; Jankowich, R.; Harrison, R.W.; Hay, D.; Fremes, S.; Duncan, A.; et al. Levosimendan in Patients with Left Ventricular Dysfunction Undergoing Cardiac Surgery. N. Engl. J. Med. 2017, 376, 2032–2042. [Google Scholar] [CrossRef]
  178. Uhlig, C.; Bluth, T.; Schwarz, K.; Deckert, S.; Heinrich, L.; De Hert, S.; Landoni, G.; Serpa Neto, A.; Schultz, M.J.; Pelosi, P.; et al. Effects of Volatile Anesthetics on Mortality and Postoperative Pulmonary and Other Complications in Patients Undergoing Surgery: A Systematic Review and Meta-analysis. Anesthesiology 2016, 124, 1230–1245. [Google Scholar] [CrossRef]
  179. Smit, K.F.; Oei, G.T.M.L.; Brevoord, D.; Stroes, E.S.; Nieuwland, R.; Schlack, W.S.; Hollmann, M.W.; Weber, N.C.; Preckel, B. Helium Induces Preconditioning in Human Endothelium In Vivo. Anesthesiology 2013, 118, 95–104. [Google Scholar] [CrossRef]
  180. Smit, K.F.; Brevoord, D.; de Hert, S.; de Mol, B.A.; Kerindongo, R.P.; Van Dieren, S.; Schlack, W.S.; Hollmann, M.W.; Weber, N.C.; Preckel, B. Effect of helium pre- or postconditioning on signal transduction kinases in patients undergoing coronary artery bypass graft surgery. J. Transl. Med. 2016, 14, 294. [Google Scholar] [CrossRef]
  181. Arola, O.; Saraste, A.; Laitio, R.; Airaksinen, J.; Hynninen, M.; Bäcklund, M.; Ylikoski, E.; Wennervirta, J.; Pietilä, M.; Roine, R.O.; et al. Inhaled Xenon Attenuates Myocardial Damage in Comatose Survivors of Out-of-Hospital Cardiac Arrest: The Xe-Hypotheca Trial. J. Am. Coll. Cardiol. 2017, 70, 2652–2660. [Google Scholar] [CrossRef] [PubMed]
  182. Saraste, A.; Ballo, H.; Arola, O.; Laitio, R.; Airaksinen, J.; Hynninen, M.; Bäcklund, M.; Ylikoski, E.; Wennervirta, J.; Pietilä, M.; et al. Effect of Inhaled Xenon on Cardiac Function in Comatose Survivors of Out-of-Hospital Cardiac Arrest—A Substudy of the Xenon in Combination With Hypothermia After Cardiac Arrest Trial. Crit. Care Explor. 2021, 3, e0502. [Google Scholar] [CrossRef]
  183. Derwall, M.; Ebeling, A.; Nolte, K.W.; Weis, J.; Rossaint, R.; Ichinose, F.; Nix, C.; Fries, M.; Brücken, A. Inhaled nitric oxide improves transpulmonary blood flow and clinical outcomes after prolonged cardiac arrest: A large animal study. Crit. Care 2015, 19, 328. [Google Scholar] [CrossRef] [PubMed]
  184. Liu, X.; Huang, Y.; Pokreisz, P.; Vermeersch, P.; Marsboom, G.; Swinnen, M.; Verbeken, E.; Santos, J.; Pellens, M.; Gillijns, H.; et al. Nitric Oxide Inhalation Improves Microvascular Flow and Decreases Infarction Size After Myocardial Ischemia and Reperfusion. J. Am. Coll. Cardiol. 2007, 50, 808–817. [Google Scholar] [CrossRef] [PubMed]
  185. Minamishima, S.; Kida, K.; Tokuda, K.; Wang, H.; Sips, P.Y.; Kosugi, S.; Mandeville, J.B.; Buys, E.S.; Brouckaert, P.; Liu, P.K.; et al. Inhaled Nitric Oxide Improves Outcomes After Successful Cardiopulmonary Resuscitation in Mice. Circulation 2011, 124, 1645–1653. [Google Scholar] [CrossRef]
  186. Russo, I.; Barale, C.; Melchionda, E.; Penna, C.; Pagliaro, P. Platelets and Cardioprotection: The Role of Nitric Oxide and Carbon Oxide. Int. J. Mol. Sci. 2023, 24, 6107. [Google Scholar] [CrossRef]
  187. Hayashida, K.; Sano, M.; Ohsawa, I.; Shinmura, K.; Tamaki, K.; Kimura, K.; Endo, J.; Katayama, T.; Kawamura, A.; Kohsaka, S.; et al. Inhalation of hydrogen gas reduces infarct size in the rat model of myocardial ischemia–reperfusion injury. Biochem. Biophys. Res. Commun. 2008, 373, 30–35. [Google Scholar] [CrossRef]
  188. Janssens, S.P.; Bogaert, J.; Zalewski, J.; Toth, A.; Adriaenssens, T.; Belmans, A.; Bennett, J.; Claus, P.; Desmet, W.; Dubois, C.; et al. Nitric oxide for inhalation in ST-elevation myocardial infarction (NOMI): A multicentre, double-blind, randomized controlled trial. Eur. Heart J. 2018, 39, 2717–2725. [Google Scholar] [CrossRef]
  189. Yuan, T.; Zhao, J.N.; Bao, N.R. Hydrogen applications: Advances in the field of medical therapy. Med. Gas Res. 2023, 13, 99–107. [Google Scholar]
  190. Ohsawa, I.; Ishikawa, M.; Takahashi, K.; Watanabe, M.; Nishimaki, K.; Yamagata, K.; Katsura, K.; Katayama, Y.; Asoh, S.; Ohta, S. Hydrogen acts as a therapeutic antioxidant by selectively reducing cytotoxic oxygen radicals. Nat. Med. 2007, 13, 688–694. [Google Scholar] [CrossRef]
  191. Chambers, D.J.; Fallouh, H.B. Cardioplegia and cardiac surgery: Pharmacological arrest and cardioprotection during global ischemia and reperfusion. Pharmacol. Ther. 2010, 127, 41–52. [Google Scholar] [CrossRef]
  192. Francica, A.; Tonelli, F.; Rossetti, C.; Tropea, I.; Luciani, G.B.; Faggian, G.; Dobson, G.P.; Onorati, F. Cardioplegia between Evolution and Revolution: From Depolarized to Polarized Cardiac Arrest in Adult Cardiac Surgery. J. Clin. Med. 2021, 10, 4485. [Google Scholar] [CrossRef]
  193. Pinto, A.R.; Ilinykh, A.; Ivey, M.J.; Kuwabara, J.T.; D’Antoni, M.L.; Debuque, R.; Chandran, A.; Wang, L.; Arora, K.; Rosenthal, N.A.; et al. Revisiting Cardiac Cellular Composition. Circ. Res. 2016, 118, 400–409. [Google Scholar] [CrossRef]
  194. Colliva, A.; Braga, L.; Giacca, M.; Zacchigna, S. Endothelial cell–cardiomyocyte crosstalk in heart development and disease. J. Physiol. 2019, 598, 2923–2939. [Google Scholar] [CrossRef] [PubMed]
  195. Os, M.M.D.; Brom, C.E.V.D.; van Leeuwen, A.L.I.; Dekker, N.A.M. Microcirculatory perfusion disturbances following cardiopulmonary bypass: A systematic review. Crit. Care 2020, 24, 218. [Google Scholar] [CrossRef]
  196. Dekker, N.A.M.; van Leeuwen, A.L.I.; van de Ven, P.M.; de Vries, R.; Hordijk, P.L.; Boer, C.; van den Brom, C.E. Pharmacological interventions to reduce edema following cardiopulmonary bypass: A systematic review and meta-analysis. J. Crit. Care 2020, 56, 63–72. [Google Scholar] [CrossRef] [PubMed]
  197. Shen, X.; Tan, Z.; Zhong, X.; Tian, Y.; Wang, X.; Yu, B.; Ramirez-Correa, G.; Murphy, A.; Gabrielson, K.; Paolocci, N.; et al. Endocardial endothelium is a key determinant of force-frequency relationship in rat ventricular myocardium. J. Appl. Physiol. 2013, 115, 383–393. [Google Scholar] [CrossRef]
  198. Kuramochi, Y.; Cote, G.M.; Guo, X.; Lebrasseur, N.K.; Cui, L.; Liao, R.; Sawyer, D.B. Cardiac endothelial cells regulate reactive oxygen species-induced cardiomyocyte apoptosis through neuregulin-1beta/erbB4 signaling. J. Biol. Chem. 2004, 279, 51141–51147. [Google Scholar] [CrossRef]
  199. Hernández-Reséndiz, S.; Muñoz-Vega, M.; E Contreras, W.; E Crespo-Avilan, G.; Rodriguez-Montesinos, J.; Arias-Carrión, O.; Pérez-Méndez, O.; A Boisvert, W.; Preissner, K.T.; A Cabrera-Fuentes, H. Responses of Endothelial Cells Towards Ischemic Conditioning Following Acute Myocardial Infarction. Cond. Med. 2018, 1, 247–258. [Google Scholar] [PubMed]
  200. Zakkar, M.; Guida, G.; Suleiman, M.-S.; Angelini, G.D. Cardiopulmonary Bypass and Oxidative Stress. Oxidative Med. Cell. Longev. 2015, 2015, 189863. [Google Scholar] [CrossRef]
  201. Kiziltepe, U.; Tunçtan, B.; Eyileten, Z.B.; Sirlak, M.; Arikbuku, M.; Tasoz, R.; Uysalel, A.; Ozyurda, U. Efficiency of l-arginine enriched cardioplegia and non-cardioplegic reperfusion in ischemic hearts. Int. J. Cardiol. 2004, 97, 93–100. [Google Scholar] [CrossRef]
  202. Orhan, G.; Yapici, N.; Yuksel, M.; Sargin, M.; Şenay, S.; Yalçin, A.S.; Aykaç, Z.; Aka, S.A. Effects of N-acetylcysteine on myocardial ischemia–reperfusion injury in bypass surgery. Heart Vessel. 2006, 21, 42–47. [Google Scholar] [CrossRef]
  203. Prabhu, A.; Sujatha, D.; Kanagarajan, N.; Vijayalakshmi, M.; Ninan, B. Effect of N-Acetylcysteine in Attenuating Ischemic Reperfusion Injury in Patients Undergoing Coronary Artery Bypass Grafting with Cardiopulmonary Bypass. Ann. Vasc. Surg. 2009, 23, 645–651. [Google Scholar] [CrossRef] [PubMed]
  204. Hallström, S.; Franz, M.; Gasser, H.; Vodrazka, M.; Semsroth, S.; Losert, U.M.; Haisjackl, M.; Podesser, B.K.; Malinski, T. S-nitroso human serum albumin reduces ischaemia/reperfusion injury in the pig heart after unprotected warm ischaemia. Cardiovasc. Res. 2007, 77, 506–514. [Google Scholar] [CrossRef] [PubMed]
  205. Peng, Y.-W.; Major, T.; Deatrick, K.B.; Mohammed, A.; Jeakle, M.; Charpie, J.R. Nicorandil attenuates ventricular dysfunction and organ injury after cardiopulmonary bypass. Int. J. Cardiol. 2022, 368, 62–68. [Google Scholar] [CrossRef] [PubMed]
  206. Yamamoto, S.; Yamada, T.; Kotake, Y.; Takeda, J. Cardioprotective Effects of Nicorandil in Patients Undergoing On-Pump Coronary Artery Bypass Surgery. J. Cardiothorac. Vasc. Anesth. 2008, 22, 548–553. [Google Scholar] [CrossRef] [PubMed]
  207. Kawamura, T.; Kadosaki, M.; Nara, N.; Wei, J.; Endo, S.; Inada, K. Nicorandil attenuates NF-kappaB activation, adhesion molecule expression, and cytokine production in patients with coronary artery bypass surgery. Shock 2005, 24, 103–108. [Google Scholar] [CrossRef]
  208. Yavuz, T.; Altuntas, I.; Odabasi, D.; Delibas, N.; Ocal, A.; Ibrisim, E.; Kutsal, A. Beneficial Effect of the Addition of Nitroglycerin to the Cardioplegic Solution on Lipid Peroxidation during Coronary Artery Bypass Surgery. Int. Heart J. 2005, 46, 45–55. [Google Scholar] [CrossRef]
  209. Hamarneh, A.; Ho, A.F.W.; Bulluck, H.; Sivaraman, V.; Ricciardi, F.; Nicholas, J.; Shanahan, H.; Hardman, E.A.; Wicks, P.; Ramlall, M.; et al. Negative interaction between nitrates and remote ischemic preconditioning in patients undergoing cardiac surgery: The ERIC-GTN and ERICCA studies. Basic Res. Cardiol. 2022, 117, 31. [Google Scholar] [CrossRef]
  210. Clark, D., 3rd; Tesseneer, S.; Tribble, C.G. Nitroglycerin and sodium nitroprusside: Potential contributors to postoperative bleeding? Heart Surg. Forum 2012, 15, E92–E96. [Google Scholar] [CrossRef]
  211. Rubio-Gayosso, I.; Platts, S.H.; Duling, B.R. Reactive oxygen species mediate modification of glycocalyx during ischemia-reperfusion injury. Am. J. Physiol. Heart Circ. Physiol. 2006, 290, H2247–H2256. [Google Scholar] [CrossRef]
  212. Mulivor, A.W.; Lipowsky, H.H. Inflammation- and ischemia-induced shedding of venular glycocalyx. Am. J. Physiol. Circ. Physiol. 2004, 286, H1672–H1680. [Google Scholar] [CrossRef] [PubMed]
  213. Bonfig, N.L.; Soukup, C.R.; Shah, A.A.; Olet, S.; Davidson, S.J.; Schmidt, C.W.; Peterson, R.; Henry, T.D.; Traverse, J.H. Increasing myocardial edema is associated with greater microvascular obstruction in ST-segment elevation myocardial infarction. Am. J. Physiol. Circ. Physiol. 2022, 323, H818–H824. [Google Scholar] [CrossRef]
  214. Palmier, M.; Cornet, E.; Renet, S.; Dumesnil, A.; Perzo, N.; Cohen, Q.; Richard, V.; Plissonnier, D. A Supraceliac Aortic Cross Clamping Model to Explore Remote Lung Injury and the Endothelial Glycocalyx. Ann. Vasc. Surg. 2022, S0890-5096(22)00906-2. [Google Scholar] [CrossRef] [PubMed]
  215. Dekker, N.A.M.; Veerhoek, D.; Koning, N.J.; van Leeuwen, A.L.I.; Elbers, P.W.G.; Brom, C.E.V.D.; Vonk, A.B.A.; Boer, C. Postoperative microcirculatory perfusion and endothelial glycocalyx shedding following cardiac surgery with cardiopulmonary bypass. Anaesthesia 2019, 74, 609–618. [Google Scholar] [CrossRef] [PubMed]
  216. Passov, A.; Schramko, A.; Salminen, U.-S.; Aittomäki, J.; Andersson, S.; Pesonen, E. Endothelial glycocalyx during early reperfusion in patients undergoing cardiac surgery. PLoS ONE 2021, 16, e0251747. [Google Scholar] [CrossRef]
  217. Ikonomidis, I.; Vlastos, D.; Andreadou, I.; Gazouli, M.; Efentakis, P.; Varoudi, M.; Makavos, G.; Kapelouzou, A.; Lekakis, J.; Parissis, J.; et al. Vascular conditioning prevents adverse left ventricular remodelling after acute myocardial infarction: A randomised remote conditioning study. Basic Res. Cardiol. 2021, 116, 9. [Google Scholar] [CrossRef]
  218. Annecke, T.; Chappell, D.; Chen, C.; Jacob, M.; Welsch, U.; Sommerhoff, C.P.; Rehm, M.; Conzen, P.F.; Becker, B.F. Sevoflurane preserves the endothelial glycocalyx against ischaemia-reperfusion injury. Br. J. Anaesth. 2010, 104, 414–421. [Google Scholar] [CrossRef] [PubMed]
  219. Fang, F.Q.; Sun, J.H.; Wu, Q.L.; Feng, L.Y.; Fan, Y.X.; Ye, J.X.; Gao, W.; He, G.L.; Wang, W.J. Protective effect of sevoflurane on vascular endothelial glycocalyx in patients undergoing heart valve surgery: A randomised controlled trial. Eur. J. Anaesthesiol. 2021, 38, 477–486. [Google Scholar] [CrossRef]
  220. Flameng, W.; Borgers, M.; Van Der Vusse, G.J.; Demeyere, R.; Vandermeersch, E.; Thoné, F.; Suy, R. Cardioprotective effects of lidoflazine in extensive aorta-coronary bypass grafting. J. Thorac. Cardiovasc. Surg. 1983, 85, 758–768. [Google Scholar] [CrossRef]
  221. Shen, A.C.; Jennings, R.B. Myocardial calcium and magnesium in acute ischemic injury. Am. J. Pathol. 1972, 67, 417–440. [Google Scholar] [PubMed]
  222. Serraino, G.F.; Jiritano, F.; Costa, D.; Ielapi, N.; Battaglia, D.; Bracale, U.M.; Mastroroberto, P.; Andreucci, M.; Serra, R. Metalloproteinases in Cardiac Surgery: A Systematic Review. Biomolecules 2023, 13, 113. [Google Scholar] [CrossRef] [PubMed]
  223. Gao, W.; Fang, F.; Xia, T.J.; Zhang, Y.; Sun, J.; Wu, Q.; Wang, W. Doxycycline can reduce glycocalyx shedding by inhibiting matrix metalloproteinases in patients undergoing cardiopulmonary bypass: A randomized controlled trial. Microvasc. Res. 2022, 142, 104381. [Google Scholar] [CrossRef] [PubMed]
  224. Schulze, C.J.; Castro, M.M.; Kandasamy, A.D.; Cena, J.; Bryden, C.; Wang, S.H.; Koshal, A.; Tsuyuki, R.T.; Finegan, B.A.; Schulz, R. Doxycycline Reduces Cardiac Matrix Metalloproteinase-2 Activity but Does not Ameliorate Myocardial Dysfunction During Reperfusion in Coronary Artery Bypass Patients Undergoing Cardiopulmonary Bypass. Crit. Care Med. 2013, 41, 2512–2520. [Google Scholar] [CrossRef] [PubMed]
  225. Dekker, N.A.; Veerhoek, D.; van Leeuwen, A.L.; Vonk, A.B.; Brom, C.E.V.D.; Boer, C. Microvascular Alterations during Cardiac Surgery Using a Heparin or Phosphorylcholine-Coated Circuit. J. Cardiothorac. Vasc. Anesth. 2020, 34, 912–919. [Google Scholar] [CrossRef]
  226. Selim, J.; Hamzaoui, M.; Ghemired, A.; Djerada, Z.; Chevalier, L.; Piton, N.; Besnier, E.; Clavier, T.; Dumesnil, A.; Renet, S.; et al. Priming of Cardiopulmonary Bypass with Human Albumin Decreases Endothelial Dysfunction after Pulmonary Ischemia–Reperfusion in an Animal Model. Int. J. Mol. Sci. 2022, 23, 8938. [Google Scholar] [CrossRef]
  227. Vahldieck, C.; Cianflone, E.; Fels, B.; Löning, S.; Depelmann, P.; Sabatino, J.; Salerno, N.; Karsten, C.M.; Torella, D.; Weil, J.; et al. Endothelial Glycocalyx and Cardiomyocyte Damage Is Prevented by Recombinant Syndecan-1 in Acute Myocardial Infarction. Am. J. Pathol. 2023, 193, 474–492. [Google Scholar] [CrossRef]
  228. Juffermans, N.P.; van den Brom, C.E.; Kleinveld, D.J.B. Targeting Endothelial Dysfunction in Acute Critical Illness to Reduce Organ Failure. Anesth. Analg. 2020, 131, 1708–1720. [Google Scholar] [CrossRef]
  229. Addya, S.; Shiroto, K.; Turoczi, T.; Zhan, L.; Kaga, S.; Fukuda, S.; Surrey, S.; Duan, L.-J.; Fong, G.-H.; Yamamoto, F. Ischemic preconditioning-mediated cardioprotection is disrupted in heterozygous Flt-1 (VEGFR-1) knockout mice. J. Mol. Cell. Cardiol. 2005, 38, 345–351. [Google Scholar] [CrossRef]
  230. Thirunavukkarasu, M.; Juhasz, B.; Zhan, L.; Menon, V.P.; Tosaki, A.; Otani, H.; Maulik, N. VEGFR1 (Flt-1+/-) gene knockout leads to the disruption of VEGF-mediated signaling through the nitric oxide/heme oxygenase pathway in ischemic preconditioned myocardium. Free Radic. Biol. Med. 2007, 42, 1487–1495. [Google Scholar] [CrossRef]
  231. Gonca, S.; Kiliçkan, L.; Dalçik, C.; Dalçik, H.; Bayindir, O. The cardioprotective effects of thoracal epidural anestesia are induced by the expression of vascular endothelial growth factor and inducible nitric oxide synthase in cardiopulmonary bypass surgery. J. Cardiovasc. Surg. 2007, 48, 93–102. [Google Scholar]
  232. Braile, M.; Marcella, S.; Cristinziano, L.; Galdiero, M.R.; Modestino, L.; Ferrara, A.L.; Varricchi, G.; Marone, G.; Loffredo, S. VEGF-A in Cardiomyocytes and Heart Diseases. Int. J. Mol. Sci. 2020, 21, 5294. [Google Scholar] [CrossRef] [PubMed]
  233. Tofukuji, M.; Metais, C.; Li, J.; Franklin, A.; Simons, M.; Sellke, F.W. Myocardial VEGF expression after cardiopulmonary bypass and cardioplegia. Circulation 1998, 98((Suppl. 19)), II242-6, discussion II247–II248. [Google Scholar]
  234. Koning, N.J.; Overmars, M.A.H.; Brom, C.E.V.D.; van Bezu, J.; Simon, L.E.; Vonk, A.B.A.; Girbes, A.R.J.; Amerongen, G.P.V.N.; Boer, C. Endothelial hyperpermeability after cardiac surgery with cardiopulmonary bypass as assessed using an in vitro bioassay for endothelial barrier function. Br. J. Anaesth. 2016, 116, 223–232. [Google Scholar] [CrossRef] [PubMed]
  235. Bai, Y.; Zhang, Y.; Yang, S.; Wu, M.; Fang, Y.; Feng, J.; Liu, B. Protective effect of vascular endothelial growth factor against cardiopulmonary bypass-associated acute kidney injury in beagles. Exp. Ther. Med. 2017, 15, 963–969. [Google Scholar] [CrossRef]
  236. Raza, Z.; Saleem, U.; Naureen, Z. Sphingosine 1-phosphate signaling in ischemia and reperfusion injury. Prostaglandins Other Lipid Mediat. 2020, 149, 106436. [Google Scholar] [CrossRef] [PubMed]
  237. Araibi, H.; van der Merwe, E.; Gwanyanya, A.; Kelly-Laubscher, R. The effect of sphingosine-1-phosphate on the endothelial glycocalyx during ischemia-reperfusion injury in the isolated rat heart. Microcirculation 2020, 27, e12612. [Google Scholar] [CrossRef]
  238. Samarska, I.V.; Bouma, H.R.; Buikema, H.; Mungroop, H.E.; Houwertjes, M.C.; Absalom, A.R.; Epema, A.H.; Henning, R.H. S1P1 Receptor Modulation Preserves Vascular Function in Mesenteric and Coronary Arteries after CPB in the Rat Independent of Depletion of Lymphocytes. PLoS ONE 2014, 9, e97196. [Google Scholar] [CrossRef]
  239. Greiwe, G.; Moritz, E.; Amschler, K.; Poppe, A.; Sarwari, H.; Nierhaus, A.; Kluge, S.; Reichenspurner, H.; Zoellner, C.; Schwedhelm, E.; et al. Dynamics of Vascular Protective and Immune Supportive Sphingosine-1-Phosphate During Cardiac Surgery. Front. Immunol. 2021, 12, 761475. [Google Scholar] [CrossRef]
  240. Clajus, C.; Lukasz, A.; David, S.; Hertel, B.; Lichtinghagen, R.; Parikh, S.M.; Simon, A.; Ismail, I.; Haller, H.; Kümpers, P. Angiopoietin-2 is a potential mediator of endothelial barrier dysfunction following cardiopulmonary bypass. Cytokine 2012, 60, 352–359. [Google Scholar] [CrossRef]
  241. Dekker, N.A.M.; Van Leeuwen, A.L.I.; Van Strien, W.W.J.; Majolée, J.; Szulcek, R.; Vonk, A.B.A.; Hordijk, P.L.; Boer, C.; van den Brom, C.E. Microcirculatory perfusion disturbances following cardiac surgery with cardiopulmonary bypass are associated with in vitro endothelial hyperpermeability and increased angiopoietin-2 levels. Crit. Care 2019, 23, 117. [Google Scholar] [CrossRef] [PubMed]
  242. Dekker, N.; van Meurs, M.; van Leeuwen, A.; Hofland, H.; van Slyke, P.; Vonk, A.; Boer, C.; Brom, C.V.D. Vasculotide, an angiopoietin-1 mimetic, reduces pulmonary vascular leakage and preserves microcirculatory perfusion during cardiopulmonary bypass in rats. Br. J. Anaesth. 2018, 121, 1041–1051. [Google Scholar] [CrossRef] [PubMed]
  243. Lee, S.-J.; Lee, C.-K.; Kang, S.; Park, I.; Kim, Y.H.; Kim, S.K.; Hong, S.P.; Bae, H.; He, Y.; Kubota, Y.; et al. Angiopoietin-2 exacerbates cardiac hypoxia and inflammation after myocardial infarction. J. Clin. Investig. 2018, 128, 5018–5033. [Google Scholar] [CrossRef]
  244. Hu, S.; Cao, S.; Liu, J. Role of angiopoietin-2 in the cardioprotective effect of fibroblast growth factor 21 on ischemia/reperfusion-induced injury in H9c2 cardiomyocytes. Exp. Ther. Med. 2017, 14, 771–779. [Google Scholar] [CrossRef]
  245. Aman, J.; van Bezu, J.; Damanafshan, A.; Huveneers, S.; Eringa, E.C.; Vogel, S.M.; Groeneveld, A.J.; Noordegraaf, A.V.; van Hinsbergh, V.W.; Amerongen, G.P.V.N. Effective Treatment of Edema and Endothelial Barrier Dysfunction with Imatinib. Circulation 2012, 126, 2728–2738. [Google Scholar] [CrossRef] [PubMed]
  246. Rizzo, A.N.; Aman, J.; Amerongen, G.P.V.N.; Dudek, S.M. Targeting Abl Kinases to Regulate Vascular Leak during Sepsis and Acute Respiratory Distress Syndrome. Arter. Thromb. Vasc. Biol. 2015, 35, 1071–1079. [Google Scholar] [CrossRef]
  247. Koning, N.; de Lange, F.; van Meurs, M.; Jongman, R.; Ahmed, Y.; Schwarte, L.; Amerongen, G.V.N.; Vonk, A.; Niessen, H.; Baufreton, C.; et al. Reduction of vascular leakage by imatinib is associated with preserved microcirculatory perfusion and reduced renal injury markers in a rat model of cardiopulmonary bypass. Br. J. Anaesth. 2018, 120, 1165–1175. [Google Scholar] [CrossRef]
  248. Konijnenberg, L.S.F.; Luiken, T.T.J.; Veltien, A.; Uthman, L.; Kuster, C.T.A.; Rodwell, L.; de Waard, G.A.; Lindert, M.K.-T.; Akiva, A.; Thijssen, D.H.J.; et al. Imatinib attenuates reperfusion injury in a rat model of acute myocardial infarction. Basic Res. Cardiol. 2023, 118, 2. [Google Scholar] [CrossRef]
  249. Boran, T.; Akyildiz, A.G.; Jannuzzi, A.T.; Alpertunga, B. Extended regorafenib treatment can be linked with mitochondrial damage leading to cardiotoxicity. Toxicol. Lett. 2020, 336, 39–49. [Google Scholar] [CrossRef]
  250. Raivio, P.; Lassila, R.; Petäjä, J. Thrombin in myocardial ischemia-reperfusion during cardiac surgery. Ann. Thorac. Surg. 2009, 88, 318–325. [Google Scholar] [CrossRef]
  251. Dekker, N.A.M.; van Leeuwen, A.L.I.; van Meurs, M.; Moser, J.; Pankras, J.E.; van der Wel, N.N.; Niessen, H.W.; Vervloet, M.G.; Vonk, A.B.A.; Hordijk, P.L.; et al. Preservation of renal endothelial integrity and reduction of renal edema by aprotinin does not preserve renal perfusion and function following experimental cardiopulmonary bypass. Intensiv. Care Med. Exp. 2021, 9, 30. [Google Scholar] [CrossRef] [PubMed]
  252. Khan, T.A.; Bianchi, C.; Araujo, E.; Voisine, P.; Xu, S.-H.; Feng, J.; Li, J.; Sellke, F.W. Aprotinin Preserves Cellular Junctions and Reduces Myocardial Edema After Regional Ischemia and Cardioplegic Arrest. Circulation 2005, 112, I196–I201. [Google Scholar] [CrossRef] [PubMed]
  253. Karaca, P.; Konuralp, C.; Enc, Y.; Süzer, A.; Sokullu, O.; Ayoglu, U.; Cicek, S. Cardioprotective Effect of Aprotinin on Myocardial Ischemia/Reperfusion Injury during Cardiopulmonary Bypass. Circ. J. 2006, 70, 1432–1436. [Google Scholar] [CrossRef]
  254. Bert, C.; De Buck, F.; Sergeant, P.; Van Hemelrijck, J.; Kasran, A.; Van Duppen, V.; Ceuppens, J.; Meyns, B.; Delforge, M.; Wouters, P. Aprotinin Reduces Cardiac Troponin I Release and Inhibits Apoptosis of Polymorphonuclear Cells during Off-Pump Coronary Artery Bypass Surgery. J. Cardiothorac. Vasc. Anesth. 2008, 22, 16–22. [Google Scholar] [CrossRef]
  255. Mangano, D.T.; Tudor, I.C.; Dietzel, C. The risk associated with aprotinin in cardiac surgery. N. Engl. J. Med. 2006, 354, 353–365. [Google Scholar] [CrossRef] [PubMed]
  256. Wang, X.; Li, L.; He, L.; Yao, Y. The effect of tranexamic acid on myocardial injury in cardiac surgical patients: A systematic review and meta-analysis. Blood Coagul. Fibrinolysis 2022, 33, 429–437. [Google Scholar] [CrossRef]
  257. Zhang, H.; Chen, H.; Li, J.; Bian, Y.; Song, Y.; Li, Z.; He, F.; Liu, S.; Tsai, Y. Hirudin protects against isoproternol-induced myocardial infraction by alleviating oxidative via an Nrf2 dependent manner. Int. J. Biol. Macromol. 2020, 162, 425–435. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of mechanisms during the ischemic and reperfusion period that together cause acute cardiac IRI.
Figure 1. Schematic diagram of mechanisms during the ischemic and reperfusion period that together cause acute cardiac IRI.
Cells 12 01432 g001
Figure 2. Schematic overview of Helium-induced cardioprotection: this diagram presents a summary of the established mechanisms involved in helium-induced cardioprotection, primarily through the RISK pathway, which is closely associated with changes in Caveolin-related processes. Helium is represented by a purple circle (He). Red arrows indicate activation or up-regulation, while squares denote suppression or down-regulation. Intracellularly, these mechanisms converge on the mitochondria, inhibiting the opening of the mitochondrial permeability transition pore (mPTP). Additionally, the potential pathway of “remote conditioning” by helium is illustrated on the left side of the diagram. Unknown and identified factors (potentially Caveolin, transported via exosomes) mediate protection in distant organs and enhance mitochondrial respiration in remote cells. GPCR = G-protein coupled receptor; MEK-1 = mitogen-activated protein kinase-extracellular signal-regulated kinase-1; ERK1/2 = Extracellular signal-regulated kinase 1/2; IP3 = inositol triphosphate-3; DAG = diacylglycerol; PKC-ε = protein kinase C epsilon; GSK3β = glycogen synthase kinase-3beta; PI3K = Phosphatidylinositol-3-kinase; PDK-1 = phosphoinositide-dependent protein kinase-1; PKB = protein kinase B; mTOR = mammalian target of rapamycin; P53 = Tumor protein P53; mPTP mitochondrial permeability transition pore; eNOS = endothelial nitric oxide synthase; NO = nitric oxide; L-NAME = L-NG-nitroarginine methyl ester; PKA = protein kinase A; mKCa = mitochondrial calcium-sensitive potassium channel; ROS = reactive oxygen species; Pi = inorganic phosphate; ATP = adenosine triphosphate; and ADP = adenosine diphosphate.
Figure 2. Schematic overview of Helium-induced cardioprotection: this diagram presents a summary of the established mechanisms involved in helium-induced cardioprotection, primarily through the RISK pathway, which is closely associated with changes in Caveolin-related processes. Helium is represented by a purple circle (He). Red arrows indicate activation or up-regulation, while squares denote suppression or down-regulation. Intracellularly, these mechanisms converge on the mitochondria, inhibiting the opening of the mitochondrial permeability transition pore (mPTP). Additionally, the potential pathway of “remote conditioning” by helium is illustrated on the left side of the diagram. Unknown and identified factors (potentially Caveolin, transported via exosomes) mediate protection in distant organs and enhance mitochondrial respiration in remote cells. GPCR = G-protein coupled receptor; MEK-1 = mitogen-activated protein kinase-extracellular signal-regulated kinase-1; ERK1/2 = Extracellular signal-regulated kinase 1/2; IP3 = inositol triphosphate-3; DAG = diacylglycerol; PKC-ε = protein kinase C epsilon; GSK3β = glycogen synthase kinase-3beta; PI3K = Phosphatidylinositol-3-kinase; PDK-1 = phosphoinositide-dependent protein kinase-1; PKB = protein kinase B; mTOR = mammalian target of rapamycin; P53 = Tumor protein P53; mPTP mitochondrial permeability transition pore; eNOS = endothelial nitric oxide synthase; NO = nitric oxide; L-NAME = L-NG-nitroarginine methyl ester; PKA = protein kinase A; mKCa = mitochondrial calcium-sensitive potassium channel; ROS = reactive oxygen species; Pi = inorganic phosphate; ATP = adenosine triphosphate; and ADP = adenosine diphosphate.
Cells 12 01432 g002
Figure 3. Overview of possible new therapeutics targeting the endothelium and thereby protecting the cardiomyocyte. Crosstalk between cardiomyocytes and endothelial cells is crucial for cardiac function. Endothelial cells constitute the inner lining of arteries, veins, and capillaries and form a barrier between vessels and heart. During cardiopulmonary bypass, the endothelium can be activated, leading to increased permeability of the endothelium (see top part of the schematic vessel), fluid accumulation, and tissue edema, thereby hampering oxygen exchange. Within the circulation, several proteins are up- or down-regulated during cardiopulmonary bypass. These circulating proteins can interfere with their receptors on the endothelium and also (in)directly affect the cardiomyocyte. They are, therefore, targets of interest to therapeutically protect the heart against IRI. VEGF-A = vascular endothelial growth factor A; VEGF-R2 = VEGF-receptor 2; S1PR1 = sphingosine-1-phosphate receptor 1; Tie2 = tyrosine kinase receptor; PAR1 = protease-activated receptor 1; WPB= Weibel–Palade bodies; Ca = calcium.
Figure 3. Overview of possible new therapeutics targeting the endothelium and thereby protecting the cardiomyocyte. Crosstalk between cardiomyocytes and endothelial cells is crucial for cardiac function. Endothelial cells constitute the inner lining of arteries, veins, and capillaries and form a barrier between vessels and heart. During cardiopulmonary bypass, the endothelium can be activated, leading to increased permeability of the endothelium (see top part of the schematic vessel), fluid accumulation, and tissue edema, thereby hampering oxygen exchange. Within the circulation, several proteins are up- or down-regulated during cardiopulmonary bypass. These circulating proteins can interfere with their receptors on the endothelium and also (in)directly affect the cardiomyocyte. They are, therefore, targets of interest to therapeutically protect the heart against IRI. VEGF-A = vascular endothelial growth factor A; VEGF-R2 = VEGF-receptor 2; S1PR1 = sphingosine-1-phosphate receptor 1; Tie2 = tyrosine kinase receptor; PAR1 = protease-activated receptor 1; WPB= Weibel–Palade bodies; Ca = calcium.
Cells 12 01432 g003
Table 1. Promising drugs/compounds to reduce cardiac IRI and endothelium damage.
Table 1. Promising drugs/compounds to reduce cardiac IRI and endothelium damage.
Drugs/CompoundsMechanismSetting= > Phase II Trial?
Heart
NAD+ precursors
(NR, NMN)
restoring NAD levelspreclinical models of acute cardiac IRINo
Malonatereducing ROS at early reperfusionpreclinical models of acute cardiac IRINo
NLRP3 inflammasome inhibitorspreventing inflammasome complex formationpreclinical models of cardiac IRINo
Caspase-1,4 inhibitors (VRT, emricasan)pyroptosis inhibitionpreclinical models of acute cardiac IRINo
Calpain inhibitorsproteolysis and pyroptosis inhibitionpreclinical models of cardiac IRINo
Volatile anesthetics/
Helium/Xenon
activation of Survival pathways (RISK/SAFE), Caveolin 1/3, mitophagy, and autophagypreclinical models of cardiac IRIYes
Volatile anestheticsreduced microRNAs (miRNA-155), induction of VEGFR1preclinical models of cardiac IRINo
Endothelium
Sevofluraneactivation of survival pathways (RISK/SAFE)preclinical CPB models
cardiac surgery patients with CPB
Yes
Nitric oxide donorsreducing oxidative stresspreclinical CPB models
cardiac surgery patients with CPB
Yes
Lidoflazinecalcium channel blockerpatients undergoing multiple aorta-coronary bypass graftingYes
Doxycyclineinhibition matrix metalloproteinasescardiac surgery patients with CPBYes
(Indirect) Endothelial receptor activatorsstrengthening endothelial barrierpreclinical CPB models
cardiac surgery patients with CPB
Yes, but not all
NAD = nicotinamide adenine dinucleotide; NR = nicotinamide riboside; NMN = nicotinamide mononucleotide; IRI = ischemia-reperfusion injury; NLRP3 = NOD-like receptor family pyrin domain containing 3; ROS = reactive oxygen species; VRT = active component of the caspase-1/4 inhibitor VX-765; RISK = reperfusion injury salvage kinase; SAFE = survivor activating factor enhancement; VEGFR1 = vascular endothelial growth factor-receptor 1; and CPB = cardiopulmonary bypass.
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

Wang, Q.; Zuurbier, C.J.; Huhn, R.; Torregroza, C.; Hollmann, M.W.; Preckel, B.; van den Brom, C.E.; Weber, N.C. Pharmacological Cardioprotection against Ischemia Reperfusion Injury—The Search for a Clinical Effective Therapy. Cells 2023, 12, 1432. https://doi.org/10.3390/cells12101432

AMA Style

Wang Q, Zuurbier CJ, Huhn R, Torregroza C, Hollmann MW, Preckel B, van den Brom CE, Weber NC. Pharmacological Cardioprotection against Ischemia Reperfusion Injury—The Search for a Clinical Effective Therapy. Cells. 2023; 12(10):1432. https://doi.org/10.3390/cells12101432

Chicago/Turabian Style

Wang, Qian, Coert J. Zuurbier, Ragnar Huhn, Carolin Torregroza, Markus W. Hollmann, Benedikt Preckel, Charissa E. van den Brom, and Nina C. Weber. 2023. "Pharmacological Cardioprotection against Ischemia Reperfusion Injury—The Search for a Clinical Effective Therapy" Cells 12, no. 10: 1432. https://doi.org/10.3390/cells12101432

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