Abstract
Atrial fibrillation (AF) is a prevalent and complex arrhythmia for which the pathogenesis involves various electrophysiological factors, notably the regulation of calcium channels. This article aimed to investigate the specific roles and molecular mechanisms of the L-type and T-type calcium channels, ryanodine receptors (RyRs), inositol 1,4,5-triphosphate receptors (IP3Rs), calcium release-activated calcium (CRAC) channels, and transient receptor potential (TRP) channels in the pathogenesis and persistence of AF. In addition, this article reviews recent advances in calcium channel-targeted drugs from experimental and clinical studies, offering new insights into the relationship between calcium channel regulation and AF pathology. These findings suggest promising directions for further research into the mechanisms of AF and the development of targeted therapeutic strategies.
Keywords: atrial fibrillation, calcium channels, calcium homeostasis, electrophysiology, therapeutic targets
1. Introduction
Atrial fibrillation (AF) is one of the most prevalent arrhythmias globally, affecting over 60 million individuals [1]. AF markedly elevates the risk of stroke, heart failure, and various other cardiovascular complications, imposing a substantial burden on healthcare systems and generating significant socio-economic costs [2, 3] (Fig. 1). Despite the availability of multiple clinical treatment strategies, the underlying pathogenesis of AF remains complex and not fully understood. In particular, the regulatory role of calcium channels in AF pathophysiology has received increasing attention.
Fig. 1.
Two cardiac rhythms. (A) The left electrocardiogram shows sinus rhythm. The sinoatrial (SA) node, located in the right atrium, can spontaneously generate action potentials. The electrical impulses generated by the SA node are rapidly propagated to both the left and right atria, and then transmitted to the ventricles via the atrioventricular (AV) node. The impulses then spread rapidly throughout the ventricles via the Purkinje fiber system, leading to synchronized contraction of the ventricular muscle and the expulsion of blood. (B) The right electrocardiogram shows atrial fibrillation (AF). AF is typically triggered by spontaneous firing from ectopic pacemaker sites, with the most common trigger originating from the pulmonary veins. These ectopic pacemaker sites are influenced by various factors, which promote the initiation and maintenance of AF. Electrical impulses continuously circulate within the atria, resulting in rapid and irregular electrical activity. AF can lead to various complications. Fig. 1 was drawn using BioRender.
In the pathophysiology of AF, calcium channels play a crucial role in maintaining calcium homeostasis. There are several types of calcium channels. This article specifically addresses L-type calcium channels (LTCCs), T-type calcium channels (T-channels), ryanodine receptors (RyRs), inositol 1,4,5-triphosphate receptors (IP3Rs), calcium release activated calcium (CRAC) channels, and transient receptor potential (TRP) channels, all of which are closely associated with cardiac function [4, 5, 6, 7, 8, 9]. The regulation of calcium ion balance depends on the interaction and functional maintenance of these calcium channels; dysfunction in these channels is widely regarded as a critical mechanism underlying the pathogenesis of AF [10]. During AF, abnormal expression and function of calcium channels in atrial myocytes lead to intracellular calcium imbalance, further contributing to electrophysiological disturbances and structural remodeling in the atria [11, 12, 13, 14]. Research has demonstrated that dysregulation in calcium channels are closely associated with atrial fibrosis, inflammatory responses, and electrophysiological instability, which collectively promote the initiation and persistence of AF [15, 16].
Calcium channel blockers are widely used in the management of AF, as they help mitigate electrophysiological dysregulation and reduce AF frequency by lowering intracellular calcium concentrations in atrial myocytes [17]. In recent years, novel therapeutic strategies targeting calcium channels, including calcium channel inhibitors and agents that modulate calcium signaling pathways, have demonstrated promising potential [18, 19, 20]. Consequently, in-depth exploration of these mechanisms and the development of targeted treatment strategies are crucial to improving the prognosis for AF patients.
2. The Role of Calcium Channels in Cardiac Electrophysiological Activity
AF is a persistent or intermittent arrhythmia resulting from abnormal electrical activity within the atria [21]. Calcium homeostasis, the regulation of intracellular and extracellular calcium concentrations, is crucial for cardiac function and cellular signal transduction [22]. Calcium channels play an essential role in maintaining Ca2+ levels. T-channels open during the early depolarization and repolarization phases of cardiomyocytes, allowing a relatively small influx of calcium [23]. Upon depolarization of the cardiomyocyte membrane, LTCCs facilitate the entry of Ca2+ into the cell [24]. This increase in intracellular Ca2+ concentration activates RyR2 on the sarcoplasmic reticulum (SR), releasing stored Ca2+ and triggering calcium-induced calcium release (CICR). This process drives myocardial contraction, ensuring an adequate cardiac output to prevent ischemic injury [25].
Additionally, IP3 binds to IP3Rs in the SR, further increasing intracellular Ca2+ levels [26]. To maintain calcium homeostasis, the sarcoplasmic/endoplasmic reticulum Ca2+-ATPase (SERCA) actively transports excess intracellular calcium back into the SR, thereby reducing cytosolic calcium levels [27]. Simultaneously, the sodium-calcium exchanger (NCX) contributes to calcium balance by exchanging intracellular Ca2+ for extracellular Na⁺, thus regulating calcium concentrations [28]. When intracellular calcium levels decrease, CRAC channels facilitate the entry of extracellular calcium to replenish the intracellular calcium deficit [29]. Therefore, these calcium transport mechanisms ensure normal excitation-contraction coupling and maintain cardiac rhythm. In view of the central role of calcium influx in cardiac electrophysiology, it is necessary to further explore the specific functions of various calcium channels.
2.1 LTCCs
LTCCs are voltage-gated calcium channels (VGCCs) typically activated at membrane potentials between approximately –30 mV and –20 mV, with prolonged opening times that produce a stable calcium current [30, 31] (Fig. 2). Each LTCC consists of four subunits: 1, 2, , and . The 1 subunit forms the core of the channel, facilitating ion conduction and imparting selective permeability to calcium ions [32]. Based on the genes encoding the 1 subunit, LTCCs are classified into four types: Cav1.1, Cav1.2, Cav1.3, and Cav1.4. Of these, Cav1.2 and Cav1.3 are particularly important in cardiac electrophysiological activity [33].
Fig. 2.
The mechanism of calcium channels regulating Ca2+ in cardiomyocytes. GPCR regulates the level of intracellular and extracellular Ca2+ by activating G protein; calcium channels distributed in cell membranes and organelles regulate the balance of Ca2+. The binding of Stromal Interaction Molecule 1 (STIM1) on the sarcoplasmic reticulum (SR) with Orai Calcium Release-Activated Calcium Modulator 1 (ORAI1) on the cell membrane participates in the regulation of calcium ions. The red and green arrows indicate the directions of Ca2+ and Na⁺ movement, respectively. LTCC, L-type calcium channel; RyR2, ryanodine receptor 2; TRP channel, transient receptor potential channel; NCX, sodium-calcium exchanger; IP3R, inositol 1,4,5-triphosphate receptor; SERCA, sarcoplasmic/endoplasmic reticulum Ca2+-ATPase; PLB, phospholamban; ATP, adenosine triphosphate; P, phosphate; GPCR, G-Protein coupled receptor; Ror, receptor tyrosine kinase-like orphan receptor; Wnt5a, wnt family member 5A; FZD6, frizzled class receptor 6; PLC, phospholipase C; PIP2, phosphatidylinositol 4,5-bisphosphate; IP3, inositol trisphosphate; DAG, diacylglycerol; PKC, protein kinase C. Fig. 2 was drawn using BioRender.
Cav1.2 is the predominant LTCCs subtype in cardiomyocytes, is widely distributed in atrial and ventricular cells, and plays a central role in myocardial excitation-contraction coupling (ECC) [34] (Table 1, Ref. [9, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59]). During the plateau phase (phase 2) of the action potential, Cav1.2 is activated, allowing a substantial influx of calcium ions into the cell. This influx triggers CICR via RyR2 on the SR, leading to muscle contraction [60, 61]. The synergistic interaction of Cav1.2 with other ion channels, such as voltage-gated sodium channels, IP3Rs, and TRP channels, further enhances calcium influx, supporting a rapid myocardial response and robust contraction [62, 63, 64].
Table 1.
Subtypes of calcium channels and their kinetics and pharmacology.
Ca2+ channels | Activation potential | Modulators | Ref | |
Types | Subtypes | |||
LTCC channels | CaV1.1, CaV1.2, CaV1.3, CaV1.4 | –50 to –20 mV, Slow inactivation | Verapamil, Diltiazem, Nifedipine | [33, 34, 35] |
T-channels | CaV3.1, CaV3.2, CaV3.3 | –70 to –40 mV, Rapid inactivation | Suvecaltamide, Nickel (Ni2+) | [36, 37, 38] |
RyRs | RyR1, RyR2, RyR3 | –30 mV to –20 mV | Ryanodine | [39, 40, 41] |
IP3Rs | IP3R1, IP3R2, IP3R3 | NA | IP3, Heparin, 2-APB | [42, 43, 44] |
CRAC channels | Orai1, Orai2, Orai3, STIM1, STIM2 | NA | 2-APB | [45, 46, 47] |
TRPC channels | TRPC1, TRPC3, TRPC4, TRPC5, TRPC6, TRPC7 | –60 mV to +20 mV | SKF 96365 | [9, 44, 48, 49, 50, 51, 52] |
TRPM channels | TRPM1, TRPM2, TRPM3, TRPM4, TRPM5, TRPM6, TRPM7 | –50 mV | 9-Phenanthrol | [53, 54, 55, 56] |
TRPV channels | TRPV1, TRPV2, TRPV3, TRPV4 | –30 mV to –20 mV | Capsaicin | [57, 58] |
TRPP channels | TRPP1, TRPP2 | NA | Phenamil | [50, 59] |
Notes: “NA” represents missing data. 2-APB, 2-Aminoethoxydiphenyl borate; T-channels, T-type calcium channel; CRAC, calcium release activated calcium; RyRs, ryanodine receptors.
Cav1.3 is primarily expressed in the sinoatrial (SA) and atrioventricular (AV) nodes, where it regulates the heart’s automatic pacing and conduction functions [35] (Table 1). Unlike Cav1.2, Cav1.3 has a lower activation voltage threshold, typically between –40 mV and –50 mV, enabling it to play a crucial role in the automatic depolarization of SA node cells and to trigger pacing activity at lower voltages [35, 63, 65]. In the AV node, Cav1.3 facilitates signal transmission between the atria and ventricles, coordinating atrial and ventricular contractions to maintain cardiac rhythm [66]. During repolarization, Cav1.3 contributes to signal transduction between the atria and ventricles. Additionally, LTCCs work in conjunction with small-conductance calcium-activated potassium channels (KCa2.x), facilitating the return of the cell membrane potential to its resting state by modulating potassium channel activity [67].
2.2 T-channels
T-channels are low-voltage-activated channels (LVACs), typically activated at membrane potentials near –60 mV (Fig. 2). They exhibit brief opening times, generating transient calcium currents and rapid inactivation [31]. The core structure of T-channels consists of the 1 subunit, which determines their selectivity, conductivity, and voltage dependence for calcium ions [68]. Based on the genes encoding the 1 subunit, T-channels are classified into three subtypes: Cav3.1, Cav3.2, and Cav3.3, encoded by the CACNA1G, CACNA1H, and CACNA1I genes, respectively [69].
Cav3.1 is primarily distributed in the SA nodes, AV nodes, and Purkinje fibers, where it contributes to cardiac pacing and rhythm control [36] (Table 1). Cav3.1 can be activated at low voltages near the resting membrane potential, resulting in a small calcium influx that triggers early depolarization. This activity aids the self-regulating cells of the SA nodes in reaching the threshold potential to initiate an action potential and subsequently activate Cav1.3 [70]. The expression level and functional strength of Cav3.1 determine the depolarization rate of the SA node, thereby modulating heart rate [71]. In the AV nodes and Purkinje fibers, Cav3.1 regulates calcium influx to delay the conduction of electrical signals, ensuring that atrial contraction is completed before ventricular contraction. This delay promotes synchronization between the atria and ventricles, which facilitates the coordination of cardiac function [37].
Cav3.2 is highly expressed in embryonic pacemaker cells and has been shown to contribute to rhythmogenic activity during early cardiac development [38]. However, in the adult heart, the role of Cav3.2 in normal pacemaker function remains controversial. While Cav3.2 expression is detectable in the sinoatrial and AV nodes, its contribution to adult pacemaker control appears to be limited. Genetic deletion studies indicate that Cav3.1, rather than Cav3.2, is the predominant T-type calcium channel responsible for ICaT (T-type calcium current) in adult rhythmogenic centers, as its knockout abolishes nearly all T-type calcium currents in these regions [72]. Nonetheless, Cav3.2 is more sensitive to acidic environments and oxidative stress, which may contribute to its role under pathological conditions such as AF [73]. Under stress conditions, Cav3.2 can activate delayed rectifier potassium channels, including HERG and Kir2.1, thereby accelerating cell membrane repolarization and potentially facilitating arrhythmogenic activity [74].
Although T-channels are not the primary pathway for calcium influx, they play a significant role in ectopic pacemaker activity and the highly excitable autonomic activity at the pulmonary vein ostium [75]. Studies have shown that upregulation of Cav3.1 and Cav3.2 increases the instability of atrial depolarization and enhances triggered activity, thereby promoting the occurrence of AF [76].
2.3 RyRs
RyRs are calcium release channels located on the membranes of the endoplasmic and SR, where they play a critical role in regulating intracellular calcium levels and contribute to calcium signaling and myocardial contraction [39] (Fig. 2, Table 1). RyRs are composed of four identical subunits that assemble into a large tetrameric complex, featuring three primary functional domains: the N-terminal regulatory domain, the central hub region, and the C-terminal pore domain. The N-terminal regulatory domain contains binding sites for calcium, calmodulin, and adenosine triphosphate (ATP), which modulate the channel’s opening. The central hub region relays signals to the C-terminal pore domain, thereby enabling calcium release [41].
In the heart, RyRs exist in three isoforms: RyR1, RyR2, and RyR3, with RyR2 being the predominant subtype in cardiomyocytes, and is extensively distributed on the SR membrane [40]. Upon depolarization of the cardiomyocyte membrane, Cav1.2 channels open, allowing calcium ions to enter the cytosol. This influx triggers the activation of RyR2 and initiates CICR, releasing large amounts of calcium from the SR to stimulate myocardial fiber contraction [41]. The localized calcium release from RyR2, known as “calcium sparks”, synchronizes across multiple sites to generate a calcium wave, leading to global myocardial contraction [77]. RyR2 also interacts with potassium channels, such as large-conductance calcium-activated potassium (BKCa) channels, to regulate membrane potential and modulate contraction strength via calcium sparks [78].
After contraction, calcium ions are reabsorbed into the SR by the SERCA, while the NCX extrudes a portion of calcium from the cell to maintain calcium homeostasis [79]. Calcium release from RyR2 is finely regulated through a dual mechanism: low cytosolic calcium concentrations activate RyR2, whereas elevated calcium concentrations inhibit its opening via a negative feedback mechanism, preventing calcium overload and cellular damage [80]. Additionally, calmodulin (CaM) binds to RyR2 in response to high cytosolic calcium levels, further inhibiting channel opening to protect against calcium overload [81].
Under conditions of sympathetic activation, such as stress or physical exercise, RyR2-mediated calcium release is enhanced through phosphorylation by protein kinase A (PKA) and calcium/calmodulin-dependent protein kinase II (CaMKII), leading to increased myocardial contractility [82]. However, abnormal activation of RyR2, particularly pathological calcium leakage, disrupts calcium homeostasis and destabilizes myocardial electrical activity. Studies indicate that RyR2-mediated calcium leakage can induce delayed afterdepolarizations (DADs), triggering ectopic excitation and thereby promoting the onset of AF [77]. This calcium leakage and subsequent calcium overload increase the heterogeneity of electrical activity in atrial myocytes, further exacerbating the maintenance and progression of AF [83].
2.4 IP3Rs
IP3Rs are key intracellular calcium release channels regulated by IP3 and reactive oxygen species (ROS) [84] (Fig. 2). IP3Rs are tetrameric protein complexes, each with a molecular weight of approximately 240–260 kDa, and are composed of three primary domains: the ligand-binding domain, the regulatory domain, and the pore domain [42]. Upon IP3 binding to the N-terminal ligand-binding domain, the calcium channel is activated. The regulatory domain modulates IP3R activity through interactions with calmodulin and phosphatidylinositol 4,5-bisphosphate (PIP2), while the C-terminal pore domain serves as the pathway for calcium efflux [85].
In mammals, IP3Rs exist in three isoforms: IP3R1, IP3R2, and IP3R3. IP3R2 is the predominant isoform in atrial myocytes, where it collaborates with RyR2 to regulate calcium dynamics within cardiomyocytes [42] (Table 1). In adult hearts, abnormal activation of IP3R2 is closely linked to pathological calcium overload, which can promote myocardial hypertrophy and arrhythmias [43]. Research indicates that IP3R2 plays a pivotal role in cardiac remodeling; its overactivation leads to intracellular calcium overload, exacerbating myocardial fibrosis and vascular contraction, thereby increasing the risk of AF and heart failure [86]. Recent studies suggest that targeting the structure of the mitochondria-associated endoplasmic reticulum (ER) membrane or modulating IP3R2 activity may effectively restore calcium homeostasis in cardiomyocytes, offering a potential therapeutic strategy to mitigate the progression of heart failure [87].
2.5 CRAC Channels
CRAC channels are non-voltage-dependent calcium channels (NVDCCs) primarily located in the pacemaker cells of the sinoatrial node and ventricular myocytes, where they are essential for maintaining cardiac calcium homeostasis [45] (Table 1). CRAC channels are composed of ORAI family proteins—ORAI1, ORAI2, and ORAI3—embedded in the cell membrane to facilitate extracellular calcium influx [47]. ORAI1 is the most highly expressed subtype in the heart, serving as the primary mediator of calcium signaling and homeostasis, while ORAI2 and ORAI3 function mainly in immune cells, providing complementary roles [45].
STIM1, a calcium-sensing protein located on the SR membrane, continuously monitors intracellular calcium levels [46] (Table 1). When calcium stores are depleted, STIM1 becomes activated and translocates to the cell membrane, where it binds to ORAI1 to form functional CRAC channels, enabling calcium influx through a process known as store-operated calcium entry (SOCE) [45] (Fig. 2). This influx of calcium directly affects cardiomyocyte depolarization, triggering additional calcium release and generating calcium transients essential for ECC in cardiomyocytes [88].
The cooperation between STIM1 and ORAI1, along with their interaction with VGCCs, enables precise regulation of calcium signaling. In conditions of calcium overload, CRAC channels work in concert with potassium channels to stabilize membrane potential [89]. However, excessive activation of CRAC channels in cardiomyocytes can lead to intracellular calcium overload, resulting in abnormal electrical activity and potential myocardial injury [47]. Therefore, modulating CRAC channel function may offer therapeutic benefits in maintaining calcium homeostasis and managing cardiac pathologies such as arrhythmias.
2.6 TRP Channels
The TRP channel family comprises various subtypes, including TRPC, TRPM, TRPV, and TRPP channels (Table 1). Notably, TRPC1, TRPC3, TRPC6, TRPM4, TRPM7, TRPV1, and TRPP2 are closely associated with cardiac function [9, 44, 48, 53, 54, 57, 90] (Fig. 2). Within the TRPC subfamily, TRPC1, TRPC3, and TRPC6 play key roles in cardiac contraction, myocardial hypertrophy, and arrhythmogenesis through the regulation of calcium influx. Abnormal activation of TRPC3 and TRPC6 has been linked to pathological cardiac remodeling [9, 44, 48]. TRPC3, in particular, modulates myocardial structure by influencing the proliferation and differentiation of cardiac fibroblasts, possibly via calcium influx in the extracellular signal-regulated kinase (ERK) signaling pathway [14]. Studies indicate that deletion of microRNA-26 enhances TRPC3 expression, further promoting the proliferation and differentiation of cardiac fibroblasts [91]. TRPC6 is also critical in cardiac fibroblast transformation, underscoring its role in myocardial remodeling [92].
Upon depletion of SR calcium stores, STIM1 protein activates and facilitates the interaction between ORAI2 and TRPC6, establishing a calcium signaling network that supports calcium homeostasis [93]. TRPC6 channels are activated by stimuli from G-protein-coupled receptors (GPCRs), mechanical stress, and fluctuations in sarcoplasmic calcium levels, thereby increasing calcium influx to sustain cardiomyocyte contraction and electrical activity [49]. TRPC6 activation elevates intracellular calcium concentrations, which promotes depolarization and the generation of action potentials, which contribute to pathological calcium dysfunction in the myocardium [9]. In addition to TRPC channels, TRPM4, a sodium channel, contributes to cardiomyocyte depolarization, thereby influencing cardiac automaticity and conduction [55]. TRPM7 is essential in maintaining calcium and magnesium homeostasis, and its dysregulation is linked to arrhythmias [56]. TRPM7 regulates HCN4 (hyperpolarization-activated cyclic nucleotide-gated channel 4) through epigenetic mechanisms and plays an important role in the control of heart rate, pacing and cardiac conduction [94]. HCN4 is the main pacemaker channel of SA cells, and its regulation by TRPM7 may directly affect the maintenance of heart rhythm [95]. TRPV1 is, responsive to temperature and chemical stimuli, regulates calcium influx, and affects cardiomyocyte excitability [58]. TRPP2 functions as a non-selective cation channel involved in calcium signaling, influencing cardiac structure and function [59]. The diversity of TRP channels and their regulatory roles in cardiac function make them an important area of interest. They are key to understanding the pathophysiology of cardiovascular diseases and identifying novel therapeutic targets.
3. Mechanisms of Calcium Channel Dysregulation in AF
When calcium channel function is impaired, calcium homeostasis is disrupted. Recent studies have shown that Cav1.2, Cav3.3, connexin 43, and RyR2 are highly expressed in the pulmonary veins of horses with AF, indicating a link between calcium channel dysregulation and AF [96]. Calcium homeostasis imbalance is recognized as a critical factor in the initiation and maintenance of AF. Increased calcium influx, abnormal calcium release, and impaired calcium transport mechanisms contribute to this imbalance [97]. This disruption in calcium homeostasis can lead to depolarization of the atrial myocyte membrane, enhancing atrial excitability and conductivity, thereby promoting the onset of AF [98]. Persistent calcium influx and calcium overload not only result in structural and functional remodeling of the atria—including atrial dilation, fibrosis, and impaired electrical conduction—but also facilitate the progression of AF from a paroxysmal to a persistent form, creating a vicious cycle that perpetuates AF [98, 99].
3.1 LTCCs Dysregulation
LTCCs are essential for maintaining calcium homeostasis in atrial myocytes. Under normal conditions, calcium ions enter the cell via LTCCs, triggering further calcium release from the SR to facilitate muscle contraction [60]. However, in AF, LTCCs dysfunction leads to either intracellular calcium overload or insufficient calcium influx, resulting in increased instability of calcium transients within atrial myocytes [41] (Table 2, Ref. [11, 12, 13, 14, 20, 31, 33, 41, 45, 49, 50, 51, 52, 75, 76, 79, 82, 98, 100, 101, 102, 103, 104, 105, 106, 107, 108, 109, 110, 111, 112, 113, 114, 115, 116, 117, 118, 119, 120, 121, 122, 123, 124, 125]).
Table 2.
Calcium channels dysregulation in AF.
Ca2+ channels | Activation or inactivation | Ca2+ in cytoplasm | Electrophysiological impact | Ref |
LTCC channels | Increased depolarization, prolonged action potential duration, leading to early afterdepolarizations (EADs) and triggered arrhythmias (TDP) | [11, 12, 31, 33, 100, 101] | ||
Decreased depolarization, shortened action potential duration, reduced myocardial contractility, leading to arrhythmias | [13, 41, 102, 103, 104, 105] | |||
T-channels | Increased pacemaker activity, contributes to early depolarization, may enhance arrhythmogenicity | [20, 45, 75, 76, 98, 106, 107] | ||
RyR2 | Dysregulation of calcium release, leading to increased calcium spark events, contributing to arrhythmias | [82, 108, 109, 110, 111] | ||
IP3Rs | Increased diastolic Ca2+ leak and arrhythmogenic potential | [112, 113] | ||
CRAC channels | Enhanced calcium influx, contributing to cellular depolarization and arrhythmia initiation | [14, 114, 115, 116, 117, 118, 119] | ||
TRPC channels | Increased calcium influx, contributing to pathological calcium overload and arrhythmias | [49, 51, 120, 121, 122, 123] | ||
TRPM channels | Increased calcium influx, leading to cell membrane depolarization and arrhythmogenesis | [52] | ||
TRPV channels | Increased calcium influx, contributing to cell excitability and arrhythmia | [123, 124] | ||
TRPP channels | Increased calcium influx, leading to enhanced excitability and arrhythmogenic potential | [50] | ||
NCX | Reduced calcium overload, leading to impaired contraction and potentially contributing to arrhythmia | [75, 79, 82, 125] |
Notes: “” refers to the activation of calcium channels or changes in intracellular Ca2+ concentration. “” refers to the inactivation of calcium channels.
Upregulation of LTCCs enhances calcium influx and prolongs action potential duration (APD), thereby increasing the risk of reentrant arrhythmias [33]. Excessive calcium influx disrupts calcium transients, which can increase DADs, heightening the likelihood of triggered ectopic activity [11]. Additionally, calcium overload activates calcium-dependent non-selective cation channels (NSCCs), further aggravating atrial electrical instability through DAD generation [31]. AF is commonly associated with structural and electrophysiological remodeling of the atrial myocardium. The activation of CaMKII plays a key role in this process, promoting pathological changes such as atrial fibrosis and cellular hypertrophy [12] (Fig. 3). Research indicates that oxidative stress can further activate CaMKII, resulting in sustained activation of LTCCs, disrupting calcium homeostasis, and increased susceptibility to AF [100]. In addition, sympathetic nervous system activation upregulates LTCCs expression via -adrenergic receptors and the cAMP/PKA signaling pathway, thereby enhancing calcium influx and exacerbating calcium imbalance [101].
Fig. 3.
Factors regulating calcium channels in cardiomyocytes. The octagon represents different stimulating factors, and the arrow points to the calcium channel regulated by this factor. ROS, reactive oxygen species; PKA, protein kinase A; cAMP, cyclic adenosine monophosphate; CaMKII, calcium/calmodulin-dependent protein kinase II; CaM, calmodulin; P2YR, purinergic receptor P2Y; GPCR, G-protein coupled receptor. Fig. 3 was drawn using BioRender.
In AF, the downregulation of LTCCs is primarily driven by decreased transcription and protein levels of the 1 subunit [102]. Experimental studies have demonstrated that intraperitoneal injection of lipopolysaccharide (LPS) significantly reduces LTCCs gene expression in SD rats [103]. ETV1 gene knockout in mouse atrial myocytes has been shown to decrease Cav1.2 expression, shorten APD, and increase the incidence of recurrent excitations, potentially leading to structural remodeling such as myocardial fibrosis [13]. The electrical remodeling associated with AF also contributes to the down regulation of LTCCs, reducing the probability for channels to be opened, which leads to diminished calcium influx and decreased APD stability [102]. A decrease in calcium influx directly impacts Cav1.2 downregulation, weakening the CICR mechanism and disrupting intracellular calcium homeostasis. This impairment significantly reduces myocardial contractility, particularly in atrial myocytes, where the effects are more pronounced [41]. Furthermore, reduced calcium influx increases the burden on the SERCA, depleting calcium stores, and further aggravating calcium imbalance [27].
AF progression also modulates nuclear calcium levels in cardiomyocytes through the miR-26a-regulated IP3R1/CaMKII/HDAC4 signaling pathway, which accelerates LTCCs downregulation [104]. AF-induced nuclear translocation of NFAT downregulates Cav1.2 expression and reduces L-type calcium influx. This signaling cascade, coupled with miR-26 downregulation and subsequent enhancement of upstream potassium currents, promotes the persistence of AF [105].
3.2 T-channels Dysregulation
Upregulation of T-channels is closely associated with the initiation and maintenance of AF, along with molecular mechanisms involving CaMKII, oxidative stress, and other regulatory signaling factors [20] (Table 2, Fig. 3). CaMKII plays a dual role in this process: it phosphorylates and activates T-channels, while also enhancing the activity of RyR2. This dual activation exacerbates calcium leakage and overload, establishing a pathological feedback loop that perpetuates AF [106]. Ethanol exposure further contributes to AF susceptibility by upregulating T-channel expression through the PKC/GSK3 signaling pathway—a mechanism that is seen in alcohol-induced AF patients [107].
In AF patients, Cav3.1 upregulation increases intracellular calcium load in myocytes, promoting atrial electrical remodeling [76]. Similarly, Cav3.2 upregulation in response to myocardial ischemia and hypertrophy increases calcium influx, enhancing the automaticity of atrial myocytes and increases the risk of triggered activity [107]. The low-threshold activation characteristic of Cav3.1 allows it to initiate activity early in the action potential, contributing to early depolarizations and shortening the APD [70]. This shortening of the APD enhances atrial myocyte excitability, making reentrant electrical activity more likely and thereby sustaining AF [38].
Chronic AF results in further upregulation of T-channels, leading to increased abnormal calcium influx and significant alterations in the electrophysiological properties of atrial myocytes [98]. Excessive activation of T-channels results in calcium overload, which activates the NCX, triggering both early afterdepolarizations (EADs) and DADs, thereby facilitating the initiation and perpetuation of AF [75]. The resultant calcium overload from T-channel upregulation places an increased burden on the SERCA, impairing calcium reuptake and further disrupting calcium homeostasis. This dysregulation contributes to an imbalance in intracellular calcium levels within atrial myocytes, exacerbating AF pathology, and results in further disease progression [45].
3.3 RyR2 Dysregulation
Calcium leakage through RyR2 channels is a central mechanism in the pathogenesis of AF. Under normal physiological conditions, RyR2 channels open in response to specific triggers, allowing tightly regulated calcium release from the SR [41] (Fig. 3). However, factors such as hyperphosphorylation and oxidative stress can lead to abnormal RyR2 activation, resulting in SR calcium depletion and sustained calcium leakage [108] (Table 2). Recent proteomic studies have identified that the absence of a novel regulatory subunit of protein phosphatase 1 (PP1), encoded by the PPP1R3A gene, accelerates the phosphorylation of RyR2 and phospholamban (PLN), increasing AF susceptibility in murine models [109]. Calcium leakage from RyR2 contributes to calcium overload within atrial myocytes, activating the NCX and inducing DADs during repolarization, which promotes ectopic activity and facilitates the initiation and maintenance of AF [79]. CaMKII and PKA are major pathways that contribute to RyR2 hyperphosphorylation, thereby exacerbating calcium leakage [82, 110, 111]. Elevated intracellular calcium levels further enhance CaMKII activity, which indirectly activates RyR2 by modulating LTCC and NCX function, creating a feedback loop that intensifies calcium dysregulation [82]. Studies have found that ETV1 gene knockout in mice reduces Cav1.2 expression, affecting cardiomyocyte calcium homeostasis and RyR2 activation [13].
Clinical and experimental evidence has shown that RyR2 phosphorylation levels are significantly elevated in AF patients, leading to increased calcium leakage, particularly in the early stages of condition [109]. As AF progresses, this calcium leakage accelerates the transition from paroxysmal to persistent AF [126]. The chronic disruption of calcium homeostasis also drives atrial structural remodeling, including fibrosis and cardiomyocyte apoptosis, increasing atrial heterogeneity and thereby contributing to sustained AF pathology [15].
3.4 IP3Rs Dysregulation
Upregulation of IP3Rs has been linked to increased calcium influx, and increased susceptibility to AF (Table 2). Research indicates that combined activation of CaMKII and IP3 signaling pathways can stimulate IP3Rs, thereby promoting AF in patients with heart failure [112] (Fig. 3). ER stress and mitochondrial dysfunction are recognized as central mechanisms in atrial remodeling, especially in individuals with type 2 diabetes [127]. The IP3R1/GRP75/VDAC1 complex is crucial for calcium signaling between the ER and mitochondria, facilitating oxidative stress interactions that contribute to diabetes-associated atrial remodeling [128].
Excessive activation of IP3R2 has been shown to induce electrical remodeling in atrial myocytes, further increasing the risk of AF [113]. Anti-apoptotic members of the Bcl-2 family play a dual role: stabilizing mitochondrial membrane integrity and modulating IP3R activity within the ER, thereby directly influencing calcium dynamics and homeostasis [129]. These findings suggest that targeting IP3Rs and related regulatory pathways may offer a therapeutic approach to mitigate AF risk in patients with diabetes and heart failure [104, 130].
3.5 CRAC Channels Dysregulation
The CRAC channels, primarily consisting of STIM1 and ORAI1, play pivotal roles in the pathophysiology of AF, with dysregulated expression and function contributing significantly to the disease process [45] (Table 2, Fig. 3). In pathological states such as myocardial hypertrophy and heart failure, overactivation of ORAI1 leads to substantial intracellular calcium overload, thereby amplifying the calcium homeostasis imbalance. This disruption in calcium regulation results in aberrant electrical activity and myocardial fibrosis, key factors in the development and progression of AF [114]. SOCE, through enhanced collagen secretion by atrial fibroblasts, directly participates in the formation of atrial fibrosis, a process closely linked to the development of AF [115].
The interaction between CRAC channels and TRPC3 has also been demonstrated to stimulate the expression of fibrosis-related genes in atrial fibroblasts, further contributing to the persistence of AF [116]. In animal models, increased expression of ORAI1 and enhanced SOCE-mediated calcium influx have been strongly associated with the progression of atrial fibrosis, exacerbating the arrhythmic substrate [117, 118].
Clinically, elevated levels of ORAI1 expression in atrial tissues from AF patients have been found to correlate with increased calcium ion influx, triggering electrical remodeling and functional disturbances in the atria [14]. Recent research has highlighted fibroblast growth factor 23 (FGF-23) as a key regulator, which upregulates ORAI1-mediated calcium influx via activation of the FGF receptor 1. This discovery opens promising new avenues for targeted therapies in the management of AF [119].
3.6 TRP Channels Dysregulation
The role of TRP channels, particularly TRPC6, in the pathophysiology of heart disease, especially in the onset and maintenance of AF, has received significant attention in recent years. Studies have shown that the expression of TRP channel-related genes is elevated in leukocytes of patients with non-valvular AF (NVAF) [131]. TRPC1 activity may be triggered by angiotensin II or mechanical stress, leading to cardiomyocyte hypertrophy and fibrosis [120] (Table 2, Fig. 3). In conditions such as cardiac hypertrophy or arrhythmia, the synergistic effect of TRPC1 with CRAC channels enhances calcium influx, resulting in prolonged action potentials and spontaneous calcium release, thereby promoting electrical remodeling in the myocardium [49].
Upregulation of TRPC3, TRPC6, and TRPP2 has also been associated with myocardial remodeling, which contributes to cardiomyocyte hypertrophy and accelerates the progression of heart failure [50, 51]. In atrial tissue from AF patients, elevated TRPC6 expression is closely linked to the activation of NFAT and AP-1 transcription factors, which influence the electrophysiological properties of the heart by promoting TRPC6 gene expression [121]. Vericiguat has shown promise in AF management by attenuating structural and electrical remodeling through TRPC6 upregulation [132]. In heart failure models, TRPC6 upregulation has been observed to enhance contractility in cardiomyocytes, though it also leads to electrical instability [114]. Experimental models of AF have demonstrated that specific TRPC6 inhibitors can reduce the incidence of AF, underscoring the therapeutic potential of targeting TRPC6 [9, 133]. Other TRP family members also play significant roles in AF pathophysiology.
Mutations in TRPM4, a sodium channel, have been linked to idiopathic arrhythmias and cardiac conduction blockage [52]. TRPM7 deficiency or mutation disrupts calcium and magnesium homeostasis in cardiomyocytes and influences AF pathogenesis by modulating calcium and calmodulin signaling pathways [122, 134]. TRPV1, activated by oxidative stress and inflammation, is aberrantly expressed in atrial myocytes and may contribute to increased cellular stress responses, electrical remodeling, and persistence of AF [123, 124]. Overall, TRP channels represent critical therapeutic targets in AF pathophysiology. Modulation of TRP channel function offers a promising avenue for innovative treatments in arrhythmias and cardiomyopathies.
4. Application of Calcium Channel-Targeted Therapies in AF
Calcium channel-targeted therapies have attracted significant attention in the management of AF, particularly for their potential to stabilize atrial electrical activity by modulating calcium homeostasis.
LTCC blockers, including verapamil and diltiazem, are widely used in the management of chronic AF for effective heart rate control. These agents reduce calcium influx and prolong action potential duration in atrial myocytes, offering reliable rate control with a favorable safety profile [135]. Additionally, nifedipine significantly inhibits Cav1.2 by reducing connexin 43 levels, which may further contribute to its therapeutic effects [136]. Targeting the TGF- signaling pathway and using agents such as sacubitril/valsartan or miR-155 have shown promise in downregulating Cav1.2 expression, thereby reducing AF-related fibrosis and electrical remodeling [137, 138]. Recent studies indicate that CRISPR-Cas9 gene-editing technology can selectively inhibit the CACNA1C gene, potentially correcting mutations associated with LTCC overexpression. This approach has demonstrated antifibrotic effects in human-induced pluripotent stem cell (iPSC) cardiomyocyte models, offering an innovative strategy for AF management pending further clinical validation [139, 140].
-Blockers, which indirectly downregulate Cav1.2 by inhibiting sympathetic activity, are also used in AF management, although their atrial specificity is limited [141]. A small-molecule integrated stress response inhibitor has shown efficacy in reducing macrophage infiltration and oxidative stress following myocardial infarction in rat models, thus decreasing AF susceptibility [142]. Natriuretic peptide receptor B, which targets both LTCCs and RyR2, shows potential in treating AF triggered by -adrenergic stimulation [143]. Additionally, KN-93, a CaMKII inhibitor, prevents LTCCs activation, making it a novel target for AF therapy [144].
Cav3.1 and Cav3.2 T-type calcium channels are essential in regulating cardiac automaticity. Studies have shown that TNF- reduces T-type calcium channel currents in atrial myocytes of mice, while mibefradil suppresses Cav3.1 expression by downregulating connexin 43 [136, 145]. Additionally, low concentrations of isopimaric acid, a diterpene compound, have been found to inactivate both L-type and T-type calcium channels, leading to reduced action potential frequency and the restoration of normal cardiac rhythm [146]. Despite their potential, clinical applications of T-type calcium channel blockers are currently limited due to adverse effects. Consequently, the development of selective modulators for T-type calcium channels, with improved specificity and safety profiles, presents a promising direction for future AF therapies.
In AF patients, excessive phosphorylation of RyR2 leads to spontaneous calcium leakage, resulting in DADs and reentrant electrical activity. Modulation of RyR2 and the NCX via histone deacetylase inhibitors has shown promise in AF prevention [125]. Recent research has identified several agents—such as dapansutrile, febuxostat, the selective RyR2 inhibitor ent-verticilide, and propafenone, that can reduce AF susceptibility [147, 148, 149]. M201-A has demonstrated the potential to prolong the atrial effective refractory period and improve cardiac function without compromising ventricular contraction, making it a promising candidate for AF patients with concurrent heart failure [150].
In ischemic and cholinergic stimulation-induced models of AF, inhibition of the IP3R by 2-APB has been shown to effectively reduce cytosolic calcium overload and mitigate cellular energy damage [151]. Modulation of IP3R offers a promising approach not only to alleviate abnormal calcium signaling in AF but also to pave the way for the development of precision molecular therapies for arrhythmias. This targeted approach holds potential for advancing AF treatment through innovative molecular design and tailored therapeutic strategies.
The interaction between STIM1 and Orai1 is a crucial step in the activation of the CRAC channel. Inhibitors such as CM4620 and Synta66 have been shown to effectively block Orai1-mediated calcium influx, thereby preventing the trigger calcium transients that play a pivotal role in the initiation of AF [152, 153]. Oxidative stress is one of the key factors contributing to the abnormal activation of the CRAC channel, and its modulation may represent a potential therapeutic approach for AF [154, 155]. Qingyi decoction has been reported to regulate the STIM1/Orai1 pathway, mitigating inflammation and improving calcium homeostasis, which could provide a promising strategy for AF treatment [156]. Furthermore, enhancing the activity of the calcium ATPase SERCA2a accelerates calcium reuptake, which not only mitigates calcium overload but also helps restore calcium homeostasis, offering potential therapeutic avenues for managing AF-induced electrical remodeling [157].
In experimental models of diabetes, TRPC6 expression was significantly downregulated following treatment with Tinglu Yixin granules, leading to a reduction in inflammation, oxidative stress, and myocardial fibrosis [158]. In addition, studies using SKF96365 and puerarin have further demonstrated that the downregulation of TRPC6 expression can effectively mitigate myocardial remodeling and oxidative stress in diabetic rats [121, 159]. These findings suggest that targeting TRPC6 may hold therapeutic potential for alleviating the pathological changes associated with AF.
There are not many calcium channel therapeutic drugs in the clinical treatment of AF. At present, the most widely used calcium channel blockers are verapamil and diltiazem. Therefore, it is necessary to develop additional of calcium channel therapeutic drugs. Several emerging calcium channel regulation technologies are still in the laboratory stage, and the transformation from basic research to clinical application should be accelerated. Calcium channels are distributed in both the atrium and ventricle, and calcium channel modulators with high selectivity and low side effects will optimize AF treatment.
5. Conclusions
In the pathogenesis of AF, dysregulation of calcium channels disrupts calcium homeostasis, leading to electrophysiological dysregulation. and structural remodeling in the atrial myocytes. Various calcium channels—including L-type, T-type, RyRs, IP3Rs, CRAC and TRP channels—regulate intracellular calcium dynamics in unique and interrelated ways, driving the progression of AF.
While calcium channel-targeted therapies have shown efficacy in AF management, the diversity and complexity of calcium channel regulation in different atrial cell types continue to pose challenges for the development of highly specific treatments with minimal side effects. Additionally, strategies integrating multi-target drug design, gene editing, and precision medicine approaches to selectively correct calcium channel dysfunction may offer more promising therapeutic options for AF patients. These methods hold potential for optimizing AF management in the future, providing patients with more effective and personalized treatment options.
Acknowledgment
Not applicable.
Abbreviations
AF, atrial fibrillation; LTCCs, L-type calcium channels; T-channels, T-type calcium channels; RyRs, ryanodine receptor calcium release channels; IP3Rs, inositol 1,4,5-triphosphate receptors; CRAC channels, calcium release-activated calcium channels; TRP channels, transient receptor potential channels; STIM1, The binding of Stromal Interaction Molecule 1; SR, sarcoplasmic reticulum; CICR, calcium-induced calcium release; SERCA, sarcoplasmic/endoplasmic reticulum Ca2+-ATPase; NCX, sodium-calcium exchanger; ECC, excitation-contraction coupling; VGCCs, voltage-gated calcium channels; SA, sinoatrial nodes; AV, atrioventricular nodes; KCa2.x, small-conductance calcium-activated potassium channels; LVACs, low-voltage-activated channels; ICaT, T-type calcium current; BKCa, large-conductance calcium-activated potassium; CaM, calmodulin; PKA, protein kinase A; CaMKII, calcium/calmodulin-dependent protein kinase II; DADs, delayed afterdepolarizations; ROS, reactive oxygen species; PIP2, phosphatidylinositol 4,5-bisphosphate; NVDCCs, non-voltage-dependent calcium channels; SOCE, store-operated calcium entry; GPCRs, G-protein-coupled receptors; HCN4, hyperpolarization-activated cyclic nucleotide-gated channel 4; TDP, triggered arrhythmias; APD, action potential duration; NSCCs, non-selective cation channels; LPS, lipopolysaccharide; EADs, early afterdepolarizations; PP1, protein phosphatase 1; PLN, phospholamban; ER, endoplasmic reticulum; FGF-23, fibroblast growth factor 23; NVAF, non-valvular AF; iPSC, human-induced pluripotent stem cell.
Funding Statement
This research was funded by grants from the National Natural Science Foundation of China (Grant No.82360066) and the Natural Science Foundation of Guangxi (Grant No. 2025GXNSFAA069056).
Footnotes
Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Author Contributions
ZH and CL designed the research study. ZH, CL and ZW performed the research. CL and ZW validated the study. CL analyzed the data. ZH provided resources. ZW curated the data. ZH, CL and ZW prepared the original draft. ZH and CL created visualizations. BZ formulated the research questions, designed the experimental protocols, and oversaw the overall direction of the project. BZ coordinated the research team, ensured the proper collection and management of data, and provided critical input on data collection methods. BZ is involved in the statistical analysis, interpretation of results, and discussion of the findings and drafted the manuscript and critically revised it for important intellectual content. All authors read and approved the final manuscript. All authors have participated sufficiently in the work and agreed to be accountable for all aspects of the work.
Ethics Approval and Consent to Participate
Not applicable.
Funding
This research was funded by grants from the National Natural Science Foundation of China (Grant No.82360066) and the Natural Science Foundation of Guangxi (Grant No. 2025GXNSFAA069056).
Conflict of Interest
The authors declare no conflicts of interest.
References
- [1].Elliott AD, Middeldorp ME, Van Gelder IC, Albert CM, Sanders P. Epidemiology and modifiable risk factors for atrial fibrillation. Nature Reviews. Cardiology . 2023;20:404–417. doi: 10.1038/s41569-022-00820-8. [DOI] [PubMed] [Google Scholar]
- [2].Wu J, Nadarajah R, Nakao YM, Nakao K, Wilkinson C, Mamas MA, et al. Temporal trends and patterns in atrial fibrillation incidence: A population-based study of 3•4 million individuals. The Lancet Regional Health . 2022;17:100386. doi: 10.1016/j.lanepe.2022.100386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [3].Essien UR, Kornej J, Johnson AE, Schulson LB, Benjamin EJ, Magnani JW. Social determinants of atrial fibrillation. Nature Reviews. Cardiology . 2021;18:763–773. doi: 10.1038/s41569-021-00561-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [4].Mikami A, Imoto K, Tanabe T, Niidome T, Mori Y, Takeshima H, et al. Primary structure and functional expression of the cardiac dihydropyridine-sensitive calcium channel. Nature . 1989;340:230–233. doi: 10.1038/340230a0. [DOI] [PubMed] [Google Scholar]
- [5].Bartolucci C, Mesirca P, Ricci E, Sales-Bellés C, Torre E, Louradour J, et al. Computational modelling of mouse atrio ventricular node action potential and automaticity. The Journal of Physiology . 2024;602:4821–4847. doi: 10.1113/JP285950. [DOI] [PubMed] [Google Scholar]
- [6].Yano M, Yamamoto T, Ikemoto N, Matsuzaki M. Abnormal ryanodine receptor function in heart failure. Pharmacology & Therapeutics . 2005;107:377–391. doi: 10.1016/j.pharmthera.2005.04.003. [DOI] [PubMed] [Google Scholar]
- [7].Gorza L, Schiaffino S, Volpe P. Inositol 1,4,5-trisphosphate receptor in heart: evidence for its concentration in Purkinje myocytes of the conduction system. The Journal of Cell Biology . 1993;121:345–353. doi: 10.1083/jcb.121.2.345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [8].Feske S. CRAC channels and disease - From human CRAC channelopathies and animal models to novel drugs. Cell Calcium . 2019;80:112–116. doi: 10.1016/j.ceca.2019.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [9].Chen W, Oberwinkler H, Werner F, Gaßner B, Nakagawa H, Feil R, et al. Atrial natriuretic peptide-mediated inhibition of microcirculatory endothelial Ca2+ and permeability response to histamine involves cGMP-dependent protein kinase I and TRPC6 channels. Arteriosclerosis, Thrombosis, and Vascular Biology . 2013;33:2121–2129. doi: 10.1161/ATVBAHA.113.001974. [DOI] [PubMed] [Google Scholar]
- [10].Lauder L, Mahfoud F, Azizi M, Bhatt DL, Ewen S, Kario K, et al. Hypertension management in patients with cardiovascular comorbidities. European Heart Journal . 2023;44:2066–2077. doi: 10.1093/eurheartj/ehac395. [DOI] [PubMed] [Google Scholar]
- [11].Santini L, Coppini R, Cerbai E. Ion Channel Impairment and Myofilament Ca2+ Sensitization: Two Parallel Mechanisms Underlying Arrhythmogenesis in Hypertrophic Cardiomyopathy. Cells . 2021;10:2789. doi: 10.3390/cells10102789. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [12].Zhang M, Yang X, Zimmerman RJ, Wang Q, Ross MA, Granger JM, et al. CaMKII exacerbates heart failure progression by activating class I HDACs. Journal of Molecular and Cellular Cardiology . 2020;149:73–81. doi: 10.1016/j.yjmcc.2020.09.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [13].Fang LH, Chen Q, Cheng XL, Li XQ, Zou T, Chen JQ, et al. Calcium-mediated DAD in membrane potentials and triggered activity in atrial myocytes of ETV1f/fMyHCCre/+ mice. Journal of Cellular and Molecular Medicine . 2024;28:e70005. doi: 10.1111/jcmm.70005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [14].Han L, Li J. Canonical transient receptor potential 3 channels in atrial fibrillation. European Journal of Pharmacology . 2018;837:1–7. doi: 10.1016/j.ejphar.2018.08.030. [DOI] [PubMed] [Google Scholar]
- [15].Su KN, Ma Y, Cacheux M, Ilkan Z, Raad N, Muller GK, et al. Atrial AMP-activated protein kinase is critical for prevention of dysregulation of electrical excitability and atrial fibrillation. JCI Insight . 2022;7:e141213. doi: 10.1172/jci.insight.141213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [16].Yao Y, Yang M, Liu D, Zhao Q. Immune remodeling and atrial fibrillation. Frontiers in Physiology . 2022;13:927221. doi: 10.3389/fphys.2022.927221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [17].Varró A, Tomek J, Nagy N, Virág L, Passini E, Rodriguez B, et al. Cardiac transmembrane ion channels and action potentials: cellular physiology and arrhythmogenic behavior. Physiological Reviews . 2021;101:1083–1176. doi: 10.1152/physrev.00024.2019. [DOI] [PubMed] [Google Scholar]
- [18].Shi S, Mao X, Lv J, Wang Y, Zhang X, Shou X, et al. Qi-Po-Sheng-Mai granule ameliorates Ach-CaCl2 -induced atrial fibrillation by regulating calcium homeostasis in cardiomyocytes. Phytomedicine: International Journal of Phytotherapy and Phytopharmacology . 2023;119:155017. doi: 10.1016/j.phymed.2023.155017. [DOI] [PubMed] [Google Scholar]
- [19].Grammatika Pavlidou N, Dobrev S, Beneke K, Reinhardt F, Pecha S, Jacquet E, et al. Phosphodiesterase 8 governs cAMP/PKA-dependent reduction of L-type calcium current in human atrial fibrillation: a novel arrhythmogenic mechanism. European Heart Journal . 2023;44:2483–2494. doi: 10.1093/eurheartj/ehad086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [20].Peyronnet R, Ravens U. Atria-selective antiarrhythmic drugs in need of alliance partners. Pharmacological Research . 2019;145:104262. doi: 10.1016/j.phrs.2019.104262. [DOI] [PubMed] [Google Scholar]
- [21].Moise N, Weinberg SH. Calcium Homeostatic Feedback Control Predicts Atrial Fibrillation Initiation, Remodeling, and Progression. JACC. Clinical Electrophysiology . 2025;2405:S2405–S2405. doi: 10.1016/j.jacep.2025.03.004. (online ahead of print) [DOI] [PubMed] [Google Scholar]
- [22].Barry WH, Bridge JH. Intracellular calcium homeostasis in cardiac myocytes. Circulation . 1993;87:1806–1815. doi: 10.1161/01.cir.87.6.1806. [DOI] [PubMed] [Google Scholar]
- [23].Leuranguer V, Monteil A, Bourinet E, Dayanithi G, Nargeot J. T-type calcium currents in rat cardiomyocytes during postnatal development: contribution to hormone secretion. American Journal of Physiology. Heart and Circulatory Physiology . 2000;279:H2540–H2548. doi: 10.1152/ajpheart.2000.279.5.H2540. [DOI] [PubMed] [Google Scholar]
- [24].Jaleel N, Nakayama H, Chen X, Kubo H, MacDonnell S, Zhang H, et al. Ca2+ influx through T- and L-type Ca2+ channels have different effects on myocyte contractility and induce unique cardiac phenotypes. Circulation Research . 2008;103:1109–1119. doi: 10.1161/CIRCRESAHA.108.185611. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [25].Bers DM. Calcium fluxes involved in control of cardiac myocyte contraction. Circulation Research . 2000;87:275–281. doi: 10.1161/01.res.87.4.275. [DOI] [PubMed] [Google Scholar]
- [26].Dellis O, Dedos SG, Tovey SC, Dubel SJ, Taylor CW. Ca2+ entry through plasma membrane IP3 receptors. Science (New York, N.Y.) . 2006;313:229–233. doi: 10.1126/science.1125203. [DOI] [PubMed] [Google Scholar]
- [27].Lytton J, Westlin M, Burk SE, Shull GE, MacLennan DH. Functional comparisons between isoforms of the sarcoplasmic or endoplasmic reticulum family of calcium pumps. The Journal of Biological Chemistry . 1992;267:14483–14489. [PubMed] [Google Scholar]
- [28].Blaustein MP, Lederer WJ. Sodium/calcium exchange: its physiological implications. Physiological Reviews . 1999;79:763–854. doi: 10.1152/physrev.1999.79.3.763. [DOI] [PubMed] [Google Scholar]
- [29].Zweifach A, Lewis RS. Rapid inactivation of depletion-activated calcium current (ICRAC) due to local calcium feedback. The Journal of General Physiology . 1995;105:209–226. doi: 10.1085/jgp.105.2.209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [30].Tuinte WE, Török E, Mahlknecht I, Tuluc P, Flucher BE, Campiglio M. STAC3 determines the slow activation kinetics of CaV 1.1 currents and inhibits its voltage-dependent inactivation. Journal of Cellular Physiology . 2022;237:4197–4214. doi: 10.1002/jcp.30870. [DOI] [PubMed] [Google Scholar]
- [31].Pietrobon D, Hess P. Novel mechanism of voltage-dependent gating in L-type calcium channels. Nature . 1990;346:651–655. doi: 10.1038/346651a0. [DOI] [PubMed] [Google Scholar]
- [32].Wei XY, Perez-Reyes E, Lacerda AE, Schuster G, Brown AM, Birnbaumer L. Heterologous regulation of the cardiac Ca2+ channel alpha 1 subunit by skeletal muscle beta and gamma subunits. Implications for the structure of cardiac L-type Ca2+ channels. The Journal of Biological Chemistry . 1991;266:21943–21947. [PubMed] [Google Scholar]
- [33].Striessnig J, Pinggera A, Kaur G, Bock G, Tuluc P. L-type Ca2+ channels in heart and brain. Wiley Interdisciplinary Reviews. Membrane Transport and Signaling . 2014;3:15–38. doi: 10.1002/wmts.102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [34].Yang L, Katchman A, Morrow JP, Doshi D, Marx SO. Cardiac L-type calcium channel (Cav1.2) associates with gamma subunits. FASEB Journal: Official Publication of the Federation of American Societies for Experimental Biology . 2011;25:928–936. doi: 10.1096/fj.10-172353. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [35].Torrente AG, Mesirca P, Neco P, Rizzetto R, Dubel S, Barrere C, et al. L-type Cav1.3 channels regulate ryanodine receptor-dependent Ca2+ release during sino-atrial node pacemaker activity. Cardiovascular Research . 2016;109:451–461. doi: 10.1093/cvr/cvw006. [DOI] [PubMed] [Google Scholar]
- [36].Torrente AG, Mesirca P, Bidaud I, Mangoni ME. Channelopathies of voltage-gated L-type Cav1.3/α1D and T-type Cav3.1/α1G Ca2+ channels in dysfunction of heart automaticity. Pflugers Archiv: European Journal of Physiology . 2020;472:817–830. doi: 10.1007/s00424-020-02421-1. [DOI] [PubMed] [Google Scholar]
- [37].Baudot M, Torre E, Bidaud I, Louradour J, Torrente AG, Fossier L, et al. Concomitant genetic ablation of L-type Cav1.3 (α1D) and T-type Cav3.1 (α1G) Ca2+ channels disrupts heart automaticity. Scientific Reports . 2020;10:18906. doi: 10.1038/s41598-020-76049-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [38].Marger L, Mesirca P, Alig J, Torrente A, Dubel S, Engeland B, et al. Functional roles of Ca(v)1.3, Ca(v)3.1 and HCN channels in automaticity of mouse atrioventricular cells: insights into the atrioventricular pacemaker mechanism. Channels (Austin) . 2011;5:251–261. doi: 10.4161/chan.5.3.15266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [39].Guo Y, Cao Y, Jardin BD, Zhang X, Zhou P, Guatimosim S, et al. Ryanodine receptor 2 (RYR2) dysfunction activates the unfolded protein response and perturbs cardiomyocyte maturation. Cardiovascular Research . 2023;119:221–235. doi: 10.1093/cvr/cvac077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [40].Bauerová-Hlinková V, Hajdúchová D, Bauer JA. Structure and Function of the Human Ryanodine Receptors and Their Association with Myopathies-Present State, Challenges, and Perspectives. Molecules (Basel, Switzerland) . 2020;25:4040. doi: 10.3390/molecules25184040. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [41].Kim K, Blackwell DJ, Yuen SL, Thorpe MP, Johnston JN, Cornea RL, et al. The selective RyR2 inhibitor ent-verticilide suppresses atrial fibrillation susceptibility caused by Pitx2 deficiency. Journal of Molecular and Cellular Cardiology . 2023;180:1–9. doi: 10.1016/j.yjmcc.2023.04.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [42].Woll KA, Van Petegem F. Calcium-release channels: structure and function of IP3 receptors and ryanodine receptors. Physiological Reviews . 2022;102:209–268. doi: 10.1152/physrev.00033.2020. [DOI] [PubMed] [Google Scholar]
- [43].Domeier TL, Zima AV, Maxwell JT, Huke S, Mignery GA, Blatter LA. IP3 receptor-dependent Ca2+ release modulates excitation-contraction coupling in rabbit ventricular myocytes. American Journal of Physiology. Heart and Circulatory Physiology . 2008;294:H596–H604. doi: 10.1152/ajpheart.01155.2007. [DOI] [PubMed] [Google Scholar]
- [44].Adebiyi A, Thomas-Gatewood CM, Leo MD, Kidd MW, Neeb ZP, Jaggar JH. An elevation in physical coupling of type 1 inositol 1,4,5-trisphosphate (IP3) receptors to transient receptor potential 3 (TRPC3) channels constricts mesenteric arteries in genetic hypertension. Hypertension (Dallas, Tex.: 1979) . 2012;60:1213–1219. doi: 10.1161/HYPERTENSIONAHA.112.198820. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [45].Rosenberg P, Katz D, Bryson V. SOCE and STIM1 signaling in the heart: Timing and location matter. Cell Calcium . 2019;77:20–28. doi: 10.1016/j.ceca.2018.11.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [46].Liu J, Xin L, Benson VL, Allen DG, Ju YK. Store-operated calcium entry and the localization of STIM1 and Orai1 proteins in isolated mouse sinoatrial node cells. Frontiers in Physiology . 2015;6:69. doi: 10.3389/fphys.2015.00069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [47].Prakriya M, Feske S, Gwack Y, Srikanth S, Rao A, Hogan PG. Orai1 is an essential pore subunit of the CRAC channel. Nature . 2006;443:230–233. doi: 10.1038/nature05122. [DOI] [PubMed] [Google Scholar]
- [48].Falcón D, Galeano-Otero I, Martín-Bórnez M, Fernández-Velasco M, Gallardo-Castillo I, Rosado JA, et al. TRPC Channels: Dysregulation and Ca2+ Mishandling in Ischemic Heart Disease. Cells . 2020;9:173. doi: 10.3390/cells9010173. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [49].Numaga-Tomita T, Nishida M. TRPC Channels in Cardiac Plasticity. Cells . 2020;9:454. doi: 10.3390/cells9020454. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [50].Gwanyanya A, Mubagwa K. Emerging role of transient receptor potential (TRP) ion channels in cardiac fibroblast pathophysiology. Frontiers in Physiology . 2022;13:968393. doi: 10.3389/fphys.2022.968393. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [51].Onohara N, Nishida M, Inoue R, Kobayashi H, Sumimoto H, Sato Y, et al. TRPC3 and TRPC6 are essential for angiotensin II-induced cardiac hypertrophy. The EMBO Journal . 2006;25:5305–5316. doi: 10.1038/sj.emboj.7601417. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [52].Stallmeyer B, Zumhagen S, Denjoy I, Duthoit G, Hébert JL, Ferrer X, et al. Mutational spectrum in the Ca(2+)–activated cation channel gene TRPM4 in patients with cardiac conductance disturbances. Human Mutation . 2012;33:109–117. doi: 10.1002/humu.21599. [DOI] [PubMed] [Google Scholar]
- [53].Kecskés M, Jacobs G, Kerselaers S, Syam N, Menigoz A, Vangheluwe P, et al. The Ca(2+)-activated cation channel TRPM4 is a negative regulator of angiotensin II-induced cardiac hypertrophy. Basic Research in Cardiology . 2015;110:43. doi: 10.1007/s00395-015-0501-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [54].Yu Y, Chen S, Xiao C, Jia Y, Guo J, Jiang J, et al. TRPM7 is involved in angiotensin II induced cardiac fibrosis development by mediating calcium and magnesium influx. Cell Calcium . 2014;55:252–260. doi: 10.1016/j.ceca.2014.02.019. [DOI] [PubMed] [Google Scholar]
- [55].Hu Y, Cang J, Hiraishi K, Fujita T, Inoue R. The Role of TRPM4 in Cardiac Electrophysiology and Arrhythmogenesis. International Journal of Molecular Sciences . 2023;24:11798. doi: 10.3390/ijms241411798. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [56].Gwanyanya A, Andriulė I, Istrate BM, Easmin F, Mubagwa K, Mačianskienė R. Modulation of the Cardiac Myocyte Action Potential by the Magnesium-Sensitive TRPM6 and TRPM7-like Current. International Journal of Molecular Sciences . 2021;22:8744. doi: 10.3390/ijms22168744. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [57].Castrejón-Téllez V, Del Valle-Mondragón L, Pérez-Torres I, Guarner-Lans V, Pastelín-Hernández G, Ruiz-Ramírez A, et al. TRPV1 Contributes to Modulate the Nitric Oxide Pathway and Oxidative Stress in the Isolated and Perfused Rat Heart during Ischemia and Reperfusion. Molecules (Basel, Switzerland) . 2022;27:1031. doi: 10.3390/molecules27031031. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [58].Shuba YM. Beyond Neuronal Heat Sensing: Diversity of TRPV1 Heat-Capsaicin Receptor-Channel Functions. Frontiers in Cellular Neuroscience . 2021;14:612480. doi: 10.3389/fncel.2020.612480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [59].Vassilev PM, Guo L, Chen XZ, Segal Y, Peng JB, Basora N, et al. Polycystin-2 is a novel cation channel implicated in defective intracellular Ca(2+) homeostasis in polycystic kidney disease. Biochemical and Biophysical Research Communications . 2001;282:341–350. doi: 10.1006/bbrc.2001.4554. [DOI] [PubMed] [Google Scholar]
- [60].Tang M, Zhang X, Li Y, Guan Y, Ai X, Szeto C, et al. Enhanced basal contractility but reduced excitation-contraction coupling efficiency and beta-adrenergic reserve of hearts with increased Cav1.2 activity. American Journal of Physiology. Heart and Circulatory Physiology . 2010;299:H519–H528. doi: 10.1152/ajpheart.00265.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [61].Lehnart SE, Wehrens XHT, Kushnir A, Marks AR. Cardiac ryanodine receptor function and regulation in heart disease. Annals of the New York Academy of Sciences . 2004;1015:144–159. doi: 10.1196/annals.1302.012. [DOI] [PubMed] [Google Scholar]
- [62].Belkacemi A, Hui X, Wardas B, Laschke MW, Wissenbach U, Menger MD, et al. IP3 Receptor-Dependent Cytoplasmic Ca2+ Signals Are Tightly Controlled by Cavβ3. Cell Reports . 2018;22:1339–1349. doi: 10.1016/j.celrep.2018.01.010. [DOI] [PubMed] [Google Scholar]
- [63].Dixon RE, Moreno CM, Yuan C, Opitz-Araya X, Binder MD, Navedo MF, et al. Graded Ca2+/calmodulin-dependent coupling of voltage-gated CaV1.2 channels. eLife . 2015;4:e05608. doi: 10.7554/eLife.05608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [64].Ghosh D, Nieves-Cintrón M, Tajada S, Brust-Mascher I, Horne MC, Hell JW, et al. Dynamic L-type CaV1.2 channel trafficking facilitates CaV1.2 clustering and cooperative gating. Biochimica et Biophysica Acta. Molecular Cell Research . 2018;1865:1341–1355. doi: 10.1016/j.bbamcr.2018.06.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [65].Tuluc P, Flucher BE. Divergent biophysical properties, gating mechanisms, and possible functions of the two skeletal muscle Ca(V)1.1 calcium channel splice variants. Journal of Muscle Research and Cell Motility . 2011;32:249–256. doi: 10.1007/s10974-011-9270-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [66].Christel CJ, Cardona N, Mesirca P, Herrmann S, Hofmann F, Striessnig J, et al. Distinct localization and modulation of Cav1.2 and Cav1.3 L-type Ca2+ channels in mouse sinoatrial node. The Journal of Physiology . 2012;590:6327–6342. doi: 10.1113/jphysiol.2012.239954. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [67].Zhang XD, Thai PN, Lieu DK, Chiamvimonvat N. Cardiac small-conductance calcium-activated potassium channels in health and disease. Pflugers Archiv: European Journal of Physiology . 2021;473:477–489. doi: 10.1007/s00424-021-02535-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [68].Perez-Reyes E. Molecular characterization of T-type calcium channels. Cell Calcium . 2006;40:89–96. doi: 10.1016/j.ceca.2006.04.012. [DOI] [PubMed] [Google Scholar]
- [69].Lory P, Nicole S, Monteil A. Neuronal Cav3 channelopathies: recent progress and perspectives. Pflugers Archiv: European Journal of Physiology . 2020;472:831–844. doi: 10.1007/s00424-020-02429-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [70].Louradour J, Bortolotti O, Torre E, Bidaud I, Lamb N, Fernandez A, et al. L-Type Cav1.3 Calcium Channels Are Required for Beta-Adrenergic Triggered Automaticity in Dormant Mouse Sinoatrial Pacemaker Cells. 3 Calcium Channels Are Required for Beta-Adrenergic Triggered Automaticity in Dormant Mouse Sinoatrial Pacemaker Cells . 2022;11:1114. doi: 10.3390/cells11071114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [71].Mavroidis M, Athanasiadis NC, Rigas P, Kostavasili I, Kloukina I, Te Rijdt WP, et al. Desmin is essential for the structure and function of the sinoatrial node: implications for increased arrhythmogenesis. American Journal of Physiology. Heart and Circulatory Physiology . 2020;319:H557–H570. doi: 10.1152/ajpheart.00594.2019. [DOI] [PubMed] [Google Scholar]
- [72].Niwa N, Yasui K, Opthof T, Takemura H, Shimizu A, Horiba M, et al. Cav3.2 subunit underlies the functional T-type Ca2+ channel in murine hearts during the embryonic period. American Journal of Physiology. Heart and Circulatory Physiology . 2004;286:H2257–H2263. doi: 10.1152/ajpheart.01043.2003. [DOI] [PubMed] [Google Scholar]
- [73].Nam YW, Im D, Garcia ASC, Tringides ML, Nguyen HM, Liu Y, et al. Cryo-EM structures of the small-conductance Ca2+-activated KCa2.2 channel. Nature Communications . 2025;16:3690. doi: 10.1038/s41467-025-59061-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [74].Chen L, He Y, Wang X, Ge J, Li H. Ventricular voltage-gated ion channels: Detection, characteristics, mechanisms, and drug safety evaluation. Clinical and Translational Medicine . 2021;11:e530. doi: 10.1002/ctm2.530. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [75].Chan CS, Lin YK, Chen YC, Lu YY, Chen SA, Chen YJ. Heart Failure Differentially Modulates Natural (Sinoatrial Node) and Ectopic (Pulmonary Veins) Pacemakers: Mechanism and Therapeutic Implication for Atrial Fibrillation. International Journal of Molecular Sciences . 2019;20:3224. doi: 10.3390/ijms20133224. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [76].Zhang X, Gao Y, Zhou Y, Liu Z, Liu R. Pharmacological mechanism of natural drugs and their active ingredients in the treatment of arrhythmia via calcium channel regulation. Biomedicine & Pharmacotherapy = Biomedecine & Pharmacotherapie . 2023;160:114413. doi: 10.1016/j.biopha.2023.114413. [DOI] [PubMed] [Google Scholar]
- [77].Cheng H, Lederer WJ, Cannell MB. Calcium sparks: elementary events underlying excitation-contraction coupling in heart muscle. Science (New York, N.Y.) . 1993;262:740–744. doi: 10.1126/science.8235594. [DOI] [PubMed] [Google Scholar]
- [78].Shi L, Zhang Y, Liu Y, Gu B, Cao R, Chen Y, et al. Exercise Prevents Upregulation of RyRs-BKCa Coupling in Cerebral Arterial Smooth Muscle Cells From Spontaneously Hypertensive Rats. Arteriosclerosis, Thrombosis, and Vascular Biology . 2016;36:1607–1617. doi: 10.1161/ATVBAHA.116.307745. [DOI] [PubMed] [Google Scholar]
- [79].He H, Giordano FJ, Hilal-Dandan R, Choi DJ, Rockman HA, McDonough PM, et al. Overexpression of the rat sarcoplasmic reticulum Ca2+ ATPase gene in the heart of transgenic mice accelerates calcium transients and cardiac relaxation. The Journal of Clinical Investigation . 1997;100:380–389. doi: 10.1172/JCI119544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [80].Federico M, Valverde CA, Mattiazzi A, Palomeque J. Unbalance Between Sarcoplasmic Reticulum Ca2+ Uptake and Release: A First Step Toward Ca2+ Triggered Arrhythmias and Cardiac Damage. Frontiers in Physiology . 2020;10:1630. doi: 10.3389/fphys.2019.01630. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [81].Joshi P, Estes S, DeMazumder D, Knollmann BC, Dey S. Ryanodine receptor 2 inhibition reduces dispersion of cardiac repolarization, improves contractile function, and prevents sudden arrhythmic death in failing hearts. eLife . 2023;12:RP88638. doi: 10.7554/eLife.88638. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [82].Wehrens XHT, Lehnart SE, Reiken S, Vest JA, Wronska A, Marks AR. Ryanodine receptor/calcium release channel PKA phosphorylation: a critical mediator of heart failure progression. Proceedings of the National Academy of Sciences of the United States of America . 2006;103:511–518. doi: 10.1073/pnas.0510113103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [83].Lkhagva B, Lin YK, Chen YC, Cheng WL, Higa S, Kao YH, et al. ZFHX3 knockdown dysregulates mitochondrial adaptations to tachypacing in atrial myocytes through enhanced oxidative stress and calcium overload. Acta Physiologica (Oxford, England) . 2021;231:e13604. doi: 10.1111/apha.13604. [DOI] [PubMed] [Google Scholar]
- [84].Kanaporis G, Blatter LA. Increased Risk for Atrial Alternans in Rabbit Heart Failure: The Role of Ca2+/Calmodulin-Dependent Kinase II and Inositol-1,4,5-trisphosphate Signaling. Biomolecules . 2023;14:53. doi: 10.3390/biom14010053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [85].Li C, Chan J, Haeseleer F, Mikoshiba K, Palczewski K, Ikura M, et al. Structural insights into Ca2+-dependent regulation of inositol 1,4,5-trisphosphate receptors by CaBP1. The Journal of Biological Chemistry . 2009;284:2472–2481. doi: 10.1074/jbc.M806513200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [86].Liu H, Liu X, Zhuang H, Fan H, Zhu D, Xu Y, et al. Mitochondrial Contact Sites in Inflammation-Induced Cardiovascular Disease. Frontiers in Cell and Developmental Biology . 2020;8:692. doi: 10.3389/fcell.2020.00692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [87].Piamsiri C, Fefelova N, Pamarthi SH, Gwathmey JK, Chattipakorn SC, Chattipakorn N, et al. Potential Roles of IP3 Receptors and Calcium in Programmed Cell Death and Implications in Cardiovascular Diseases. Biomolecules . 2024;14:1334. doi: 10.3390/biom14101334. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [88].Mosqueira M, Konietzny R, Andresen C, Wang C, H A Fink R. Cardiomyocyte depolarization triggers NOS-dependent NO transient after calcium release, reducing the subsequent calcium transient. Basic Research in Cardiology . 2021;116:18. doi: 10.1007/s00395-021-00860-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [89].Derler I, Butorac C, Krizova A, Stadlbauer M, Muik M, Fahrner M, et al. Authentic CRAC channel activity requires STIM1 and the conserved portion of the Orai N terminus. The Journal of Biological Chemistry . 2018;293:1259–1270. doi: 10.1074/jbc.M117.812206. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [90].Almanaitytė M, Jurevičius J, Mačianskienė R. Effect of Carvacrol, TRP Channels Modulator, on Cardiac Electrical Activity. BioMed Research International . 2020:6456805. doi: 10.1155/2020/6456805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [91].Harada M, Luo X, Qi XY, Tadevosyan A, Maguy A, Ordog B, et al. Transient receptor potential canonical-3 channel-dependent fibroblast regulation in atrial fibrillation. Circulation . 2012;126:2051–2064. doi: 10.1161/CIRCULATIONAHA.112.121830. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [92].Davis J, Burr AR, Davis GF, Birnbaumer L, Molkentin JD. A TRPC6-dependent pathway for myofibroblast transdifferentiation and wound healing in vivo. Developmental Cell . 2012;23:705–715. doi: 10.1016/j.devcel.2012.08.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [93].Ong HL, Subedi KP, Son GY, Liu X, Ambudkar IS. Tuning store-operated calcium entry to modulate Ca2+-dependent physiological processes. Biochimica et Biophysica Acta. Molecular Cell Research . 2019;1866:1037–1045. doi: 10.1016/j.bbamcr.2018.11.018. [DOI] [PubMed] [Google Scholar]
- [94].Sah R, Mesirca P, Van den Boogert M, Rosen J, Mably J, Mangoni ME, et al. Ion channel-kinase TRPM7 is required for maintaining cardiac automaticity. Proceedings of the National Academy of Sciences of the United States of America . 2013;110:E3037–E3046. doi: 10.1073/pnas.1311865110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [95].Sah R, Mesirca P, Mason X, Gibson W, Bates-Withers C, Van den Boogert M, et al. Timing of myocardial trpm7 deletion during cardiogenesis variably disrupts adult ventricular function, conduction, and repolarization. Circulation . 2013;128:101–114. doi: 10.1161/CIRCULATIONAHA.112.000768. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [96].Kjeldsen ST, Nissen SD, Saljic A, Hesselkilde EM, Carstensen H, Sattler SM, et al. Structural and electro-anatomical characterization of the equine pulmonary veins: implications for atrial fibrillation. Journal of Veterinary Cardiology: the Official Journal of the European Society of Veterinary Cardiology . 2024;52:1–13. doi: 10.1016/j.jvc.2024.01.001. [DOI] [PubMed] [Google Scholar]
- [97].Morciano G, Rimessi A, Patergnani S, Vitto VAM, Danese A, Kahsay A, et al. Calcium dysregulation in heart diseases: Targeting calcium channels to achieve a correct calcium homeostasis. Pharmacological Research . 2022;177:106119. doi: 10.1016/j.phrs.2022.106119. [DOI] [PubMed] [Google Scholar]
- [98].Liu T, Xiong F, Qi XY, Xiao J, Villeneuve L, Abu-Taha I, et al. Altered calcium handling produces reentry-promoting action potential alternans in atrial fibrillation-remodeled hearts. JCI Insight . 2020;5:e133754. doi: 10.1172/jci.insight.133754. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [99].Denham NC, Pearman CM, Caldwell JL, Madders GWP, Eisner DA, Trafford AW, et al. Calcium in the Pathophysiology of Atrial Fibrillation and Heart Failure. Frontiers in Physiology . 2018;9:1380. doi: 10.3389/fphys.2018.01380. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [100].Sadredini M, Haugsten Hansen M, Frisk M, Louch WE, Lehnart SE, Sjaastad I, et al. CaMKII inhibition has dual effects on spontaneous Ca2+ release and Ca2+ alternans in ventricular cardiomyocytes from mice with a gain-of-function RyR2 mutation. American Journal of Physiology. Heart and Circulatory Physiology . 2021;321:H446–H460. doi: 10.1152/ajpheart.00011.2021. [DOI] [PubMed] [Google Scholar]
- [101].Yang HQ, Zhou P, Wang LP, Zhao YT, Ren YJ, Guo YB, et al. Compartmentalized β1-adrenergic signalling synchronizes excitation-contraction coupling without modulating individual Ca2+ sparks in healthy and hypertrophied cardiomyocytes. Cardiovascular Research . 2020;116:2069–2080. doi: 10.1093/cvr/cvaa013. [DOI] [PubMed] [Google Scholar]
- [102].Christ T, Boknik P, Wöhrl S, Wettwer E, Graf EM, Bosch RF, et al. L-type Ca2+ current downregulation in chronic human atrial fibrillation is associated with increased activity of protein phosphatases. Circulation . 2004;110:2651–2657. doi: 10.1161/01.CIR.0000145659.80212.6A. [DOI] [PubMed] [Google Scholar]
- [103].Okazaki R, Iwasaki YK, Miyauchi Y, Hirayama Y, Kobayashi Y, Katoh T, et al. lipopolysaccharide induces atrial arrhythmogenesis via down-regulation of L-type Ca2+ channel genes in rats. International Heart Journal . 2009;50:353–363. doi: 10.1536/ihj.50.353. [DOI] [PubMed] [Google Scholar]
- [104].Qi XY, Vahdati Hassani F, Hoffmann D, Xiao J, Xiong F, Villeneuve LR, et al. Inositol Trisphosphate Receptors and Nuclear Calcium in Atrial Fibrillation. Circulation Research . 2021;128:619–635. doi: 10.1161/CIRCRESAHA.120.317768. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [105].Luo X, Pan Z, Shan H, Xiao J, Sun X, Wang N, et al. MicroRNA-26 governs profibrillatory inward-rectifier potassium current changes in atrial fibrillation. The Journal of Clinical Investigation . 2013;123:1939–1951. doi: 10.1172/JCI62185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [106].Njegic A, Wilson C, Cartwright EJ. Targeting Ca2+ Handling Proteins for the Treatment of Heart Failure and Arrhythmias. Frontiers in Physiology . 2020;11:1068. doi: 10.3389/fphys.2020.01068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [107].Wang Y, Morishima M, Li D, Takahashi N, Saikawa T, Nattel S, et al. Binge Alcohol Exposure Triggers Atrial Fibrillation Through T-Type Ca2+ Channel Upregulation via Protein Kinase C (PKC) / Glycogen Synthesis Kinase 3β (GSK3β) / Nuclear Factor of Activated T-Cells (NFAT) Signaling - An Experimental Account of Holiday Heart Syndrome. Circulation Journal: Official Journal of the Japanese Circulation Society . 2020;84:1931–1940. doi: 10.1253/circj.CJ-20-0288. [DOI] [PubMed] [Google Scholar]
- [108].Xia HJ, Dai DZ, Dai Y. Up-regulated inflammatory factors endothelin, NFkappaB, TNFalpha and iNOS involved in exaggerated cardiac arrhythmias in l-thyroxine-induced cardiomyopathy are suppressed by darusentan in rats. Life Sciences . 2006;79:1812–1819. doi: 10.1016/j.lfs.2006.06.007. [DOI] [PubMed] [Google Scholar]
- [109].Alsina KM, Hulsurkar M, Brandenburg S, Kownatzki-Danger D, Lenz C, Urlaub H, et al. Loss of Protein Phosphatase 1 Regulatory Subunit PPP1R3A Promotes Atrial Fibrillation. Circulation . 2019;140:681–693. doi: 10.1161/CIRCULATIONAHA.119.039642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [110].Fischer TH, Eiringhaus J, Dybkova N, Förster A, Herting J, Kleinwächter A, et al. Ca(2+) /calmodulin-dependent protein kinase II equally induces sarcoplasmic reticulum Ca(2+) leak in human ischaemic and dilated cardiomyopathy. European Journal of Heart Failure . 2014;16:1292–1300. doi: 10.1002/ejhf.163. [DOI] [PubMed] [Google Scholar]
- [111].Tateishi H, Yano M, Mochizuki M, Suetomi T, Ono M, Xu X, et al. Defective domain-domain interactions within the ryanodine receptor as a critical cause of diastolic Ca2+ leak in failing hearts. Cardiovascular Research . 2009;81:536–545. doi: 10.1093/cvr/cvn303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [112].Kanaporis G, Blatter LA. Increased Risk for Atrial Alternans in Rabbit Heart Failure: The Role of Ca2+/Calmodulin-Dependent Kinase II and Inositol-1,4,5-trisphosphate Signaling. Biomolecules . 2023;14:53. doi: 10.3390/biom14010053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [113].Zhang X, Shi S, Du Y, Chai R, Guo Z, Duan C, et al. Shaping cardiac destiny: the role of post-translational modifications on endoplasmic reticulum - mitochondria crosstalk in cardiac remodeling. Frontiers in Pharmacology . 2024;15:1423356. doi: 10.3389/fphar.2024.1423356. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [114].Gammons J, Trebak M, Mancarella S. Cardiac-Specific Deletion of Orai3 Leads to Severe Dilated Cardiomyopathy and Heart Failure in Mice. Journal of the American Heart Association . 2021;10:e019486. doi: 10.1161/JAHA.120.019486. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [115].Reese-Petersen AL, Olesen MS, Karsdal MA, Svendsen JH, Genovese F. Atrial fibrillation and cardiac fibrosis: A review on the potential of extracellular matrix proteins as biomarkers. Matrix Biology: Journal of the International Society for Matrix Biology . 2020;91:188–203. doi: 10.1016/j.matbio.2020.03.005. [DOI] [PubMed] [Google Scholar]
- [116].He R, Zhang J, Luo D, Yu Y, Chen T, Yang Y, et al. Upregulation of Transient Receptor Potential Canonical Type 3 Channel via AT1R/TGF-β1/Smad2/3 Induces Atrial Fibrosis in Aging and Spontaneously Hypertensive Rats. Oxidative Medicine and Cellular Longevity . 2019;2019:4025496. doi: 10.1155/2019/4025496. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [117].Zhang B, Jiang J, Yue Z, Liu S, Ma Y, Yu N, et al. Store-Operated Ca2+ Entry (SOCE) contributes to angiotensin II-induced cardiac fibrosis in cardiac fibroblasts. Journal of Pharmacological Sciences . 2016;132:171–180. doi: 10.1016/j.jphs.2016.05.008. [DOI] [PubMed] [Google Scholar]
- [118].Sabourin J, Beauvais A, Luo R, Montani D, Benitah JP, Masson B, et al. The SOCE Machinery: An Unbalanced Knowledge between Left and Right Ventricular Pathophysiology. Cells . 2022;11:3282. doi: 10.3390/cells11203282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [119].Lee TW, Chung CC, Lee TI, Lin YK, Kao YH, Chen YJ. Fibroblast Growth Factor 23 Stimulates Cardiac Fibroblast Activity through Phospholipase C-Mediated Calcium Signaling. International Journal of Molecular Sciences . 2021;23:166. doi: 10.3390/ijms23010166. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [120].Camacho Londoño JE, Marx A, Kraft AE, Schürger A, Richter C, Dietrich A, et al. Angiotensin-II-Evoked Ca2+ Entry in Murine Cardiac Fibroblasts Does Not Depend on TRPC Channels. Cells . 2020;9:322. doi: 10.3390/cells9020322. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [121].Yan J, Honglei Y, Yun W, Sheng D, Yun H, Anhua Z, et al. Puerarin ameliorates myocardial remodeling of spontaneously hypertensive rats through inhibiting TRPC6-CaN-NFATc3 pathway. European Journal of Pharmacology . 2022;933:175254. doi: 10.1016/j.ejphar.2022.175254. [DOI] [PubMed] [Google Scholar]
- [122].Rios FJ, Zou ZG, Harvey AP, Harvey KY, Camargo LL, Neves KB, et al. TRPM7 deficiency exacerbates cardiovascular and renal damage induced by aldosterone-salt. Communications Biology . 2022;5:746. doi: 10.1038/s42003-022-03715-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [123].Bao J, Gao Z, Hu Y, Ye L, Wang L. Transient receptor potential vanilloid type 1: cardioprotective effects in diabetic models. Channels (Austin, Tex.) . 2023;17:2281743. doi: 10.1080/19336950.2023.2281743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [124].Yang C, Guo W, He R, Meng X, Fu J, Lu Y. Dietary capsaicin attenuates cardiac injury after myocardial infarction in type 2 diabetic mice by inhibiting ferroptosis through activation of TRPV1 and Nrf2/HMOX1 pathway. International Immunopharmacology . 2024;140:112852. doi: 10.1016/j.intimp.2024.112852. [DOI] [PubMed] [Google Scholar]
- [125].Lkhagva B, Chang SL, Chen YC, Kao YH, Lin YK, Chiu CTH, et al. Histone deacetylase inhibition reduces pulmonary vein arrhythmogenesis through calcium regulation. International Journal of Cardiology . 2014;177:982–989. doi: 10.1016/j.ijcard.2014.09.175. [DOI] [PubMed] [Google Scholar]
- [126].Ni L, Lahiri SK, Nie J, Pan X, Abu-Taha I, Reynolds JO, et al. Genetic inhibition of nuclear factor of activated T-cell c2 prevents atrial fibrillation in CREM transgenic mice. Cardiovascular Research . 2022;118:2805–2818. doi: 10.1093/cvr/cvab325. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [127].Yuan M, Gong M, Zhang Z, Meng L, Tse G, Zhao Y, et al. Hyperglycemia Induces Endoplasmic Reticulum Stress in Atrial Cardiomyocytes, and Mitofusin-2 Downregulation Prevents Mitochondrial Dysfunction and Subsequent Cell Death. Oxidative Medicine and Cellular Longevity . 2020;2020:6569728. doi: 10.1155/2020/6569728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [128].Yuan M, Gong M, He J, Xie B, Zhang Z, Meng L, et al. IP3R1/GRP75/VDAC1 complex mediates endoplasmic reticulum stress-mitochondrial oxidative stress in diabetic atrial remodeling. Redox Biology . 2022;52:102289. doi: 10.1016/j.redox.2022.102289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [129].Ivanova H, Vervliet T, Monaco G, Terry LE, Rosa N, Baker MR, et al. Bcl-2-Protein Family as Modulators of IP3 Receptors and Other Organellar Ca2+ Channels. Cold Spring Harbor Perspectives in Biology . 2020;12:a035089. doi: 10.1101/cshperspect.a035089. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [130].Luo Y, Liu X, Ma R, Wang Y, Zimering M, Pan Z. Circulating IgGs in Type 2 Diabetes with Atrial Fibrillation Induce IP3-Mediated Calcium Elevation in Cardiomyocytes. iScience . 2020;23:101036. doi: 10.1016/j.isci.2020.101036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [131].Düzen IV, Yavuz F, Vuruskan E, Saracoglu E, Poyraz F, Göksülük H, et al. Leukocyte TRP channel gene expressions in patients with non-valvular atrial fibrillation. Scientific Reports . 2017;7:9272. doi: 10.1038/s41598-017-10039-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [132].Lou Q, Li L, Liu G, Li T, Zhang L, Zang Y, et al. Vericiguat reduces electrical and structural remodeling in a rabbit model of atrial fibrillation. Journal of Cardiovascular Pharmacology and Therapeutics . 2023;28:10742484231185252. doi: 10.1177/10742484231185252. [DOI] [PubMed] [Google Scholar]
- [133].Nikolova-Krstevski V, Wagner S, Yu ZY, Cox CD, Cvetkovska J, Hill AP, et al. Endocardial TRPC-6 Channels Act as Atrial Mechanosensors and Load-Dependent Modulators of Endocardial/Myocardial Cross-Talk. JACC. Basic to Translational Science . 2017;2:575–590. doi: 10.1016/j.jacbts.2017.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [134].Li S, Li M, Yi X, Guo F, Zhou Y, Chen S, et al. TRPM7 channels mediate the functional changes in cardiac fibroblasts induced by angiotensin II. International Journal of Molecular Medicine . 2017;39:1291–1298. doi: 10.3892/ijmm.2017.2943. [DOI] [PubMed] [Google Scholar]
- [135].Rajput FA, Du L, Chappell RJ, Berei TJ, Goldberger ZD, Wright JM. Comparative Efficacy and Safety of Intravenous Verapamil and Diltiazem for Rate Control in Rapidly Conducted Atrial Fibrillation and Atrial Flutter. Journal of General Internal Medicine . 2020;35:3721–3723. doi: 10.1007/s11606-020-06051-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [136].Peng DW, Lai YY, Luo XS, Li X, Deng CY, Guo HM, et al. Connexin 43 participates in atrial electrical remodelling through colocalization with calcium channels in atrial myocytes. Clinical and Experimental Pharmacology & Physiology . 2022;49:25–34. doi: 10.1111/1440-1681.13580. [DOI] [PubMed] [Google Scholar]
- [137].Wang J, Ye Q, Bai S, Chen P, Zhao Y, Ma X, et al. Inhibiting microRNA-155 attenuates atrial fibrillation by targeting CACNA1C. Journal of Molecular and Cellular Cardiology . 2021;155:58–65. doi: 10.1016/j.yjmcc.2021.02.008. [DOI] [PubMed] [Google Scholar]
- [138].Li Q, Fang Y, Peng DW, Li LA, Deng CY, Yang H, et al. Sacubitril/valsartan reduces susceptibility to atrial fibrillation by improving atrial remodeling in spontaneously hypertensive rats. European Journal of Pharmacology . 2023;952:175754. doi: 10.1016/j.ejphar.2023.175754. [DOI] [PubMed] [Google Scholar]
- [139].Kashiwa A, Makiyama T, Kohjitani H, Maurissen TL, Ishikawa T, Yamamoto Y, et al. Disrupted CaV1.2 selectivity causes overlapping long QT and Brugada syndrome phenotypes in the CACNA1C-E1115K iPS cell model. Heart Rhythm . 2023;20:89–99. doi: 10.1016/j.hrthm.2022.08.021. [DOI] [PubMed] [Google Scholar]
- [140].Kashiwa A, Makiyama T, Kohjitani H, Maurissen TL, Ishikawa T, Yamamoto Y, et al. Disrupted CaV1.2 selectivity causes overlapping long QT and Brugada syndrome phenotypes in the CACNA1C-E1115K iPS cell model. Heart Rhythm . 2023;20:89–99. doi: 10.1016/j.hrthm.2022.08.021. [DOI] [PubMed] [Google Scholar]
- [141].Jiménez-Sábado V, Casabella-Ramón S, Llach A, Gich I, Casellas S, Ciruela F, et al. Beta-blocker treatment of patients with atrial fibrillation attenuates spontaneous calcium release-induced electrical activity. Biomedicine & Pharmacotherapy = Biomedecine & Pharmacotherapie . 2023;158:114169. doi: 10.1016/j.biopha.2022.114169. [DOI] [PubMed] [Google Scholar]
- [142].Zhang T, Wu Y, Hu Z, Xing W, Kun LV, Wang D, et al. Small-Molecule Integrated Stress Response Inhibitor Reduces Susceptibility to Postinfarct Atrial Fibrillation in Rats via the Inhibition of Integrated Stress Responses. The Journal of Pharmacology and Experimental Therapeutics . 2021;378:197–206. doi: 10.1124/jpet.121.000491. [DOI] [PubMed] [Google Scholar]
- [143].Dorey TW, Liu Y, Jansen HJ, Bohne LJ, Mackasey M, Atkinson L, et al. Natriuretic Peptide Receptor B Protects Against Atrial Fibrillation by Controlling Atrial cAMP Via Phosphodiesterase 2. Circulation. Arrhythmia and Electrophysiology . 2023;16:e012199. doi: 10.1161/CIRCEP.123.012199. [DOI] [PubMed] [Google Scholar]
- [144].Liang F, Fan P, Jia J, Yang S, Jiang Z, Karpinski S, et al. Inhibitions of late INa and CaMKII act synergistically to prevent ATX-II-induced atrial fibrillation in isolated rat right atria. Journal of Molecular and Cellular Cardiology . 2016;94:122–130. doi: 10.1016/j.yjmcc.2016.04.001. [DOI] [PubMed] [Google Scholar]
- [145].Rao F, Xue YM, Wei W, Yang H, Liu FZ, Chen SX, et al. Role of tumour necrosis factor-a in the regulation of T-type calcium channel current in HL-1 cells. Clinical and Experimental Pharmacology & Physiology . 2016;43:706–711. doi: 10.1111/1440-1681.12585. [DOI] [PubMed] [Google Scholar]
- [146].Salari S, Silverå Ejneby M, Brask J, Elinder F. Isopimaric acid - a multi-targeting ion channel modulator reducing excitability and arrhythmicity in a spontaneously beating mouse atrial cell line. Acta Physiologica (Oxford, England) . 2018;222 doi: 10.1111/apha.12895. [DOI] [PubMed] [Google Scholar]
- [147].Shoemaker MB, Yoneda ZT, Crawford DM, Akers WS, Richardson T, Montgomery JA, et al. A Mechanistic Clinical Trial Using (R)- Versus (S)-Propafenone to Test RyR2 (Ryanodine Receptor) Inhibition for the Prevention of Atrial Fibrillation Induction. Circulation. Arrhythmia and Electrophysiology . 2022;15:e010713. doi: 10.1161/CIRCEP.121.010713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [148].Xu D, Murakoshi N, Tajiri K, Duo F, Okabe Y, Murakata Y, et al. Xanthine oxidase inhibitor febuxostat reduces atrial fibrillation susceptibility by inhibition of oxidized CaMKII in Dahl salt-sensitive rats. Clinical Science (London, England: 1979) . 2021;135:2409–2422. doi: 10.1042/CS20210405. [DOI] [PubMed] [Google Scholar]
- [149].Yang H, Zhu J, Fu H, Shuai W. Dapansutrile Ameliorates Atrial Inflammation and Vulnerability to Atrial Fibrillation in HFpEF Rats. Heart, Lung & Circulation . 2024;33:65–77. doi: 10.1016/j.hlc.2023.09.017. [DOI] [PubMed] [Google Scholar]
- [150].Kambayashi R, Goto A, Shinozaki M, Izumi-Nakaseko H, Takei Y, Iwata K, et al. In vivo cardiovascular profile of ryanodine receptor 2 inhibitor M201-A: Utility as an anti-atrial fibrillatory drug for patients suffering from heart failure with preserved ejection fraction. Journal of Pharmacological Sciences . 2024;156:171–179. doi: 10.1016/j.jphs.2024.08.004. [DOI] [PubMed] [Google Scholar]
- [151].Xiao J, Liang D, Zhao H, Liu Y, Zhang H, Lu X, et al. 2-Aminoethoxydiphenyl borate, a inositol 1,4,5-triphosphate receptor inhibitor, prevents atrial fibrillation. Experimental Biology and Medicine (Maywood, N.J.) . 2010;235:862–868. doi: 10.1258/ebm.2010.009362. [DOI] [PubMed] [Google Scholar]
- [152].Zhang J, Dong F, Ju G, Pan X, Mao X, Zhang X. Sodium Houttuyfonate Alleviates Monocrotaline-induced Pulmonary Hypertension by Regulating Orai1 and Orai2. American Journal of Respiratory Cell and Molecular Biology . 2024;71:332–342. doi: 10.1165/rcmb.2023-0015OC. [DOI] [PubMed] [Google Scholar]
- [153].Souza Bomfim GH, Niemeyer BA, Lacruz RS, Lis A. On the Connections between TRPM Channels and SOCE. Cells . 2022;11:1190. doi: 10.3390/cells11071190. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [154].Lee CJ, Lee SH, Kang BS, Park MK, Yang HW, Woo SY, et al. Effects of L-Type Voltage-Gated Calcium Channel (LTCC) Inhibition on Hippocampal Neuronal Death after Pilocarpine-Induced Seizure. Antioxidants (Basel, Switzerland) . 2024;13:389. doi: 10.3390/antiox13040389. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [155].Feng S, Gui J, Qin B, Ye J, Zhao Q, Guo A, et al. Resveratrol Inhibits VDAC1-Mediated Mitochondrial Dysfunction to Mitigate Pathological Progression in Parkinson’s Disease Model. Molecular Neurobiology . 2024 doi: 10.1007/s12035-024-04234-0. (online ahead of print) [DOI] [PubMed] [Google Scholar]
- [156].Li L, Li YQ, Sun ZW, Xu CM, Wu J, Liu GL, et al. Qingyi decoction protects against myocardial injuries induced by severe acute pancreatitis. World Journal of Gastroenterology . 2020;26:1317–1328. doi: 10.3748/wjg.v26.i12.1317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [157].Quan C, Li M, Du Q, Chen Q, Wang H, Campbell D, et al. SPEG Controls Calcium Reuptake Into the Sarcoplasmic Reticulum Through Regulating SERCA2a by Its Second Kinase-Domain. Circulation Research . 2019;124:712–726. doi: 10.1161/CIRCRESAHA.118.313916. [DOI] [PubMed] [Google Scholar]
- [158].Zhang M, Sun X, Zhao F, Chen Z, Liu M, Wang P, et al. Tinglu Yixin granule inhibited fibroblast-myofibroblast transdifferentiation to ameliorate myocardial fibrosis in diabetic mice. Journal of Ethnopharmacology . 2024;337:118980. doi: 10.1016/j.jep.2024.118980. [DOI] [PubMed] [Google Scholar]
- [159].Liu HZ, Wu W. Effect of SKF96365 on Myocardial Fibrosis in Type-II Diabetic Rats. Frontiers in Bioscience (Landmark Edition) . 2023;28:231. doi: 10.31083/j.fbl2809231. [DOI] [PubMed] [Google Scholar]