Next Article in Journal
Nanoscale Iron-Based Metal–Organic Frameworks: Incorporation of Functionalized Drugs and Degradation in Biological Media
Next Article in Special Issue
Synergy between Membrane Currents Prevents Severe Bradycardia in Mouse Sinoatrial Node Tissue
Previous Article in Journal
Endogenous miRNA-Based Innate-Immunity against SARS-CoV-2 Invasion of the Brain
Previous Article in Special Issue
Knockdown of Amyloid Precursor Protein Increases Ion Channel Expression and Alters Ca2+ Signaling Pathways
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Calcium Handling in Inherited Cardiac Diseases: A Focus on Catecholaminergic Polymorphic Ventricular Tachycardia and Hypertrophic Cardiomyopathy

1
Marseille Medical Genetics, INSERM, Aix Marseille University, U1251 Marseille, France
2
Department of Congenital and Hereditary Diseases, Charles Nicolle Hospital, Tunis 1006, Tunisia
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(4), 3365; https://doi.org/10.3390/ijms24043365
Submission received: 5 January 2023 / Revised: 2 February 2023 / Accepted: 4 February 2023 / Published: 8 February 2023
(This article belongs to the Special Issue Calcium Handling 2.0)

Abstract

:
Calcium (Ca2+) is the major mediator of cardiac contractile function. It plays a key role in regulating excitation–contraction coupling and modulating the systolic and diastolic phases. Defective handling of intracellular Ca2+ can cause different types of cardiac dysfunction. Thus, the remodeling of Ca2+ handling has been proposed to be a part of the pathological mechanism leading to electrical and structural heart diseases. Indeed, to ensure appropriate electrical cardiac conduction and contraction, Ca2+ levels are regulated by several Ca2+-related proteins. This review focuses on the genetic etiology of cardiac diseases related to calcium mishandling. We will approach the subject by focalizing on two clinical entities: catecholaminergic polymorphic ventricular tachycardia (CPVT) as a cardiac channelopathy and hypertrophic cardiomyopathy (HCM) as a primary cardiomyopathy. Further, this review will illustrate the fact that despite the genetic and allelic heterogeneity of cardiac defects, calcium-handling perturbations are the common pathophysiological mechanism. The newly identified calcium-related genes and the genetic overlap between the associated heart diseases are also discussed in this review.

1. Introduction

In the heart, calcium (Ca2+) is fundamental for the generation of the contractile force that gives rise to the heartbeat. Indeed, Ca2+ plays a key role in excitation–contraction coupling (ECC) and electrophysiological signaling in the heart [1]. Of note, the sarcoplasmic reticulum (SR) acts as an intracellular Ca2+ storage in the cardiomyocytes. The regulation of the release of Ca2+ ions from SR Ca2+ stores is mediated by the ECC, which then propagates and binds to the myofilaments to initiate the systole phase (contraction). The diastole phase (relaxation) is then ensured by the reuptake of Ca2+ into the SR Ca2+ store via sarcoendoplasmic reticulum Ca2+ ATPase (SERCA2a) or extruded from the cell via the Na+/Ca2+ exchanger [1,2,3]. This process is known as Ca2+-induced Ca2+ release (CICR) and is mediated by the ryanodine receptors (Ca2+ activation of Ca2+ release channels) [4]. Intracellular Ca2+ homeostasis in cardiomyocytes is regulated by the phosphorylation and dephosphorylation of several key Ca2+-handling proteins.
Given that Ca2+ is the cornerstone of cardiac electrophysiology and contraction, calcium mishandling has been associated with contractile dysfunction, arrhythmia, and cellular hypertrophy [5].
In this review, we focus on the genetic etiology of cardiac diseases related to calcium mishandling. We will approach the subject by focalizing on two clinical entities: catecholaminergic polymorphic ventricular tachycardia (CPVT) as a cardiac channelopathy and hypertrophic cardiomyopathy (HCM) as a primary cardiomyopathy. HCM is considered the most common primary cardiomyopathy, usually due to heterozygous mutations in sarcomeric genes. HCM may lead to severe complications such as heart failure and sudden cardiac death (SCD). On the other hand, catecholaminergic polymorphic ventricular tachycardia (CPVT) is a rare genetic arrhythmia that is mainly defined by bidirectional or polymorphic ventricular tachycardia occurring during physical activity or emotional stress. CPVT patients are predisposed to life-threatening arrhythmias. CPVT is caused by mutations in Ca2+-regulating genes. This review will define both diseases and describe their genetic etiology by reporting the different Ca2+-related genes.

2. Methods

We conducted a literature search on PubMed for cardiac disorders related to calcium handling with a special focus on catecholaminergic polymorphic ventricular tachycardia (CPVT) and hypertrophic cardiomyopathy (HCM). Different keyword combinations were applied: “catecholaminergic polymorphic ventricular tachycardia” AND “calcium” OR “calcium handling” OR “calcium mishandling” OR “calcium homeostasis”, “catecholaminergic polymorphic ventricular tachycardia” AND “RYR2, CASQ2, TECRL, TRDN, CALM1, CALM2, CALM3, PKP2 OR ANK2”,“hypertrophic cardiomyopathy” AND “calcium” OR “calcium handling” OR “calcium mishandling” OR “calcium homeostasis”, “hypertrophic cardiomyopathy” AND “calcium” OR “calcium handling” OR “calcium mishandling” OR “calcium homeostasis”, “hypertrophic cardiomyopathy” AND “RYR2, ALPK3, TNNC1, JPH2 OR PLN”. The relevance of the articles was determined by screening titles and abstracts. About sixty articles were selected for each heart disease.
Gene data, namely the gene symbol, location, exon count, protein function, and associated OMIM diseases, were collected from several databases: NCBI-gene (https://www.ncbi.nlm.nih.gov; accessed 15 September 2022), Uniprot (https://www.uniprot.org; accessed 15 September 2022), GeneCards (https://www.genecards.org/; accessed 15 September 2022), GeneMANIA (https://genemania.org; accessed 2 November 2022), and OMIM (https://www.omim.org; accessed 2 November 2022).
All variants were described following the recommendations of the Human Genome Variation Society (http://varnomen.hgvs.org/recommendations; accessed 22 January 2023).

3. Results and Discussion

3.1. Catecholaminergic Polymorphic Ventricular Tachycardia

CPVT is a genetic stress-induced cardiac channelopathy characterized by adrenergically mediated polymorphic ventricular tachyarrhythmias that may lead to SCD, particularly in pediatric and young adult cases [6,7,8]. As the name implies, CPVT is triggered by increased catecholamines during exercise or emotional stress in healthy individuals with structurally normal hearts and a regular resting ECG [6].
The prevalence of CPVT is about 1:10,000 in the general population [8,9]. The average age of onset of clinical symptoms is between 7 and 9 years [7,10].
Clinically, CPVT is diagnosed in the absence of structural cardiac anomalies, normal ECG, and unexplained catecholamine-induced or stress-induced bidirectional VT, polymorphic premature ventricular beats, or VT in persons younger than 40 years [11].
Usually, a definite CPVT diagnosis is made after an average delay of 2 years from the first arrhythmogenic event due to a primary suspicion of vasovagal discomfort or a neurological etiology [12]. The early detection of CPVT is crucial to prevent SCD. Hence, molecular genetic screening is necessary to confirm an uncertain clinical diagnosis of CPVT and identify asymptomatic family members.
The life-threatening arrhythmias in CPVT are mainly caused by unregulated Ca2+ release from the SR [13]. Indeed, the electrocardiographic pattern of the ventricular tachycardia observed in CPVT patients is very similar to the arrhythmias linked to intracellular Ca2+ overload and the delayed afterdepolarizations (DADs) detected in digitalis toxicity [14,15]. Thus, DADs and triggered physical activity have been proposed as the underlying arrhythmogenic mechanism in CPVT [14,15]. Consistently, mutations in calcium-handling proteins implicated in the release of Ca2+ from the SR are associated with the CPVT phenotype, namely the RYR2 and CASQ2 proteins [3,16]. Of note, RYR2 and CASQ2 are parts of the multimolecular Ca2+ release channel complex located in the SR. Although RYR2 serves as a Ca2+ release channel, the SR Ca2+-binding protein, CASQ2, plays a dual role by regulating RYR2 function and serving as a buffer for SR Ca2+ [13].
To date, a handful of Ca2+-handling genes have been linked to CPVT including mainly the RYR2 gene and, less frequently, the CASQ2, TRDN, CALM1, CALM2, CALM3, TECRL, PKP2, and ANK2 genes [3,11]. Indeed, the incidence of RYR2 mutations in CPVT patients ranges from 35 to 79%, whereas mutations in the CASQ2 gene account for 3–5% of CPVT cases [11]. The CPVT-associated genes with their main functions are summarized in Table 1.

3.1.1. Major CPVT Genes

Ryanodine Receptor 2

The ryanodine receptor 2 (RYR2) gene encodes the cardiac ryanodine receptor in the SR. The encoded RYR2 protein is the major Ca2+ channel protein in the membrane of the SR, which acts as intracellular Ca2+ storage in cardiomyocytes. Ca2+ can be released from the SR to the cytosol by the RYR2 channels [17]. Ca2+ release in cardiomyocytes is triggered by increased Ca2+ levels due to the activation of CACNA1C channels. A dysfunctional RYR2 channel leads to diastolic Ca2+ leak from the SR and contributes to the development of DADs [15].
During normal cardiac contraction, RYR2 is activated by cytosolic Ca2+, whereas under the conditions of storage overload, RYR2 opening is regulated by SR Ca2+ [18]. In addition to RYR2 activation, SR luminal Ca2+ also determines the RYR2 channel closing [18].
During exercise, RYR2 phosphorylation by protein kinase A partially dissociates FK-binding protein 12.6 (FKBP12.6) from the RYR2 channel, leading to an increase in intracellular Ca2+ release and cardiac contractility [19]. Functional studies using Fkbp12.6 −/− mice showed exercise-induced cardiac ventricular arrhythmias resulting in SCD. Indeed, RYR2 mutations linked to exercise-induced arrhythmias in patients with CPVT reduced the affinity of FKBP12.6 for RYR2 channels and increased single-channel activity during exercise. These findings suggested that ‘leaky’ RYR2 channels can trigger malignant arrhythmias, likely causing CPVT [19].
Gain-of-function mutations in RYR2 are found in approximately 79% to 95 % of CPVT1 cases with an autosomal dominant pattern of inheritance [11,20]. Loss-of-function RYR2 mutations are less frequent and linked to other ventricular arrhythmia syndromes [7].
CPVT-linked RYR2 mutations increase the likelihood of spontaneous RYR2 openings and Ca2+ leak from the SR during diastole, triggering malignant arrhythmias [20]. To date, over 150 RYR2 mutations have been associated with CPVT [16]. The majority of RYR2 mutations are located in four well-conserved domains including the pore, pseudo-voltage sensor, and central domains. Indeed, these domains are implicated in channel activation and gating. A potential link between mutation localization and phenotype severity has been emphasized [18,21]. Therefore, mutations located in the C-terminus of the RYR2 protein have been correlated with sudden death during sleep [22].

Calsequestrin 2

The calsequestrin 2 (CASQ2) gene encodes the calsequestrin protein localized in the SR of cardiac and slow skeletal muscle cells. CASQ2 is Ca2+-binding protein that stores Ca2+ for muscle function. Ca2+ ions are bound by clusters of acidic residues at the protein surface [23]. CASQ2 plays a pivotal role in triggering muscle contraction by regulating the release of lumenal Ca2+ through the RYR2 channel. Thus, CASQ2 significantly contributes to the cardiac ECC and regulates the heartbeat rate [23].
Mutations in CASQ2 cause approximately 2 to 5% of CPVT with an autosomal recessive pattern of inheritance. Rarely is an autosomal dominant model associated with CASQ2 mutations [8]. CASQ2 c.539A>G; p.(Lys180Arg) was the first reported mutation associated with an autosomal-dominant inheritance of CPVT [24].
Mutations in CASQ2 result in a lack of control of the RYR2 channel and, consequently, a constant release of Ca2+ into the cytoplasm, leading to arrhythmias [13]. Indeed, to regulate Ca2+ release, calsequestrin is anchored to RYR2 by triadin and junction proteins. It has been proposed that the interaction between CASQ2 and RYR2 may contribute to the refractory period of Ca2+ release occurring after each physiological CICR but the mechanism is not fully understood [13,20]. The most plausible CPVT mechanism linked to nonsense CASQ2 mutations is impaired Ca2+ buffering. However, missense mutations, such as the CASQ2 c.98G>A; p.(Arg33Gln), have been linked to Ca2+ buffering decrease and an alteration of CASQ2/RYR2 interaction [20].
Casq2-null mice exhibited normal Ca2+ release from the SR and contractile function under basal conditions. However, mutant mice had an increase in SR volume and an absence of Casq2-binding proteins such as triadin-1 and junction [25]. Exposure to catecholamines in Casq2-null myocytes induced increased diastolic SR Ca2+ leak, which resulted in premature spontaneous SR Ca2+ release and triggered arrhythmias [25].

Trans-2,3-Enoyl-CoA Reductase-like

The trans-2,3-enoyl-CoA reductase-like (TECRL) gene encodes the trans-2,3-enoyl-CoA reductase protein, which is an endoplasmic reticulum protein mainly expressed in the heart and skeletal muscle [26]. The TECRL protein consists of 363-amino acid with an N-terminal ubiquitin-like domain, 3 transmembrane regions, and a C-terminal 3-oxo-5-alpha steroid 4-dehydrogenase domain and plays a crucial role in intracellular Ca2+ homeostasis [27]. Indeed, the concentration of the expressed TECRL protein is critical for Ca2+ regulation of major cardiac proteins such as RYR2, CASQ2, and CALM [26,27].
Bhuiyan et al. in 2007 described the first CPVT phenotype associated with the TECRL gene according to an autosomal recessive pattern of inheritance, mapped to chromosome locus 7p22–p14 [28]. The members of this family were diagnosed with an early-onset and highly malignant form of CPVT with a history of SCD during physical activity [27,28]. Using exome sequencing, Devalla et al. (2016) identified a homozygous loss-of-function mutation in the TECRL gene in all the affected members of this same family [27].
To assess the functional consequence of the TECRL c.331+1G>A mutation, human induced pluripotent stem cells (hiPSCs) from a 5-year-old homozygous patient (TECRLHom-hiPSCs), his heterozygous father (TECRLHet-hiPSCs), and a non-carrier family member (CTRL-hiPSCs) were generated [27]. hiPSCs were differentiated into cardiomyocytes (CMs) and analyzed in vitro. Using this in vitro model, the authors showed that the c.331+1G>A mutation in TECRL leads to the skipping of exon 3. Moreover, the TECRLHomhiPSC-CMs closely replicated the disease phenotype and the mutant cells showed an increase in triggered electrical activity upon catecholaminergic stimulation [27].
Analysis of intracellular calcium dynamics of the TECRLHom-hiPSCs revealed altered Ca2+ properties, including a high diastolic Ca2+, smaller amplitude and slower decay of cytosolic Ca2+ transients, and a prolonged action potential duration [26,27]. This study reported two additional families with overlapping clinical features of long QT syndrome (LQTS) and CPVT carrying TECRL mutations [27]. Interestingly, both heterozygous and homozygous TECRL c.587G>A; p.(Arg196Gln) mutations can cause a decrease in Ca2+ stores in the SR and an increase in diastolic cytoplasmic Ca2+ concentration in cardiomyocytes, leading to a CVPT phenotype [27].
The TECRL c.331+1G>A mutation was subsequently reported by Jaouadi et al. (2020) in a consanguineous family with three deceased children, each at 8 years old [29]. The three SCD events occurred during normal daily activities (playing, slow walking, and at school). Exome sequencing of the family revealed the presence of the homozygous TECRL c.331+1G>A mutation in the last deceased child. Both parents were found to be heterozygous for the variant. The father was asymptomatic with a structurally normal heart and no history of cardiac arrhythmias, whereas the mother had a history of syncopes and a clinical suspicion of Brugada syndrome [29]. Intriguingly, no CPVT or LQTS features were noted in this family [29].
Although the first reported patients with TECRL mutations displayed strict CPVT features or CPVT-specific features combined with a long QT interval but not an isolated LQTS [27,30], the newly identified cases carrying TECRL mutations displayed divergent cardiac phenotypes within a single genetic locus [30].
Moscu-Gregor et al. (2020) have identified four additional mutations in the TECRL gene in CPVT patients with severe and early-onset clinical presentation at the homozygous and compound heterozygous state (c.415C>T; p.(Gln139*), c.893T>C; p.(Val298Ala), c.926C>A; p.(Ser309*), and c.869C>A; p.(Pro290His)). The authors concluded that variants in TECRL may be causative of up to 5% of CPVT patients [31].
Overall, patients with TECRL mutations presented a highly lethal form of arrhythmias, with a median age of symptom onset at 8 years of age [30,32].

Triadin

Triadin (TRDN) is one of the major transmembrane proteins located in the junctional SR playing a role in ECC regulation and Ca2+ influx in the calcium release complex [33]. Mutations in the TRDN gene lead to a significant decrease in protein expression causing Ca2+ overload in the SR, which may explain the development of CVPT [34,35].
From a cohort of 97 CPVT patients with no mutations in the RYR2 and CASQ2 genes, Roux-Buisson et al. (2012) identified three mutations in the TRDN gene, which cosegregated with the disease according to an autosomal recessive pattern in two families: a c.del53_56ACAG; p.(Asp18Alafs*13) homozygous deletion in the first family and compound heterozygous mutations c.176C>G; p.(Thr59Arg) and c.613C>T; p.(Gln205*) in the second family [35]. Thereafter, Rooryck et al. (2015) identified compound heterozygous pathogenic mutations (c.613C>T; p.(Gln205*) and c.22 + 29 A>G) in two sisters with CPVT [36]. Overall, the TRDN gene was associated with less than 1% of CPVT5 cases [35].

Calmodulin

Calmodulin proteins (CALM) are members of the EF-hand calcium-binding protein family and play an essential role in Ca2+ sensing and signal transducing. Three distinct calmodulin genes (CALM1, CALM2, and CALM3) are distributed within the human genome that encode the identical protein but differ at the nucleotide level. The three calmodulin genes share about an 80% identity within their coding regions. Calcium-induced activation of calmodulin modulates the function of cardiac ion channels including CACNA1C, SCN5A, and RYR2 [37,38].
Mutations in CALM1, CALM2, and CALM3 genes have been associated with CPVT.
Nyegaard et al. (2012) identified a heterozygous CALM1 c.161A>T; p.(Asn53Ile) mutation that segregated with the disease in 10 affected family members [38]. A de novo missense mutation in CALM1, c.293A>G; p.(Asn97Ser), was subsequently identified in a 23-year-old woman with a history of resuscitated cardiac arrest at 4 years of age due to ventricular fibrillation while running. Both substitutions showed compromised calcium binding [38].
Makita et al. (2014) identified two heterozygous missense CALM2 mutations in two patients with overlapping features of LQTS and CPVT [39]. The c.396T>G; p.(Asp132Glu) mutation was identified in a 29-year-old woman who was initially diagnosed with neonatal LQTS and later with exercise-induced polymorphic ventricular ectopy. The second CALM2 variant, c.407A>C; p.(Gln136Pro), was identified in an 8-year-old girl with a presumptive diagnosis of LQTS and CPVT who died suddenly during exercise despite treatment with β-blockers [39]. The two mutations were de novo. The encoded mutant calmodulin proteins impaired C-domain Ca2+-binding affinity, likely causing Ca2+ signaling dysfunction [39].
Gomez-Hurtado et al. (2016) identified a heterozygous CALM3 mutation, c.308C>T; p.(Ala103Val), in a 31-year-old woman among a cohort of 12 CPVT patients with no mutations in the other known CPVT genes. The CALM3 mutation was shown to activate RYR2 Ca2+ release channels, generating Ca2+ waves and depleting the SR Ca2+ store [40,41]. Moreover, it has been shown that CPVT calmodulin mutants tend to bind to RYR2 with higher affinity than wild-type, which can explain their autosomal-dominant mode of action [40,41].
Thus, both de novo and inherited mutations have been reported and patients harboring CALM mutations may present overlapping features of LQTS and CPVT.

3.1.2. Minor CPVT Genes

Plakophilin-2

The plakophilin-2 (PKP2) gene encodes the desmosomal plakophilin-2 protein. Mutations in the PKP2 gene were associated primarily with arrhythmogenic right ventricular cardiomyopathy (ARVC) [42]. Subsequently, Cerrone et al. (2017) have used conditional mouse deletion of Pkp2 in cardiomyocytes to demonstrate that the lack of Pkp2 reduces expression levels of Ryr2, Ank2, Cacna1c, Trdn, and Casq2 proteins, leading to disruption of intracellular Ca2+ homeostasis and isoproterenol-induced arrhythmias, even in the absence of overt structural heart disease [43]. Tester et al. (2019) screened the PKP2 gene in genotype-negative patients with CPVT. PKP2 mutations were found in 27.7% of CPVT cases and 5.3% of exercise-related sudden unexplained death in the young cases. Cardiac imaging or autopsy demonstrated a structurally normal heart in all patients [44].

Ankyrin-2

The ankyrin 2 (ANK2) gene encodes an ankyrin-B protein that is located mainly in the transverse-tubule SR sites of the cardiomyocytes [45]. This protein is a crucial part of the Na+/Ca2+ exchanger, Na+/K+ ATPase, and inositol trisphosphate (InsP3) receptor. Thus, ankyrins have key roles in membrane trafficking and regulation of different ion channels in the heart [46]. A loss-of-function mutation of ankyrin-B was initially associated with LQTS 4 [45]. Subsequently, Mohler et al., (2004 and 2007), identified nine ANK2 loss-of-function mutations in patients with variable expressivity of cardiac dysfunction including bradycardia, sinus arrhythmia, idiopathic ventricular fibrillation, and CPVT [47,48]. The CPVT patients carried the ANK2 c.4864C>A; p.(Leu1622Ile), c.5437G>A; p.(Glu1813Lys), and c.T4547A; p.(Val1516Asp) mutations [47,48]. The ANK2 mutations were shown to abolish the ability of ankyrin-B to restore defective Ca2+ dynamics. The authors also noted abnormal localization and expression of the Na+/Ca2+ exchanger, Na+/K+ ATPase, and InsP3 [47].
Table 1. Calcium-related genes linked to catecholaminergic polymorphic ventricular tachycardia.
Table 1. Calcium-related genes linked to catecholaminergic polymorphic ventricular tachycardia.
HGNC_Gene SymbolFull NameLocationExon CountProtein Function OMIM DiseaseOMIM IDsInheritanceReferences
RYR2Ryanodine receptor 21q43105Calcium release channels from the sarcoplasmic reticulum into the cytoplasm by ER and SR.
Activates and modulates small-conductance Ca2+-activated K+ channels in cardiac myocytes.
Regulates cardiac muscle contraction by calcium ion signaling.
ARVD2600996AD[18,49,50]
VA115000AD
CPVT1604772AD
CASQ2Calsequestrin 21p13.111Major Ca2+-binding protein in the SR.
Key SR Ca2+ storage protein essential for SR Ca2+ release in the heart.
Structural organization of the SR with TRDN.
Facilitates high rates of Ca2+ release through RYR2 during systole.
Plays a critical role in mobilizing Ca2+ release from ER/SR lumens.
Role in ECC in the heart and regulation of heart rate beats.
Regulates cardiac muscle conduction and contraction by calcium ion signaling.
Regulates ryanodine-sensitive calcium-release channel activity.
CPVT2611938AR/
AD
[51,52]
TECRLTrans-2,3-Enoyl-CoA Reductase Like4q13.112ER protein.
Role in intracellular Ca2+ homeostasis.
Regulates heart contraction.
CPVT3614021AR[26,27,30,31]
TRDNTriadin6q22.3141Contributes to the regulation of lumenal Ca2+ release via the SR calcium release channels RYR1 and RYR2, a key step in triggering skeletal and heart muscle contraction.
Regulates the release of sequestered calcium ions into the cytosol by SR.
Cell–cell signaling involved in cardiac conduction.
Anchors calsequestrin to the junctional SR, allowing its functional coupling with the ryanodine receptor.
Indirect role of triadin in regulating myoplasmic Ca2+ homeostasis and organizing the molecular complex of the triad but not in regulating skeletal-type excitation–contraction coupling.
CPVT5615441AR[35,53]
CALM1Calmodulin 114q32.116Intracellular Ca2+ transducer involved in numerous activities in a broad Ca2+ signaling network.
Regulates RYR1 and RYR2 by binding to a single, highly conserved calmodulin binding site.
Regulates tail-anchored insertion into the ER membrane in a Ca2+-dependent manner.
LQTS14616247AD[54,55]
CPVT4614916AD
CALM2Calmodulin 22p216Phosphorylase kinase, delta, calcium-modulated protein.
Mediates the control of a large number of enzymes and other proteins by Ca2+.
Plays a crucial role in the processes of Ca2+-induced neuronal cell death.
LQTS 15616249AD[37,56]
CALM3Calmodulin 319q13.326Calcium-modulated protein.
Mediates the control of a large number of enzymes by Ca2+, protein kinases, and phosphatases.
CPVT6618782AD[37,40,57]
LQTS16
PKP2Plakophilin-212p11.2113Regulates the signaling activity of beta-catenin.
Maintains the transcription of genes that control intracellular calcium cycling including RYR2, ANK2, TRDN, and CACNA1C.
Regulates cardiac muscle cell contraction and cell action potential.
Regulates actin filament-based movement/cardiac muscle tissue development/cell–cell junction organization.
ARVD 9609040AD[42]
ANK2Ankyrin 24q25-q2646Required for coordinated assembly of Na/Ca exchanger, Na/K ATPase, and inositol trisphosphate INSP3 receptor at transverse-tubule/sarcoplasmic reticulum sites in cardiomyocytes.
Regulates cardiac muscle contraction by calcium ion signaling.
Role in normal cardiac electric activity and cardiac automaticity.
Regulates KCNJ5 channel gating.
LQTS4600919AD[58,59,60]
ARVD: arrhythmogenic right ventricular dysplasia; AD: autosomal dominant; AR: autosomal recessive; CPVT: catecholaminergic polymorphic ventricular tachycardia; DCM: dilated cardiomyopathy; ECC: excitation–contraction coupling ER: endoplasmic reticulum; SR: sarcoplasmic reticulum; HCM: hypertrophic cardiomyopathy; LQTS: long QT syndrome.

3.2. Hypertrophic Cardiomyopathy

HCM is a primary cardiac disorder characterized by an increased left ventricular wall thickness in the absence of other loading conditions [61]. HCM is the most common inherited heart disease, with a prevalence of 1/200 to 1/500, and is mostly inherited in an autosomal dominant manner [62,63]. Patients with HCM have a higher risk of developing clinical complications such as progressive heart failure, arrhythmia, and SCD [63]. Molecular genetic studies have demonstrated that HCM is mainly caused by mutations in sarcomeric genes encoding contractile myofilament proteins [62,64,65,66]. The frequency of mutations within sarcomeric genes varies from 25% to 65% of patients [67].
Increased calcium buffering has been proposed as the causal mechanism leading to the alteration of intracellular Ca2+ cycling and triggering Ca2+-dependent hypertrophy [68]. In vivo experiments using LV guinea pig cardiomyocytes expressing mutations in the TNNT2, TNNI1, and TPM1 genes demonstrated increased diastolic Ca2+ and Ca2+ reuptake [68]. Moreover, mutations in the MYBPC3 gene may increase myofilament Ca2+ sensitivity and promote cardiac hypertrophy due to the inability to release Ca2+ and relax from contraction [69].
Nevertheless, there is a lack of strong evidence to show whether these alterations are a causal factor of HCM or consequential clinical manifestations. With the advent of patient-specific iPSC models, Lan et al. (2013) have provided evidence that the elevation of intracellular Ca2+ is the initial factor in HCM development [5]. Indeed, time-based gene expression data analysis of single iPSC-CMs carrying the MYH7: Arg663His mutation revealed that the downstream effectors of cardiac hypertrophy (e.g., GATA4 and MEF2) were expressed in a Ca2+-dependent manner [5]. Furthermore, the authors concluded that elevated cardiomyocyte Ca2+ loading seems to contribute to both cardiac hypertrophy and arrhythmogenesis [5].
Given the aforementioned observations and key role of Ca2+ in ECC, several studies have investigated the implication of mutations in calcium-related genes in HCM pathogenesis. The TNNC1 gene was among the first genes encoding a Ca2+-sensitive/handling protein to be neatly linked to HCM [70]. Subsequently, several mutations in other genes, such as ALPK3, RYR2, PLN, and JPH2, have been implicated in the pathogenesis of HCM [67]. It should be noted that sarcomere mutations in HCM either directly or indirectly influence intracellular calcium [71]. The HCM-associated genes linked to calcium with their main functions are summarized in Table 2.

3.2.1. Troponin C1

Troponin is a central regulatory protein of striated muscle contraction located on the actin filament. The cardiac troponin complex is a heterotrimeric myofilament composed of three subunits: an elongated troponin T subunit (TNNT2), an inhibitory troponin I subunit (TNNI3), and a Ca2+-sensitive troponin C subunit (TNNC1). TNNC1 is a sarcomeric Ca2+ sensor that binds to the cytosolic divalent cation at the specific Ca2+ binding site to enhance its interaction with TNNI3. This complex reduces the inhibitory function of TNNI3, releasing it from actin, and causes the troponin–tropomyosin complex to move into the actin groove, exposing myosin binding sites [72]. Accordingly, TNNC1 plays a critical molecular role in the initiation of myofilament contraction [67].
Mutations in the TNNC1 gene are rare, occurring in ~0.4% of HCM patients [73]. The first HCM-associated mutation in TNNC1, c.86T>A; p.(Leu29Gln), was identified by Hoffmann et al. in 2001 [70]. Using in vitro and in situ models, this mutation was found to affect the Ca2+-dependent structural change in cardiac TnC in trabeculae under basal conditions and abolish the effect of force-generating myosin cross-bridges. [74]. Six additional mutations in the TNNC1 gene (c.23C>T; p.(Ala8Val), c.91G>T; p.(Ala31Ser), c.251G>A; p.(Cys84Tyr), c.402G>T; p.(Glu134Asp), c.435C>A; p.(Asp145Glu), and c.363dupG; p.(Gln122Alafs*30)) have been associated with HCM by the screening of the TNNC1 gene in a cohort of 1025 HCM patients [73,75,76,77].
Functional studies showed an increased Ca2+ sensitivity of force development for c.23C>T, c.251G>A, and c.435C>A and force recovery for c.23C>T and c.435C>A mutations. The TNNC1 c.402G>T mutation showed no changes in these parameters [73]. The frameshift mutation TNNC1 c.363dupG is located in the EF-hand 3 domain and was found to destroy the H-helices of troponin C that are required for the interaction with troponin I [78]. Moreover, the functional analysis suggested that the TNNC1 c.91G>T mutation directly affects Ca2+ sensitivity and may alter Ca2+ handling, leading to arrhythmogenesis [67].

3.2.2. Ryanodine Receptor Type 2

Mutations in the ryanodine receptor type 2 (RYR2) gene are typically associated with CPVT, ventricular arrhythmias due to calcium release deficiency syndrome, and arrhythmogenic right ventricular cardiomyopathy/ dysplasia (Table 1). In 2006, Fujino et al. reported the first RYR2 mutation potentially involved in HCM: c.3320C>T; p.(Thr1107Met) [79]. This mutation was subsequently identified in exome-sequencing cohorts with a relatively high frequency (MAF=0.0004) [80] and in CPVT patients [81], questioning its involvement in HCM. Recently, Alvarado et al. (2019) identified a novel RYR2 mutation c.3372G>A; p.(Pro1124Leu), in an HCM patient who did not have a sarcomeric mutation [82]. Functional studies have shown that homozygous mice for this mutation presented mild cardiac hypertrophy combined with an increase in the expression of calmodulin, a classical inhibitor of RYR2 [82].

3.2.3. Alpha Kinase 3

The alpha kinase 3 (ALPK3) gene encodes the alpha-protein kinase 3 that may act as a transcriptional regulator through the phosphorylation of cardiac transcription factors [83]. The ALPK3 gene is early expressed in the cardiac crescent and later remains highly expressed in cardiomyocytes throughout life [84,85]. Recently, ALPK3 was identified as an important cardiac pseudokinase that inserts into the nuclear envelope and M-band of the sarcomere [85].
Homozygous ALPK3-truncating mutations were initially associated with early-onset cardiomyopathy [84]. Pediatric cases reported thereafter consistently displayed a severe clinical presentation with irregular cardiac remodeling and an inconsistent syndromic pattern [84,86,87,88]. Indeed, pediatric cases with biallelic ALPK3-truncating mutations showed a variable HCM phenotype with atypical distribution of hypertrophy (concentric/ apical/ asymmetric septal hypertrophy/right ventricular dysfunction) and variable facio-thoraco-skeletal features [84,85,86,87,88,89]. It is noteworthy that homozygous ALPK3 carriers diagnosed in utero or at birth presented a mixed DCM/HCM phenotype with progression to an HCM phenotype with age [87,89]. Although the majority of heterozygous parents or relatives of the reported cases were found to be healthy or with mild HCM, recent case reports and studies extended the phenotype and genotype spectrum of ALPK3 mutations to include (i) missense mutations, (ii) compound-heterozygous and autosomal-dominant patterns of inheritance, and (iii) adult-onset HCM with a less severe clinical presentation [89,90].
The ALPK3 gene appears to be essential for the normal formation of the intercalated disc and the organization of cardiomyofibril in humans and mice [86,88,91]. It has been shown that hiPSCs-derived cardiomyocytes containing a homozygous ALPK3- truncating mutation displayed defective Ca2+ handling in addition to sarcomeric disorganization and impaired intercalated disc integrity [84,88].
A functional study of the Ca2+ flux during contraction demonstrated that ALPK3-mutant cardiomyocytes displayed several changes in intracellular Ca2+ including significantly increased irregular Ca2+-transients, which may explain the cellular hypertrophy [88]. Indeed, using multi-electrode array analysis, ALPK3–hiPSCs-derived cardiomyocytes demonstrated an extended extracellular field potential duration, indicating that the loss of ALPK3 disrupts membrane repolarization [88].
These results suggest that intracellular Ca2+ is elevated in ALPK3-deficient cardiomyocytes, agreeing with what has been observed in other hiPSC models of HCM [84,88].

3.2.4. Junctophilin 2

The junctophilin 2 (JPH2) gene encodes junctophilin 2, a major component of the junctional membrane complex involved in Ca2+ homeostasis and ECC [92]. Junctophilins are a family of proteins found in all excitable cells [93]. JPH2 plays a crucial role in maintaining the proper structure of the cardiac dyad, which is necessary for effective CICR [94]. The implication of the JPH2 gene in HCM was first reported by Landstrom et al. (2007) with the identification of three mutations (c.301A>C; p.(Ser101Arg), c.421T>C; p.(Tyr141His), and c.494C>T; p.(Ser165Phe), in three unrelated patients with HCM negative for sarcomeric or Z-disc mutations [95]. The characterization of these mutations using an in vitro model of myocyte culture showed a decrease in CICR amplitude and disruption of cellular ultrastructure. The c.421T>C and c.494C>T mutations were found to induce cellular hypertrophy [95]. Of note, mutations in JPH2 are considered a rare cause of HCM and are found in less than 1% of index cases [67].

3.2.5. Phospholamban

The phospholamban (PLN) gene encodes phospholamban, a major substrate for the cAMP-dependent protein kinase in cardiac muscle. In its unphosphorylated state, PLN is an inhibitor of cardiac muscle SERCA2a. This inhibition is abolished upon phosphorylation of the PLN protein. The consecutive activation of the Ca2+ pump results in enhanced muscle relaxation rates. Thus, PLN is a key regulator of cardiac diastolic function [96,97,98].
Mutations in the PLN gene are known to cause DCM [99]. However, rare promoter mutations have been identified in multiple independent cohorts of HCM patients [100,101]. Nevertheless, only the rare truncating PLN c.116T>G; p.(Leu39Ter) mutation is recognized as a causative HCM mutation [102]. This mutation cosegregated with HCM in a multigenerational family and the truncated protein is likely to impair PLN and SERCA2a interactions [103]. The authors also estimated an overall yield of PLN–HCM mutations of 0.65% by comparing different studies reporting PLN mutations in HCM cohorts [102].
Table 2. Calcium-related genes linked to hypertrophic cardiomyopathy.
Table 2. Calcium-related genes linked to hypertrophic cardiomyopathy.
HGNC_Gene SymbolFull NameLocationExon CountProtein Function OMIM DiseaseOMIM IDsInheritanceReferences
RYR2Ryanodine receptor 21q43105Regulates muscle hypertrophy and actin filament-based movement.ARVD2600996AD[104,105]
VA115000AD
CPVT1604772AD
ALPK3Alpha kinase 315q25.314Striated muscle cell development and differentiation.
Actomyosin structure organization.
Heart development and morphogenesis.
Regulates the expression and localization of critical proteins in both the sarcomere M-band and nuclear envelope of cardiomyocytes.
HCM,27618052AR[84,85]
TNNC1Troponin C1, slow skeletal and cardiac type3p21.16Ca2+ sensor and key regulator of cardiac contraction.
Its Ca2+-binding properties modulate the rate of cardiac muscle contraction at submaximal levels of Ca2+ activation.
Modulates the Ca2+-binding properties.
Calcium-binding subunit of the troponin complex responsible for initiating striated muscle contraction in response to calcium influx.
DCM, 1Z611879AD[106,107,108]
HCM13613243AD
JPH2Junctophilin 220q13.125Mediates cross-talk between the cell surface and ER.
Cellular Ca2+ signaling in excitable cells form junctional membrane complexes between the plasma membrane and the ER/SR.
DCM, 2E619492AR[109,110,111]
HCM,17613873AD
PLNPhospholamban6q22.312Crucial Ca2+ cycling protein and a primary mediator of the beta-adrenergic effects, resulting in enhanced cardiac activity.
Regulates the activity of the sarcoplasmic Ca2+ ATPase isoform, a regulator of the kinetics of cardiac contraction.
Unphosphorylated PLN reduces ATP2A1 affinity for Ca2+ and affects enzymatic turnover.
DCM, 1P609909AD[97,112,113]
HCM, 18613874AD
ARVD: arrhythmogenic right ventricular dysplasia; AD: autosomal dominant; AR: autosomal recessive; CPVT: catecholaminergic polymorphic ventricular tachycardia; DCM: dilated cardiomyopathy; ER: endoplasmic reticulum; SR: sarcoplasmic reticulum; HCM: hypertrophic cardiomyopathy; LQTS: long QT syndrome; VA: ventricular arrhythmias due to cardiac ryanodine receptor calcium release deficiency syndrome.

4. Conclusions

This review aimed to summarize the Ca2+-related genes linked to CPVT and HCM with their underlying physiopathological mechanisms. Reporting HCM patients with mutations in Ca2+-related genes stems from the fact that abnormal Ca2+ handling is a common feature of structural heart diseases. Studying calcium handling in healthy and diseased hearts may improve our understanding of Ca2+-mediated cardiac diseases and aid the advancement of therapeutic strategies.

Author Contributions

Writing—Original draft, S.Z. and L.K.; Writing—review and editing, H.J.; Supervision, H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no specific grant from any funding agency in the public, commercial, or not-for-profit sectors.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bers, D.M. Cardiac excitation–contraction coupling. Nature 2002, 415, 198–205. [Google Scholar] [CrossRef] [PubMed]
  2. 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] [CrossRef]
  3. Blayney, L.M.; Lai, T. Ryanodine receptor-mediated arrhythmias and sudden cardiac death. Pharmacol. Ther. 2009, 123, 151–177. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, Y.-T.; Valdivia, C.R.; Gurrola, G.B.; Hernández, J.J.; Valdivia, H.H. Arrhythmogenic mechanisms in ryanodine receptor channelopathies. Sci. China Life Sci. 2014, 58, 54–58. [Google Scholar] [CrossRef] [PubMed]
  5. Lan, F.; Lee, A.S.; Liang, P.; Sanchez-Freire, V.; Nguyen, P.K.; Wang, L.; Han, L.; Yen, M.; Wang, Y.; Sun, N.; et al. Abnormal Calcium Handling Properties Underlie Familial Hypertrophic Cardiomyopathy Pathology in Patient-Specific Induced Pluripotent Stem Cells. Cell Stem Cell 2013, 12, 101–113. [Google Scholar] [CrossRef]
  6. Kim, C.W.; Aronow, W.S.; Dutta, T.; Frenkel, D.; Frishman, W.H. Catecholaminergic Polymorphic Ventricular Tachycardia. Cardiol. Rev. 2020, 28, 325–331. [Google Scholar] [CrossRef]
  7. Roston, T.M.; Yuchi, Z.; Kannankeril, P.J.; Hathaway, J.; Vinocur, J.M.; Etheridge, S.P.; E Potts, J.; Maginot, K.R.; Salerno, J.C.; I Cohen, M.; et al. The clinical and genetic spectrum of catecholaminergic polymorphic ventricular tachycardia: Findings from an international multicentre registry. EP Eur. 2017, 20, 541–547. [Google Scholar] [CrossRef]
  8. Kallas, D.; Lamba, A.; Roston, T.M.; Arslanova, A.; Franciosi, S.; Tibbits, G.F.; Sanatani, S. Pediatric Catecholaminergic Polymorphic Ventricular Tachycardia: A Translational Perspective for the Clinician-Scientist. Int. J. Mol. Sci. 2021, 22, 9293. [Google Scholar] [CrossRef]
  9. Priori, S.G.; Wilde, A.A.; Horie, M.; Cho, Y.; Behr, E.R.; Berul, C.; Blom, N.; Brugada, J.; Chiang, C.-E.; Huikuri, H.; et al. HRS/EHRA/APHRS Expert Consensus Statement on the Diagnosis and Management of Patients with Inherited Primary Arrhythmia Syndromes. Heart Rhythm 2013, 10, 1932–1963. [Google Scholar] [CrossRef]
  10. Jaouadi, H.; Chabrak, S.; Lahbib, S.; Abdelhak, S.; Zaffran, S. Identification of two variants in AGRN and RPL3L genes in a patient with catecholaminergic polymorphic ventricular tachycardia suggesting new candidate disease genes and digenic inheritance. Clin. Case Rep. 2022, 10. [Google Scholar] [CrossRef]
  11. Sumitomo, N. Current topics in catecholaminergic polymorphic ventricular tachycardia. J. Arrhythmia 2016, 32, 344–351. [Google Scholar] [CrossRef]
  12. Liu, N.; Ruan, Y.; Priori, S.G. Catecholaminergic Polymorphic Ventricular Tachycardia. Prog. Cardiovasc. Dis. 2008, 51, 23–30. [Google Scholar] [CrossRef]
  13. Györke, S. Molecular basis of catecholaminergic polymorphic ventricular tachycardia. Heart Rhythm 2009, 6, 123–129. [Google Scholar] [CrossRef]
  14. Priori, S.G.; Napolitano, C.; Memmi, M.; Colombi, B.; Drago, F.; Gasparini, M.; DeSimone, L.; Coltorti, F.; Bloise, R.; Keegan, R.; et al. Clinical and Molecular Characterization of Patients With Catecholaminergic Polymorphic Ventricular Tachycardia. Circulation 2002, 106, 69–74. [Google Scholar] [CrossRef]
  15. Xie, L.-H.; Weiss, J.N. Arrhythmogenic consequences of intracellular calcium waves. Am. J. Physiol. Circ. Physiol. 2009, 297, H997–H1002. [Google Scholar] [CrossRef]
  16. Zhao, Y.-T.; Valdivia, C.R.; Gurrola, G.B.; Powers, P.P.; Willis, B.C.; Moss, R.L.; Jalife, J.; Valdivia, H.H. Arrhythmogenesis in a catecholaminergic polymorphic ventricular tachycardia mutation that depresses ryanodine receptor function. Proc. Natl. Acad. Sci. USA 2015, 112, E1669–E1677. [Google Scholar] [CrossRef]
  17. Seidel, M.; Lai, F.A.; Zissimopoulos, S. Structural and functional interactions within ryanodine receptor. Biochem. Soc. Trans. 2015, 43, 377–383. [Google Scholar] [CrossRef]
  18. Zhang, J.Z.; Waddell, H.M.; Jones, P.P. Regulation of RYR2 by sarcoplasmic reticulum Ca2+. Clin. Exp. Pharmacol. Physiol. 2015, 42, 720–726. [Google Scholar] [CrossRef]
  19. Wehrens, X.H.; Lehnart, S.E.; Huang, F.; Vest, J.A.; Reiken, S.R.; Mohler, P.J.; Sun, J.; Guatimosim, S.; Song, L.-S.; Rosemblit, N.; et al. FKBP12.6 Deficiency and Defective Calcium Release Channel (Ryanodine Receptor) Function Linked to Exercise-Induced Sudden Cardiac Death. Cell 2003, 113, 829–840. [Google Scholar] [CrossRef]
  20. Wleklinski, M.J.; Kannankeril, P.J.; Knollmann, B.C. Molecular and tissue mechanisms of catecholaminergic polymorphic ventricular tachycardia. J. Physiol. 2020, 598, 2817–2834. [Google Scholar] [CrossRef]
  21. Sleiman, Y.; Lacampagne, A.; Meli, A.C. “Ryanopathies” and RyR2 dysfunctions: Can we further decipher them using in vitro human disease models? Cell Death Dis. 2021, 12, 1–19. [Google Scholar] [CrossRef] [PubMed]
  22. Olubando, D.; Hopton, C.; Eden, J.; Caswell, R.; Thomas, N.L.; Roberts, S.A.; Morris-Rosendahl, D.; Venetucci, L.; Newman, W.G. Classification and correlation of RYR2 missense variants in individuals with catecholaminergic polymorphic ventricular tachycardia reveals phenotypic relationships. J. Hum. Genet. 2020, 65, 531–539. [Google Scholar] [CrossRef] [PubMed]
  23. Rossi, D.; Gamberucci, A.; Pierantozzi, E.; Amato, C.; Migliore, L.; Sorrentino, V. Calsequestrin, a key protein in striated muscle health and disease. J. Muscle Res. Cell Motil. 2020, 42, 267–279. [Google Scholar] [CrossRef] [PubMed]
  24. Gray, B.; Bagnall, R.D.; Lam, L.; Ingles, J.; Turner, C.; Haan, E.; Davis, A.; Yang, P.-C.; Clancy, C.E.; Sy, R.W.; et al. A novel heterozygous mutation in cardiac calsequestrin causes autosomal dominant catecholaminergic polymorphic ventricular tachycardia. Heart Rhythm 2016, 13, 1652–1660. [Google Scholar] [CrossRef]
  25. 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]
  26. Perry, M.D.; I Vandenberg, J. TECRL: Connecting sequence to consequence for a new sudden cardiac death gene. EMBO Mol. Med. 2016, 8, 1364–1365. [Google Scholar] [CrossRef]
  27. Devalla, H.D.; Gélinas, R.; Aburawi, E.H.; Beqqali, A.; Goyette, P.; Freund, C.; Chaix, M.; Tadros, R.; Jiang, H.; Le Béchec, A.; et al. TECRL, a new life-threatening inherited arrhythmia gene associated with overlapping clinical features of both LQTS and CPVT. EMBO Mol. Med. 2016, 8, 1390–1408. [Google Scholar] [CrossRef]
  28. Bhuiyan, Z.A.; Hamdan, M.A.; Shamsi, E.T.; Postma, A.V.; Mannens, M.M.; Wilde, A.A.M.; Al-Gazali, L. A Novel Early Onset Lethal Form of Catecholaminergic Polymorphic Ventricular Tachycardia Maps to Chromosome 7p14-p22. J. Cardiovasc. Electrophysiol. 2007, 18, 1060–1066. [Google Scholar] [CrossRef]
  29. Jaouadi, H.; Bouyacoub, Y.; Chabrak, S.; Kraoua, L.; Zaroui, A.; Elouej, S.; Nagara, M.; Dallali, H.; Delague, V.; Levy, N.; et al. Multiallelic rare variants support an oligogenic origin of sudden cardiac death in the young. Herz 2020, 46, 94–102. [Google Scholar] [CrossRef]
  30. Webster, G.; Aburawi, E.H.; A Chaix, M.; Chandler, S.; Foo, R.; Islam, A.K.M.M.; A E Kammeraad, J.; Rioux, J.D.; Al-Gazali, L.; Sayeed, Z.; et al. Life-threatening arrhythmias with autosomal recessive TECRL variants. EP Eur. 2020, 23, 781–788. [Google Scholar] [CrossRef]
  31. Msc, A.M.; Marschall, C.; Müntjes, C.; Schönecker, A.; Schuessler-Hahn, F.; Hohendanner, F.; Parwani, A.S.; Boldt, L.; Ott, C.; Bennewiz, A.; et al. Novel variants in TECRL cause recessive inherited CPVT type 3 with severe and variable clinical symptoms. J. Cardiovasc. Electrophysiol. 2020, 31, 1527–1535. [Google Scholar] [CrossRef]
  32. Jaouadi, H.; Bouyacoub, Y.; Chabrak, S.; Elouej, S.; Delague, V.; Levy, N.; Nagara, M.; Dallali, H.; Delague, V.; Levy, N.; et al. An oligogenic inheritance pattern in a Tunisian family with sudden cardiac death in the young n.d.:19. Herz 2021, 46 (Suppl. 1), 94–102. [Google Scholar] [CrossRef]
  33. Chopra, N.; Knollmann, B.C. Triadin regulates cardiac muscle couplon structure and microdomain Ca2+ signalling: A path towards ventricular arrhythmias. Cardiovasc. Res. 2013, 98, 187–191. [Google Scholar] [CrossRef]
  34. Landstrom, A.P.; Dobrev, D.; Wehrens, X.H. Calcium Signaling and Cardiac Arrhythmias. Circ. Res. 2017, 120, 1969–1993. [Google Scholar] [CrossRef]
  35. Roux-Buisson, N.; Cacheux, M.; Fourest-Lieuvin, A.; Fauconnier, J.; Brocard, J.; Denjoy, I.; Durand, P.; Guicheney, P.; Kyndt, F.; Leenhardt, A.; et al. Absence of triadin, a protein of the calcium release complex, is responsible for cardiac arrhythmia with sudden death in human. Hum. Mol. Genet. 2012, 21, 2759–2767. [Google Scholar] [CrossRef]
  36. Rooryck, C.; Kyndt, F.; Bozon, D.; Roux-Buisson, N.; Sacher, F.; Probst, V.; Thambo, J.-B. New Family With Catecholaminergic Polymorphic Ventricular Tachycardia Linked to the Triadin Gene. J. Cardiovasc. Electrophysiol. 2015, 26, 1146–1150. [Google Scholar] [CrossRef]
  37. Crotti, L.; Spazzolini, C.; Tester, D.J.; Ghidoni, A.; Baruteau, A.-E.; Beckmann, B.-M.; Behr, E.R.; Bennett, J.S.; Bezzina, C.R.; A Bhuiyan, Z.; et al. Calmodulin mutations and life-threatening cardiac arrhythmias: Insights from the International Calmodulinopathy Registry. Eur. Heart J. 2019, 40, 2964–2975. [Google Scholar] [CrossRef]
  38. Nyegaard, M.; Overgaard, M.T.; Søndergaard, M.T.; Vranas, M.; Behr, E.R.; Hildebrandt, L.L.; Lund, J.; Hedley, P.L.; Camm, A.J.; Wettrell, G.; et al. Mutations in Calmodulin Cause Ventricular Tachycardia and Sudden Cardiac Death. Am. J. Hum. Genet. 2012, 91, 703–712. [Google Scholar] [CrossRef]
  39. Makita, N.; Yagihara, N.; Crotti, L.; Johnson, C.N.; Beckmann, B.-M.; Roh, M.S.; Shigemizu, D.; Lichtner, P.; Ishikawa, T.; Aiba, T.; et al. Novel Calmodulin Mutations Associated With Congenital Arrhythmia Susceptibility. Circ. Cardiovasc. Genet. 2014, 7, 466–474. [Google Scholar] [CrossRef]
  40. Gomez-Hurtado, N.; Boczek, N.J.; Kryshtal, D.O.; Johnson, C.N.; Sun, J.; Nitu, F.R.; Cornea, R.L.; Chazin, W.J.; Calvert, M.L.; Tester, D.J.; et al. Novel CPVT-Associated Calmodulin Mutation in CALM3 (CALM3-A103V) Activates Arrhythmogenic Ca Waves and Sparks. Circ. Arrhythmia Electrophysiol. 2016, 9. [Google Scholar] [CrossRef] [Green Version]
  41. Hwang, H.S.; Nitu, F.; Yang, Y.; Walweel, K.; Pereira, L.; Johnson, C.N.; Faggioni, M.; Chazin, W.J.; Laver, D.; George, A.L.; et al. Divergent regulation of ryanodine receptor 2 calcium release channels by arrhythmogenic human calmodulin missense mutants. Circ. Res. 2014, 114, 1114–1124. [Google Scholar] [CrossRef] [PubMed]
  42. Alcalde, M.; Campuzano, O.; Berne, P.; García-Pavía, P.; Doltra, A.; Arbelo, E.; Sarquella-Brugada, G.; Iglesias, A.; Alonso-Pulpón, L.; Brugada, J.; et al. Stop-Gain Mutations in PKP2 Are Associated with a Later Age of Onset of Arrhythmogenic Right Ventricular Cardiomyopathy. PLoS ONE 2014, 9, e100560. [Google Scholar] [CrossRef] [PubMed]
  43. Cerrone, M.; Montnach, J.; Lin, X.; Zhao, Y.-T.; Zhang, M.; Agullo-Pascual, E.; Leo-Macias, A.; Alvarado, F.J.; Dolgalev, I.; Karathanos, T.V.; et al. Plakophilin-2 is required for transcription of genes that control calcium cycling and cardiac rhythm. Nat. Commun. 2017, 8, 1–16. [Google Scholar] [CrossRef] [PubMed]
  44. Tester, D.J.; Ackerman, J.P.; Giudicessi, J.R.; Ackerman, N.C.; Cerrone, M.; Delmar, M.; Ackerman, M.J. Plakophilin-2 Truncation Variants in Patients Clinically Diagnosed With Catecholaminergic Polymorphic Ventricular Tachycardia and Decedents With Exercise-Associated Autopsy Negative Sudden Unexplained Death in the Young. JACC Clin. Electrophysiol. 2018, 5, 120–127. [Google Scholar] [CrossRef]
  45. Mohler, P.J.; Schott, J.-J.; Gramolini, A.O.; Dilly, K.W.; Guatimosim, S.; Dubell, W.H.; Song, L.-S.; Haurogné, K.; Kyndt, F.; Ali, M.E.; et al. Ankyrin-B mutation causes type 4 long-QT cardiac arrhythmia and sudden cardiac death. Nature 2003, 421, 634–639. [Google Scholar] [CrossRef]
  46. Mohler, P.J.; Davis, J.Q.; Bennett, V. Ankyrin-B Coordinates the Na/K ATPase, Na/Ca Exchanger, and InsP3 Receptor in a Cardiac T-Tubule/SR Microdomain. PLoS Biol. 2005, 3, e423. [Google Scholar] [CrossRef]
  47. Mohler, P.J.; Splawski, I.; Napolitano, C.; Bottelli, G.; Sharpe, L.; Timothy, K.; Priori, S.G.; Keating, M.T.; Bennett, V. A cardiac arrhythmia syndrome caused by loss of ankyrin-B function. Proc. Natl. Acad. Sci. USA 2004, 101, 9137–9142. [Google Scholar] [CrossRef]
  48. Mohler, P.J.; Le Scouarnec, S.; Denjoy, I.; Lowe, J.S.; Guicheney, P.; Caron, L.; Driskell, I.M.; Schott, J.-J.; Norris, K.; Leenhardt, A.; et al. Defining the Cellular Phenotype of “Ankyrin-B Syndrome” Variants. Circulation 2007, 115, 432–441. [Google Scholar] [CrossRef]
  49. Seidel, M.; de Meritens, C.R.; Johnson, L.; Parthimos, D.; Bannister, M.; Thomas, N.L.; Ozekhome-Mike, E.; Lai, F.A.; Zissimopoulos, S. Identification of an amino-terminus determinant critical for ryanodine receptor/Ca2+ release channel function. Cardiovasc. Res. 2020, 117, 780–791. [Google Scholar] [CrossRef]
  50. Wescott, A.P.; Jafri, M.S.; Lederer, W.; Williams, G.S. Ryanodine receptor sensitivity governs the stability and synchrony of local calcium release during cardiac excitation-contraction coupling. J. Mol. Cell. Cardiol. 2016, 92, 82–92. [Google Scholar] [CrossRef] [Green Version]
  51. Wei, L.; Hanna, A.D.; Beard, N.A.; Dulhunty, A.F. Unique isoform-specific properties of calsequestrin in the heart and skeletal muscle. Cell Calcium 2009, 45, 474–484. [Google Scholar] [CrossRef]
  52. Guo, A.; Cala, S.E.; Song, L.-S. Calsequestrin Accumulation in Rough Endoplasmic Reticulum Promotes Perinuclear Ca2+ Release. J. Biol. Chem. 2012, 287, 16670–16680. [Google Scholar] [CrossRef]
  53. Shen, X.; Franzini-Armstrong, C.; Lopez, J.R.; Jones, L.R.; Kobayashi, Y.M.; Wang, Y.; Kerrick, W.G.L.; Caswell, A.H.; Potter, J.D.; Miller, T.; et al. Triadins Modulate Intracellular Ca2+ Homeostasis but Are Not Essential for Excitation-Contraction Coupling in Skeletal Muscle. J. Biol. Chem. 2007, 282, 37864–37874. [Google Scholar] [CrossRef]
  54. Zou, J.; Salarian, M.; Chen, Y.; Zhuo, Y.; Brown, N.E.; Hepler, J.R.; Yang, J.J. Direct visualization of interaction between calmodulin and connexin. Biochem. J. 2017, 474, 4035–4051. [Google Scholar] [CrossRef]
  55. Haßdenteufel, S.; Schäuble, N.; Cassella, P.; Leznicki, P.; Müller, A.; High, S.; Jung, M.; Zimmermann, R. Ca2+-calmodulin inhibits tail-anchored protein insertion into the mammalian endoplasmic reticulum membrane. FEBS Lett. 2011, 585, 3485–3490. [Google Scholar] [CrossRef]
  56. Jiménez-Jáimez, J.; Doza, J.P.; Ortega, Á.; Macías-Ruiz, R.; Perin, F.; Del Rey, M.M.R.-V.; Ortiz-Genga, M.; Monserrat, L.; Barriales-Villa, R.; Blanca, E.; et al. Calmodulin 2 Mutation N98S Is Associated with Unexplained Cardiac Arrest in Infants Due to Low Clinical Penetrance Electrical Disorders. PLoS ONE 2016, 11, e0153851. [Google Scholar] [CrossRef]
  57. Reed, G.J.; Boczek, N.J.; Etheridge, S.P.; Ackerman, M.J. CALM3 mutation associated with long QT syndrome. Heart Rhythm 2014, 12, 419–422. [Google Scholar] [CrossRef]
  58. Glukhov, A.V.; Fedorov, V.V.; Anderson, M.E.; Mohler, P.J.; Efimov, I.R. Functional anatomy of the murine sinus node: High-resolution optical mapping of ankyrin-B heterozygous mice. Am. J. Physiol. Circ. Physiol. 2010, 299, H482–H491. [Google Scholar] [CrossRef]
  59. Li, J.; Kline, C.F.; Hund, T.J.; Anderson, M.E.; Mohler, P.J. Ankyrin-B Regulates Kir6.2 Membrane Expression and Function in Heart. J. Biol. Chem. 2010, 285, 28723–28730. [Google Scholar] [CrossRef]
  60. Sedlacek, K.; Stark, K.; Cunha, S.R.; Pfeufer, A.; Weber, S.; Berger, I.; Perz, S.; Kääb, S.; Wichmann, H.-E.; Mohler, P.J.; et al. Common Genetic Variants in ANK2 Modulate QT Interval. Circ. Cardiovasc. Genet. 2008, 1, 93–99. [Google Scholar] [CrossRef] [Green Version]
  61. Elliott, P.M.; Anastasakis, A.; Borger, M.A.; Borggrefe, M.; Cecchi, F.; Charron, P.; Hagege, A.A.; Lafont, A.; Limongelli, G.; Mahrholdt, H.; et al. 2014 ESC Guidelines on diagnosis and management of hypertrophic cardiomyopathy: The Task Force for the Diagnosis and Management of Hypertrophic Cardiomyopathy of the European Society of Cardiology (ESC). Eur. Heart J. 2014, 35, 2733–2779. [Google Scholar] [CrossRef] [PubMed]
  62. Sabater-Molina, M.; Pérez-Sánchez, I.; Hernandez Del Rincon, J.P.; Gimeno, J.R. Genetics of hypertrophic cardiomyopathy: A review of current state. Clin. Genet. 2018, 93, 3–14. [Google Scholar] [CrossRef]
  63. Maron, B.J.; Maron, M.S.; Semsarian, C. Double or compound sarcomere mutations in hypertrophic cardiomyopathy: A potential link to sudden death in the absence of conventional risk factors. Heart Rhythm 2012, 9, 57–63. [Google Scholar] [CrossRef] [PubMed]
  64. Geisterfer-Lowrance, A.A.; Kass, S.; Tanigawa, G.; Vosberg, H.-P.; McKenna, W.; Seidman, C.E.; Seidman, J. A molecular basis for familial hypertrophic cardiomyopathy: A β cardiac myosin heavy chain gene missense mutation. Cell 1990, 62, 999–1006. [Google Scholar] [CrossRef]
  65. Seidman, J.; Seidman, C. The Genetic Basis for Cardiomyopathy: From Mutation Identification to Mechanistic Paradigms. Cell 2001, 104, 557–567. [Google Scholar] [CrossRef] [PubMed]
  66. Alfares, A.A.; Kelly, M.A.; McDermott, G.; Funke, B.H.; Lebo, M.S.; Baxter, S.B.; Shen, J.; McLaughlin, H.M.; Clark, E.H.; Babb, L.J.; et al. Results of clinical genetic testing of 2,912 probands with hypertrophic cardiomyopathy: Expanded panels offer limited additional sensitivity. Anesth. Analg. 2015, 17, 880–888. [Google Scholar] [CrossRef]
  67. Landstrom, A.P.; Ackerman, M.J. Beyond the Cardiac Myofilament: Hypertrophic Cardiomyopathy- Associated Mutations in Genes that Encode Calcium-Handling Proteins. Curr. Mol. Med. 2012, 12, 507–518. [Google Scholar] [CrossRef]
  68. Robinson, P.; Liu, X.; Sparrow, A.; Patel, S.; Zhang, Y.-H.; Casadei, B.; Watkins, H.; Redwood, C. Hypertrophic cardiomyopathy mutations increase myofilament Ca2+ buffering, alter intracellular Ca2+ handling, and stimulate Ca2+-dependent signaling. J. Biol. Chem. 2018, 293, 10487–10499. [Google Scholar] [CrossRef]
  69. Fraysse, B.; Weinberger, F.; Bardswell, S.C.; Cuello, F.; Vignier, N.; Geertz, B.; Starbatty, J.; Krämer, E.; Coirault, C.; Eschenhagen, T.; et al. Increased myofilament Ca2+ sensitivity and diastolic dysfunction as early consequences of Mybpc3 mutation in heterozygous knock-in mice. J. Mol. Cell. Cardiol. 2012, 52, 1299–1307. [Google Scholar] [CrossRef]
  70. Hoffmann, B.; Schmidt-Traub, H.; Perrot, A.; Osterziel, K.J.; Geßner, R. First mutation in cardiac troponin C, L29Q, in a patient with hypertrophic cardiomyopathy. Hum. Mutat. 2001, 17, 524. [Google Scholar] [CrossRef]
  71. Helms, A.S.; Alvarado, F.J.; Yob, J.; Tang, V.T.; Pagani, F.; Russell, M.W.; Valdivia, H.H.; Day, S.M. Genotype-Dependent and -Independent Calcium Signaling Dysregulation in Human Hypertrophic Cardiomyopathy. Circulation 2016, 134, 1738–1748. [Google Scholar] [CrossRef]
  72. Gordon, A.M.; Homsher, E.; Regnier, M. Regulation of Contraction in Striated Muscle. Physiol. Rev. 2000, 80, 853–924. [Google Scholar] [CrossRef]
  73. Landstrom, A.P.; Parvatiyar, M.S.; Pinto, J.R.; Marquardt, M.L.; Bos, J.M.; Tester, D.J.; Ommen, S.R.; Potter, J.D.; Ackerman, M.J. Molecular and functional characterization of novel hypertrophic cardiomyopathy susceptibility mutations in TNNC1-encoded troponin C. J. Mol. Cell. Cardiol. 2008, 45, 281–288. [Google Scholar] [CrossRef]
  74. Robertson, I.M.; Sevrieva, I.; Li, M.X.; Irving, M.; Sun, Y.-B.; Sykes, B.D. The structural and functional effects of the familial hypertrophic cardiomyopathy-linked cardiac troponin C mutation, L29Q. J. Mol. Cell. Cardiol. 2015, 87, 257–269. [Google Scholar] [CrossRef]
  75. Chung, W.K.; Kitner, C.; Maron, B.J. Novel frameshift mutation in Troponin C (TNNC1) associated with hypertrophic cardiomyopathy and sudden death. Cardiol. Young 2011, 21, 345–348. [Google Scholar] [CrossRef]
  76. Parvatiyar, M.; Landstrom, A.; Figueiredo-Freitas, C.; Potter, J.D.; Ackerman, M.J.; Pinto, J.R. A Mutation in TNNC1-encoded Cardiac Troponin C, TNNC1-A31S, Predisposes to Hypertrophic Cardiomyopathy and Ventricular Fibrillation. J. Biol. Chem. 2012, 287, 31845–31855. [Google Scholar] [CrossRef]
  77. Jaafar, N.; Girolami, F.; Zairi, I.; Kraiem, S.; Hammami, M.; Olivotto, I. Genetic profile of hypertrophic cardiomyopathy in Tunisia: Is it different? Glob. Cardiol. Sci. Pract. 2015, 2015, 16. [Google Scholar] [CrossRef] [PubMed]
  78. Lindhout, D.A.; Sykes, B.D. Structure and Dynamics of the C-domain of Human Cardiac Troponin C in Complex with the Inhibitory Region of Human Cardiac Troponin I. J. Biol. Chem. 2003, 278, 27024–27034. [Google Scholar] [CrossRef] [PubMed]
  79. Fujino, N.; Ino, H.; Hayashi, K.; Uchiyama, K.; Nagata, M.; Konno, T.; Katoh, H.; Sakamoto, Y.; Tsubokawa, T.; Ohsato, K.; et al. Abstract 915: A Novel Missense Mutation in Cardiac Ryanodine Receptor Gene as a Possible Cause of Hypertrophic Cardiomyopathy: Evidence From Familial Analysis. Circulation 2006, 114, II_165. [Google Scholar] [CrossRef]
  80. Landstrom, A.P.; Dailey-Schwartz, A.L.; Rosenfeld, J.A.; Yang, Y.; McLean, M.J.; Miyake, C.Y.; Valdes, S.O.; Fan, Y.; Allen, H.D.; Penny, D.J.; et al. Interpreting Incidentally Identified Variants in Genes Associated With Catecholaminergic Polymorphic Ventricular Tachycardia in a Large Cohort of Clinical Whole-Exome Genetic Test Referrals. Circ. Arrhythmia Electrophysiol. 2017, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Medeiros-Domingo, A.; Bhuiyan, Z.A.; Tester, D.J.; Hofman, N.; Bikker, H.; van Tintelen, J.P.; Mannens, M.M.; Wilde, A.A.; Ackerman, M.J. The RYR2-Encoded Ryanodine Receptor/Calcium Release Channel in Patients Diagnosed Previously With Either Catecholaminergic Polymorphic Ventricular Tachycardia or Genotype Negative, Exercise-Induced Long QT Syndrome: A Comprehensive Open Reading Frame Mutational Analysis. J. Am. Coll. Cardiol. 2009, 54, 2065–2074. [Google Scholar] [CrossRef] [PubMed]
  82. Alvarado, F.J.; Bos, J.M.; Yuchi, Z.; Valdivia, C.R.; Hernández, J.J.; Zhao, Y.-T.; Henderlong, D.S.; Chen, Y.; Booher, T.R.; Marcou, C.A.; et al. Cardiac hypertrophy and arrhythmia in mice induced by a mutation in ryanodine receptor. J. Clin. Investig. 2019, 4. [Google Scholar] [CrossRef]
  83. Hosoda, T.; Monzen, K.; Hiroi, Y.; Oka, T.; Takimoto, E.; Yazaki, Y.; Nagai, R.; Komuro, I. A Novel Myocyte-specific Gene MidoriPromotes the Differentiation of P19CL6 Cells into Cardiomyocytes. J. Biol. Chem. 2001, 276, 35978–35989. [Google Scholar] [CrossRef]
  84. Almomani, R.; Verhagen, J.M.; Herkert, J.C.; Brosens, E.; van Spaendonck-Zwarts, K.Y.; Asimaki, A.; van der Zwaag, P.A.; Frohn-Mulder, I.M.; Bertoli-Avella, A.M.; Boven, L.G.; et al. Biallelic Truncating Mutations in ALPK3 Cause Severe Pediatric Cardiomyopathy. J. Am. Coll. Cardiol. 2016, 67, 515–525. [Google Scholar] [CrossRef]
  85. Agarwal, R.; Wakimoto, H.; Paulo, J.A.; Zhang, Q.; Reichart, D.; Toepfer, C.; Sharma, A.; Tai, A.C.; Lun, M.; Gorham, J.; et al. Pathogenesis of Cardiomyopathy Caused by Variants in ALPK3, an Essential Pseudokinase in the Cardiomyocyte Nucleus and Sarcomere. Circulation 2022, 146, 1674–1693. [Google Scholar] [CrossRef]
  86. Çağlayan, A.O.; Sezer, R.G.; Kaymakçalan, H.; Ulgen, E.; Yavuz, T.; Baranoski, J.F.; Bozaykut, A.; Harmanci, A.S.; Yalcin, Y.; Youngblood, M.W.; et al. ALPK3 gene mutation in a patient with congenital cardiomyopathy and dysmorphic features. Mol. Case Stud. 2017, 3, a001859. [Google Scholar] [CrossRef]
  87. Jaouadi, H.; Kraoua, L.; Chaker, L.; Atkinson, A.; Delague, V.; Levy, N.; Benkhalifa, R.; Mrad, R.; Abdelhak, S.; Zaffran, S. Novel ALPK3 mutation in a Tunisian patient with pediatric cardiomyopathy and facio-thoraco-skeletal features. J. Hum. Genet. 2018, 63, 1077–1082. [Google Scholar] [CrossRef]
  88. Phelan, D.G.; Anderson, D.J.; Howden, S.; Wong, R.C.-B.; Hickey, P.; Pope, K.; Wilson, G.R.; Pébay, A.; Davis, A.M.; Petrou, S.; et al. ALPK3-deficient cardiomyocytes generated from patient-derived induced pluripotent stem cells and mutant human embryonic stem cells display abnormal calcium handling and establish that ALPK3 deficiency underlies familial cardiomyopathy. Eur. Heart J. 2016, 37, 2586–2590. [Google Scholar] [CrossRef]
  89. Walsh, R.; Bezzina, C.R. ALPK3: A full spectrum cardiomyopathy gene? Eur. Heart J. 2021, 42, 3074–3077. [Google Scholar] [CrossRef]
  90. Lopes, L.R.; Garcia-Hernández, S.; Lorenzini, M.; Futema, M.; Chumakova, O.; Zateyshchikov, D.; Isidoro-Garcia, M.; Villacorta, E.; Escobar-Lopez, L.; Garcia-Pavia, P.; et al. Alpha-protein kinase 3 (ALPK3) truncating variants are a cause of autosomal dominant hypertrophic cardiomyopathy. Eur. Heart J. 2021, 42, 3063–3073. [Google Scholar] [CrossRef]
  91. Van Sligtenhorst, I.; Ding, Z.-M.; Shi, Z.-Z.; Read, R.W.; Hansen, G.; Vogel, P. Cardiomyopathy in α-Kinase 3 (ALPK3)–Deficient Mice. Vet. Pathol. 2011, 49, 131–141. [Google Scholar] [CrossRef]
  92. Marian, A. Molecular Genetic Basis of Hypertrophic Cardiomyopathy. Circ. Res. 2021, 128, 1533–1553. [Google Scholar] [CrossRef] [PubMed]
  93. Garbino, A.; Van Oort, R.J.; Dixit, S.S.; Landstrom, A.P.; Ackerman, M.J.; Wehrens, X.H.T. Molecular evolution of the junctophilin gene family. Physiol. Genom. 2009, 37, 175–186. [Google Scholar] [CrossRef] [PubMed]
  94. Ziman, A.P.; Gómez-Viquez, N.L.; Bloch, R.J.; Lederer, W. Excitation–contraction coupling changes during postnatal cardiac development. J. Mol. Cell. Cardiol. 2010, 48, 379–386. [Google Scholar] [CrossRef]
  95. Landstrom, A.P.; Weisleder, N.; Batalden, K.B.; Bos, J.M.; Tester, D.J.; Ommen, S.R.; Wehrens, X.H.; Claycomb, W.C.; Ko, J.-K.; Hwang, M.; et al. Mutations in JPH2-encoded junctophilin-2 associated with hypertrophic cardiomyopathy in humans. J. Mol. Cell. Cardiol. 2007, 42, 1026–1035. [Google Scholar] [CrossRef]
  96. Oxenoid, K.; Chou, J.J. The structure of phospholamban pentamer reveals a channel-like architecture in membranes. Proc. Natl. Acad. Sci. USA 2005, 102, 10870–10875. [Google Scholar] [CrossRef]
  97. Gustavsson, M.; Traaseth, N.J.; Veglia, G. Activating and Deactivating Roles of Lipid Bilayers on the Ca2+-ATPase/Phospholamban Complex. Biochemistry 2011, 50, 10367–10374. [Google Scholar] [CrossRef] [PubMed]
  98. Koss, K.L.; Kranias, E.G. Phospholamban: A Prominent Regulator of Myocardial Contractility. Circ. Res. 1996, 79, 1059–1063. [Google Scholar] [CrossRef]
  99. Haghighi, K.; Kolokathis, F.; Pater, L.; Lynch, R.A.; Asahi, M.; Gramolini, A.O.; Fan, G.-C.; Tsiapras, D.; Hahn, H.S.; Adamopoulos, S.; et al. Human phospholamban null results in lethal dilated cardiomyopathy revealing a critical difference between mouse and human. J. Clin. Investig. 2003, 111, 869–876. [Google Scholar] [CrossRef]
  100. Medin, M.; Hermida-Prieto, M.; Monserrat, L.; Laredo, R.; Rodríguez-Rey, J.C.; Fernandez, X.; Castro-Beiras, A. Mutational screening of phospholamban gene in hypertrophic and idiopathic dilated cardiomyopathy and functional study of the PLN -42 C>G mutation. Eur. J. Heart Fail. 2007, 9, 37–43. [Google Scholar] [CrossRef]
  101. Minamisawa, S.; Sato, Y.; Tatsuguchi, Y.; Fujino, T.; Imamura, S.-I.; Uetsuka, Y.; Nakazawa, M.; Matsuoka, R. Mutation of the phospholamban promoter associated with hypertrophic cardiomyopathy. Biochem. Biophys. Res. Commun. 2003, 304, 1–4. [Google Scholar] [CrossRef] [PubMed]
  102. Landstrom, A.P.; Adekola, B.A.; Bos, J.M.; Ommen, S.R.; Ackerman, M.J. PLN-encoded phospholamban mutation in a large cohort of hypertrophic cardiomyopathy cases: Summary of the literature and implications for genetic testing. Am. Heart J. 2011, 161, 165–171. [Google Scholar] [CrossRef]
  103. Chiu, C.L.; Tebo, M.; Ingles, J.; Yeates, L.; Arthur, J.W.; Lind, J.M.; Semsarian, C. Genetic screening of calcium regulation genes in familial hypertrophic cardiomyopathy. J. Mol. Cell. Cardiol. 2007, 43, 337–343. [Google Scholar] [CrossRef]
  104. Ding, Z.; Yuan, J.; Liang, Y.; Wu, J.; Gong, H.; Ye, Y.; Jiang, G.; Yin, P.; Li, Y.; Zhang, G.; et al. Ryanodine Receptor Type 2 Plays a Role in the Development of Cardiac Fibrosis under Mechanical Stretch Through TGFβ-1. Int. Heart J. 2017, 58, 957–961. [Google Scholar] [CrossRef]
  105. Zou, Y.; Liang, Y.; Gong, H.; Zhou, N.; Ma, H.; Guan, A.; Sun, A.; Wang, P.; Niu, Y.; Jiang, H.; et al. Ryanodine Receptor Type 2 Is Required for the Development of Pressure Overload-Induced Cardiac Hypertrophy. Hypertension 2011, 58, 1099–1110. [Google Scholar] [CrossRef]
  106. Norman, C.; Rall, J.A.; Tikunova, S.B.; Davis, J.P. Modulation of the rate of cardiac muscle contraction by troponin C constructs with various calcium binding affinities. Am. J. Physiol. Circ. Physiol. 2007, 293, H2580–H2587. [Google Scholar] [CrossRef]
  107. Liu, B.; Tikunova, S.B.; Kline, K.P.; Siddiqui, J.K.; Davis, J.P. Disease-Related Cardiac Troponins Alter Thin Filament Ca2+ Association and Dissociation Rates. PLoS ONE 2012, 7, e38259. [Google Scholar] [CrossRef] [PubMed]
  108. Cordina, N.; Liew, C.K.; Gell, D.; Fajer, P.G.; Mackay, J.; Brown, L.J. Effects of Calcium Binding and the Hypertrophic Cardiomyopathy A8V Mutation on the Dynamic Equilibrium between Closed and Open Conformations of the Regulatory N-Domain of Isolated Cardiac Troponin C. Biochemistry 2013, 52, 1950–1962. [Google Scholar] [CrossRef] [PubMed]
  109. Yamazaki, D.; Yamazaki, T.; Takeshima, H. New molecular components supporting ryanodine receptor-mediated Ca2+ release: Roles of junctophilin and TRIC channel in embryonic cardiomyocytes. Pharmacol. Ther. 2009, 121, 265–272. [Google Scholar] [CrossRef]
  110. Vanninen, S.U.M.; Leivo, K.; Seppälä, E.H.; Aalto-Setälä, K.; Pitkänen, O.; Suursalmi, P.; Annala, A.-P.; Anttila, I.; Alastalo, T.-P.; Myllykangas, S.; et al. Heterozygous junctophilin-2 (JPH2) p.(Thr161Lys) is a monogenic cause for HCM with heart failure. PLoS ONE 2018, 13, e0203422. [Google Scholar] [CrossRef]
  111. Landstrom, A.P.; Kellen, C.A.; Dixit, S.S.; van Oort, R.J.; Garbino, A.; Weisleder, N.; Ma, J.; Wehrens, X.H.; Ackerman, M.J. Junctophilin-2 Expression Silencing Causes Cardiocyte Hypertrophy and Abnormal Intracellular Calcium-Handling. Circ. Heart Fail. 2011, 4, 214–223. [Google Scholar] [CrossRef] [PubMed]
  112. Chen, L.T.L.; Yao, Q.; Soares, T.; Squier, T.C.; Bigelow, D.J. Phospholamban Modulates the Functional Coupling between Nucleotide Domains in Ca-ATPase Oligomeric Complexes in Cardiac Sarcoplasmic Reticulum. Biochemistry 2009, 48, 2411–2421. [Google Scholar] [CrossRef] [PubMed]
  113. Medeiros, A.; Biagi, D.G.; Sobreira, T.J.; de Oliveira, P.S.L.; Negrão, C.E.; Mansur, A.J.; Krieger, J.E.; Brum, P.C.; Pereira, A.C. Mutations in the human phospholamban gene in patients with heart failure. Am. Hear. J. 2011, 162, 1088–1095.e1. [Google Scholar] [CrossRef] [PubMed]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zaffran, S.; Kraoua, L.; Jaouadi, H. Calcium Handling in Inherited Cardiac Diseases: A Focus on Catecholaminergic Polymorphic Ventricular Tachycardia and Hypertrophic Cardiomyopathy. Int. J. Mol. Sci. 2023, 24, 3365. https://doi.org/10.3390/ijms24043365

AMA Style

Zaffran S, Kraoua L, Jaouadi H. Calcium Handling in Inherited Cardiac Diseases: A Focus on Catecholaminergic Polymorphic Ventricular Tachycardia and Hypertrophic Cardiomyopathy. International Journal of Molecular Sciences. 2023; 24(4):3365. https://doi.org/10.3390/ijms24043365

Chicago/Turabian Style

Zaffran, Stéphane, Lilia Kraoua, and Hager Jaouadi. 2023. "Calcium Handling in Inherited Cardiac Diseases: A Focus on Catecholaminergic Polymorphic Ventricular Tachycardia and Hypertrophic Cardiomyopathy" International Journal of Molecular Sciences 24, no. 4: 3365. https://doi.org/10.3390/ijms24043365

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