Abstract
Oxidative phosphorylation is regulated by mitochondrial calcium (Ca2+) in health and disease. In physiological states, Ca2+ enters via the mitochondrial Ca2+ uniporter and rapidly enhances NADH and ATP production. However, maintaining Ca2+ homeostasis is critical: insufficient Ca2+ impairs stress adaptation, and Ca2+ overload can trigger cell death. In this review, we delve into recent insights further defining the relationship between mitochondrial Ca2+ dynamics and oxidative phosphorylation. Our focus is on how such regulation affects cardiac function in health and disease, including heart failure, ischemia-reperfusion, arrhythmias, catecholaminergic polymorphic ventricular tachycardia, mitochondrial cardiomyopathies, Barth syndrome, and Friedreich’s ataxia. Several themes emerge from recent data. First, mitochondrial Ca2+ regulation is critical for fuel substrate selection, metabolite import, and matching of ATP supply to demand. Second, mitochondrial Ca2+ regulates both the production and response to reactive oxygen species (ROS), and the balance between its pro- and antioxidant effects is key to how it contributes to physiological and pathological states. Third, Ca2+ exerts localized effects on the electron transport chain (ETC), not through traditional allosteric mechanisms but rather indirectly. These effects hinge on specific transporters, such as the uniporter or the Na+/Ca2+ exchanger, and may not be noticeable acutely, contributing differently to phenotypes depending on whether Ca2+ transporters are acutely or chronically modified. Perturbations in these novel relationships during disease states may either serve as compensatory mechanisms or exacerbate impairments in oxidative phosphorylation. Consequently, targeting mitochondrial Ca2+ holds promise as a therapeutic strategy for a variety of cardiac diseases characterized by contractile failure or arrhythmias.
Keywords: heart failure, ischemia-reperfusion injury, MCU, mitochondrial calcium transport, mitochondrial cardiomyopathy
Introduction
Calcium ions (Ca2+) are fundamental for a multitude of cellular processes, allowing the rapid transmission and interconversion between electrical and biochemical signals during electrical conduction, muscular contraction, and metabolic signaling (1). The endoplasmic reticulum (ER) is the primary Ca2+ store in the cell (200–500 µM free Ca2+), whereas levels within the mitochondrial matrix are lower (<100 µM free Ca2+, though the resting levels have been the topic of some controversy) (2–5). Thus, when released from the ER, Ca2+ levels rapidly increase in nearby mitochondria. Within mitochondria, Ca2+ signaling has long been recognized to increase NADH, double ATP synthesis rates, and regulate reactive oxygen species (ROS) production (6–8). In this review, we focus on more recent investigations conducted in murine models of normal or diseased cardiac function that have expanded our understanding of mitochondrial Ca2+ signaling in relation to oxidative phosphorylation (OXPHOS) and oxidative stress. Beyond direct binding and allosteric effects on mitochondrial enzymes, these studies have revealed a multitude of indirect mechanisms that fine-tune Ca2+ regulation of OXPHOS. Notably, these novel forms of regulation often depend on local interactions between Ca2+ transporters and mitochondrial complexes, emphasizing the localized nature of Ca2+ signaling (FIGURE 1). Over the course of the review, we also highlight critical unanswered questions and controversies regarding the mechanisms by which Ca2+ regulates OXPHOS.
FIGURE 1.
The localized nature of mitochondrial Ca2+ signaling In the heart, Ca2+ transfer from sarcoplasmic reticulum (SR) to mitochondria occurs at sites of close contact between the SR and mitochondria, allowing rapid transport through RyR2 (cardiac ryanodine receptor), voltage-dependent anion channel (VDAC), and MCU [mitochondrial Ca2+ uniporter (pore-forming subunit)] into the mitochondrial matrix. Ca2+ export back to the SR occurs in a more distal location, with Ca2+ exchange for Na+ via NCLX (Na+/Ca2+ exchanger) and finally uptake back into the SR via sarco(endo)plasmic reticulum Ca2+ ATPase (SERCA) pumps. IMM, inner mitochondrial membrane; IMS, intermembrane space; MICU1/2, mitochondrial Ca2+ uptake (gating subunit); OMM, outer mitochondrial membrane.
Mitochondrial Ca2+ Transporters and Their Spatial Distribution
An overview of the molecules involved in mitochondrial Ca2+ transport is included in Table 1.
Table 1.
Key components of Ca2+ transporting proteins
Gene | Protein | Tissue Expression | Localization | Comments | Reviews/ References |
---|---|---|---|---|---|
VDAC1 | Voltage-Dependent Anion Channel 1 | Ubiquitous, abundant in heart, liver, kidney, and brain | OMM | Responsible for Ca2+ transport across the OMM. Regulated by cholesterol, fatty acids, and NADH | (9, 10) |
VDAC2 | Voltage-Dependent Anion Channel 2 | Heart, skeletal muscle, brain, lung, pancreas | OMM | Responsible for Ca2+ transport across the OMM. Targeted by cyclophilin D in the mitochondrial permeability transition pore opening | (9, 10) |
VDAC3 | Voltage-Dependent Anion Channel 3 | Testis, spleen, ovary, brain | OMM | Responsible for Ca2+ transport across the OMM. Regulated by ceramide, calcium, and pH response to oxidative stress, altered mitochondrial respiration | (9, 10) |
MCU | Mitochondrial Calcium Uniporter | Ubiquitous, high levels in heart, muscle, and brain | IMM | Pore-forming subunit of the mitochondrial Ca2+ uniporter, allowing Ca2+ influx across the IMM. Inhibited by Ru360, ruthenium red, mitoxantrone, and others | (11, 12) |
MCUB | Mitochondrial Calcium Uniporter dominant negative subunit beta | Broad expression. Low levels in heart but upregulated during injury. | IMM | Homologous to MCU but inhibits Ca2+ uptake when incorporated into the channel pore in a dominant-negative fashion | (11–14) |
MICU1 | Mitochondrial Calcium Uptake 1 | Ubiquitous, highest in heart, muscle, and brain | IMS | Component of the uniporter. Acts as a gatekeeper, modulating MCU activity in response to Ca²+ levels | (11, 12) |
MICU2 | Mitochondrial Calcium Uptake 2 | Ubiquitous | IMS | Component of the uniporter. Collaborates with MICU1 in regulating MCU activity | (11, 12) |
MICU3 | Mitochondrial Calcium Uptake 3 | Ubiquitous | IMS | Component of the uniporter. Collaborates with MICU1 in regulating MCU activity | (11, 12) |
SMDT1 (EMRE) | Essential MCU regulator | Ubiquitous | IMM | Transmembrane protein that surrounds MCU, opening the channel and helping bridge the interaction with MICU1 | (11, 12) |
MCUR1 | Mitochondrial calcium uniporter regulator 1 | Ubiquitous | IMM | Positive regulator of mitochondrial Ca2+ uptake through the uniporter | (11, 12) |
SLC8B1 (NCLX) | Mitochondrial Na+/Ca2+ exchanger | Ubiquitous, high expression in heart, brain, and muscle | IMM | Mediates the efflux of Ca2+ from the matrix via exchange with Na+. Helps prevent mitochondrial Ca2+ overload | (15, 16) |
LETM1 | Leucine zipper and EF-hand containing transmembrane protein 1 | Ubiquitous | IMM | Mediates the efflux of Ca2+ from the matrix via exchange with H+. May also be involved in K+/H+ exchange | (17, 18) |
TMEM65 | Transmembrane protein 65 | Widely expressed, with higher levels in brain, heart, and skeletal muscle | IMM | New potential component of Na+/Ca2+ exchanger | (19–21) |
TMBIM5 | Transmembrane BAX inhibitor motif-containing protein 5 | Ubiquitous expression, higher levels in brain, heart, and muscle tissues | IMM | New potential component of the Ca2+/H+ exchanger | (22–24) |
IMM, inner mitochondrial membrane; IMS, intermembrane space; OMM, outer mitochondrial membrane.
Mitochondrial Outer Membrane
Voltage-dependent anion channel.
The outer membrane of the mitochondria is not felt to be a significant barrier to calcium diffusion because of the presence of voltage-dependent anion channel (VDAC) channels. These channels have an array of antiparallel beta strands that create a large pore (9). Even when these channels constrict to block diffusion of larger molecules, small ions such as Ca2+ can diffuse through them freely. Although the channels themselves do not form critical barriers to Ca2+ diffusion, by altering their localization across the outer membrane they can facilitate Ca2+ transfer from sites of release in the ER or sarcoplasmic reticulum (SR) to the inner membrane (10). In fact, cardiac-specific loss of VDAC isoforms can lead to dilated cardiomyopathy due to insufficient Ca2+ transfer to mitochondria (25). Thus, processes that alter the clustering of VDAC channels can also alter the rates of Ca2+ influx into the mitochondria.
Mitochondrial Inner Membrane Ca2+ Import
Mitochondrial Ca2+ uniporter.
High-capacity Ca2+ influx, ascribed to a uniporter driven by mitochondrial membrane potential (ΔΨ), occurs through a highly selective Ca2+ channel (26–28). Whereas initial efforts with limited control over ΔΨ and Ca2+ buffering had suggested that this pathway was low affinity (29), more precise recent measurements have shown the ability of the channel to conduct Ca2+ at submicromolar levels, especially in the heart (30–33). Efforts in the past decade culminated with the discovery of the genes encoding this multisubunit channel, known as the mitochondrial Ca2+ uniporter (11, 34–40). From here onward, we refer to the entire channel as the “uniporter,” whereas individual subunits {e.g., MCU [mitochondrial Ca2+ uniporter (pore-forming subunit)]} are specified by their protein or gene abbreviation. The pore-forming subunit is encoded by the MCU gene. Recent structural studies revealed a tetrameric organization of MCU, centered around a narrow pore lined with concentric rings of glutamate residues that create a highly negatively charged selectivity filter with nanomolar affinity for Ca2+ (12, 41–46). A highly homologous subunit, MCUB, appears to act in a dominant-negative manner, inhibiting Ca2+ transport in channels into which it incorporates (39). In metazoans, the transmembrane protein EMRE (SMDT1 gene) surrounds the MCU channel, helping to open the pore where it exits into the matrix and bridging interactions between MCU and the accessory MICU1 subunits (45, 47, 48). The MICU1–3 subunits are located in the intermembrane space, contain Ca2+-binding EF-hands, gate the channel by blocking the pore, and help the channel dimerize (43, 49). Micromolar Ca2+ binds to the MICU EF-hands, releasing them from the pore, allowing Ca2+ passage and turning the uniporter into a Ca2+-gated Ca2+ channel. These core components of the uniporter can also bind to other mitochondrial proteins to regulate Ca2+ import. Although compelling evidence described below indicates that genetic deletion of MCU appears to largely prevent Ca2+ import, basal mitochondrial Ca2+ levels are often only mildly affected. It may be that other transporters, either the known Ca2+ exchangers described below or others, may compensate (50, 51).
As with VDAC channels on the outer membrane, the spatial localization of the uniporter helps regulate mitochondrial metabolism. MICU1 subunits appear restricted to the boundary membrane, the portion of the inner mitochondrial membrane closest to the outer mitochondrial membrane; MCU itself is more homogeneously distributed and is found in cristae as well (52). During Ca2+ signals, the distribution of these two shows much greater overlap. Superresolution microscopy has revealed that the uniporter tends to cluster at the edges of cardiomyocyte mitochondria in the vicinity of T tubules and the sarcoplasmic reticulum (SR), where cardiac Ca2+ signals originate (53). Thus, within the heart, the uniporter appears to be clustered at sites of active Ca2+ signaling, localizing close to the outer membrane VDAC and SR-embedded ryanodine receptors (RyRs) and guaranteeing seamless Ca2+ transfer from SR to mitochondrial matrix (FIGURE 1).
Outstanding questions: 1) From recent structural studies, it is clear that uniporter channel dimerizes via both the MCU NH2-terminal domain and MICU1–3 subunits. This may allow cooperative activation, with Ca2+ binding at only one set of MICU subunits, allowing opening of the adjacent channel as well, though this has not been demonstrated functionally and any impact on OXPHOS remains unexplored. 2) MCUB inhibits Ca2+ transport when incorporated into uniporter channels, but the mechanism of inhibition remains unclear. When incorporated into lipid bilayers, it may form Na+ channels, but how it inhibits the native uniporter complex remains unanswered. 3) Given the restricted localization of MICU subunits relative to the homogeneous MCU distribution in the inner membrane, what is the mechanism for MCU-MICU redistribution during Ca2+ signals, and how does this affect the time course of uptake during Ca2+ signaling?
Mitochondrial Inner Membrane Ca2+ Export
Na+/Ca2+ exchange.
Ca2+ export from the mitochondrial matrix has not been studied to the degree that Ca2+ import has. However, there are two well-characterized pathways mediating such export. One pathway involves the influx of Na+ ions in exchange for Ca2+ efflux in a possibly electroneutral ratio (54, 55). The exchanger is encoded by the NCLX protein (SLC8B1 gene) (15). Loss of this transmembrane protein is lethal within cardiomyocytes, likely because of overwhelming Ca2+ overload when this efflux pathway is blocked (16). The spatial distribution of NCLX may also contribute to its effects on Ca2+ regulation. Within cardiomyocytes NCLX seems to be distributed further away from sources of Ca2+, and closer to the sarco(endo)plasmic Ca2+-ATPase (SERCA), which takes up Ca2+ back into the SR (56) (FIGURE 1). In recently posted preprints, TMEM65 also appears to contribute to Na+/Ca2+ exchange, though it is unclear whether this protein interacts with NCLX or encodes an entirely separate pathway (19–21).
Ca2+/H+ exchange.
The second pathway for Ca2+ export from the mitochondrial matrix involves the electroneutral exchange of Ca2+ for H+ via the Ca2+/H+ exchanger. One protein mediating Ca2+/H+ exchange has been identified as LETM1, though it may also be involved in the exchange of K+ and H+ (17, 18, 57–60). Within heart cells, despite expression of the protein and closely related homologs, this activity is thought to be less important for the export of Ca2+ compared with NCLX, since it cannot compensate for the loss of NCLX in knockout (KO) mice (16). Whether LETM1 homologs (LETMD1 and LETM2) are also involved in Ca2+/H+ exchange has also not been well established. Although loss of LETM1 is lethal, the precise consequences of altered Ca2+/H+ exchange on mitochondrial mechanisms are not well understood, nor do we know whether there is heterogeneity in spatial distribution along the mitochondrial inner membrane. Finally, very recent reports have also identified TMBIM5 as critical for Ca2+/H+ exchange, potentially via direct interaction with LETM1 (22–24), though the precise mechanisms by which it contributes to Ca2+/H+ exchange remain to be established.
Mitochondrial Ca2+ Buffering
Some of the most poorly understood aspects of mitochondrial Ca2+ physiology are the identities, mechanisms, and localization of matrix Ca2+ buffers. The critical difficulty appears to be that, unlike the ER, there are no well-characterized high-capacity Ca2+ binding proteins in the mitochondrial matrix that would allow genetic or pharmacological modification. Nonetheless, the mitochondrial matrix has remarkable capacity to buffer Ca2+ influx, with total matrix Ca2+ thought to be 103- to 105-fold greater than free Ca2+ concentrations (61–63). This substantial buffering power comes from the presence of amorphous calcium phosphate, a substance strongly regulated by pH (64–66). Within the matrix, phosphate can oligomerize into complex polymers with high-energy bonds and high capacity for binding Ca2+. Cleavage of these polymers in cultured cells can reduce ΔΨ, increase NADH, and produce a shift away from OXPHOS toward glycolysis, though whether these effects depend on altering Ca2+ levels is unclear (67, 68). In addition, Ca2+ phosphate complexes can form insoluble granules within the matrix (62). How these localized structures regulate Ca2+ physiology remains difficult to study experimentally, though in modeling studies they may interfere with Complex I-dependent respiration (69).
Mitochondrial Ca2+ buffering can potentially alter the timing and strength of cytoplasmic Ca2+ signals. These altered cytoplasmic Ca2+ levels can engage a variety of signaling pathways, such as 5′ AMP-activated protein kinase or the cAMP response element-binding protein, in a variety of tissues (70, 71). Such effects have been reviewed elsewhere (72–74). In the heart, Ca2+ extrusion from the cytoplasm is dominated by transporters at the plasma membrane and SR, so there is limited direct effect on the action potential from mitochondrial Ca2+ buffering (75, 76). However, chronic alteration of mitochondrial Ca2+ uptake can alter action potential dynamics, contributing to arrhythmias in both the atrium and the ventricle, as described below in Catecholaminergic Polymorphic Ventricular Tachycardia (77).
Ca2+ in Mitochondrial Metabolism
Allosteric Ca2+ Binding Enhances the Activity of Metabolite Import
Ca2+-activated mitochondrial carriers (CaMCs) are mitochondrial inner membrane transporters of metabolites, which are regulated by Ca2+ via intermembrane space-facing EF-hands, similar to MICU1 (78) (FIGURE 2). When Ca2+ levels rise within the cytoplasm and intermembrane space, Ca2+ binding to these carriers enhances metabolite transport rates. The aspartate/glutamate exchangers (AGCs, aralar and citrin) possess longer NH2-terminal extensions with multiple EF-hand copies. Both AGCs are expressed in the heart, are involved in the aspartate-malate shuttle, and, although active at baseline, can be further enhanced by relatively modest increases in Ca2+ (<500 nM) (79). In contrast, the shorter CaMCs are ATP-Mg2+/Pi exchangers that need higher concentrations of Ca2+ to become active. These CaMCs also have NH2-terminal extensions containing EF-hands that more closely resemble those in calmodulin (78). In recent studies, Ca2+-dependent enhancement of substrate transport via these carriers was a primary mechanism by which cytoplasmic Ca2+ could accelerate delivery of substrates for OXPHOS during cardiac mitochondrial metabolism (80).
FIGURE 2.
Overview of established and novel forms of mitochondrial metabolism regulation by Ca2+ MCU [mitochondrial Ca2+ uniporter (pore-forming subunit)] imports Ca2+ into the matrix, whereas export mechanisms include Ca2+/H+ exchange via LETM1/TMBIM5 and Na+/Ca2+ exchange via NCLX/TMEM65. Well-established mechanisms of Ca2+ regulation are shown with solid red arrows and include allosteric regulation of aspartate-glutamate carriers (AGC), glycerol-3-phosphate dehydrogenase 2 (GPD2), isocitrate dehydrogenase (IDH), α-ketoglutarate dehydrogenase (αKDH), pyruvate dehydrogenase (PDH) phosphatase, and short Ca2+-binding mitochondrial carriers (SCaMC). Newer indirect mechanisms are shown as dashed red arrows and involve regulation of Complex I, coenzyme Q (CoQ), Complex IV, and ATP synthase. Cyt c, cytochrome c; IMM, inner mitochondrial membrane; IMS, intermembrane space; MCU, mitochondrial Ca2+ uniporter (pore-forming subunit); MPC, mitochondrial pyruvate carrier; OMM, outer mitochondrial membrane.
The glycerol-3-phosphate shuttle can also be subject to Ca2+ regulation (81). This pathway provides another route for electrons to enter the ETC. In this pathway, dihydroxyacetone conversion into glycerol-3-phosphate is catalyzed by cytoplasmic glycerol-3-phosphate dehydrogenase (GPD). Glycerol-3-phosphate can then diffuse into the intermembrane space, where it is oxidized by GPD2, transferring electrons to coenzyme Q (CoQ) upstream of Complex III. Modest increases in Ca2+ (100–250 nM) in the intermembrane space can bind to the COOH-terminal EF-hands within GPD2 to enhance its activity (81–84). This pathway appears to be minimally active in the healthy heart but may be upregulated during ischemic disease, particularly when Complex I of the ETC is dysfunctional (81, 85–88).
Ca2+ Regulation of Pyruvate Dehydrogenase Alters Fuel Substrate Choice
Much of the work on the uniporter over the past decade has shown that it is a critical determinant of mitochondrial fuel substrate preference, which we have reviewed recently and summarize briefly here (89–92). The key nexus for this regulation is the pyruvate dehydrogenase (PDH) complex, a multisubunit enzyme that transforms pyruvate into acetyl CoA. PDH activity is inhibited when PDH kinase phosphorylates it and enhanced when PDH phosphatases remove these modifications. Ca2+ stimulates PDH indirectly, by allosterically binding (K0.5 ∼1 µM) and promoting the activity of PDH phosphatase (93, 94) (FIGURE 2).
In recent studies of cardiac and skeletal muscle-specific knockouts of MCU, flux through PDH was reduced (89–91). Because pyruvate contributes to the tricarboxylic acid (TCA) cycle downstream of glycolysis, deletion of MCU produced a shift of substrate utilization away from glucose toward fatty acids. Moreover, in reciprocal fashion, inhibition of the import of pyruvate into mitochondria was associated with inhibition of mitochondrial Ca2+ uptake (95). This mechanism may prevent mitochondrial Ca2+ overload when nutrient depletion causes bioenergetic stress.
Allosteric Ca2+ Binding Stimulates Multiple TCA Cycle Dehydrogenases
Since the discovery of Ca2+ import into the mitochondria, the best-established mechanism of metabolic regulation has been through binding and allosteric enhancement of several key dehydrogenases of the TCA cycle, which have been reviewed extensively (6, 93, 96) (FIGURE 2). Ca2+ activates isocitrate dehydrogenase (IDH) (97, 98) [K0.5 at Ca2+ 5–50 µM (99, 100)] and α-ketoglutarate dehydrogenase (AKGDH) (101, 102) [KM at Ca2+ 0.2 µM with ADP or ∼2 µM with ATP (99)]. Surprisingly, despite substantial evidence supporting this traditional model, recent investigations using knockouts of the mitochondrial calcium uniporter, especially in the heart, found relatively preserved respiration and ATP production (103–108). This suggests either that the well-established effects of Ca2+ on TCA cycle dehydrogenases may be less important compared with its upstream effects on metabolite transport and fuel substrate choice or that the absence of Ca2+-induced TCA cycle enhancement may be compensated by other mechanisms, such as product inhibition. Importantly, although uniporter deficiency minimally affects basal cardiac function, it is clear that the uniporter is required to boost ATP synthesis rapidly during periods of stress (see Ischemia-Reperfusion Injury and Systolic Heart Failure) (105, 107, 109).
Ca2+ Alters Oxidative Phosphorylation via Indirect Mechanisms
Ca2+ regulation of the electron transport chain (ETC) and F1-FO-ATP synthase is an unsettled question. Direct Ca2+ regulation of the ETC via allosteric binding has never been established at physiological Ca2+ concentrations, though a recent preprint suggests that Ca2+ may bind multiple subunits of ETC Complexes I–IV and the ATP synthase (110). Regardless, various studies have shown Ca2+-dependent increases in activity of the ETC and ATP synthase in isolated mitochondria or purified enzyme systems (6). The key benefit of Ca2+ regulation of OXPHOS is its very rapid onset, occurring within ∼250 ms of Ca2+ entry (111). In two recent studies of OXPHOS in cardiac and skeletal mitochondria, Ca2+-induced increases in ATP production were explained by its effects on substrate import, TCA cycle flux, and increased ΔΨ rather than direct allosteric effects on the ETC or ATP synthase (33, 80). However, a different study examining skeletal muscle mitochondria revealed more than twofold increases in OXPHOS rates induced by Ca2+ at ETC Complexes I, III, and IV, even after controlling for the effects of Ca2+ on substrate import, ΔΨ, and NADH production (112). The use of different assays in these studies, including respirometry, cytochrome redox absorbance, and fluorescent reporters, may account for some of the discrepant findings. Furthermore, the use of the creatine kinase clamp to mimic resistance to ATP production under physiological conditions may have affected the results (80, 112), as the effects of Ca2+ on respiration are minimized when saturating levels of ADP are used (113). Despite conflicting data on direct Ca2+ allosteric regulation of OXPHOS, there is a growing list of recent studies described below that show that mitochondrial Ca2+ is a critical indirect regulator of OXPHOS (FIGURE 2).
Complex I (NADH:ubiquinone oxidoreductase).
Complex I is the largest component of the ETC, consisting of an L-shaped structure with multiple transmembrane subunits responsible for H+ pumping and a matrix-facing soluble arm that is responsible for transferring electrons from NADH to coenzyme Q (CoQ). Until a very recent preprint (see future directions and conclusions), there were no data to suggest that Ca2+ binds directly to Complex I. Nevertheless, our group has recently discovered that Complex I interacts with the uniporter and regulates its stability, via a mechanism we term Complex I-induced protein turnover (CLIPT) (114) (FIGURE 3). Under normal physiology, during NADH oxidation, some electrons leak away from Complex I at a low rate (<1%) (115). Such electrons rapidly react with molecular oxygen to produce reactive oxygen species (ROS). These species are highly reactive and primarily disrupt Complex I locally (116). We identified the NH2-terminal domain of MCU as an interactor with Complex I, precisely near the electron transfer sites of its soluble arm. Under normal conditions, the electron leak from Complex I oxidizes MCU and leads to its degradation by quality-control proteases (114). Therefore, MCU is constantly turning over and potentially protecting Complex I from oxidation caused by this harmful electron leak. When MCU levels decline, loss of this protective function may produce long-term decreases in Complex I levels, supported by recent studies that show loss of Complex I when the uniporter is disrupted in cardiac tissue (89, 109).
FIGURE 3.
New mechanisms involving mitochondrial Ca2+ A: Complex I interacts with MCU [mitochondrial Ca2+ uniporter (pore-forming subunit)] and increases its turnover because of reactive oxygen species (ROS) leak (CLIPT), which helps protect Complex I from oxidation. B: matrix acidification during ischemic injury solubilizes Ca2+ from Ca2+ phosphate precipitates. The increased matrix Ca2+ via NCLX boosts matrix Na+, thereby decreasing inner mitochondrial membrane (IM) fluidity and slowing coenzyme Q (CoQ) diffusion and the activity of Complexes II + III. C: excess Ca2+ may trigger rearrangement of the F1 domain of the ATP synthase to produce a channel in the membrane. IMS, intermembrane space; NTD, NH2-terminal domain of MCU.
Complex I is the ETC complex most affected by congenital mutations. Our group found that cardiac mitochondrial Ca2+ increases when Complex I is impaired, as a potential compensatory mechanism needed for survival. Impairment in Complex I led to a loss of the interaction with MCU. Absent this interaction, MCU became stabilized, with a consequent increase in channel levels (114, 117). This effect was extremely well conserved, evident in Drosophila, mice, and induced pluripotent stem cells (IPSCs) derived from an infant with Complex I deficiency. Moreover, we found that the uniporter activity was essential for survival. In Drosophila with Complex I deficiency, suppressing uniporter expression led to severe developmental mortality, whereas overexpressing the uniporter rescued both survival and function. Such results suggest that altering mitochondrial Ca2+ may be a potential therapeutic target in the setting of Complex I dysfunction.
Complex II (succinate dehydrogenase).
Complex II is a component of the TCA cycle, producing FADH2, which is subsequently oxidized to transfer electrons to coenzyme Q in the ETC. There has been no evidence of direct or indirect Ca2+ regulation of this complex under normal physiology, though loss of Complex II may impair the ability to take up mitochondrial Ca2+ in cell lines (118).
Complex III (CoQH2-cytochrome c reductase).
Complex III is a dimeric structure that transfers electrons from reduced CoQ (CoQH2) to cytochrome c, through a loop known as the Q cycle. In this process, a net two protons are pumped into the intermembrane space. As with Complex II, physiological levels of Ca2+ do not appear to directly affect Complex III activity, although there is at least one publication that suggests that Ca2+ at very low submicromolar concentrations may alter flux through an unclear mechanism (119).
Despite the absence of direct effects of Ca2+ on Complex III during normal physiology, a recent publication details a novel mechanism by which Ca2+ affects the flux of CoQ from Complex II to Complex III (120) (FIGURE 3). Ischemia-reperfusion injury leads to increases in cytoplasmic Na+ levels, due to limited activity of the Na+-K+-ATPase, and acidification of the mitochondrial matrix. This acidification releases Ca2+ from calcium phosphate granules located in the matrix. Then, increases in Ca2+ in the matrix as well as Na+ in the cytoplasm cause overactivation of NCLX, contributing to Na+ overload within the mitochondrial matrix. These increases in Na+ within the mitochondria are thought to decrease the fluidity of the inner mitochondrial membrane and slow the diffusion of CoQH2 from Complex II to Complex III, decreasing flux through the ETC. Moreover, the Complex III semiquinone is prevented from completing the Q cycle, and its excess electrons can leak out as ROS, promoting oxidative damage. Here, rather than compensating for deficits in the ETC, as during Complex I impairment, increases in mitochondrial Ca2+ may contribute to oxidative damage.
Complex IV (cytochrome-c oxidase).
Complex IV transfers electrons from cytochrome c to oxygen. This redox reaction is coupled to the translocation of four protons into the intermembrane space, contributing to setting ΔΨ. Notably, the intermembrane space side of complex IV has a binding site for Na+ and Ca2+ (121). Under resting Ca2+ levels, the site is normally bound to Na+. However, during Ca2+ signals (∼1 µM, though prior reports suggested >20 µM), the Na+ is displaced, inhibiting activity of the complex with an IC50 of 0.5 µM. It appears that this inhibition may only be evident when fuel substrates are limited. In addition, Complex IV may also be indirectly regulated by the uniporter. MCU regulator 1 (MCUR1) appears to be an assembly factor for Complex IV as well as a scaffold factor for the uniporter (122–125). MCUR1 inhibition slows mitochondrial Ca2+ uptake and may destabilize Complex IV. Moreover, during periods of mitochondrial Ca2+ overload, MCUR1 appears to be a critical sensor of the Ca2+ threshold triggering the permeability transition (126), a process described in more detail below. The mechanistic basis for these effects remains unexplored.
Supercomplexes.
Complex I binds to two subunits of Complex III and a single subunit of Complex IV to assemble a supercomplex known as the respirasome (127, 128). Although these respirasomes are seen repeatedly in biochemical assays, whether they exert functional effects on energetic flux or reflect optimal packing density within mitochondrial cristae is an ongoing topic of research (129–131). In fact, recent studies suggest that disrupting supercomplex formation has limited effects on cardiac function (132). Nevertheless, multiple interventions that alter mitochondrial Ca2+ uptake also affect supercomplex formation, though the association between Ca2+ uptake and supercomplex formation is not consistent across studies (133–135). For example, in one study sprint interval training exercise led to increases in mitochondrial Ca2+ and supercomplex formation (136). In contrast, in another report decreased levels of mitochondrial Ca2+ in female cardiomyocytes, relative to males, were associated with increased supercomplex formation through the COX7RP interfacial subunit, though respiration was unaffected (137). This effect was associated with lower ROS levels and thought to partly explain the lower propensity for arrhythmias in premenopausal females.
F1-FO-ATP synthase.
The F1-FO-ATP synthase contains a circular membrane-embedded FO module and a rotary, soluble F1 module, connected by central and peripheral stalks (138, 139). This allows the catalytic F1 portion to synthesize ATP from ADP and phosphate, driven by H+ flux through an oligomeric ring in FO. No evidence exists for direct allosteric regulation by Ca2+ during normal physiology (33, 140). Intriguingly, in trypanosomal parasites the uniporter may directly interact with the ATP synthase c subunit, the highly conserved protein that forms the oligomeric ring of FO (141).
In pathophysiological states, however, it is possible that Ca2+ exerts a profound effect on ATP synthase structure and function. It is well established that excessive Ca2+ uptake opens a cyclosporine-sensitive, large-conductance channel known as the mitochondrial permeability transition (MPT) pore (142–145). When this channel opens for prolonged periods it leads to mitochondrial depolarization, swelling, and membrane disruption and can ultimately lead to cell death. One current model for the MPT suggests that it occurs via abnormal changes in ATP synthase conformation (142) (FIGURE 3). Mg2+ typically binds at the interface between the α- and β-subunits of the soluble F1 module, helping to coordinate ATP or ADP within the nucleotide-binding pocket. In the ATP synthase model of the MPT, at high concentrations Ca2+ displaces Mg2+ at this site, leading to the transmission of abnormal strain to the transmembrane FO module and opening of a large-pore channel within the ring formed by c subunits. This model has received some support in a recent preprint showing Ca2+ binding to the F1 module β-subunit (110).
Outstanding question: Because opening of the MPT appears to be one of the earliest irreversible events in the pathophysiology of injury during heart attacks, strokes, and other forms of ischemic injury, deciphering its mechanism remains one of the most significant unanswered questions in mitochondrial Ca2+ signaling. How Ca2+ produces MPT through the ADP-ATP translocase, the other major complex involved, has not been resolved (146, 147).
Mitochondrial Ca2+ Transport Regulation of OXPHOS, Oxidative Stress, and Mitochondrial Dysfunction in Heart Disease
Ischemia-Reperfusion Injury and Systolic Heart Failure
Acute cardiac injury occurs most frequently during acute coronary syndromes, leading to ischemia-reperfusion injury. Both this type of insult as well as damage caused by other processes such as hypertension (excessive neurohormonal signals), drugs, or genetic disorders can lead to chronic systolic heart failure, characterized by impaired cardiac contractile function. In rodent models, these pathologies have been investigated by either injuring the heart through nongenetic manipulations or studying transgenic animals with disease-causing mutations. In studies of nongenetic heart disease, three rodent models have been used repeatedly: 1) acute ischemia-reperfusion, 2) chronic pressure-overload-induced heart failure, and 3) chronic β-adrenergic stress-induced heart failure. The phenotypes associated with uniporter deletion in these models of heart disease have been reviewed recently (148–151). Here, we focus on four interrelated mechanistic concepts that can be generalized from investigating mitochondrial Ca2+ uptake in these heart injury disease models (FIGURE 4).
FIGURE 4.
Mitochondrial Ca2+ pathways involved in health and disease A: by regulating multiple enzymes involved in metabolite transport, the tricarboxylic acid (TCA) cycle, and the electron transport chain (ETC), Ca2+ allows rapid matching of ATP supply to demand. B: Ca2+ overload produces permeability transition, in which the ATP synthase, ATP/ADP translocase (ANT), and mitochondrial Ca2+ uniporter (MCU) regulator 1 (MCUR1) are implicated. C: by boosting flux through the TCA cycle, Ca2+ can produce both prooxidant effects [via NADH oxidation and reactive oxygen species (ROS) leak through Complex I] and antioxidant effects [enhancing the supply of NADPH by indirectly increasing substrates to malic enzyme 3 (ME3), IDH2 (isocitrate dehydrogenase 2), and nicotinamide nucleotide transhydrogenase (NNT)]. The balance of these effects determines the ultimate consequence of Ca2+ on oxidative stress. D: metabolism can be altered differently depending on the chronicity of uniporter inhibition. Acute inhibition abrogates supply-demand matching and Ca2+ overload during ischemia-reperfusion injury. However, chronic inhibition can potentially activate other cell death pathways, cause compensation of Ca2+ influx by reversing the direction of Ca2+/H+ and Na+/Ca2+ exchangers, or cause the loss of protective uniporter functions such as its interaction with Complex I. All of these chronic mechanisms can offset the benefit on Ca2+ overload of uniporter inhibition and may underlie the benefits of boosting uniporter activity in chronic cardiac disease. GSSH, reduced glutathione.
Concept 1: Ca2+-dependent enhancement of OXPHOS in heart injury.
First, mitochondrial Ca2+ uptake during cardiac function appears to be important for matching ATP supply and demand, via the effects of Ca2+ on OXPHOS and fuel choice described in ca2+ in mitochondrial metabolism. Such contributions to energetic homeostasis are most evident when cardiac workload increases, since mice lacking uniporter components are unable to rapidly increase heart rate or cardiac contractility during pharmacological or physical stress or exercise (105, 107, 108). In addition, overexpression of the inhibitory uniporter subunit MCUB produces a transient cardiomyopathy, suggesting that acute loss of Ca2+ uptake prevents supply-demand matching (13).
As heart failure progresses, energy supply fails to meet demand, and individuals with more severe supply-demand mismatch have reduced survival (152). Therefore, overexpression of MCU has been suggested as a potential therapy for the failing heart. Of note, in several of the studies listed below cardiac energetics was only partly investigated. However, global changes in ATP are a late manifestation of energetic dysfunction, whereas one of the earliest markers of cardiac energetic failure is diminished activity of SERCA, the protein responsible for reloading the SR with Ca2+ (153). Therefore, finding decreases in SR Ca2+ leak and rescue of SR Ca2+ uptake are useful surrogates for improved cardiac function during heart failure. In a guinea pig pressure-overload model of heart failure, acute viral transduction of MCU increased uniporter levels by ∼50% and improved the contractile response. Although the consequence to energetics was not directly measured, it is notable that oxidative stress was decreased and Ca2+ leak from the SR was diminished (154). Furthermore, in experiments where heart failure was induced in mice by chronic isoproterenol infusion, which leads to excessive β-adrenergic stimulation, overexpressing MCU fourfold with an inducible transgene approach also ameliorated the decline in cardiac function (155). Critically, the increase in Ca2+ uptake was associated with improved Ca2+ uptake into the SR, suggesting that MCU overexpression improved cellular energy homeostasis in this heart failure model. Taken together, these studies suggest that modest increases in mitochondrial Ca2+ uptake can help boost cardiac energy synthesis and preserve function during heart failure. What remains to be established in these scenarios is to what degree the changes in metabolism occur at the level of ATP synthesis/OXPHOS, altered fuel choice via changes to pyruvate dehydrogenase activity, or the alteration of other matrix targets of Ca2+.
Concepts 2 and 3: Ca2+-dependent injury via the regulation of reactive oxygen species and the mitochondrial permeability transition.
In contrast to the potential beneficial effects of Ca2+ on energetics in heart failure, in acute cardiac injury the greater concern is mitochondrial Ca2+ overload, via two toxic downstream mechanisms involving the OXPHOS machinery. These are the excessive production of ROS and the potential activation of the MPT. Ca2+-induced activation of the MPT has been discussed above (in F1-FO-ATP synthase). Ca2+-induced boosting of ROS can occur via its effects on NADH production. During ischemic injury, hypoxia leads to stalling of forward flow of electrons down the ETC, and these can leak out as ROS, particularly during reperfusion, when mitochondrial dysfunction leads to reverse electron flow through Complex I (156). Concurrently, cytoplasmic Ca2+ overload occurs because of the failure of Ca2+ recycling into the SR and increased permeability of the plasma membrane (153, 157, 158) and subsequently enters the matrix via the uniporter to produce mitochondrial Ca2+ overload (159). Thus the combined effect of increased Ca2+-mediated NADH production and dysfunctional ETC activity is to increase pathological ROS production. Because of the combined effects on pathological ROS and MPT, mice with whole body or cardiac-specific knockout of MCU, pharmacological inhibition of MCU, or cardiac deletion of EMRE all had reduced cardiomyocyte necrosis during experimental ischemia-reperfusion injury (105, 107, 109, 159, 160). The benefit of preventing Ca2+ overload could also be replicated by increasing NCLX expression approximately threefold, which increased mitochondrial Ca2+ efflux and improved survival (16). Inhibition of MCU and EMRE and enhancement of NCLX have all been associated with reduced susceptibility to the MPT in these cardiac models, and less ROS generation has been noted in MCU KO mouse embryonic fibroblasts subject to ischemia-reperfusion injury (16, 105, 107). Moreover, ischemic cardiac injury leads to endogenous downregulation of mitochondrial Ca2+ uptake, as multiple studies have shown that the inhibitory uniporter subunit MCUB, usually poorly expressed in the heart, is substantially induced in the postischemic period (13, 14). Consistent with this effect, transgenic overexpression of MCUB reduces the degree of injury following ischemia-reperfusion, whereas deletion of MCUB increases both injury and susceptibility to MPT (13, 14).
Nevertheless, the effects of Ca2+ on ROS are not entirely straightforward, as Ca2+ can also indirectly contribute to the antioxidant response (8). For example, increasing matrix Ca2+ via inhibition of NCLX appears to prevent the production of pathological ROS in failing cardiomyocytes (161). The mitochondrial glutathione and thioredoxin-dependent enzymes involved in the antioxidant response rely on the NADPH/NADP+ redox potential (149). Mitochondrial Ca2+ can indirectly contribute to boosting this potential by its effects on the TCA cycle. The TCA cycle metabolites malate and isocitrate are substrates for NADP+-dependent malic enzyme (ME3) and isocitrate dehydrogenase 2 (IDH2) (149, 162), both of which have robust cardiac expression. Though these enzymes are not known to bind Ca2+ directly, increases in their substrates via Ca2+-induced boosting of TCA cycle flux can lead to increased NADPH and subsequent antioxidant pathways. Similarly, enhanced Ca2+-induced NADH production can be interconverted to NADPH via the nicotinamide nucleotide transhydrogenase (NNT), a ΔΨ-dependent transmembrane protein. In pressure-overload heart failure models, it appears that reversed NNT activity depletes NADPH to maintain the NADH pool (163), and thus Ca2+ boosting of NADH production may also contribute to maintaining NADPH levels via this mechanism. The loss of this protective mechanism may offset the benefit of preventing Ca2+ overload in pressure-overload models and explain why deletion of the uniporter has no effect on survival in this form of myocardial injury (103, 105). In contrast, as discussed above, modest MCU overexpression may in fact prove beneficial in pressure-overload models, implying that the benefit conferred to energetics and redox homeostasis may outweigh concerns about Ca2+ overload (154).
Outstanding questions: Since mitochondrial Ca2+ can indirectly promote both an antioxidant response via the contribution to NADPH as well as a prooxidant response via NADH oxidation at Complex I, what factors contribute to the balance of these effects in health and disease? Furthermore, there has been only limited investigation into whether Ca2+ can directly regulate the enzymes involved in the mitochondrial antioxidant response (164).
Concept 4: Chronic uniporter inhibition may affect metabolism beyond mitochondrial Ca2+ uptake.
The final mechanistic concept arising from studies investigating therapeutic depletion of the uniporter is perhaps the most intriguing: the effects vary greatly depending on the timing of the uniporter inhibition relative to the onset of injury. In the studies listed above, acute deletion of MCU or EMRE in the weeks before ischemia-reperfusion injury led to improved outcomes in uniporter-inhibited animals. In contrast, germline and whole body deletion of uniporter components have consistently shown minimal effect on outcomes, despite loss of mitochondrial Ca2+ uptake, in the same ischemia-reperfusion models (103, 106, 108, 109, 148). In a recent study, these differential effects were shown to be dependent on the chronicity of uniporter loss relative to the injury. The authors induced cardiac-specific loss of the EMRE subunit with tamoxifen and studied the effects of heart injury either 3 wk or 3 mo after the deletion. Whereas loss of EMRE was protective when injury followed shortly thereafter, the effects were absent when chronic loss of EMRE preceded the injury.
Outstanding question: The mechanisms for the adaptation to uniporter loss remain to be established. It appears that either 1) other mechanisms compensate for the loss of uniporter-mediated mitochondrial Ca2+ uptake or 2) the benefits of preventing mitochondrial Ca2+ overload are offset in the long term by detrimental effects of absent Ca2+ influx on energetics. Evidence for both these possibilities exists, as alterations in the activity of NCLX are seen in in vitro models of ischemia-reperfusion injury (165), suggesting it may compensate as an alternative Ca2+ flux pathway (though see Ref. 104), whereas chronic loss of the uniporter leads to reductions in Complex I levels and an increase in mitochondrial oxidative modifications, suggesting potential harm from chronic uniporter loss (89, 109, 165).
Arrhythmias Induced by Heart Injury
Beyond assessment of overall cardiac function, investigations have also focused on the susceptibility to arrhythmias, a separate mechanism for cardiac death. In several heart failure models, regulation of NCLX function appears critical for arrhythmia generation (FIGURE 5A). Both acute and chronic pharmacological inhibition of NCLX by transfusion of the inhibitor CGP-37157 reduced ventricular tachyarrhythmias in failing hearts. In one study, acute inhibition of NCLX in a pig coronary ischemia-reperfusion model resulted in suppression of ventricular tachyarrhythmias, though no effect was seen on cardiac infarct size (166). Similarly, in a study of guinea pigs subjected to combined aortic constriction and chronic β-adrenergic stress, control animals developed heart failure and fatal tachyarrhythmias, whereas chronic infusion of CGP-37157 led to a suppression of both functional decline as well as arrhythmias (161). The mechanism was presumed to be the prevention of excess Ca2+ depletion, leading to beneficial effects on ATP production via classical allosteric mechanisms, though these were not tested directly. Notably, in the guinea pig study, prolonged cardiac stress leads to cytoplasmic Na+ overload, which is then imported into mitochondria via NCLX in exchange for Ca2+, leading to Ca2+ depletion. In this model, cardiac deficits result from insufficient Ca2+ to support NADH production through the TCA cycle. The increased NAD+-to-NADH ratio impairs electron flow both for ATP production (via the electron transport chain) as well as for interconversion to NADPH, which is critical for the redox response, as discussed above (Concepts 2 and 3: Ca2+-dependent injury via the regulation of reactive oxygen species and the mitochondrial permeability transition). Thus, inhibiting NCLX preserves Ca2+ for both energetics and combating oxidative stress. However, the therapeutic window for this intervention may well be limited, as loss of NCLX is lethal (16).
FIGURE 5.
Mitochondrial Ca2+ alterations involved in cardiac disease A and B: contributions to arrhythmogenesis after cardiac injury. A: depletion of mitochondrial Ca2+ by excessive Na+/Ca2+ exchange can lead to energetic impairment via insufficient NADH production and promote arrhythmias. B: Ca2+ overload in states of overnutrition promotes the production of mitochondrial reactive oxygen species (mROS), promoting oxidative injury to ryanodine receptor (RyR)2, diastolic Ca2+ leak, action potential (AP) lengthening, and consequent arrhythmias. C: the balance between the effects described in A and B may also determine whether mitochondrial Ca2+ promotes or inhibits arrhythmias in CPVT. D: in mitochondrial cardiomyopathies, injury to Complex I prevents interaction and oxidation of MCU (mitochondrial Ca2+ uniporter), leading to buildup in uniporter levels and subsequent energetic compensation. E: cardiolipin (blue) can bind within a fenestration between the MCU transmembrane domains. Loss of cardiolipin in Barth syndrome may alter MCU structure to inhibit interaction with accessory subunits and prevent Ca2+ influx. IM, inner mitochondrial membrane; IMS, intermembrane space.
Arrhythmias induced by diabetes or overnutrition represent another area where mitochondrial Ca2+ may be contributing by exerting prooxidant effects (FIGURE 5B). In a mouse model of diabetes induced by high-fat diet, excess mitochondrial ROS was noted and led to oxidation of RyR2 receptors and diastolic Ca2+ leak, which can prolong the cardiac action potential and trigger arrhythmias (167). In a prediabetic model induced by chronic fructose feeding, pharmacological uniporter enhancement exacerbated arrhythmias, whereas MCU ablation reduced them (168). In another study, control and MCU KO cardiomyocytes were exposed to palmitic acid as a model for metabolic stress. Whereas the control cardiomyocytes had increases in mitochondrial Ca2+, ROS, and depolarization of mitochondrial membranes, these changes were abrogated in MCU KO cells. In animals exposed to a high-fat diet for 4 wk, cardiomyocytes also had prolonged action potentials and easily induced ventricular tachyarrhythmias, whereas cardiac-specific deletion of MCU abrogated these changes (169). However, not all investigations have been consistent with mitochondrial Ca2+ promoting these pathologies. In particular, in a study of diabetes induced by overexpression of human amylin in rat pancreas, mitochondrial Ca2+ levels were low and pharmacological inhibition of NCLX prevented arrhythmias (170). Nevertheless, the bulk of studies suggest that increases in mitochondrial Ca2+ may have proarrhythmic effects in diabetes/overnutrition.
The effects of mitochondrial Ca2+ on cardiac arrhythmias in heart failure versus diabetes/obesity models are strikingly discordant. In the heart failure models, mitochondrial Ca2+ appears to decreasing ROS by its effects on NAD(P)H, whereas in the overfeeding or diabetic models, mitochondrial Ca2+ appears to boost ROS through Complex I.
Outstanding question: As discussed above, this raises the question of what mechanisms alter the balance between the pro- and antioxidant effects of mitochondrial Ca2+ and whether Ca2+ contributes to these beyond the established mechanisms.
Catecholaminergic Polymorphic Ventricular Tachycardia
Catecholaminergic polymorphic ventricular tachycardia (CPVT) is a rare inherited arrhythmia syndrome that can cause sudden cardiac death in young people. It is characterized by exercise- or emotional stress-induced bidirectional or polymorphic ventricular tachycardia in the setting of a structurally normal heart and a normal ECG. CPVT is caused by mutations affecting proteins that regulate cardiomyocyte SR Ca2+ levels, most commonly RyR2. The resulting intracellular Ca2+ overload leads to delayed afterdepolarizations and triggered activity, which can induce ventricular tachycardia and fibrillation (171, 172).
Given the centrality of Ca2+ transfer between RyRs and mitochondria for energy supply-demand matching, several studies have investigated whether aberrant Ca2+ release from CPVT RyR2 mutants affects mitochondrial function (FIGURE 5C). In one study (173), RyR Ca2+ leak induced by caffeine led to increased mitochondrial ROS production, setting up a feedback loop where mitochondria-derived ROS promoted further Ca2+ leak via RyR oxidation. Attenuating the mitochondrial production of ROS and inhibiting mitochondrial Ca2+ uptake by viral expression of dominant-negative MCU both decreased arrhythmogenesis parameters in isolated cardiomyocytes, whereas cardiomyocytes from CPVT mice had elevated expression of the inhibitory MCUB subunit. How such interventions may affect mitochondrial Ca2+-dependent energetics was not explored but remains an intriguing question, because triggered activity does not always lead to clinical arrhythmias. That such a disconnect may exist between ROS promotion of triggered activity versus altered energetics promotion of an arrhythmogenic substrate comes from contrasting data on drugs that promote, rather than inhibit, mitochondrial Ca2+ uptake. In two studies in zebrafish models of CPVT, enhancement of SR-to-mitochondrial Ca2+ transfer, via either the VDAC2 agonist efsevin or the MCU agonists ezetimibe and disulfiram, restored normal cardiac rhythms by more than threefold (174, 175). Strategies that promote SR-to-mitochondria Ca2+ transfer were also successful in preventing arrhythmias in live RyR2R4496C/WT mice. When efsevin, the VDAC2 agonist, or kaempferol, a weak MCU agonist, was infused via osmotic minipumps in this model, there was a ∼40% reduction in the number of mice with ventricular tachyarrhythmias, an effect also replicated in cardiomyocytes derived from induced pluripotent stem cells (176). Surprisingly, boosting MCU Ca2+ uptake with kaempferol actually reduced ROS in the calsequestrin KO model of CPVT, an effect in part attributed to global increases in the ability of mitochondria to buffer Ca2+ (168). Conversely, inhibiting MCU with a tamoxifen-inducible knockout drastically worsened arrhythmias, leading to rapid lethality (168). The evidence for Ca2+ overload as a mediator of injury in CPVT is also somewhat controversial, as calsequestrin KOs that were crossed with mice with inhibited MPT demonstrated increased ROS production and an increased propensity to arrhythmias (177).
Outstanding question: How to reconcile these studies with surprisingly opposite outcomes? One possibility is the use of pharmacological versus genetic manipulations, though it should be noted that even genetic manipulations led to apparently opposite outcomes [decreased arrhythmogenesis with a dominant-negative MCU in rat ventricular myocytes vs increased lethality with MCU KO (168, 173)]. An intriguing possibility is the disconnect between in vitro phenotypes, which have variably shown both worsened and improved arrhythmogenesis and susceptibility to Ca2+ overload, versus whole animal outcomes, which have consistently shown improvements in outcomes with Ca2+ uniporter enhancement. In vitro assays typically measure parameters in isolated cardiomyocytes without energetic or contractile resistance, as occurs in vivo. Therefore, it may be that the Ca2+ contribution to cardiac energetic homeostasis, poorly evaluated in vitro, may be a key determinant to long-term benefit during dysregulated SR Ca2+ release in CPVT.
Mitochondrial Cardiomyopathies
Energetic failure can be a primary cause of cardiomyopathy in mitochondrial diseases. Such diseases involve deficient oxidative phosphorylation (OXPHOS) and arise from mutations in mitochondrial proteins encoded by either the nuclear or the mitochondrial genome (mtDNA) (178, 179). These diseases typically occur in infants and children, with a prevalence of 1 in 5,000 (180). Because cardiac contraction requires high energy supply, cardiac dysfunction occurs in 20–40% of children with mitochondrial diseases and increases their mortality nearly threefold (181). In these disorders, the reduction in OXPHOS produces heart failure due to the energy supply-demand mismatch.
Investigators have previously studied OXPHOS deficiency caused by mutations (118, 182–186), deletion of Tfam (the transcription factor for mtDNA) (187), or chemical mtDNA depletion (188). The typical finding is reduced or unchanged Ca2+ uptake in the presence of diminished ΔΨ, though increased Ca2+ was seen in skeletal muscle (187). Diminished ΔΨ correlated with disease severity. This presents a technical problem in assessing mitochondrial Ca2+ transport pathways. Ca2+ influx through the uniporter is driven by ΔΨ, so a change to either ΔΨ or uniporter activity can alter the magnitude of Ca2+ influx. Imaging assays are the most commonly used method to measure mitochondrial Ca2+ influx, but they cannot distinguish whether changes in flux are due to altered ΔΨ or uniporter activity. Analysis becomes especially difficult when these two factors change in opposite directions. In OXPHOS-deficient mitochondria, precisely such diminished ΔΨ may mask any compensatory increases in uniporter activity in typical in vitro imaging assays. We used complementary imaging and electrophysiological assays to measure uniporter activity directly in mice with cardiac-specific deletion of Tfam (Tfam cKO) (117), a rodent model that best mimics pediatric mitochondrial cardiomyopathies (189, 190). These animals had substantial reductions in cardiac mtDNA-encoded OXPHOS subunits and ETC activity. In these mice, we noted a doubling in the rate of Ca2+ uptake and a three- to fivefold increase in the size of mitochondrial Ca2+ currents, associated with a doubling of protein, but not transcript, levels of multiple uniporter subunits. We found that these changes in Ca2+ influx boosted mitochondrial respiration and ATP synthesis. In the absence of Ca2+, oxygen consumption rates in Tfam cKO mitochondria were half of wild-type values, but incubation in Ca2+ restored these rates to near control levels. The mechanism for such energetic compensation was CLIPT, the regulation of protein lifetime by Complex I described above in Complex I (NADH:ubiquinone oxidoreductase) (FIGURE 5D). These studies reinforce the concept that mitochondrial Ca2+ can be a critical compensatory contributor to energetic homeostasis during ETC injury.
Outstanding question: Given the benefits afforded by mitochondrial Ca2+ in these diseases, are there therapeutic approaches that can maximize the stability of uniporter channels without overactivating oxidant responses or triggering MPT? In general, such an approach may be applicable not only to these diseases but to a range of the other cardiac disorders described here.
Barth Syndrome
Barth syndrome is a rare X-linked genetic disorder characterized by cardiomyopathy, skeletal myopathy, growth retardation, neutropenia, and elevated urinary excretion of 3-methylglutaconic acid (3-MGA) (191). The disease results from mutations in the tafazzin gene (TAZ). Tafazzin plays a crucial role in the remodeling of cardiolipin, a unique phospholipid predominantly found in the inner mitochondrial membrane. Mutations in tafazzin lead to reductions in cardiolipin levels compared with precursors. Because cardiolipin binds to multiple mitochondrial membrane proteins, including components of the ETC, this imbalance disrupts mitochondrial bioenergetics and membrane integrity (191, 192). In a mouse model of Barth syndrome generated by RNA interference against Taz, a systolic cardiomyopathy develops over 10–20 wk, leading to reductions in OXPHOS (193). In hearts from these mice, as well as human cardiac tissue and inducible pluripotent stem cell-derived cardiomyocytes (CM-IPSCs) obtained from patients with Barth syndrome, there was >50% reduction in MCU expression, along with a modest ∼10% increase in MCUB levels (193). In separate studies, loss of cardiolipin also affected the stability of MICU1 posttranscriptionally in several model systems (194, 195). The mechanism for loss of MCU appears to be direct, with structures of the uniporter revealing cardiolipin molecules buried in the fenestration between MCU transmembrane domains (49) (FIGURE 5E). Notably, the lack of mitochondrial Ca2+ uptake produced pronounced effects during increases in cardiac workload, leading to diminished respiratory capacity, excess oxidation of NAD(P)H, and insufficient ATP generation (193). In contrast to other examples of dysregulated cardiac mitochondrial Ca2+ homeostasis, there were no obvious differences in ROS measurements, likely because of an upregulation of endogenous ROS metabolizers. [It is worth noting that in a separate study using cardiac-specific Taz knockout an increase in mitochondrial ROS, mediated by excessive RyR leak and elevated diastolic Ca2+, was found (196).] To restore mitochondrial Ca2+ levels, investigators used in vivo pharmacological inhibition of Na+/Ca2+ exchange. Although this reversed arrhythmic parameters, the inotropic response remained blunted. This modest recovery engendered by mitochondrial Ca2+ boosting is expected, since cardiolipin binding regulates not just MCU but also multiple components of the ETC (133, 134). However, given the known differential localization of the uniporter relative to NCLX, this result also may further support the tenet that mitochondrial Ca2+ regulation of OXPHOS may be a highly local phenomenon, not adequately reflected in global assessments of mitochondrial Ca2+ concentrations.
Friedreich’s Ataxia
Friedreich’s ataxia is a peripheral and central neurodegenerative disorder that frequently affects the heart, producing a hypertrophic cardiomyopathy. In this disorder, triplet expansion repeats within the frataxin (FXN) gene cause mitochondrial dysfunction, primarily affecting iron-sulfur cluster biogenesis and heme production. Deficiencies in these central processes disrupt cellular iron homeostasis, impair ATP synthesis, and cause substantial oxidative stress (197, 198). Initial studies in cultured SH-SY5Y (neuroblastoma-like) cells revealed that mitochondrial depolarization and the primary energetic deficiency prevented efficient mitochondrial Ca2+ uptake, though whether this contributed to energetic deficiency was not explored (199). Such results were replicated in a separate study, directly revealing diminished Ca2+ uptake into isolated FTX-deficient SH-SY5Y cells (200). Here, the decreased accumulation of Ca2+ was attributed to dysregulation in ER-mitochondrial contacts, perhaps directly by a variant form of frataxin. Artificially restoring these contacts rescued mitochondrial Ca2+ uptake. In a further study, neonatal rat cardiomyocytes acutely depleted of frataxin developed increased sensitivity to Ca2+ overload, with swollen mitochondria (201). As in SH-SY5Y cells (200), MCU levels were unchanged. However, in the cardiomyocytes, NCLX levels were decreased, implying that mitochondrial Ca2+ overload may lead to toxicity via the MPT. In fact, global chelation of Ca2+ or treatment with cyclosporine A, a MPT inhibitor, reduced mitochondrial swelling in these cardiomyocytes (201).
Outstanding questions: What remains unknown is whether mitochondrial Ca2+ levels are increased in FXN-deficient cardiomyocytes and whether these produce compensatory boosts to energetics during disease. It is notable that treatment with Ca2+ chelation had a less robust phenotype compared to inhibition of the MPT, which may suggest that modest increases in Ca2+ may produce salutary effects on mitochondrial energetics before the onset of overload and MPT. In fact, when Ca2+ uptake in neurons was enhanced in fruit flies deficient in FTX via overexpression of MCU, locomotor function and survival increased substantially (200). Another interesting point is that interventions to protect mitochondria, such as maintaining organisms under chronic hypoxia, reverse oxidative stress signaling and neurological deficits but fail to affect cardiomyopathy (202), suggesting that the energetic deficits are a key factor for this organ and compensatory Ca2+ signaling may offer an independent means of promoting survival.
Future Directions and Conclusions
Because most of the studies conducted on Ca2+ regulation of OXPHOS have been conducted in the heart and skeletal muscle, investigators need to establish whether these mechanisms are more generally conserved across organs or if novel tissue-specific Ca2+-dependent pathways exist. Moreover, although knockout studies have provided valuable insights, the lack of highly specific compounds for acute manipulation of mitochondrial Ca2+ levels hinders investigations during physiological activities. If these can be combined with novel sensors for Ca2+, ATP, NAD+, and other metabolites, multimodal assessments will reveal how physiological outcomes correlate with changes in critical mitochondrial parameters.
Additionally, emerging technologies are poised to deepen our understanding of mitochondrial Ca2+ regulation. The use of isotope-labeled tracers will hopefully offer more precise assessments of how Ca2+ influences metabolic fluxes and enzymatic activity, beyond global measurement of respiration. In addition, a recent preprint appears poised to greatly expand the repertoire of targets of mitochondrial Ca2+ regulation (110). In this study, assessment of proteomic thermostability in the presence or absence of Ca2+ identified >200 potentially Ca2+-sensitive mitochondrial proteins, including 11 Complex I, 4 Complex III, 1 Complex IV, and 8 F1-FO-ATP synthase subunits, along with 2 Complex II assembly factors. Moreover, Ca2+ binding to 2,4-Dienoyl-CoA Reductase 1, a rate-limiting enzyme in mitochondrial polyunsaturated fatty acid oxidation, suggests that Ca2+ effects on fuel substrate choice may extend to lipids as well. Such discoveries are overturning the dogma of limited effects of mitochondrial Ca2+ on metabolism.
In summary, mitochondrial Ca2+ is essential for the cardiac response to physiological and pathological stress. Although it enhances ATP production and shifts the balance in mitochondrial fuel preference, its indirect effects on OXPHOS are also emerging. Recent discoveries of binding partners between ETC complexes and Ca2+ regulators provide insights into cellular management of mitochondrial Ca2+. As with cytoplasmic Ca2+ regulation, a central principle of mitochondrial Ca2+ signaling is that these effects are highly localized. Ultimately, understanding how Ca2+ governs OXPHOS may have therapeutic implications, since mitochondrial damage is a frequent pathophysiological event across organs and diseases.
Acknowledgments
Support is from the National Institutes of Health (HL165797 to D.C., HL141353 to D.C., HL165806 to S.H.J.L.) and the Nora Eccles Treadwell Foundation (to D.C.).
The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institutes of Health.
No conflicts of interest, financial or otherwise, are declared by the authors.
S.H.J.L., N.K.R., D.M.M., and D.C. prepared figures; E.B., S.H.J.L., H.E.D., and D.C. drafted manuscript; E.B., S.H.J.L., N.K.R., H.E.D., and D.C. edited and revised manuscript; E.B., S.H.J.L., N.K.R., H.E.D., and D.C. approved final version of manuscript.
References
- 1. Clapham DE. Calcium signaling. Cell 131: 1047–1058, 2007. doi: 10.1016/j.cell.2007.11.028. [DOI] [PubMed] [Google Scholar]
- 2. Fernandez-Sanz C, De la Fuente S, Sheu SS. Mitochondrial Ca2+ concentrations in live cells: quantification methods and discrepancies. FEBS Lett 593: 1528–1541, 2019. doi: 10.1002/1873-3468.13427. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Lu X, Ginsburg KS, Kettlewell S, Bossuyt J, Smith GL, Bers DM. Measuring local gradients of intramitochondrial [Ca2+] in cardiac myocytes during sarcoplasmic reticulum Ca2+ release. Circ Res 112: 424–431, 2013. doi: 10.1161/CIRCRESAHA.111.300501. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Marban E, Rink TJ, Tsien RW, Tsien RY. Free calcium in heart muscle at rest and during contraction measured with Ca2+-sensitive microelectrodes. Nature 286: 845–850, 1980. doi: 10.1038/286845a0. [DOI] [PubMed] [Google Scholar]
- 5. Tsien RY, Rink TJ. Neutral carrier ion-selective microelectrodes for measurement of intracellular free calcium. Biochim Biophys Acta 599: 623–638, 1980. doi: 10.1016/0005-2736(80)90205-9. [DOI] [PubMed] [Google Scholar]
- 6. Glancy B, Balaban RS. Role of mitochondrial Ca2+ in the regulation of cellular energetics. Biochemistry 51: 2959–2973, 2012. doi: 10.1021/bi2018909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Williams GS, Boyman L, Lederer WJ. Mitochondrial calcium and the regulation of metabolism in the heart. J Mol Cell Cardiol 78: 35–45, 2015. doi: 10.1016/j.yjmcc.2014.10.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Kohlhaas M, Nickel AG, Maack C. Mitochondrial energetics and calcium coupling in the heart. J Physiol 595: 3753–3763, 2017. doi: 10.1113/JP273609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9. Zeth K, Zachariae U. ten years of high resolution structural research on the voltage dependent anion channel (VDAC)–recent developments and future directions. Front Physiol 9: 108, 2018. doi: 10.3389/fphys.2018.00108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Rosencrans WM, Rajendran M, Bezrukov SM, Rostovtseva TK. VDAC regulation of mitochondrial calcium flux: from channel biophysics to disease. Cell Calcium 94: 102356, 2021. doi: 10.1016/j.ceca.2021.102356. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Kamer KJ, Mootha VK. The molecular era of the mitochondrial calcium uniporter. Nat Rev Mol Cell Biol 16: 545–553, 2015. doi: 10.1038/nrm4039. [DOI] [PubMed] [Google Scholar]
- 12. Balderas E, Sommakia S, Eberhardt D, Lee S, Chaudhuri D. The structural era of the mitochondrial calcium uniporter. In: Calcium Signals. Bristol, UK: IOP Publishing, 2023, p. 12–1–12-22. [Google Scholar]
- 13. Lambert JP, Luongo TS, Tomar D, Jadiya P, Gao E, Zhang X, Lucchese AM, Kolmetzky DW, Shah NS, Elrod JW. MCUB regulates the molecular composition of the mitochondrial calcium uniporter channel to limit mitochondrial calcium overload during stress. Circulation 140: 1720–1733, 2019. doi: 10.1161/CIRCULATIONAHA.118.037968. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Huo J, Lu S, Kwong JQ, Bround MJ, Grimes KM, Sargent MA, Brown ME, Davis ME, Bers DM, Molkentin JD. MCUb induction protects the heart from postischemic remodeling. Circ Res 127: 379–390, 2020. doi: 10.1161/CIRCRESAHA.119.316369. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Palty R, Silverman WF, Hershfinkel M, Caporale T, Sensi SL, Parnis J, Nolte C, Fishman D, Shoshan-Barmatz V, Herrmann S, Khananshvili D, Sekler I. NCLX is an essential component of mitochondrial Na+/Ca2+ exchange. Proc Natl Acad Sci USA 107: 436–441, 2010. doi: 10.1073/pnas.0908099107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Luongo TS, Lambert JP, Gross P, Nwokedi M, Lombardi AA, Shanmughapriya S, Carpenter AC, Kolmetzky D, Gao E, van Berlo JH, Tsai EJ, Molkentin JD, Chen X, Madesh M, Houser SR, Elrod JW. The mitochondrial Na+/Ca2+ exchanger is essential for Ca2+ homeostasis and viability. Nature 545: 93–97, 2017. doi: 10.1038/nature22082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Jiang D, Zhao L, Clapham DE. Genome-wide RNAi screen identifies Letm1 as a mitochondrial Ca2+/H+ antiporter. Science 326: 144–147, 2009. doi: 10.1126/science.1175145. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Froschauer E, Nowikovsky K, Schweyen RJ. Electroneutral K+/H+ exchange in mitochondrial membrane vesicles involves Yol027/Letm1 proteins. Biochim Biophys Acta 1711: 41–48, 2005. [DOI] [PubMed] [Google Scholar]
- 19. Garbincius JF, Salik O, Cohen HM, Choya-Foces C, Mangold AS, Makhoul AD, Schmidt AE, Khalil DY, Doolittle JJ, Wilkinson AS, Murray EK, Lazaropoulos MP, Hildebrand AN, Tomar D, Elrod JW. TMEM65 regulates NCLX-dependent mitochondrial calcium efflux (Preprint). bioRxiv 2023.10.06.561062, 2023. doi: 10.1101/2023.10.06.561062. [DOI]
- 20. Vetralla M, Wischhof L, Cadenelli V, Scifo E, Ehninger D, Rizzuto R, Bano D, Stefani DD. TMEM65-dependent Ca2+ extrusion safeguards mitochondrial homeostasis (Preprint). bioRxiv 2023.10.10.561661, 2023. doi: 10.1101/2023.10.10.561661. [DOI]
- 21. Zhang Y, Reyes L, Sun J, Liu C, Springer D, Noguchi A, Aponte AM, Munasinghe J, Covian R, Murphy E, Glancy B. Loss of TMEM65 causes mitochondrial disease mediated by mitochondrial calcium (Preprint). bioRxiv 2022.08.02.502535, 2023. doi: 10.1101/2022.08.02.502535. [DOI]
- 22. Austin S, Mekis R, Mohammed SE, Scalise M, Wang WA, Galluccio M, Pfeiffer C, Borovec T, Parapatics K, Vitko D, Dinhopl N, Demaurex N, Bennett KL, Indiveri C, Nowikovsky K. TMBIM5 is the Ca2+/H+ antiporter of mammalian mitochondria. EMBO Rep 23: e54978, 2022. doi: 10.15252/embr.202254978. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Patron M, Tarasenko D, Nolte H, Kroczek L, Ghosh M, Ohba Y, Lasarzewski Y, Ahmadi ZA, Cabrera-Orefice A, Eyiama A, Kellermann T, Rugarli EI, Brandt U, Meinecke M, Langer T. Regulation of mitochondrial proteostasis by the proton gradient. EMBO J 41: e110476, 2022. doi: 10.15252/embj.2021110476. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Zhang L, Dietsche F, Seitaj B, Rojas-Charry L, Latchman N, Tomar D, Wüst RC, Nickel A, Frauenknecht KB, Schoser B, Schumann S, Schmeisser MJ, Vom Berg J, Buch T, Finger S, Wenzel P, Maack C, Elrod JW, Parys JB, Bultynck G, Methner A. TMBIM5 loss of function alters mitochondrial matrix ion homeostasis and causes a skeletal myopathy. Life Sci Alliance 5: e202201478, 2022. doi: 10.26508/lsa.202201478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Shankar TS, Ramadurai DK, Steinhorst K, Sommakia S, Badolia R, Thodou Krokidi A, Calder D, Navankasattusas S, Sander P, Kwon OS, Aravamudhan A, Ling J, Dendorfer A, Xie C, Kwon O, Cheng EH, Whitehead KJ, Gudermann T, Richardson RS, Sachse FB, Schredelseker J, Spitzer KW, Chaudhuri D, Drakos SG. Cardiac-specific deletion of voltage dependent anion channel 2 leads to dilated cardiomyopathy by altering calcium homeostasis. Nat Commun 12: 4583, 2021. doi: 10.1038/s41467-021-24869-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Deluca HF, Engstrom GW. Calcium uptake by rat kidney mitochondria. Proc Natl Acad Sci USA 47: 1744–1750, 1961. doi: 10.1073/pnas.47.11.1744. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Vasington FD, Murphy JV. Ca ion uptake by rat kidney mitochondria and its dependence on respiration and phosphorylation. J Biol Chem 237: 2670–2677, 1962. doi: 10.1016/S0021-9258(19)73805-8. [DOI] [PubMed] [Google Scholar]
- 28. Kirichok Y, Krapivinsky G, Clapham DE. The mitochondrial calcium uniporter is a highly selective ion channel. Nature 427: 360–364, 2004. doi: 10.1038/nature02246. [DOI] [PubMed] [Google Scholar]
- 29. Gunter KK, Gunter TE. Transport of calcium by mitochondria. J Bioenerg Biomembr 26: 471–485, 1994. doi: 10.1007/BF00762732. [DOI] [PubMed] [Google Scholar]
- 30. Csordas G, Golenar T, Seifert EL, Kamer KJ, Sancak Y, Perocchi F, Moffat C, Weaver D, Perez SF, Bogorad R, Koteliansky V, Adijanto J, Mootha VK, Hajnoczky G. MICU1 controls both the threshold and cooperative activation of the mitochondrial Ca2+ uniporter. Cell Metab 17: 976–987, 2013. doi: 10.1016/j.cmet.2013.04.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Garg V, Suzuki J, Paranjpe I, Unsulangi T, Boyman L, Milescu LS, Lederer WJ, Kirichok Y. The mechanism of MICU-dependent gating of the mitochondrial Ca2+ uniporter. eLife 10: e69312, 2021. doi: 10.7554/eLife.69312. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32. Paillard M, Csordas G, Szanda G, Golenar T, Debattisti V, Bartok A, Wang N, Moffat C, Seifert EL, Spat A, Hajnoczky G. Tissue-specific mitochondrial decoding of cytoplasmic Ca2+ signals is controlled by the stoichiometry of MICU1/2 and MCU. Cell Rep 18: 2291–2300, 2017. doi: 10.1016/j.celrep.2017.02.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Wescott AP, Kao JP, Lederer WJ, Boyman L. Voltage-energized calcium-sensitive ATP production by mitochondria. Nat Metab 1: 975–984, 2019. doi: 10.1038/s42255-019-0126-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Baughman JM, Perocchi F, Girgis HS, Plovanich M, Belcher-Timme CA, Sancak Y, Bao XR, Strittmatter L, Goldberger O, Bogorad RL, Koteliansky V, Mootha VK. Integrative genomics identifies MCU as an essential component of the mitochondrial calcium uniporter. Nature 476: 341–345, 2011. doi: 10.1038/nature10234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Chaudhuri D, Sancak Y, Mootha VK, Clapham DE. MCU encodes the pore conducting mitochondrial calcium currents. eLife 2: e00704, 2013. doi: 10.7554/eLife.00704. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. De Stefani D, Raffaello A, Teardo E, Szabò I, Rizzuto R. A forty-kilodalton protein of the inner membrane is the mitochondrial calcium uniporter. Nature 476: 336–340, 2011. doi: 10.1038/nature10230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. Perocchi F, Gohil VM, Girgis HS, Bao XR, McCombs JE, Palmer AE, Mootha VK. MICU1 encodes a mitochondrial EF hand protein required for Ca2+ uptake. Nature 467: 291–296, 2010. doi: 10.1038/nature09358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Sancak Y, Markhard AL, Kitami T, Kovacs-Bogdan E, Kamer KJ, Udeshi ND, Carr SA, Chaudhuri D, Clapham DE, Li AA, Calvo SE, Goldberger O, Mootha VK. EMRE is an essential component of the mitochondrial calcium uniporter complex. Science 342: 1379–1382, 2013. doi: 10.1126/science.1242993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. 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 32: 2362–2376, 2013. doi: 10.1038/emboj.2013.157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Plovanich M, Bogorad RL, Sancak Y, Kamer KJ, Strittmatter L, Li AA, Girgis HS, Kuchimanchi S, De Groot J, Speciner L, Taneja N, Oshea J, Koteliansky V, Mootha VK. MICU2, a paralog of MICU1, resides within the mitochondrial uniporter complex to regulate calcium handling. PLoS One 8: e55785, 2013. doi: 10.1371/journal.pone.0055785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Baradaran R, Wang C, Siliciano AF, Long SB. Cryo-EM structures of fungal and metazoan mitochondrial calcium uniporters. Nature 559: 580–584, 2018. doi: 10.1038/s41586-018-0331-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Fan C, Fan M, Orlando BJ, Fastman NM, Zhang J, Xu Y, Chambers MG, Xu X, Perry K, Liao M, Feng L. X-ray and cryo-EM structures of the mitochondrial calcium uniporter. Nature 559: 575–579, 2018. doi: 10.1038/s41586-018-0330-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Fan M, Zhang J, Tsai CW, Orlando BJ, Rodriguez M, Xu Y, Liao M, Tsai MF, Feng L. Structure and mechanism of the mitochondrial Ca2+ uniporter holocomplex. Nature 582: 129–133, 2020. doi: 10.1038/s41586-020-2309-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Nguyen NX, Armache JP, Lee C, Yang Y, Zeng W, Mootha VK, Cheng Y, Bai XC, Jiang Y. Cryo-EM structure of a fungal mitochondrial calcium uniporter. Nature 559: 570–574, 2018. doi: 10.1038/s41586-018-0333-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Wang Y, Nguyen NX, She J, Zeng W, Yang Y, Bai XC, Jiang Y. Structural mechanism of EMRE-dependent gating of the human mitochondrial calcium uniporter. Cell 177: 1252–1261.e13, 2019. doi: 10.1016/j.cell.2019.03.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Yoo J, Wu M, Yin Y, Herzik MA Jr, Lander GC, Lee SY. Cryo-EM structure of a mitochondrial calcium uniporter. Science 361: 506–511, 2018. doi: 10.1126/science.aar4056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. MacEwen MJ, Markhard AL, Bozbeyoglu M, Bradford F, Goldberger O, Mootha VK, Sancak Y. Evolutionary divergence reveals the molecular basis of EMRE dependence of the human MCU. Life Sci Alliance 3: e202000718, 2020. doi: 10.26508/lsa.202000718. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. Van Keuren AM, Tsai CW, Balderas E, Rodriguez MX, Chaudhuri D, Tsai MF. Mechanisms of EMRE-dependent MCU opening in the mitochondrial calcium uniporter complex. Cell Rep 33: 108486, 2020. doi: 10.1016/j.celrep.2020.108486. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. 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 12: 220–229, 2021. doi: 10.1007/s13238-020-00776-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Ryu SY, Beutner G, Dirksen RT, Kinnally KW, Sheu SS. Mitochondrial ryanodine receptors and other mitochondrial Ca2+ permeable channels. FEBS Lett 584: 1948–1955, 2010. doi: 10.1016/j.febslet.2010.01.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Bround MJ, Abay E, Huo J, Havens JR, York AJ, Bers DM, Molkentin JD. MCU-independent Ca2+ uptake mediates mitochondrial Ca2+ overload and necrotic cell death in a mouse model of Duchenne muscular dystrophy. Sci Rep 14: 6751, 2024. doi: 10.1038/s41598-024-57340-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Gottschalk B, Klec C, Leitinger G, Bernhart E, Rost R, Bischof H, Madreiter-Sokolowski CT, Radulović S, Eroglu E, Sattler W, Waldeck-Weiermair M, Malli R, Graier WF. MICU1 controls cristae junction and spatially anchors mitochondrial Ca2+ uniporter complex. Nat Commun 10: 3732, 2019. doi: 10.1038/s41467-019-11692-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. De La Fuente S, Fernandez-Sanz C, Vail C, Agra EJ, Holmstrom K, Sun J, Mishra J, Williams D, Finkel T, Murphy E, Joseph SK, Sheu SS, Csordás G. Strategic positioning and biased activity of the mitochondrial calcium uniporter in cardiac muscle. J Biol Chem 291: 23343–23362, 2016. doi: 10.1074/jbc.M116.755496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Carafoli E, Tiozzo R, Lugli G, Crovetti F, Kratzing C. The release of calcium from heart mitochondria by sodium. J Mol Cell Cardiol 6: 361–371, 1974. doi: 10.1016/0022-2828(74)90077-7. [DOI] [PubMed] [Google Scholar]
- 55. Crompton M, Capano M, Carafoli E. The sodium-induced efflux of calcium from heart mitochondria. Eur J Biochem 69: 453–462, 1976. doi: 10.1111/j.1432-1033.1976.tb10930.x. [DOI] [Google Scholar]
- 56. Takeuchi A, Matsuoka S. Spatial and functional crosstalk between the mitochondrial Na+-Ca2+ exchanger NCLX and the sarcoplasmic reticulum Ca2+ pump SERCA in cardiomyocytes. Int J Mol Sci 23: 7948, 2022. doi: 10.3390/ijms23147948. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57. Jiang D, Zhao L, Clish CB, Clapham DE. Letm1, the mitochondrial Ca2+/H+ antiporter, is essential for normal glucose metabolism and alters brain function in Wolf-Hirschhorn syndrome. Proc Natl Acad Sci USA 110: E2249–E2254, 2013. doi: 10.1073/pnas.1308558110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Doonan PJ, Chandramoorthy HC, Hoffman NE, Zhang X, Cárdenas C, Shanmughapriya S, Rajan S, Vallem S, Chen X, Foskett JK, Cheung JY, Houser SR, Madesh M. LETM1-dependent mitochondrial Ca2+ flux modulates cellular bioenergetics and proliferation. FASEB J 28: 4936–4949, 2014. doi: 10.1096/fj.14-256453. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Huang E, Qu D, Huang T, Rizzi N, Boonying W, Krolak D, Ciana P, Woulfe J, Klein C, Slack RS, Figeys D, Park DS. PINK1-mediated phosphorylation of LETM1 regulates mitochondrial calcium transport and protects neurons against mitochondrial stress. Nat Commun 8: 1399, 2017. doi: 10.1038/s41467-017-01435-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Tsai MF, Jiang D, Zhao L, Clapham D, Miller C. Functional reconstitution of the mitochondrial Ca2+/H+ antiporter Letm1. J Gen Physiol 143: 67–73, 2014. doi: 10.1085/jgp.201311096. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Bazil JN, Blomeyer CA, Pradhan RK, Camara AK, Dash RK. Modeling the calcium sequestration system in isolated guinea pig cardiac mitochondria. J Bioenerg Biomembr 45: 177–188, 2013. doi: 10.1007/s10863-012-9488-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Wolf SG, Mutsafi Y, Dadosh T, Ilani T, Lansky Z, Horowitz B, Rubin S, Elbaum M, Fass D. 3D visualization of mitochondrial solid-phase calcium stores in whole cells. eLife 6: e29929, 2017. doi: 10.7554/eLife.29929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Chinopoulos C, Adam-Vizi V. Mitochondrial Ca2+ sequestration and precipitation revisited. FEBS J 277: 3637–3651, 2010. doi: 10.1111/j.1742-4658.2010.07755.x. [DOI] [PubMed] [Google Scholar]
- 64. Blomeyer CA, Bazil JN, Stowe DF, Pradhan RK, Dash RK, Camara AK. Dynamic buffering of mitochondrial Ca2+ during Ca2+ uptake and Na+-induced Ca2+ release. J Bioenerg Biomembr 45: 189–202, 2013. doi: 10.1007/s10863-012-9483-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Chalmers S, Nicholls DG. The relationship between free and total calcium concentrations in the matrix of liver and brain mitochondria. J Biol Chem 278: 19062–19070, 2003. doi: 10.1074/jbc.M212661200. [DOI] [PubMed] [Google Scholar]
- 66. Wei AC, Liu T, Winslow RL, O’Rourke B. Dynamics of matrix-free Ca2+ in cardiac mitochondria: two components of Ca2+ uptake and role of phosphate buffering. J Gen Physiol 139: 465–478, 2012. doi: 10.1085/jgp.201210784. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Abramov AY, Fraley C, Diao CT, Winkfein R, Colicos MA, Duchen MR, French RJ, Pavlov E. Targeted polyphosphatase expression alters mitochondrial metabolism and inhibits calcium-dependent cell death. Proc Natl Acad Sci USA 104: 18091–18096, 2007. doi: 10.1073/pnas.0708959104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68. Solesio ME, Xie L, McIntyre B, Ellenberger M, Mitaishvili E, Bhadra-Lobo S, Bettcher LF, Bazil JN, Raftery D, Jakob U, Pavlov EV. Depletion of mitochondrial inorganic polyphosphate (polyP) in mammalian cells causes metabolic shift from oxidative phosphorylation to glycolysis. Biochem J 478: 1631–1646, 2021. doi: 10.1042/BCJ20200975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69. Malyala S, Zhang Y, Strubbe JO, Bazil JN. Calcium phosphate precipitation inhibits mitochondrial energy metabolism. PLoS Comput Biol 15: e1006719, 2019. doi: 10.1371/journal.pcbi.1006719. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70. Tomar D, Jaña F, Dong Z, Quinn WJ 3rd, Jadiya P, Breves SL, , et al. Blockade of MCU-mediated Ca2+ uptake perturbs lipid metabolism via PP4-dependent AMPK dephosphorylation. Cell Rep 26: 3709–3725.e7, 2019. doi: 10.1016/j.celrep.2019.02.107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71. Wheeler DG, Groth RD, Ma H, Barrett CF, Owen SF, Safa P, Tsien RW. CaV1 and CaV2 channels engage distinct modes of Ca2+ signaling to control CREB-dependent gene expression. Cell 149: 1112–1124, 2012. doi: 10.1016/j.cell.2012.03.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Romero-Garcia S, Prado-Garcia H. Mitochondrial calcium: transport and modulation of cellular processes in homeostasis and cancer. Int J Oncol 54: 1155–1167, 2019. doi: 10.3892/ijo.2019.4696. [DOI] [PubMed] [Google Scholar]
- 73. Singh S, Mabalirajan U. Mitochondrial calcium in command of juggling myriads of cellular functions. Mitochondrion 57: 108–118, 2021. doi: 10.1016/j.mito.2020.12.011. [DOI] [PubMed] [Google Scholar]
- 74. Wacquier B, Combettes L, Dupont G. Cytoplasmic and mitochondrial calcium signaling: a two-way relationship. Cold Spring Harb Perspect Biol 11: a035139, 2019. doi: 10.1101/cshperspect.a035139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Zima AV, Pabbidi MR, Lipsius SL, Blatter LA. Effects of mitochondrial uncoupling on Ca2+ signaling during excitation-contraction coupling in atrial myocytes. Am J Physiol Heart Circ Physiol 304: H983–H993, 2013. doi: 10.1152/ajpheart.00932.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Williams GS, Boyman L, Chikando AC, Khairallah RJ, Lederer WJ. Mitochondrial calcium uptake. Proc Natl Acad Sci USA 110: 10479–10486, 2013. doi: 10.1073/pnas.1300410110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Oropeza-Almazán Y, Blatter LA. Mitochondrial calcium uniporter complex activation protects against calcium alternans in atrial myocytes. Am J Physiol Heart Circ Physiol 319: H873–H881, 2020. doi: 10.1152/ajpheart.00375.2020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Del Arco A, Contreras L, Pardo B, Satrustegui J. Calcium regulation of mitochondrial carriers. Biochim Biophys Acta 1863: 2413–2421, 2016. doi: 10.1016/j.bbamcr.2016.03.024. [DOI] [PubMed] [Google Scholar]
- 79. Contreras L, Gomez-Puertas P, Iijima M, Kobayashi K, Saheki T, Satrústegui J. Ca2+ activation kinetics of the two aspartate-glutamate mitochondrial carriers, aralar and citrin: role in the heart malate-aspartate NADH shuttle. J Biol Chem 282: 7098–7106, 2007. doi: 10.1074/jbc.M610491200. [DOI] [PubMed] [Google Scholar]
- 80. Fisher-Wellman KH, Davidson MT, Narowski TM, Lin CT, Koves TR, Muoio DM. Mitochondrial diagnostics: a multiplexed assay platform for comprehensive assessment of mitochondrial energy fluxes. Cell Rep 24: 3593–3606.e10, 2018. doi: 10.1016/j.celrep.2018.08.091. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Mráček T, Drahota Z, Houštěk J. The function and the role of the mitochondrial glycerol-3-phosphate dehydrogenase in mammalian tissues. Biochim Biophys Acta 1827: 401–410, 2013. doi: 10.1016/j.bbabio.2012.11.014. [DOI] [PubMed] [Google Scholar]
- 82. Nichols BJ, Denton RM. Towards the molecular basis for the regulation of mitochondrial dehydrogenases by calcium ions. Mol Cell Biochem 149-150: 203–212, 1995. doi: 10.1007/BF01076578. [DOI] [PubMed] [Google Scholar]
- 83. Orr AL, Quinlan CL, Perevoshchikova IV, Brand MD. A refined analysis of superoxide production by mitochondrial sn-glycerol 3-phosphate dehydrogenase. J Biol Chem 287: 42921–42935, 2012. doi: 10.1074/jbc.M112.397828. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Beleznai Z, Szalay L, Jancsik V. Ca2+ and Mg2+ as modulators of mitochondrial L-glycerol-3-phosphate dehydrogenase. Eur J Biochem 170: 631–636, 1988. doi: 10.1111/j.1432-1033.1988.tb13744.x. [DOI] [PubMed] [Google Scholar]
- 85. Terburgh K, Lindeque Z, Mason S, van der Westhuizen F, Louw R. Metabolomics of Ndufs4-/- skeletal muscle: adaptive mechanisms converge at the ubiquinone-cycle. Biochim Biophys Acta Mol Basis Dis 1865: 98–106, 2019. doi: 10.1016/j.bbadis.2018.10.034. [DOI] [PubMed] [Google Scholar]
- 86. Barrow JJ, Balsa E, Verdeguer F, Tavares CD, Soustek MS, Hollingsworth LR, Jedrychowski M, Vogel R, Paulo JA, Smeitink J, Gygi SP, Doench J, Root DE, Puigserver P. Bromodomain inhibitors correct bioenergetic deficiency caused by mitochondrial disease Complex I mutations. Mol Cell 64: 163–175, 2016. doi: 10.1016/j.molcel.2016.08.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Zieliński ŁP, Smith AC, Smith AG, Robinson AJ. Metabolic flexibility of mitochondrial respiratory chain disorders predicted by computer modelling. Mitochondrion 31: 45–55, 2016. doi: 10.1016/j.mito.2016.09.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88. Ishihama S, Yoshida S, Yoshida T, Mori Y, Ouchi N, Eguchi S, Sakaguchi T, Tsuda T, Kato K, Shimizu Y, Ohashi K, Okumura T, Bando YK, Yagyu H, Wettschureck N, Kubota N, Offermanns S, Kadowaki T, Murohara T, Takefuji M. LPL/AQP7/GPD2 promotes glycerol metabolism under hypoxia and prevents cardiac dysfunction during ischemia. FASEB J 35: e22048, 2021. doi: 10.1096/fj.202100882R. [DOI] [PubMed] [Google Scholar]
- 89. Altamimi TR, Karwi QG, Uddin GM, Fukushima A, Kwong JQ, Molkentin JD, Lopaschuk GD. Cardiac-specific deficiency of the mitochondrial calcium uniporter augments fatty acid oxidation and functional reserve. J Mol Cell Cardiol 127: 223–231, 2019. doi: 10.1016/j.yjmcc.2018.12.019. [DOI] [PubMed] [Google Scholar]
- 90. Gherardi G, Nogara L, Ciciliot S, Fadini GP, Blaauw B, Braghetta P, Bonaldo P, De Stefani D, Rizzuto R, Mammucari C. Loss of mitochondrial calcium uniporter rewires skeletal muscle metabolism and substrate preference. Cell Death Differ 26: 362–381, 2019. doi: 10.1038/s41418-018-0191-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Kwong JQ, Huo J, Bround MJ, Boyer JG, Schwanekamp JA, Ghazal N, Maxwell JT, Jang YC, Khuchua Z, Shi K, Bers DM, Davis J, Molkentin JD. The mitochondrial calcium uniporter underlies metabolic fuel preference in skeletal muscle. JCI Insight 3: e121689, 2018. doi: 10.1172/jci.insight.121689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92. Lee SH, Duron HE, Chaudhuri D. Beyond the TCA cycle: new insights into mitochondrial calcium regulation of oxidative phosphorylation. Biochem Soc Trans 51: 1661–1673, 2023. doi: 10.1042/BST20230012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93. Denton RM. Regulation of mitochondrial dehydrogenases by calcium ions. Biochim Biophys Acta 1787: 1309–1316, 2009. doi: 10.1016/j.bbabio.2009.01.005. [DOI] [PubMed] [Google Scholar]
- 94. Denton RM, Randle PJ, Martin BR. Stimulation by calcium ions of pyruvate dehydrogenase phosphate phosphatase. Biochem J 128: 161–163, 1972. doi: 10.1042/bj1280161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Nemani N, Dong Z, Daw CC, Madaris TR, Ramachandran K, Enslow BT, Rubannelsonkumar CS, Shanmughapriya S, Mallireddigari V, Maity S, SinghMalla P, Natarajanseenivasan K, Hooper R, Shannon CE, Tourtellotte WG, Singh BB, Reeves WB, Sharma K, Norton L, Srikantan S, Soboloff J, Madesh M. Mitochondrial pyruvate and fatty acid flux modulate MICU1-dependent control of MCU activity. Sci Signal 13: eaaz6206, 2020. doi: 10.1126/scisignal.aaz6206. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96. McCormack JG, Halestrap AP, Denton RM. Role of calcium ions in regulation of mammalian intramitochondrial metabolism. Physiol Rev 70: 391–425, 1990. doi: 10.1152/physrev.1990.70.2.391. [DOI] [PubMed] [Google Scholar]
- 97. Denton RM, Richards DA, Chin JG. Calcium ions and the regulation of NAD+-linked isocitrate dehydrogenase from the mitochondria of rat heart and other tissues. Biochem J 176: 899–906, 1978. doi: 10.1042/bj1760899. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. McCormack JG, Denton RM. Role of calcium ions in the regulation of intramitochondrial metabolism. Properties of the Ca2+-sensitive dehydrogenases within intact uncoupled mitochondria from the white and brown adipose tissue of the rat. Biochem J 190: 95–105, 1980. doi: 10.1042/bj1900095. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99. Rutter GA, Denton RM. Regulation of NAD+-linked isocitrate dehydrogenase and 2-oxoglutarate dehydrogenase by Ca2+ ions within toluene-permeabilized rat heart mitochondria. Interactions with regulation by adenine nucleotides and NADH/NAD+ ratios. Biochem J 252: 181–189, 1988. doi: 10.1042/bj2520181. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100. Rutter GA, Denton RM. The binding of Ca2+ ions to pig heart NAD+-isocitrate dehydrogenase and the 2-oxoglutarate dehydrogenase complex. Biochem J 263: 453–462, 1989. doi: 10.1042/bj2630453. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Denton RM, McCormack JG, Edgell NJ. Role of calcium ions in the regulation of intramitochondrial metabolism. Effects of Na+, Mg2+ and ruthenium red on the Ca2+-stimulated oxidation of oxoglutarate and on pyruvate dehydrogenase activity in intact rat heart mitochondria. Biochem J 190: 107–117, 1980. doi: 10.1042/bj1900107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Panov AV, Scaduto RC Jr.. Substrate specific effects of calcium on metabolism of rat heart mitochondria. Am J Physiol Heart Circ Physiol 270: H1398–H1406, 1996. doi: 10.1152/ajpheart.1996.270.4.H1398. [DOI] [PubMed] [Google Scholar]
- 103. Holmström KM, Pan X, Liu JC, Menazza S, Liu J, Nguyen TT, Pan H, Parks RJ, Anderson S, Noguchi A, Springer D, Murphy E, Finkel T. Assessment of cardiac function in mice lacking the mitochondrial calcium uniporter. J Mol Cell Cardiol 85: 178–182, 2015. doi: 10.1016/j.yjmcc.2015.05.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104. Kosmach A, Roman B, Sun J, Femnou A, Zhang F, Liu C, Combs CA, Balaban RS, Murphy E. Monitoring mitochondrial calcium and metabolism in the beating MCU-KO heart. Cell Rep 37: 109846, 2021. doi: 10.1016/j.celrep.2021.109846. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105. Kwong JQ, Lu X, Correll RN, Schwanekamp JA, Vagnozzi RJ, Sargent MA, York AJ, Zhang J, Bers DM, Molkentin JD. The mitochondrial calcium uniporter selectively matches metabolic output to acute contractile stress in the heart. Cell Rep 12: 15–22, 2015. doi: 10.1016/j.celrep.2015.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106. Liu JC, Syder NC, Ghorashi NS, Willingham TB, Parks RJ, Sun J, Fergusson MM, Liu J, Holmström KM, Menazza S, Springer DA, Liu C, Glancy B, Finkel T, Murphy E. EMRE is essential for mitochondrial calcium uniporter activity in a mouse model. JCI Insight 5: e134063, 2020. doi: 10.1172/jci.insight.134063. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Luongo TS, Lambert JP, Yuan A, Zhang X, Gross P, Song J, Shanmughapriya S, Gao E, Jain M, Houser SR, Koch WJ, Cheung JY, Madesh M, Elrod JW. The mitochondrial calcium uniporter matches energetic supply with cardiac workload during stress and modulates permeability transition. Cell Rep 12: 23–34, 2015. doi: 10.1016/j.celrep.2015.06.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Pan X, Liu J, Nguyen T, Liu C, Sun J, Teng Y, Fergusson MM, Rovira II, Allen M, Springer DA, Aponte AM, Gucek M, Balaban RS, Murphy E, Finkel T. The physiological role of mitochondrial calcium revealed by mice lacking the mitochondrial calcium uniporter. Nat Cell Biol 15: 1464–1472, 2013. doi: 10.1038/ncb2868. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109. Chapoy Villanueva H, Sung JH, Stevens JA, Zhang MJ, Nelson PM, Denduluri LS, Feng F, O’Connell TD, Townsend D, Liu JC. Distinct effects of cardiac mitochondrial calcium uniporter inactivation via EMRE deletion in the short and long term. J Mol Cell Cardiol 181: 33–45, 2023. doi: 10.1016/j.yjmcc.2023.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110. Locke TM, Fields R, Gizinski H, Otto GM, Shechner DM, Berg MD, Villen J, Sancak Y, Schweppe D. High-throughput identification of calcium regulated proteins across diverse proteomes (Preprint). bioRxiv 2024.01.18.575273, 2024. doi: 10.1101/2024.01.18.575273. [DOI]
- 111. Territo PR, French SA, Dunleavy MC, Evans FJ, Balaban RS. Calcium activation of heart mitochondrial oxidative phosphorylation: rapid kinetics of mVO2, NADH, and light scattering. J Biol Chem 276: 2586–2599, 2001. doi: 10.1074/jbc.M002923200. [DOI] [PubMed] [Google Scholar]
- 112. Glancy B, Willis WT, Chess DJ, Balaban RS. Effect of calcium on the oxidative phosphorylation cascade in skeletal muscle mitochondria. Biochemistry 52: 2793–2809, 2013. doi: 10.1021/bi3015983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113. Rutter GA, McCormack JG, Halestrap AP, Denton RM. The roles of cytosolic and intramitochondrial Ca2+ and the mitochondrial Ca2+-uniporter (MCU) in the stimulation of mammalian oxidative phosphorylation. J Biol Chem 295: 10506, 2020. doi: 10.1074/jbc.L120.013975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. Balderas E, Eberhardt DR, Lee S, Pleinis JM, Sommakia S, Balynas AM, Yin X, Parker MC, Maguire CT, Cho S, Szulik MW, Bakhtina A, Bia RD, Friederich MW, Locke TM, Van Hove JL, Drakos SG, Sancak Y, Tristani-Firouzi M, Franklin S, Rodan AR, Chaudhuri D. Mitochondrial calcium uniporter stabilization preserves energetic homeostasis during Complex I impairment. Nat Commun 13: 2769, 2022. doi: 10.1038/s41467-022-30236-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115. Kussmaul L, Hirst J. The mechanism of superoxide production by NADH:ubiquinone oxidoreductase (complex I) from bovine heart mitochondria. Proc Natl Acad Sci USA 103: 7607–7612, 2006. doi: 10.1073/pnas.0510977103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Pryde KR, Taanman JW, Schapira AH. A LON-ClpP proteolytic axis degrades Complex I to extinguish ROS production in depolarized mitochondria. Cell Rep 17: 2522–2531, 2016. doi: 10.1016/j.celrep.2016.11.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117. Sommakia S, Houlihan PR, Deane SS, Simcox JA, Torres NS, Jeong MY, Winge DR, Villanueva CJ, Chaudhuri D. Mitochondrial cardiomyopathies feature increased uptake and diminished efflux of mitochondrial calcium. J Mol Cell Cardiol 113: 22–32, 2017. doi: 10.1016/j.yjmcc.2017.09.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118. Jaña F, Bustos G, Rivas J, Cruz P, Urra F, Basualto-Alarcón C, Sagredo E, Ríos M, Lovy A, Dong Z, Cerda O, Madesh M, Cardenas C. Complex I and II are required for normal mitochondrial Ca2+ homeostasis. Mitochondrion 49: 73–82, 2019. doi: 10.1016/j.mito.2019.07.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Murphy AN, Kelleher JK, Fiskum G. Submicromolar Ca2+ regulates phosphorylating respiration by normal rat liver and AS-30D hepatoma mitochondria by different mechanisms. J Biol Chem 265: 10527–10534, 1990. doi: 10.1016/S0021-9258(18)86979-4. [DOI] [PubMed] [Google Scholar]
- 120. Hernansanz-Agustín P, Choya-Foces C, Carregal-Romero S, Ramos E, Oliva T, Villa-Piña T, Moreno L, Izquierdo-Álvarez A, Cabrera-Garcia JD, Cortes A, Lechuga-Vieco AV, Jadiya P, Navarro E, Parada E, Palomino-Antolin A, Tello D, Acin-Perez R, Rodriguez-Aguilera JC, Navas P, Cogolludo A, Lopez-Montero I, Martinez-Del-Pozo A, Egea J, Lopez MG, Elrod JW, Ruiz-Cabello J, Bogdanova A, Enriquez JA, Martinez-Ruiz A. Na+ controls hypoxic signalling by the mitochondrial respiratory chain. Nature 586: 287–291, 2020. doi: 10.1038/s41586-020-2551-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121. Vygodina T, Kirichenko A, Konstantinov AA. Direct regulation of cytochrome c oxidase by calcium ions. PLoS One 8: e74436, 2013. doi: 10.1371/journal.pone.0074436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122. Mallilankaraman K, Cárdenas C, Doonan PJ, Chandramoorthy HC, Irrinki KM, Golenár T, Csordás G, Madireddi P, Yang J, Müller M, Miller R, Kolesar JE, Molgó J, Kaufman B, Hajnóczky G, Foskett JK, Madesh M. MCUR1 is an essential component of mitochondrial Ca2+ uptake that regulates cellular metabolism. Nat Cell Biol 14: 1336–1343, 2012. doi: 10.1038/ncb2622. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123. Paupe V, Prudent J, Dassa EP, Rendon OZ, Shoubridge EA. CCDC90A (MCUR1) is a cytochrome c oxidase assembly factor and not a regulator of the mitochondrial calcium uniporter. Cell Metab 21: 109–116, 2015. doi: 10.1016/j.cmet.2014.12.004. [DOI] [PubMed] [Google Scholar]
- 124. Tomar D, Dong Z, Shanmughapriya S, Koch DA, Thomas T, Hoffman NE, , et al. MCUR1 Is a scaffold factor for the MCU complex function and promotes mitochondrial bioenergetics. Cell Rep 15: 1673–1685, 2016. doi: 10.1016/j.celrep.2016.04.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Vais H, Tanis JE, Müller M, Payne R, Mallilankaraman K, Foskett JK. MCUR1, CCDC90A, is a regulator of the mitochondrial calcium uniporter. Cell Metab 22: 533–535, 2015. doi: 10.1016/j.cmet.2015.09.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126. Chaudhuri D, Artiga DJ, Abiria SA, Clapham DE. Mitochondrial calcium uniporter regulator 1 (MCUR1) regulates the calcium threshold for the mitochondrial permeability transition. Proc Natl Acad Sci USA 113: E1872–E1880, 2016. doi: 10.1073/pnas.1602264113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127. Letts JA, Fiedorczuk K, Sazanov LA. The architecture of respiratory supercomplexes. Nature 537: 644–648, 2016. doi: 10.1038/nature19774. [DOI] [PubMed] [Google Scholar]
- 128. Wu M, Gu J, Guo R, Huang Y, Yang M. Structure of mammalian respiratory supercomplex I1III2IV1. Cell 167: 1598–1609.e10, 2016. doi: 10.1016/j.cell.2016.11.012. [DOI] [PubMed] [Google Scholar]
- 129. Milenkovic D, Blaza JN, Larsson NG, Hirst J. The enigma of the respiratory chain supercomplex. Cell Metab 25: 765–776, 2017. doi: 10.1016/j.cmet.2017.03.009. [DOI] [PubMed] [Google Scholar]
- 130. Sousa JS, Mills DJ, Vonck J, Kühlbrandt W. Functional asymmetry and electron flow in the bovine respirasome. eLife 5: e21290, 2016. doi: 10.7554/eLife.21290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Fernández-Vizarra E, López-Calcerrada S, Sierra-Magro A, Pérez-Pérez R, Formosa LE, Hock DH, Illescas M, Peñas A, Brischigliaro M, Ding S, Fearnley IM, Tzoulis C, Pitceathly RD, Arenas J, Martín MA, Stroud DA, Zeviani M, Ryan MT, Ugalde C. Two independent respiratory chains adapt OXPHOS performance to glycolytic switch. Cell Metab 34: 1792–1808.e6, 2022. doi: 10.1016/j.cmet.2022.09.005. [DOI] [PubMed] [Google Scholar]
- 132. Milenkovic D, Misic J, Hevler JF, Molinié T, Chung I, Atanassov I, Li X, Filograna R, Mesaros A, Mourier A, Heck AJ, Hirst J, Larsson NG. Preserved respiratory chain capacity and physiology in mice with profoundly reduced levels of mitochondrial respirasomes. Cell Metab 35: 1799–1813.e7, 2023. doi: 10.1016/j.cmet.2023.07.015. [DOI] [PubMed] [Google Scholar]
- 133. Kasahara T, Kubota-Sakashita M, Nagatsuka Y, Hirabayashi Y, Hanasaka T, Tohyama K, Kato T. Cardiolipin is essential for early embryonic viability and mitochondrial integrity of neurons in mammals. FASEB J 34: 1465–1480, 2020. doi: 10.1096/fj.201901598R. [DOI] [PubMed] [Google Scholar]
- 134. Jang S, Lewis TS, Powers C, Khuchua Z, Baines CP, Wipf P, Javadov S. Elucidating mitochondrial electron transport chain supercomplexes in the heart during ischemia-reperfusion. Antioxid Redox Signal 27: 57–69, 2017. doi: 10.1089/ars.2016.6635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135. Jang S, Javadov S. Association between ROS production, swelling and the respirasome integrity in cardiac mitochondria. Arch Biochem Biophys 630: 1–8, 2017. doi: 10.1016/j.abb.2017.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136. Zanou N, Dridi H, Reiken S, Imamura de Lima T, Donnelly C, De Marchi U, Ferrini M, Vidal J, Sittenfeld L, Feige JN, Garcia-Roves PM, Lopez-Mejia IC, Marks AR, Auwerx J, Kayser B, Place N. Acute RyR1 Ca2+ leak enhances NADH-linked mitochondrial respiratory capacity. Nat Commun 12: 7219, 2021. doi: 10.1038/s41467-021-27422-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137. Clements RT, Terentyeva R, Hamilton S, Janssen PM, Roder K, Martin BY, Perger F, Schneider T, Nichtova Z, Das AS, Veress R, Lee BS, Kim DG, Koren G, Stratton MS, Csordas G, Accornero F, Belevych AE, Gyorke S, Terentyev D. Sexual dimorphism in bidirectional SR-mitochondria crosstalk in ventricular cardiomyocytes. Basic Res Cardiol 118: 15, 2023. doi: 10.1007/s00395-023-00988-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138. Abrahams JP, Leslie AG, Lutter R, Walker JE. Structure at 2.8 A resolution of F1-ATPase from bovine heart mitochondria. Nature 370: 621–628, 1994. doi: 10.1038/370621a0. [DOI] [PubMed] [Google Scholar]
- 139. Stock D, Leslie AG, Walker JE. Molecular architecture of the rotary motor in ATP synthase. Science 286: 1700–1705, 1999. doi: 10.1126/science.286.5445.1700. [DOI] [PubMed] [Google Scholar]
- 140. Territo PR, Mootha VK, French SA, Balaban RS. Ca2+ activation of heart mitochondrial oxidative phosphorylation: role of the F0/F1-ATPase. Am J Physiol Cell Physiol 278: C423–C435, 2000. doi: 10.1152/ajpcell.2000.278.2.C423. [DOI] [PubMed] [Google Scholar]
- 141. Huang G, Docampo R. The mitochondrial calcium uniporter interacts with subunit c of the ATP synthase of trypanosomes and humans. mBio 11: 10.1128/mbio.00268-20, 2020. doi: 10.1128/mBio.00268-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142. Bernardi P, Carraro M, Lippe G. The mitochondrial permeability transition: recent progress and open questions. FEBS J 289: 7051–7074, 2022. doi: 10.1111/febs.16254. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143. Fournier N, Ducet G, Crevat A. Action of cyclosporine on mitochondrial calcium fluxes. J Bioenerg Biomembr 19: 297–303, 1987. doi: 10.1007/BF00762419. [DOI] [PubMed] [Google Scholar]
- 144. Hunter DR, Haworth RA, Southard JH. Relationship between configuration, function, and permeability in calcium-treated mitochondria. J Biol Chem 251: 5069–5077, 1976. doi: 10.1016/S0021-9258(17)33220-9. [DOI] [PubMed] [Google Scholar]
- 145. Eberhardt DR, Lee SH, Yin X, Balynas AM, Rekate EC, Kraiss JN, Lang MJ, Walsh MA, Streiff ME, Corbin AC, Li Y, Funai K, Sachse FB, Chaudhuri D. EFHD1 ablation inhibits cardiac mitoflash activation and protects cardiomyocytes from ischemia. J Mol Cell Cardiol 167: 1–14, 2022. doi: 10.1016/j.yjmcc.2022.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146. Brustovetsky N. The role of adenine nucleotide translocase in the mitochondrial permeability transition. Cells 9: 2686, 2020. doi: 10.3390/cells9122686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147. Karch J, Bround MJ, Khalil H, Sargent MA, Latchman N, Terada N, Peixoto PM, Molkentin JD. Inhibition of mitochondrial permeability transition by deletion of the ANT family and CypD. Sci Adv 5: eaaw4597, 2019. doi: 10.1126/sciadv.aaw4597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148. Garbincius JF, Luongo TS, Elrod JW. The debate continues-what is the role of MCU and mitochondrial calcium uptake in the heart? J Mol Cell Cardiol 143: 163–174, 2020. doi: 10.1016/j.yjmcc.2020.04.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149. O’Rourke B, Ashok D, Liu T. Mitochondrial Ca2+ in heart failure: not enough or too much? J Mol Cell Cardiol 151: 126–134, 2021. doi: 10.1016/j.yjmcc.2020.11.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150. Wang P, Fernandez-Sanz C, Wang W, Sheu SS. Why don’t mice lacking the mitochondrial Ca2+ uniporter experience an energy crisis? J Physiol 598: 1307–1326, 2020. doi: 10.1113/JP276636. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151. Balderas E, Chaudhuri D. Absence (of the uniporter) makes the heart grow fonder: the cardiac response to injury adapts after prolonged EMRE inhibition. J Mol Cell Cardiol 181: 31–32, 2023. doi: 10.1016/j.yjmcc.2023.05.006. [DOI] [PubMed] [Google Scholar]
- 152. Ingwall JS, Weiss RG. Is the failing heart energy starved? On using chemical energy to support cardiac function. Circ Res 95: 135–145, 2004. doi: 10.1161/01.RES.0000137170.41939.d9. [DOI] [PubMed] [Google Scholar]
- 153. Tian R, Ingwall JS. Energetic basis for reduced contractile reserve in isolated rat hearts. Am J Physiol Heart Circ Physiol 270: H1207–H1216, 1996. doi: 10.1152/ajpheart.1996.270.4.H1207. [DOI] [PubMed] [Google Scholar]
- 154. Liu T, Yang N, Sidor A, O’Rourke B. MCU overexpression rescues inotropy and reverses heart failure by reducing SR Ca2+ leak. Circ Res 128: 1191–1204, 2021. doi: 10.1161/CIRCRESAHA.120.318562. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155. Wang P, Xu S, Xu J, Xin Y, Lu Y, Zhang H, Zhou B, Xu H, Sheu SS, Tian R, Wang W. Elevated MCU expression by CaMKIIdeltaB limits pathological cardiac remodeling. Circulation 145: 1067–1083, 2022. doi: 10.1161/CIRCULATIONAHA.121.055841. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156. Chouchani ET, Pell VR, Gaude E, Aksentijević D, Sundier SY, Robb EL, Logan A, Nadtochiy SM, Ord EN, Smith AC, Eyassu F, Shirley R, Hu CH, Dare AJ, James AM, Rogatti S, Hartley RC, Eaton S, Costa AS, Brookes PS, Davidson SM, Duchen MR, Saeb-Parsy K, Shattock MJ, Robinson AJ, Work LM, Frezza C, Krieg T, Murphy MP. Ischaemic accumulation of succinate controls reperfusion injury through mitochondrial ROS. Nature 515: 431–435, 2014. doi: 10.1038/nature13909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157. Sciuto KJ, Deng SW, Moreno A, Zaitsev AV. Chronology of critical events in neonatal rat ventricular myocytes occurring during reperfusion after simulated ischemia. PLoS One 14: e0212076, 2019. doi: 10.1371/journal.pone.0212076. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158. Sciuto KJ, Deng SW, Venable PW, Warren M, Warren JS, Zaitsev AV. Cyclosporine-insensitive mode of cell death after prolonged myocardial ischemia: evidence for sarcolemmal permeabilization as the pivotal step. PLoS One 13: e0200301, 2018. doi: 10.1371/journal.pone.0200301. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159. Petersen CE, Sun J, Silva K, Kosmach A, Balaban RS, Murphy E. Increased mitochondrial free Ca2+ during ischemia is suppressed, but not eliminated by, germline deletion of the mitochondrial Ca2+ uniporter. Cell Rep 42: 112735, 2023. doi: 10.1016/j.celrep.2023.112735. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160. García-Rivas GD, Carvajal K, Correa F, Zazueta C. Ru360, a specific mitochondrial calcium uptake inhibitor, improves cardiac post-ischaemic functional recovery in rats in vivo. Br J Pharmacol 149: 829–837, 2006. doi: 10.1038/sj.bjp.0706932. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161. Liu T, Takimoto E, Dimaano VL, DeMazumder D, Kettlewell S, Smith G, Sidor A, Abraham TP, O’Rourke B. Inhibiting mitochondrial Na+/Ca2+ exchange prevents sudden death in a Guinea pig model of heart failure. Circ Res 115: 44–54, 2014. doi: 10.1161/CIRCRESAHA.115.303062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162. Balsa E, Perry EA, Bennett CF, Jedrychowski M, Gygi SP, Doench JG, Puigserver P. Defective NADPH production in mitochondrial disease complex I causes inflammation and cell death. Nat Commun 11: 2714, 2020. doi: 10.1038/s41467-020-16423-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163. Nickel AG, von Hardenberg A, Hohl M, Löffler JR, Kohlhaas M, Becker J, Reil JC, Kazakov A, Bonnekoh J, Stadelmaier M, Puhl SL, Wagner M, Bogeski I, Cortassa S, Kappl R, Pasieka B, Lafontaine M, Lancaster CR, Blacker TS, Hall AR, Duchen MR, Kästner L, Lipp P, Zeller T, Müller C, Knopp A, Laufs U, Böhm M, Hoth M, Maack C. Reversal of mitochondrial transhydrogenase causes oxidative stress in heart failure. Cell Metab 22: 472–484, 2015. doi: 10.1016/j.cmet.2015.07.008. [DOI] [PubMed] [Google Scholar]
- 164. Rigobello MP, Vianello F, Folda A, Roman C, Scutari G, Bindoli A. Differential effect of calcium ions on the cytosolic and mitochondrial thioredoxin reductase. Biochem Biophys Res Commun 343: 873–878, 2006. doi: 10.1016/j.bbrc.2006.03.050. [DOI] [PubMed] [Google Scholar]
- 165. Ashok D, Papanicolaou K, Sidor A, Wang M, Solhjoo S, Liu T, O’Rourke B. Mitochondrial membrane potential instability on reperfusion after ischemia does not depend on mitochondrial Ca2+ uptake. J Biol Chem 299: 104708, 2023. doi: 10.1016/j.jbc.2023.104708. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166. Sventzouri S, Nanas I, Vakrou S, Kapelios C, Sousonis V, Sfakianaki T, Papalois A, Manolis AS, Nanas JN, Malliaras K. Pharmacologic inhibition of the mitochondrial Na+/Ca2+ exchanger protects against ventricular arrhythmias in a porcine model of ischemia-reperfusion. Hellenic J Cardiol 59: 217–222, 2018. doi: 10.1016/j.hjc.2017.12.009. [DOI] [PubMed] [Google Scholar]
- 167. Liu H, Zhao Y, Xie A, Kim TY, Terentyeva R, Liu M, Shi G, Feng F, Choi BR, Terentyev D, Hamilton S, Dudley SC Jr.. Interleukin-1β, oxidative stress, and abnormal calcium handling mediate diabetic arrhythmic risk. JACC Basic Transl Sci 6: 42–52, 2021. doi: 10.1016/j.jacbts.2020.11.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168. Tow BD, Deb A, Neupane S, Patel SM, Reed M, Loper AB, Eliseev RA, Knollmann BC, Györke S, Liu B. SR-mitochondria crosstalk shapes Ca signalling to impact pathophenotype in disease models marked by dysregulated intracellular Ca release. Cardiovasc Res 118: 2819–2832, 2022. doi: 10.1093/cvr/cvab324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169. Joseph LC, Reyes MV, Homan EA, Gowen B, Avula UM, Goulbourne CN, Wan EY, Elrod JW, Morrow JP. The mitochondrial calcium uniporter promotes arrhythmias caused by high-fat diet. Sci Rep 11: 17808, 2021. doi: 10.1038/s41598-021-97449-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170. Velmurugan S, Liu T, Chen KC, Despa F, O’Rourke B, Despa S. Distinct effects of mitochondrial Na+/Ca2+ exchanger inhibition and Ca2+ uniporter activation on Ca2+ sparks and arrhythmogenesis in diabetic rats. J Am Heart Assoc 12: e029997, 2023. doi: 10.1161/JAHA.123.029997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171. Abbas M, Miles C, Behr E. Catecholaminergic polymorphic ventricular tachycardia. Arrhythm Electrophysiol Rev 11: e20, 2022. doi: 10.15420/aer.2022.09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172. Zhao YT, Valdivia CR, Gurrola GB, Powers PP, Willis BC, Moss RL, Jalife J, Valdivia HH. Arrhythmogenesis in a catecholaminergic polymorphic ventricular tachycardia mutation that depresses ryanodine receptor function. Proc Natl Acad Sci USA 112: E1669–E1677, 2015. doi: 10.1073/pnas.1419795112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173. Hamilton S, Terentyeva R, Martin B, Perger F, Li J, Stepanov A, Bonilla IM, Knollmann BC, Radwański PB, Györke S, Belevych AE, Terentyev D. Increased RyR2 activity is exacerbated by calcium leak-induced mitochondrial ROS. Basic Res Cardiol 115: 38, 2020. doi: 10.1007/s00395-020-0797-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174. Sander P, Feng M, Schweitzer MK, Wilting F, Gutenthaler SM, Arduino DM, Fischbach S, Dreizehnter L, Moretti A, Gudermann T, Perocchi F, Schredelseker J. Approved drugs ezetimibe and disulfiram enhance mitochondrial Ca2+ uptake and suppress cardiac arrhythmogenesis. Br J Pharmacol 178: 4518–4532, 2021. doi: 10.1111/bph.15630. [DOI] [PubMed] [Google Scholar]
- 175. Shimizu H, Schredelseker J, Huang J, Lu K, Naghdi S, Lu F, Franklin S, Fiji HD, Wang K, Zhu H, Tian C, Lin B, Nakano H, Ehrlich A, Nakai J, Stieg AZ, Gimzewski JK, Nakano A, Goldhaber JI, Vondriska TM, Hajnóczky G, Kwon O, Chen JN. Mitochondrial Ca2+ uptake by the voltage-dependent anion channel 2 regulates cardiac rhythmicity. eLife 4: e04801, 2015. doi: 10.7554/eLife.04801. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176. Schweitzer MK, Wilting F, Sedej S, Dreizehnter L, Dupper NJ, Tian Q, Moretti A, My I, Kwon O, Priori SG, Laugwitz KL, Storch U, Lipp P, Breit A, Mederos YS, Gudermann T, Schredelseker J. Suppression of arrhythmia by enhancing mitochondrial Ca2+ uptake in catecholaminergic ventricular tachycardia models. JACC Basic Transl Sci 2: 737–747, 2017. doi: 10.1016/j.jacbts.2017.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177. Deb A, Tow BD, Qing Y, Walker M, Hodges ER, Stewart JA Jr, Knollmann BC, Zheng Y, Wang Y, Liu B. Genetic inhibition of mitochondrial permeability transition pore exacerbates ryanodine receptor 2 dysfunction in arrhythmic disease. Cells 12: 204, 2023. doi: 10.3390/cells12020204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178. El-Hattab AW, Scaglia F. Mitochondrial cardiomyopathies. Front Cardiovasc Med 3: 25, 2016. doi: 10.3389/fcvm.2016.00025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 179. Meyers DE, Basha HI, Koenig MK. Mitochondrial cardiomyopathy: pathophysiology, diagnosis, and management. Tex Heart Inst J 40: 385–394, 2013. [PMC free article] [PubMed] [Google Scholar]
- 180. Schaefer AM, Taylor RW, Turnbull DM, Chinnery PF. The epidemiology of mitochondrial disorders–past, present and future. Biochim Biophys Acta 1659: 115–120, 2004. doi: 10.1016/j.bbabio.2004.09.005. [DOI] [PubMed] [Google Scholar]
- 181. Holmgren D, Wåhlander H, Eriksson BO, Oldfors A, Holme E, Tulinius M. Cardiomyopathy in children with mitochondrial disease; clinical course and cardiological findings. Eur Heart J 24: 280–288, 2003. doi: 10.1016/s0195-668x(02)00387-1. [DOI] [PubMed] [Google Scholar]
- 182. Willems PH, Valsecchi F, Distelmaier F, Verkaart S, Visch HJ, Smeitink JA, Koopman WJ. Mitochondrial Ca2+ homeostasis in human NADH:ubiquinone oxidoreductase deficiency. Cell Calcium 44: 123–133, 2008. doi: 10.1016/j.ceca.2008.01.002. [DOI] [PubMed] [Google Scholar]
- 183. Visch HJ, Rutter GA, Koopman WJ, Koenderink JB, Verkaart S, de Groot T, Varadi A, Mitchell KJ, van den Heuvel LP, Smeitink JA, Willems PH. Inhibition of mitochondrial Na+-Ca2+ exchange restores agonist-induced ATP production and Ca2+ handling in human complex I deficiency. J Biol Chem 279: 40328–40336, 2004. doi: 10.1074/jbc.M408068200. [DOI] [PubMed] [Google Scholar]
- 184. Granatiero V, Giorgio V, Calì T, Patron M, Brini M, Bernardi P, Tiranti V, Zeviani M, Pallafacchina G, De Stefani D, Rizzuto R. Reduced mitochondrial Ca2+ transients stimulate autophagy in human fibroblasts carrying the 13514A>G mutation of the ND5 subunit of NADH dehydrogenase. Cell Death Differ 23: 231–241, 2016. doi: 10.1038/cdd.2015.84. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185. Brini M, Pinton P, King MP, Davidson M, Schon EA, Rizzuto R. A calcium signaling defect in the pathogenesis of a mitochondrial DNA inherited oxidative phosphorylation deficiency. Nat Med 5: 951–954, 1999. doi: 10.1038/11396. [DOI] [PubMed] [Google Scholar]
- 186. Valsecchi F, Esseling JJ, Koopman WJ, Willems PH. Calcium and ATP handling in human NADH:ubiquinone oxidoreductase deficiency. Biochim Biophys Acta 1792: 1130–1137, 2009. doi: 10.1016/j.bbadis.2009.01.001. [DOI] [PubMed] [Google Scholar]
- 187. Aydin J, Andersson DC, Hänninen SL, Wredenberg A, Tavi P, Park CB, Larsson NG, Bruton JD, Westerblad H. Increased mitochondrial Ca2+ and decreased sarcoplasmic reticulum Ca2+ in mitochondrial myopathy. Hum Mol Genet 18: 278–288, 2009. doi: 10.1093/hmg/ddn355. [DOI] [PubMed] [Google Scholar]
- 188. Sherer TB, Trimmer PA, Parks JK, Tuttle JB. Mitochondrial DNA-depleted neuroblastoma (Rho degrees) cells exhibit altered calcium signaling. Biochim Biophys Acta 1496: 341–355, 2000. doi: 10.1016/s0167-4889(00)00027-6. [DOI] [PubMed] [Google Scholar]
- 189. Li H, Wang J, Wilhelmsson H, Hansson A, Thoren P, Duffy J, Rustin P, Larsson NG. Genetic modification of survival in tissue-specific knockout mice with mitochondrial cardiomyopathy. Proc Natl Acad Sci USA 97: 3467–3472, 2000. doi: 10.1073/pnas.97.7.3467. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190. Wang J, Wilhelmsson H, Graff C, Li H, Oldfors A, Rustin P, Brüning JC, Kahn CR, Clayton DA, Barsh GS, Thorén P, Larsson NG. Dilated cardiomyopathy and atrioventricular conduction blocks induced by heart-specific inactivation of mitochondrial DNA gene expression. Nat Genet 21: 133–137, 1999. doi: 10.1038/5089. [DOI] [PubMed] [Google Scholar]
- 191. Clarke SL, Bowron A, Gonzalez IL, Groves SJ, Newbury-Ecob R, Clayton N, Martin RP, Tsai-Goodman B, Garratt V, Ashworth M, Bowen VM, McCurdy KR, Damin MK, Spencer CT, Toth MJ, Kelley RI, Steward CG. Barth syndrome. Orphanet J Rare Dis 8: 23, 2013. doi: 10.1186/1750-1172-8-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 192. Sabbah HN. Barth syndrome cardiomyopathy: targeting the mitochondria with elamipretide. Heart Fail Rev 26: 237–253, 2021. doi: 10.1007/s10741-020-10031-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193. Bertero E, Nickel A, Kohlhaas M, Hohl M, Sequeira V, Brune C, Schwemmlein J, Abeßer M, Schuh K, Kutschka I, Carlein C, Münker K, Atighetchi S, Müller A, Kazakov A, Kappl R, von der Malsburg K, van der Laan M, Schiuma AF, Böhm M, Laufs U, Hoth M, Rehling P, Kuhn M, Dudek J, von der Malsburg A, Prates Roma L, Maack C. Loss of mitochondrial Ca2+ uniporter limits inotropic reserve and provides trigger and substrate for arrhythmias in Barth syndrome cardiomyopathy. Circulation 144: 1694–1713, 2021. doi: 10.1161/CIRCULATIONAHA.121.053755. [DOI] [PubMed] [Google Scholar]
- 194. Ghosh S, Zulkifli M, Joshi A, Venkatesan M, Cristel A, Vishnu N, Madesh M, Gohil VM. MCU-complex-mediated mitochondrial calcium signaling is impaired in Barth syndrome. Hum Mol Genet 31: 376–385, 2022. doi: 10.1093/hmg/ddab254. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195. Ghosh S, Basu Ball W, Madaris TR, Srikantan S, Madesh M, Mootha VK, Gohil VM. An essential role for cardiolipin in the stability and function of the mitochondrial calcium uniporter. Proc Natl Acad Sci USA 117: 16383–16390, 2020. doi: 10.1073/pnas.2000640117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196. Liu X, Wang S, Guo X, Li Y, Ogurlu R, Lu F, Prondzynski M, de la Serna Buzon S, Ma Q, Zhang D, Wang G, Cotton J, Guo Y, Xiao L, Milan DJ, Xu Y, Schlame M, Bezzerides VJ, Pu WT. Increased reactive oxygen species-mediated Ca2+/calmodulin-dependent protein kinase II Activation contributes to calcium handling abnormalities and impaired contraction in Barth syndrome. Circulation 143: 1894–1911, 2021. doi: 10.1161/CIRCULATIONAHA.120.048698. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197. Chiang S, Kalinowski DS, Jansson PJ, Richardson DR, Huang ML. Mitochondrial dysfunction in the neuro-degenerative and cardio-degenerative disease, Friedreich’s ataxia. Neurochem Int 117: 35–48, 2018. doi: 10.1016/j.neuint.2017.08.002. [DOI] [PubMed] [Google Scholar]
- 198. Tamarit J, Britti E, Delaspre F, Medina-Carbonero M, Sanz-Alcázar A, Cabiscol E, Ros J. Mitochondrial iron and calcium homeostasis in Friedreich ataxia. IUBMB Life 73: 543–553, 2021. doi: 10.1002/iub.2457. [DOI] [PubMed] [Google Scholar]
- 199. Bolinches-Amorós A, Mollá B, Pla-Martín D, Palau F, González-Cabo P. Mitochondrial dysfunction induced by frataxin deficiency is associated with cellular senescence and abnormal calcium metabolism. Front Cell Neurosci 8: 124, 2014. doi: 10.3389/fncel.2014.00124. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200. Rodríguez LR, Calap-Quintana P, Lapeña-Luzón T, Pallardó FV, Schneuwly S, Navarro JA, Gonzalez-Cabo P. Oxidative stress modulates rearrangement of endoplasmic reticulum-mitochondria contacts and calcium dysregulation in a Friedreich’s ataxia model. Redox Biol 37: 101762, 2020. doi: 10.1016/j.redox.2020.101762. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201. Purroy R, Britti E, Delaspre F, Tamarit J, Ros J. Mitochondrial pore opening and loss of Ca2+ exchanger NCLX levels occur after frataxin depletion. Biochim Biophys Acta Mol Basis Dis 1864: 618–631, 2018. doi: 10.1016/j.bbadis.2017.12.005. [DOI] [PubMed] [Google Scholar]
- 202. Ast T, Meisel JD, Patra S, Wang H, Grange RM, Kim SH, Calvo SE, Orefice LL, Nagashima F, Ichinose F, Zapol WM, Ruvkun G, Barondeau DP, Mootha VK. Hypoxia rescues frataxin loss by restoring iron sulfur cluster biogenesis. Cell 177: 1507–1521.e16, 2019. doi: 10.1016/j.cell.2019.03.045. [DOI] [PMC free article] [PubMed] [Google Scholar]