Investigation
Investigation
Review
Physiological and Pathophysiological Roles of Mitochondrial
Na+-Ca2+ Exchanger, NCLX, in Hearts
Ayako Takeuchi 1,2, *              and Satoshi Matsuoka 1,2
                                          1   Department of Integrative and Systems Physiology, Faculty of Medical Sciences, University of Fukui,
                                              Fukui 910-1193, Japan; smatsuok@u-fukui.ac.jp
                                          2   Life Science Innovation Center, University of Fukui, Fukui 910-1193, Japan
                                          *   Correspondence: atakeuti@u-fukui.ac.jp; Tel.: +81-776-61-8311
                                          Abstract: It has been over 10 years since SLC24A6/SLC8B1, coding the Na+ /Ca2+ /Li+ exchanger
                                          (NCLX), was identified as the gene responsible for mitochondrial Na+ -Ca2+ exchange, a major
                                          Ca2+ efflux system in cardiac mitochondria. This molecular identification enabled us to determine
                                          structure–function relationships, as well as physiological/pathophysiological contributions, and our
                                          understandings have dramatically increased. In this review, we provide an overview of the recent
                                          achievements in relation to NCLX, focusing especially on its heart-specific characteristics, biophysical
                                          properties, and spatial distribution in cardiomyocytes, as well as in cardiac mitochondria. In addition,
                                          we discuss the roles of NCLX in cardiac functions under physiological and pathophysiological
                                          conditions—the generation of rhythmicity, the energy metabolism, the production of reactive oxygen
                                          species, and the opening of mitochondrial permeability transition pores.
                                          Keywords: mitochondria; heart; mitochondrial Na+ -Ca2+ exchanger; NCLX; metabolism; Ca2+ signaling
         
         
                              it has been well described functionally and structurally in other tissues or cell types (see
                              review [21]). These heart-specific characteristics of the mitochondrial Ca2+ influx system
                              may contribute to preventing mitochondrial Ca2+ overload in the heart, where cytosolic
                              Ca2+ periodically rises.
                                   In order to balance against the mitochondrial Ca2+ influx, NCXmit and H+ -Ca2+ ex-
                              change (HCXmit ) extrude Ca2+ from mitochondria, with the former accounting for the
                              major component in excitable tissues such as the heart and brain, and the latter being domi-
                              nant in non-excitable tissues such as the liver and kidney (see review [22]). Rysted et al. [23]
                              quantitatively compared the NCXmit activity in mitochondria isolated from mouse brains,
                              livers, and hearts. By evaluating extra-mitochondrial Ca2+ using Calcium Green-5N, they
                              demonstrated that the rate of Na+ -dependent Ca2+ efflux from mitochondria was ~3-fold
                              larger in the brain than in the heart. This well agrees with the lower CUmit activity in the
                              heart compared with other tissues [16,18]. Interestingly, the NCXmit activity in the liver
                              was negligible, despite the fact that it has the highest mRNA expression level of NCLX.
                              The authors attributed this to extra-mitochondrial expression of NCLX protein in the liver.
                                   In the heart, the fraction of Na+ -dependent Ca2+ efflux to total Ca2+ efflux is 60–100%,
                              depending on species and experimental conditions [8,23–25]. The remaining fraction
                              should be mediated by HCXmit , though its contribution in the heart has been controversial.
                              Leucine-zipper-EF hand-containing transmembrane (Letm1), which was initially shown to
                              mediate H+ -dependent Ca2+ influx into mitochondria [26,27], was proposed as the gene re-
                              sponsible for HCXmit . Natarajan et al. [28] detected H+ -induced Ca2+ efflux from rat cardiac
                              mitochondria, which were dependent on the free matrix Ca2+ concentration. Furthermore,
                              they confirmed Letm1-mediated Ca2+ efflux from mitochondria by demonstrating a dimin-
                              ished Ca2+ efflux rate in permeabilized H9c2 cells due to Letm1 knockdown. Interestingly,
                              they found that the expression level of the Letm1 protein in mitochondria was higher
                              in the heart than in the liver, though the functional contribution of HCXmit was much
                              higher in the liver than in the heart. Post-translational modifications or extra-mitochondrial
                              localization of Letm1 protein in the heart, just as reported for NCLX [23], may explain the
                              disparity between the expression level and function.
                                 substituted resides causing a NCLX-like phenotype, and found that peptides 248–255 were
                                 sensitive only to Li+ binding, but not to Na+ nor Ca2+ binding. Therefore, it is reasonable
                                 that the efficacy of exchanging for Ca2+ was different between Na+ and Li+ . Although
                                 the three-dimensional (3D) structure of NCLX has not been solved yet, recent advances
                                 in artificial intelligence-based structure prediction methods makes it possible to easily
                                 visualize a putative 3D structure of NCLX. Figure 1 shows a putative 3D structure of
                                 human NCLX (Q6J4K2), predicted using AlphaFold [36], with specific residues highlighted
                                 that are suggested to be functionally important.
                                       In the whole-mitoplast patch clamp experiments, the NCXmit current in reverse mode—
                                 an extra-mitochondrial Ca2+ -induced outward current with Na+ in the pipette—could not be
                                 recorded [8]. This was rather surprising to us because the reverse mode of NCXmit activity
                                 was previously reported to exist in mitochondria of rat cardiomyocytes [32,37]. Further
                                 evaluation of intra-mitochondrial Ca2+ using Fluo-8 in isolated mitochondria revealed that
                                 the reverse mode of NCXmit activity did exist in the heart. That is, CGP-37157-sensitive and
                                 intra-mitochondrial Na+ -dependent Ca2+ influx was detected, but the rate was too slow to
                                 be recorded electrophysiologically [8]. What is the mechanism underlying the slow NCXmit
                                 activity in reverse mode? One possible explanation may be an allosteric regulation of NCLX
                                 by ∆Ψ, as reported in SH-SY5Y neuronal cells and in HEK-293T cells [38]. The authors
                                 showed that mild ∆Ψ depolarization inhibited NCXmit via two clusters of positively charged
                                 residues, which are putatively located in the regulatory loop around the inner membrane
                                 (yellow sticks in Figure 1). They also showed that phosphorylation of S258 in human
                                 NCLX, known to be a protein kinase A (PKA) target site [39] (blue sticks in Figure 1), could
                                 override the regulation. Since mitoplasts and isolated mitochondria were free of cytosolic
                                 ingredients, it could be possible that phosphorylation at the residue was not sufficient to
                                 override the depolarization-mediated inhibition under the experimental conditions of [8].
                                 The unfavorable reversal of NCXmit was also reported in leukotriene C4 -stimulated mast
                                 cells with depolarized mitochondria [40]. Interestingly, however, mitochondrial fusion
                                 protein mitofusin (MFN) 2 knockdown caused repetitive reversal of NCXmit even under
                                 depolarized conditions, resulting in mitochondrial and cytosolic Ca2+ oscillation. It is worth
                                 examining phosphorylation status at NCLX S258 in MFN2-knockdown cells.
      Figure 1. Putative three-dimensional structure of human NCLX (UniProtKB accession number Q6J4K2) predicted using
      AlfaFold [36]. The pdb file (AF-Q6J4K2-F1-model_v1) was downloaded from the AlphaFold Protein Structure Database
      (https://alphafold.ebi.ac.uk/ accessed on
                                              on05
                                                 05 November 2021) and graphics were prepared using PyMOL v.2.1.0. (A) Side
      view, (B) bottom view. Putative mitochondria transit peptide and two sodium/calcium exchanger membrane regions are
      shown in green and pale and dark pink, respectively. Putative protein kinase A (PKA) phosphorylation ∆Ψ   site, S258 [39],
      is shown as blue sticks. Putative amino acids rendering Li+ selectivity, Na+ selectivity [34], and those sensitive to ∆Ψ
      depolarization [38] are shown as light blue, red, and yellow sticks, respectively.
Biomolecules 2021, 11, 1876                                                                                          4 of 14
                              it was suggested that CUmit activity was higher at mitochondria facing junctional SR (jSR)
                              than at those facing bulk cytosol, whereas Ca2+ efflux activity was comparable. This is
                              reasonable because mitochondria–jSR association creates high Ca2+ microdomains near the
                              dyadic space, which enables them to meet the low affinity of CUmit for Ca2+ uptake (see
                              review [54]).
                                   De La Fuente et al. [55,56] further explored the spatial heterogeneities of mitochondrial
                              Ca2+ dynamics. Using conventional and super-resolution immunofluorescence analyses of
                              isolated cardiac mitochondria and isolated cardiomyocytes, they demonstrated that about
                              50% of MCU were closely co-localized with the SR Ca2+ release channel ryanodine receptor
                              (RyR) 2 [55]. The authors explained that the divergence of this biased MCU distribution
                              from the previously reported uniform distribution [53] was attributable to the antibodies
                              chosen, since one used in [53] gave non-specific signals in MCU knockout mouse-derived
                              cardiomyocytes. Supporting the idea of MCU-RyR2 colocalization, MCU and EMRE, which
                              are essential CUmit regulator proteins, were more abundant in crude mitochondria than
                              in Percoll-purified mitochondria, and were also found in jSR [55]. On the other hand, the
                              NCLX protein was more abundant in pure mitochondria than in crude mitochondria, and
                              was not found in jSR [56]. Moreover, the authors strengthened their findings on distinct
                              distributions of MCU and NCLX by means of functional assays. CUmit activity—CUmit
                              inhibitor Ru360-sensitive 45 Ca2+ uptake corrected with citrate synthase activity—was
                              much higher in isolated jSR than that in isolated mitochondria. On the contrary, 45 Ca2+
                              retention assays revealed that Na+ - and CGP-37157-sensitive mitochondrial Ca2+ efflux
                              activity was much higher in pure mitochondria than that in jSR. This 45 Ca2+ efflux activity
                              became larger and smaller in heart-specific NCLX overexpressing and knockout mice,
                              respectively. The authors proposed that the spatially separated distribution of MCU-
                              RyR2 and NCLX contributes to minimizing the energy cost for maintaining ∆Ψ. In other
                              word, if MCU-RyR2 were near NCLX, ∆Ψ would depolarize both due to Ca2+ influx via
                              CUmit and due to Ca2+ efflux via NCXmit . Accordingly, the spatial separation of MCU-
                              RyR2 and NCLX should be necessary for optimizing mitochondrial Ca2+ signals and
                              energy cost. Interestingly, it was demonstrated that NCLX efficiently supplies Ca2+ from
                              mitochondria to the SR/endoplasmic reticulum (ER) via SERCA, thereby regulating the
                              automaticity of HL-1 cardiomyocytes, as well as antigen receptor-mediated Ca2+ signaling
                              of B lymphocytes [3,5]. It is worth evaluating the physical coupling of NCLX and SERCA
                              in cardiomyocytes, which would fill in the last piece in our understanding of the efficient
                              Ca2+ cycling between SR and mitochondria.
                                    Considering that HL-1 cells are derived from atrial myocytes, which are quiescent
                              under physiological conditions, NCXmit may be involved in abnormal automaticity of
                              atria, such as atrial flutter and atrial ectopic tachycardia. In addition, it may also be
                              plausible that abnormal NCXmit function causes ventricular arrhythmias. In fact, the
                              involvement of abnormal NCXmit activity in altered rhythmicity was suggested in mouse
                              embryonic stem cell-derived as well as in human induced pluripotent stem cell-derived
                              ventricular myocytes, where the “Ca2+ clock” drives the automaticity [59]. In addition,
                              arrhythmic evens with QRS interval widening were observed in tamoxifen-induced heart-
                              specific conditional NCLX-knockout mice, though the events only occurred immediately
                              before death [25].
                                    The question of whether NCXmit participates in the automaticity of normal pacemaker
                              cells, i.e., sinoatrial (SA) node cells, is still a big issue. The automaticity of SA node cells
                              has been proposed to be driven by a “coupled-clock” pacemaking system, which is com-
                              posed of a sarcolemmal ion channel/transporter-derived rhythm (“membrane clock”) and
                              subsarcolemmal Ca2+ release (LCR)-related rhythm (“Ca2+ clock”) [60–62]. In the former,
                              pacemaker channels such as the hyperpolarization-activated cation channel and various
                              other inward membrane currents at the plasma membrane drive diastolic depolarization.
                              In the latter, LCR from SR activates the inward current via sarcolemmal Na+ -Ca2+ exchange
                              to drive diastolic depolarization. NCXmit may modulate the “Ca2+ clock” part in SA node
                              cells, as observed in HL-1 cells [5]. In fact, application of an NCXmit inhibitor, CGP-37157,
                              slowed the firing rate of rabbit as well as mouse SA node cells [63,64]. However, recent
                              imaging studies of mouse SA node preparations revealed marked heterogeneity of LCR
                              and action potential-induced Ca2+ transients within and among SA node cells [65]. That
                              is, some SA node cells generated only LCR and did not fire; some only generated action
                              potential-induced Ca2+ transients and did not generate LCR; and some generated LCR
                              during the diastolic phase before an occurrence of action potential-induced Ca2+ transients.
                              These data suggest that the coupling degree of the “coupled-clock” system may differ
                              among SA node cells in vivo. Our model analyses suggested that NCXmit reduction in an
                              SA node cell which is solely driven by the “membrane clock” accelerates, instead of decel-
                              erating, the firing rate [11]. NCXmit reduction-mediated slowing of automaticity in “Ca2+
                              clock”-driven cells may be compromised by NCXmit -mediated acceleration of automaticity
                              in “membrane clock” cells in the SA node region. In fact, tamoxifen-induced NCLX deletion
                              in the adult mouse heart, with a 70% reduction of NCLX protein 3 days after tamoxifen
                              treatment, did not show altered sinus rhythms except for on the date of death, 8–10 days
                              after tamoxifen treatment [25]. In vivo imaging of the SA node of NCLX-knockout mice
                              would clarify the quantitative roles of NCXmit in pacemaking activity.
      Figure 2. An overview of the NCLX-mediated physiological and pathophysiological functions in a cardiomyocyte. ATPsyn,
                                                                                   ∆Ψ
      F1 Fo -ATP synthase; Cav 1.2, L-type Ca2+ channel; cytc, cytochrome c; ∆Ψ, mitochondrial  membrane potential; MCU,
      mitochondrial Ca2+ uniporter complex; mPTP, mitochondrial permeability transition pores; Nav 1.5, voltage-dependent
      Na+ channel; ROS, reactive oxygen species; RyR2, ryanodine receptor 2; SERCA, sarcoplasmic reticulum Ca2+ pump;
      SR, sarcoplasmic reticulum; TCA, tricarboxylic acid.
                                      This was first shown in whole-cell patch clamp experiments using guinea pig ven-
                                tricular myocytes loaded with Rhod-2 for evaluating mitochondrial Ca2+ changes [68].
                                It was demonstrated that when NCXmit became more active with 15 mM compared with
                                5 mM Na+ in the pipette, the mitochondrial Ca2+ increase induced by an abrupt workload
                                increase (3–4 Hz pacing in the presence of isoproterenol) was diminished. At the same
                                time, under the condition of 15 mM Na+ in the pipette, NADH autofluorescence decreased
                                upon the workload increase, indicating that mitochondrial Ca2+ was not sufficient enough
                                to activate NADH production by mitochondrial dehydrogenases. An NCXmit inhibitor,
                                CGP-37157, restored the workload-induced Ca2+ accumulation in mitochondria and at-
                                tenuated the NADH decrease [69]. Since a cytosolic Na+ increase and energy starvation
                                are characteristic properties of failing heart [70], the authors further studied a guinea pig
                                model of heart failure which was induced by aortic constriction with/without β-adrenergic
                                receptor stimulation [69,71]. In ventricular myocytes from failing hearts, where cytosolic
                                Na+ evaluated from SBFI ratio image was ~15 mM compared to ~5 mM in sham myocytes,
                                the abrupt workload increase caused essentially the same responses of mitochondrial Ca2+
                                (Rhod-2 or Myticam) and NADH as those reported with 15 mM Na+ in the pipette [68,69]—
                                the diminished increase of mitochondrial Ca2+ and the subsequent NADH starvation upon
                                the workload increase. More importantly, the changes were restored in the presence of
                                an NCXmit inhibitor, CGP-37157, to levels similar to those observed in sham myocytes.
                                These results suggested a causative role of NCXmit in the energy starvation of the failing
                                heart. In addition, as will be explained in Section 5.3, chronic treatment of the animals with
                                CGP-37157 partially prevented cardiac dysfunctions. Accordingly, the authors proposed
                                that blocking of NCXmit is a novel strategy for treating heart failure [71].
                                      However, the contribution of NCXmit to cardiac energetics in the failing heart may not
                                be as large as that expected from experiments using cardiomyocytes, where an extreme
                                workload change was applied—rapid 3–4 Hz pacing from a quiescent state, which hearts
                                in situ never experience [68,69,71]. Recently, the effects of chronic and acute myocardial
                                Na+ loads on cardiac energetics were extensively studied in Langendorff-perfused mouse
                                hearts with 23 Na, 31 P, 13 C NMR, and 1 H-NMR metabolomic profiling [72]. Chronic (phosp-
Biomolecules 2021, 11, 1876                                                                                           8 of 14
                              holemman PLM3SA mouse) and acute (treatment with ouabain and blebbistatin) inhibition
                              of Na+ -K+ ATPase, as well as pressure-overload-induced cardiac hypertrophy caused a cy-
                              tosolic Na+ increase, and switched the substrate preference from fatty acid to carbohydrate
                              oxidation, which are characteristic features frequently observed in failing hearts [70,73].
                              The acute Na+ elevation resulted in the most severe metabolic alterations, such as de-
                              creased metabolite levels of tricarboxylic acid (TCA) cycle intermediates downstream
                              from OGDH (succinate, fumarate, and malate), suggesting the reduced Ca2+ -dependent
                              activation of TCA cycle dehydrogenases. However, regardless of the strategy for cytosolic
                              Na+ elevations, the energy supply was maintained, as is evident from the preserved ATP,
                              phosphocreatine (PCr), PCr/ATP ratio, NADH, and pH. Metabolome profiles obtained
                              with NMR, as well as in silico predictions using CardiNet, revealed that they were achieved
                              at the cost of extensive metabolic flux remodeling. Therefore, the impact of impaired
                              cytosolic Na+ homeostasis on mitochondrial ATP production should be mechanistically
                              more complex than what has been suggested in isolated cardiomyocytes. In all three sets
                              of hearts with elevated cytosolic Na+ , treatment with CGP-37157 reversed the substrate
                              preference from carbohydrate to fatty acid oxidation with normalized levels of the depleted
                              metabolites. This suggests a therapeutic potential for CGP-37157 in the treatment of the
                              metabolic reprograming that occurs before energetic impairments.
                                   To the contrary, detrimental contributions of NCXmit to cardiac energetics were not
                              suggested in heart-specific NCLX-overexpression mice [25]. There were no apparent dif-
                              ferences between control and NCLX-overexpression mice’s ventricular myocytes in terms
                              of the NAD+ /NADH ratio, oxidative phosphorylation evaluated by seahorse analyses
                              with either pyruvate or palmitate as energy substrates (basal, ATP-linked and maximum
                              respirations, spare capacity, and proton leak), nor in the phosphorylation level of mitochon-
                              drial Ca2+ -responsive PDHC. These findings suggested that NCLX overexpression had
                              marginal effects on cardiac energetics. Rather, as will be explained in Section 5.3, NCLX
                              overexpression prevented the cardiac dysfunctions of ischemia-reperfusion injury and is-
                              chemic heart failure. It should be noted that the basal mitochondrial Ca2+ level, evaluated as
                              carbonyl cyanide 4-(trifluoromethoxy)phenylhydrazone)-responsive Fura-2 intensity, was
                              comparable between control and NCLX-overexpressing cardiomyocytes, indicating that
                              NCLX overexpression did not cause excessive deprivation of mitochondrial Ca2+ , in spite
                              of an increase in mitochondrial Ca2+ efflux rate by 88%. Our model analyses suggested
                              that cytosolic Ca2+ within its physiological range, 100 nM–2 µM, does not largely affect
                              steady-state levels of energy substrates, though a lower cytosolic Ca2+ level collapsed the
                              system because of mitochondrial Ca2+ deprivation [74,75]. Therefore, some compensation or
                              backup mechanisms may work to prevent mitochondrial Ca2+ deprivation via NCLX overex-
                              pression. It would be informative to further evaluate the mitochondrial Ca2+ level, cytosolic
                              Na+ level, and metabolome profiles in failing hearts with or without NCLX overexpression.
                              Ca2+ dynamics have been implicated to be closely associated with ROS dynamics in failing
                              and injured hearts [79,80].
                                    Hamilton et al. [81] demonstrated the involvement of NCXmit in ROS production,
                              SR Ca2+ handling, and arrhythmogenesis in rat ventricular myocytes. They monitored
                              mitochondrial Ca2+ using a biosensor mtRCamp1h, and showed that NCXmit inhibi-
                              tion by CGP-37157 decelerated mitochondrial Ca2+ decay, thereby enhancing mitochon-
                              drial Ca2+ accumulation triggered by 2 Hz electrical stimulation in the presence of iso-
                              proterenol. This resulted in larger ∆Ψ depolarization monitored by TMRM, increased
                              ROS in the mitochondria-SR microdomain evaluated using ER-tuned redox sensor ER-
                              roGFP_iE, increased RyR oxidation as evident from increased immunoprecipitation with
                              anti-dinitrophenyl-antibody, and increased proarrhythmic Ca2+ waves. The authors also
                              showed that this cascade further exacerbated proarrhythmic-triggered activity in hypertro-
                              phied hearts, which were induced by thoracic aortic banding.
                                    The detrimental consequences of NCXmit inhibition were more prominent in NCLX-
                              knockout mice. The germline NCLX knockout was unsuccessful, and adult acute heart-
                              specific NCLX knockout, in which NCLX protein expression was reduced by ~70%, caused
                              ~87% lethality within 2 weeks due to severe myocardial dysfunction accompanying
                              increased ROS, evaluated with dihydroethidium and MitoSox red, and mitochondrial
                              swelling [25]. This lethality was attributable to mitochondrial Ca2+ overload-mediated
                              mPTP opening, because the depletion of the mPTP component cyclophilin D on the NCLX
                              conditional knockout background rescued the myocardial dysfunction and lethality fol-
                              lowing tamoxifen-induced NCLX ablation. Those authors suggested that NCLX-mediated
                              Ca2+ efflux was necessary to maintain an appropriate mitochondrial Ca2+ level, which
                              was vital for preventing mPTP opening and excessive ROS production, and for survival.
                              The idea was further confirmed in heart-specific NCLX-overexpression mice subjected to
                              ischemia-reperfusion [25]. Accordingly, NCLX overexpression reduced the ROS level eval-
                              uated using dihydroethidium in hearts with 40 min-left coronary artery ligation followed
                              by 24 h reperfusion, and tended to decrease it 4 weeks after permanent occlusion of the
                              left coronary artery. In addition, cardiac dysfunctions characterized by TUNEL-positive
                              interstitial cells, fibrosis, and contractile dysfunction were all improved by NCLX over-
                              expression. The above findings clearly indicated beneficial contributions of NCXmit in
                              ischemia-induced failing hearts.
                                    However, a contradictory mechanism was proposed by Liu et al. [71]. As explained in
                              Section 5.2, an abrupt workload increase resulted in a diminished increase in mitochondrial
                              Ca2+ , followed by NADH starvation, in failing ventricular myocytes, possibly because
                              elevated cytosolic Na+ excessively extruded Ca2+ from mitochondria via NCXmit [71].
                              Interestingly, dichlorodihydrofluorescein diacetate oxidation, an index of the ROS level,
                              was dramatically increased upon an abrupt workload increase in the failing cardiomyocytes
                              but not in the sham cardiomyocytes and this ROS production was completely diminished
                              in the presence of an NCXmit inhibitor, CGP-37157. Moreover, chronic treatment of the
                              animals with CGP-37157 using an osmotic pump partially prevented the animals from
                              developing heart failure, as evident from improved hypertrophic remodeling, interstitial
                              fibrosis, contractile dysfunction, and occurrence of arrhythmia. The authors attributed the
                              mechanism to reduced ROS scavenging capacity due to the reduced NAD(P)H levels in
                              failing cardiomyocytes. Accordingly, these findings indicated a detrimental contribution of
                              NCXmit in failing hearts.
                                    The abovementioned contradictory roles of NCXmit in failing hearts suggested that
                              mitochondrial Ca2+ did not simply correlate with ROS production. Recently, a brand-new
                              mechanism underlying hypoxia-induced ROS production via NCXmit was proposed—
                              Na+ –phospholipid interaction-mediated ROS regulation [82] (Figure 2). The authors first
                              confirmed that NCXmit was involved in hypoxia-induced ROS production in primary
                              bovine aortic endothelial cells and mouse embryonic fibroblasts. Pharmacological inhi-
                              bition with CGP-37157 or genetical reduction (siRNA or knockout) of NCLX diminished
                              the cytosolic Ca2+ increase and cytosolic Na+ decrease, attenuated the reduction of the
Biomolecules 2021, 11, 1876                                                                                             10 of 14
                              inner mitochondrial membrane fluidity and the mitochondrial ROS production caused
                              by exposure of the cells to hypoxia (exposure of the cells to 1% O2 ). Then, the authors
                              showed that hypoxia-induced matrix acidification via Complex I inhibition caused Ca2+
                              solubilization from calcium phosphate precipitation in the matrix, as evident from mor-
                              phological (electron microscopy images) as well as from functional assays (measurements
                              of free mitochondrial Ca2+ in isolated mitochondria as well as in cells). Since a mitochon-
                              drial free Ca2+ increase enables NCXmit to extrude Ca2+ in exchange for Na+ , the authors
                              then focused on the roles of matrix Na+ on electron transport chains and found that only
                              Complex II-dependent respirations were decreased by Na+ , which was NCXmit -dependent,
                              resulting in increased ROS production. The authors filled in the final piece by showing
                              that Na+ directly bound to the phospholipid bilayer, as evident from infrared spectroscopy,
                              which reduced the fluidity of the inner mitochondrial membrane for ubiquinone diffusion
                              in the inner mitochondrial membrane, increasing the ROS production. Taken together,
                              these findings clarified a distinct scheme of ROS production—regulation by matrix Na+ via
                              NCXmit —from those proposed in previous reports.
                              6. Future Perspectives
                                   As has been described so far, knowledge on the biophysical properties, distributions,
                              and the physiological and pathophysiological significance of NCXmit in the heart is rapidly
                              increasing. The more knowledge is accumulated, the more complicated systems are elu-
                              cidated, sometimes introducing difficulties into our understanding as a whole. Taking
                              ROS dynamics under pathological conditions as an example, some experimental evidence
                              supports the roles of NCXmit in increasing ROS production [82], whereas others support its
                              preventive roles in relation to ROS increases [25,81]. NCXmit directly modulates and is af-
                              fected by cytosolic and mitochondrial concentrations of Na+ and Ca2+ ions, and ∆Ψ, which
                              are associated with ROS balance regulation via different pathways (Figure 2). Therefore,
                              differences in ionic conditions and mitochondrial viability under different experimental
                              conditions or diseased states would result in different contributions of NCXmit .
                                   In order to understand these complicated networks, the integration of NCXmit activity
                              and cellular/mitochondrial functions with mathematical modeling could be a powerful tool.
                              Very recently, Cortassa et al. [83] succeeded in reconciling the apparently paradoxical roles
                              of NCXmit in ROS dynamics (see details [83]). In brief, they built two scenarios, the “Na+ -
                              driven oxidized scenario” and the “Ca2+ -driven reduced scenario”, and demonstrated
                              that variations in redox status, cytoplasmic Na+ concentrations and energetic capacity
                              resulted in different mitochondrial Ca2+ levels and bioenergetic responses driving ATP
                              supply and oxidative stress. The former scenario could be represented by heart failure with
                              a reduced ejection fraction (HFrEF) in which considerable cytosolic Na+ overload occurs,
                              and the latter by heart failure with a preserved or moderate ejection fraction (HFpEF,
                              HFmEF) in which only a modest Na+ increase is expected. Integrating the model of matrix
                              Ca2+ solubilization and precipitation from and to calcium phosphate [84] into Cortassa’s
                              model [83] would further facilitate our understandings in this area.
                                   The discrepancies in experimental findings obtained from isolated mitochondria,
                              isolated cardiomyocytes, and whole hearts are other issues that remain to be solved.
                              Recent advances in imaging techniques used to evaluate electrophysiological and metabolic
                              properties of single cells and even organelles in tissue are promising [65,85]. By utilizing
                              these techniques, it is expected that our understandings of the roles of NCXmit in healthy
                              as well as in failing hearts will be further deepened.
                              Author Contributions: Conceptualization, A.T. and S.M.; writing—original draft preparation, A.T.;
                              writing—review and editing, S.M.; visualization, A.T. and S.M.; funding acquisition, A.T. and S.M.
                              All authors have read and agreed to the published version of the manuscript.
                              Funding: This research was funded by JSPS KAKENHI (grant number 18K06869 (A.T.) and 19H03400
                              (S.M.)) and by Research Grant from University of Fukui (grant number LSI21205 (S.M.)).
                              Institutional Review Board Statement: Not applicable.
Biomolecules 2021, 11, 1876                                                                                                                       11 of 14
References
1.    Carafoli, E.; Tiozzo, R.; Lugli, G.; Crovetti, F.; Kratzing, C. The release of calcium from heart mitochondria by sodium. J. Mol. Cell
      Cardiol. 1974, 6, 361–371. [CrossRef]
2.    Palty, R.; Silverman, W.F.; Hershfinkel, M.; Caporale, T.; Sensi, S.L.; Parnis, J.; Nolte, C.; Fishman, D.; Shoshan-Barmatz, V.; Herrmann, S.; et al.
      NCLX is an essential component of mitochondrial Na+/Ca2+ exchange. Proc. Natl. Acad. Sci. USA 2010, 107, 436–441. [CrossRef] [PubMed]
3.    Kim, B.; Takeuchi, A.; Koga, O.; Hikida, M.; Matsuoka, S. Pivotal role of mitochondrial Na+ -Ca2+ exchange in antigen receptor
      mediated Ca2+ signalling in DT40 and A20 B lymphocytes. J. Physiol. 2012, 590, 459–474. [CrossRef] [PubMed]
4.    Nita, I.I.; Hershfinkel, M.; Fishman, D.; Ozeri, E.; Rutter, G.A.; Sensi, S.L.; Khananshvili, D.; Lewis, E.C.; Sekler, I. The
      mitochondrial Na+ /Ca2+ exchanger upregulates glucose dependent Ca2+ signalling linked to insulin secretion. PLoS ONE 2012,
      7, e46649. [CrossRef] [PubMed]
5.    Takeuchi, A.; Kim, B.; Matsuoka, S. The mitochondrial Na+ -Ca2+ exchanger, NCLX, regulates automaticity of HL-1 cardiomy-
      ocytes. Sci. Rep. 2013, 3, 2766. [CrossRef] [PubMed]
6.    Parnis, J.; Montana, V.; Delgado-Martinez, I.; Matyash, V.; Parpura, V.; Kettenmann, H.; Sekler, I.; Nolte, C. Mitochondrial
      exchanger NCLX plays a major role in the intracellular Ca2+ signaling, gliotransmission, and proliferation of astrocytes. J. Neurosci.
      2013, 33, 7206–7219. [CrossRef]
7.    Kim, B.; Takeuchi, A.; Hikida, M.; Matsuoka, S. Roles of the mitochondrial Na+ -Ca2+ exchanger, NCLX, in B lymphocyte
      chemotaxis. Sci. Rep. 2016, 6, 28378. [CrossRef] [PubMed]
8.    Islam, M.M.; Takeuchi, A.; Matsuoka, S. Membrane current evoked by mitochondrial Na+ -Ca2+ exchange in mouse heart.
      J. Physiol. Sci. 2020, 70, 24. [CrossRef]
9.    Assali, E.A.; Jones, A.E.; Veliova, M.; Acin-Perez, R.; Taha, M.; Miller, N.; Shum, M.; Oliveira, M.F.; Las, G.; Liesa, M.; et al. NCLX
      prevents cell death during adrenergic activation of the brown adipose tissue. Nat. Commun. 2020, 11, 3347. [CrossRef] [PubMed]
10.   Takeuchi, A.; Matsuoka, S. Minor contribution of NCX to Na+ -Ca2+ exchange activity in brain mitochondria. Cell Calcium. 2021,
      96, 102386. [CrossRef] [PubMed]
11.   Takeuchi, A.; Kim, B.; Matsuoka, S. The destiny of Ca2+ released by mitochondria. J. Physiol. Sci. 2015, 65, 11–24. [CrossRef]
      [PubMed]
12.   Takeuchi, A.; Kim, B.; Matsuoka, S. Physiological functions of mitochondrial Na+ -Ca2+ exchanger, NCLX, in lymphocytes. Cell
      Calcium 2020, 85, 102114. [CrossRef]
13.   Katoshevski, T.; Ben-Kasus Nissim, T.; Sekler, I. Recent studies on NCLX in health and diseases. Cell Calcium 2021, 94, 102345.
      [CrossRef]
14.   Brown, D.A.; Perry, J.B.; Allen, M.E.; Sabbah, H.N.; Stauffer, B.L.; Shaikh, S.R.; Cleland, J.G.; Colucci, W.S.; Butler, J.; Voors, A.A.; et al.
      Expert consensus document: Mitochondrial function as a therapeutic target in heart failure. Nat. Rev. Cardiol. 2017, 14, 238–250.
      [CrossRef] [PubMed]
15.   O’Rourke, B.; Ashok, D.; Liu, T. Mitochondrial Ca2+ in heart failure: Not enough or too much? J. Mol. Cell Cardiol. 2021, 151, 126–134.
      [CrossRef]
16.   Fieni, F.; Lee, S.B.; Jan, Y.N.; Kirichok, Y. Activity of the mitochondrial calcium uniporter varies greatly between tissues. Nat.
      Commun. 2012, 3, 1317. [CrossRef] [PubMed]
17.   Raffaello, A.; De Stefani, D.; Sabbadin, D.; Teardo, E.; Merli, G.; Picard, A.; Checchetto, V.; Moro, S.; Szabò, I.; Rizzuto, R. The
      mitochondrial calcium uniporter is a multimer that can include a dominant-negative pore-forming subunit. EMBO J. 2013, 32,
      2362–2376. [CrossRef]
18.   Paillard, M.; Csordás, G.; Szanda, G.; Golenár, T.; Debattisti, V.; Bartok, A.; Wang, N.; Moffat, C.; Seifert, E.L.; Spät, A.; et al.
      Tissue-specific mitochondrial decoding of cytoplasmic Ca2+ signals is controlled by the stoichiometry of MICU1/2 and MCU.
      Cell Rep. 2017, 18, 2291–2300. [CrossRef] [PubMed]
19.   Huo, J.; Lu, S.; Kwong, J.Q.; Bround, M.J.; Grimes, K.M.; Sargent, M.A.; Brown, M.E.; Davis, M.E.; Bers, D.M.; Molkentin, J.D.
      MCUb induction protects the heart from postischemic remodeling. Circ. Res. 2020, 127, 379–390. [CrossRef]
20.   Wescott, A.P.; Kao, J.P.Y.; Lederer, W.J.; Boyman, L. Voltage-energized calcium-sensitive ATP production by mitochondria. Nat.
      Metab. 2019, 1, 975–984. [CrossRef] [PubMed]
21.   Pallafacchina, G.; Zanin, S.; Rizzuto, R. From the identification to the dissection of the physiological role of the mitochondrial
      calcium uniporter: An ongoing story. Biomolecules 2021, 11. [CrossRef] [PubMed]
22.   Bernardi, P. Mitochondrial transport of cations: Channels, exchangers, and permeability transition. Physiol. Rev. 1999, 79, 1127–1155.
      [CrossRef] [PubMed]
23.   Rysted, J.E.; Lin, Z.; Walters, G.C.; Rauckhorst, A.J.; Noterman, M.; Liu, G.; Taylor, E.B.; Strack, S.; Usachev, Y.M. Distinct
      properties of Ca2+ efflux from brain, heart and liver mitochondria: The effects of Na+ , Li+ and the mitochondrial Na+ /Ca2+
      exchange inhibitor CGP37157. Cell Calcium 2021, 96, 102382. [CrossRef]
Biomolecules 2021, 11, 1876                                                                                                                12 of 14
24.   Crompton, M.; Künzi, M.; Carafoli, E. The calcium-induced and sodium-induced effluxes of calcium from heart mitochondria.
      Evidence for a sodium-calcium carrier. Eur. J. Biochem. 1977, 79, 549–558. [CrossRef] [PubMed]
25.   Luongo, T.S.; Lambert, J.P.; Gross, P.; Nwokedi, M.; Lombardi, A.A.; Shanmughapriya, S.; Carpenter, A.C.; Kolmetzky, D.; Gao, E.;
      van Berlo, J.H.; et al. The mitochondrial Na+ /Ca2+ exchanger is essential for Ca2+ homeostasis and viability. Nature 2017, 545, 93–97.
      [CrossRef] [PubMed]
26.   Jiang, D.; Zhao, L.; Clapham, D.E. Genome-wide RNAi screen identifies Letm1 as a mitochondrial Ca2+ /H+ antiporter. Science
      2009, 326, 144–147. [CrossRef]
27.   Jiang, D.; Zhao, L.; Clish, C.B.; Clapham, D.E. 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 2013, 110, E2249–E2254. [CrossRef]
28.   Natarajan, G.K.; Glait, L.; Mishra, J.; Stowe, D.F.; Camara, A.K.S.; Kwok, W.M. Total matrix Ca2+ modulates Ca2+ efflux via the
      Ca2+ /H+ exchanger in cardiac mitochondria. Front. Physiol. 2020, 11, 510600. [CrossRef] [PubMed]
29.   Crompton, M.; Capano, M.; Carafoli, E. The sodium-induced efflux of calcium from heart mitochondria. A possible mechanism
      for the regulation of mitochondrial calcium. Eur. J. Biochem. 1976, 69, 453–462. [CrossRef]
30.   Affolter, H.; Carafoli, E. The Ca2+ -Na+ antiporter of heart mitochondria operates electroneutrally. Biochem. Biophys. Res. Commun.
      1980, 95, 193–196. [CrossRef]
31.   Jung, D.W.; Baysal, K.; Brierley, G.P. The sodium-calcium antiport of heart mitochondria is not electroneutral. J. Biol. Chem. 1995,
      270, 672–678. [CrossRef] [PubMed]
32.   Kim, B.; Matsuoka, S. Cytoplasmic Na+ -dependent modulation of mitochondrial Ca2+ via electrogenic mitochondrial Na+ -Ca2+
      exchange. J. Physiol. 2008, 586, 1683–1697. [CrossRef]
33.   Cox, D.A.; Conforti, L.; Sperelakis, N.; Matlib, M.A. Selectivity of inhibition of Na+ -Ca2+ exchange of heart mitochondria by
      benzothiazepine CGP-37157. J. Cardiovasc. Pharm. 1993, 21, 595–599. [CrossRef] [PubMed]
34.   Roy, S.; Dey, K.; Hershfinkel, M.; Ohana, E.; Sekler, I. Identification of residues that control Li+ versus Na+ dependent Ca2+
      exchange at the transport site of the mitochondrial NCLX. Biochim Biophys. Acta Mol. Cell Res. 2017, 1864, 997–1008. [CrossRef]
35.   Giladi, M.; Lee, S.Y.; Refaeli, B.; Hiller, R.; Chung, K.Y.; Khananshvili, D. Structure-dynamic and functional relationships in a
      Li+ -transporting sodium-calcium exchanger mutant. Biochim Biophys. Acta Bioenerg. 2019, 1860, 189–200. [CrossRef]
36.   Jumper, J.; Evans, R.; Pritzel, A.; Green, T.; Figurnov, M.; Ronneberger, O.; Tunyasuvunakool, K.; Bates, R.; Židek, A.; Potapenko, A.; et al.
      Highly accurate protein structure prediction with alphafold. Nature 2021, 596, 583–589. [CrossRef] [PubMed]
37.   Griffiths, E.J. Reversal of mitochondrial Na/Ca exchange during metabolic inhibition in rat cardiomyocytes. FEBS Lett. 1999, 453, 400–404.
      [CrossRef]
38.   Kostic, M.; Katoshevski, T.; Sekler, I. Allosteric regulation of NCLX by mitochondrial membrane potential links the metabolic
      state and Ca2+ signaling in mitochondria. Cell Rep. 2018, 25, 3465–3475.e3464. [CrossRef]
39.   Kostic, M.; Ludtmann, M.H.; Bading, H.; Hershfinkel, M.; Steer, E.; Chu, C.T.; Abramov, A.Y.; Sekler, I. PKA phosphorylation
      of NCLX reverses mitochondrial calcium overload and depolarization, promoting survival of PINK1-deficient dopaminergic
      neurons. Cell Rep. 2015, 13, 376–386. [CrossRef]
40.   Samanta, K.; Mirams, G.R.; Parekh, A.B. Sequential forward and reverse transport of the Na+ Ca2+ exchanger generates Ca2+
      oscillations within mitochondria. Nat. Commun. 2018, 9, 156. [CrossRef]
41.   Gandhi, S.; Wood-Kaczmar, A.; Yao, Z.; Plun-Favreau, H.; Deas, E.; Klupsch, K.; Downward, J.; Latchman, D.S.; Tabrizi, S.J.;
      Wood, N.W.; et al. PINK1-associated Parkinson’s disease is caused by neuronal vulnerability to calcium-induced cell death. Mol.
      Cell 2009, 33, 627–638. [CrossRef] [PubMed]
42.   Pickrell, A.M.; Youle, R.J. The roles of PINK1, Parkin, and mitochondrial fidelity in Parkinson’s disease. Neuron 2015, 85, 257–273.
      [CrossRef]
43.   Kang, C.; Badr, M.A.; Kyrychenko, V.; Eskelinen, E.L.; Shirokova, N. Deficit in PINK1/PARKIN-mediated mitochondrial
      autophagy at late stages of dystrophic cardiomyopathy. Cardiovasc. Res. 2018, 114, 90–102. [CrossRef] [PubMed]
44.   Zhou, Q.; Xie, M.; Zhu, J.; Yi, Q.; Tan, B.; Li, Y.; Ye, L.; Zhang, X.; Zhang, Y.; Tian, J.; et al. PINK1 contained in huMSC-derived
      exosomes prevents cardiomyocyte mitochondrial calcium overload in sepsis via recovery of mitochondrial Ca2+ efflux. Stem Cell
      Res. 2021, 12, 269. [CrossRef]
45.   Glancy, B.; Hartnell, L.M.; Malide, D.; Yu, Z.X.; Combs, C.A.; Connelly, P.S.; Subramaniam, S.; Balaban, R.S. Mitochondrial
      reticulum for cellular energy distribution in muscle. Nature 2015, 523, 617–620. [CrossRef]
46.   Glancy, B.; Hartnell, L.M.; Combs, C.A.; Femnou, A.; Sun, J.; Murphy, E.; Subramaniam, S.; Balaban, R.S. Power grid protection of
      the muscle mitochondrial reticulum. Cell Rep. 2017, 19, 487–496. [CrossRef] [PubMed]
47.   Romashko, D.N.; Marban, E.; O’Rourke, B. Subcellular metabolic transients and mitochondrial redox waves in heart cells. Proc.
      Natl. Acad. Sci. USA 1998, 95, 1618–1623. [CrossRef] [PubMed]
48.   Kuznetsov, A.V.; Usson, Y.; Leverve, X.; Margreiter, R. Subcellular heterogeneity of mitochondrial function and dysfunction:
      Evidence obtained by confocal imaging. Mol. Cell Biochem. 2004, 256–257, 359–365. [CrossRef] [PubMed]
49.   Lu, X.; Thai, P.N.; Lu, S.; Pu, J.; Bers, D.M. Intrafibrillar and perinuclear mitochondrial heterogeneity in adult cardiac myocytes.
      J. Mol. Cell Cardiol. 2019, 136, 72–84. [CrossRef] [PubMed]
50.   Perez-Hernández, M.; Leo-Macias, A.; Keegan, S.; Jouni, M.; Kim, J.C.; Agullo-Pascual, E.; Vermij, S.; Zhang, M.; Liang, F.X.;
      Burridge, P.; et al. Structural and functional characterization of a Nav 1.5-mitochondrial couplon. Circ. Res. 2021, 128, 419–432.
      [CrossRef]
Biomolecules 2021, 11, 1876                                                                                                               13 of 14
51.   Aon, M.A.; Cortassa, S.; Marban, E.; O’Rourke, B. Synchronized whole cell oscillations in mitochondrial metabolism triggered by
      a local release of reactive oxygen species in cardiac myocytes. J. Biol. Chem. 2003, 278, 44735–44744. [CrossRef] [PubMed]
52.   Skogestad, J.; Lines, G.T.; Louch, W.E.; Sejersted, O.M.; Sjaastad, I.; Aronsen, J.M. Evidence for heterogeneous subsarcolemmal
      Na+ levels in rat ventricular myocytes. Am. J. Physiol. Heart Circ. Physiol. 2019, 316, H941–H957. [CrossRef]
53.   Lu, X.; Ginsburg, K.S.; Kettlewell, S.; Bossuyt, J.; Smith, G.L.; Bers, D.M. Measuring local gradients of intramitochondrial [Ca2+ ]
      in cardiac myocytes during sarcoplasmic reticulum Ca2+ release. Circ. Res. 2013, 112, 424–431. [CrossRef]
54.   Rizzuto, R.; De Stefani, D.; Raffaello, A.; Mammucari, C. Mitochondria as sensors and regulators of calcium signalling. Nat. Rev.
      Mol. Cell Biol. 2012, 13, 566–578. [CrossRef]
55.   De La Fuente, S.; Fernandez-Sanz, C.; Vail, C.; Agra, E.J.; Holmstrom, K.; Sun, J.; Mishra, J.; Williams, D.; Finkel, T.; Murphy, E.; et al.
      Strategic positioning and biased activity of the mitochondrial calcium uniporter in cardiac muscle. J. Biol. Chem. 2016, 291, 23343–23362.
      [CrossRef] [PubMed]
56.   De La Fuente, S.; Lambert, J.P.; Nichtova, Z.; Fernandez-Sanz, C.; Elrod, J.W.; Sheu, S.S.; Csordás, G. Spatial separation of
      mitochondrial calcium uptake and extrusion for energy-efficient mitochondrial calcium signaling in the heart. Cell Rep. 2018, 24,
      3099–3107.e3094. [CrossRef]
57.   Bassani, R.A.; Bassani, J.W.; Bers, D.M. Relaxation in ferret ventricular myocytes: Unusual interplay among calcium transport
      systems. J. Physiol. 1994, 476, 295–308. [CrossRef]
58.   Bassani, J.W.; Bassani, R.A.; Bers, D.M. Relaxation in rabbit and rat cardiac cells: Species-dependent differences in cellular
      mechanisms. J. Physiol. 1994, 476, 279–293. [CrossRef]
59.   Xie, A.; Zhou, A.; Liu, H.; Shi, G.; Liu, M.; Boheler, K.R.; Dudley, S.C., Jr. Mitochondrial Ca2+ flux modulates spontaneous
      electrical activity in ventricular cardiomyocytes. PLoS ONE 2018, 13, e0200448. [CrossRef]
60.   Mangoni, M.E.; Nargeot, J. Genesis and regulation of the heart automaticity. Physiol. Rev. 2008, 88, 919–982. [CrossRef]
61.   Lakatta, E.G.; Maltsev, V.A.; Vinogradova, T.M. A coupled system of intracellular Ca2+ clocks and surface membrane voltage
      clocks controls the timekeeping mechanism of the heart’s pacemaker. Circ. Res. 2010, 106, 659–673. [CrossRef]
62.   Monfredi, O.; Maltsev, V.A.; Lakatta, E.G. Modern concepts concerning the origin of the heartbeat. Physiology 2013, 28, 74–92.
      [CrossRef]
63.   Yaniv, Y.; Spurgeon, H.A.; Lyashkov, A.E.; Yang, D.; Ziman, B.D.; Maltsev, V.A.; Lakatta, E.G. Crosstalk between mitochondrial
      and sarcoplasmic reticulum Ca2+ cycling modulates cardiac pacemaker cell automaticity. PLoS ONE 2012, 7, e37582. [CrossRef]
64.   Takeda, Y.; Matsuoka, S. Impact of mitochondria on local calcium release in murine sinoatrial nodal cells. J. Mol. Cell Cardiol.
      2021, 164, 42–50. [CrossRef] [PubMed]
65.   Bychkov, R.; Juhaszova, M.; Tsutsui, K.; Coletta, C.; Stern, M.D.; Maltsev, V.A.; Lakatta, E.G. Synchronized cardiac impulses
      emerge from heterogeneous local calcium signals within and among cells of pacemaker tissue. JACC Clin. Electrophysiol. 2020, 6,
      907–931. [CrossRef]
66.   Katz, A. Physiology of the Heart; Lippincott Williams & Wilkins: Philadelphia, PA, USA, 2010.
67.   McCormack, J.G.; Halestrap, A.P.; Denton, R.M. Role of calcium ions in regulation of mammalian intramitochondrial metabolism.
      Physiol. Rev. 1990, 70, 391–425. [CrossRef] [PubMed]
68.   Maack, C.; Cortassa, S.; Aon, M.A.; Ganesan, A.N.; Liu, T.; O’Rourke, B. Elevated cytosolic Na+ decreases mitochondrial Ca2+
      uptake during excitation-contraction coupling and impairs energetic adaptation in cardiac myocytes. Circ. Res. 2006, 99, 172–182.
      [CrossRef] [PubMed]
69.   Liu, T.; O’Rourke, B. Enhancing mitochondrial Ca2+ uptake in myocytes from failing hearts restores energy supply and demand
      matching. Circ. Res. 2008, 103, 279–288. [CrossRef]
70.   Ventura-Clapier, R.; Garnier, A.; Veksler, V.; Joubert, F. Bioenergetics of the failing heart. Biochim Biophys. Acta 2011, 1813,
      1360–1372. [CrossRef]
71.   Liu, T.; Takimoto, E.; Dimaano, V.L.; DeMazumder, D.; Kettlewell, S.; Smith, G.; Sidor, A.; Abraham, T.P.; O’Rourke, B. Inhibiting
      mitochondrial Na+ /Ca2+ exchange prevents sudden death in a guinea pig model of heart failure. Circ. Res. 2014, 115, 44–54.
      [CrossRef]
72.   Aksentijević, D.; Karlstaedt, A.; Basalay, M.V.; O’Brien, B.A.; Sanchez-Tatay, D.; Eminaga, S.; Thakker, A.; Tennant, D.A.; Fuller,
      W.; Eykyn, T.R.; et al. Intracellular sodium elevation reprograms cardiac metabolism. Nat. Commun. 2020, 11, 4337. [CrossRef]
73.   Carley, A.N.; Taegtmeyer, H.; Lewandowski, E.D. Matrix revisited: Mechanisms linking energy substrate metabolism to the
      function of the heart. Circ. Res. 2014, 114, 717–729. [CrossRef]
74.   Saito, R.; Takeuchi, A.; Himeno, Y.; Inagaki, N.; Matsuoka, S. A simulation study on the constancy of cardiac energy metabolites
      during workload transition. J. Physiol. 2016, 594, 6929–6945. [CrossRef]
75.   Takeuchi, A.; Matsuoka, S. Integration of mitochondrial energetics in heart with mathematical modelling. J. Physiol. 2020, 598,
      1443–1457. [CrossRef]
76.   Chen, Y.R.; Zweier, J.L. Cardiac mitochondria and reactive oxygen species generation. Circ. Res. 2014, 114, 524–537. [CrossRef]
77.   Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol. Rev.
      2014, 94, 909–950. [CrossRef]
78.   Dey, S.; DeMazumder, D.; Sidor, A.; Foster, D.B.; O’Rourke, B. Mitochondrial ROS drive sudden cardiac death and chronic
      proteome remodeling in heart failure. Circ. Res. 2018, 123, 356–371. [CrossRef]
Biomolecules 2021, 11, 1876                                                                                                         14 of 14
79.   Santulli, G.; Xie, W.; Reiken, S.R.; Marks, A.R. Mitochondrial calcium overload is a key determinant in heart failure. Proc. Natl.
      Acad. Sci. USA 2015, 112, 11389–11394. [CrossRef]
80.   Beretta, M.; Santos, C.X.; Molenaar, C.; Hafstad, A.D.; Miller, C.C.; Revazian, A.; Betteridge, K.; Schröder, K.; Streckfuss-Bömeke,
      K.; Doroshow, J.H.; et al. Nox4 regulates InsP3 receptor-dependent Ca2+ release into mitochondria to promote cell survival.
      EMBO J. 2020, 39, e103530. [CrossRef]
81.   Hamilton, S.; Terentyeva, R.; Kim, T.Y.; Bronk, P.; Clements, R.T.; O-Uchi, J.; Csordás, G.; Choi, B.R.; Terentyev, D. Pharmacological
      modulation of mitochondrial Ca2+ content regulates sarcoplasmic reticulum Ca2+ release via oxidation of the ryanodine receptor
      by mitochondria-derived reactive oxygen species. Front. Physiol. 2018, 9, 1831. [CrossRef]
82.   Hernansanz-Agustín, P.; Choya-Foces, C.; Carregal-Romero, S.; Ramos, E.; Oliva, T.; Villa-Piña, T.; Moreno, L.; Izquierdo-Álvarez,
      A.; Cabrera-García, J.D.; Cortés, A.; et al. Na+ controls hypoxic signalling by the mitochondrial respiratory chain. Nature 2020,
      586, 287–291. [CrossRef]
83.   Cortassa, S.; Juhaszova, M.; Aon, M.A.; Zorov, D.B.; Sollott, S.J. Mitochondrial Ca2+ , redox environment and ROS emission in
      heart failure: Two sides of the same coin? J. Mol. Cell Cardiol. 2021, 151, 113–125. [CrossRef]
84.   Malyala, S.; Zhang, Y.; Strubbe, J.O.; Bazil, J.N. Calcium phosphate precipitation inhibits mitochondrial energy metabolism. PLoS
      Comput. Biol. 2019, 15, e1006719. [CrossRef]
85.   Kosmach, A.; Roman, B.; Sun, J.; Femnou, A.; Zhang, F.; Liu, C.; Combs, C.A.; Balaban, R.S.; Murphy, E. Monitoring mitochondrial
      calcium and metabolism in the beating MCU-KO heart. Cell Rep. 2021, 37, 109846. [CrossRef]