Next Article in Journal
The Role of G Protein-Coupled Estrogen Receptor (GPER) in Vascular Pathology and Physiology
Next Article in Special Issue
Novel Biomarkers and Advanced Cardiac Imaging in Aortic Stenosis: Old and New
Previous Article in Journal
Multi-Omics Analyses Reveal Mitochondrial Dysfunction Contributing to Temozolomide Resistance in Glioblastoma Cells
Previous Article in Special Issue
PFKFB3 Inhibitor 3PO Reduces Cardiac Remodeling after Myocardial Infarction by Regulating the TGF-β1/SMAD2/3 Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial Calcium Overload Plays a Causal Role in Oxidative Stress in the Failing Heart

1
Department of Physiology and Cellular Biophysics, Clyde and Helen Wu Center for Molecular Cardiology, Columbia University Vagelos College of Physicians & Surgeons, New York, NY 10032, USA
2
Department of Medicine, Division of Cardiology, Wilf Family Cardiovascular Research Institute, Albert Einstein College of Medicine, New York, NY 10461, USA
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(9), 1409; https://doi.org/10.3390/biom13091409
Submission received: 17 August 2023 / Revised: 13 September 2023 / Accepted: 17 September 2023 / Published: 19 September 2023

Abstract

:
Heart failure is a serious global health challenge, affecting more than 6.2 million people in the United States and is projected to reach over 8 million by 2030. Independent of etiology, failing hearts share common features, including defective calcium (Ca2+) handling, mitochondrial Ca2+ overload, and oxidative stress. In cardiomyocytes, Ca2+ not only regulates excitation–contraction coupling, but also mitochondrial metabolism and oxidative stress signaling, thereby controlling the function and actual destiny of the cell. Understanding the mechanisms of mitochondrial Ca2+ uptake and the molecular pathways involved in the regulation of increased mitochondrial Ca2+ influx is an ongoing challenge in order to identify novel therapeutic targets to alleviate the burden of heart failure. In this review, we discuss the mechanisms underlying altered mitochondrial Ca2+ handling in heart failure and the potential therapeutic strategies.

1. Introduction

Heart failure (HF) is a complex clinical syndrome characterized by the inability of the heart to pump blood at a rate and with adequate force to meet the metabolic demands of the body. This condition can result from various underlying causes, including damage to the heart muscle (such as from a heart attack), hypertension, valvular heart disease, or other cardiac conditions. HF can affect either the left, right, or both ventricles of the heart and can manifest as systolic dysfunction (impaired ability of the heart to contract and eject blood) or diastolic dysfunction (impaired ability of the heart to relax and fill with blood). The hallmark symptoms of HF include shortness of breath, fatigue, fluid retention (edema), and reduced exercise tolerance. It is a chronic condition that requires ongoing management and treatment to improve the patient’s quality of life and reduce morbidity and mortality. HF shows a steadily growing prevalence and remains the leading cause of death in developed countries [1,2]. Although initially adaptive [3,4,5], cardiac responses to pressure or volume overload are associated with deep molecular changes, eventually leading to fetal gene expression [6,7,8,9], impaired contractile function [10,11], abnormal vascularization [12,13], altered extracellular matrix composition [14,15,16,17], fibrosis [18,19], and profound metabolic abnormalities [20,21,22,23,24,25,26,27], all of which unavoidably affect myocardial contraction and eventually progress towards overt HF [28].
The heart consumes a significant amount of energy supplied almost entirely by mitochondrial ATP production derived from oxidative metabolism [29]. Nearly a third of each cardiomyocyte, by volume, is taken up by mitochondria [30], which are located in close proximity to the sarcoplasmic reticulum (SR), allowing the two organelles to exchange metabolites and interconnect [31,32,33,34]. Calcium (Ca2+) and reactive oxygen species (ROS) are generally considered the main transduction signals linking the SR and mitochondrion and permit them to adapt their responses to external stress in a highly regulated manner [35,36,37,38,39,40]. Independent of etiology, a key hallmark of HF is Ca2+ dyshomeostasis, altered bioenergetics, changes in mitochondrial respiration, increased oxidative stress, and impaired energy production and consumption [41,42,43,44]. During HF, impaired Ca2+ uptake due to dysregulation in several calcium cycling proteins and increased SR Ca2+ leak through type-2 ryanodine receptors (RyR2) resulting in reduced intracellular Ca2+ stores required for cardiomyocyte contraction and increased baseline cytosolic Ca2+ [45,46,47,48,49,50]. Increased mitochondrial Ca2+ uptake follows, which contributes to mitochondrial Ca2+ overload, dysfunction, and oxidative stress [51]. Mitochondrial dysfunction increases mitochondrial ROS production, which has secondary effects on a broad range of cellular functions and accelerates HF progression [52,53,54,55,56].
In this review, we summarize the recent evidence demonstrating the crucial role of mitochondrial Ca2+ overload in inducing oxidative stress during HF and highlight the potential therapeutic targets as well as new outlooks for future studies.

2. Mitochondrial Function in Cardiomyocytes

The heart primarily depends on mitochondrial oxidative metabolism to fulfill approximately 95% of its energy demands [29], with a particular reliance on ATP usage during the process of excitation–contraction (EC) coupling. Within the mitochondria, oxidative metabolism is chiefly propelled by redox reactions occurring in the electron transport chain (ETC). The reduced forms of nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide (FADH2) serve as electron donors, and these electrons traverse a series of increasingly reduced electron carriers, ultimately reaching oxygen, which serves as the final electron acceptor [57,58]. The energy generated from this exergonic flow of electrons is utilized to move protons from the mitochondrial matrix to the intermembrane space, thus establishing a chemical and electrical gradient across the inner mitochondrial membranes [59,60,61,62]. This gradient, impermeable to ions, represents the proton motive force, which is then harnessed by ATP synthase to convert ADP into ATP through oxidative phosphorylation [63]. A continuous flow of electrons along the ETC is sustained by the dehydrogenases found in the tricarboxylic acid cycle, also known as the Krebs cycle, which consistently supplies NADH and FADH2 [64]. In the intermembrane space, mitochondrial creatine kinase (CK) transforms creatine transported from the cytosol into phosphocreatine, serving as a spatial and temporal reservoir of high-energy phosphates [28,65].
Increased ATP consumption in cardiomyocytes requires additional reducing equivalents, and Ca2+ ions are essential to this process. Indeed, increased cytosolic Ca2+ levels with elevated cardiac contractions allow Ca2+ to enter the mitochondria and stimulate pyruvate dehydrogenase, α-ketoglutarate, and isocitrate dehydrogenase to regenerate reducing equivalents [66], as well as signal ATP synthase and complex III of the ETC to accelerate ATP production [67]. However, persistently high levels of Ca2+ entry into the mitochondria, as in HF, may be deleterious, leading to mitochondrial Ca2+ overload, increased ROS production, and oxidative stress [68]. The fundamental molecular mechanisms leading to mitochondrial Ca2+ overload during HF are discussed in the following sections.

3. Ca2+ Homeostasis Dysregulation and Mitochondrial Ca2+ Overload in HF

3.1. Cardiac Ca2+ Handling

Cardiac muscle contraction is regulated beat-to-beat by Ca2+ stored and released within cardiomyocytes in EC coupling. During the depolarizing phase of the cardiac action potential, Ca2+ enters the cardiomyocyte through voltage-activated L-type Ca2+ channels and triggers RyR2 to open and release Ca2+ from SR Ca2+ stores through calcium-induced calcium release (CICR), raising cytosolic Ca2+ levels about ten-fold to ~1 μM [46]. Ca2+ then activates specific proteins of the contractile apparatus and induces contraction of cardiac cells [49,69,70]. Subsequent relaxation occurs as Ca2+ is removed from the cytosol via three primary mechanisms: type-2a SR Ca2+-ATPase (SERCA2a) transport, sarcolemmal Na+/Ca2+ exchanger (NCX) extrusion, and mitochondrial Ca2+ uptake [71,72,73,74,75].
During HF, the widespread dysregulation of Ca2+ cycling leading to significant increases in cytosolic baseline Ca2+ has been observed [76]. First, the increased activation and open probabilities of voltage-gated L-type Ca2+ channels in failing compared with non-failing human ventricles was confirmed, which resulted in an excessive calcium influx into the cytosol [77,78]. Moreover, the hyperactivity of type-2 ryanodine receptors (RyR2) due to increased CICR and post-translational modification, which culminated in a significant SR Ca2+ leak, was also shown in human and mice cardiomyocytes of failing hearts [76]. Furthermore, a reduced effectiveness of the sarcolemmal Na+/Ca2+ exchanger (NCX) in extruding cytosolic calcium was demonstrated in a mouse model of heart failure [79]. Finally, reduced SERCA2a expression and activity was observed in several animal models of heart failure and was shown to decrease SR calcium uptake from the cytosol [80,81]. These defects result in increased cytosolic Ca2+ and potentially increase Ca2+ entry into the mitochondria, causing a detrimental mitochondrial Ca2+ overload in cardiomyocytes [82,83].

3.2. Inflammation-Mediated Ca2+ Dyshomeostasis in HF

Inflammation has emerged as a major pathophysiological feature of HF progression, regardless of etiology. Numerous pro-inflammatory signaling molecules are produced and released in HF, initiating vicious cascades involving oxidative stress, mitochondrial dysfunction, myocardial weakened contractility, and, importantly, Ca2+ dyshomeostasis [84,85]. Elevated levels pro-inflammatory cytokines, including tumor necrosis factor (TNF-⍺), interleukin-17 (IL-17), interleukin-6 (IL-6), interleukin-1 (IL-1ß), and transforming growth factor (TGF-ß), found in HF were all associated with the dysregulation of several important Ca2+ cycling proteins [84,85]. TNF-⍺ was found to be significantly elevated in the plasma of heart failure patients, serving as an effective predictor of HF severity and diastolic dysfunction [86,87]. Several studies have shown that TNF-⍺ suppresses SERCA2a gene expression, worsening cytosolic Ca2+ overload [86,87]. Increased levels of IL-17, IL-6, and IL-1ß in HF were all associated with left ventricular dysfunction, causing further SERCA2a gene downregulation [88,89]. Finally, TGF-ß was demonstrated to be significantly upregulated in HF, associated with cardiac injury, cardiomyocyte apoptosis, hypertrophy, and fibrosis [84,85]. Waning et al. report that this increased TGF-ß activity can lead to NADPH Oxidase 2 (NOX2) localization to RyR2, causing channel oxidation due to superoxide radicals (O2·) and resulting in an SR Ca2+ leak [90]. The elevated release of these pro-inflammatory cytokines in HF would increase cytosolic baseline Ca2+ and mitochondrial Ca2+ overload, leading to ROS formation and oxidative stress. The signaling cascades associating the inflammatory response to SR/mitochondrial dysfunction in HF still remain to be elucidated.

3.3. Mitochondrial Ca2+ Handling

3.3.1. Role of the Mitochondrial Ca2+ Uniporter

Mitochondrial Ca2+ influx occurs primarily via the mitochondrial Ca2+ uniporter (MCU), whereas Ca2+ efflux occurs mainly via the mitochondrial Na+-Ca2+—Li+ exchanger (mNCLX) [91,92,93,94]. MCU, previously known as the coiled-coil domain containing 109A (CCD109A) and C10 or f42, is a ruthenium-sensitive Ca2+ channel in the mitochondrial inner membrane that facilitates Ca2+ transport down its electrochemical gradient without coupling Ca2+ transport with the movement of other ions [95,96,97,98,99,100,101,102,103,104,105,106,107,108,109]. The mitochondrial Ca2+ uptake channel is a large holocomplex consisting of MCU as the pore-forming unit and multiple regulatory proteins, including the essential MCU regulator (EMRE) [110,111,112,113], and EF-hand proteins MICU1 [114,115,116,117,118,119,120,121,122,123,124], MICU2 [125,126,127,128], and MICU3 [127,128,129,130,131,132]. The study of the cardiac phenotype of murine models lacking MCU yielded controversial results. First and foremost, the germline mutation of MCU was embryonically lethal in the C57BL/6 background but not in the CD background. The latter mice showed modest differences in the myocardial contractile response to isoproterenol and did not seem to be protected from cardiac ischemia–reperfusion injury [103]. These results were further confirmed in a cardiac-specific dominant negative MCU mouse [133]. However, the cardiac-specific inducible MCU mouse exhibited a significant difference in the contractile response to adrenergic stimulation [134,135]; however, that finding was not reproduced when the hearts were studied in an isolated system [136].
During HF, increased cytosolic Ca2+ leads to excessive Ca2+ entry into the mitochondria and causes mitochondrial Ca2+ overload, which is detrimental to the cardiomyocyte [82,83,137]. Considering that excessive mitochondrial Ca2+ uptake causes mitochondrial Ca2+ overload and oxidative stress, MCU modulators can be an interesting therapeutic option to rescue the subsequent detrimental signaling cascades observed in HF [122,136,138,139,140,141,142,143,144,145,146]. Several studies have reported that the inhibition of mitochondrial Ca2+ uptake during acute stress can be reduced by the deletion of the MCU gene or by the MCU blocker, Ru360 [147,148], while basal mitochondrial Ca2+ levels remain unchanged [134,135]. Furthermore, MCU deletion was shown to not completely inhibit mitochondrial Ca2+ uptake [149]. These findings suggest the existence of other mechanisms of mitochondrial Ca2+ uptake with different levels of sensitivity to cytosolic Ca2+. Possible mediators of MCU-independent mitochondrial Ca2+ uptake include the Ca2+/H+ exchanger Leucine Zipper And EF-Hand Containing Transmembrane Protein 1 (Letm1) [150,151,152,153,154,155,156] and the Transient Receptor Potential Canonical 3 (TRPC3) [130,157,158,159,160,161,162,163]. These aspects are not surprising since MCU Ca2+ sensitivity has been shown [164] to be quite low (Kd~20–30 μMol) and, as such, resting cytosolic Ca2+ can be insufficient to trigger MCU activity [120,139,165,166]. On the other hand, it is important to define what the basal cytosolic Ca2+ concentration in cardiomyocytes is, as well as the difference between basal cytosolic Ca2+ in normal versus failing cardiomyocytes when associated with a defective RyR2 leak or reduced SERCA2a activity. Nonetheless, is basal cytosolic Ca2+ in HF sufficient to enter mitochondria and cause Ca2+ overload? This dilemma represents an important question that needs to be clarified by future investigations.

3.3.2. Role of the mNCLX

Recently, Luongo and collaborators revealed a novel mechanism, independent of MCU, responsible for mitochondrial Ca2+ homeostasis [167]. Since mitochondrial Ca2+ uptake has to be balanced with extrusion in homeostatic conditions, they hypothesized that the key channel responsible for mitochondrial efflux Ca2+, mNCLX [94,167], would play a crucial role in mitochondrial Ca2+ handling [167]. Using a murine model harboring a tamoxifen-induced deletion of mNCLX, they demonstrated that these mice developed left ventricular dilation and decreased cardiac function. This was mirrored by mitochondrial remodeling and dysfunction followed by death within 2 weeks of tamoxifen delivery. On the other hand, mice overexpressing mitochondrial mNCLX, which enhanced mitochondrial Ca2+ efflux by 38%, displayed an attenuated infarct size after MI or I/R injury, improved cardiac function, and reduced superoxide generation [167]. Despite the promising outcomes of mNCLX overexpression, whether mitochondria Ca2+ overload in HF is due to enhanced Ca2+ uptake or repressed extrusion has yet to be further elucidated. Furthermore, the following queries need to be investigated: does increased mNCLX expression/activity affect cytosolic Ca2+ levels by increasing mitochondrial Ca2+ extrusion? Moreover, what are the long-term outcomes of enhanced mNCLX activity in mice? Will those mice be more vulnerable to arrhythmias due to cytosolic Ca2+ overload? We believe that the long-term outcomes of this enhanced mNCLX activity are deteriorated when cardiac disease is associated with leaky RyR2 channels. Although, these hypotheses remain to be verified, recent studies by Garbincius point towards the deleterious effect of mNCLX cardiomyocyte-specific overexpression in mice. Indeed, survival was reduced when mNCLX cardiomyocyte-specific overexpressing mice were subjected to severe neurohormonal stress with angiotensin II and phenylephrine [168].

3.3.3. Factors influencing Mitochondrial Ca2+ Uptake

It is well established that MCU does not transport Ca2+ when the Ca2+ concentration is below a threshold of 200 nM [169]. However, such a low affinity might be overcome when MCU is juxtaposed with the SR Ca2+ release channels RyR2 and/or 1,4,5-trisphosphate receptors (IP3Rs) [51,170,171]. Remarkably, hyper-physiological Ca2+ concentrations (10–100 µM) are required to activate Ca2+ uptake into isolated mitochondria; however, in intact cardiomyocytes, the cytosolic Ca2+ concentration after Ca2+ release from the SR remains less than 10 µM, yet Ca2+ is still internalized by mitochondria [172,173].
Such a discrepancy between isolated mitochondria and intact cells was partially resolved by the finding of high cytosolic Ca2+ concentrations in microdomains between mitochondria and the SR. For example, RyR2 Ca2+ release during systole creates microdomains of high, localized Ca2+ concentrations (~30 µM) in the vicinity of the mitochondrial membrane, leading to mitochondrial Ca2+ uptake [174,175,176,177,178,179,180,181,182,183]. This is an exciting theory; however, it is missing important pieces of the SR–mitochondria Ca2+ transfer puzzle that mostly concern the complexity of this highly regulated process, which involves two dynamic organelles. Due, in part, to the low Ca2+ affinity of the MCU, we believe that the formation of Ca2+-rich microdomains alone is not sufficient to allow Ca2+ to enter the mitochondria, especially in the presence of competing sarcolemmal NCX and SERCA2a pumps with a higher Ca2+ affinity and faster uptake rate.
Thus, a very close proximity between the two organelles seems to be necessary to allow Ca2+ to transfer to the mitochondria. This hypothesis was recently investigated and two major, non-mutually exclusive interactions between the SR and mitochondria were identified. The functional, tight coupling between the SR and mitochondria is attributed to the inter-organelle tether protein involving mitofusin2 (MFN2) [184,185,186] and/or the IP3R-Grp75-VDAC protein complex [33,143,187,188]. The relevance of each theory to the mitochondrial Ca2+ overload observed in HF is discussed below.

4. Physical SR–Mitochondria Interaction

4.1. Mitofusin-Mediated Tethering

Recently, the purification of subcellular fractions corresponding to endoplasmic reticulum (ER)–mitochondria contacts, referred to as mitochondria-associated membranes (MAMs), led to the identification of proteins enriched in those membrane domains [189,190,191]. Among them, MFN2 is located on both the ER and mitochondrial membranes [192,193], where it forms homo- and heterotypic interactions with mitofusin1 (MFN1) [194]. In addition, a relevant fraction of MFN2 (7–10%) exists in the ER and its ablation alters the structure of this organelle, causing its fragmentation [195,196]. The selective reconstitution of the ER pool of MFN2 in the same cells completely restores the reticular nature of the ER [184], suggesting an essential role of MFN2 in the modulation of ER shape.
MFN2-KO cells were used to investigate how tethering between the ER and mitochondria would impact Ca2+ transfer to the mitochondria. The lack of MFN2 to tether the ER to mitochondria resulted in a significant reduction in mitochondrial Ca2+ uptake, without any impairment of mitochondrial capacity to uptake Ca2+. This impaired uptake reflects an increased distance between the two organelles and, therefore, a limited generation of Ca2+ microdomains. Therefore, MFN2-KO cells provide experimental proof for the importance of Ca2+ microdomains between the ER/SR and mitochondria, a theory postulated by Rizzuto and Pozzan almost 30 years ago [175,197,198]. Thus, the physical juxtaposition of the SR to mitochondria by MFN2 appears to be essential for normal inter-organelle Ca2+ signaling in the myocardium. This is consistent with a requirement for SR–mitochondrial Ca2+ signaling through microdomains to maintain a competent bioenergetic feedback response in cardiomyocytes [192].
Interestingly, quantitative analyses have shown an increase, not a decrease, in the close contacts between the ER and mitochondria in MFN2-KO mouse embryonic fibroblasts, as compared with WT cells [199,200]. As previously described, the tethering of the SR to mitochondria in cardiac cells would be expected to facilitate the exchange of Ca2+ signals through inter-organelle microdomains, where the signaling factor is prevented from diffusing into the cytosol. However, as mitochondria are the main source of ATP in cardiomyocytes and the SR is the main source of Ca2+, it is reasonable to postulate the existence of a physical, dynamic interaction that facilitates direct Ca2+ exchange between the two organelles. Increased beat-to-beat SR Ca2+ release can then be rapidly sensed by the mitochondria, inducing mitochondrial Ca2+ uptake, stimulating mitochondrial dehydrogenases, increasing the production of NADH for the ETC, and increasing ATP production according to the physiological demand (see Figure 1). Furthermore, mitochondrial ATP would then be available to the SR to drive energy-intensive and phasic Ca2+ cycling. Nevertheless, the relevance of this finding for cardiomyocytes experiencing HF continues to be debated.

4.2. IP3R-Grp75-VDAC Complex and Mitochondrial Ca2+ Uptake

In the same vein as the close proximity between the SR and mitochondria described above, recent studies have identified several proteins enriched at the SR–mitochondria interface, highlighting the emerging role of inter-organelle interactions within the cell [184,201]. One important protein complex composed of the voltage-dependent ion channel (VDAC), chaperone glucose-regulated protein 75 (Grp75), and IP3R1 has been identified as a key regulator of direct Ca2+ transfer from the ER/SR to the mitochondria [33]. Specifically, VDAC1 was shown to be physically linked to IP3R1 through the chaperone protein Grp75 [192]. The functional interaction between these channels and the mitochondrion was verified by the recombinant expression of only the ligand-binding domain (LBD) of IP3R1 on the ER. Without any IP3R1 Ca2+ channel functionality, the LBD was able to directly enhance Ca2+ accumulation in the mitochondria through the outer mitochondrial membrane. Knocking down Grp75 abolished this stimulatory effect, highlighting the importance of chaperone-mediated conformational coupling between IP3R and the mitochondrial Ca2+ uptake machinery [33]. This finding was recently confirmed by other methods, including an in situ proximity ligation assay [202]. However, most studies, including ours, indicated that the activation of IP3Rs in cardiomyocytes, unlike RyR2, did not significantly contribute to EC coupling [51,203]. Furthermore, the relevance of IP3R to mitochondrial Ca2+ overload in HF remains mostly unclear. The removal of IP3R2 channels from cardiomyocytes did not have any impact on the progression of heart failure or on factors, such as Ca2+ sparks, SR Ca2+ load, mitochondrial Ca2+ levels, or the production of reactive oxygen species (ROS). This suggests that these channels do not play a role in the malfunctioning of Ca2+ regulation in ventricular cardiomyocytes [51]. Interestingly, when IP3R1 was eliminated, it led to a reduction in vasoconstriction in the coronary arteries, which was mediated by the activation of vascular smooth muscle cells, and consequently, it slowed down the advancement of HF [202]. On another note, our recent research demonstrated a strong association between mitochondrial Ca2+ overload in HF and the leakage of Ca2+ from RyR2 channels [51]. This discovery was further validated using mouse models with persistently leaky RyR2 channels, which displayed changes in the mitochondrial structure, a decrease in ATP content, and an increase in the production of mitochondrial ROS [51].
Despite the important progress achieved in the investigation of the pathophysiologic implications of mitochondrial Ca2+ overload, the precise quantity of Ca2+ taken up by mitochondria with each heartbeat in both normal and failing hearts remains uncertain. Furthermore, there is an ongoing debate regarding whether the mitochondrion can function as a primary reservoir for Ca2+ ions [204,205,206,207]. However, numerous studies have used genetic methods to demonstrate substantial mitochondrial dysfunction when Ca2+ homeostasis is manipulated [110,111]. Indeed, an excess of cytosolic Ca2+ in various animal models has led to a mitochondrial phenotype resembling that observed in heart failure. This phenotype includes compromised ATP generation, heightened mitochondrial oxidative stress, and, ultimately, the death of cardiomyocytes [208,209].
Other proteins facilitating a physical connection between the ER and mitochondria have been proposed, including PDZ-domain-containing 8 (PDZD8) [210,211], protrudin [212], Rab7 [213], FUN14-domain-containing 1 (FUNDC1), and protein tyrosine phosphatase-interacting protein 51 (PTPIP51) [214]. Christopher Miller’s lab demonstrated the interaction between vesicle-associated membrane protein-associated protein B (VAPB), expressed on the ER, and PTPIP51, located on the outer mitochondrial membrane [215]. The disruption of this interaction was shown to profoundly alter inter-organelle Ca2+ fluxes between the ER and mitochondria [216]; however, the exact role of RyR Ca2+ leak was not investigated in these models.

5. Mitochondrial Ca2+ Overload and Mitochondrial ROS Production

Numerous studies have shown that excessive mitochondrial Ca2+ levels lead to the production of mitochondrial reactive oxygen species in various diseases, including heart failure [51,68]. ROS are chemically reactive species based on oxygen (O2) that can rapidly interact with cellular components, such as lipids [217], proteins [218], and nucleic acids [219]. This category encompasses highly reactive molecules, including free radicals, such as superoxide (O2·) and hydroxyl radical (·OH), as well as non-radical species, such as hydrogen peroxide (H2O2) and hypochlorite (OCl) [220,221]. In mitochondria, the presence of Ca2+ stimulates the Krebs cycle, resulting in the generation of NADH and FADH2, both of which provide electrons to the electron transport chain (ETC). As electrons progress through a series of redox reactions along the ETC, a proton gradient is established, facilitating ATP production, as described earlier [222]. Depending on the cellular respiratory state, as much as 20% of the electrons within the ETC may experience a leakage to molecular oxygen to form superoxide anions [223,224,225]. In the mitochondria, much of the superoxide anions are dismutated to hydrogen peroxide (H2O2) by superoxide dismutase (SOD), with SOD2 being the primary mitochondrial isoform [226,227]. Mitochondrial H2O2 is then removed by the antioxidant systems of peroxiredoxin (Prx) and glutathione peroxidase (Gpx) in the mitochondria [228,229]. In HF, increased O2· generation from complex I of the ETC caused excessive formation of H2O2 and ·OH, which became deleterious when their production overwhelmed detoxification [230,231]. Since several ion channels are regulated by redox reactions, oxidative stress can further impact EC coupling either by directly targeting ion channels involved in EC coupling or by activating stress kinases, which may contribute to contractile dysfunction and the progression of HF [49,232]. In the following section, we focus on the key proteins involved in EC coupling that are targeted by mitochondrial ROS production and causative of increased mitochondrial Ca2+ overload. This overload further amplifies mitochondrial oxidative stress, propagating a vicious cycle.

6. Redox-Regulated Proteins of the Ca2+ Handling Machinery and HF

6.1. Ryanodine-Receptor Ca2+ Release Channel

Increased mitochondrial oxidative stress during HF, in human patients as well as in animal models [233,234], is commonly associated with an SR Ca2+ leak through the hyperactivity of RyR2 [46]. As HF progresses, oxidative stress worsens due to the increasing energy demand and workload of the failing heart [229]. Increased RyR2 oxidation and Ca2+ leak not only augments cytosolic Ca2+ concentrations by depleting Ca2+ stores but also increases mitochondria Ca2+ overload, which further produces ROS [51] and worsens the prognosis of HF. A deleterious cycle is therefore perpetuated.
Many cysteine residues in RyR can be subjected to oxidative post-translational modification; however, the exact nature of these modifications for available cysteines and how they affect the activity of the channel remains to be determined. Recent work suggests that, in the presence of H2O2, RyR undergoes covalent disulfide cross-linking between the N-terminal domains of the four neighboring protomers, adopting an open state conformation and generating an SR Ca2+ leak [235]. We and others have published extensive studies showing that the oxidation of the channel dissociates the stabilizing subunit, calstabin2, which results in increased activity and Ca2+ leak through the channel [50,51,236,237]. As such, the oxidation of RyR2 in HF is far from a simple signaling pathway and needs extensive research. Identifying the oxidized residues following HF, as well as the responsible ROS species, would enhance our understanding of RyR2 oxidation-mediated HF. Recent advances in resolving the structure of RyR2 in HF at a high resolution [45,238,239,240,241] using single-particle cryogenic-electron microscopy (cryo-EM) would be the perfect tool for solving the complete structure of RyR at an atomic resolution in order to clarify the architecture of multiple functional states, ligand-binding sites, and gating mechanisms, including RyR2 abnormalities, during HF. In the following section, we provide structural insights and future directions for the investigation of the redox-sensitive residues within RyR2 channels.

Potentially Oxidized Residues in RyR2

RyR channels comprise several surface-exposed cysteine residues with free thiol groups prone to oxidation by reactive oxygen and nitrogen species. Oxidized cysteines that render RyR2 channels leaky to Ca2+ have been implicated as a key factor for the development of age-related cardiac disorders, HF, and atrial fibrillation [157]. Assessing the cellular redox state of these cysteine residues is experimentally challenging because free thiol groups exhibit a high sensitivity to the redox potential of the experimental environment and may be altered accordingly (they can be oxidized or reduced during tissue isolation, preparation, storage, experiments, etc.). This poses an experimental limitation in defining a specific target or mechanism of action for RyR oxidation. Using a maleimide derivative to mask and label reactive cysteines, Voss et al. identified seven hyper-reactive cysteines (rabbit: C1040, C1303, C2436, C2565, C2606, C2611, and C3635) in RyR1, the skeletal muscle isoform [242]. On the other hand, Aracena-Parks et al. identified 12 hyper-active cysteines [243] in RyR1 (rabbit: C36, C253, C315, C811, C906, C1040, C1303, C1591, C2326, C3193, and C3635), three of which overlapped with those identified by Voss et al. (i.e., C1040, C1303, and C3635). These hyper-active cysteines are evolutionarily conserved in RyR1 across mammalian species and 9 of those are also found to be conserved in the RyR2 isoform (human: C36, C822, C917, C1582, C2402, C2572, C2577, C3158, and C3602). Although the functional significance of these oxygen reactive cysteines in RyR1 remains largely unknown, both studies identify cysteine RyR1-C3635/RyR2-C3602 as the most hyper-reactive residue. This particularly important residue is located in the calmodulin-binding domain (see Figure 2). These studies raised an important question regarding the functional significance of the corresponding RyR2 cysteine residue that is evolutionarily conserved in humans, C3602. Subsequently, Mi et al. found that the mutation of cysteine to alanine (C3602A) suppressed the store-overload-induced Ca2+ release (SOICR) by raising the activation threshold and delaying the termination of Ca2+ release. Furthermore, the C3602A mutation markedly increased fractional Ca2+ release and diminished the inhibitory effect of N-ethylmaleimide on Ca2+ release, but had no effect on the stimulatory action of 4,4′-dithiodipyridine on Ca2+ release [244]. Moreover, the C3602A mutation did not abolish the effect of calmodulin on Ca2+ release. Likewise, Nikolaienko et al. reported that this particular residue did not participate in inter-subunit cross-linking and played a limited role in RyR2 regulation by calmodulin during oxidative stress [245]. Based on our experience with high-resolution structures of RyR1 [246] and our ongoing research on the structure of human RyR2 [247], we mapped these previously identified residues to the 3D structure of RyR2 and discussed their potential relevance to RyR2 function (Figure 2). The 3D mapping confirmed that C3602 was located in the shell-core linker at the interface of the calmodulin binding site.
The other hyper-reactive cysteines C36, C253, and C315 are located in the N-terminal domain (NTD) of RyR1, while only C36 is conserved in RyR2. The N-terminal β-trefoil domains form the central cytosolic vestibule mediating inter-subunit interactions in the resting state of the channel. Since these interactions are disrupted during channel activation, it has been proposed that NTD is involved in the allosteric regulation of channel gating in both RyR1 and RyR2. Importantly, NTD is a known disease hot spot for pathogenic CPVT mutations weakening the RyR2 stability of the closed state [248,249,250]. However, it is unknown if oxidized residues, such as C36, in the NTD would have a similar CPVT-like pathology.
The hyper-active cysteine C822 is located in the SPRY1 (SPIa kinase and ryanodine receptor) domain, C917 in the tandem-repeat domain RY1&2, and C1582 in the SPRY3 region of RyR2. The SPRY domains represent important protein–protein interaction scaffolds in the periphery of the cytosolic shell of RyR. Interestingly, C1582 is located near the calstabin-binding site. Its oxidation can be important to modulating calstabin binding to the channel, which is critical for stabilizing the closed state of RyR and reducing spontaneous Ca2+ release. The RY1&2 domain is highly dynamic and its oxidation at C917 may contribute to the local regulation of this region.
The hyper-active cysteines C2402, C2572, C2577, and C3158 are located in the bridging solenoid (Bsol) of RyR2, another highly flexible and dynamic region of the protein. BSol encompasses the helical domains BSol1, BSol2, and BSol3. Between the first two is the RY3&4 domain that contains the PKA phosphorylation site (S2808) important for the adrenergic regulation of Ca2+ release by RyR2. Interestingly, helical domain 1 in the BSol1 domain is where C2402, C2572, and C2577 are located, and it exhibits a domain–domain interface with NTD. The disruption of NTD–Bsol interactions leads to inappropriate channel activation [251,252]. Similar to NTD, BSol1 represents another hotspot for pathogenic CPVT mutations, highlighting the important role of the NTD–Bsol interaction network in the allosteric modulation of the channel conformations. Residue C3158 is located in the BSol2 domain, which interacts with the RY1&2 and SPRY2 domains of the adjacent protomer.
Figure 2. Cysteine residues of ryanodine receptors prone to oxidation by ROS. (A) Evolutionary conservation of the hyper-active cysteines (red) across the RyR isoforms 1 and 2 in a rabbit, pig, mouse, and human. (B) Schematic diagram of the domain architecture of RyR2. Red circle in the diagram indicates the cysteines prone to oxidation for each RyR2 domain according to [243]. The same applies for the yellow circles, but according to [253] (C) Human RyR2 structure (PDB: 7UA3) showing the top and side views (left). The discussed oxidized cysteines are indicated in the close-up views (right). * means conserved residue among species.
Figure 2. Cysteine residues of ryanodine receptors prone to oxidation by ROS. (A) Evolutionary conservation of the hyper-active cysteines (red) across the RyR isoforms 1 and 2 in a rabbit, pig, mouse, and human. (B) Schematic diagram of the domain architecture of RyR2. Red circle in the diagram indicates the cysteines prone to oxidation for each RyR2 domain according to [243]. The same applies for the yellow circles, but according to [253] (C) Human RyR2 structure (PDB: 7UA3) showing the top and side views (left). The discussed oxidized cysteines are indicated in the close-up views (right). * means conserved residue among species.
Biomolecules 13 01409 g002
Nonetheless, through the application of genetic techniques to induce specific truncations and modifications in channel residues and domains, a recent study pinpointed that only two cysteine residues within each RyR2 subunit were accountable for nearly half of RyR2’s functional response when exposed to oxidative stress, specifically cysteine 1078 within the Ry1&2-SPRY2 linker and cysteine 2991 in the BSol loop function as a redox-sensitive pair. They engage in the formation of inter-subunit disulfide cross-links between adjacent subunits when confronted with oxidative stress, ultimately leading to an aberrant RyR2 Ca2+ leakage [253]. This finding aligns with the observation of elevated levels of RyR2 cross-linking in ventricular myocytes extracted from failing hearts [254].
Additionally, as previously discussed, while the previous research has pointed to cysteine 3635 in RyR1 or its equivalent, cysteine 3602, as the most hyperactive redox-sensitive residue, Nikolaienko et al. demonstrated that this particular residue did not contribute to channel cross-linking or the activation of Ca2+ leakage due to oxidative stress [245,253]. In concordance with previous discoveries, Nikolaienko et al. also showed that although none of the cysteine residues in the N-terminal domain (NTD) directly participated in cross-linking, they did play a pivotal role in allosterically supporting inter-subunit cross-linking within RyR2. Yet, none of the discussed residues were shown to be oxidized in human or animal models of HF. A further investigation into the specific mechanisms of RyR2 post-translational modification following oxidative stress, as well as the generation of cryo-EM structures with increasing resolution, is critical to understanding and designing therapeutic interventions for oxidative stress-mediated RyR2 pathological Ca2+ leak.
It has been observed that RyR channels self-organize into higher-order clusters forming checkerboard-like arrangements in the terminal cisternae of SR membranes. Peripheral domains likely involved in mediating clustering interactions between RyR channels are SPRY1, RY1&2 domains, and the C-terminal Bsol region [255,256]. Importantly, RyR clustering influences channel gating due to inter-RyR coupling [257,258]. Thus, we speculated that oxidation of hyper-active cysteines in these regions (C917, C822, and C3158) could modify RyR2 clustering and thereby its coordinated activity.

6.2. PKA and CamKII Oxidation and Modulation of SR Ca2+ Efflux

RyR2 is a large scaffold on which several regulatory proteins and enzymes can be assembled. Among them, two major regulatory kinases, protein kinase A (PKA) and Ca2+/calmodulin-dependent protein kinase II (CamKII), both known to be redox sensitive, have been shown to phosphorylate RyR2 at different sites and promote RyR2 Ca2+ leak. PKA and CamKII regulate RyR2 function and play an important role in regulating cardiac contractility and arrhythmogenesis [259,260,261,262,263,264,265,266,267,268,269]. Although it has been debated for decades, most studies lean toward the predominant role of the PKA-mediated phosphorylation of S2808 in HF. This post-translational modification depletes calstabin2 from RyR2 channels, causing Ca2+ leak, which in turn reduces cardiac contractility and ultimately furthers HF progression. Chronic PKA activation in HF is caused by the hyper-adrenergic state observed in HF patients, as well as in animal models of HF [4].
Furthermore, the type-I negative regulatory subunits of PKA are also subject to oxidation by ROS [270]. The oxidation of cysteines 17 and 38 leads to an inter-subunit disulfide bond formation between the two regulatory subunits and dissociation of the PKA holoenzyme complex. These processes allow PKA to phosphorylate several targets and increase cellular contractility without elevating 3′,5′-cyclic adenosine monophosphate (cAMP) levels [270], triggering pathological Ca2+ leak, which may lead to mitochondrial Ca2+ overload and dysfunction. To date, however, the exact physiological relevance of this ROS-dependent activation of PKA under physiological and pathological conditions, especially in comparison to adrenergic-dependent activation, remains mostly unknown.
Since CamKII has also demonstrated autonomous activation following the oxidation of its regulatory domain, the role of CamMKII oxidation in cardiovascular disease, including HF, has garnered significant attention [271]. The Anderson laboratory was one of the first to report that oxidized CamKII levels increased after myocardial infarction or angiotensin II injection, but not after isoproterenol [272]. They developed a panel of mutant CamKII to test the functional role of ROS-modifiable amino acids that were conserved across the four known CamKII isoforms. The authors suggested that CamKII could be activated by ROS as a consequence of M281/282 oxidation [272]. Animal models with a targeted loss of oxidized CamKII, in which M281/282 were replaced with valine, displayed decreased ROS induced-CamKII activation, reduced myocardial death, and improved left ventricular function after I/R injury. However, they retained other features of wild-type CamKII, including Ca2+/CaM autonomous activity due to T287 auto-phosphorylation [273]. Recently, Joiner and collaborators proposed that CamKII regulated mitochondrial Ca2+ uptake in the heart by phosphorylating the MCU, with subsequent impacts on mPTP activation and programmed cell death [274]. However, these findings were challenged by Fieni and coworkers, who did not detect significant effects of CamKII on MCU current density in experiments performed via the patch clamp of isolated mitochondria [275]. These data were consolidated using genetic approaches involving two mouse models with either the global deletion of CamKIIσ or the cardiomyocyte-specific deletion of CamKIIσ/ϒ, showing no significant effects on Ca2+ uptake, respiration, or ROS emission in isolated cardiac mitochondria, nor in isolated cardiac myocytes [276].
In addition to RyR2, PKA, and CamKII, SERCA2a has also been shown to experience dysregulation following oxidation by ROS. By treating cardiomyocytes with various reactive oxygen species, several studies have demonstrated decreased SERCA2a Ca2+ uptake into the SR, which may increase mitochondrial Ca2+ overload and further ROS production [277,278,279].

7. Mitochondrial Ca2+ Overload and Cell Death

Elevated mitochondrial matrix Ca2+ triggers the opening of the mitochondrial permeability transition pore (mPTP), a non-specific pore in the inner membrane of the mitochondria [280,281], converting them from organelles, whose production of ATP sustains the cell life, to instruments of cell death [282]. Thus, water and all solutes <1500 Da enter the mitochondrial matrix, leading to mitochondrial swelling, the unfolding of inner mitochondrial membrane cristae, and the disruption of the outer mitochondrial membrane, causing the release of cytochrome c and SMAC/Diablo into the cytosol and, eventually, apoptosis [282,283,284,285]. Pathological conditions that can open the mPTP include Ca2+ overload, production of ROS, depletion of ATP and ADP, and increases in inorganic phosphate [283], all of which are generally present during HF.

8. Therapies Targeting Ca2+ Defects/Mitochondrial Oxidative Stress in HF

As previously discussed, defective Ca2+ handling, mitochondrial dysfunction, and oxidative stress have been linked to the development and progression of HF, independent of etiology. Therefore, the SR and mitochondria may serve as two of the most promising therapeutic targets for HF [286,287,288,289,290,291]. Several drugs targeting mitochondria are currently under evaluation in clinical trials for HF patients. In light of these targeted therapies, an understanding of impaired mitochondrial Ca2+ handling, as well as its relation to altered cytosolic Ca2+, appears pivotal. Several strategies targeting mitochondria or the SR, including small molecules, peptides, and antioxidants that have been tested in preclinical and clinical studies in HF, are summarized in Table 1. We discuss in detail the properties and outcomes of two recent compounds that target the SR and mitochondria, Rycals and SS31, respectively.

8.1. Ca2+ Dysregulation-Targeted Therapies: Rycals

As mentioned above, RyR2 is an essential player in several functions of cardiomyocytes, from EC coupling to the activation of numerous pathways, gene expression, as well as mitochondrial Ca2+ overload and subsequent oxidative stress, thereby controlling the function and fate of cardiomyocytes. The impairment of RyR2 Ca2+ release has emerged as a major mechanism underlying HF and arrhythmogenesis as a consequence of genetic mutations and/or stress-mediated RyR2 post-translational modifications, loss of channel stability, and impairment of coupled gating.
A family of compounds known as Rycals, a small, orally available molecule, prevent RyR2 Ca2+ leak by preventing calstabin2 depletion without blocking the channel pore [246,247]. Promising results from the preclinical studies show the improvement of cardiac function in mice with dysfunctional RyR2. In rat and mouse models of post-myocardial infarction, RyR stabilizing compounds attenuated systolic and diastolic cardiac dysfunctions and preserved cardiac structure [292,293]. This finding was supported by recent experiments involving Rycal treatment of hiPS-derived cardiomyocytes obtained from skin fibroblasts or hair keratinocytes of patients with defective RyR2 mutations [283].

8.2. Mitochondria-Targeted Antioxidant SS31

SS31, a potent antioxidant peptide with the ability to penetrate cell membranes, is located within the inner mitochondrial membrane, where it exhibits protective properties for mitochondria [294,295]. Experimental studies conducted before the clinical trials have shown that SS31 provides protection in models of heart failure induced by pressure overload and hypertensive cardiomyopathy [186,296]. In the latter investigation, SS31 was found to prevent the accumulation of reactive oxygen species in mitochondria and alleviate angiotensin II-induced cardiac hypertrophy and diastolic dysfunction [294]. Furthermore, a newer version of SS31, known as Elamipretide (also referred to as Bendavia), was subjected to testing in two clinical trials involving a small group of patients [297]. One trial involved patients with heart failure characterized by a reduced ejection fraction (HFrEF, with 36 patients enrolled—NCT02388464) and the other focused on patients with an ST-segment elevation myocardial infarction (STEMI, with 300 patients enrolled—NCT01572909-EMBRACE) [298,299]. Both trials indicated acceptable safety and tolerability. However, the phase-2a trial EMBRACE did not demonstrate a significant improvement in the primary endpoint, which was myocardial infarct size reduction, following treatment. Subsequently, two modified versions of SS31, named mtCPP and mtgCPP, were ben developed, showcasing a 2–3-times-greater efficacy and antioxidant capacity compared to SS31 [300,301]. Nevertheless, whether these modified peptides possess the potential to emerge as new drug candidates for the treatment of heart failure in clinical settings remains to be described. Furthermore, despite the promising data and outcomes provided by SS31, there are no studies showing its effect on cardiac Ca2+ homeostasis, known as the major mechanism in cardiomyopathies, arrythmias, and death.
Table 1. Oxidative stress/Ca2+ dysregulation-targeted therapies in cardiac disease.
Table 1. Oxidative stress/Ca2+ dysregulation-targeted therapies in cardiac disease.
AgentTargetActionDiseasesDose/SpeciesDevelopmentOutcomesReferences
AAV1/SERCA2aSERCA2aSERCA2a overexpressionHeart failure
Ischemic cardiomyopathy
Patients
1 × 1011–1 × 1013 DNase particles
Terminated
clinical trials
No effect/
reduced cardiac events 1 year after
[277,302]
NCT00534703
NCT01966887
NCT02346422
Alda-1Mito
ALDH2
Increased ALDH2 activityI/R injury
HF post-MI
Rat 8.5 mg/kgPre-clinicalReduced infarct size by 60%[303,304,305,306]
Cyclosporine AMTPMTP inhibitionI/R injury
HF post-MI
Patients 2.5 mg/Kg
IV
Phase 3Reduced infarct size[307,308,309]
EUK-8SODSOD/catalase mimeticDCM
Pressure overload-induced HF
Mice 30 mg/kg/day
IP
Pre-clinicalPrevented DCM in mice
Ameliorated systolic function and survival
[310,311]
IdebenoneCoenzyme Q10Free radical scavengerMitochondrial cardiomyopathy225 mg/day
In patients
Pre-clinical phase 3Increased EF by >50%
(a case report)
[312]
M40403SODSOD/catalase mimeticI/R heart injuryRats 1 to 10 mg/kg
IV
Pre-clinicalProtected tissue damage after I/R in rats[313,314]
mitoTEMPOMitochondrial
nitroxide
ROS scavengerDiabetic cardiomyopathy
Hypertension
Mice 0.7 mg/kg/day
IP
Pre-clinicalReduced myocardial hypertrophy[315,316,317]
MCI-186
(Edaravone)
Free radicalsFree radical scavengerHF
Acute ischemic stroke
patients
30 mg
IV
Pre-clinical
phases 2–4
Reduced enzymatic infarcts/better clinical outcomes[318,319]
MitoQETCFree radical scavengerPressure overload
Cardiovascular function
Mice 100 uM/DW
Rats
Unavailable for patients
Pre-clinical
2 clinical trials
Decreased heart dysfunction[320,321]
NCT03506633
NCT03586414
MetforminETCETC inhibitionI/R injury
HF post-MI
HFpEF
Mice and rats 200–250 mg/kgPhase 2Improved cardiac function (rats and mice)[322,323]
NCT03629340
Rycals (S107-ARM210)Ryanodine receptor (RyR)Stabilizing RyRI/R injury
HF post-MI
Mice 20–75 mg/kg oralPreclinical studyImproved cardiac function/reduced arrhythmias[324]
NCT04141670
SS31MitochondriaCardiolipin protectionHFrEF, HFpEF
Congestive HF
Patients 0.25 mg/kg/hClinical phases 1 and 2Improved cardiac volumes[298]
NCT02814097
NCT02914665
TRO-40303MTPMTP inhibitionI/R injury
HF post MI
Patients 2.5 mg/kgPhase 2No effect
Reduced infarct size by 38%
[325,326,327]
XJB-5-131Mitochondrial
nitroxide
ROS scavengerI/R heart injuryRats 3 mg/Kg
IP
Pre-clinicalImproved post-ischemic recovery[278,279]

9. Conclusions

Beyond their crucial role as an energy source, cardiac mitochondria serve as Ca2+ reservoirs. They accumulate Ca2+ ions during cytosolic Ca2+ elevations in cardiomyocytes. Elevations in intramitochondrial Ca2+ levels trigger multiple enzymes within the mitochondrial matrix. These enzymes, in turn, boost ATP generation, modify the patterns of intracellular Ca2+ signaling across both space and time, and assume a pivotal role in instigating mitochondrial oxidative stress and pathways leading to cell death. It is clear that, in HF, an excessive SR Ca2+ leak, mainly through RyR2 channels, plays a crucial role in the pathophysiology of mitochondrial Ca2+ accumulation and dysfunction and constitutes a feedback loop of alterations between the SR and mitochondria, which contribute to the impairment of cardiac function. Future investigations should focus on breaking this feedback loop by simultaneously targeting the SR leak and mitochondrial Ca2+ overload/oxidative stress using RyR2 stabilizing drugs and mitochondrial antioxidants.

Author Contributions

All authors contributed to the writing and revision of the paper. H.D. supervised the work. A.R.M. supervised the work and acquired Funding. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the NIH to A.R.M., grant number (T32HL120826, R01HL145473, R01DK118240, R01HL142903, R01HL140934, R01AR070194, R25HL156002, and RF1NS114570).

Conflicts of Interest

Andrew R. Marks and Columbia University own shares in ARMGO Pharma, Inc., a biotechnology company developing RyR targeted drugs. All the remaining authors declare no conflict of interest.

Abbreviations

HFHeart failure
HFrEFHeart Failure with reduced ejection fraction
SRSarcoplasmic reticulum
EREndoplasmic reticulum
ATPAdenosine triphosphate
Ca2+Calcium
ROSReactive oxygen species
RyR2Ryanodine receptor type 2
IP3Rs1,4,5-trisphosphate receptors
ECExcitation–contraction
ETCElectron transport chain
NADHNicotinamide adenine dinucleotide
FADH2Flavin adenine dinucleotide
ADPAdenosine diphosphate
CKCreatine kinase
CICRCalcium-induced calcium release
SERCA2aSarco(endo)plasmic reticulum calcium-ATPase 2a
NCXSarcolemmal Na+/Ca2+ exchanger
TNF-⍺Tumor necrosis factor
IL-17Interleukin-17
IL-6Interleukin-6
IL-1ßInterleukin-1
TGF-ßTransforming growth factor
NOX2NADPH oxidase 2
O2·Superoxide radical
MCUMitochondrial Ca2+ uniporter
mNCLXMitochondrial Na+-Ca2+—Li+ exchanger
Letm1Mitochondrial proton/calcium exchanger protein
TRPC3Transient receptor potential canonical 3
MFN2Mitofusin2
MFN1Mitofusin1
Grp75Chaperone glucose-regulated protein 75
VDACVoltage-dependent anion channel
mPTPMitochondrial permeability transition pore
MAMMitochondria-associated membrane
LBDLigand-binding domain (LBD)
PDZD8PDZ-domain-containing 8
Rab7Ras-related protein Rab-7a
FUNDC1FUN14-domain-containing 1
PTPIP51Protein tyrosine phosphatase-interacting protein 51
VAPBVesicle-associated membrane protein-associated protein B
·OHHydroxyl radical
H2O2Hydrogen peroxide
SODSuperoxide dismutase
PrxPeroxiredoxin
GpxGlutathione peroxidase
Cryo-EMCryogenic-electron microscopy
NTDN-terminal domain
CPVTCatecholaminergic polymorphic ventricular tachycardia
SPRY1SPIa kinase and ryanodine receptor
BsolBridging solenoid
PKAProtein kinase A
CamKIICa2+/calmodulin-dependent protein kinase II
cAMP3′,5′-cyclic adenosine monophosphate
hiPSHuman-induced pluripotent stem

References

  1. Black, H.R. The burden of cardiovascular disease: Following the link from hypertension to myocardial infarction and heart failure. Am. J. Hypertens. 2003, 16 Pt 2, 4S–6S. [Google Scholar] [CrossRef] [PubMed]
  2. Benjamin, E.J.; Muntner, P.; Alonso, A.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Chang, A.R.; Cheng, S.; Das, S.R.; et al. Heart Disease and Stroke Statistics-2019 Update: A Report from the American Heart Association. Circulation 2019, 139, e56–e528. [Google Scholar] [CrossRef]
  3. Grossman, W.; Jones, D.; McLaurin, L.P. Wall stress and patterns of hypertrophy in the human left ventricle. J. Clin. Investig. 1975, 56, 56–64. [Google Scholar] [CrossRef]
  4. Santulli, G. Sympathetic nervous system signaling in heart failure and cardiac aging. In Pathophysiology and Pharmacotherapy of Cardiovascular Disease; Adis: Cham, Switzerland, 2015; pp. 83–105. [Google Scholar] [CrossRef]
  5. Burkart, F.; Kiowski, W. Circulatory abnormalities and compensatory mechanisms in heart failure. Am. J. Med. 1991, 90, 19S–22S. [Google Scholar] [CrossRef]
  6. Arai, M.; Alpert, N.R.; MacLennan, D.H.; Barton, P.; Periasamy, M. Alterations in sarcoplasmic reticulum gene expression in human heart failure. A possible mechanism for alterations in systolic and diastolic properties of the failing myocardium. Circ. Res. 1993, 72, 463–469. [Google Scholar] [CrossRef] [PubMed]
  7. Iemitsu, M.; Miyauchi, T.; Maeda, S.; Sakai, S.; Kobayashi, T.; Fujii, N.; Miyazaki, H.; Matsuda, M.; Yamaguchi, I.; Krzesiak, A.; et al. Physiological and pathological cardiac hypertrophy induce different molecular phenotypes in the rat. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2001, 281, R2029–R2036. [Google Scholar] [CrossRef]
  8. Bernardo, B.C.; Weeks, K.L.; Pretorius, L.; McMullen, J.R. Molecular distinction between physiological and pathological cardiac hypertrophy: Experimental findings and therapeutic strategies. Pharmacol. Ther. 2010, 128, 191–227. [Google Scholar] [CrossRef]
  9. Takahashi, T.; Allen, P.D.; Izumo, S. Expression of A-, B-, and C-type natriuretic peptide genes in failing and developing human ventricles. Correlation with expression of the Ca2+-ATPase gene. Circ. Res. 1992, 71, 9–17. [Google Scholar] [CrossRef] [PubMed]
  10. Mittmann, C.; Eschenhagen, T.; Scholz, H. Cellular and molecular aspects of contractile dysfunction in heart failure. Cardiovasc. Res. 1998, 39, 267–275. [Google Scholar] [CrossRef] [PubMed]
  11. Sachse, F.B.; Torres, N.S.; Savio-Galimberti, E.; Aiba, T.; Kass, D.A.; Tomaselli, G.F.; Bridge, J.H. Subcellular structures and function of myocytes impaired during heart failure are restored by cardiac resynchronization therapy. Circ. Res. 2012, 110, 588–597. [Google Scholar] [CrossRef]
  12. Mohammed, S.F.; Hussain, S.; Mirzoyev, S.A.; Edwards, W.D.; Maleszewski, J.J.; Redfield, M.M. Coronary microvascular rarefaction and myocardial fibrosis in heart failure with preserved ejection fraction. Circulation 2015, 131, 550–559. [Google Scholar] [CrossRef] [PubMed]
  13. Gogiraju, R.; Bochenek, M.L.; Schafer, K. Angiogenic Endothelial Cell Signaling in Cardiac Hypertrophy and Heart Failure. Front. Cardiovasc. Med. 2019, 6, 20. [Google Scholar] [CrossRef]
  14. Belviso, I.; Angelini, F.; Di Meglio, F.; Picchio, V.; Sacco, A.M.; Nocella, C.; Romano, V.; Nurzynska, D.; Frati, G.; Maiello, C.; et al. The Microenvironment of Decellularized Extracellular Matrix from Heart Failure Myocardium Alters the Balance between Angiogenic and Fibrotic Signals from Stromal Primitive Cells. Int. J. Mol. Sci. 2020, 21, 7903. [Google Scholar] [PubMed]
  15. Zile, M.R.; O’Meara, E.; Claggett, B.; Prescott, M.F.; Solomon, S.D.; Swedberg, K.; Packer, M.; McMurray, J.J.; Shi, V.; Lefkowitz, M.; et al. Effects of Sacubitril/Valsartan on Biomarkers of Extracellular Matrix Regulation in Patients with HFrEF. J. Am. Coll. Cardiol. 2019, 73, 795–806. [Google Scholar] [CrossRef]
  16. Richter, M.; Kostin, S. The failing human heart is characterized by decreased numbers of telocytes as result of apoptosis and altered extracellular matrix composition. J. Cell Mol. Med. 2015, 19, 2597–2606. [Google Scholar] [CrossRef] [PubMed]
  17. Frangogiannis, N.G. The Extracellular Matrix in Ischemic and Nonischemic Heart Failure. Circ. Res. 2019, 125, 117–146. [Google Scholar] [CrossRef]
  18. Frangogiannis, N.G. Cardiac fibrosis. Cardiovasc. Res. 2021, 117, 1450–1488. [Google Scholar] [CrossRef]
  19. Kong, P.; Christia, P.; Frangogiannis, N.G. The pathogenesis of cardiac fibrosis. Cell. Mol. Life Sci. 2014, 71, 549–574. [Google Scholar]
  20. Gargiulo, P.; Marsico, F.; Renga, F.; Dell’Aversana, S.; Esposito, I.; Marciano, C.; Dellegrottaglie, S.; Perrone-Filardi, P.; Paolillo, S. The metabolic syndrome in heart failure: Insights to specific mechanisms. Heart Fail. Rev. 2019, 25, 1–7. [Google Scholar] [CrossRef]
  21. Harada, T.; Sunaga, H.; Sorimachi, H.; Yoshida, K.; Kato, T.; Kurosawa, K.; Nagasaka, T.; Koitabashi, N.; Iso, T.; Kurabayashi, M.; et al. Pathophysiological role of fatty acid-binding protein 4 in Asian patients with heart failure and preserved ejection fraction. ESC Heart Fail. 2020, 7, 4256–4266. [Google Scholar] [CrossRef]
  22. Hunter, W.G.; Kelly, J.P.; McGarrah, R.W., 3rd; Khouri, M.G.; Craig, D.; Haynes, C.; Ilkayeva, O.; Stevens, R.D.; Bain, J.R.; Muehlbauer, M.J.; et al. Metabolomic Profiling Identifies Novel Circulating Biomarkers of Mitochondrial Dysfunction Differentially Elevated in Heart Failure with Preserved Versus Reduced Ejection Fraction: Evidence for Shared Metabolic Impairments in Clinical Heart Failure. J. Am. Heart Assoc. 2016, 5, e003190. [Google Scholar] [CrossRef]
  23. Murray, A.J.; Cole, M.A.; Lygate, C.A.; Carr, C.A.; Stuckey, D.J.; Little, S.E.; Neubauer, S.; Clarke, K. Increased mitochondrial uncoupling proteins, respiratory uncoupling and decreased efficiency in the chronically infarcted rat heart. J. Mol. Cell. Cardiol. 2008, 44, 694–700. [Google Scholar] [CrossRef] [PubMed]
  24. De Rosa, M.; Gambardella, J.; Shu, J.; Santulli, G. Dietary fat is a key determinant in balancing mitochondrial dynamics in heart failure: A novel mechanism underlying the obesity paradox. Cardiovasc. Res. 2018, 114, 925–927. [Google Scholar] [CrossRef]
  25. Seki, M.; Powers, J.C.; Maruyama, S.; Zuriaga, M.A.; Wu, C.L.; Kurishima, C.; Kim, L.; Johnson, J.; Poidomani, A.; Wang, T.; et al. Acute and Chronic Increases of Circulating FSTL1 Normalize Energy Substrate Metabolism in Pacing-Induced Heart Failure. Circ. Heart Fail. 2018, 11, e004486. [Google Scholar] [CrossRef]
  26. Fukushima, A.; Lopaschuk, G.D. Acetylation control of cardiac fatty acid beta-oxidation and energy metabolism in obesity, diabetes, and heart failure. Biochim. Biophys. Acta 2016, 1862, 2211–2220. [Google Scholar] [CrossRef] [PubMed]
  27. Abdurrachim, D.; Luiken, J.J.; Nicolay, K.; Glatz, J.F.; Prompers, J.J.; Nabben, M. Good and bad consequences of altered fatty acid metabolism in heart failure: Evidence from mouse models. Cardiovasc. Res. 2015, 106, 194–205. [Google Scholar] [CrossRef] [PubMed]
  28. Neubauer, S. The failing heart—An engine out of fuel. N. Engl. J. Med. 2007, 356, 1140–1151. [Google Scholar] [CrossRef]
  29. Doenst, T.; Nguyen, T.D.; Abel, E.D. Cardiac metabolism in heart failure: Implications beyond ATP production. Circ. Res. 2013, 113, 709–724. [Google Scholar] [CrossRef]
  30. Schaper, J.; Meiser, E.; Stammler, G. Ultrastructural morphometric analysis of myocardium from dogs, rats, hamsters, mice, and from human hearts. Circ. Res. 1985, 56, 377–391. [Google Scholar] [CrossRef]
  31. Booth, D.M.; Enyedi, B.; Geiszt, M.; Varnai, P.; Hajnoczky, G. Redox Nanodomains Are Induced by and Control Calcium Signaling at the ER-Mitochondrial Interface. Mol. Cell 2016, 63, 240–248. [Google Scholar] [CrossRef]
  32. Ramesh, V.; Sharma, V.K.; Sheu, S.S.; Franzini-Armstrong, C. Structural proximity of mitochondria to calcium release units in rat ventricular myocardium may suggest a role in Ca2+ sequestration. Ann. N. Y. Acad. Sci. 1998, 853, 341–344. [Google Scholar] [CrossRef]
  33. Szabadkai, G.; Bianchi, K.; Varnai, P.; De Stefani, D.; Wieckowski, M.R.; Cavagna, D.; Nagy, A.I.; Balla, T.; Rizzuto, R. Chaperone-mediated coupling of endoplasmic reticulum and mitochondrial Ca2+ channels. J. Cell Biol. 2006, 175, 901–911. [Google Scholar] [CrossRef]
  34. Boncompagni, S.; Rossi, A.E.; Micaroni, M.; Beznoussenko, G.V.; Polishchuk, R.S.; Dirksen, R.T.; Protasi, F. Mitochondria are linked to calcium stores in striated muscle by developmentally regulated tethering structures. Mol. Biol. Cell 2009, 20, 1058–1067. [Google Scholar] [CrossRef] [PubMed]
  35. Li, Q.; Su, D.; O’Rourke, B.; Pogwizd, S.M.; Zhou, L. Mitochondria-derived ROS bursts disturb Ca2+ cycling and induce abnormal automaticity in guinea pig cardiomyocytes: A theoretical study. Am. J. Physiol. Heart Circ. Physiol. 2015, 308, H623–H636. [Google Scholar] [CrossRef]
  36. Zhou, L.; Aon, M.A.; Liu, T.; O’Rourke, B. Dynamic modulation of Ca2+ sparks by mitochondrial oscillations in isolated guinea pig cardiomyocytes under oxidative stress. J. Mol. Cell. Cardiol. 2011, 51, 632–639. [Google Scholar] [CrossRef] [PubMed]
  37. Yan, Y.; Liu, J.; Wei, C.; Li, K.; Xie, W.; Wang, Y.; Cheng, H. Bidirectional regulation of Ca2+ sparks by mitochondria-derived reactive oxygen species in cardiac myocytes. Cardiovasc. Res. 2008, 77, 432–441. [Google Scholar] [CrossRef]
  38. Bragadin, M.; Pozzan, T.; Azzone, G.F. Kinetics of Ca2+ carrier in rat liver mitochondria. Biochemistry 1979, 18, 5972–5978. [Google Scholar] [CrossRef]
  39. Massari, S.; Pozzan, T. The interaction of organic cations with the mitochondrial membrane. Experientia 1976, 32, 868–869. [Google Scholar] [CrossRef] [PubMed]
  40. Massari, S.; Pozzan, T. The accumulation ratio of K+, Na+, Ca2+ and tetrapropylammonium in steady-state Mitochondria. Arch. Biochem. Biophys. 1976, 173, 332–340. [Google Scholar] [CrossRef]
  41. Tromp, J.; Khan, M.A.; Klip, I.T.; Meyer, S.; de Boer, R.A.; Jaarsma, T.; Hillege, H.; van Veldhuisen, D.J.; van der Meer, P.; Voors, A.A. Biomarker Profiles in Heart Failure Patients with Preserved and Reduced Ejection Fraction. J. Am. Heart Assoc. 2017, 6, e003989. [Google Scholar] [CrossRef]
  42. Bianchi, V.E. Nutrition in chronic heart failure patients: A systematic review. Heart Fail. Rev. 2020, 25, 1017–1026. [Google Scholar] [CrossRef] [PubMed]
  43. Padmanabhan, A.; Alexanian, M.; Linares-Saldana, R.; Gonzalez-Teran, B.; Andreoletti, G.; Huang, Y.; Connolly, A.J.; Kim, W.; Hsu, A.; Duan, Q.; et al. BRD4 Interacts with GATA4 to Govern Mitochondrial Homeostasis in Adult Cardiomyocytes. Circulation 2020, 142, 2338–2355. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, L.; Song, J.; Hu, S. Metabolic remodeling of substrate utilization during heart failure progression. Heart Fail. Rev. 2019, 24, 143–154. [Google Scholar] [CrossRef]
  45. Santulli, G.; Lewis, D.; des Georges, A.; Marks, A.R.; Frank, J. Ryanodine Receptor Structure and Function in Health and Disease. Subcell Biochem. 2018, 87, 329–352. [Google Scholar]
  46. Marx, S.O.; Marks, A.R. Dysfunctional ryanodine receptors in the heart: New insights into complex cardiovascular diseases. J. Mol. Cell. Cardiol. 2013, 58, 225–231. [Google Scholar] [CrossRef]
  47. Marks, A.R. Calcium cycling proteins and heart failure: Mechanisms and therapeutics. J. Clin. Investig. 2013, 123, 46–52. [Google Scholar] [CrossRef] [PubMed]
  48. Huang, F.; Shan, J.; Reiken, S.; Wehrens, X.H.; Marks, A.R. Analysis of calstabin2 (FKBP12.6)-ryanodine receptor interactions: Rescue of heart failure by calstabin2 in mice. Proc. Natl. Acad. Sci. USA 2006, 103, 3456–3461. [Google Scholar] [CrossRef]
  49. Gambardella, J.; Trimarco, B.; Iaccarino, G.; Santulli, G. New Insights in Cardiac Calcium Handling and Excitation-Contraction Coupling. Adv. Exp. Med. Biol. 2018, 1067, 373–385. [Google Scholar]
  50. Shan, J.; Kushnir, A.; Betzenhauser, M.J.; Reiken, S.; Li, J.; Lehnart, S.E.; Lindegger, N.; Mongillo, M.; Mohler, P.J.; Marks, A.R. Phosphorylation of the ryanodine receptor mediates the cardiac fight or flight response in mice. J. Clin. Investig. 2010, 120, 4388–4398. [Google Scholar] [CrossRef]
  51. Santulli, G.; Xie, W.; Reiken, S.R.; Marks, A.R. Mitochondrial calcium overload is a key determinant in heart failure. Proc. Natl. Acad. Sci. USA 2015, 112, 11389–11394. [Google Scholar] [CrossRef]
  52. Schwarz, K.; Siddiqi, N.; Singh, S.; Neil, C.J.; Dawson, D.K.; Frenneaux, M.P. The breathing heart—Mitochondrial respiratory chain dysfunction in cardiac disease. Int. J. Cardiol. 2014, 171, 134–143. [Google Scholar] [CrossRef]
  53. Forini, F.; Nicolini, G.; Kusmic, C.; Iervasi, G. Protective Effects of Euthyroidism Restoration on Mitochondria Function and Quality Control in Cardiac Pathophysiology. Int. J. Mol. Sci. 2019, 20, 3377. [Google Scholar] [CrossRef]
  54. Li, X.; Lu, W.J.; Li, Y.; Wu, F.; Bai, R.; Ma, S.; Dong, T.; Zhang, H.; Lee, A.S.; Wang, Y.; et al. MLP-deficient human pluripotent stem cell derived cardiomyocytes develop hypertrophic cardiomyopathy and heart failure phenotypes due to abnormal calcium handling. Cell Death Dis. 2019, 10, 610. [Google Scholar] [CrossRef]
  55. Hamilton, D.J.; Zhang, A.; Li, S.; Cao, T.N.; Smith, J.A.; Vedula, I.; Cordero-Reyes, A.M.; Youker, K.A.; Torre-Amione, G.; Gupte, A.A.; et al. Combination of angiotensin II and l-NG-nitroarginine methyl ester exacerbates mitochondrial dysfunction and oxidative stress to cause heart failure. Am. J. Physiol. Heart Circ. Physiol. 2016, 310, H667–H680. [Google Scholar] [CrossRef] [PubMed]
  56. Rosca, M.G.; Vazquez, E.J.; Kerner, J.; Parland, W.; Chandler, M.P.; Stanley, W.; Sabbah, H.N.; Hoppel, C.L. Cardiac mitochondria in heart failure: Decrease in respirasomes and oxidative phosphorylation. Cardiovasc. Res. 2008, 80, 30–39. [Google Scholar] [CrossRef]
  57. Leverve, X.M. Mitochondrial function and substrate availability. Crit. Care Med. 2007, 35 (Suppl. S9), S454–S460. [Google Scholar] [CrossRef]
  58. Erecinska, M.; Wilson, D.F. Regulation of cellular energy metabolism. J. Membr. Biol. 1982, 70, 1–14. [Google Scholar] [CrossRef]
  59. Kampjut, D.; Sazanov, L.A. Structure and mechanism of mitochondrial proton-translocating transhydrogenase. Nature 2019, 573, 291–295. [Google Scholar] [CrossRef] [PubMed]
  60. Kory, N.; Uit de Bos, J.; van der Rijt, S.; Jankovic, N.; Gura, M.; Arp, N.; Pena, I.A.; Prakash, G.; Chan, S.H.; Kunchok, T.; et al. MCART1/SLC25A51 is required for mitochondrial NAD transport. Sci. Adv. 2020, 6, eabe5310. [Google Scholar] [CrossRef] [PubMed]
  61. Nowinski, S.M.; Solmonson, A.; Rusin, S.F.; Maschek, J.A.; Bensard, C.L.; Fogarty, S.; Jeong, M.Y.; Lettlova, S.; Berg, J.A.; Morgan, J.T.; et al. Mitochondrial fatty acid synthesis coordinates oxidative metabolism in mammalian mitochondria. eLife 2020, 9, e58041. [Google Scholar] [CrossRef]
  62. Balsa, E.; Perry, E.A.; Bennett, C.F.; Jedrychowski, M.; Gygi, S.P.; Doench, J.G.; Puigserver, P. Defective NADPH production in mitochondrial disease complex I causes inflammation and cell death. Nat. Commun. 2020, 11, 2714. [Google Scholar] [CrossRef] [PubMed]
  63. Jouaville, L.S.; Pinton, P.; Bastianutto, C.; Rutter, G.A.; Rizzuto, R. Regulation of mitochondrial ATP synthesis by calcium: Evidence for a long-term metabolic priming. Proc. Natl. Acad. Sci. USA 1999, 96, 13807–13812. [Google Scholar] [CrossRef]
  64. Gong, G.; Liu, X.; Wang, W. Regulation of metabolism in individual mitochondria during excitation-contraction coupling. J. Mol. Cell. Cardiol. 2014, 76, 235–246. [Google Scholar] [CrossRef] [PubMed]
  65. Piquereau, J.; Veksler, V.; Novotova, M.; Ventura-Clapier, R. Energetic Interactions between Subcellular Organelles in Striated Muscles. Front. Cell Dev. Biol. 2020, 8, 581045. [Google Scholar] [CrossRef]
  66. Cortassa, S.; Aon, M.A.; Marban, E.; Winslow, R.L.; O’Rourke, B. An integrated model of cardiac mitochondrial energy metabolism and calcium dynamics. Biophys. J. 2003, 84, 2734–2755. [Google Scholar] [CrossRef]
  67. Glancy, B.; Balaban, R.S. Role of mitochondrial Ca2+ in the regulation of cellular energetics. Biochemistry 2012, 51, 2959–2973. [Google Scholar] [CrossRef]
  68. Yang, K.C.; Kyle, J.W.; Makielski, J.C.; Dudley, S.C., Jr. Mechanisms of sudden cardiac death: Oxidants and metabolism. Circ. Res. 2015, 116, 1937–1955. [Google Scholar] [CrossRef] [PubMed]
  69. Guo, A.; Wang, Y.; Chen, B.; Wang, Y.; Yuan, J.; Zhang, L.; Hall, D.; Wu, J.; Shi, Y.; Zhu, Q.; et al. E-C coupling structural protein junctophilin-2 encodes a stress-adaptive transcription regulator. Science 2018, 362, eaan3303. [Google Scholar] [CrossRef]
  70. Ziman, A.P.; Gomez-Viquez, N.L.; Bloch, R.J.; Lederer, W.J. Excitation-contraction coupling changes during postnatal cardiac development. J. Mol. Cell. Cardiol. 2010, 48, 379–386. [Google Scholar] [CrossRef]
  71. Firth, J.M.; Yang, H.Y.; Francis, A.J.; Islam, N.; MacLeod, K.T. The Effect of Estrogen on Intracellular Ca2+ and Na+ Regulation in Heart Failure. JACC Basic Transl. Sci. 2020, 5, 901–912. [Google Scholar] [CrossRef]
  72. Arbel-Ganon, L.; Behar, J.A.; Gomez, A.M.; Yaniv, Y. Distinct mechanisms mediate pacemaker dysfunction associated with catecholaminergic polymorphic ventricular tachycardia mutations: Insights from computational modeling. J. Mol. Cell. Cardiol. 2020, 143, 85–95. [Google Scholar] [CrossRef]
  73. Abu-Khousa, M.; Fiegle, D.J.; Sommer, S.T.; Minabari, G.; Milting, H.; Heim, C.; Weyand, M.; Tomasi, R.; Dendorfer, A.; Volk, T.; et al. The Degree of t-System Remodeling Predicts Negative Force-Frequency Relationship and Prolonged Relaxation Time in Failing Human Myocardium. Front. Physiol. 2020, 11, 182. [Google Scholar] [CrossRef] [PubMed]
  74. Scranton, K.; John, S.; Escobar, A.; Goldhaber, J.I.; Ottolia, M. Modulation of the cardiac Na+-Ca2+ exchanger by cytoplasmic protons: Molecular mechanisms and physiological implications. Cell Calcium 2020, 87, 102140. [Google Scholar] [CrossRef]
  75. Li, S.; Chopra, A.; Keung, W.; Chan, C.W.Y.; Costa, K.D.; Kong, C.W.; Hajjar, R.J.; Chen, C.S.; Li, R.A. Sarco/endoplasmic reticulum Ca2+-ATPase is a more effective calcium remover than sodium-calcium exchanger in human embryonic stem cell-derived cardiomyocytes. Am. J. Physiol. Heart Circ. Physiol. 2019, 317, H1105–H1115. [Google Scholar] [CrossRef] [PubMed]
  76. Dridi, H.; Kushnir, A.; Zalk, R.; Yuan, Q.; Melville, Z.; Marks, A.R. Intracellular calcium leak in heart failure and atrial fibrillation: A unifying mechanism and therapeutic target. Nat. Rev. Cardiol. 2020, 17, 732–747. [Google Scholar] [PubMed]
  77. Schroder, F.; Handrock, R.; Beuckelmann, D.J.; Hirt, S.; Hullin, R.; Priebe, L.; Schwinger, R.H.G.; Weil, J.; Herzig, S. Increased availability and open probability of single L-type calcium channels from failing compared with nonfailing human ventricle. Circulation 1998, 98, 969–976. [Google Scholar] [CrossRef] [PubMed]
  78. Kamp, T.J.; Hell, J.W. Regulation of cardiac L-type calcium channels by protein kinase A and protein kinase C. Circ. Res. 2000, 87, 1095–1102. [Google Scholar] [CrossRef]
  79. Sipido, K.R.; Volders, P.G.; Vos, M.A.; Verdonck, F. Altered Na/Ca exchange activity in cardiac hypertrophy and heart failure: A new target for therapy? Cardiovasc. Res. 2002, 53, 782–805. [Google Scholar] [CrossRef]
  80. Hasenfuss, G.; Reinecke, H.; Studer, R.; Meyer, M.; Pieske, B.; Holtz, J.; Holubarsch, C.; Posival, H.; Just, H.; Drexler, H. Relation between myocardial function and expression of sarcoplasmic reticulum Ca2+-ATPase in failing and nonfailing human myocardium. Circ. Res. 1994, 75, 434–442. [Google Scholar] [CrossRef]
  81. Hobai, I.A.; O’Rourke, B. Decreased sarcoplasmic reticulum calcium content is responsible for defective excitation-contraction coupling in canine heart failure. Circulation 2001, 103, 1577–1584. [Google Scholar] [CrossRef]
  82. Tu, C.C.; Wan, B.Y.; Zeng, Y. STIM2 knockdown protects against ischemia/reperfusion injury through reducing mitochondrial calcium overload and preserving mitochondrial function. Life Sci. 2020, 247, 116560. [Google Scholar] [CrossRef]
  83. Santin, Y.; Fazal, L.; Sainte-Marie, Y.; Sicard, P.; Maggiorani, D.; Tortosa, F.; Yücel, Y.Y.; Teyssedre, L.; Rouquette, J.; Marcellin, M.; et al. Mitochondrial 4-HNE derived from MAO-A promotes mitoCa2+ overload in chronic postischemic cardiac remodeling. Cell Death Differ. 2020, 27, 1907–1923. [Google Scholar] [CrossRef]
  84. Stanciu, A.E. Cytokines in heart failure. Adv. Clin. Chem. 2019, 93, 63–113. [Google Scholar] [PubMed]
  85. Szabo, T.M.; Frigy, A.; Nagy, E.E. Targeting Mediators of Inflammation in Heart Failure: A Short Synthesis of Experimental and Clinical Results. Int. J. Mol. Sci. 2021, 22, 13053. [Google Scholar] [CrossRef]
  86. Wu, C.K.; Lee, J.K.; Chiang, F.T.; Yang, C.H.; Huang, S.W.; Hwang, J.J.; Lin, J.-L.; Tseng, C.-D.; Chen, J.-J.; Tsai, C.-T. Plasma levels of tumor necrosis factor-alpha and interleukin-6 are associated with diastolic heart failure through downregulation of sarcoplasmic reticulum Ca2+ ATPase. Crit. Care Med. 2011, 39, 984–992. [Google Scholar] [CrossRef]
  87. Tsai, C.T.; Wu, C.K.; Lee, J.K.; Chang, S.N.; Kuo, Y.M.; Wang, Y.C.; Lai, L.P.; Chiang, F.T.; Hwang, J.J.; Lin, J.L. TNF-alpha down-regulates sarcoplasmic reticulum Ca2+ ATPase expression and leads to left ventricular diastolic dysfunction through binding of NF-kappaB to promoter response element. Cardiovasc. Res. 2015, 105, 318–329. [Google Scholar] [CrossRef]
  88. Villegas, S.; Villarreal, F.J.; Dillmann, W.H. Leukemia Inhibitory Factor and Interleukin-6 downregulate sarcoplasmic reticulum Ca2+ ATPase (SERCA2) in cardiac myocytes. Basic Res. Cardiol. 2000, 95, 47–54. [Google Scholar] [CrossRef] [PubMed]
  89. Xue, G.L.; Li, D.S.; Wang, Z.Y.; Liu, Y.; Yang, J.M.; Li, C.Z.; Li, X.D.; Ma, M.J.; Zhang, M.M.; Liu, Y.J.; et al. Interleukin-17 upregulation participates in the pathogenesis of heart failure in mice via NF-kappaB-dependent suppression of SERCA2a and Cav1.2 expression. Acta Pharmacol. Sin. 2021, 42, 1780–1789. [Google Scholar] [CrossRef] [PubMed]
  90. Waning, D.L.; Mohammad, K.S.; Reiken, S.; Xie, W.; Andersson, D.C.; John, S.; Chiechi, A.; Wright, L.E.; Umanskaya, A.; Niewolna, M.; et al. Excess TGF-beta mediates muscle weakness associated with bone metastases in mice. Nat. Med. 2015, 21, 1262–1271. [Google Scholar] [CrossRef]
  91. Campanella, M.; Pinton, P.; Rizzuto, R. Mitochondrial Ca2+ homeostasis in health and disease. Biol. Res. 2004, 37, 653–660. [Google Scholar] [CrossRef]
  92. Brown, D.A.; O’Rourke, B. Cardiac mitochondria and arrhythmias. Cardiovasc. Res. 2010, 88, 241–249. [Google Scholar] [CrossRef]
  93. Crompton, M.; Heid, I. The cycling of calcium, sodium, and protons across the inner membrane of cardiac mitochondria. Eur. J. Biochem. 1978, 91, 599–608. [Google Scholar] [CrossRef]
  94. Palty, R.; Silverman, W.F.; Hershfinkel, M.; Caporale, T.; Sensi, S.L.; Parnis, J.; Nolte, C.; Fishman, D.; Shoshan-Barmatz, V.; Herrmann, S.; et al. NCLX is an essential component of mitochondrial Na+/Ca2+ exchange. Proc. Natl. Acad. Sci. USA 2010, 107, 436–441. [Google Scholar] [CrossRef] [PubMed]
  95. De Stefani, D.; Raffaello, A.; Teardo, E.; Szabo, I.; Rizzuto, R. A forty-kilodalton protein of the inner membrane is the mitochondrial calcium uniporter. Nature 2011, 476, 336–340. [Google Scholar] [CrossRef]
  96. Fan, M.; Zhang, J.; Tsai, C.W.; Orlando, B.J.; Rodriguez, M.; Xu, Y.; Liao, M.; Tsai, M.-F.; Feng, L. Structure and mechanism of the mitochondrial Ca2+ uniporter holocomplex. Nature 2020, 582, 129–133. [Google Scholar] [CrossRef]
  97. Raffaello, A.; De Stefani, D.; Sabbadin, D.; Teardo, E.; Merli, G.; Picard, A.; Checchetto, V.; Moro, S.; Szabò, I.; Rizzuto, R. The mitochondrial calcium uniporter is a multimer that can include a dominant-negative pore-forming subunit. EMBO J. 2013, 32, 2362–2376. [Google Scholar] [CrossRef]
  98. Baughman, J.M.; Perocchi, F.; Girgis, H.S.; Plovanich, M.; Belcher-Timme, C.A.; Sancak, Y.; Bao, X.R.; Strittmatter, L.; Goldberger, O.; Bogorad, R.L.; et al. Integrative genomics identifies MCU as an essential component of the mitochondrial calcium uniporter. Nature 2011, 476, 341–345. [Google Scholar] [CrossRef] [PubMed]
  99. Selwyn, M.J.; Dawson, A.P.; Dunnett, S.J. Calcium transport in mitochondria. FEBS Lett. 1970, 10, 1–5. [Google Scholar] [CrossRef]
  100. Vais, H.; Payne, R.; Paudel, U.; Li, C.; Foskett, J.K. Coupled transmembrane mechanisms control MCU-mediated mitochondrial Ca2+ uptake. Proc. Natl. Acad. Sci. USA 2020, 117, 21731–21739. [Google Scholar] [CrossRef] [PubMed]
  101. Boyman, L.; Lederer, W.J. How the mitochondrial calcium uniporter complex (MCUcx) works. Proc. Natl. Acad. Sci. USA 2020, 117, 22634–22636. [Google Scholar] [CrossRef] [PubMed]
  102. Fieni, F.; Lee, S.B.; Jan, Y.N.; Kirichok, Y. Activity of the mitochondrial calcium uniporter varies greatly between tissues. Nat. Commun. 2012, 3, 1317. [Google Scholar] [CrossRef]
  103. Pan, X.; Liu, J.; Nguyen, T.; Liu, C.; Sun, J.; Teng, Y.; Fergusson, M.M.; Rovira, I.I.; Allen, M.; Springer, D.A.; et al. The physiological role of mitochondrial calcium revealed by mice lacking the mitochondrial calcium uniporter. Nat. Cell Biol. 2013, 15, 1464–1472. [Google Scholar] [CrossRef]
  104. Holmstrom, K.M.; Pan, X.; Liu, J.C.; Menazza, S.; Liu, J.; Nguyen, T.T.; Pan, H.; Parks, R.J.; Anderson, S.; Noguchi, A.; et al. Assessment of cardiac function in mice lacking the mitochondrial calcium uniporter. J. Mol. Cell. Cardiol. 2015, 85, 178–182. [Google Scholar] [CrossRef]
  105. Dong, Z.; Shanmughapriya, S.; Tomar, D.; Siddiqui, N.; Lynch, S.; Nemani, N.; Breves, S.L.; Zhang, X.; Tripathi, A.; Palaniappan, P.; et al. Mitochondrial Ca2+ Uniporter Is a Mitochondrial Luminal Redox Sensor that Augments MCU Channel Activity. Mol. Cell 2017, 65, 1014–1028.e7. [Google Scholar] [CrossRef]
  106. Pan, L.; Huang, B.J.; Ma, X.E.; Wang, S.Y.; Feng, J.; Lv, F.; Liu, Y.; Liu, Y.; Li, C.-M.; Liang, D.-D.; et al. MiR-25 protects cardiomyocytes against oxidative damage by targeting the mitochondrial calcium uniporter. Int. J. Mol. Sci. 2015, 16, 5420–5433. [Google Scholar] [CrossRef]
  107. Kovacs-Bogdan, E.; Sancak, Y.; Kamer, K.J.; Plovanich, M.; Jambhekar, A.; Huber, R.J.; Myre, M.A.; Blower, M.D.; Mootha, V.K. Reconstitution of the mitochondrial calcium uniporter in yeast. Proc. Natl. Acad. Sci. USA 2014, 111, 8985–8990. [Google Scholar] [CrossRef] [PubMed]
  108. Oxenoid, K.; Dong, Y.; Cao, C.; Cui, T.; Sancak, Y.; Markhard, A.L.; Grabarek, Z.; Kong, L.; Liu, Z.; Ouyang, B.; et al. Architecture of the mitochondrial calcium uniporter. Nature 2016, 533, 269–273. [Google Scholar] [CrossRef]
  109. Zhuo, W.; Zhou, H.; Guo, R.; Yi, J.; Zhang, L.; Yu, L.; Sui, Y.; Zeng, W.; Wang, P.; Yang, M. Structure of intact human MCU supercomplex with the auxiliary MICU subunits. Protein Cell 2020, 12, 220–229. [Google Scholar] [CrossRef] [PubMed]
  110. Sancak, Y.; Markhard, A.L.; Kitami, T.; Kovacs-Bogdan, E.; Kamer, K.J.; Udeshi, N.D.; Carr, S.A.; Chaudhuri, D.; Clapham, D.E.; Li, A.A.; et al. EMRE is an essential component of the mitochondrial calcium uniporter complex. Science 2013, 342, 1379–1382. [Google Scholar] [CrossRef]
  111. Pittis, A.A.; Goh, V.; Cebrian-Serrano, A.; Wettmarshausen, J.; Perocchi, F.; Gabaldon, T. Discovery of EMRE in fungi resolves the true evolutionary history of the mitochondrial calcium uniporter. Nat. Commun. 2020, 11, 4031. [Google Scholar] [CrossRef] [PubMed]
  112. Liu, J.C.; Syder, N.C.; Ghorashi, N.S.; Willingham, T.B.; Parks, R.J.; Sun, J.; Fergusson, M.M.; Liu, J.; Holmström, K.M.; Menazza, S.; et al. EMRE is essential for mitochondrial calcium uniporter activity in a mouse model. JCI Insight 2020, 5, e134063. [Google Scholar] [CrossRef]
  113. Wang, Y.; Nguyen, N.X.; She, J.; Zeng, W.; Yang, Y.; Bai, X.C.; Jiang, Y. Structural Mechanism of EMRE-Dependent Gating of the Human Mitochondrial Calcium Uniporter. Cell 2019, 177, 1252–1261.e13. [Google Scholar] [CrossRef] [PubMed]
  114. Mallilankaraman, K.; Doonan, P.; Cardenas, C.; Chandramoorthy, H.C.; Muller, M.; Miller, R.; Hoffman, N.E.; Gandhirajan, R.K.; Molgo, J.; Birnbaum, M.J.; et al. MICU1 is an essential gatekeeper for MCU-mediated mitochondrial Ca2+ uptake that regulates cell survival. Cell 2012, 151, 630–644. [Google Scholar] [CrossRef] [PubMed]
  115. Nemani, N.; Dong, Z.; Daw, C.C.; Madaris, T.R.; Ramachandran, K.; Enslow, B.T.; Rubannelsonkumar, C.S.; Shanmughapriya, S.; Mallireddigari, V.; Maity, S.; et al. Mitochondrial pyruvate and fatty acid flux modulate MICU1-dependent control of MCU activity. Sci. Signal 2020, 13, eaaz6206. [Google Scholar] [CrossRef]
  116. Shah, S.I.; Ullah, G. The Function of Mitochondrial Calcium Uniporter at the Whole-Cell and Single Mitochondrion Levels in WT, MICU1 KO, and MICU2 KO Cells. Cells 2020, 9, 1520. [Google Scholar] [CrossRef]
  117. Gottschalk, B.; Klec, C.; Leitinger, G.; Bernhart, E.; Rost, R.; Bischof, H.; Madreiter-Sokolowski, C.T.; Radulović, S.; Eroglu, E.; Sattler, W.; et al. MICU1 controls cristae junction and spatially anchors mitochondrial Ca2+ uniporter complex. Nat. Commun. 2019, 10, 3732. [Google Scholar] [CrossRef] [PubMed]
  118. Phillips, C.B.; Tsai, C.W.; Tsai, M.F. The conserved aspartate ring of MCU mediates MICU1 binding and regulation in the mitochondrial calcium uniporter complex. eLife 2019, 8, e41112. [Google Scholar] [CrossRef] [PubMed]
  119. Marchi, S.; Corricelli, M.; Branchini, A.; Vitto, V.A.M.; Missiroli, S.; Morciano, G.; Perrone, M.; Ferrarese, M.; Giorgi, C.; Pinotti, M.; et al. Akt-mediated phosphorylation of MICU1 regulates mitochondrial Ca2+ levels and tumor growth. EMBO J. 2019, 38, e99435. [Google Scholar] [CrossRef]
  120. Paillard, M.; Csordas, G.; Huang, K.T.; Varnai, P.; Joseph, S.K.; Hajnoczky, G. MICU1 Interacts with the D-Ring of the MCU Pore to Control Its Ca2+ Flux and Sensitivity to Ru360. Mol. Cell 2018, 72, 778–785.e3. [Google Scholar] [CrossRef]
  121. Kamer, K.J.; Sancak, Y.; Fomina, Y.; Meisel, J.D.; Chaudhuri, D.; Grabarek, Z.; Mootha, V.K. MICU1 imparts the mitochondrial uniporter with the ability to discriminate between Ca2+ and Mn2+. Proc. Natl. Acad. Sci. USA 2018, 115, E7960–E7969. [Google Scholar] [CrossRef] [PubMed]
  122. Vecellio Reane, D.; Vallese, F.; Checchetto, V.; Acquasaliente, L.; Butera, G.; De Filippis, V.; Szabò, I.; Zanotti, G.; Rizzuto, R.; Raffaello, A. A MICU1 Splice Variant Confers High Sensitivity to the Mitochondrial Ca2+ Uptake Machinery of Skeletal Muscle. Mol. Cell 2016, 64, 760–773. [Google Scholar] [CrossRef] [PubMed]
  123. Logan, C.V.; Szabadkai, G.; Sharpe, J.A.; Parry, D.A.; Torelli, S.; Childs, A.M.; Kriek, M.; Phadke, R.; A Johnson, C.; Roberts, N.Y.; et al. Loss-of-function mutations in MICU1 cause a brain and muscle disorder linked to primary alterations in mitochondrial calcium signaling. Nat. Genet. 2014, 46, 188–193. [Google Scholar] [CrossRef]
  124. Alam, M.R.; Groschner, L.N.; Parichatikanond, W.; Kuo, L.; Bondarenko, A.I.; Rost, R.; Waldeck-Weiermair, M.; Malli, R.; Graier, G.; Pinotti, M.; et al. Mitochondrial Ca2+ uptake 1 (MICU1) and mitochondrial Ca2+ uniporter (MCU) contribute to metabolism-secretion coupling in clonal pancreatic beta-cells. J. Biol. Chem. 2012, 287, 34445–34454. [Google Scholar] [CrossRef] [PubMed]
  125. Patron, M.; Checchetto, V.; Raffaello, A.; Teardo, E.; Vecellio Reane, D.; Mantoan, M.; Granatiero, V.; Szabò, I.; De Stefani, D.; Rizzuto, R. MICU1 and MICU2 finely tune the mitochondrial Ca2+ uniporter by exerting opposite effects on MCU activity. Mol. Cell 2014, 53, 726–737. [Google Scholar] [CrossRef] [PubMed]
  126. Petrungaro, C.; Zimmermann, K.M.; Kuttner, V.; Fischer, M.; Dengjel, J.; Bogeski, I.; Riemer, J. The Ca2+-Dependent Release of the Mia40-Induced MICU1-MICU2 Dimer from MCU Regulates Mitochondrial Ca2+ Uptake. Cell Metab. 2015, 22, 721–733. [Google Scholar] [CrossRef]
  127. Plovanich, M.; Bogorad, R.L.; Sancak, Y.; Kamer, K.J.; Strittmatter, L.; Li, A.A.; Girgis, H.S.; Kuchimanchi, S.; De Groot, J.; Speciner, L.; et al. MICU2, a paralog of MICU1, resides within the mitochondrial uniporter complex to regulate calcium handling. PLoS ONE 2013, 8, e55785. [Google Scholar] [CrossRef]
  128. Wu, W.; Shen, Q.; Lei, Z.; Qiu, Z.; Li, D.; Pei, H.; Zheng, J.; Jia, Z. The crystal structure of MICU2 provides insight into Ca2+ binding and MICU1-MICU2 heterodimer formation. EMBO Rep. 2019, 20, e47488. [Google Scholar] [CrossRef] [PubMed]
  129. Patron, M.; Granatiero, V.; Espino, J.; Rizzuto, R.; De Stefani, D. MICU3 is a tissue-specific enhancer of mitochondrial calcium uptake. Cell Death Differ. 2019, 26, 179–195. [Google Scholar] [CrossRef]
  130. Wang, B.; Xiong, S.; Lin, S.; Xia, W.; Li, Q.; Zhao, Z.; Wei, X.; Lu, Z.; Wei, X.; Gao, P.; et al. Enhanced Mitochondrial Transient Receptor Potential Channel, Canonical Type 3-Mediated Calcium Handling in the Vasculature from Hypertensive Rats. J. Am. Heart Assoc. 2017, 6, e005812. [Google Scholar] [CrossRef]
  131. Puente, B.N.; Sun, J.; Parks, R.J.; Fergusson, M.M.; Liu, C.; Springer, D.; Aponte, A.M.; Liu, J.C.; Murphy, E. MICU3 Plays an Important Role in Cardiovascular Function. Circ. Res. 2020, 127, 1571–1573. [Google Scholar] [CrossRef]
  132. Ashrafi, G.; de Juan-Sanz, J.; Farrell, R.J.; Ryan, T.A. Molecular Tuning of the Axonal Mitochondrial Ca2+ Uniporter Ensures Metabolic Flexibility of Neurotransmission. Neuron 2020, 105, 678–687.e5. [Google Scholar] [CrossRef] [PubMed]
  133. Rasmussen, T.P.; Wu, Y.; Joiner, M.L.; Koval, O.M.; Wilson, N.R.; Luczak, E.D.; Wang, Q.; Chen, B.; Gao, Z.; Zhu, Z.; et al. Inhibition of MCU forces extramitochondrial adaptations governing physiological and pathological stress responses in heart. Proc. Natl. Acad. Sci. USA 2015, 112, 9129–9134. [Google Scholar] [CrossRef] [PubMed]
  134. Kwong, J.Q.; Lu, X.; Correll, R.N.; Schwanekamp, J.A.; Vagnozzi, R.J.; Sargent, M.A.; York, A.J.; Zhang, J.; Bers, D.M.; Molkentin, J.D. The Mitochondrial Calcium Uniporter Selectively Matches Metabolic Output to Acute Contractile Stress in the Heart. Cell Rep. 2015, 12, 15–22. [Google Scholar] [CrossRef] [PubMed]
  135. Luongo, T.S.; Lambert, J.P.; Yuan, A.; Zhang, X.; Gross, P.; Song, J.; Shanmughapriya, S.; Gao, E.; Jain, M.; Houser, S.R.; et al. The Mitochondrial Calcium Uniporter Matches Energetic Supply with Cardiac Workload during Stress and Modulates Permeability Transition. Cell Rep. 2015, 12, 23–34. [Google Scholar] [CrossRef]
  136. Altamimi, T.R.; Karwi, Q.G.; Uddin, G.M.; Fukushima, A.; Kwong, J.Q.; Molkentin, J.D.; Lopaschuk, G.D. Cardiac-specific deficiency of the mitochondrial calcium uniporter augments fatty acid oxidation and functional reserve. J. Mol. Cell. Cardiol. 2019, 127, 223–231. [Google Scholar] [CrossRef]
  137. Joseph, L.C.; Barca, E.; Subramanyam, P.; Komrowski, M.; Pajvani, U.; Colecraft, H.M.; Hirano, M.; Morrow, J.P. Inhibition of NAPDH Oxidase 2 (NOX2) Prevents Oxidative Stress and Mitochondrial Abnormalities Caused by Saturated Fat in Cardiomyocytes. PLoS ONE 2016, 11, e0145750. [Google Scholar] [CrossRef]
  138. Yu, Z.; Chen, R.; Li, M.; Yu, Y.; Liang, Y.; Han, F.; Qin, S.; Chen, X.; Su, Y.; Ge, J. Mitochondrial calcium uniporter inhibition provides cardioprotection in pressure overload-induced heart failure through autophagy enhancement. Int. J. Cardiol. 2018, 271, 161–168. [Google Scholar] [CrossRef]
  139. Zhang, B.; Jia, K.; Tian, J.; Du, H. Cyclophilin D counterbalances mitochondrial calcium uniporter-mediated brain mitochondrial calcium uptake. Biochem. Biophys. Res. Commun. 2020, 529, 314–320. [Google Scholar] [CrossRef]
  140. Bermont, F.; Hermant, A.; Benninga, R.; Chabert, C.; Jacot, G.; Santo-Domingo, J.; Kraus, M.R.-C.; Feige, J.N.; De Marchi, U. Targeting Mitochondrial Calcium Uptake with the Natural Flavonol Kaempferol, to Promote Metabolism/Secretion Coupling in Pancreatic beta-cells. Nutrients 2020, 12, 538. [Google Scholar] [CrossRef]
  141. Woods, J.J.; Nemani, N.; Shanmughapriya, S.; Kumar, A.; Zhang, M.; Nathan, S.R.; Thomas, M.; Carvalho, E.; Ramachandran, K.; Srikantan, S.; et al. A Selective and Cell-Permeable Mitochondrial Calcium Uniporter (MCU) Inhibitor Preserves Mitochondrial Bioenergetics after Hypoxia/Reoxygenation Injury. ACS Cent. Sci. 2019, 5, 153–166. [Google Scholar] [CrossRef]
  142. Hamilton, S.; Terentyeva, R.; Kim, T.Y.; Bronk, P.; Clements, R.T.; O-Uchi, J.; Csordás, G.; Choi, B.-R.; Terentyev, D. Pharmacological Modulation of Mitochondrial Ca2+ Content Regulates Sarcoplasmic Reticulum Ca2+ Release via Oxidation of the Ryanodine Receptor by Mitochondria-Derived Reactive Oxygen Species. Front. Physiol. 2018, 9, 1831. [Google Scholar] [CrossRef]
  143. Xu, H.; Guan, N.; Ren, Y.L.; Wei, Q.J.; Tao, Y.H.; Yang, G.S.; Liu, X.-Y.; Bu, D.-F.; Zhang, Y.; Zhu, S.-N. IP3R-Grp75-VDAC1-MCU calcium regulation axis antagonists protect podocytes from apoptosis and decrease proteinuria in an Adriamycin nephropathy rat model. BMC Nephrol. 2018, 19, 140. [Google Scholar] [CrossRef]
  144. Arduino, D.M.; Wettmarshausen, J.; Vais, H.; Navas-Navarro, P.; Cheng, Y.; Leimpek, A.; Ma, Z.; Delrio-Lorenzo, A.; Giordano, A.; Garcia-Perez, C.; et al. Systematic Identification of MCU Modulators by Orthogonal Interspecies Chemical Screening. Mol. Cell 2017, 67, 711–723.e7. [Google Scholar] [CrossRef]
  145. Kon, N.; Murakoshi, M.; Isobe, A.; Kagechika, K.; Miyoshi, N.; Nagayama, T. DS16570511 is a small-molecule inhibitor of the mitochondrial calcium uniporter. Cell Death Discov. 2017, 3, 17045. [Google Scholar] [CrossRef]
  146. Ren, T.; Zhang, H.; Wang, J.; Zhu, J.; Jin, M.; Wu, Y.; Guo, X.; Ji, L.; Huang, Q.; Zhang, H.; et al. MCU-dependent mitochondrial Ca2+ inhibits NAD(+)/SIRT3/SOD2 pathway to promote ROS production and metastasis of HCC cells. Oncogene 2017, 36, 5897–5909. [Google Scholar] [CrossRef]
  147. Pan, S.; Ryu, S.Y.; Sheu, S.S. Distinctive characteristics and functions of multiple mitochondrial Ca2+ influx mechanisms. Sci. China Life Sci. 2011, 54, 763–769. [Google Scholar] [CrossRef] [PubMed]
  148. Canato, M.; Capitanio, P.; Cancellara, L.; Leanza, L.; Raffaello, A.; Reane, D.V.; Marcucci, L.; Michelucci, A.; Protasi, F.; Reggiani, C. Excessive Accumulation of Ca2+ in Mitochondria of Y522S-RYR1 Knock-in Mice: A Link between Leak from the Sarcoplasmic Reticulum and Altered Redox State. Front. Physiol. 2019, 10, 1142. [Google Scholar] [CrossRef] [PubMed]
  149. Hamilton, J.; Brustovetsky, T.; Rysted, J.E.; Lin, Z.; Usachev, Y.M.; Brustovetsky, N. Deletion of mitochondrial calcium uniporter incompletely inhibits calcium uptake and induction of the permeability transition pore in brain mitochondria. J. Biol. Chem. 2018, 293, 15652–15663. [Google Scholar] [CrossRef]
  150. Natarajan, G.K.; Glait, L.; Mishra, J.; Stowe, D.F.; Camara, A.K.S.; Kwok, W.M. Total Matrix Ca2+ Modulates Ca2+ Efflux via the Ca2+/H+ Exchanger in Cardiac Mitochondria. Front. Physiol. 2020, 11, 510600. [Google Scholar] [CrossRef] [PubMed]
  151. Tsai, M.F.; Jiang, D.; Zhao, L.; Clapham, D.; Miller, C. Functional reconstitution of the mitochondrial Ca2+/H+ antiporter Letm1. J. Gen. Physiol. 2014, 143, 67–73. [Google Scholar] [CrossRef]
  152. Jiang, D.; Zhao, L.; Clapham, D.E. Genome-wide RNAi screen identifies Letm1 as a mitochondrial Ca2+/H+ antiporter. Science 2009, 326, 144–147. [Google Scholar] [CrossRef]
  153. Austin, S.; Nowikovsky, K. LETM1: Essential for Mitochondrial Biology and Cation Homeostasis? Trends Biochem Sci. 2019, 44, 648–658. [Google Scholar] [CrossRef] [PubMed]
  154. Durigon, R.; Mitchell, A.L.; Jones, A.W.; Manole, A.; Mennuni, M.; Hirst, E.M.; Houlden, H.; Maragni, G.; Lattante, S.; Doronzio, P.N.; et al. LETM1 couples mitochondrial DNA metabolism and nutrient preference. EMBO Mol. Med. 2018, 10, e8550. [Google Scholar] [CrossRef] [PubMed]
  155. Austin, S.; Tavakoli, M.; Pfeiffer, C.; Seifert, J.; Mattarei, A.; De Stefani, D.; Zoratti, M.; Nowikovsky, K. LETM1-Mediated K+ and Na+ Homeostasis Regulates Mitochondrial Ca2+ Efflux. Front. Physiol. 2017, 8, 839. [Google Scholar] [CrossRef] [PubMed]
  156. Huang, E.; Qu, D.; Huang, T.; Rizzi, N.; Boonying, W.; Krolak, D.; Ciana, P.; Woulfe, J.; Klein, C.; Slack, R.S.; et al. PINK1-mediated phosphorylation of LETM1 regulates mitochondrial calcium transport and protects neurons against mitochondrial stress. Nat. Commun. 2017, 8, 1399. [Google Scholar] [CrossRef]
  157. Feng, S.; Li, H.; Tai, Y.; Huang, J.; Su, Y.; Abramowitz, J.; Zhu, M.X.; Birnbaumer, L.; Wang, Y. Canonical transient receptor potential 3 channels regulate mitochondrial calcium uptake. Proc. Natl. Acad. Sci. USA 2013, 110, 11011–11016. [Google Scholar] [CrossRef]
  158. He, X.; Li, S.; Liu, B.; Susperreguy, S.; Formoso, K.; Yao, J.; Kang, J.; Shi, A.; Birnbaumer, L.; Liao, Y. Major contribution of the 3/6/7 class of TRPC channels to myocardial ischemia/reperfusion and cellular hypoxia/reoxygenation injuries. Proc. Natl. Acad. Sci. USA 2017, 114, E4582–E4591. [Google Scholar] [CrossRef] [PubMed]
  159. Khayyat, N.H.; Tomilin, V.N.; Zaika, O.; Pochynyuk, O. Polymodal roles of TRPC3 channel in the kidney. Channels 2020, 14, 257–267. [Google Scholar] [CrossRef]
  160. Han, J.W.; Kang, C.; Kim, Y.; Lee, M.G.; Kim, J.Y. Isoproterenol-induced hypertrophy of neonatal cardiac myocytes and H9c2 cell is dependent on TRPC3-regulated CaV1.2 expression. Cell Calcium 2020, 92, 102305. [Google Scholar] [CrossRef]
  161. Chen, X.; Sooch, G.; Demaree, I.S.; White, F.A.; Obukhov, A.G. Transient Receptor Potential Canonical (TRPC) Channels: Then and Now. Cells 2020, 9, 1983. [Google Scholar] [CrossRef]
  162. Nishiyama, K.; Tanaka, T.; Nishimura, A.; Nishida, M. TRPC3-based protein signaling complex as a therapeutic target of myocardial atrophy. Curr. Mol. Pharmacol. 2020, 14, 123–131. [Google Scholar] [CrossRef]
  163. Ma, T.; Lin, S.; Wang, B.; Wang, Q.; Xia, W.; Zhang, H.; Cui, Y.; He, C.; Wu, H.; Sun, F.; et al. TRPC3 deficiency attenuates high salt-induced cardiac hypertrophy by alleviating cardiac mitochondrial dysfunction. Biochem. Biophys. Res. Commun. 2019, 519, 674–681. [Google Scholar] [CrossRef] [PubMed]
  164. Kirichok, Y.; Krapivinsky, G.; Clapham, D.E. The mitochondrial calcium uniporter is a highly selective ion channel. Nature 2004, 427, 360–364. [Google Scholar] [CrossRef] [PubMed]
  165. Parks, R.J.; Menazza, S.; Holmstrom, K.M.; Amanakis, G.; Fergusson, M.; Ma, H.; Aponte, A.M.; Bernardi, P.; Finkel, T.; Murphy, E. Cyclophilin D-mediated regulation of the permeability transition pore is altered in mice lacking the mitochondrial calcium uniporter. Cardiovasc. Res. 2019, 115, 385–394. [Google Scholar] [CrossRef]
  166. Marta, K.; Hasan, P.; Rodriguez-Prados, M.; Paillard, M.; Hajnoczky, G. Pharmacological inhibition of the mitochondrial Ca2+ uniporter: Relevance for pathophysiology and human therapy. J. Mol. Cell. Cardiol. 2020, 151, 135–144. [Google Scholar] [CrossRef]
  167. Luongo, T.S.; Lambert, J.P.; Gross, P.; Nwokedi, M.; Lombardi, A.A.; Shanmughapriya, S.; Carpenter, A.C.; Kolmetzky, D.; Gao, E.; van Berlo, J.H.; et al. The mitochondrial Na+/Ca2+ exchanger is essential for Ca2+ homeostasis and viability. Nature 2017, 545, 93–97. [Google Scholar] [CrossRef]
  168. Garbincius, J.F.; Luongo, T.S.; Jadiya, P.; Hildebrand, A.N.; Kolmetzky, D.W.; Mangold, A.S.; Roy, R.; Ibetti, J.; Nwokedi, M.; Koch, W.J.; et al. Enhanced NCLX-dependent mitochondrial Ca2+ efflux attenuates pathological remodeling in heart failure. J. Mol. Cell. Cardiol. 2022, 167, 52–66. [Google Scholar] [CrossRef]
  169. Sparagna, G.C.; Gunter, K.K.; Sheu, S.S.; Gunter, T.E. Mitochondrial calcium uptake from physiological-type pulses of calcium. A description of the rapid uptake mode. J. Biol. Chem. 1995, 270, 27510–27515. [Google Scholar] [CrossRef] [PubMed]
  170. Santulli, G.; Marks, A.R. Essential Roles of Intracellular Calcium Release Channels in Muscle, Brain, Metabolism, and Aging. Curr. Mol. Pharmacol. 2015, 8, 206–222. [Google Scholar] [CrossRef]
  171. Cardenas, C.; Miller, R.A.; Smith, I.; Bui, T.; Molgo, J.; Muller, M.; Vais, H.; Cheung, K.-H.; Yang, J.; Parker, I.; et al. Essential regulation of cell bioenergetics by constitutive InsP3 receptor Ca2+ transfer to mitochondria. Cell 2010, 142, 270–283. [Google Scholar] [CrossRef]
  172. Rizzuto, R.; Simpson, A.W.; Brini, M.; Pozzan, T. Rapid changes of mitochondrial Ca2+ revealed by specifically targeted recombinant aequorin. Nature 1992, 358, 325–327. [Google Scholar] [CrossRef]
  173. Jou, M.J.; Peng, T.I.; Sheu, S.S. Histamine induces oscillations of mitochondrial free Ca2+ concentration in single cultured rat brain astrocytes. J. Physiol. 1996, 497 Pt 2, 299–308. [Google Scholar] [CrossRef]
  174. Llinas, R.; Sugimori, M.; Silver, R.B. Microdomains of high calcium concentration in a presynaptic terminal. Science 1992, 256, 677–679. [Google Scholar] [CrossRef]
  175. Rizzuto, R.; Brini, M.; Murgia, M.; Pozzan, T. Microdomains with high Ca2+ close to IP3-sensitive channels that are sensed by neighboring mitochondria. Science 1993, 262, 744–747. [Google Scholar] [CrossRef] [PubMed]
  176. Rizzuto, R.; Pinton, P.; Brini, M.; Chiesa, A.; Filippin, L.; Pozzan, T. Mitochondria as biosensors of calcium microdomains. Cell Calcium 1999, 26, 193–199. [Google Scholar] [CrossRef] [PubMed]
  177. Szalai, G.; Csordas, G.; Hantash, B.M.; Thomas, A.P.; Hajnoczky, G. Calcium signal transmission between ryanodine receptors and mitochondria. J. Biol. Chem. 2000, 275, 15305–15313. [Google Scholar] [CrossRef] [PubMed]
  178. Beaumont, V.; Llobet, A.; Lagnado, L. Expansion of calcium microdomains regulates fast exocytosis at a ribbon synapse. Proc. Natl. Acad. Sci. USA 2005, 102, 10700–10705. [Google Scholar] [CrossRef] [PubMed]
  179. Scriven, D.R.; Klimek, A.; Lee, K.L.; Moore, E.D. The molecular architecture of calcium microdomains in rat cardiomyocytes. Ann. N. Y. Acad. Sci. 2002, 976, 488–499. [Google Scholar] [CrossRef]
  180. Bautista, D.M.; Lewis, R.S. Modulation of plasma membrane calcium-ATPase activity by local calcium microdomains near CRAC channels in human T cells. J. Physiol. 2004, 556 Pt 3, 805–817. [Google Scholar] [CrossRef]
  181. Dedkova, E.N.; Blatter, L.A. Mitochondrial Ca2+ and the heart. Cell Calcium 2008, 44, 77–91. [Google Scholar] [CrossRef]
  182. Dedkova, E.N.; Seidlmayer, L.K.; Blatter, L.A. Mitochondria-mediated cardioprotection by trimetazidine in rabbit heart failure. J. Mol. Cell. Cardiol. 2013, 59, 41–54. [Google Scholar] [CrossRef]
  183. Moshkforoush, A.; Ashenagar, B.; Tsoukias, N.M.; Alevriadou, B.R. Modeling the role of endoplasmic reticulum-mitochondria microdomains in calcium dynamics. Sci. Rep. 2019, 9, 17072. [Google Scholar] [CrossRef] [PubMed]
  184. de Brito, O.M.; Scorrano, L. Mitofusin 2 tethers endoplasmic reticulum to mitochondria. Nature 2008, 456, 605–610. [Google Scholar] [CrossRef]
  185. Garcia-Perez, C.; Schneider, T.G.; Hajnoczky, G.; Csordas, G. Alignment of sarcoplasmic reticulum-mitochondrial junctions with mitochondrial contact points. Am. J. Physiol. Heart Circ. Physiol. 2011, 301, H1907–H1915. [Google Scholar] [CrossRef] [PubMed]
  186. Zacharioudakis, E.; Biris, N.; Garner, T.; Chen, Y.; Pekson, R.; Dhingra, R.; Santulli, G.; Kirshenbaum, L.A.; Kitsis, R.A.; Gavathiotis, E. Direct Small Molecule Activation of Mitofusins. BioRxiv 2018. [Google Scholar] [CrossRef]
  187. Paillard, M.; Tubbs, E.; Thiebaut, P.A.; Gomez, L.; Fauconnier, J.; Da Silva, C.C.; Teixeira, G.; Mewton, N.; Belaidi, E.; Durand, A.; et al. Depressing mitochondria-reticulum interactions protects cardiomyocytes from lethal hypoxia-reoxygenation injury. Circulation 2013, 128, 1555–1565. [Google Scholar] [CrossRef]
  188. Basso, V.; Marchesan, E.; Ziviani, E. A trio has turned into a quartet: DJ-1 interacts with the IP3R-Grp75-VDAC complex to control ER-mitochondria interaction. Cell Calcium 2020, 87, 102186. [Google Scholar] [CrossRef]
  189. Marchi, S.; Patergnani, S.; Pinton, P. The endoplasmic reticulum-mitochondria connection: One touch, multiple functions. Biochim. Biophys. Acta 2014, 1837, 461–469. [Google Scholar] [CrossRef]
  190. Giacomello, M.; Drago, I.; Bortolozzi, M.; Scorzeto, M.; Gianelle, A.; Pizzo, P.; Pozzan, T. Ca2+ hot spots on the mitochondrial surface are generated by Ca2+ mobilization from stores, but not by activation of store-operated Ca2+ channels. Mol. Cell 2010, 38, 280–290. [Google Scholar] [CrossRef]
  191. Rossini, M.; Filadi, R. Sarcoplasmic Reticulum-Mitochondria Kissing in Cardiomyocytes: Ca2+, ATP, and Undisclosed Secrets. Front. Cell Dev. Biol. 2020, 8, 532. [Google Scholar] [CrossRef]
  192. Chen, Y.; Csordas, G.; Jowdy, C.; Schneider, T.G.; Csordas, N.; Wang, W.; Liu, Y.; Kohlhaas, M.; Meiser, M.; Bergem, S.; et al. Mitofusin 2-containing mitochondrial-reticular microdomains direct rapid cardiomyocyte bioenergetic responses via interorganelle Ca2+ crosstalk. Circ. Res. 2012, 111, 863–875. [Google Scholar] [CrossRef] [PubMed]
  193. Liu, Y.; Zhu, X. Endoplasmic reticulum-mitochondria tethering in neurodegenerative diseases. Transl. Neurodegener. 2017, 6, 21. [Google Scholar] [CrossRef]
  194. Chen, H.; Detmer, S.A.; Ewald, A.J.; Griffin, E.E.; Fraser, S.E.; Chan, D.C. Mitofusins Mfn1 and Mfn2 coordinately regulate mitochondrial fusion and are essential for embryonic development. J. Cell Biol. 2003, 160, 189–200. [Google Scholar] [CrossRef]
  195. Sood, A.; Jeyaraju, D.V.; Prudent, J.; Caron, A.; Lemieux, P.; McBride, H.M.; Laplante, M.; Tóth, K.; Pellegrini, L. A Mitofusin-2-dependent inactivating cleavage of Opa1 links changes in mitochondria cristae and ER contacts in the postprandial liver. Proc. Natl. Acad. Sci. USA 2014, 111, 16017–16022. [Google Scholar] [CrossRef]
  196. Munoz, J.P.; Ivanova, S.; Sanchez-Wandelmer, J.; Martinez-Cristobal, P.; Noguera, E.; Sancho, A.; Díaz-Ramos, A.; Hernández-Alvarez, M.I.; Sebastián, D.; Mauvezin, C.; et al. Mfn2 modulates the UPR and mitochondrial function via repression of PERK. EMBO J. 2013, 32, 2348–2361. [Google Scholar] [CrossRef] [PubMed]
  197. Rizzuto, R.; Bastianutto, C.; Brini, M.; Murgia, M.; Pozzan, T. Mitochondrial Ca2+ homeostasis in intact cells. J. Cell Biol. 1994, 126, 1183–1194. [Google Scholar] [CrossRef]
  198. Rizzuto, R.; Pinton, P.; Carrington, W.; Fay, F.S.; Fogarty, K.E.; Lifshitz, L.M.; Tuft, R.A.; Pozzan, T. Close contacts with the endoplasmic reticulum as determinants of mitochondrial Ca2+ responses. Science 1998, 280, 1763–1766. [Google Scholar] [CrossRef] [PubMed]
  199. Cosson, P.; Marchetti, A.; Ravazzola, M.; Orci, L. Mitofusin-2 independent juxtaposition of endoplasmic reticulum and mitochondria: An ultrastructural study. PLoS ONE 2012, 7, e46293. [Google Scholar] [CrossRef]
  200. Filadi, R.; Greotti, E.; Turacchio, G.; Luini, A.; Pozzan, T.; Pizzo, P. Mitofusin 2 ablation increases endoplasmic reticulum-mitochondria coupling. Proc. Natl. Acad. Sci. USA 2015, 112, E2174–E2181. [Google Scholar] [CrossRef]
  201. Giorgi, C.; Ito, K.; Lin, H.K.; Santangelo, C.; Wieckowski, M.R.; Lebiedzinska, M.; Bononi, A.; Bonora, M.; Duszynski, J.; Bernardi, R.; et al. PML regulates apoptosis at endoplasmic reticulum by modulating calcium release. Science 2010, 330, 1247–1251. [Google Scholar] [CrossRef]
  202. Tubbs, E.; Rieusset, J. Study of Endoplasmic Reticulum and Mitochondria Interactions by In Situ Proximity Ligation Assay in Fixed Cells. J. Vis. Exp. 2016, 118, e54899. [Google Scholar]
  203. Santulli, G.; Nakashima, R.; Yuan, Q.; Marks, A.R. Intracellular calcium release channels: An update. J. Physiol. 2017, 595, 3041–3051. [Google Scholar] [CrossRef]
  204. Rizzuto, R.; De Stefani, D.; Raffaello, A.; Mammucari, C. Mitochondria as sensors and regulators of calcium signalling. Nat. Rev. Mol. Cell Biol. 2012, 13, 566–578. [Google Scholar] [CrossRef]
  205. Williams, G.S.; Boyman, L.; Lederer, W.J. Mitochondrial calcium and the regulation of metabolism in the heart. J. Mol. Cell. Cardiol. 2015, 78, 35–45. [Google Scholar] [CrossRef]
  206. Wust, R.C.; Helmes, M.; Martin, J.L.; van der Wardt, T.J.; Musters, R.J.; van der Velden, J.; Stienen, G.J.M. Rapid frequency-dependent changes in free mitochondrial calcium concentration in rat cardiac myocytes. J. Physiol. 2017, 595, 2001–2019. [Google Scholar] [CrossRef] [PubMed]
  207. Drago, I.; De Stefani, D.; Rizzuto, R.; Pozzan, T. Mitochondrial Ca2+ uptake contributes to buffering cytoplasmic Ca2+ peaks in cardiomyocytes. Proc. Natl. Acad. Sci. USA 2012, 109, 12986–12991. [Google Scholar] [CrossRef] [PubMed]
  208. Bernardi, P.; Di Lisa, F. The mitochondrial permeability transition pore: Molecular nature and role as a target in cardioprotection. J. Mol. Cell. Cardiol. 2015, 78, 100–106. [Google Scholar] [CrossRef]
  209. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol. Rev. 2014, 94, 909–950. [Google Scholar] [CrossRef]
  210. Hirabayashi, Y.; Kwon, S.K.; Paek, H.; Pernice, W.M.; Paul, M.A.; Lee, J.; Erfani, P.; Raczkowski, A.; Petrey, D.S.; Pon, L.A.; et al. ER-mitochondria tethering by PDZD8 regulates Ca2+ dynamics in mammalian neurons. Science 2017, 358, 623–630. [Google Scholar] [CrossRef] [PubMed]
  211. Wideman, J.G.; Balacco, D.L.; Fieblinger, T.; Richards, T.A. PDZD8 is not the ‘functional ortholog’ of Mmm1, it is a paralog. F1000Res 2018, 7, 1088. [Google Scholar] [CrossRef]
  212. Shirane, M.; Wada, M.; Morita, K.; Hayashi, N.; Kunimatsu, R.; Matsumoto, Y.; Matsuzaki, F.; Nakatsumi, H.; Ohta, K.; Tamura, Y.; et al. Protrudin and PDZD8 contribute to neuronal integrity by promoting lipid extraction required for endosome maturation. Nat. Commun. 2020, 11, 4576. [Google Scholar] [CrossRef]
  213. Elbaz-Alon, Y.; Guo, Y.; Segev, N.; Harel, M.; Quinnell, D.E.; Geiger, T.; Avinoam, O.; Li, D.; Nunnari, J. PDZD8 interacts with Protrudin and Rab7 at ER-late endosome membrane contact sites associated with mitochondria. Nat. Commun. 2020, 11, 3645. [Google Scholar] [CrossRef] [PubMed]
  214. Qiao, X.; Jia, S.; Ye, J.; Fang, X.; Zhang, C.; Cao, Y.; Xu, C.; Zhao, L.; Zhu, Y.; Wang, L.; et al. PTPIP51 regulates mouse cardiac ischemia/reperfusion through mediating the mitochondria-SR junction. Sci. Rep. 2017, 7, 45379. [Google Scholar] [CrossRef]
  215. De Vos, K.J.; Morotz, G.M.; Stoica, R.; Tudor, E.L.; Lau, K.F.; Ackerley, S.; Warley, A.; Shaw, C.E.; Miller, C.C.J. VAPB interacts with the mitochondrial protein PTPIP51 to regulate calcium homeostasis. Hum. Mol. Genet. 2012, 21, 1299–1311. [Google Scholar] [CrossRef]
  216. Stoica, R.; De Vos, K.J.; Paillusson, S.; Mueller, S.; Sancho, R.M.; Lau, K.F.; Vizcay-Barrena, G.; Lin, W.-L.; Xu, Y.-F.; Lewis, J.; et al. ER-mitochondria associations are regulated by the VAPB-PTPIP51 interaction and are disrupted by ALS/FTD-associated TDP-43. Nat. Commun. 2014, 5, 3996. [Google Scholar] [CrossRef] [PubMed]
  217. Spiteller, G. Are changes of the cell membrane structure causally involved in the aging process? Ann. N. Y. Acad. Sci. 2002, 959, 30–44. [Google Scholar] [CrossRef] [PubMed]
  218. Cadenas, E.; Davies, K.J. Mitochondrial free radical generation, oxidative stress, and aging. Free Radic. Biol. Med. 2000, 29, 222–230. [Google Scholar] [CrossRef]
  219. Ozawa, T. Mitochondrial genome mutation in cell death and aging. J. Bioenergy Biomembr. 1999, 31, 377–390. [Google Scholar] [CrossRef]
  220. Bode, D.; Wen, Y.; Hegemann, N.; Primessnig, U.; Parwani, A.; Boldt, L.H.; Pieske, B.; Heinzel, F.R.; Hohendanner, F. Oxidative Stress and Inflammatory Modulation of Ca2+ Handling in Metabolic HFpEF-Related Left Atrial Cardiomyopathy. Antioxidants 2020, 9, 860. [Google Scholar] [CrossRef]
  221. Chang, J.C.; Lien, C.F.; Lee, W.S.; Chang, H.R.; Hsu, Y.C.; Luo, Y.P.; Jeng, J.-R.; Hsieh, J.-C.; Yang, K.-T. Intermittent Hypoxia Prevents Myocardial Mitochondrial Ca2+ Overload and Cell Death during Ischemia/Reperfusion: The Role of Reactive Oxygen Species. Cells 2019, 8, 564. [Google Scholar] [CrossRef]
  222. Mitchell, P. Coupling of phosphorylation to electron and hydrogen transfer by a chemi-osmotic type of mechanism. Nature 1961, 191, 144–148. [Google Scholar] [CrossRef] [PubMed]
  223. Chance, B.; Sies, H.; Boveris, A. Hydroperoxide metabolism in mammalian organs. Physiol. Rev. 1979, 59, 527–605. [Google Scholar] [CrossRef]
  224. Aon, M.A.; Cortassa, S.; Marban, E.; O’Rourke, B. Synchronized whole cell oscillations in mitochondrial metabolism triggered by a local release of reactive oxygen species in cardiac myocytes. J. Biol. Chem. 2003, 278, 44735–44744. [Google Scholar] [CrossRef]
  225. Brand, M.D. The sites and topology of mitochondrial superoxide production. Exp. Gerontol. 2010, 45, 466–472. [Google Scholar] [CrossRef] [PubMed]
  226. Sakamoto, T.; Imai, H. Hydrogen peroxide produced by superoxide dismutase SOD-2 activates sperm in Caenorhabditis elegans. J. Biol. Chem. 2017, 292, 14804–14813. [Google Scholar] [CrossRef] [PubMed]
  227. Kang, P.T.; Chen, C.L.; Ohanyan, V.; Luther, D.J.; Meszaros, J.G.; Chilian, W.M.; Chen, Y.-R. Overexpressing superoxide dismutase 2 induces a supernormal cardiac function by enhancing redox-dependent mitochondrial function and metabolic dilation. J. Mol. Cell. Cardiol. 2015, 88, 14–28. [Google Scholar] [CrossRef] [PubMed]
  228. Ribas, V.; Garcia-Ruiz, C.; Fernandez-Checa, J.C. Glutathione and mitochondria. Front. Pharmacol. 2014, 5, 151. [Google Scholar] [CrossRef] [PubMed]
  229. Lopert, P.; Patel, M. Brain mitochondria from DJ-1 knockout mice show increased respiration-dependent hydrogen peroxide consumption. Redox. Biol. 2014, 2, 667–672. [Google Scholar] [CrossRef]
  230. Ide, T.; Tsutsui, H.; Kinugawa, S.; Utsumi, H.; Kang, D.; Hattori, N.; Uchida, K.; Arimura, K.-I.; Egashira, K.; Takeshita, A. Mitochondrial electron transport complex I is a potential source of oxygen free radicals in the failing myocardium. Circ. Res. 1999, 85, 357–363. [Google Scholar] [CrossRef]
  231. Ide, T.; Tsutsui, H.; Kinugawa, S.; Suematsu, N.; Hayashidani, S.; Ichikawa, K.; Utsumi, H.; Machida, Y.; Egashira, K.; Takeshita, A. Direct evidence for increased hydroxyl radicals originating from superoxide in the failing myocardium. Circ. Res. 2000, 86, 152–157. [Google Scholar] [CrossRef]
  232. Wagner, S.; Rokita, A.G.; Anderson, M.E.; Maier, L.S. Redox regulation of sodium and calcium handling. Antioxid. Redox. Signal 2013, 18, 1063–1077. [Google Scholar] [CrossRef] [PubMed]
  233. Kiyuna, L.A.; Albuquerque, R.P.E.; Chen, C.H.; Mochly-Rosen, D.; Ferreira, J.C.B. Targeting mitochondrial dysfunction and oxidative stress in heart failure: Challenges and opportunities. Free Radic. Biol. Med. 2018, 129, 155–168. [Google Scholar] [CrossRef] [PubMed]
  234. Palaniyandi, S.S.; Qi, X.; Yogalingam, G.; Ferreira, J.C.; Mochly-Rosen, D. Regulation of mitochondrial processes: A target for heart failure. Drug Discov. Today Dis. Mech. 2010, 7, e95–e102. [Google Scholar] [CrossRef]
  235. Mazurek, S.R.; Bovo, E.; Zima, A.V. Regulation of sarcoplasmic reticulum Ca2+ release by cytosolic glutathione in rabbit ventricular myocytes. Free Radic. Biol. Med. 2014, 68, 159–167. [Google Scholar] [CrossRef] [PubMed]
  236. Shan, J.; Betzenhauser, M.J.; Kushnir, A.; Reiken, S.; Meli, A.C.; Wronska, A.; Dura, M.; Chen, B.; Marks, A.R. Role of chronic ryanodine receptor phosphorylation in heart failure and beta-adrenergic receptor blockade in mice. J. Clin. Investig. 2010, 120, 4375–4387. [Google Scholar] [CrossRef] [PubMed]
  237. Goonasekera, S.A.; Hammer, K.; Auger-Messier, M.; Bodi, I.; Chen, X.; Zhang, H.; Reiken, S.; Elrod, J.W.; Correll, R.N.; York, A.J.; et al. Decreased cardiac L-type Ca2+ channel activity induces hypertrophy and heart failure in mice. J. Clin. Investig. 2012, 122, 280–290. [Google Scholar] [CrossRef]
  238. Zalk, R.; Clarke, O.B.; des Georges, A.; Grassucci, R.A.; Reiken, S.; Mancia, F.; Hendrickson, W.A.; Frank, J.; Marks, A.R. Structure of a mammalian ryanodine receptor. Nature 2015, 517, 44–49. [Google Scholar] [CrossRef]
  239. Efremov, R.G.; Leitner, A.; Aebersold, R.; Raunser, S. Architecture and conformational switch mechanism of the ryanodine receptor. Nature 2015, 517, 39–43. [Google Scholar] [CrossRef]
  240. Yan, Z.; Bai, X.; Yan, C.; Wu, J.; Li, Z.; Xie, T.; Peng, W.; Yin, C.-C.; Li, X.; Scheres, S.H.W.; et al. Structure of the rabbit ryanodine receptor RyR1 at near-atomic resolution. Nature 2015, 517, 50–55. [Google Scholar] [CrossRef]
  241. des Georges, A.; Clarke, O.B.; Zalk, R.; Yuan, Q.; Condon, K.J.; Grassucci, R.A.; Hendrickson, W.A.; Marks, A.R.; Frank, J. Structural Basis for Gating and Activation of RyR1. Cell 2016, 167, 145–157.e17. [Google Scholar] [CrossRef]
  242. Voss, A.A.; Lango, J.; Ernst-Russell, M.; Morin, D.; Pessah, I.N. Identification of hyperreactive cysteines within ryanodine receptor type 1 by mass spectrometry. J. Biol. Chem. 2004, 279, 34514–34520. [Google Scholar] [CrossRef]
  243. Aracena-Parks, P.; Goonasekera, S.A.; Gilman, C.P.; Dirksen, R.T.; Hidalgo, C.; Hamilton, S.L. Identification of cysteines involved in S-nitrosylation, S-glutathionylation, and oxidation to disulfides in ryanodine receptor type 1. J. Biol. Chem. 2006, 281, 40354–40368. [Google Scholar] [CrossRef]
  244. Mi, T.; Xiao, Z.; Guo, W.; Tang, Y.; Hiess, F.; Xiao, J.; Wang, Y.; Zhang, J.Z.; Zhang, L.; Wang, R.; et al. Role of Cys3602 in the function and regulation of the cardiac ryanodine receptor. Biochem. J. 2015, 467, 177–190. [Google Scholar] [CrossRef] [PubMed]
  245. Nikolaienko, R.; Bovo, E.; Rebbeck, R.T.; Kahn, D.; Thomas, D.D.; Cornea, R.L.; Zima, A.V. The functional significance of redox-mediated intersubunit cross-linking in regulation of human type 2 ryanodine receptor. Redox. Biol. 2020, 37, 101729. [Google Scholar] [CrossRef]
  246. Melville, Z.; Dridi, H.; Yuan, Q.; Reiken, S.; Wronska, A.; Liu, Y.; Clarke, O.; Marks, A. A drug and ATP binding site in type 1 ryanodine receptor. Structure 2022, 30, 1025–1034.e4. [Google Scholar] [CrossRef]
  247. Miotto, M.C.; Weninger, G.; Dridi, H.; Yuan, Q.; Liu, Y.; Wronska, A.; Melville, Z.; Sittenfeld, L.; Reiken, S.; Marks, A.R. Structural analyses of human ryanodine receptor type 2 channels reveal the mechanisms for sudden cardiac death and treatment. Sci. Adv. 2022, 8, eabo1272. [Google Scholar] [CrossRef]
  248. Priori, S.G.; Chen, S.R. Inherited dysfunction of sarcoplasmic reticulum Ca2+ handling and arrhythmogenesis. Circ. Res. 2011, 108, 871–883. [Google Scholar] [CrossRef] [PubMed]
  249. Iyer, K.A.; Hu, Y.; Nayak, A.R.; Kurebayashi, N.; Murayama, T.; Samso, M. Structural mechanism of two gain-of-function cardiac and skeletal RyR mutations at an equivalent site by cryo-EM. Sci. Adv. 2020, 6, eabb2964. [Google Scholar] [CrossRef] [PubMed]
  250. Zhong, X.; Liu, Y.; Zhu, L.; Meng, X.; Wang, R.; Van Petegem, F.; Wagenknecht, T.; Chen, S.W.; Liu, Z. Conformational dynamics inside amino-terminal disease hotspot of ryanodine receptor. Structure 2013, 21, 2051–2060. [Google Scholar] [CrossRef] [PubMed]
  251. Lamb, G.D.; Posterino, G.S.; Yamamoto, T.; Ikemoto, N. Effects of a domain peptide of the ryanodine receptor on Ca2+ release in skinned skeletal muscle fibers. Am. J. Physiol. Cell Physiol. 2001, 281, C207–C214. [Google Scholar] [CrossRef]
  252. Bannister, M.L.; Hamada, T.; Murayama, T.; Harvey, P.J.; Casarotto, M.G.; Dulhunty, A.F.; Ikemoto, N. Malignant hyperthermia mutation sites in the Leu2442-Pro2477 (DP4) region of RyR1 (ryanodine receptor 1) are clustered in a structurally and functionally definable area. Biochem. J. 2007, 401, 333–339. [Google Scholar] [CrossRef]
  253. Nikolaienko, R.; Bovo, E.; Kahn, D.; Gracia, R.; Jamrozik, T.; Zima, A.V. Cysteines 1078 and 2991 cross-linking plays a critical role in redox regulation of cardiac ryanodine receptor (RyR). Nat. Commun. 2023, 14, 4498. [Google Scholar] [CrossRef] [PubMed]
  254. Bovo, E.; Mazurek, S.R.; Zima, A.V. The role of RyR2 oxidation in the blunted frequency-dependent facilitation of Ca2+ transient amplitude in rabbit failing myocytes. Pflugers Arch. 2018, 470, 959–968. [Google Scholar] [CrossRef]
  255. Walker, M.A.; Williams, G.S.B.; Kohl, T.; Lehnart, S.E.; Jafri, M.S.; Greenstein, J.L.; Lederer, W.; Winslow, R.L. Superresolution modeling of calcium release in the heart. Biophys. J. 2014, 107, 3018–3029. [Google Scholar] [CrossRef] [PubMed]
  256. Cabra, V.; Murayama, T.; Samso, M. Ultrastructural Analysis of Self-Associated RyR2s. Biophys. J. 2016, 110, 2651–2662. [Google Scholar] [CrossRef] [PubMed]
  257. Marx, S.O.; Ondrias, K.; Marks, A.R. Coupled gating between individual skeletal muscle Ca2+ release channels (ryanodine receptors). Science 1998, 281, 818–821. [Google Scholar] [CrossRef]
  258. Marx, S.O.; Gaburjakova, J.; Gaburjakova, M.; Henrikson, C.; Ondrias, K.; Marks, A.R. Coupled gating between cardiac calcium release channels (ryanodine receptors). Circ. Res. 2001, 88, 1151–1158. [Google Scholar] [CrossRef]
  259. Currie, S. Cardiac ryanodine receptor phosphorylation by CaM Kinase II: Keeping the balance right. Front. Biosci. 2009, 14, 5134–5156. [Google Scholar] [CrossRef]
  260. Campbell, H.M.; Quick, A.P.; Abu-Taha, I.; Chiang, D.Y.; Kramm, C.F.; Word, T.A.; Brandenburg, S.; Hulsurkar, M.; Alsina, K.M.; Liu, H.-B.; et al. Loss of SPEG Inhibitory Phosphorylation of Ryanodine Receptor Type-2 Promotes Atrial Fibrillation. Circulation 2020, 142, 1159–1172. [Google Scholar] [CrossRef]
  261. Eiringhaus, J.; Herting, J.; Schatter, F.; Nikolaev, V.O.; Sprenger, J.; Wang, Y.; Köhn, M.; Zabel, M.; El-Armouche, A.; Hasenfuss, G.; et al. Protein kinase/phosphatase balance mediates the effects of increased late sodium current on ventricular calcium cycling. Basic Res. Cardiol. 2019, 114, 13. [Google Scholar] [CrossRef]
  262. Llach, A.; Mazevet, M.; Mateo, P.; Villejouvert, O.; Ridoux, A.; Rucker-Martin, C.; Ribeiro, M.; Fischmeister, R.; Crozatier, B.; Benitah, J.-P.; et al. Progression of excitation-contraction coupling defects in doxorubicin cardiotoxicity. J. Mol. Cell Cardiol. 2019, 126, 129–139. [Google Scholar] [CrossRef]
  263. Dries, E.; Santiago, D.J.; Johnson, D.M.; Gilbert, G.; Holemans, P.; Korte, S.M.; Roderick, H.L.; Sipido, K.R. Calcium/calmodulin-dependent kinase II and nitric oxide synthase 1-dependent modulation of ryanodine receptors during beta-adrenergic stimulation is restricted to the dyadic cleft. J. Physiol. 2016, 594, 5923–5939. [Google Scholar] [CrossRef] [PubMed]
  264. Greiser, M.; Kerfant, B.G.; Williams, G.S.; Voigt, N.; Harks, E.; Dibb, K.M.; Giese, A.; Meszaros, J.; Verheule, S.; Ravens, U.; et al. Tachycardia-induced silencing of subcellular Ca2+ signaling in atrial myocytes. J. Clin. Investig. 2014, 124, 4759–4772. [Google Scholar] [CrossRef]
  265. Di Carlo, M.N.; Said, M.; Ling, H.; Valverde, C.A.; De Giusti, V.C.; Sommese, L.; Palomeque, J.; Aiello, E.A.; Skapura, D.G.; Rinaldi, G.; et al. CaMKII-dependent phosphorylation of cardiac ryanodine receptors regulates cell death in cardiac ischemia/reperfusion injury. J. Mol. Cell. Cardiol. 2014, 74, 274–283. [Google Scholar] [CrossRef]
  266. Gambardella, J.; Sorriento, D.; Ciccarelli, M.; Del Giudice, C.; Fiordelisi, A.; Napolitano, L.; Trimarco, B.; Iaccarino, G.; Santulli, G. Functional Role of Mitochondria in Arrhythmogenesis. Adv. Exp. Med. Biol. 2017, 982, 191–202. [Google Scholar] [PubMed]
  267. Kushnir, A.; Santulli, G.; Reiken, S.R.; Coromilas, E.; Godfrey, S.J.; Brunjes, D.L.; Colombo, P.C.; Yuzefpolskaya, M.; Sokol, S.I.; Kitsis, R.N.; et al. Ryanodine Receptor Calcium Leak in Circulating B-Lymphocytes as a Biomarker in Heart Failure. Circulation 2018, 138, 1144–1154. [Google Scholar] [CrossRef]
  268. Yuan, Q.; Chen, Z.; Santulli, G.; Gu, L.; Yang, Z.G.; Yuan, Z.Q.; Zhao, Y.-T.; Xin, H.-B.; Deng, K.-Y.; Wang, S.-Q.; et al. Functional role of Calstabin2 in age-related cardiac alterations. Sci. Rep. 2014, 4, 7425. [Google Scholar] [CrossRef]
  269. Xie, W.; Santulli, G.; Reiken, S.R.; Yuan, Q.; Osborne, B.W.; Chen, B.X.; Marks, A.R. Mitochondrial oxidative stress promotes atrial fibrillation. Sci. Rep. 2015, 5, 11427. [Google Scholar] [CrossRef]
  270. Brennan, J.P.; Bardswell, S.C.; Burgoyne, J.R.; Fuller, W.; Schroder, E.; Wait, R.; Begum, S.; Kentish, J.C.; Eaton, P. Oxidant-induced activation of type I protein kinase A is mediated by RI subunit interprotein disulfide bond formation. J. Biol. Chem. 2006, 281, 21827–21836. [Google Scholar] [CrossRef] [PubMed]
  271. Erickson, J.R. Mechanisms of CaMKII Activation in the Heart. Front. Pharmacol. 2014, 5, 59. [Google Scholar] [CrossRef]
  272. Erickson, J.R.; Joiner, M.L.; Guan, X.; Kutschke, W.; Yang, J.; Oddis, C.V.; Bartlett, R.K.; Lowe, J.S.; O’Donnell, S.E.; Aykin-Burns, N.; et al. A dynamic pathway for calcium-independent activation of CaMKII by methionine oxidation. Cell 2008, 133, 462–474. [Google Scholar] [CrossRef]
  273. Wu, Y.; Wang, Q.; Feng, N.; Granger, J.M.; Anderson, M.E. Myocardial death and dysfunction after ischemia-reperfusion injury require CaMKIIdelta oxidation. Sci. Rep. 2019, 9, 9291. [Google Scholar] [CrossRef]
  274. Joiner, M.L.; Koval, O.M.; Li, J.; He, B.J.; Allamargot, C.; Gao, Z.; Luczak, E.D.; Hall, D.D.; Fink, B.D.; Chen, B.; et al. CaMKII determines mitochondrial stress responses in heart. Nature 2012, 491, 269–273. [Google Scholar] [CrossRef] [PubMed]
  275. Fieni, F.; Johnson, D.E.; Hudmon, A.; Kirichok, Y. Mitochondrial Ca2+ uniporter and CaMKII in heart. Nature 2014, 513, E1–E2. [Google Scholar] [CrossRef]
  276. Nickel, A.G.; Kohlhaas, M.; Bertero, E.; Wilhelm, D.; Wagner, M.; Sequeira, V.; Kreusser, M.M.; Dewenter, M.; Kappl, R.; Hoth, M.; et al. CaMKII does not control mitochondrial Ca2+ uptake in cardiac myocytes. J. Physiol. 2019, 598, 1361–1376. [Google Scholar] [CrossRef]
  277. Hulot, J.S.; Salem, J.E.; Redheuil, A.; Collet, J.P.; Varnous, S.; Jourdain, P.; Logeart, D.; Gandjbakhch, E.; Bernard, C.; Hatem, S.N.; et al. Effect of intracoronary administration of AAV1/SERCA2a on ventricular remodelling in patients with advanced systolic heart failure: Results from the AGENT-HF randomized phase 2 trial. Eur. J. Heart Fail. 2017, 19, 1534–1541. [Google Scholar] [CrossRef]
  278. Escobales, N.; Nunez, R.E.; Jang, S.; Parodi-Rullan, R.; Ayala-Pena, S.; Sacher, J.R.; Skoda, E.M.; Wipf, P.; Frontera, W.; Javadov, S. Mitochondria-targeted ROS scavenger improves post-ischemic recovery of cardiac function and attenuates mitochondrial abnormalities in aged rats. J. Mol. Cell. Cardiol. 2014, 77, 136–146. [Google Scholar] [CrossRef]
  279. Fink, M.P.; Macias, C.A.; Xiao, J.; Tyurina, Y.Y.; Delude, R.L.; Greenberger, J.S.; Kagan, V.E.; Wipf, P. Hemigramicidin-TEMPO conjugates: Novel mitochondria-targeted antioxidants. Crit. Care Med. 2007, 35 (Suppl. S9), S461–S467. [Google Scholar] [CrossRef] [PubMed]
  280. Bernardi, P.; Vassanelli, S.; Veronese, P.; Colonna, R.; Szabo, I.; Zoratti, M. Modulation of the mitochondrial permeability transition pore. Effect of protons and divalent cations. J. Biol. Chem. 1992, 267, 2934–2939. [Google Scholar] [CrossRef] [PubMed]
  281. Giorgio, V.; Burchell, V.; Schiavone, M.; Bassot, C.; Minervini, G.; Petronilli, V.; Argenton, F.; Forte, M.; Tosatto, S.; Lippe, G.; et al. Ca2+ binding to F-ATP synthase beta subunit triggers the mitochondrial permeability transition. EMBO Rep. 2017, 18, 1065–1076. [Google Scholar] [CrossRef]
  282. Halestrap, A.P. The mitochondrial permeability transition: Its molecular mechanism and role in reperfusion injury. Biochem. Soc Symp. 1999, 66, 181–203. [Google Scholar]
  283. Crompton, M. The mitochondrial permeability transition pore and its role in cell death. Biochem. J. 1999, 341 Pt 2, 233–249. [Google Scholar] [CrossRef] [PubMed]
  284. Scarabelli, T.M.; Stephanou, A.; Pasini, E.; Gitti, G.; Townsend, P.; Lawrence, K.; Chen-Scarabelli, C.; Saravolatz, L.; Latchman, D.; Knight, R.; et al. Minocycline inhibits caspase activation and reactivation, increases the ratio of XIAP to smac/DIABLO, and reduces the mitochondrial leakage of cytochrome C and smac/DIABLO. J. Am. Coll. Cardiol. 2004, 43, 865–874. [Google Scholar] [CrossRef]
  285. Li, J.; Li, Y.; Qin, D.; von Harsdorf, R.; Li, P. Mitochondrial fission leads to Smac/DIABLO release quenched by ARC. Apoptosis 2010, 15, 1187–1196. [Google Scholar] [CrossRef] [PubMed]
  286. Brown, D.A.; Perry, J.B.; Allen, M.E.; Sabbah, H.N.; Stauffer, B.L.; Shaikh, S.R.; Cleland, J.G.; Colucci, W.S.; Butler, J.; Voors, A.A.; et al. Expert consensus document: Mitochondrial function as a therapeutic target in heart failure. Nat. Rev. Cardiol. 2017, 14, 238–250. [Google Scholar] [CrossRef] [PubMed]
  287. Smith, R.A.; Hartley, R.C.; Cocheme, H.M.; Murphy, M.P. Mitochondrial pharmacology. Trends Pharmacol. Sci. 2012, 33, 341–352. [Google Scholar] [CrossRef] [PubMed]
  288. Kloner, R.A.; Brown, D.A.; Csete, M.; Dai, W.; Downey, J.M.; Gottlieb, R.A.; Hale, S.L.; Shi, J. New and revisited approaches to preserving the reperfused myocardium. Nat. Rev. Cardiol. 2017, 14, 679–693. [Google Scholar] [CrossRef] [PubMed]
  289. Fu, J.D.; Li, J.; Tweedie, D.; Yu, H.M.; Chen, L.; Wang, R.; Riordon, D.R.; Brugh, S.A.; Wang, S.; Boheler, K.R.; et al. Crucial role of the sarcoplasmic reticulum in the developmental regulation of Ca2+ transients and contraction in cardiomyocytes derived from embryonic stem cells. FASEB J. 2006, 20, 181–183. [Google Scholar] [CrossRef]
  290. Sharma, V.K.; Ramesh, V.; Franzini-Armstrong, C.; Sheu, S.S. Transport of Ca2+ from sarcoplasmic reticulum to mitochondria in rat ventricular myocytes. J. Bioenergy Biomembr. 2000, 32, 97–104. [Google Scholar] [CrossRef]
  291. Rimessi, A.; Pozzato, C.; Carparelli, L.; Rossi, A.; Ranucci, S.; De Fino, I.; Cigana, C.; Talarico, A.; Wieckowski, M.R.; Ribeiro, C.M.P.; et al. Pharmacological modulation of mitochondrial calcium uniporter controls lung inflammation in cystic fibrosis. Sci. Adv. 2020, 6, eaax9093. [Google Scholar] [CrossRef]
  292. Yano, M.; Kobayashi, S.; Kohno, M.; Doi, M.; Tokuhisa, T.; Okuda, S.; Suetsugu, M.; Hisaoka, T.; Obayashi, M.; Ohkusa, T.; et al. FKBP12.6-mediated stabilization of calcium-release channel (ryanodine receptor) as a novel therapeutic strategy against heart failure. Circulation 2003, 107, 477–484. [Google Scholar] [CrossRef] [PubMed]
  293. Wehrens, X.H.; Lehnart, S.E.; Reiken, S.; van der Nagel, R.; Morales, R.; Sun, J.; Cheng, Z.; Deng, S.-X.; de Windt, L.J.; Landry, D.W.; et al. Enhancing calstabin binding to ryanodine receptors improves cardiac and skeletal muscle function in heart failure. Proc. Natl. Acad. Sci. USA 2005, 102, 9607–9612. [Google Scholar] [CrossRef] [PubMed]
  294. Dai, D.F.; Chen, T.; Szeto, H.; Nieves-Cintron, M.; Kutyavin, V.; Santana, L.F.; Rabinovitch, P.S. Mitochondrial targeted antioxidant Peptide ameliorates hypertensive cardiomyopathy. J. Am. Coll. Cardiol. 2011, 58, 73–82. [Google Scholar] [CrossRef] [PubMed]
  295. Rohani, L.; Machiraju, P.; Sabouny, R.; Meng, G.; Liu, S.; Zhao, T.; Iqbal, F.; Wang, X.; Ravandi, A.; Wu, J.C.; et al. Reversible Mitochondrial Fragmentation in iPSC-Derived Cardiomyocytes from Children with DCMA, a Mitochondrial Cardiomyopathy. Can. J. Cardiol. 2020, 36, 554–563. [Google Scholar] [CrossRef]
  296. Dai, D.F.; Hsieh, E.J.; Chen, T.; Menendez, L.G.; Basisty, N.B.; Tsai, L.; Beyer, R.P.; Crispin, D.A.; Shulman, N.J.; Szeto, H.H.; et al. Global proteomics and pathway analysis of pressure-overload-induced heart failure and its attenuation by mitochondrial-targeted peptides. Circ. Heart Fail. 2013, 6, 1067–1076. [Google Scholar] [CrossRef]
  297. Ghionzoli, N.; Gentile, F.; Del Franco, A.M.; Castiglione, V.; Aimo, A.; Giannoni, A.; Burchielli, S.; Cameli, M.; Emdin, M.; Vergaro, G. Current and emerging drug targets in heart failure treatment. Heart Fail. Rev. 2022, 27, 1119–1136. [Google Scholar] [CrossRef]
  298. Daubert, M.A.; Yow, E.; Dunn, G.; Marchev, S.; Barnhart, H.; Douglas, P.S.; O’Connor, C.; Goldstein, S.; Udelson, J.E.; Sabbah, H.N. Novel Mitochondria-Targeting Peptide in Heart Failure Treatment: A Randomized, Placebo-Controlled Trial of Elamipretide. Circ. Heart Fail. 2017, 10, e004389. [Google Scholar] [CrossRef] [PubMed]
  299. Gibson, C.M.; Giugliano, R.P.; Kloner, R.A.; Bode, C.; Tendera, M.; Janosi, A.; Merkely, B.; Godlewski, J.; Halaby, R.; Korjian, S.; et al. EMBRACE STEMI study: A Phase 2a trial to evaluate the safety, tolerability, and efficacy of intravenous MTP-131 on reperfusion injury in patients undergoing primary percutaneous coronary intervention. Eur. Heart J. 2016, 37, 1296–1303. [Google Scholar] [CrossRef]
  300. Cerrato, C.P.; Pirisinu, M.; Vlachos, E.N.; Langel, U. Novel cell-penetrating peptide targeting mitochondria. FASEB J. 2015, 29, 4589–4599. [Google Scholar] [CrossRef]
  301. Cerrato, C.P.; Langel, U. Effect of a Fusion Peptide by Covalent Conjugation of a Mitochondrial Cell-Penetrating Peptide and a Glutathione Analog Peptide. Mol. Ther. Methods Clin. Dev. 2017, 5, 221–231. [Google Scholar] [CrossRef] [PubMed]
  302. Zsebo, K.; Yaroshinsky, A.; Rudy, J.J.; Wagner, K.; Greenberg, B.; Jessup, M.; Hajjar, R.J. Long-term effects of AAV1/SERCA2a gene transfer in patients with severe heart failure: Analysis of recurrent cardiovascular events and mortality. Circ. Res. 2014, 114, 101–108. [Google Scholar] [CrossRef] [PubMed]
  303. Chen, C.H.; Budas, G.R.; Churchill, E.N.; Disatnik, M.H.; Hurley, T.D.; Mochly-Rosen, D. Activation of aldehyde dehydrogenase-2 reduces ischemic damage to the heart. Science 2008, 321, 1493–1495. [Google Scholar] [CrossRef] [PubMed]
  304. Gomes, K.M.; Campos, J.C.; Bechara, L.R.; Queliconi, B.; Lima, V.M.; Disatnik, M.H.; Magno, P.; Chen, C.-H.; Brum, P.C.; Kowaltowski, A.J.; et al. Aldehyde dehydrogenase 2 activation in heart failure restores mitochondrial function and improves ventricular function and remodelling. Cardiovasc. Res. 2014, 103, 498–508. [Google Scholar] [CrossRef]
  305. Gomes, K.M.; Bechara, L.R.; Lima, V.M.; Ribeiro, M.A.; Campos, J.C.; Dourado, P.M.; Kowaltowski, A.J.; Mochly-Rosen, D.; Ferreira, J.C. Aldehydic load and aldehyde dehydrogenase 2 profile during the progression of post-myocardial infarction cardiomyopathy: Benefits of Alda-1. Int. J. Cardiol. 2015, 179, 129–138. [Google Scholar] [CrossRef]
  306. Luo, X.J.; Liu, B.; Ma, Q.L.; Peng, J. Mitochondrial aldehyde dehydrogenase, a potential drug target for protection of heart and brain from ischemia/reperfusion injury. Curr. Drug Targets 2014, 15, 948–955. [Google Scholar] [CrossRef]
  307. Piot, C.; Croisille, P.; Staat, P.; Thibault, H.; Rioufol, G.; Mewton, N.; Elbelghiti, R.; Cung, T.T.; Bonnefoy, E.; Angoulvant, D.; et al. Effect of cyclosporine on reperfusion injury in acute myocardial infarction. N. Engl. J. Med. 2008, 359, 473–481. [Google Scholar] [CrossRef] [PubMed]
  308. Ottani, F.; Latini, R.; Staszewsky, L.; La Vecchia, L.; Locuratolo, N.; Sicuro, M.; Masson, S.; Barlera, S.; Milani, V.; Lombardi, M.; et al. Cyclosporine A in Reperfused Myocardial Infarction: The Multicenter, Controlled, Open-Label CYCLE Trial. J. Am. Coll. Cardiol. 2016, 67, 365–374. [Google Scholar] [CrossRef]
  309. Mewton, N.; Croisille, P.; Gahide, G.; Rioufol, G.; Bonnefoy, E.; Sanchez, I.; Cung, T.T.; Sportouch, C.; Angoulvant, D.; Finet, G.; et al. Effect of cyclosporine on left ventricular remodeling after reperfused myocardial infarction. J. Am. Coll. Cardiol. 2010, 55, 1200–1205. [Google Scholar] [CrossRef]
  310. Kawakami, S.; Matsuda, A.; Sunagawa, T.; Noda, Y.; Kaneko, T.; Tahara, S.; Hiraumi, Y.; Adachi, S.; Matsui, H.; Ando, K.; et al. Antioxidant, EUK-8, prevents murine dilated cardiomyopathy. Circ. J. 2009, 73, 2125–2134. [Google Scholar] [CrossRef]
  311. van Empel, V.P.; Bertrand, A.T.; van Oort, R.J.; van der Nagel, R.; Engelen, M.; van Rijen, H.V.; Doevendans, P.A.; Crijns, H.J.; Ackerman, S.L.; Sluiter, W.; et al. EUK-8, a superoxide dismutase and catalase mimetic, reduces cardiac oxidative stress and ameliorates pressure overload-induced heart failure in the harlequin mouse mutant. J. Am. Coll. Cardiol. 2006, 48, 824–832. [Google Scholar] [CrossRef]
  312. Lerman-Sagie, T.; Rustin, P.; Lev, D.; Yanoov, M.; Leshinsky-Silver, E.; Sagie, A.; Ben-Gal, T.; Munnich, A. Dramatic improvement in mitochondrial cardiomyopathy following treatment with idebenone. J. Inherit. Metab. Dis. 2001, 24, 28–34. [Google Scholar] [CrossRef]
  313. Salvemini, D.; Wang, Z.Q.; Zweier, J.L.; Samouilov, A.; Macarthur, H.; Misko, T.P.; Currie, M.G.; Cuzzocrea, S.; Sikorski, J.A.; Riley, D.P.; et al. A nonpeptidyl mimic of superoxide dismutase with therapeutic activity in rats. Science 1999, 286, 304–306. [Google Scholar] [CrossRef]
  314. Masini, E.; Cuzzocrea, S.; Mazzon, E.; Marzocca, C.; Mannaioni, P.F.; Salvemini, D. Protective effects of M40403, a selective superoxide dismutase mimetic, in myocardial ischaemia and reperfusion injury in vivo. Br. J. Pharmacol. 2002, 136, 905–917. [Google Scholar] [CrossRef]
  315. Ni, R.; Cao, T.; Xiong, S.; Ma, J.; Fan, G.C.; Lacefield, J.C.; Lu, Y.; Le Tissier, S.; Peng, T. Therapeutic inhibition of mitochondrial reactive oxygen species with mito-TEMPO reduces diabetic cardiomyopathy. Free Radic. Biol. Med. 2016, 90, 12–23. [Google Scholar] [CrossRef] [PubMed]
  316. Dikalova, A.E.; Bikineyeva, A.T.; Budzyn, K.; Nazarewicz, R.R.; McCann, L.; Lewis, W.; Harrison, D.G.; Dikalova, A. Therapeutic targeting of mitochondrial superoxide in hypertension. Circ. Res. 2010, 107, 106–116. [Google Scholar] [CrossRef] [PubMed]
  317. Dikalova, A.E.; Kirilyuk, I.A.; Dikalov, S.I. Antihypertensive effect of mitochondria-targeted proxyl nitroxides. Redox. Biol. 2015, 4, 355–362. [Google Scholar] [CrossRef] [PubMed]
  318. Tsujita, K.; Shimomura, H.; Kawano, H.; Hokamaki, J.; Fukuda, M.; Yamashita, T.; Hida, S.; Nakamura, Y.; Nagayoshi, Y.; Sakamoto, T.; et al. Effects of edaravone on reperfusion injury in patients with acute myocardial infarction. Am. J. Cardiol. 2004, 94, 481–484. [Google Scholar] [CrossRef]
  319. Edaravone Acute Infarction Study Group. Effect of a novel free radical scavenger, edaravone (MCI-186), on acute brain infarction. Randomized, placebo-controlled, double-blind study at multicenters. Cerebrovasc. Dis. 2003, 15, 222–229. [Google Scholar] [CrossRef]
  320. Junior, R.F.R.; Dabkowski, E.R.; Shekar, K.C.; O’Connell, K.A.; Hecker, P.A.; Murphy, M.P. MitoQ improves mitochondrial dysfunction in heart failure induced by pressure overload. Free Radic. Biol. Med. 2018, 117, 18–29. [Google Scholar] [CrossRef]
  321. Adlam, V.J.; Harrison, J.C.; Porteous, C.M.; James, A.M.; Smith, R.A.; Murphy, M.P.; Sammut, I.A. Targeting an antioxidant to mitochondria decreases cardiac ischemia-reperfusion injury. FASEB J. 2005, 19, 1088–1095. [Google Scholar] [CrossRef]
  322. Yin, M.; van der Horst, I.C.; van Melle, J.P.; Qian, C.; van Gilst, W.H.; Sillje, H.H.; de Boer, R.A. Metformin improves cardiac function in a nondiabetic rat model of post-MI heart failure. Am. J. Physiol. Heart Circ. Physiol. 2011, 301, H459–H468. [Google Scholar] [CrossRef]
  323. Sun, D.; Yang, F. Metformin improves cardiac function in mice with heart failure after myocardial infarction by regulating mitochondrial energy metabolism. Biochem. Biophys. Res. Commun. 2017, 486, 329–335. [Google Scholar] [CrossRef] [PubMed]
  324. Dridi, H.; Wu, W.; Reiken, S.R.; Ofer, R.M.; Liu, Y.; Yuan, Q.; Sittenfeld, L.; Kushner, J.; Muchir, A.; Worman, H.J.; et al. Ryanodine receptor remodeling in cardiomyopathy and muscular dystrophy caused by lamin A/C gene mutation. Hum. Mol. Genet. 2021, 29, 3919–3934. [Google Scholar] [CrossRef] [PubMed]
  325. Atar, D.; Arheden, H.; Berdeaux, A.; Bonnet, J.L.; Carlsson, M.; Clemmensen, P.; Cuvier, V.; Danchin, N.; Dubois-Randé, J.-L.; Engblom, H.; et al. Effect of intravenous TRO40303 as an adjunct to primary percutaneous coronary intervention for acute ST-elevation myocardial infarction: MITOCARE study results. Eur. Heart J. 2015, 36, 112–119. [Google Scholar] [CrossRef]
  326. Schaller, S.; Paradis, S.; Ngoh, G.A.; Assaly, R.; Buisson, B.; Drouot, C.; Ostuni, M.A.; Lacapere, J.J.; Bassissi, F.; Bordet, T.; et al. TRO40303, a new cardioprotective compound, inhibits mitochondrial permeability transition. J. Pharmacol. Exp. Ther. 2010, 333, 696–706. [Google Scholar] [CrossRef] [PubMed]
  327. Butt, N.; Bache-Mathiesen, L.K.; Nordrehaug, J.E.; Tuseth, V.; Munk, P.S.; Bonarjee, V.; Hall, T.S.; Jensen, S.E.; Halvorsen, S.; Firat, H.; et al. Administration of the Mitochondrial Permeability Transition Pore Inhibitor, TRO40303, prior to Primary Percutaneous Coronary Intervention, Does Not Affect the Levels of Pro-Inflammatory Cytokines or Acute-Phase Proteins. Cardiology 2017, 138, 122–132. [Google Scholar] [CrossRef]
Figure 1. Physical interaction and Ca2+ transfer from the SR to the mitochondria in cardiomyocytes: during the depolarization phase of the cardiac action potential, Ca2+ enters the cardiomyocyte through voltage-activated L-type Ca2+ channels (CaV1.2) and triggers RyR2 to open and release Ca2+ from SR Ca2+ stores through calcium-induced calcium release (CICR). RyR2 activity can be increased by CaMKII or PKA phosphorylation, in response to stress. Ca2+ release from the SR by RyR2 creates high [Ca2+]cyt microdomains for mitochondrial Ca2+ uptake. Close contact points between the SR and mitochondria are controlled by SR mitofusin 2 (MFN2) tethering to MFN2 and MFN1 on the outer mitochondrial membrane (OMM), further creating microdomains of a high Ca2+ concentration in the vicinity of mitochondria. Ca2+ release from the SR by IP3R transits through the protein complex formed by IP3R/GRP75/VDAC and enters the mitochondria through the MCU, later being removed by mNCLX. Ca2+ activates complex III of the electron transport chain, ATP Synthase, several dehydrogenases in the TCA cycle, mPTP, CaMKII, and contraction of the sarcomeres. To induce repolarization, SERCA2a and NCX cycle Ca2+ out of the cytosol.
Figure 1. Physical interaction and Ca2+ transfer from the SR to the mitochondria in cardiomyocytes: during the depolarization phase of the cardiac action potential, Ca2+ enters the cardiomyocyte through voltage-activated L-type Ca2+ channels (CaV1.2) and triggers RyR2 to open and release Ca2+ from SR Ca2+ stores through calcium-induced calcium release (CICR). RyR2 activity can be increased by CaMKII or PKA phosphorylation, in response to stress. Ca2+ release from the SR by RyR2 creates high [Ca2+]cyt microdomains for mitochondrial Ca2+ uptake. Close contact points between the SR and mitochondria are controlled by SR mitofusin 2 (MFN2) tethering to MFN2 and MFN1 on the outer mitochondrial membrane (OMM), further creating microdomains of a high Ca2+ concentration in the vicinity of mitochondria. Ca2+ release from the SR by IP3R transits through the protein complex formed by IP3R/GRP75/VDAC and enters the mitochondria through the MCU, later being removed by mNCLX. Ca2+ activates complex III of the electron transport chain, ATP Synthase, several dehydrogenases in the TCA cycle, mPTP, CaMKII, and contraction of the sarcomeres. To induce repolarization, SERCA2a and NCX cycle Ca2+ out of the cytosol.
Biomolecules 13 01409 g001
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

Dridi, H.; Santulli, G.; Bahlouli, L.; Miotto, M.C.; Weninger, G.; Marks, A.R. Mitochondrial Calcium Overload Plays a Causal Role in Oxidative Stress in the Failing Heart. Biomolecules 2023, 13, 1409. https://doi.org/10.3390/biom13091409

AMA Style

Dridi H, Santulli G, Bahlouli L, Miotto MC, Weninger G, Marks AR. Mitochondrial Calcium Overload Plays a Causal Role in Oxidative Stress in the Failing Heart. Biomolecules. 2023; 13(9):1409. https://doi.org/10.3390/biom13091409

Chicago/Turabian Style

Dridi, Haikel, Gaetano Santulli, Laith Bahlouli, Marco C. Miotto, Gunnar Weninger, and Andrew R. Marks. 2023. "Mitochondrial Calcium Overload Plays a Causal Role in Oxidative Stress in the Failing Heart" Biomolecules 13, no. 9: 1409. https://doi.org/10.3390/biom13091409

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