Next Article in Journal
Which Low-Abundance Proteins are Present in the Human Milieu of Gamete/Embryo Maternal Interaction?
Next Article in Special Issue
Renal Ca2+ and Water Handling in Response to Calcium Sensing Receptor Signaling: Physiopathological Aspects and Role of CaSR-Regulated microRNAs
Previous Article in Journal
The Positive Effects of Grifola frondosa Heteropolysaccharide on NAFLD and Regulation of the Gut Microbiota
Previous Article in Special Issue
ALS-Associated SOD1(G93A) Decreases SERCA Pump Levels and Increases Store-Operated Ca2+ Entry in Primary Spinal Cord Astrocytes from a Transgenic Mouse Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pathophysiology of Calcium Mediated Ventricular Arrhythmias and Novel Therapeutic Options with Focus on Gene Therapy

1
Department of Internal Medicine II, Paracelsus Medical University, 5020 Salzburg, Austria
2
Division of Cardiology, Hospital Maria Hilf Moenchengladbach, 41063 Moenchengladbach, Germany
3
Department of Internal Diseases, Bashkir State Medical University, 450000 Ufa, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2019, 20(21), 5304; https://doi.org/10.3390/ijms20215304
Submission received: 14 September 2019 / Revised: 16 October 2019 / Accepted: 21 October 2019 / Published: 24 October 2019
(This article belongs to the Special Issue Calcium Signaling in Human Health and Diseases 2.0)

Abstract

:
Cardiac arrhythmias constitute a major health problem with a huge impact on mortality rates and health care costs. Despite ongoing research efforts, the understanding of the molecular mechanisms and processes responsible for arrhythmogenesis remains incomplete. Given the crucial role of Ca2+-handling in action potential generation and cardiac contraction, Ca2+ channels and Ca2+ handling proteins represent promising targets for suppression of ventricular arrhythmias. Accordingly, we report the different roles of Ca2+-handling in the development of congenital as well as acquired ventricular arrhythmia syndromes. We highlight the therapeutic potential of gene therapy as a novel and innovative approach for future arrhythmia therapy. Furthermore, we discuss various promising cellular and mitochondrial targets for therapeutic gene transfer currently under investigation.

Cardiac arrhythmias constitute a major public health concern worldwide [1]. Several severe types of arrhythmias are responsible for syncope and sudden cardiac death (SCD). Ventricular fibrillation and tachycardia are either based on genetic mutations concerning ion channel expression and function or on arrhythmogenic substrates that may be caused by heart failure (HF), infarction or drugs that modify ion channel behavior [2,3]. The incomplete understanding of arrhythmia formation and their underlying mechanisms in the failing heart has hindered the development of safe and effective pharmacological treatments [4]. In fact, some pharmacological agents even constitute a pro-arrhythmic tendency, which may lead to arrhythmia-related deaths. For instance, the Cardiac Arrhythmia Suppression Trial (CAST), which was terminated prematurely after a substantial increase in arrhythmic deaths of patients, treated with either encainide or flecainide compared to placebo [5]. These pro-arrhythmic effects may be based upon the heterogeneity of cardiac ion channel expression and function within the different regions and layers of the heart [6,7], the unspecific nature of multiple pharmacological agents targeting ion channels, the complexity of ion channel behavior and their cross-talk, or the dynamic cellular and environmental remodeling properties caused by cardiac disease progression [8].
Therefore, investigations on ionic currents and channels, and their influence on action potential (AP) waveforms are of major interest for understanding arrhythmogenesis. Calcium (Ca2+) homeostasis plays a crucial role in AP generation and correct cardiac contraction. Several studies have shown that cellular calcium handling is altered in the pathophysiology of heart diseases, such as HF and cardiomyopathies [9]. Additionally, Ca2+ handling may also be altered by inherited mutations that directly affect Ca2+ channels or Ca2+ handling proteins, such as ryanodine receptor type 2 (RyR2) [10], calsequestrin (CASQ) [11] and calmodulin (CaM) [12]. Furthermore, the role of myocardial mitochondria in Ca2+ handling gets more into focus of current studies [13,14,15]. In this review we will focus on ventricular arrhythmias caused by defective Ca2+ currents or channels and novel therapeutic options.

1. The Role of Calcium in Action Potential Generation

Cardiac contraction and relaxation are mediated by a precise and coordinated linkage of electrical activation (excitation) and intracellular Ca2+ homeostasis, resulting in a so-called excitation-contraction coupling. Each ventricular AP starts with the influx of sodium (Na+) through voltage-gated Na+ channels that depolarizes the cell. When a certain threshold is reached voltage-sensitive L-type Ca2+ channels (LTCC) are activated that allow the influx of Ca2+ into the cytosol, triggering the much larger Ca2+ release from the sarcoplasmic reticulum (SR), the main intracellular Ca2+ storage organelle [16,17]. SR Ca2+ release is mediated by RyR2, which is a massive structure comprising the largest known ion channel-bearing macromolecular complex [18,19]. This process is called Ca2+-induced Ca2+-release (CICR) and is the fundamental link between electrical and mechanical activation in the heart [20]. The cytosolic Ca2+ then binds to troponin C and initiates myocardial contraction, forming the cardiac systole [16].
During diastole, myofilaments relax, either due to calcium re-uptake into the SR by the SR Ca2+-ATPase type-2a (SERCA2a) that pumps Ca2+ back into the SR stores, or Ca2+ is extruded and released into the extracellular space through Na+/Ca2+ exchanger (NCX) that exchanges approximately three Na+ ions entering the cell in exchange for one Ca2+ ion leaving [21]. Thus, the NCX removes Ca2+ by generating a net inward depolarizing current, called the transient inward current (Iti). Approximately 63% of cytosolic Ca2+ are taken up by SERCA2a and around 37% Ca2+ are extruded by NCX in humans [22,23].
The depolarization and release of Ca2+ into the cytosol and the rapid re-uptake or extrusion results in a Ca2+ wave, which is known as the Ca2+ transient. The amount of Ca2+ released from the SR directly correlates with the Ca2+ transient amplitude and is responsible for the strength of systolic contraction [16]. Consequently, any perturbations in intracellular Ca2+ handling could result in changes of electrical stability and cardiac contractility, which may lead to malignant ventricular arrhythmias.

1.1. The Role of β-Adrenergic Receptor Stimulation

Additionally, the activation of the adrenergic nervous system through β-adrenergic receptor (β-AR) stimulation has a great impact on Ca2+ handling. β-AR activates a crucial cascade of events, resulting in the phosphorylation of Ca2+-mediating proteins. In principal, two effects follow β-AR signaling: (1) an increased activity of LTCC and (2) an enhanced Ca2+ concentration in the SR due to stimulation of SERCA2a [24]. Both channels are activated through phosphorylation activities of the two enzymes protein kinase A (PKA) [25] or Ca2+/calmodulin-dependent protein kinase type II (CaMKII) [26]. Hence, the emerging phosphorylation of LTCC increases the amplitude of the current. Additionally, phosphorylation of phospholamban (PLN) hampers its inhibiting properties on SERCA2a, resulting in an increased SR Ca2+ re-uptake. On the other hand, the activity of PKA and CaMKII phosphorylates RyR2 and causes an activation of the Ca2+ releasing channel. In physiological conditions, this phosphorylation property, due to β-AR stimulation, is an important mechanism to enable SR Ca2+ release. This leads to an improvement of myocardial contractility in response to environmental stressors [27,28]. The influence of β-AR stimulation in pathophysiological conditions is described in the following section.

1.2. Cardiac Mitochondrial Ca2+ (mCa2+) Handling

For correct AP progression and myocardial contraction, several adenosine triphosphate (ATP)-consuming proteins, or so-called ion-pumps, such as SERCA2a (as mentioned above), are necessary. Therefore, a continuous supply of energy, in form of ATP, is required and provided by oxidative phosphorylation in myocardial mitochondria [29]. Cardiac mitochondria occupy about 20–30% of the cardiomyocyte’s volume and are located near the SR [30]. In humans, it is estimated that approximately 6 kg (kg) of ATP are hydrolyzed in the heart per day [31]. Mitochondrial Ca2+ uptake from the cytosol directly correlates with the amount of energy being produced [32]. As the inner mitochondrial membrane is rather impermeable for almost all ions, Ca2+ uptake is thought to be mostly mediated by the mitochondrial Ca2+ uniporter (MCU) [13,33,34]. The channel is known to be blocked by nanomolar concentrations of ruthenium red (RuR) or by its more specific derivate, ruthenium 360 (Ru360). In addition, it has been suggested that adenine nucleotides tend to suppress MCU activity, with ATP being the most effective one. Consequently, if a huge amount of ATP is available, Ca2+ uptake and ATP synthesis in mitochondria is inhibited [33,35,36].
The MCU is a highly Ca2+-selective protein complex that consists of the pore-forming mitochondrial Ca2+ uniporter protein [36,37], the essential MCU regulator (EMRE), and the mitochondrial calcium uptake 1 and 2 (MICU1/2) [38,39,40,41,42]. Furthermore, besides the pore forming units, other mitochondrial proteins seem to modulate the MCU activity [13,43,44,45]. Indeed, the mitochondrial uncoupling proteins (UCPs) 2 and 3, which are located in the cardiac inner mitochondrial membrane, are suggested to influence MCU function in an ATP dependent manner [45,46,47,48,49,50]. This interaction results in UCP2-dependent modulatory effects on cardiac excitation contraction coupling via altered LTCC activity. LTCC amplitude was shown to be decreased in UCP knock out cardiomyocytes vs. control. The authors speculate this was possibly due to Ca2+-dependent inactivation of LTCCs during increased dyadic cleft Ca2+ in UCP2 knock out cardiomyocytes [45,51].
The mitochondrial NCX (mNCX or NCLX) constitutes the major mitochondrial Ca2+ efflux pathway in the heart [52]. Similar to the sarcolemmal form it transports stoichiometry 3 Na+:1 Ca2+ [53]. This channel is essential to prevent mitochondrial Ca2+ overload with consequent effects on ROS generation [54]. As it is the case for sarcolemmal NCX, also a reverse mode for NCLX has been described. These effects may play an important role in diseased hearts with known alterations in Na+ and Ca2+ concentrations, such as HF [34,55,56].
With respect to cytosolic Ca2+ handling, there is still considerable debate whether mitochondrial Ca2+-uptake might affect cytosolic Ca2+-homeostasis [34,45].

2. Pathophysiology of Calcium Mediated Ventricular Arrhythmia

Malignant ventricular tachyarrhythmias are caused by either (1) enhanced automaticity, (2) triggered activity or (3) reentry. The first two conditions refer to cellular phenomena, whereas the latter is a matter of the cardiac electrophysiological network [57,58,59].
Firstly, “enhanced automaticity” is the acceleration of the spontaneous firing rate of cardiomyocytes (CMs). If ventricular CMs, normally quiescent, automatically generate APs due to disease or genetic modifications, irregular activation patterns arise. These abnormal wave fronts then collide against or compete with normal waves originating from the sinoatrial node. Secondly, “triggered activity” refers to Ca2+-mediated premature APs that result from early or delayed afterdepolarization (EADs, or DADs). These premature cellular depolarizations may provoke ventricular contraction or short runs of ventricular tachycardia (VT). The third and most common manifestation is “reentry” where one or several wave fronts circulate around zones of refractory tissue or rotate as spiral waves [58,59,60,61,62].
Pathophysiological conditions resulting in Ca2+-mediated ventricular arrhythmia and VT may either be caused by heritable mutations in cardiac ion channels and channel-interacting proteins or develop through acquired diseases of the myocardium (Figure 1).

2.1. Spontaneous Ca2+ Release Events (SCRs)

In SCRs, Ca2+ release is not triggered by an AP, but evoked by channelopathies (pathophysiological remodeling and mutations in ion channels) or due to acquired conditions, such as chronic HF, injury from myocardial ischemia or ischemia reperfusion (I/R) [63]. The probability of SCR dependent Ca2+ waves is related to the balance between the SR Ca2+ concentration and the minimum level of Ca2+, which is able to induce Ca2+ release from the SR, the so-called calcium threshold [64]. RyR2 and its inhibitory protein calsequestrin 2 (CASQ2) play the crucial roles in maintaining this threshold.
Notably, in inherited genetic mutations, such as catecholaminergic polymorphic VT (CPVT), the fundamental step of arrhythmogenesis is SCR due to mutations in either RyR2 (CPVT-1) or CASQ2 (CPVT-2) [65,66]. Several studies have shown that both, mutations in RyR2 [67,68] or CASQ2 [69], cause a significant increase in RyR2 activity, facilitating the occurrence of SCRs.
However, β-AR activity seems to be crucial for the generation of SCRs. As mentioned above, β-AR is responsible for the activation of LTCCs and constitutes a trigger for SR Ca2+ re-uptake due to SERCA2a phosphorylation. However, in pathological conditions, the RyR2 channel can be chronically hyperphosphorylated. Consequently, channel opening probability increases, leading to diastolic Ca2+ release, the so-called Ca2+ leak, from the SR [70,71]. However, even if activation of RyR2 increases, arrhythmias are only observed during β-AR stimulation [72]. On the other hand, some studies have elucidated PKA- and CaMKII-related phosphorylation of RyR2 to be the essential arrhythmogenic step [73]. Of note, CaMKII increases LTCC activity, thus removes the inhibitory effect of PLN on SERCA2a, and activates RyR2 [74]. LTCC activity is also affected by cardiac remodeling processes. In their elegant trials, Brooksby et al. and Cerbai et al., revealed that Ca2+ channel expression is highly dependent on the stage of HF [75,76], as the density of LTCCs is elevated in mild to moderate hypertrophy and decreases in more severe stages of hypertrophy and HF [77,78]. Of note, CMs from failing hearts exhibit a slight alteration in responsiveness to β-AR stimulation, resulting in slowing of LTCC inactivation and AP prolongation [79,80,81].
Furthermore, mutations and genetic defects in the genes CACNA1C, CACNB2, and CACNA2D1 (encoding components of LTCC) may result in various arrhythmogenic syndromes like the Brugada syndrome [82,83], Brugada syndrome with short QT duration [83,84], short QT syndrome [83,84,85], early repolarization syndrome [82,86] and idiopathic ventricular fibrillation (VF). All these mutations result in a loss-of-function of LTCCs [87].

2.2. Action Potential Prolongation and Afterdepolarizations

In the diseased heart, due to pathophysiological alterations like HF or genetic mutations, cellular remodeling is often accompanied by changes in Ca2+ channel expression and function. When abnormal Ca2+ release occurs, intracellular Ca2+ levels increase, potentially leading to cytosolic Ca2+ overload. To maintain Ca2+ homeostasis counterbalancing mechanisms, such as the NCX channel, need to be enhanced [88]. This current in turn produces transient membrane depolarization at the plateau phase or after the completion of the AP, leading to EADs or DADs, respectively [89]. When a certain threshold is reached at the plateau phase of the AP, due to EAD amplitude, LTCC channels are again activated, leading to AP prolongation. The longer the AP duration, the higher the possibility for EADs to occur due to the recovery of LTCCs from inactivation [61]. The generated inward current leads to a secondary membrane depolarization that interrupts the repolarization phase of the previous beat. This can further prolong AP and may be the origin for more EADs. In case of DADs, the net inward current generated by SCRs or NCX activity may trigger Na+ channel opening, resulting in a premature beat. Furthermore, if a certain threshold is reached, the electrical wave could spread to downstream cells and trigger premature contraction [90]. Of note, EADs, which occur during the so-called vulnerable window of repolarization, have a great potential to initiate sustained arrhythmias by reentrant mechanisms [91]. However, these events are mostly triggered by the presence of a pathologic electrophysiological substrate, which enables the formation of a unidirectional conduction block. Most commonly fibrotic formations and action potential duration (APD) gradients, which are present in the failing heart, promote such reentrant mechanisms [90].

2.3. The Role of Mitochondria in Ca2+ Handling of Diseased Heart

As mentioned above, dysregulation of Ca2+ handling, leading to a reduced cytosolic Ca2+ transient during excitation on the one hand but increased baseline cytosolic Ca2+ level on the other hand is a hallmark of HF. Since mitochondria are located in close proximity of the SR, they are speculated to act as a cytosolic Ca2+ sink in cardiac pathologies. The MCU has a low affinity but a high capacity to transport Ca2+. Therefore, under physiological conditions a rapid uptake of cytosolic Ca2+ into the mitochondria at the plateau phase, and not under basal conditions of the AP, is accomplished. Consequently, with increased cytosolic Ca2+ level at baseline, MCU related Ca2+ uptake could play a crucial role in balancing Ca2+ homeostasis in the failing heart [92,93]. However, this mechanism might promote mitochondrial Ca2+ overload and contribute to mitochondrial dysfunction [94,95].
Furthermore, since mitochondrial Ca2+ handling regulates mitochondrial energy generation, further Ca2+ dependent alterations might be suggested in mitochondria from the diseased heart. Besides ATP, also reactive oxygen species (ROS) emerge through the reaction of electrons with oxygen, forming superoxide anions (O2) [96]. In healthy cardiac tissue, redox balance is maintained by efficient antioxidant mechanisms that prevent excessive ROS accumulation. However, in diseased myocardium, ROS may be enhanced due to a defective scavenging system or an enhanced production, leading to oxidative stress (OS). Several studies [97,98,99,100,101] have revealed the mechanism of ROS-induced ROS-release (RIRR). They observed that local ROS injury may rapidly gather and get above a critical threshold to cause myocardial OS. During this process cardiac mitochondria react to an elevated ROS level by the production of more ROS, the so-called RIRR [98]. Enhanced mitochondrial-derived RIRR bursts influence RyR2 and SERCA2a activity, resulting in increased cytosolic Ca2+. Either regulated by the inner membrane anion channel (IMAC) or the Ca2+-dependent PTP, RIRR contributes to mitochondrial dysfunction with consequent generation of an arrhythmogenic substrate [102,103]. For extended information, Gambardella and colleagues summarize the role of mitochondria in the generation of cardiac arrhythmias [104].
In summary, mitochondrial dysfunction, including mitochondria-initiated cell death and altered mitochondrial Ca2+ handling, seem to be a hallmark of the failing heart [105]. Nevertheless, since the underlying mechanisms are not fully understood yet, the ability of mitochondria to compensate for extensive Ca2+ levels remains a matter of debate [94,106].

3. Novel Therapeutic Options for Calcium Mediated Ventricular Arrhythmias

As summarized in the previous sections and previous reviews [107,108,109,110], cellular Ca2+ imbalance is one of the main triggers for the generation of malignant ventricular arrhythmias [111]. Therefore, various approaches have focused on restoring physiological Ca2+handling in the diseased heart.
Approximately 50 years ago, Fleckenstein [112] was the first to review Ca2+ channel blockers as new drugs for the treatment of coronary diseases. Indeed, drugs affecting intracellular Ca2+handling have the ability to decrease automaticity of Ca2+ related AP generation and have emerging uses as anti-arrhythmic agents. Classical drugs that target proteins responsible for cellular Ca2+-handling include β-blockers (inhibitor of β-AR activation) [113], and the LTCC inhibitors verapamil [114] and diltiazem [115,116], as well as the class I anti-arrhythmic flecainide. Although this agent is an anti-arrhythmic drug with Na+ channel blocking properties, in mouse models and patients with CPVT it has been shown that it also inhibits RyR2 [117,118].
Unfortunately, current anti-arrhythmic drug therapy often increases or ideally has a neutral effect on cardiac-related mortality [5,119,120,121,122,123]. While device therapies, such as implantable cardioverter defibrillators (ICDs), have a positive effect on correct heart rhythm, they do not treat or cure the cause of arrhythmia. Of note, they change the electrical signaling to convert tachyarrhythmias following their onset. Besides its preventative effects, ICD implantation may lead to psychopathological disorders and may therefore reduce patient’s quality of life. Furthermore, several surgical complications are associated with ICD implantations, as well as device and lead failures [124].
Due to the reasons mentioned above and in order to achieve a potential curative effect, a high demand for novel therapies to treat the origin of ventricular arrhythmia is present. In this review, we want to summarize novel therapeutic strategies and targets to combat Ca2+ related VT, such as gene therapy. Of note, by using various viral vectors (Table 1) this experimental approach is able to target specifically arrhythmia related proteins. Therefore, cardiac gene therapy is one of the emerging therapeutic approaches that can offer a new therapeutic strategy. Indeed, various experimental studies have already investigated the potential effects of gene therapy aimed to target complexes involved in Ca2+ related ventricular arrhythmias.

3.1. Gene Therapy Targeting LTCC

Due to their correlative effect on contractility or blood pressure, in patients suffering from ventricular arrhythmias and HF, strategies of blocking LTCC are not optimal yet [142]. For this reason, gene therapy to shorten APD without impairing influences on systemic physiological behavior is intended. Table 2 lists up gene therapy strategies targeting LTCC in vivo and in vitro.
To the best of our knowledge, Muarata et al. [143] was the first to create a biological effective genetic Ca2+ channel blocker by the overexpression of the ras-related small G-protein Gem. This was achieved by adenovirus (AD) mediated gene transfer of Gem, resulting in a significant decreased LTCC current density in ventricular myocytes. Consequently, this approach successfully promoted shortening of APD with consequent abbreviation of electrocardiographic QTc interval [143]. In 2007, Cingolani et al. [144] seized on this topic and created a genetically knockdown of LTCC accessory β-subunit gene by a short hairpin RNA template sequence. In vivo injection of the lentiviral vector with the RNA template partially inhibited LTCC current and reduced Ca2+ transient amplitude in neonatal rats. Furthermore, the hypertrophic response in vivo and in vitro was attenuated, without affecting systolic performance [144].
As expression dependent manipulation of large cardiac genes (>6 kb) is difficult, due to the limited packaging capacity of viral vectors (4–7 kb), improvements of the delivery approaches were anticipated. In their elegant proof of concept study, Subramanyam et al. [145] developed a new technique targeting the 6.6 kb pore-forming α1C-subunit of LTCC. Concerning this, the so-called split-intein protein transsplicing was used. Split-intein-tagged α1C fragments encoding dihydropyridine-resistant channels were incorporated into AD and applied to adult rat cardiomyocytes in vitro. Of note, triggered recombinant LTCC channel expression promoted Ca2+ transients and supported β-adrenergic regulation of excitation-contraction coupling [145].

3.2. SERCA2a Gene Therapy

In diseased heart, Ca2+ reuptake by SERCA2a is commonly decreased and causes major defects in excitation-contraction coupling [146]. Since SERCA2a is the major pump that transfers cytosolic Ca2+ back to the sarcoplasmic reticulum during diastole, promoting SERCA2a expression seems to be a promising goal in HF. Indeed, multiple preclinical studies using SERCA2a adeno-associated virus (AAV) mediated gene transfer [45,147,148,149] encouraged broader application of this approach in the HF population. However, while various preclinical trials [42,48,49,50,51] supported beneficial outcomes mostly by restoring mechanical function, the CUPID2 trial, most probably due to technical problems with gene delivery, failed to elucidate therapeutic benefits in the clinical setting [59]. Nevertheless, pathophysiology of intracellular Ca2+ handling is complex. Consequently, besides affecting the heart’s mechanical function, SERCA2a expression might also influence cardiac electrophysiology. Theoretically, the restoration of SERCA2a activity and prevention of cytotoxic Ca2+ overload could affect the generation of beat-to-beat repolarization alternans and consequent DADs.
Nevertheless, it remains unclear whether an increase in SERCA2a expression by gene therapy could alter the electrophysiological substrate that triggers the generation of cardiac arrhythmias. Furthermore, SERCA2a overexpression dependent Ca2+ overload of the SR could promote a Ca2+ leak through RYR2 and contribute to an increased susceptibility to arrhythmias.
However, in contrast to these speculations, in rodent models, SERCA2a overexpression did not exacerbate SR Ca2+ leak through RYR2 and has suggested protection against, not promotion of, arrhythmias (Table 3) [150,151,152,153,154,155]. Indeed, Lyon et al. explored an established rat chronic HF model, which is characterized by a high burden of spontaneous ventricular arrhythmia [153]. In treated rats, SR Ca2+ load was stabilized, and arrhythmia was significantly reduced. Furthermore, Ca2+ leak decreased to a level not statistically different from control animals. Of note, RYR2 phosphorylation at Ser2815, which is accomplished by CaMKII, was lowered, resulting in an increase in the threshold for SR Ca2+ release with consequent lower rates of SCR events and cellular triggered activity in vitro and in vivo [153]. These findings are supported by previous data. Xie et al. reported a suppression of Ca2+ alternans in cultured rabbit CMs [156]. Other studies in healthy [152] but also in a pressure-overload induced HF model in the guinea pig [154] demonstrated prevention of Ca2+ and APD alternans with consequent suppression of pacing-induced ventricular arrhythmias. Furthermore, in small as well as large animal models the incidence of I/R arrhythmias was suppressed [150,151].
More recently, a novel mechanism how SERCA2a gene therapy might affect cardiac electrophysiology was explored [160]. In a large animal model, for the first time, the authors demonstrated that SERCA2a gene transfer could prevent cellular remodeling in an advanced stage of ischemic HF. The animals were treated by gene therapy after HF was confirmed, one month after myocardial infarction. Even in this pronounced stage, SERCA2a gene therapy prevented the incidence of dobutamine dependent ventricular arrhythmias in vivo. Importantly, while gene therapy was not related to improvements of hemodynamic function, SERCA2a related prevention of major electrophysiological remodeling processes was observed. SERCA2a expression increased conduction velocity reserve, likely by preventing CAMKII overactivation with consequent increase in the reserve of cardiac excitability. Of note, in HF electrophysiological remodeling promotes reduction of the conduction velocity reserve leading to an increased susceptibility for reentry mechanisms at higher pacing rates [162]. Consequently, in this study, HF dependent prolongation of QRS duration on ECG in vivo as well as the rate of pacing induced sustained VT and ventricular fibrillation (ex vivo) were decreased. Therefore, this study indicated primary effects of SERCA2a gene therapy on myocardial excitability, independently of altered mechanical function [160].
Strauss and colleagues [161] investigated potential effects of SERCA2a gene delivery on the arrhythmogenic substrate in pulmonary arterial hypertension (PAH). Of note, this potential, fatal pathology promotes right heart failure with consequent malignant electrophysiological remodeling. In their elegant study, male rats developed advanced PAH after subcutaneous injection of monocrotaline. This approach was followed by aerosolized delivery of AAV1/SERCA2a after three weeks. Indeed, delivery of SERCA2a ameliorated myocardial electrophysiological remodeling including fibrosis and altered ion channel expression. These findings coincided with increased rate of VT at rapid heart rates as well as major electrophysiological alterations in the non-treated PAH animals (including prolonged APD, increased APD heterogeneity, a reversal in the trans-epicardial APD gradient and marked conduction slowing), which was not described in the treated group. Thus, for the first time Strauss et al. depicted a non-cardiac gene delivery approach with successful suppression of ventricular arrhythmias [161].

3.3. Gene Therapy Targeting RyR2 Complex

CPVT is the most prominent form of genetic mutations concerning changes particularly in RyR2 and CASQ2. Mutations in RyR2 are mostly autosomal-dominant and lead to unzipping of CaM from RyR2 protein, thus repealing the inhibitory effect of CaM on RyR2. These results in SCRs, promoting arrhythmia and HF [19,163,164]. Indeed, a recent study demonstrated ryanodine receptor Ca2+ leak to act as a useful biomarker in patients with HF [165].
The RyR2 channel itself is targeted rarely by experimental gene therapy (Table 3), as several compounds have been developed to reduce Ca2+ leak through RyR2, such as JTV519 [166]. Nevertheless, two studies aimed RyR2 channel modification by gene therapy (Table 4), achieved by the new genome editing technique CRISPR/CAS9 (clustered regularly interspaced short palindromic repeats with caspase 9). As in CPVT RyR2 channels containing one mutant monomer can promote a Ca2+ leak, it is hypothesized that modification of even one allele, such as silencing or removing of a fraction of the mutant subunits, could restore cardiac physiology [167,168].
Consequently, Bongianino et al. investigated allele-specific gene silencing in the autosomal-dominant form of CPVT [167]. They used a short-interfering ribonucleic acid (siRNA) to specifically silence an allele of RyR2 that bears a dominant negative (R4496C) mutation in the CPVT mouse model (Table 4). Indeed, by using this elegant technique, the authors were able to reduce isoproterenol-induced DADs and triggered activity. Importantly, this was followed by a decreased susceptibility for adrenergically mediated VT. These findings were consistent with reverted ultrastructural changes of SR and transverse tubules, as well as attenuated mitochondrial abnormalities [167]. Similar findings were confirmed by others. CRISPR/Cas9 genomic allele silencing of R176Q was achieved by single subcutaneous injection at postnatal day 10. This highly specific editing normalized the incidence of Ca2+ sparks to normal levels and abolished the genesis of ventricular arrhythmias in the treated animals [168].
Successful results were also achieved by targeting regulatory proteins like CaMKII. Of note, CaMKII dependent phosphorylation is crucial for RyR2-mediated SR Ca2+ release. Consequently, pharmacological inhibition reduces Ca2+ handling abnormalities and arrhythmias in human iPSC-CMs [174]. Importantly, CaMKII phosphorylation of RyR2 at serine 2814 is evident to prevent the latent arrhythmic potential of RyR2 mutations in CPVT-1 indicating antiarrhythmic effects [73,175]. Therefore, efforts were started to explore the genetic inhibition of CaMKII in a CPVT-1 mouse model (RYR2R176Q/+) [173]. CaMKII inhibition was achieved by AAV9 gated CaMKII inhibitory peptide autocamitide-2-related inhibitory peptide (AIP), which was fused to a green fluorescent protein (GFP). The outcomes showed a robust expression of the peptide solely in the heart. This successful approach was able to prevent VTs either after adrenergic stimulation or programmed ventricular pacing [173]. These results were supported by concomitant in vitro studies in human induced pluripotent stem cells cardiomyocytes (iPSC-CMs), derived from two patients with different RyR2 mutations. In this cellular model CPVT-induced abnormal Ca2+ release events were efficiently suppressed by AIP [173].
Further studies have addressed the recessive form of CPVT, namely CPVT-2. This form is characterized by mutations in CASQ2, which is an essential regulating part of the RyR2 macromolecular complex. RyR2 channels lacking the CASQ2 protein tend to open spontaneously, without the need for LTCC dependent trigger Ca2+ influx [176]. Consequently, in animal models with mutant CASQ2 protein a higher frequency of Ca2+ dependent ventricular arrhythmias is observed [177]. In CPVT several missense, deletion or nonsense mutations, which result in a severe reduction, or complete loss of the protein have been detected [178]. Therefore, various studies have focused on restoring CASQ2 expression using diverse gene transfer techniques.
In iPSCs from a patient carrying the homozygous CASQ2-G112+5X mutation, AAV9 dependent gene delivery was able to normalize cellular Ca2+ transient and reduce the incidence of DADs [170]. These optimistic results are supported by further animal studies. Denegri and colleagues have investigated a murine CPVT knock-model of CASQ2 (CASQ2R33Q/R33Q (R33Q) mutation) [169]. Successful transfection of the wild-type CASQ2 gene was achieved by an AAV9 vector (AAV9-CASQ2), which was delivered in neonatal mice on day 3. In vivo and in vitro investigations of this model have revealed astonishing curative effects of this therapeutic approach. Restoration of physiological expression and interaction of CASQ2, junctin and triadin was followed by normalization of electrophysiological and ultrastructural abnormalities of cellular Ca2+-handling. Importantly, due to successful transfection life-threatening arrhythmias have been abrogated [169]. Further findings in a murine CASQ2D307H model supported these enthusiastic results. Interestingly, in the study by Kurtzwald-Josefson et al. [171] antiarrhythmic efficacy was dependent on the CASQ2 level expression (at least greater than 33% of normal CASQ2).
Further studies have successfully focused on other regulatory proteins of the RyR2 complex, namely CaM. CaM gene transfer into the CPVT knock-in mouse model carrying a CASQ2R33Q/R33Q (R33Q) mutation was also able to restore defective Ca2+ handling and prevent consequent isoproterenol-triggered VT [172].

3.4. Cardiac NCX Constituting a Target for Gene Therapy?

Although transgenic mouse models overexpressing NCX have been studied for a long time [179,180,181], NCX currently fails to be a direct and single target for efficient gene therapy approaches. This may be due to the great dependence of NCX function on several other channels, such as SERCA2a and LTCC. Overexpression of cardiac NCX results in cardiac hypertrophy and increases the risk for HF presumably by alterations in excitation-contraction coupling [182]. On the other hand, homozygous NCX-deficiency results in premature death of embryonic mice by days 9 and 10 due to defect pacemaker function [183]. Studies addressing targeted NCX gene transfer are summarized in Table 5.
As in most cardiac pathologies NCX activity seems to be reduced [189,190,191], restoration of the exchanger function might be an attractive therapy approach. However, therapeutic increase in NCX activity revealed adverse pathologically altered phenotypes. Due to potential dependence on Na+/K+-ATPase (NKA) activity, NCX overexpression resulted in systolic and diastolic dysfunction. These potential alterations observed in transgenic mouse model as well as in transfected CMs are probably related to decreased SR Ca2+ stores with concomitant cellular Ca2+ overload [184,187].
On the other hand, in some pathological manifestations decreased NCX activity is accompanied by reduction in SERCA2a function. Therefore, the usage of a dual therapy including simultaneous elevation of SERCA2a and NCX function seems favorable. Terracciano and colleagues [159] have investigated this approach by overexpressing cardiac NCX in a transgenic mouse model with decreased SERCA2a activity. They were able to conclude that overexpression of NCX may compensate SERCA2a inhibition and restore normal Ca2+ homeostasis [159].
In vitro studies downregulating NCX in rat cardiomyocytes were performed by Tadros et al. [186]. They showed a reduced cellular influx and efflux of Ca2+ that is contingent on reduced NCX function.

3.5. Mitochondrial Ca2+ Channels—Possible Targets for Gene Therapy

A new proof-of-principle study has been published by Gammage et al. [192]. This elegant trial addressed the question if mutations in the mitochondrial genome (mtDNA) could be corrected by mitochondrially targeted zinc finger-nucleases (mtZFNs). The authors successfully targeted specific mtDNA, which resulted in the restoration of molecular and physiological disease phenotypes in the heart tissue [192]. Consequently, mitochondria are becoming an attractive target for novel genetic therapy approaches.
As already described, mitochondria have the ability to transport huge amounts of cellular Ca2+. This vital process has major regulatory effects on mitochondrial function (generation of ATP, genesis of ROS, etc.) but also on the whole cellular metabolism including cellular Ca2+ handling [29,30,31,32]. Therefore, targeting proteins involved in mitochondrial Ca2+ handling seems to be a promising strategy to suppress cardiac arrhythmias. This concept is supported by several studies, in which MCU activity was suppressed via pharmacological or genetical approaches, resulting in a diminished myocardial damage after I/R [29,193]. Table 6 presents an overview of recent trials addressing genetic modifications in MCU and further mitochondrial proteins involved in mitochondrial Ca2+ handling.
Wu et al. [194] were the first to create a transgenic MCU knock-out mouse. The group used an AD vector to transduce dominant negative (DN) MCU into mouse embryonic stem cells, which were further inserted into pseudo-pregnant females. DN-MCU mice showed normal resting heart rates. However, they were incapable of accelerating fight or flight heart rate acceleration. Thus, for the first time this elegant technique revealed chronotropic function of the MCU indicating this channel as a promising target for cardiac arrhythmias [194]. Others applied siRNA to target MCU in CMs. This approach prevented mitochondrial Ca2+ overload with consequent permeability pore opening, leading to reduced hypoxia-reoxygenation injury indicating a potential strategy for arrhythmia prevention [195].
Since mitochondrial Ca2+ uptake is decreased in diabetic cardiomyocytes [197,198], genetic therapies aimed to increase MCU expression in diabetic mice. For this purpose, AAV9-MCU was generated and injected into the jugular vein of diseased animals. Indeed, after 4–6 weeks, normalized MCU levels as well as a restoration of mitochondrial Ca2+ handling were detected. These findings support the idea that abnormal Ca2+ handling in CMs could be restored via MCU transgene expression [196].
Further studies explored the function of UCPs. As already mentioned, the isoforms UCP2 and UCP3 are suggested to influence MCU activity [45,46,47,48,49,50]. Consequently, in UCP2 knockout mice altered excitation contraction coupling with compensatory reduced LTCC-activity is observed [45,51]. Indeed, UCP2-dependent modulations promoted APD shortening, increased slope factor of action potential upstrokes and more hyperpolarized resting membrane potential leading to altered ECG parameters (PR and QRS as well as shortening of the QTc interval). Importantly, modifications of electrophysiology were followed by increased incidence of DAD and a higher susceptibility to Ca2+ mediated ventricular arrhythmias [51]. Therefore, this study identified mitochondrial proteins involved in Ca2+ handling as attractive targets to influence Ca2+ related ventricular arrhythmias.
Since HF is accompanied by alterations in Na+ and Ca2+ concentrations, the mitochondrial NCX (NCLX) may also constitute a promising target for gene therapy [34,55,56]. Indeed, chronic administration of the NCLX inhibitor CGP leads to cellular remodeling, fibrosis and alters the susceptibility for ventricular arrhythmias and SCD [199]. Consequently, preservation of physiological NCX expression and function seems evident. The elegant work of Luongo and colleagues [193] addressed this issue by deleting the gene encoding NCLX (Slc8b1). Due to mitochondrial Ca2+ overload, they observed the development of HF with consequent SCD. On the contrary, overexpression of NCLX resulted in increased mitochondrial Ca2+ clearance with reduced permeability transition formation and decreased tissue necrosis [193]. Therefore, enhancing NCLX function in diseased hearts might be a promising therapeutic strategy in the future.

4. New Targets for Ca2+ Mediated Ventricular Arrhythmias

Besides the genetic targets mentioned above, further regulatory proteins that mediate cardiac Ca2+ homeostasis could constitute a potential point of application for gene therapy. These potential therapeutic targets are summarized in the following paragraph.
The small ubiquitin-related modifier 1 (SUMO-1) was uncovered as a crucial regulatory part of the SERCA2a complex in HF. Under physiological conditions, SUMO-1 binds to SERCA2a in a process called SUMOylation affecting the ATP-binding domain of SERCA2a. SUMOylation is followed by an increased ATP-binding affinity of SERCA2a with concomitant stabilization of the protein. However, in HF, SUMO-1 expression is decreased. Therefore, multiple studies have aimed for the expression of SUMO using genetic strategies in various HF animal models [200,201,202]. Indeed, the authors were able to restore cardiac hemodynamic function, indicating overexpression of SUMO-1 as a promising strategy in HF [200,201,202]. Nevertheless, further trials need to investigate the impact of SUMO-1 on the incidence of cardiac arrhythmias.
The Ca2+ binding protein S100 calcium-binding protein A1 (S100A1) is a further important regulator of cardiac performance. It directly interacts with SERCA2a and RyR2. Thus, it contributes to improved Ca2+ handling and contractile performance. Consequently, it plays a crucial role in the pathophysiology of HF and might provide a novel therapeutic target for treating acute and chronic cardiac dysfunction [203]. As suspected, in various HF model genetic enhancing of S100A1 expression was able to improve contractile function. Importantly, this was achieved by normalization of physiological Ca2+ transients and SR load due to an increase in SR Ca2+ uptake and reduced SR Ca2+ leak [204,205,206,207] Therefore, this strategy might be suspected to provide a potential tool to prevent Ca2+ related ventricular arrhythmias.
The regulation of Ca2+ influx into the cytoplasm seems not only to be limited to activation of the LTCC. Interestingly, when Ca2+ is depleted from intracellular stores, such as ER, store-operated Ca2+ entry channels (SOC) open. The most important SOC channel is the Ca2+ release-activated Ca2+ (CRAC) channel, which consist of multimers of ORAI (Ca2+ release-activated Ca2+ channel protein) family proteins, there of the best characterized is ORAI1. ORAI1/CRAC channels are regulated by the stromal interaction molecule 1 (STIM1). Of note, STIM1 is an ER transmembrane protein, which is activated through ER Ca2+ store depletion [208,209,210,211,212,213]. Through its effects on cellular Ca2+ cycling this protein is involved in various cellular pathologies including cancerogenesis [214,215,216,217,218,219,220,221,222,223]. However, investigations have also revealed regulatory function in several important cardiac physiological and pathophysiological processes including angiogenesis, pacemaker function and cardiac hypertrophy [218,219,223,224,225,226,227]. Therefore, this protein might constitute an interesting target for cardiac gene therapy including prevention of Ca2+ mediated ventricular arrhythmias [218,224,228,229,230].
Further potential gene therapy targets could consist of the Ca2+ regulating proteins PLN [231,232,233,234], junctophilin- (JPH2) [235,236,237] and the inhibitor of protein phosphatase 1 (I-1c) [238]. In experimental models all have been shown to improve cardiac Ca2+ handling and cardiac performance.

5. Conclusions

A defect in Ca2+ handling is one the major reasons for congenital but also acquired malignant ventricular arrhythmia syndromes. In recent years, major progresses were achieved to understand the underlying mechanisms of Ca2+ mediated arrhythmogenesis. Consequently, various cellular and mitochondrial targets were identified. The strategy of gene transfer is able specifically to aim disease relevant proteins. Indeed, targeting cardiac Ca2+ handling by somatic gene transfer has been demonstrated to be effective in experimental models. Therefore, various studies have explored this strategy to prevent Ca2+ related arrhythmogenesis. By targeting regulatory proteins of Ca2+ handling, these investigations were already able to uncover promising results by suppressing Ca2+ related congenital but also acquired arrhythmia mechanisms. Therefore, gene transfer to suppress Ca2+ related ventricular arrhythmias seems to be an attractive strategy in in the future. Nevertheless, first investigations in humans will be necessary to uncover the efficiency of this therapeutic approach.

Author Contributions

Conceptualization, V.P. and L.J.M.; methodology, not applicable; software, not applicable; validation, V.P. and L.J.M.; formal analysis, not applicable; investigation, V.P. and L.J.M.; resources, not applicable; data curation, not applicable; writing—original draft preparation, V.P., P.J., R.L., N.S.Z., M.C.B., M.L., U.C.H., L.J.M.; writing—review and editing, V.P., P.J., R.L. and L.J.M.; visualization, not applicable.; supervision, L.J.M.; project administration, L.J.M.; funding acquisition, not applicable.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AAVAdeno-associated virus
ADAdenovirus
AIPAutocamitide-2-related inhibitory peptide
APAction potential
APDAction potential duration
ATPAdenosine triphosphate
β-ARβ-adrenergic receptor
Ca2+Calcium
CGPCGP317157
CaMCalmodulin
CaMKIICalcium/calmodulin-dependent protein kinase type II
CASQCalsequestrin
CASTCardiac Arrhythmia Suppression Trial
CICRCalcium-induced calcium-release
CMCardiomyocyte
CPVTCatecholaminergic polymorphic ventricular tachycardia
CRACCa2+ release-activated Ca2+
CRISPR/Cas9Clustered regularly interspaced short palindromic repeats with caspase 9
DADDelayed afterdepolarizations
DNDominant negative
DNADeoxyribonucleic acid
DRDownregulation
EADEarly afterdepolarizations
EMREEssential mitochondrial calcium uniporter regulator
GFPGreen fluorescent protein
HFHeart failure
I-1cInhibitor of protein phosphatase 1
ICa-LL-type calcium current
ICDImplantable cardioverter defibrillator
IMACInner membrane anion channel
iPSCInduced pluripotent stem cell
I/RIschemia reperfusion
ItiTransient inward current
JPH2Junctophilin 2
kbKilobases
kgKilograms
LTCCL-type calcium channel
LVLeft ventricle
MCUMitochondrial calcium uniporter
MICU1/2Mitochondrial calcium uptake 1 and 2
mNCXMitochondrial sodium/calcium exchanger
Na+Sodium
NAbNeutralizing antibodies
NCXSodium/calcium exchanger
NCLXMitochondrial sodium/calcium exchanger
NKASodium/potassium-ATPase
O2Superoxide anions
OEOverexpression
ORAICalcium release-activated calcium channel protein
OSOxidative stress
PAHPulmonary arterial hypertension
PKAProtein kinase A
PLNPhospholamban
PTPPermeability transition pore
RIRRReactive oxygen species-induced reactive oxygen species-release
RNARibonucleic acid
ROSReactive oxygen species
Ru360Ruthenium 360
RuRRuthenium red
RyR2Ryanodine receptor type 2
S100A1S100 calcium-binding protein A1
SCDSudden cardiac death
SCRSpontaneous calcium release events
SERCA2aSarcoplasmic reticulum calcium-ATPase type-2a
siRNAShort-interfering ribonucleic-acid
SOCStore-operated Ca2+ entry channels
SRSarcoplasmic reticulum
STIM1Stromal interaction molecule 1
SUMO-1Small ubiquitin-like modifier type 1
TACTransverse aortic constriction
UCPUncoupling protein
VFVentricular fibrillation
VTVentricular tachycardia
wtWild-type

References

  1. Schocken, D.D.; Arrieta, M.I.; Leaverton, P.E.; Ross, E.A. Prevalence and mortality rate of congestive heart failure in the United States. J. Am. Coll. Cardiol. 1992, 20, 301–306. [Google Scholar] [CrossRef] [Green Version]
  2. Napolitano, C.; Bloise, R.; Monteforte, N.; Priori, S.G. Sudden cardiac death and genetic ion channelopathies: Long QT, Brugada, short QT, catecholaminergic polymorphic ventricular tachycardia, and idiopathic ventricular fibrillation. Circulation 2012, 125, 2027–2034. [Google Scholar] [CrossRef] [PubMed]
  3. Kaur, K.; Jalife, J. Is TGF-β1 (Transforming Growth Factor-β1) an Enabler of Myofibroblast-Cardiomyocyte Cross Talk? Circ. Arrhythm. Electrophysiol. 2017, 10, e005289. [Google Scholar] [CrossRef] [PubMed]
  4. Estes, N.A.M.; Weinstock, J.; Wang, P.J.; Homoud, M.K.; Link, M.S. Use of antiarrhythmics and implantable cardioverter-defibrillators in congestive heart failure. Am. J. Cardiol. 2003, 91, 45D–52D. [Google Scholar] [CrossRef]
  5. Echt, D.S.; Liebson, P.R.; Mitchell, L.B.; Peters, R.W.; Obias-Manno, D.; Barker, A.H.; Arensberg, D.; Baker, A.; Friedman, L.; Greene, H.L. Mortality and morbidity in patients receiving encainide, flecainide, or placebo. The Cardiac Arrhythmia Suppression Trial. N. Engl. J. Med. 1991, 324, 781–788. [Google Scholar] [CrossRef] [PubMed]
  6. Schram, G.; Pourrier, M.; Melnyk, P.; Nattel, S. Differential distribution of cardiac ion channel expression as a basis for regional specialization in electrical function. Circ. Res. 2002, 90, 939–950. [Google Scholar] [CrossRef]
  7. Akar, F.G.; Wu, R.C.; Juang, G.J.; Tian, Y.; Burysek, M.; Disilvestre, D.; Xiong, W.; Armoundas, A.A.; Tomaselli, G.F. Molecular mechanisms underlying K+ current downregulation in canine tachycardia-induced heart failure. Am. J. Physiol. Heart Circ. Physiol. 2005, 288, H2887–H2896. [Google Scholar] [CrossRef]
  8. Jin, H.; Chemaly, E.R.; Lee, A.; Kho, C.; Hadri, L.; Hajjar, R.J.; Akar, F.G. Mechanoelectrical remodeling and arrhythmias during progression of hypertrophy. FASEB J. 2010, 24, 451–463. [Google Scholar] [CrossRef]
  9. Yano, M.; Ikeda, Y.; Matsuzaki, M. Altered intracellular Ca2+ handling in heart failure. J. Clin. Investig. 2005, 115, 556–564. [Google Scholar] [CrossRef]
  10. Betzenhauser, M.J.; Pitt, G.S.; Antzelevitch, C. Calcium Channel Mutations in Cardiac Arrhythmia Syndromes. Curr. Mol. Pharmacol. 2015, 8, 133–142. [Google Scholar] [CrossRef] [Green Version]
  11. Györke, I.; Hester, N.; Jones, L.R.; Györke, S. The role of calsequestrin, triadin, and junctin in conferring cardiac ryanodine receptor responsiveness to luminal calcium. Biophys. J. 2004, 86, 2121–2128. [Google Scholar] [CrossRef]
  12. Kotta, M.-C.; Sala, L.; Ghidoni, A.; Badone, B.; Ronchi, C.; Parati, G.; Zaza, A.; Crotti, L. Calmodulinopathy: A Novel, Life-Threatening Clinical Entity Affecting the Young. Front. Cardiovasc. Med. 2018, 5, 175. [Google Scholar] [CrossRef] [PubMed]
  13. Michels, G.; Khan, I.F.; Endres-Becker, J.; Rottlaender, D.; Herzig, S.; Ruhparwar, A.; Wahlers, T.; Hoppe, U.C. Regulation of the human cardiac mitochondrial Ca2+ uptake by 2 different voltage-gated Ca2+ channels. Circulation 2009, 119, 2435–2443. [Google Scholar] [CrossRef] [PubMed]
  14. Gong, G.; Liu, X.; Wang, W. Regulation of metabolism in individual mitochondria during excitation-contraction coupling. J. Mol. Cell. Cardiol. 2014, 76, 235–246. [Google Scholar] [CrossRef] [PubMed]
  15. Zhou, B.; Tian, R. Mitochondrial dysfunction in pathophysiology of heart failure. J. Clin. Investig. 2018, 128, 3716–3726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Bers, D.M. Cardiac excitation-contraction coupling. Nature 2002, 415, 198–205. [Google Scholar] [CrossRef] [PubMed]
  17. Gambardella, J.; Trimarco, B.; Iaccarino, G.; Santulli, G. New Insights in Cardiac Calcium Handling and Excitation-Contraction Coupling. Adv. Exp. Med. Biol. 2018, 1067, 373–385. [Google Scholar]
  18. Santulli, G.; Lewis, D.; des Georges, A.; Marks, A.R.; Frank, J. Ryanodine Receptor Structure and Function in Health and Disease. Subcell. Biochem. 2018, 87, 329–352. [Google Scholar] [Green Version]
  19. Santulli, G.; Nakashima, R.; Yuan, Q.; Marks, A.R. Intracellular calcium release channels: An update. J. Physiol. 2017, 595, 3041–3051. [Google Scholar] [CrossRef]
  20. Fabiato, A. Calcium-induced release of calcium from the cardiac sarcoplasmic reticulum. Am. J. Physiol. 1983, 245, C1–C14. [Google Scholar] [CrossRef]
  21. Stern, M.D.; Cheng, H. Putting out the fire: What terminates calcium-induced calcium release in cardiac muscle? Cell Calcium 2004, 35, 591–601. [Google Scholar] [CrossRef] [PubMed]
  22. 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. [Google Scholar] [CrossRef] [PubMed]
  23. Terentyev, D.; Viatchenko-Karpinski, S.; Valdivia, H.H.; Escobar, A.L.; Györke, S. Luminal Ca2+ controls termination and refractory behavior of Ca2+-induced Ca2+ release in cardiac myocytes. Circ. Res. 2002, 91, 414–420. [Google Scholar] [CrossRef] [PubMed]
  24. Ginsburg, K.S.; Bers, D.M. Modulation of excitation-contraction coupling by isoproterenol in cardiomyocytes with controlled SR Ca2+ load and Ca2+ current trigger. J. Physiol. 2004, 556, 463–480. [Google Scholar] [CrossRef] [PubMed]
  25. Hussain, M.; Orchard, C.H. Sarcoplasmic reticulum Ca2+ content, L-type Ca2+ current and the Ca2+ transient in rat myocytes during beta-adrenergic stimulation. J. Physiol. 1997, 505 Pt 2, 385–402. [Google Scholar] [CrossRef]
  26. Grimm, M.; Brown, J.H. Beta-adrenergic receptor signaling in the heart: Role of CaMKII. J. Mol. Cell. Cardiol. 2010, 48, 322–330. [Google Scholar] [CrossRef]
  27. Reiken, S.; Gaburjakova, M.; Guatimosim, S.; Gomez, A.M.; D’Armiento, J.; Burkhoff, D.; Wang, J.; Vassort, G.; Lederer, W.J.; Marks, A.R. Protein kinase A phosphorylation of the cardiac calcium release channel (ryanodine receptor) in normal and failing hearts. Role of phosphatases and response to isoproterenol. J. Biol. Chem. 2003, 278, 444–453. [Google Scholar] [CrossRef]
  28. Wehrens, X.H.T.; Lehnart, S.E.; Reiken, S.R.; Marks, A.R. Ca2+/calmodulin-dependent protein kinase II phosphorylation regulates the cardiac ryanodine receptor. Circ. Res. 2004, 94, e61–e70. [Google Scholar] [CrossRef]
  29. Kwong, J.Q. The mitochondrial calcium uniporter in the heart: Energetics and beyond. J. Physiol. 2017, 595, 3743–3751. [Google Scholar] [CrossRef]
  30. Ramesh, V.; Sharma, V.K.; Sheu, S.S.; Franzini-Armstrong, C. Structural proximity of mitochondria to calcium release units in rat ventricular myocardium may suggest a role in Ca2+ sequestration. Ann. N. Y. Acad. Sci. 1998, 853, 341–344. [Google Scholar] [CrossRef]
  31. Weiss, R.G.; Gerstenblith, G.; Bottomley, P.A. ATP flux through creatine kinase in the normal, stressed, and failing human heart. Proc. Natl. Acad. Sci. USA 2005, 102, 808–813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Jouaville, L.S.; Pinton, P.; Bastianutto, C.; Rutter, G.A.; Rizzuto, R. Regulation of mitochondrial ATP synthesis by calcium: Evidence for a long-term metabolic priming. Proc. Natl. Acad. Sci. USA 1999, 96, 13807–13812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Kirichok, Y.; Krapivinsky, G.; Clapham, D.E. The mitochondrial calcium uniporter is a highly selective ion channel. Nature 2004, 427, 360–364. [Google Scholar] [CrossRef] [PubMed]
  34. 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. [Google Scholar] [CrossRef] [PubMed]
  35. Litsky, M.L.; Pfeiffer, D.R. Regulation of the mitochondrial Ca2+ uniporter by external adenine nucleotides: The uniporter behaves like a gated channel which is regulated by nucleotides and divalent cations. Biochemistry 1997, 36, 7071–7080. [Google Scholar] [CrossRef] [PubMed]
  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 2011, 476, 336–340. [Google Scholar] [CrossRef] [PubMed]
  37. Baughman, J.M.; Perocchi, F.; Girgis, H.S.; Plovanich, M.; Belcher-Timme, C.A.; Sancak, Y.; Bao, X.R.; Strittmatter, L.; Goldberger, O.; Bogorad, R.L.; et al. Integrative genomics identifies MCU as an essential component of the mitochondrial calcium uniporter. Nature 2011, 476, 341–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Perocchi, F.; Gohil, V.M.; Girgis, H.S.; Bao, X.R.; McCombs, J.E.; Palmer, A.E.; Mootha, V.K. MICU1 encodes a mitochondrial EF hand protein required for Ca2+ uptake. Nature 2010, 467, 291–296. [Google Scholar] [CrossRef] [PubMed]
  39. Mallilankaraman, K.; Doonan, P.; Cárdenas, C.; Chandramoorthy, H.C.; Müller, M.; Miller, R.; Hoffman, N.E.; Gandhirajan, R.K.; Molgó, J.; Birnbaum, M.J.; et al. MICU1 is an essential gatekeeper for MCU-mediated mitochondrial Ca2+ uptake that regulates cell survival. Cell 2012, 151, 630–644. [Google Scholar] [CrossRef]
  40. Plovanich, M.; Bogorad, R.L.; Sancak, Y.; Kamer, K.J.; Strittmatter, L.; Li, A.A.; Girgis, H.S.; Kuchimanchi, S.; De Groot, J.; Speciner, L.; et al. MICU2, a paralog of MICU1, resides within the mitochondrial uniporter complex to regulate calcium handling. PLoS ONE 2013, 8, e55785. [Google Scholar] [CrossRef]
  41. Ahuja, M.; Muallem, S. The gatekeepers of mitochondrial calcium influx: MICU1 and MICU2. EMBO Rep. 2014, 15, 205–206. [Google Scholar] [CrossRef] [PubMed]
  42. Waldeck-Weiermair, M.; Malli, R.; Parichatikanond, W.; Gottschalk, B.; Madreiter-Sokolowski, C.T.; Klec, C.; Rost, R.; Graier, W.F. Rearrangement of MICU1 multimers for activation of MCU is solely controlled by cytosolic Ca2+. Sci. Rep. 2015, 5, 15602. [Google Scholar] [CrossRef] [PubMed]
  43. Hoppe, U.C. Mitochondrial calcium channels. FEBS Lett. 2010, 584, 1975–1981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Bondarenko, A.I.; Parichatikanond, W.; Madreiter, C.T.; Rost, R.; Waldeck-Weiermair, M.; Malli, R.; Graier, W.F. UCP2 modulates single-channel properties of a MCU-dependent Ca2+ inward current in mitochondria. Pflügers Arch.-Eur. J. Physiol. 2015, 467, 2509–2518. [Google Scholar] [CrossRef]
  45. Motloch, L.J.; Larbig, R.; Gebing, T.; Reda, S.; Schwaiger, A.; Leitner, J.; Wolny, M.; Eckardt, L.; Hoppe, U.C. By Regulating Mitochondrial Ca2+-Uptake UCP2 Modulates Intracellular Ca2+. PLoS ONE 2016, 11, e0148359. [Google Scholar] [CrossRef] [PubMed]
  46. Brookes, P.S.; Parker, N.; Buckingham, J.A.; Vidal-Puig, A.; Halestrap, A.P.; Gunter, T.E.; Nicholls, D.G.; Bernardi, P.; Lemasters, J.J.; Brand, M.D. UCPs—Unlikely calcium porters. Nat. Cell Biol. 2008, 10, 1235–1237. [Google Scholar] [CrossRef]
  47. Motloch, L.J.; Reda, S.; Wolny, M.; Hoppe, U.C. UCP2 Modulates Cardioprotective Effects of Ru360 in Isolated Cardiomyocytes during Ischemia. Pharm. Basel Switz. 2015, 8, 474–482. [Google Scholar] [CrossRef] [Green Version]
  48. Motloch, L.J.; Gebing, T.; Reda, S.; Schwaiger, A.; Wolny, M.; Hoppe, U.C. UCP3 Regulates Single-Channel Activity of the Cardiac mCa1. J. Membr. Biol. 2016, 249, 577–584. [Google Scholar] [CrossRef]
  49. Waldeck-Weiermair, M.; Malli, R.; Naghdi, S.; Trenker, M.; Kahn, M.J.; Graier, W.F. The contribution of UCP2 and UCP3 to mitochondrial Ca2+ uptake is differentially determined by the source of supplied Ca2+. Cell Calcium 2010, 47, 433–440. [Google Scholar] [CrossRef]
  50. Waldeck-Weiermair, M.; Duan, X.; Naghdi, S.; Khan, M.J.; Trenker, M.; Malli, R.; Graier, W.F. Uncoupling protein 3 adjusts mitochondrial Ca2+ uptake to high and low Ca2+ signals. Cell Calcium 2010, 48, 288–301. [Google Scholar] [CrossRef]
  51. Larbig, R.; Reda, S.; Paar, V.; Trost, A.; Leitner, J.; Weichselbaumer, S.; Motloch, K.A.; Wernly, B.; Arrer, A.; Strauss, B.; et al. Through modulation of cardiac Ca2+ handling, UCP2 affects cardiac electrophysiology and influences the susceptibility for Ca2+-mediated arrhythmias. Exp. Physiol. 2017, 102, 650–662. [Google Scholar] [CrossRef] [PubMed]
  52. 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. [Google Scholar] [CrossRef]
  53. 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. [Google Scholar] [CrossRef] [PubMed]
  54. Kohlhaas, M.; Liu, T.; Knopp, A.; Zeller, T.; Ong, M.F.; Böhm, M.; O’Rourke, B.; Maack, C. Elevated cytosolic Na+ increases mitochondrial formation of reactive oxygen species in failing cardiac myocytes. Circulation 2010, 121, 1606–1613. [Google Scholar] [CrossRef]
  55. Smets, I.; Caplanusi, A.; Despa, S.; Molnar, Z.; Radu, M.; VandeVen, M.; Ameloot, M.; Steels, P. Ca2+ uptake in mitochondria occurs via the reverse action of the Na+/Ca2+ exchanger in metabolically inhibited MDCK cells. Am. J. Physiol. Renal Physiol. 2004, 286, F784–F794. [Google Scholar] [CrossRef]
  56. Kim, B.; Matsuoka, S. Cytoplasmic Na+-dependent modulation of mitochondrial Ca2+ via electrogenic mitochondrial Na+-Ca2+ exchange. J. Physiol. 2008, 586, 1683–1697. [Google Scholar] [CrossRef]
  57. Wit, A.L.; Rosen, M.R. Pathophysiologic mechanisms of cardiac arrhythmias. Am. Heart J. 1983, 106, 798–811. [Google Scholar] [CrossRef]
  58. Akar, F.G.; Hajjar, R.J. Gene therapies for arrhythmias in heart failure. Pflugers Arch. 2014, 466, 1211–1217. [Google Scholar] [CrossRef]
  59. Motloch, L.J.; Akar, F.G. Gene therapy to restore electrophysiological function in heart failure. Expert Opin. Biol. Ther. 2015, 15, 803–817. [Google Scholar] [CrossRef]
  60. Marban, E.; Robinson, S.W.; Wier, W.G. Mechanisms of arrhythmogenic delayed and early afterdepolarizations in ferret ventricular muscle. J. Clin. Investig. 1986, 78, 1185–1192. [Google Scholar] [CrossRef]
  61. January, C.T.; Moscucci, A. Cellular mechanisms of early afterdepolarizations. Ann. N. Y. Acad. Sci. 1992, 644, 23–32. [Google Scholar] [CrossRef] [PubMed]
  62. Tse, G. Mechanisms of cardiac arrhythmias. J. Arrhythm. 2016, 32, 75–81. [Google Scholar] [CrossRef] [PubMed]
  63. Venetucci, L.A.; Trafford, A.W.; O’Neill, S.C.; Eisner, D.A. The sarcoplasmic reticulum and arrhythmogenic calcium release. Cardiovasc. Res. 2008, 77, 285–292. [Google Scholar] [CrossRef] [PubMed]
  64. Venetucci, L.A.; Trafford, A.W.; Eisner, D.A. Increasing ryanodine receptor open probability alone does not produce arrhythmogenic calcium waves: Threshold sarcoplasmic reticulum calcium content is required. Circ. Res. 2007, 100, 105–111. [Google Scholar] [CrossRef] [PubMed]
  65. Laitinen, P.J.; Brown, K.M.; Piippo, K.; Swan, H.; Devaney, J.M.; Brahmbhatt, B.; Donarum, E.A.; Marino, M.; Tiso, N.; Viitasalo, M.; et al. Mutations of the cardiac ryanodine receptor (RyR2) gene in familial polymorphic ventricular tachycardia. Circulation 2001, 103, 485–490. [Google Scholar] [CrossRef] [PubMed]
  66. Postma, A.V.; Denjoy, I.; Hoorntje, T.M.; Lupoglazoff, J.-M.; Da Costa, A.; Sebillon, P.; Mannens, M.M.A.M.; Wilde, A.A.M.; Guicheney, P. Absence of calsequestrin 2 causes severe forms of catecholaminergic polymorphic ventricular tachycardia. Circ. Res. 2002, 91, e21–e26. [Google Scholar] [CrossRef] [PubMed]
  67. Jiang, D.; Xiao, B.; Yang, D.; Wang, R.; Choi, P.; Zhang, L.; Cheng, H.; Chen, S.R.W. RyR2 mutations linked to ventricular tachycardia and sudden death reduce the threshold for store-overload-induced Ca2+ release (SOICR). Proc. Natl. Acad. Sci. USA 2004, 101, 13062–13067. [Google Scholar] [CrossRef]
  68. Jiang, D.; Wang, R.; Xiao, B.; Kong, H.; Hunt, D.J.; Choi, P.; Zhang, L.; Chen, S.R.W. Enhanced store overload-induced Ca2+ release and channel sensitivity to luminal Ca2+ activation are common defects of RyR2 mutations linked to ventricular tachycardia and sudden death. Circ. Res. 2005, 97, 1173–1181. [Google Scholar] [CrossRef]
  69. Terentyev, D.; Nori, A.; Santoro, M.; Viatchenko-Karpinski, S.; Kubalova, Z.; Gyorke, I.; Terentyeva, R.; Vedamoorthyrao, S.; Blom, N.A.; Valle, G.; et al. Abnormal interactions of calsequestrin with the ryanodine receptor calcium release channel complex linked to exercise-induced sudden cardiac death. Circ. Res. 2006, 98, 1151–1158. [Google Scholar] [CrossRef]
  70. Marx, S.O.; Reiken, S.; Hisamatsu, Y.; Jayaraman, T.; Burkhoff, D.; Rosemblit, N.; Marks, A.R. PKA phosphorylation dissociates FKBP12.6 from the calcium release channel (ryanodine receptor): Defective regulation in failing hearts. Cell 2000, 101, 365–376. [Google Scholar] [CrossRef]
  71. Ai, X.; Curran, J.W.; Shannon, T.R.; Bers, D.M.; Pogwizd, S.M. Ca2+/calmodulin-dependent protein kinase modulates cardiac ryanodine receptor phosphorylation and sarcoplasmic reticulum Ca2+ leak in heart failure. Circ. Res. 2005, 97, 1314–1322. [Google Scholar] [CrossRef] [PubMed]
  72. Eisner, D.A.; Trafford, A.W.; Díaz, M.E.; Overend, C.L.; O’Neill, S.C. The control of Ca release from the cardiac sarcoplasmic reticulum: Regulation versus autoregulation. Cardiovasc. Res. 1998, 38, 589–604. [Google Scholar] [CrossRef]
  73. Liu, N.; Ruan, Y.; Denegri, M.; Bachetti, T.; Li, Y.; Colombi, B.; Napolitano, C.; Coetzee, W.A.; Priori, S.G. Calmodulin kinase II inhibition prevents arrhythmias in RyR2(R4496C+/-) mice with catecholaminergic polymorphic ventricular tachycardia. J. Mol. Cell. Cardiol. 2011, 50, 214–222. [Google Scholar] [CrossRef] [PubMed]
  74. Maier, L.S.; Bers, D.M. Role of Ca2+/calmodulin-dependent protein kinase (CaMK) in excitation-contraction coupling in the heart. Cardiovasc. Res. 2007, 73, 631–640. [Google Scholar] [CrossRef] [PubMed]
  75. Brooksby, P.; Levi, A.J.; Jones, J.V. The electrophysiological characteristics of hypertrophied ventricular myocytes from the spontaneously hypertensive rat. J. Hypertens. 1993, 11, 611–622. [Google Scholar] [CrossRef]
  76. Cerbai, E.; Barbieri, M.; Li, Q.; Mugelli, A. Ionic basis of action potential prolongation of hypertrophied cardiac myocytes isolated from hypertensive rats of different ages. Cardiovasc. Res. 1994, 28, 1180–1187. [Google Scholar] [CrossRef]
  77. Richard, S.; Leclercq, F.; Lemaire, S.; Piot, C.; Nargeot, J. Ca2+ currents in compensated hypertrophy and heart failure. Cardiovasc. Res. 1998, 37, 300–311. [Google Scholar] [CrossRef]
  78. Hill, J.A. Electrical remodeling in cardiac hypertrophy. Trends Cardiovasc. Med. 2003, 13, 316–322. [Google Scholar] [CrossRef]
  79. Ryder, K.O.; Bryant, S.M.; Hart, G. Membrane current changes in left ventricular myocytes isolated from guinea pigs after abdominal aortic coarctation. Cardiovasc. Res. 1993, 27, 1278–1287. [Google Scholar] [CrossRef]
  80. Ouadid, H.; Albat, B.; Nargeot, J. Calcium currents in diseased human cardiac cells. J. Cardiovasc. Pharmacol. 1995, 25, 282–291. [Google Scholar] [CrossRef]
  81. de Lucia, C.; Eguchi, A.; Koch, W.J. New Insights in Cardiac β-Adrenergic Signaling During Heart Failure and Aging. Front. Pharmacol. 2018, 9, 904. [Google Scholar] [CrossRef] [PubMed]
  82. Letsas, K.P.; Sacher, F.; Probst, V.; Weber, R.; Knecht, S.; Kalusche, D.; Haïssaguerre, M.; Arentz, T. Prevalence of early repolarization pattern in inferolateral leads in patients with Brugada syndrome. Heart Rhythm 2008, 5, 1685–1689. [Google Scholar] [CrossRef] [PubMed]
  83. Burashnikov, E.; Pfeiffer, R.; Barajas-Martinez, H.; Delpón, E.; Hu, D.; Desai, M.; Borggrefe, M.; Häissaguerre, M.; Kanter, R.; Pollevick, G.D.; et al. Mutations in the cardiac L-type calcium channel associated with inherited J-wave syndromes and sudden cardiac death. Heart Rhythm 2010, 7, 1872–1882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Templin, C.; Ghadri, J.-R.; Rougier, J.-S.; Baumer, A.; Kaplan, V.; Albesa, M.; Sticht, H.; Rauch, A.; Puleo, C.; Hu, D.; et al. Identification of a novel loss-of-function calcium channel gene mutation in short QT syndrome (SQTS6). Eur. Heart J. 2011, 32, 1077–1088. [Google Scholar] [CrossRef] [PubMed]
  85. Bjerregaard, P.; Nallapaneni, H.; Gussak, I. Short QT interval in clinical practice. J. Electrocardiol. 2010, 43, 390–395. [Google Scholar] [CrossRef] [PubMed]
  86. Watanabe, H.; Makiyama, T.; Koyama, T.; Kannankeril, P.J.; Seto, S.; Okamura, K.; Oda, H.; Itoh, H.; Okada, M.; Tanabe, N.; et al. High prevalence of early repolarization in short QT syndrome. Heart Rhythm 2010, 7, 647–652. [Google Scholar] [CrossRef] [PubMed]
  87. Napolitano, C.; Antzelevitch, C. Phenotypical manifestations of mutations in the genes encoding subunits of the cardiac voltage-dependent L-type calcium channel. Circ. Res. 2011, 108, 607–618. [Google Scholar] [CrossRef] [PubMed]
  88. Bers, D.M.; Pogwizd, S.M.; Schlotthauer, K. Upregulated Na/Ca exchange is involved in both contractile dysfunction and arrhythmogenesis in heart failure. Basic Res. Cardiol. 2002, 97 (Suppl. 1), I36–I42. [Google Scholar] [CrossRef] [PubMed]
  89. Lederer, W.J.; Tsien, R.W. Transient inward current underlying arrhythmogenic effects of cardiotonic steroids in Purkinje fibres. J. Physiol. 1976, 263, 73–100. [Google Scholar] [CrossRef]
  90. Rahm, A.-K.; Lugenbiel, P.; Schweizer, P.A.; Katus, H.A.; Thomas, D. Role of ion channels in heart failure and channelopathies. Biophys. Rev. 2018, 10, 1097–1106. [Google Scholar] [CrossRef]
  91. Shaw, R.M.; Rudy, Y. The vulnerable window for unidirectional block in cardiac tissue: Characterization and dependence on membrane excitability and intercellular coupling. J. Cardiovasc. Electrophysiol. 1995, 6, 115–131. [Google Scholar] [CrossRef] [PubMed]
  92. Patron, M.; Checchetto, V.; Raffaello, A.; Teardo, E.; Vecellio Reane, D.; Mantoan, M.; Granatiero, V.; Szabò, I.; De Stefani, D.; Rizzuto, R. MICU1 and MICU2 finely tune the mitochondrial Ca2+ uniporter by exerting opposite effects on MCU activity. Mol. Cell 2014, 53, 726–737. [Google Scholar] [CrossRef] [PubMed]
  93. Sommakia, S.; Houlihan, P.R.; Deane, S.S.; Simcox, J.A.; Torres, N.S.; Jeong, M.-Y.; Winge, D.R.; Villanueva, C.J.; Chaudhuri, D. Mitochondrial cardiomyopathies feature increased uptake and diminished efflux of mitochondrial calcium. J. Mol. Cell. Cardiol. 2017, 113, 22–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Santulli, G.; Xie, W.; Reiken, S.R.; Marks, A.R. Mitochondrial calcium overload is a key determinant in heart failure. Proc. Natl. Acad. Sci. USA 2015, 112, 11389–11394. [Google Scholar] [CrossRef] [Green Version]
  95. Odagiri, K.; Katoh, H.; Kawashima, H.; Tanaka, T.; Ohtani, H.; Saotome, M.; Urushida, T.; Satoh, H.; Hayashi, H. Local control of mitochondrial membrane potential, permeability transition pore and reactive oxygen species by calcium and calmodulin in rat ventricular myocytes. J. Mol. Cell. Cardiol. 2009, 46, 989–997. [Google Scholar] [CrossRef] [Green Version]
  96. Turrens, J.F. Mitochondrial formation of reactive oxygen species. J. Physiol. 2003, 552, 335–344. [Google Scholar] [CrossRef]
  97. Zorov, D.B.; Filburn, C.R.; Klotz, L.O.; Zweier, J.L.; Sollott, S.J. Reactive oxygen species (ROS)-induced ROS release: A new phenomenon accompanying induction of the mitochondrial permeability transition in cardiac myocytes. J. Exp. Med. 2000, 192, 1001–1014. [Google Scholar] [CrossRef]
  98. Zorov, D.B.; Juhaszova, M.; Sollott, S.J. Mitochondrial ROS-induced ROS release: An update and review. Biochim. Biophys. Acta 2006, 1757, 509–517. [Google Scholar] [CrossRef] [Green Version]
  99. Aon, M.A.; Cortassa, S.; Marbán, E.; O’Rourke, B. Synchronized whole cell oscillations in mitochondrial metabolism triggered by a local release of reactive oxygen species in cardiac myocytes. J. Biol. Chem. 2003, 278, 44735–44744. [Google Scholar] [CrossRef]
  100. Aon, M.A.; Cortassa, S.; O’Rourke, B. The Fundamental Organization of Cardiac Mitochondria as a Network of Coupled Oscillators. Biophys. J. 2006, 91, 4317–4327. [Google Scholar] [CrossRef] [Green Version]
  101. Aon, M.A.; Cortassa, S.; Maack, C.; O’Rourke, B. Sequential opening of mitochondrial ion channels as a function of glutathione redox thiol status. J. Biol. Chem. 2007, 282, 21889–21900. [Google Scholar] [CrossRef] [PubMed]
  102. Hopper, R.K.; Carroll, S.; Aponte, A.M.; Johnson, D.T.; French, S.; Shen, R.-F.; Witzmann, F.A.; Harris, R.A.; Balaban, R.S. Mitochondrial matrix phosphoproteome: Effect of extra mitochondrial calcium. Biochemistry 2006, 45, 2524–2536. [Google Scholar] [CrossRef] [PubMed]
  103. Motloch, L.J.; Hu, J.; Akar, F.G. The Mitochondrial Translocator Protein and Arrhythmogenesis in Ischemic Heart Disease. Oxid. Med. Cell. Longev. 2015. [Google Scholar] [CrossRef] [PubMed]
  104. Gambardella, J.; Sorriento, D.; Ciccarelli, M.; Del Giudice, C.; Fiordelisi, A.; Napolitano, L.; Trimarco, B.; Iaccarino, G.; Santulli, G. Functional Role of Mitochondria in Arrhythmogenesis. Adv. Exp. Med. Biol. 2017, 982, 191–202. [Google Scholar]
  105. Nakayama, H.; Chen, X.; Baines, C.P.; Klevitsky, R.; Zhang, X.; Zhang, H.; Jaleel, N.; Chua, B.H.L.; Hewett, T.E.; Robbins, J.; et al. Ca2+-and mitochondrial-dependent cardiomyocyte necrosis as a primary mediator of heart failure. J. Clin. Investig. 2007, 117, 2431–2444. [Google Scholar] [CrossRef]
  106. Rizzuto, R.; De Stefani, D.; Raffaello, A.; Mammucari, C. Mitochondria as sensors and regulators of calcium signalling. Nat. Rev. Mol. Cell Biol. 2012, 13, 566–578. [Google Scholar] [CrossRef]
  107. Fozzard, H.A. Cardiac sodium and calcium channels: A history of excitatory currents. Cardiovasc. Res. 2002, 55, 1–8. [Google Scholar] [CrossRef]
  108. Landstrom, A.P.; Dobrev, D.; Wehrens, X.H.T. Calcium Signaling and Cardiac Arrhythmias. Circ. Res. 2017, 120, 1969–1993. [Google Scholar] [CrossRef]
  109. Briot, J.; Tétreault, M.-P.; Bourdin, B.; Parent, L. Inherited Ventricular Arrhythmias: The Role of the Multi-Subunit Structure of the L-Type Calcium Channel Complex. Adv. Exp. Med. Biol. 2017, 966, 55–64. [Google Scholar]
  110. Hamilton, S.; Terentyev, D. Altered Intracellular Calcium Homeostasis and Arrhythmogenesis in the Aged Heart. Int. J. Mol. Sci. 2019, 20, 2386. [Google Scholar] [CrossRef]
  111. Tomaselli, G.F.; Beuckelmann, D.J.; Calkins, H.G.; Berger, R.D.; Kessler, P.D.; Lawrence, J.H.; Kass, D.; Feldman, A.M.; Marban, E. Sudden cardiac death in heart failure. The role of abnormal repolarization. Circulation 1994, 90, 2534–2539. [Google Scholar] [CrossRef] [PubMed]
  112. Fleckenstein, A. History of calcium antagonists. Circ. Res. 1983, 52, I3–I16. [Google Scholar]
  113. Rosso, R.; Kalman, J.M.; Rogowski, O.; Diamant, S.; Birger, A.; Biner, S.; Belhassen, B.; Viskin, S. Calcium channel blockers and beta-blockers versus beta-blockers alone for preventing exercise-induced arrhythmias in catecholaminergic polymorphic ventricular tachycardia. Heart Rhythm 2007, 4, 1149–1154. [Google Scholar] [CrossRef] [PubMed]
  114. Singh, B.N. A fourth class of anti-dysrhythmic action? Effect of verapamil on ouabain toxicity, on atrial and ventricular intracellular potentials, and on other features of cardiac function. Cardiovasc. Res. 1972, 6, 109–119. [Google Scholar] [CrossRef] [PubMed]
  115. Rodríguez Padial, L.; Barón-Esquivias, G.; Hernández Madrid, A.; Marzal Martín, D.; Pallarés-Carratalá, V.; de la Sierra, A. Clinical Experience with Diltiazem in the Treatment of Cardiovascular Diseases. Cardiol. Ther. 2016, 5, 75–82. [Google Scholar] [CrossRef] [PubMed]
  116. Page, R.L.; Joglar, J.A.; Caldwell, M.A.; Calkins, H.; Conti, J.B.; Deal, B.J.; Estes, N.A.M.; Field, M.E.; Goldberger, Z.D.; Hammill, S.C.; et al. 2015 ACC/AHA/HRS Guideline for the Management of Adult Patients With Supraventricular Tachycardia: A Report of the American College of Cardiology/American Heart Association Task Force on Clinical Practice Guidelines and the Heart Rhythm Society. J. Am. Coll. Cardiol. 2016, 67, e27–e115. [Google Scholar] [CrossRef] [PubMed]
  117. Watanabe, H.; Chopra, N.; Laver, D.; Hwang, H.S.; Davies, S.S.; Roach, D.E.; Duff, H.J.; Roden, D.M.; Wilde, A.A.M.; Knollmann, B.C. Flecainide prevents catecholaminergic polymorphic ventricular tachycardia in mice and humans. Nat. Med. 2009, 15, 380–383. [Google Scholar] [CrossRef]
  118. van der Werf, C.; Kannankeril, P.J.; Sacher, F.; Krahn, A.D.; Viskin, S.; Leenhardt, A.; Shimizu, W.; Sumitomo, N.; Fish, F.A.; Bhuiyan, Z.A.; et al. Flecainide therapy reduces exercise-induced ventricular arrhythmias in patients with catecholaminergic polymorphic ventricular tachycardia. J. Am. Coll. Cardiol. 2011, 57, 2244–2254. [Google Scholar] [CrossRef]
  119. Singh, S.N.; Fletcher, R.D.; Fisher, S.G.; Singh, B.N.; Lewis, H.D.; Deedwania, P.C.; Massie, B.M.; Colling, C.; Lazzeri, D. Amiodarone in patients with congestive heart failure and asymptomatic ventricular arrhythmia. Survival Trial of Antiarrhythmic Therapy in Congestive Heart Failure. N. Engl. J. Med. 1995, 333, 77–82. [Google Scholar] [CrossRef]
  120. Waldo, A.L.; Camm, A.J.; deRuyter, H.; Friedman, P.L.; MacNeil, D.J.; Pauls, J.F.; Pitt, B.; Pratt, C.M.; Schwartz, P.J.; Veltri, E.P. Effect of d-sotalol on mortality in patients with left ventricular dysfunction after recent and remote myocardial infarction. The SWORD Investigators. Survival With Oral d-Sotalol. Lancet Lond. Engl. 1996, 348, 7–12. [Google Scholar] [CrossRef]
  121. Køber, L.; Thomsen, P.E.B.; Møller, M.; Torp-Pedersen, C.; Carlsen, J.; Sandøe, E.; Egstrup, K.; Agner, E.; Videbaek, J.; Marchant, B.; et al. Effect of dofetilide in patients with recent myocardial infarction and left-ventricular dysfunction: A randomised trial. Lancet Lond. Engl. 2000, 356, 2052–2058. [Google Scholar] [CrossRef]
  122. Kuck, K.H.; Cappato, R.; Siebels, J.; Rüppel, R. Randomized comparison of antiarrhythmic drug therapy with implantable defibrillators in patients resuscitated from cardiac arrest: The Cardiac Arrest Study Hamburg (CASH). Circulation 2000, 102, 748–754. [Google Scholar] [CrossRef] [PubMed]
  123. Donahue, J.K. Gene therapy for ventricular tachyarrhythmias. Gene Ther. 2012, 19, 600–605. [Google Scholar] [CrossRef] [PubMed]
  124. Gould, P.A.; Krahn, A.D. Canadian Heart Rhythm Society Working Group on Device Advisories Complications associated with implantable cardioverter-defibrillator replacement in response to device advisories. JAMA 2006, 295, 1907–1911. [Google Scholar] [CrossRef] [PubMed]
  125. Bonci, D.; Cittadini, A.; Latronico, M.V.G.; Borello, U.; Aycock, J.K.; Drusco, A.; Innocenzi, A.; Follenzi, A.; Lavitrano, M.; Monti, M.G.; et al. “Advanced” generation lentiviruses as efficient vectors for cardiomyocyte gene transduction in vitro and in vivo. Gene Ther. 2003, 10, 630–636. [Google Scholar] [CrossRef]
  126. Cockrell, A.S.; Kafri, T. Gene delivery by lentivirus vectors. Mol. Biotechnol. 2007, 36, 184–204. [Google Scholar] [CrossRef]
  127. Bongianino, R.; Priori, S.G. Gene therapy to treat cardiac arrhythmias. Nat. Rev. Cardiol. 2015, 12, 531–546. [Google Scholar] [CrossRef]
  128. Hulot, J.-S.; Ishikawa, K.; Hajjar, R.J. Gene therapy for the treatment of heart failure: Promise postponed. Eur. Heart J. 2016, 37, 1651–1658. [Google Scholar] [CrossRef]
  129. Greenberg, B. Gene therapy for heart failure. Trends Cardiovasc. Med. 2017, 27, 216–222. [Google Scholar] [CrossRef]
  130. Alba, R.; Bosch, A.; Chillon, M. Gutless adenovirus: Last-generation adenovirus for gene therapy. Gene Ther. 2005, 12 (Suppl. 1), S18–S27. [Google Scholar] [CrossRef]
  131. Liu, Z.; Donahue, J.K. The Use of Gene Therapy for Ablation of Atrial Fibrillation. Arrhythmia Electrophysiol. Rev. 2014, 3, 139–144. [Google Scholar] [CrossRef] [PubMed]
  132. Zincarelli, C.; Soltys, S.; Rengo, G.; Rabinowitz, J.E. Analysis of AAV serotypes 1–9 mediated gene expression and tropism in mice after systemic injection. Mol. Ther. J. Am. Soc. Gene Ther. 2008, 16, 1073–1080. [Google Scholar] [CrossRef] [PubMed]
  133. Wang, D.; Gao, G. State-of-the-art human gene therapy: Part I. Gene delivery technologies. Discov. Med. 2014, 18, 67–77. [Google Scholar] [PubMed]
  134. Wang, D.; Gao, G. State-of-the-art human gene therapy: Part II. Gene therapy strategies and clinical applications. Discov. Med. 2014, 18, 151–161. [Google Scholar] [PubMed]
  135. Fish, K.M.; Ishikawa, K. Advances in gene therapy for heart failure. Discov. Med. 2015, 19, 285–291. [Google Scholar] [PubMed]
  136. Hammoudi, N.; Ishikawa, K.; Hajjar, R.J. Adeno-associated virus-mediated gene therapy in cardiovascular disease. Curr. Opin. Cardiol. 2015, 30, 228–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Cong, L.; Ran, F.A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P.D.; Wu, X.; Jiang, W.; Marraffini, L.A.; et al. Multiplex genome engineering using CRISPR/Cas systems. Science 2013, 339, 819–823. [Google Scholar] [CrossRef]
  138. Gori, J.L.; Hsu, P.D.; Maeder, M.L.; Shen, S.; Welstead, G.G.; Bumcrot, D. Delivery and Specificity of CRISPR-Cas9 Genome Editing Technologies for Human Gene Therapy. Hum. Gene Ther. 2015, 26, 443–451. [Google Scholar] [CrossRef]
  139. Nelson, C.E.; Hakim, C.H.; Ousterout, D.G.; Thakore, P.I.; Moreb, E.A.; Castellanos Rivera, R.M.; Madhavan, S.; Pan, X.; Ran, F.A.; Yan, W.X.; et al. In vivo genome editing improves muscle function in a mouse model of Duchenne muscular dystrophy. Science 2016, 351, 403–407. [Google Scholar] [CrossRef] [Green Version]
  140. Keeler, A.M.; ElMallah, M.K.; Flotte, T.R. Gene Therapy 2017: Progress and Future Directions. Clin. Transl. Sci. 2017, 10, 242–248. [Google Scholar] [CrossRef] [Green Version]
  141. Nguyen, A.H.; Marsh, P.; Schmiess-Heine, L.; Burke, P.J.; Lee, A.; Lee, J.; Cao, H. Cardiac tissue engineering: State-of-the-art methods and outlook. J. Biol. Eng. 2019, 13, 57. [Google Scholar] [CrossRef] [PubMed]
  142. Goldstein, R.E.; Boccuzzi, S.J.; Cruess, D.; Nattel, S. Diltiazem increases late-onset congestive heart failure in postinfarction patients with early reduction in ejection fraction. The Adverse Experience Committee; and the Multicenter Diltiazem Postinfarction Research Group. Circulation 1991, 83, 52–60. [Google Scholar] [CrossRef] [PubMed]
  143. Murata, M.; Cingolani, E.; McDonald, A.D.; Donahue, J.K.; Marbán, E. Creation of a genetic calcium channel blocker by targeted gem gene transfer in the heart. Circ. Res. 2004, 95, 398–405. [Google Scholar] [CrossRef] [PubMed]
  144. Cingolani, E.; Ramirez Correa, G.A.; Kizana, E.; Murata, M.; Cho, H.C.; Marbán, E. Gene therapy to inhibit the calcium channel beta subunit: Physiological consequences and pathophysiological effects in models of cardiac hypertrophy. Circ. Res. 2007, 101, 166–175. [Google Scholar] [CrossRef] [PubMed]
  145. Subramanyam, P.; Chang, D.D.; Fang, K.; Xie, W.; Marks, A.R.; Colecraft, H.M. Manipulating L-type calcium channels in cardiomyocytes using split-intein protein transsplicing. Proc. Natl. Acad. Sci. USA 2013, 110, 15461–15466. [Google Scholar] [CrossRef] [Green Version]
  146. O’Rourke, B.; Kass, D.A.; Tomaselli, G.F.; Kääb, S.; Tunin, R.; Marbán, E. Mechanisms of altered excitation-contraction coupling in canine tachycardia-induced heart failure, I: Experimental studies. Circ. Res. 1999, 84, 562–570. [Google Scholar] [CrossRef]
  147. Kawase, Y.; Ly, H.Q.; Prunier, F.; Lebeche, D.; Shi, Y.; Jin, H.; Hadri, L.; Yoneyama, R.; Hoshino, K.; Takewa, Y.; et al. Reversal of cardiac dysfunction after long-term expression of SERCA2a by gene transfer in a pre-clinical model of heart failure. J. Am. Coll. Cardiol. 2008, 51, 1112–1119. [Google Scholar] [CrossRef]
  148. Mariani, J.A.; Smolic, A.; Preovolos, A.; Byrne, M.J.; Power, J.M.; Kaye, D.M. Augmentation of left ventricular mechanics by recirculation-mediated AAV2/1-SERCA2a gene delivery in experimental heart failure. Eur. J. Heart Fail. 2011, 13, 247–253. [Google Scholar] [CrossRef] [Green Version]
  149. Byrne, M.J.; Power, J.M.; Preovolos, A.; Mariani, J.A.; Hajjar, R.J.; Kaye, D.M. Recirculating cardiac delivery of AAV2/1SERCA2a improves myocardial function in an experimental model of heart failure in large animals. Gene Ther. 2008, 15, 1550–1557. [Google Scholar] [CrossRef] [Green Version]
  150. del Monte, F.; Lebeche, D.; Guerrero, J.L.; Tsuji, T.; Doye, A.A.; Gwathmey, J.K.; Hajjar, R.J. Abrogation of ventricular arrhythmias in a model of ischemia and reperfusion by targeting myocardial calcium cycling. Proc. Natl. Acad. Sci. USA 2004, 101, 5622–5627. [Google Scholar] [CrossRef] [Green Version]
  151. Prunier, F.; Kawase, Y.; Gianni, D.; Scapin, C.; Danik, S.B.; Ellinor, P.T.; Hajjar, R.J.; Del Monte, F. Prevention of ventricular arrhythmias with sarcoplasmic reticulum Ca2+ ATPase pump overexpression in a porcine model of ischemia reperfusion. Circulation 2008, 118, 614–624. [Google Scholar] [CrossRef] [PubMed]
  152. Cutler, M.J.; Wan, X.; Laurita, K.R.; Hajjar, R.J.; Rosenbaum, D.S. Targeted SERCA2a gene expression identifies molecular mechanism and therapeutic target for arrhythmogenic cardiac alternans. Circ. Arrhythm. Electrophysiol. 2009, 2, 686–694. [Google Scholar] [CrossRef] [PubMed]
  153. Lyon, A.R.; Bannister, M.L.; Collins, T.; Pearce, E.; Sepehripour, A.H.; Dubb, S.S.; Garcia, E.; O’Gara, P.; Liang, L.; Kohlbrenner, E.; et al. SERCA2a gene transfer decreases sarcoplasmic reticulum calcium leak and reduces ventricular arrhythmias in a model of chronic heart failure. Circ. Arrhythm. Electrophysiol. 2011, 4, 362–372. [Google Scholar] [CrossRef] [PubMed]
  154. Cutler, M.J.; Wan, X.; Plummer, B.N.; Liu, H.; Deschenes, I.; Laurita, K.R.; Hajjar, R.J.; Rosenbaum, D.S. Targeted sarcoplasmic reticulum Ca2+ ATPase 2a gene delivery to restore electrical stability in the failing heart. Circulation 2012, 126, 2095–2104. [Google Scholar] [CrossRef] [PubMed]
  155. Davia, K.; Bernobich, E.; Ranu, H.K.; del Monte, F.; Terracciano, C.M.; MacLeod, K.T.; Adamson, D.L.; Chaudhri, B.; Hajjar, R.J.; Harding, S.E. SERCA2A overexpression decreases the incidence of aftercontractions in adult rabbit ventricular myocytes. J. Mol. Cell. Cardiol. 2001, 33, 1005–1015. [Google Scholar] [CrossRef]
  156. Xie, L.-H.; Sato, D.; Garfinkel, A.; Qu, Z.; Weiss, J.N. Intracellular Ca alternans: Coordinated regulation by sarcoplasmic reticulum release, uptake, and leak. Biophys. J. 2008, 95, 3100–3110. [Google Scholar] [CrossRef]
  157. Giordano, F.J.; He, H.; McDonough, P.; Meyer, M.; Sayen, M.R.; Dillmann, W.H. Adenovirus-mediated gene transfer reconstitutes depressed sarcoplasmic reticulum Ca2+-ATPase levels and shortens prolonged cardiac myocyte Ca2+ transients. Circulation 1997, 96, 400–403. [Google Scholar] [CrossRef]
  158. Hajjar, R.J.; Kang, J.X.; Gwathmey, J.K.; Rosenzweig, A. Physiological effects of adenoviral gene transfer of sarcoplasmic reticulum calcium ATPase in isolated rat myocytes. Circulation 1997, 95, 423–429. [Google Scholar] [CrossRef]
  159. Terracciano, C.M.N.; Hajjar, R.J.; Harding, S.E. Overexpression of SERCA2a accelerates repolarisation in rabbit ventricular myocytes. Cell Calcium 2002, 31, 299–305. [Google Scholar] [CrossRef]
  160. Motloch, L.J.; Cacheux, M.; Ishikawa, K.; Xie, C.; Hu, J.; Aguero, J.; Fish, K.M.; Hajjar, R.J.; Akar, F.G. Primary Effect of SERCA 2a Gene Transfer on Conduction Reserve in Chronic Myocardial Infarction. J. Am. Heart Assoc. 2018, 7, e009598. [Google Scholar] [CrossRef]
  161. Strauss, B.; Sassi, Y.; Bueno-Beti, C.; Ilkan, Z.; Raad, N.; Cacheux, M.; Bisserier, M.; Turnbull, I.C.; Kohlbrenner, E.; Hajjar, R.J.; et al. Intra-tracheal gene delivery of aerosolized SERCA2a to the lung suppresses ventricular arrhythmias in a model of pulmonary arterial hypertension. J. Mol. Cell. Cardiol. 2019, 127, 20–30. [Google Scholar] [CrossRef] [PubMed]
  162. Motloch, L.J.; Ishikawa, K.; Xie, C.; Hu, J.; Aguero, J.; Fish, K.M.; Hajjar, R.J.; Akar, F.G. Increased afterload following myocardial infarction promotes conduction-dependent arrhythmias that are unmasked by hypokalemia. JACC Basic Transl. Sci. 2017, 2, 258–269. [Google Scholar] [CrossRef] [PubMed]
  163. Xu, X.; Yano, M.; Uchinoumi, H.; Hino, A.; Suetomi, T.; Ono, M.; Tateishi, H.; Oda, T.; Okuda, S.; Doi, M.; et al. Defective calmodulin binding to the cardiac ryanodine receptor plays a key role in CPVT-associated channel dysfunction. Biochem. Biophys. Res. Commun. 2010, 394, 660–666. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Ono, M.; Yano, M.; Hino, A.; Suetomi, T.; Xu, X.; Susa, T.; Uchinoumi, H.; Tateishi, H.; Oda, T.; Okuda, S.; et al. Dissociation of calmodulin from cardiac ryanodine receptor causes aberrant Ca2+ release in heart failure. Cardiovasc. Res. 2010, 87, 609–617. [Google Scholar] [CrossRef] [PubMed]
  165. Kushnir, A.; Santulli, G.; Reiken, S.R.; Coromilas, E.; Godfrey, S.J.; Brunjes, D.L.; Colombo, P.C.; Yuzefpolskaya, M.; Sokol, S.I.; Kitsis, R.N.; et al. Ryanodine Receptor Calcium Leak in Circulating B-Lymphocytes as a Biomarker in Heart Failure. Circulation 2018, 138, 1144–1154. [Google Scholar] [CrossRef] [PubMed]
  166. Hasenfuss, G.; Teerlink, J.R. Cardiac inotropes: Current agents and future directions. Eur. Heart J. 2011, 32, 1838–1845. [Google Scholar] [CrossRef]
  167. Bongianino, R.; Denegri, M.; Mazzanti, A.; Lodola, F.; Vollero, A.; Boncompagni, S.; Fasciano, S.; Rizzo, G.; Mangione, D.; Barbaro, S.; et al. Allele-Specific Silencing of Mutant mRNA Rescues Ultrastructural and Arrhythmic Phenotype in Mice Carriers of the R4496C Mutation in the Ryanodine Receptor Gene (RYR2). Circ. Res. 2017, 121, 525–536. [Google Scholar] [CrossRef]
  168. Pan, X.; Philippen, L.; Lahiri, S.K.; Lee, C.; Park, S.H.; Word, T.A.; Li, N.; Jarrett, K.E.; Gupta, R.; Reynolds, J.O.; et al. In Vivo Ryr2 Editing Corrects Catecholaminergic Polymorphic Ventricular Tachycardia. Circ. Res. 2018, 123, 953–963. [Google Scholar] [CrossRef]
  169. Denegri, M.; Bongianino, R.; Lodola, F.; Boncompagni, S.; De Giusti, V.C.; Avelino-Cruz, J.E.; Liu, N.; Persampieri, S.; Curcio, A.; Esposito, F.; et al. Single delivery of an adeno-associated viral construct to transfer the CASQ2 gene to knock-in mice affected by catecholaminergic polymorphic ventricular tachycardia is able to cure the disease from birth to advanced age. Circulation 2014, 129, 2673–2681. [Google Scholar] [CrossRef]
  170. Lodola, F.; Morone, D.; Denegri, M.; Bongianino, R.; Nakahama, H.; Rutigliano, L.; Gosetti, R.; Rizzo, G.; Vollero, A.; Buonocore, M.; et al. Adeno-associated virus-mediated CASQ2 delivery rescues phenotypic alterations in a patient-specific model of recessive catecholaminergic polymorphic ventricular tachycardia. Cell Death Dis. 2016, 7, e2393. [Google Scholar] [CrossRef]
  171. Kurtzwald-Josefson, E.; Yadin, D.; Harun-Khun, S.; Waldman, M.; Aravot, D.; Shainberg, A.; Eldar, M.; Hochhauser, E.; Arad, M. Viral delivered gene therapy to treat catecholaminergic polymorphic ventricular tachycardia (CPVT2) in mouse models. Heart Rhythm 2017, 14, 1053–1060. [Google Scholar] [CrossRef] [PubMed]
  172. Liu, B.; Walton, S.D.; Ho, H.-T.; Belevych, A.E.; Tikunova, S.B.; Bonilla, I.; Shettigar, V.; Knollmann, B.C.; Priori, S.G.; Volpe, P.; et al. Gene Transfer of Engineered Calmodulin Alleviates Ventricular Arrhythmias in a Calsequestrin-Associated Mouse Model of Catecholaminergic Polymorphic Ventricular Tachycardia. J. Am. Heart Assoc. 2018, 7, e008155. [Google Scholar] [CrossRef] [PubMed]
  173. Bezzerides, V.J.; Caballero, A.; Wang, S.; Ai, Y.; Hylind, R.J.; Lu, F.; Heims-Waldron, D.A.; Chambers, K.D.; Zhang, D.; Abrams, D.J.; et al. Gene Therapy for Catecholaminergic Polymorphic Ventricular Tachycardia by Inhibition of Ca2+/Calmodulin-Dependent Kinase II. Circulation 2019, 140, 405–419. [Google Scholar] [CrossRef] [PubMed]
  174. Park, S.-J.; Zhang, D.; Qi, Y.; Li, Y.; Lee, K.Y.; Bezzerides, V.J.; Yang, P.; Xia, S.; Kim, S.L.; Liu, X.; et al. Insights Into the Pathogenesis of Catecholaminergic Polymorphic Ventricular Tachycardia From Engineered Human Heart Tissue. Circulation 2019, 140, 390–404. [Google Scholar] [CrossRef]
  175. Di Pasquale, E.; Lodola, F.; Miragoli, M.; Denegri, M.; Avelino-Cruz, J.E.; Buonocore, M.; Nakahama, H.; Portararo, P.; Bloise, R.; Napolitano, C.; et al. CaMKII inhibition rectifies arrhythmic phenotype in a patient-specific model of catecholaminergic polymorphic ventricular tachycardia. Cell Death Dis. 2013, 4, e843. [Google Scholar] [CrossRef] [PubMed]
  176. Knollmann, B.C.; Chopra, N.; Hlaing, T.; Akin, B.; Yang, T.; Ettensohn, K.; Knollmann, B.E.C.; Horton, K.D.; Weissman, N.J.; Holinstat, I.; et al. Casq2 deletion causes sarcoplasmic reticulum volume increase, premature Ca2+ release, and catecholaminergic polymorphic ventricular tachycardia. J. Clin. Investig. 2006, 116, 2510–2520. [Google Scholar] [CrossRef] [PubMed]
  177. Viatchenko-Karpinski, S.; Terentyev, D.; Györke, I.; Terentyeva, R.; Volpe, P.; Priori, S.G.; Napolitano, C.; Nori, A.; Williams, S.C.; Györke, S. Abnormal calcium signaling and sudden cardiac death associated with mutation of calsequestrin. Circ. Res. 2004, 94, 471–477. [Google Scholar] [CrossRef]
  178. Faggioni, M.; Kryshtal, D.O.; Knollmann, B.C. Calsequestrin mutations and catecholaminergic polymorphic ventricular tachycardia. Pediatr. Cardiol. 2012, 33, 959–967. [Google Scholar] [CrossRef]
  179. Adachi-Akahane, S.; Lu, L.; Li, Z.; Frank, J.S.; Philipson, K.D.; Morad, M. Calcium Signaling in Transgenic Mice Overexpressing Cardiac Na+-Ca2+ Exchanger. J. Gen. Physiol. 1997, 109, 717–729. [Google Scholar] [CrossRef]
  180. Yao, A.; Su, Z.; Nonaka, A.; Zubair, I.; Lu, L.; Philipson, K.D.; Bridge, J.H.; Barry, W.H. Effects of overexpression of the Na+-Ca2+ exchanger on [Ca2+] i transients in murine ventricular myocytes. Circ. Res. 1998, 82, 657–665. [Google Scholar] [CrossRef]
  181. Terracciano, C.M.; Souza, A.I.; Philipson, K.D.; MacLeod, K.T. Na+-Ca2+ exchange and sarcoplasmic reticular Ca2+ regulation in ventricular myocytes from transgenic mice overexpressing the Na+-Ca2+ exchanger. J. Physiol. 1998, 512 Pt 3, 651–667. [Google Scholar] [CrossRef] [PubMed]
  182. Reuter, H.; Han, T.; Motter, C.; Philipson, K.D.; Goldhaber, J.I. Mice overexpressing the cardiac sodium-calcium exchanger: Defects in excitation–contraction coupling. J. Physiol. 2004, 554, 779–789. [Google Scholar] [CrossRef] [PubMed]
  183. Wakimoto, K.; Kobayashi, K.; Kuro, O.M.; Yao, A.; Iwamoto, T.; Yanaka, N.; Kita, S.; Nishida, A.; Azuma, S.; Toyoda, Y.; et al. Targeted disruption of Na+/Ca2+ exchanger gene leads to cardiomyocyte apoptosis and defects in heartbeat. J. Biol. Chem. 2000, 275, 36991–36998. [Google Scholar] [CrossRef] [PubMed]
  184. Schillinger, W.; Janssen, P.M.; Emami, S.; Henderson, S.A.; Ross, R.S.; Teucher, N.; Zeitz, O.; Philipson, K.D.; Prestle, J.; Hasenfuss, G. Impaired contractile performance of cultured rabbit ventricular myocytes after adenoviral gene transfer of Na+-Ca2+ exchanger. Circ. Res. 2000, 87, 581–587. [Google Scholar] [CrossRef] [PubMed]
  185. Ranu, H.K.; Terracciano, C.M.N.; Davia, K.; Bernobich, E.; Chaudhri, B.; Robinson, S.E.; Bin Kang, Z.; Hajjar, R.J.; MacLeod, K.T.; Harding, S.E. Effects of Na+/Ca2+-exchanger overexpression on excitation-contraction coupling in adult rabbit ventricular myocytes. J. Mol. Cell. Cardiol. 2002, 34, 389–400. [Google Scholar] [CrossRef]
  186. Tadros, G.M.; Zhang, X.-Q.; Song, J.; Carl, L.L.; Rothblum, L.I.; Tian, Q.; Dunn, J.; Lytton, J.; Cheung, J.Y. Effects of Na+/Ca2+ exchanger downregulation on contractility and [Ca2+]i transients in adult rat myocytes. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H1616–H1626. [Google Scholar] [CrossRef]
  187. Schillinger, W.; Ohler, A.; Emami, S.; Müller, F.; Christians, C.; Janssen, P.M.L.; Kögler, H.; Teucher, N.; Pieske, B.; Seidler, T.; et al. The functional effect of adenoviral Na+/Ca2+ exchanger overexpression in rabbit myocytes depends on the activity of the Na+/K+-ATPase. Cardiovasc. Res. 2003, 57, 996–1003. [Google Scholar] [CrossRef]
  188. Bölck, B.; Münch, G.; Mackenstein, P.; Hellmich, M.; Hirsch, I.; Reuter, H.; Hattebuhr, N.; Weig, H.-J.; Ungerer, M.; Brixius, K.; et al. Na+/Ca2+ exchanger overexpression impairs frequency—And ouabain—Dependent cell shortening in adult rat cardiomyocytes. Am. J. Physiol. Heart Circ. Physiol. 2004, 287, H1435–H1445. [Google Scholar] [CrossRef]
  189. Shamraj, O.I.; Grupp, I.L.; Grupp, G.; Melvin, D.; Gradoux, N.; Kremers, W.; Lingrel, J.B.; De Pover, A. Characterisation of Na/K-ATPase, its isoforms, and the inotropic response to ouabain in isolated failing human hearts. Cardiovasc. Res. 1993, 27, 2229–2237. [Google Scholar] [CrossRef] [Green Version]
  190. Nørgaard, A.; Bagger, J.P.; Bjerregaard, P.; Baandrup, U.; Kjeldsen, K.; Thomsen, P.E. Relation of left ventricular function and Na,K-pump concentration in suspected idiopathic dilated cardiomyopathy. Am. J. Cardiol. 1988, 61, 1312–1315. [Google Scholar] [CrossRef]
  191. Schwinger, R.H.; Wang, J.; Frank, K.; Müller-Ehmsen, J.; Brixius, K.; McDonough, A.A.; Erdmann, E. Reduced sodium pump alpha1, alpha3, and beta1-isoform protein levels and Na+, K+-ATPase activity but unchanged Na+-Ca2+ exchanger protein levels in human heart failure. Circulation 1999, 99, 2105–2112. [Google Scholar] [CrossRef] [PubMed]
  192. Gammage, P.A.; Viscomi, C.; Simard, M.-L.; Costa, A.S.H.; Gaude, E.; Powell, C.A.; Van Haute, L.; McCann, B.J.; Rebelo-Guiomar, P.; Cerutti, R.; et al. Genome editing in mitochondria corrects a pathogenic mtDNA mutation in vivo. Nat. Med. 2018, 24, 1691–1695. [Google Scholar] [CrossRef] [PubMed]
  193. Luongo, T.S.; Lambert, J.P.; Gross, P.; Nwokedi, M.; Lombardi, A.A.; Shanmughapriya, S.; Carpenter, A.C.; Kolmetzky, D.; Gao, E.; van Berlo, J.H.; et al. The mitochondrial Na+/Ca2+ exchanger is essential for Ca2+ homeostasis and viability. Nature 2017, 545, 93–97. [Google Scholar] [CrossRef] [PubMed]
  194. Wu, Y.; Rasmussen, T.P.; Koval, O.M.; Joiner, M.-L.A.; Hall, D.D.; Chen, B.; Luczak, E.D.; Wang, Q.; Rokita, A.G.; Wehrens, X.H.T.; et al. The mitochondrial uniporter controls fight or flight heart rate increases. Nat. Commun. 2015, 6, 6081. [Google Scholar] [CrossRef]
  195. Oropeza-Almazán, Y.; Vázquez-Garza, E.; Chapoy-Villanueva, H.; Torre-Amione, G.; García-Rivas, G. Small Interfering RNA Targeting Mitochondrial Calcium Uniporter Improves Cardiomyocyte Cell Viability in Hypoxia/Reoxygenation Injury by Reducing Calcium Overload. Oxid. Med. Cell. Longev. 2017, 2017, 5750897. [Google Scholar] [CrossRef]
  196. Suarez, J.; Cividini, F.; Scott, B.T.; Lehmann, K.; Diaz-Juarez, J.; Diemer, T.; Dai, A.; Suarez, J.A.; Jain, M.; Dillmann, W.H. Restoring mitochondrial calcium uniporter expression in diabetic mouse heart improves mitochondrial calcium handling and cardiac function. J. Biol. Chem. 2018, 293, 8182–8195. [Google Scholar] [CrossRef] [Green Version]
  197. Flarsheim, C.E.; Grupp, I.L.; Matlib, M.A. Mitochondrial dysfunction accompanies diastolic dysfunction in diabetic rat heart. Am. J. Physiol. 1996, 271, H192–H202. [Google Scholar] [CrossRef]
  198. Tanaka, Y.; Konno, N.; Kako, K.J. Mitochondrial dysfunction observed in situ in cardiomyocytes of rats in experimental diabetes. Cardiovasc. Res. 1992, 26, 409–414. [Google Scholar] [CrossRef]
  199. 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. [Google Scholar] [CrossRef]
  200. Kho, C.; Lee, A.; Jeong, D.; Oh, J.G.; Chaanine, A.H.; Kizana, E.; Park, W.J.; Hajjar, R.J. SUMO1-dependent modulation of SERCA2a in heart failure. Nature 2011, 477, 601–605. [Google Scholar] [CrossRef]
  201. Tilemann, L.; Lee, A.; Ishikawa, K.; Aguero, J.; Rapti, K.; Santos-Gallego, C.; Kohlbrenner, E.; Fish, K.M.; Kho, C.; Hajjar, R.J. SUMO-1 gene transfer improves cardiac function in a large-animal model of heart failure. Sci. Transl. Med. 2013, 5. [Google Scholar] [CrossRef] [PubMed]
  202. Lee, A.; Jeong, D.; Mitsuyama, S.; Oh, J.G.; Liang, L.; Ikeda, Y.; Sadoshima, J.; Hajjar, R.J.; Kho, C. The role of SUMO-1 in cardiac oxidative stress and hypertrophy. Antioxid. Redox Signal. 2014, 21, 1986–2001. [Google Scholar] [CrossRef] [PubMed]
  203. Rohde, D.; Ritterhoff, J.; Voelkers, M.; Katus, H.A.; Parker, T.G.; Most, P. S100A1: A multifaceted therapeutic target in cardiovascular disease. J. Cardiovasc. Transl. Res. 2010, 3, 525–537. [Google Scholar] [CrossRef] [PubMed]
  204. Most, P.; Pleger, S.T.; Völkers, M.; Heidt, B.; Boerries, M.; Weichenhan, D.; Löffler, E.; Janssen, P.M.L.; Eckhart, A.D.; Martini, J.; et al. Cardiac adenoviral S100A1 gene delivery rescues failing myocardium. J. Clin. Investig. 2004, 114, 1550–1563. [Google Scholar] [CrossRef] [PubMed]
  205. Pleger, S.T.; Remppis, A.; Heidt, B.; Völkers, M.; Chuprun, J.K.; Kuhn, M.; Zhou, R.-H.; Gao, E.; Szabo, G.; Weichenhan, D.; et al. S100A1 gene therapy preserves in vivo cardiac function after myocardial infarction. Mol. Ther. 2005, 12, 1120–1129. [Google Scholar] [CrossRef]
  206. Pleger, S.T.; Most, P.; Boucher, M.; Soltys, S.; Chuprun, J.K.; Pleger, W.; Gao, E.; Dasgupta, A.; Rengo, G.; Remppis, A.; et al. Stable myocardial-specific AAV6-S100A1 gene therapy results in chronic functional heart failure rescue. Circulation 2007, 115, 2506–2515. [Google Scholar] [CrossRef]
  207. Pleger, S.T.; Shan, C.; Ksienzyk, J.; Bekeredjian, R.; Boekstegers, P.; Hinkel, R.; Schinkel, S.; Leuchs, B.; Ludwig, J.; Qiu, G.; et al. Cardiac AAV9-S100A1 gene therapy rescues post-ischemic heart failure in a preclinical large animal model. Sci. Transl. Med. 2011, 3. [Google Scholar] [CrossRef]
  208. Liou, J.; Kim, M.L.; Heo, W.D.; Jones, J.T.; Myers, J.W.; Ferrell, J.E.; Meyer, T. STIM is a Ca2+ sensor essential for Ca2+-store-depletion-triggered Ca2+ influx. Curr. Biol. CB 2005, 15, 1235–1241. [Google Scholar] [CrossRef]
  209. Prakriya, M.; Feske, S.; Gwack, Y.; Srikanth, S.; Rao, A.; Hogan, P.G. Orai1 is an essential pore subunit of the CRAC channel. Nature 2006, 443, 230–233. [Google Scholar] [CrossRef]
  210. Bergmeier, W.; Weidinger, C.; Zee, I.; Feske, S. Emerging roles of store-operated Ca2+ entry through STIM and ORAI proteins in immunity, hemostasis and cancer. Channels 2013, 7, 379–391. [Google Scholar] [CrossRef]
  211. Ong, H.L.; Liu, X.; Sharma, A.; Hegde, R.S.; Ambudkar, I.S. Intracellular Ca2+ release via the ER translocon activates store-operated calcium entry. Pflugers Arch. 2007, 453, 797–808. [Google Scholar] [CrossRef] [PubMed]
  212. Moccia, F.; Zuccolo, E.; Soda, T.; Tanzi, F.; Guerra, G.; Mapelli, L.; Lodola, F.; D’Angelo, E. Stim and Orai proteins in neuronal Ca2+ signaling and excitability. Front. Cell. Neurosci. 2015, 9, 153. [Google Scholar] [CrossRef] [PubMed]
  213. Maus, M.; Cuk, M.; Patel, B.; Lian, J.; Ouimet, M.; Kaufmann, U.; Yang, J.; Horvath, R.; Hornig-Do, H.-T.; Chrzanowska-Lightowlers, Z.M.; et al. Store-Operated Ca2+ Entry Controls Induction of Lipolysis and the Transcriptional Reprogramming to Lipid Metabolism. Cell Metab. 2017, 25, 698–712. [Google Scholar] [CrossRef] [PubMed]
  214. Lodola, F.; Laforenza, U.; Bonetti, E.; Lim, D.; Dragoni, S.; Bottino, C.; Ong, H.L.; Guerra, G.; Ganini, C.; Massa, M.; et al. Store-operated Ca2+ entry is remodelled and controls in vitro angiogenesis in endothelial progenitor cells isolated from tumoral patients. PLoS ONE 2012, 7, e42541. [Google Scholar] [CrossRef]
  215. Dragoni, S.; Laforenza, U.; Bonetti, E.; Reforgiato, M.; Poletto, V.; Lodola, F.; Bottino, C.; Guido, D.; Rappa, A.; Pareek, S.; et al. Enhanced expression of Stim, Orai, and TRPC transcripts and proteins in endothelial progenitor cells isolated from patients with primary myelofibrosis. PLoS ONE 2014, 9, e91099. [Google Scholar] [CrossRef]
  216. Zuccolo, E.; Laforenza, U.; Ferulli, F.; Pellavio, G.; Scarpellino, G.; Tanzi, M.; Turin, I.; Faris, P.; Lucariello, A.; Maestri, M.; et al. Stim and Orai mediate constitutive Ca2+ entry and control endoplasmic reticulum Ca2+ refilling in primary cultures of colorectal carcinoma cells. Oncotarget 2018, 9, 31098–31119. [Google Scholar] [CrossRef]
  217. Xia, J.; Wang, H.; Huang, H.; Sun, L.; Dong, S.; Huang, N.; Shi, M.; Bin, J.; Liao, Y.; Liao, W. Elevated Orai1 and STIM1 expressions upregulate MACC1 expression to promote tumor cell proliferation, metabolism, migration, and invasion in human gastric cancer. Cancer Lett. 2016, 381, 31–40. [Google Scholar] [CrossRef]
  218. Moccia, F.; Berra-Romani, R.; Rosti, V. Manipulating Intracellular Ca2+ Signals to Stimulate Therapeutic Angiogenesis in Cardiovascular Disorders. Curr. Pharm. Biotechnol. 2018, 19, 686–699. [Google Scholar] [CrossRef]
  219. Collins, H.E.; Pat, B.M.; Zou, L.; Litovsky, S.H.; Wende, A.R.; Young, M.E.; Chatham, J.C. Novel role of the ER/SR Ca2+ sensor STIM1 in the regulation of cardiac metabolism. Am. J. Physiol. Heart Circ. Physiol. 2019, 316, H1014–H1026. [Google Scholar] [CrossRef]
  220. Bartoli, F.; Sabourin, J. Cardiac Remodeling and Disease: Current Understanding of STIM1/Orai1-Mediated Store-Operated Ca2+ Entry in Cardiac Function and Pathology. Adv. Exp. Med. Biol. 2017, 993, 523–534. [Google Scholar]
  221. Ji, Y.; Guo, X.; Zhang, Z.; Huang, Z.; Zhu, J.; Chen, Q.-H.; Gui, L. CaMKIIδ meditates phenylephrine induced cardiomyocyte hypertrophy through store-operated Ca2+ entry. Cardiovasc. Pathol. Off. J. Soc. Cardiovasc. Pathol. 2017, 27, 9–17. [Google Scholar] [CrossRef] [PubMed]
  222. Zhang, B.; Jiang, J.; Yue, Z.; Liu, S.; Ma, Y.; Yu, N.; Gao, Y.; Sun, S.; Chen, S.; Liu, P. Store-Operated Ca2+ Entry (SOCE) contributes to angiotensin II-induced cardiac fibrosis in cardiac fibroblasts. J. Pharmacol. Sci. 2016, 132, 171–180. [Google Scholar] [CrossRef] [PubMed]
  223. Zhao, G.; Li, T.; Brochet, D.X.P.; Rosenberg, P.B.; Lederer, W.J. STIM1 enhances SR Ca2+ content through binding phospholamban in rat ventricular myocytes. Proc. Natl. Acad. Sci. USA 2015, 112, E4792–E4801. [Google Scholar] [CrossRef] [PubMed]
  224. Dai, F.; Zhang, Y.; Wang, Q.; Li, D.; Yang, Y.; Ma, S.; Yang, D. Overexpression of SARAF Ameliorates Pressure Overload-Induced Cardiac Hypertrophy Through Suppressing STIM1-Orai1 in Mice. Cell. Physiol. Biochem. 2018, 47, 817–826. [Google Scholar] [CrossRef]
  225. Sabourin, J.; Boet, A.; Rucker-Martin, C.; Lambert, M.; Gomez, A.-M.; Benitah, J.-P.; Perros, F.; Humbert, M.; Antigny, F. Ca2+ handling remodeling and STIM1L/Orai1/TRPC1/TRPC4 upregulation in monocrotaline-induced right ventricular hypertrophy. J. Mol. Cell. Cardiol. 2018, 118, 208–224. [Google Scholar] [CrossRef]
  226. Troupes, C.D.; Wallner, M.; Borghetti, G.; Zhang, C.; Mohsin, S.; von Lewinski, D.; Berretta, R.M.; Kubo, H.; Chen, X.; Soboloff, J.; et al. Role of STIM1 (Stromal Interaction Molecule 1) in Hypertrophy-Related Contractile Dysfunction. Circ. Res. 2017, 121, 125–136. [Google Scholar] [CrossRef]
  227. Zhang, H.; Sun, A.Y.; Kim, J.J.; Graham, V.; Finch, E.A.; Nepliouev, I.; Zhao, G.; Li, T.; Lederer, W.J.; Stiber, J.A.; et al. STIM1-Ca2+ signaling modulates automaticity of the mouse sinoatrial node. Proc. Natl. Acad. Sci. USA 2015, 112, E5618–E5627. [Google Scholar] [CrossRef]
  228. Li, P.; Bian, X.-Y.; Chen, Q.; Yao, X.-F.; Wang, X.-D.; Zhang, W.-C.; Tao, Y.-J.; Jin, R.; Zhang, L. Blocking of stromal interaction molecule 1 expression influence cell proliferation and promote cell apoptosis in vitro and inhibit tumor growth in vivo in head and neck squamous cell carcinoma. PLoS ONE 2017, 12, e0177484. [Google Scholar] [CrossRef]
  229. Schmidt, S.; Liu, G.; Liu, G.; Yang, W.; Honisch, S.; Pantelakos, S.; Stournaras, C.; Hönig, A.; Lang, F. Enhanced Orai1 and STIM1 expression as well as store operated Ca2+ entry in therapy resistant ovary carcinoma cells. Oncotarget 2014, 5, 4799–4810. [Google Scholar] [CrossRef]
  230. Moccia, F.; Lodola, F.; Dragoni, S.; Bonetti, E.; Bottino, C.; Guerra, G.; Laforenza, U.; Rosti, V.; Tanzi, F. Ca2+ signalling in endothelial progenitor cells: A novel means to improve cell-based therapy and impair tumour vascularisation. Curr. Vasc. Pharmacol. 2014, 12, 87–105. [Google Scholar] [CrossRef]
  231. del Monte, F.; Harding, S.E.; Dec, G.W.; Gwathmey, J.K.; Hajjar, R.J. Targeting phospholamban by gene transfer in human heart failure. Circulation 2002, 105, 904–907. [Google Scholar] [CrossRef] [PubMed]
  232. Kaye, D.M.; Preovolos, A.; Marshall, T.; Byrne, M.; Hoshijima, M.; Hajjar, R.; Mariani, J.A.; Pepe, S.; Chien, K.R.; Power, J.M. Percutaneous cardiac recirculation-mediated gene transfer of an inhibitory phospholamban peptide reverses advanced heart failure in large animals. J. Am. Coll. Cardiol. 2007, 50, 253–260. [Google Scholar] [CrossRef] [PubMed]
  233. Suckau, L.; Fechner, H.; Chemaly, E.; Krohn, S.; Hadri, L.; Kockskämper, J.; Westermann, D.; Bisping, E.; Ly, H.; Wang, X.; et al. Long-term cardiac-targeted RNA interference for the treatment of heart failure restores cardiac function and reduces pathological hypertrophy. Circulation 2009, 119, 1241–1252. [Google Scholar] [CrossRef] [PubMed]
  234. Ziolo, M.T.; Martin, J.L.; Bossuyt, J.; Bers, D.M.; Pogwizd, S.M. Adenoviral gene transfer of mutant phospholamban rescues contractile dysfunction in failing rabbit myocytes with relatively preserved SERCA function. Circ. Res. 2005, 96, 815–817. [Google Scholar] [CrossRef] [PubMed]
  235. Reynolds, J.O.; Quick, A.P.; Wang, Q.; Beavers, D.L.; Philippen, L.E.; Showell, J.; Barreto-Torres, G.; Thuerauf, D.J.; Doroudgar, S.; Glembotski, C.C.; et al. Junctophilin-2 gene therapy rescues heart failure by normalizing RyR2-mediated Ca2+ release. Int. J. Cardiol. 2016, 225, 371–380. [Google Scholar] [CrossRef] [PubMed]
  236. Beavers, D.L.; Landstrom, A.P.; Chiang, D.Y.; Wehrens, X.H.T. Emerging roles of junctophilin-2 in the heart and implications for cardiac diseases. Cardiovasc. Res. 2014, 103, 198–205. [Google Scholar] [CrossRef]
  237. Guo, A.; Zhang, X.; Iyer, V.R.; Chen, B.; Zhang, C.; Kutschke, W.J.; Weiss, R.M.; Franzini-Armstrong, C.; Song, L.-S. Overexpression of junctophilin-2 does not enhance baseline function but attenuates heart failure development after cardiac stress. Proc. Natl. Acad. Sci. USA. 2014, 111, 12240–12245. [Google Scholar] [CrossRef] [Green Version]
  238. Watanabe, S.; Ishikawa, K.; Fish, K.; Oh, J.G.; Motloch, L.J.; Kohlbrenner, E.; Lee, P.; Xie, C.; Lee, A.; Liang, L.; et al. Protein Phosphatase Inhibitor-1 Gene Therapy in a Swine Model of Nonischemic Heart Failure. J. Am. Coll. Cardiol. 2017, 70, 1744–1756. [Google Scholar] [CrossRef]
Figure 1. Calcium handling in cardiomyocytes illustrating the most important calcium-related organelles and channels. Disturbed ion channel functions or calcium mishandling results in an increased cytosolic calcium concentration. This further promotes the generation of early and delayed afterdepolarizations, leading to Ca2+-induced ventricular tachycardia. Ca2+: calcium; ICa: calcium current; MCU: mitochondrial calcium uniporter; MF: myofilament; NCLX: mitochondrial potassium-calcium exchanger; NCX: sodium-calcium exchanger; RyR2: ryanodine receptor type 2; VES: ventricular extrasystoles; VT: ventricular tachycardia.
Figure 1. Calcium handling in cardiomyocytes illustrating the most important calcium-related organelles and channels. Disturbed ion channel functions or calcium mishandling results in an increased cytosolic calcium concentration. This further promotes the generation of early and delayed afterdepolarizations, leading to Ca2+-induced ventricular tachycardia. Ca2+: calcium; ICa: calcium current; MCU: mitochondrial calcium uniporter; MF: myofilament; NCLX: mitochondrial potassium-calcium exchanger; NCX: sodium-calcium exchanger; RyR2: ryanodine receptor type 2; VES: ventricular extrasystoles; VT: ventricular tachycardia.
Ijms 20 05304 g001
Table 1. Advantages and limitations of different gene delivery vectors and CRISPR/Cas9 system.
Table 1. Advantages and limitations of different gene delivery vectors and CRISPR/Cas9 system.
MethodAdvantagesLimitationsReferences
LentivirusBroad host range
Infection of dividing and non-dividing cells
Low cytotoxicity
Long-term expression (integration into genome)
Insert capacity: 8 kb
No specific tropism for CMs
(requires direct injection into the heart)
Risk of insertional mutagenesis (integration into genome)
[125,126,127,128,129]
ADBroad host range
High level of gene expression
No host genome integration
Insert capacity: <35 kb
Short-term expression
Strong immunogenicity
No specific tropism for CMs
[127,129,130,131]
AAVRelatively broad host range
Low pathogenicity and toxicity
Infection of dividing and non-dividing cells
Long-term expression
Serotype modulation for organ specificity
Difficulties in high transgene expression
Delayed expression
Presence of NAbs
Small insert capacity of <5 kb
[127,129,132,133,134,135,136]
CRISPR/Cas9Targeting specific DNA sequences
Any organism
Simple and precise (compared to gene targeting)
Inactivation, integration and allele substitution possible
Reactivation of non-dividing cells
Low immunogenicity
Difficulties in off-target effects (nonspecific and mismatched genetic modifications)
Difficulties in delivery of large Cas9 sequences
[137,138,139,140,141]
AD: adenovirus; AAV: adeno-associated virus; CRISPR/Cas9: clustered regularly interspaced short palindromic repeats with caspase 9; CMs: cardiomyocytes; DNA: deoxyribonucleic acid; kb: kilobases; NAbs: neutralizing antibodies.
Table 2. Gene therapies for ventricular tachycardia targeting L-type Ca2+ channel (LTCC).
Table 2. Gene therapies for ventricular tachycardia targeting L-type Ca2+ channel (LTCC).
Author (year)VectorDelivery TechniqueGenetic InformationSpecies/ModelOutcomes
Murata et al. (2004) [143]ADInjection into LV cavityMutant Ras-related G-protein Gem W296GGuinea pig/wt↓ ICa-L in CMs
↓ QT in vivo
Cingolani et al. (2007) [144]LentivirusInjection into LV cavityHairpin RNA for β2Rat/aortic-banded model of LV hypertrophy↓ ICa-L in CMs
Subramanyam et al. (2013) [145]ADIn vitroSplit-intein-tagged α1C-fragmentsRat/wt↑ Ca2+ transients
β-adrenergic regulation
The origin of the genetic material used in the studies is indicated, if mentioned in the publication. Injection into the aortic root or the left ventricle (LV) cavity was performed during transient cross-clamping of the great vessels. Intramyocardial injection was performed after thoracotomy. ↓: decrease; ↑: support; AD: adenovirus; Ca2+: calcium; CMs: cardiomyocytes; ICa-L: L-type calcium current; LV: left ventricle; RNA: ribonucleic acid; VM: ventricular myocard; wt: wild-type.
Table 3. Gene therapies to restore cardiac electrophysiology and to prevent ventricular tachyarrhythmias targeting sarco/endo plasmic reticulum calcium ATPase (SERCA2a).
Table 3. Gene therapies to restore cardiac electrophysiology and to prevent ventricular tachyarrhythmias targeting sarco/endo plasmic reticulum calcium ATPase (SERCA2a).
Author (year)VectorDelivery TechniqueSpecies/ModelOutcomes
Giordano et al. (1997) [157]ADIn vitroRat—
↓ SERCA2a expression
↑ SERCA2a expression
↓ Ca2+ transients
↑ SR Ca2+ uptake
Hajjar et al. (1997) [158]ADIn vitroRat—
wt
↑ peak Ca2+ release
↓ resting Ca2+ levels
Terracciano et al. (2002) [159]ADIn vitroRabbit—
wt
↓ APD
↑ SR Ca2+ content
del Monte et al. (2004) [150]ADIntramyocardial injectionRat—
wt
↓ VT after I/R
Prunier et al. (2008) [151]ADAnterograde coronary injectionSwine—
wt
↓ VT after I/R
Cutler et al. (2009) [152]ADInjection into aortic rootGuinea pig—
wt
↓ APD alternans in vitro and ex vivo
↓ VT ex vivo
Lyon et al. (2011) [153]AD
AAV9
Intramyocardial (AV),
or tail vein (AAV9)
Rat—
HF
↓ VT ex vivo
↓ spontaneous and isoproterenol triggered VT in vivo
Cutler et al. (2012) [154]AAV9Injection into aortic rootGuinea pig—
HF
↓ APD alternans
↓ VT ex vivo
Motloch et al. (2018) [160]AAV1Intracoronary injectionSwine—
MI
↓ QRS duration in vivo
↓ VT in vivo and ex vivo
Strauss et al. (2019) [161]AAV1AerosolizedRat—
PAH
↓ VT in vivo
↓ APD duration
Reversed spatial APD heterogeneity
↑ Electrophysiological remodeling
The origin of the genetic material used in the studies is indicated, if indicated in the publication. Injection into the aortic root or the LV cavity was performed during transient cross-clamping of the great vessels. Intramyocardial injection was performed after thoracotomy. ↓: decrease; ↑: improvement; AD: adenovirus; AAV1: adenovirus-associated virus serotype 1; AAV9: AAV serotype 9; APD: action potential duration; EF: ejection fraction; HF: heart failure; I/R: ischemia reperfusion; LV: left ventricle; MI: myocardial infarction; PAH: pulmonary arterial hypertension; SERCA2a: sarcoplasmic reticulum Ca2+ ATPase 2a; UV: upstroke velocity; VT: ventricular tachycardia; wt: wild-type.
Table 4. Gene therapies for ventricular tachycardia targeting ryanodine receptor 2 (RyR2) and genes of RyR2-mediating proteins.
Table 4. Gene therapies for ventricular tachycardia targeting ryanodine receptor 2 (RyR2) and genes of RyR2-mediating proteins.
Author (year)VectorDelivery TechniqueGenetic InformationSpecies/ModelOutcomes
Bongianino et al. (2017) [167]AAV9Intraperitoneal injectionmiRyR2-U10Mouse/wt↓ DADs
↓ VT in vivo
Pan et al. (2018) [168]AAV9Subcutaneous injectionRyR2Mouse/CPVT (R176Q/+)↓ arrhythmias in vivo
Denegri et al. (2014) [169]AAV9Intraperitoneal injectionCASQ2Mouse/CPVT (R33Q)↓ VT in vivo
Lodola et al. (2016) [170]AAV9In vitroCASQ2Human/CPVT; iPSCs (CASQ2-G112+5X)↓ DADs
↑ Ca2+ transient amplitude and duration of Ca2+ sparks
Kurtzwald-Josefson et al. (2017) [171]AAV9Intraperitoneal injectionCASQ2Mouse/CPVT (CASQ2D307H or CASQ2Δ/Δ)↓ VT in vivo
Liu et al. (2018) [172]AAV9Intra-thoracic cavity injectionCaMMouse/CPVT (R33Q)↑ Ca2+ handling
↓ VT in vivo
Bezzerides et al. (2019) [173]AAV9Subcutaneous injection
In vitro
CaMKIIMouse/
CPVT (RYR2R176Q/+)
Human/CPVT; iPSCs (different mutations)
↓ ventricular arrhythmia in vivo
The origin of the genetic material used in the studies is indicated, if indicated in the publication. Injection into the aortic root or the LV cavity was performed during transient cross-clamping of the great vessels. Intramyocardial injection was performed after thoracotomy. ↓: decrease; ↑: improvement; AAV9: adeno-associated virus serotype 9; Ca2+: calcium; CaM: calmodulin; CaMKII: calmodulin-dependent protein kinase II; CASQ2: calsequestrin 2; CPVT: catecholaminergic polymorphic ventricular tachycardia; DADs: delayed afterdepolarizations; hiPSCs: human induced pluripotent stem cells; RyR2: ryanodine receptor type 2; VT: ventricular tachycardia; wt: wild-type.
Table 5. Downregulation or overexpression of cardiac Na+/Ca2+ exchanger (NCX).
Table 5. Downregulation or overexpression of cardiac Na+/Ca2+ exchanger (NCX).
Author (year)VectorDelivery TechniqueSpecies/ModelExpression PropertiesOutcomes
Schillinger et al. (2000) [184]ADIn vitroRabbit/wtOE↓ contractile function
Terracciano et al. (2001) [175]Transfection reagentIn vitro
Injected into nuclei
Mouse/wtOE↑ Ca2+ handling and homeostasis
Ranu et al. (2002) [185]ADIn vitroRabbit/wtOE↓ contraction amplitude
Tadros et al. (2002) [186]ADIn vitroRat/MIDR↓ Ca2+ influx and efflux
Schillinger et al. (2003) [187]ADIn vitroRabbit/wtOESystolic and diastolic dysfunction
Bölck et al. (2004) [188]ADIn vitroRat/wtOE↓ cell shortening at higher stimulation frequencies
↑ intracellular systolic Ca2+ and contractile amplitude at low stimulation rates
The origin of the genetic material used in the studies is indicated, if indicated in the publication. Injection into the aortic root or the LV cavity was performed during transient cross-clamping of the great vessels. Intramyocardial injection was performed after thoracotomy. ↓: decrease; ↑: improvement; AD: adenovirus; Ca2+: calcium; OE: overexpression; DR: downregulation; wt: wild-type.
Table 6. Gene therapies for HF and inherited cardiac diseases targeting mitochondrial proteins, such as mitochondrial calcium uniporter (MCU), mitochondrial uncoupling protein 2 (UCP2) and mitochondrial N+/Ca2+ exchanger (NCLX).
Table 6. Gene therapies for HF and inherited cardiac diseases targeting mitochondrial proteins, such as mitochondrial calcium uniporter (MCU), mitochondrial uncoupling protein 2 (UCP2) and mitochondrial N+/Ca2+ exchanger (NCLX).
Author (year)VectorDelivery TechniqueGenetic InformationSpecies/ModelOutcomes
Wu et al. (2015) [194]ADIn vitro, Mouse embryonic stem cellsDN-MCUMouse/wtMCU is necessary for physiological heart rate acceleration
Oropeza-Almazán et al. (2017) [195]Transfection reagentIn vitrosiRNA targeting MCURat/H/R injury↓ mitochondrial permeability pore opening
↓ oxidative stress
Suarez et al. (2018) [196]AAV9Direct jugular vein injectionMCUMouse/DiabeticRestoration of cardiac myocyte and heart function
Larbig et al. (2017) [51]Knock-out modelUCP2-/-Mouse/Knock-out↓ ICa-L in CM
↑ slope factor of action potential upstrokes
↑ hyperpolarized resting membrane potential
↓ PR, WRS and QTc interval
↑ after-depolarizations
↑ arrhythmias
Luongo et al. (2017) [193]Knock-out and OE modelSLC8B1 (NCLX)Mouse/Knock-out and OEmCa2+ clearance
Prevention of heart failure
The origin of the genetic material used in the studies is indicated, if indicated in the publication. Injection into the aortic root or the LV cavity was performed during transient cross-clamping of the great vessels. Intramyocardial injection was performed after thoracotomy. ↓: decrease; ↑: improvement; AD: adenovirus; AAV9: adeno-associated virus serotype 9; ATP: adenosine triphosphate; Ca2+: calcium; CM: cardiomyocyte; DN: dominant-negative; H/R: hypoxia/reoxygenation; MCU: mitochondrial calcium uniporter; OE: overexpression; siRNA: small interfering ribonucleic acid; VT: ventricular tachycardia; wt: wild-type.

Share and Cite

MDPI and ACS Style

Paar, V.; Jirak, P.; Larbig, R.; Zagidullin, N.S.; Brandt, M.C.; Lichtenauer, M.; Hoppe, U.C.; Motloch, L.J. Pathophysiology of Calcium Mediated Ventricular Arrhythmias and Novel Therapeutic Options with Focus on Gene Therapy. Int. J. Mol. Sci. 2019, 20, 5304. https://doi.org/10.3390/ijms20215304

AMA Style

Paar V, Jirak P, Larbig R, Zagidullin NS, Brandt MC, Lichtenauer M, Hoppe UC, Motloch LJ. Pathophysiology of Calcium Mediated Ventricular Arrhythmias and Novel Therapeutic Options with Focus on Gene Therapy. International Journal of Molecular Sciences. 2019; 20(21):5304. https://doi.org/10.3390/ijms20215304

Chicago/Turabian Style

Paar, Vera, Peter Jirak, Robert Larbig, Naufal Shamilevich Zagidullin, Mathias C. Brandt, Michael Lichtenauer, Uta C. Hoppe, and Lukas J. Motloch. 2019. "Pathophysiology of Calcium Mediated Ventricular Arrhythmias and Novel Therapeutic Options with Focus on Gene Therapy" International Journal of Molecular Sciences 20, no. 21: 5304. https://doi.org/10.3390/ijms20215304

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop