Open access peer-reviewed chapter

SK Channels and Heart Disease

Written By

Katherine Zhong, Shawn Kant, Frank Sellke and Jun Feng

Submitted: 06 February 2022 Reviewed: 02 March 2022 Published: 13 April 2022

DOI: 10.5772/intechopen.104115

From the Edited Volume

Ion Transporters - From Basic Properties to Medical Treatment

Edited by Zuzana Sevcikova Tomaskova

Chapter metrics overview

204 Chapter Downloads

View Full Metrics

Abstract

Extensive evidence indicates that small-conductance Ca2+-activated K+ channels (SK channels) help regulate cardiac rhythm and myocardial function in physiological and pathophysiological conditions. This chapter will begin by discussing the basic physiology of SK channel expression, localization, and activation under normal conditions, before proceeding to address the impact of SK channel dysfunction on a variety of cardiac pathologies including atrial fibrillation (AF), ventricular arrhythmias (VA), cardiac hypertrophy/heart failure (HF) and myocardial ischemia/reperfusion (IR) injury. The critical role of aberrant SK channel regulation will also be discussed to establish unifying mechanisms of SK channel pathology across these different conditions. Several animal model and human tissue experiments suggest that pharmacologic modulation of SK channel function may be beneficial in controlling AF, VA, cardiomyopathy and myocardial IR injury. Therefore, targeting SK channels may represent a promising new therapeutic avenue for treating a variety of cardiovascular disease states.

Keywords

  • SK channel
  • small-conductance Ca2+-activated K+ channels
  • cardiac rhythm
  • cardiac function
  • atrial fibrillation
  • ventricular arrhythmias
  • cardiac hypertrophy
  • heart failure
  • ischemia reperfusion injury

1. Introduction

The coordinated activity of cell membrane ion channels forms the basis of the cardiomyocyte action potential sequence, driving myocardial contractility and resulting in cardiac output. Each stage of the myocardial action potential relies on opening and closing of specific groups of ion channels to propagate excitatory stimuli through the heart [1, 2, 3, 4]. Most myocardial action potentials begin with transmission of spontaneous impulses generated by pacemaker cells (e.g., the sinoatrial node [SAN] and atrioventricular node [AVN]) to atrial and ventricular myocytes that are at resting membrane potential (phase 4, around −85 to −90 mV). Sinus node impulses reach atrial myocytes via intercellular gap junctions, leading to depolarization to threshold. Myocytes then transition to phase 0, a fast upstroke driven by opening of voltage-gated sodium channels that facilitate a strong inward sodium current that rapidly depolarizes the cardiomyocyte to a membrane potential of over 20 mV, at which point sodium channels close and transient outward potassium channels open, producing a brief repolarizing current (phase 1).

However, the strongly depolarized cardiomyocyte membrane also triggers opening of L-type voltage gated calcium channels (LTCC) that facilitate influx of calcium. Calcium influx balances potassium efflux through transient outward and delayed-rectifier potassium channels (which also open during this period), leading to the phase 2 plateau of the cardiac action potential. It is during phase 2 that calcium entering cardiomyocytes triggers calcium-induced calcium release from sarcoplasmic reticulum (SR) stores via calcium-induced calcium release, mediated by calcium interaction with ryanodine receptor (RyR) calcium channels located on the SR [2]. Increasing levels of intracellular calcium relieves troponin-tropomyosin inhibition of cardiomyocyte actin-myosin cross bridge formation, leading to cross-bridge cycling and myocardial contraction. Finally, cardiomyocytes transition to phase 3 repolarization, where calcium channels close and delayed rectifier potassium channels remain open, allowing membrane potential to return to its resting state (phase 4) and terminating the action potential.

The repolarization phase of the myocardial action potential has received extensive attention in biomedical research because aberrations in this phase are intimately related to disruptions in myocardial excitability and overall control of synchronous, regulated myocardial contraction. Until recently, phase 3 repolarization was thought to be almost entirely driven by delayed rectifier potassium channels that opened alongside calcium channels during phase 2 and remained open following calcium channel closure, subsequently driving membrane potential towards the potassium equilibrium potential.

However, a growing body of evidence suggests that another type of potassium channel, the small-conductance Ca2+-activated K+ channels (SK channel), might also have an important role in myocardial repolarization [5, 6, 7, 8]. SK channels are cell membrane potassium channels that are exquisitely sensitive to calcium, opening in response to elevated intracellular calcium and facilitating an outward potassium current. Several studies have shown that SK channels are active during the late repolarization phase of the cardiac action potential, likely assisting delayed rectifier potassium channels by amplifying the outward potassium current [5, 6]. On their own, SK channels may represent a form of feedback control of excess myocardial excitability, given their close responsiveness to calcium—which governs cardiomyocyte contraction.

Conversely, SK channel dysfunction may contribute to cardiovascular pathology in a manner reminiscent of how dysfunction of other major myocardial ion channels (especially the delayed rectifier potassium channel) is implicated in cardiovascular disease. Indeed, the past two decades have given rise to an abundance of research attempting to characterize the role of SK channels in numerous cardiovascular disease states and explore whether modulation of SK channels may be a potential therapeutic tool that deserves clinical attention. Hence the focus of this chapter, which will discuss the history and basic physiology of myocardial SK channels before delving into the current literature concerning SK channel dysfunction in four major areas: atrial fibrillation (AF), ventricular tachyarrhythmias (VA), heart failure (HF), and ischemia-reperfusion (IR) injury.

Advertisement

2. Identification of SK channels in the heart

Small-conductance Ca2+-activated K+ channels (SK channels), encoded by KCNN genes, are a family of K+ channels that have a small single channel conductance (10–20 pS with symmetrical solutions), and are gated by intracellular Ca2+ concentrations [9]. SK channels were identified using apamin, a bee venom toxin that selectively blocks SK channels [9]. SK channels were first identified in rat and human brain tissue [10]. Since their discovery, SK channels have been identified in a variety of tissues including the nervous system, blood, epithelial cells, skeletal muscle and endothelial (vessel) cells [11, 12]. SK channels are classified into three isoforms: SK1, SK2 and SK3, based on their gene origins and different sensitivities to apamin. The SK1 channel is encoded by the KCNN1 gene, located on chromosome 19, and is moderately sensitive to apamin. The SK2 channel is encoded by the KCNN2 gene located on chromosome 5 and has the strongest affinity for apamin. The SK3 channel is encoded by the KCNN3 gene located on chromosome 1 and is moderately sensitive to apamin [8, 13]. The intermediate-conductance Ca2+-activated K+ channel (IK or SK4) is similar to SK channels and is encoded by the KCNN4 gene. SK4 has a slightly larger single conductance (12–42 pS) and a higher affinity for intracellular Ca2+ than other SK channels [8, 13].

SK channels were first identified in the heart in 2003 [14, 15], and their distribution and functions in heart tissue have since been extensively studied [16, 17, 18]. SK channels have been found in both atrial and ventricular tissues in animals and humans [7, 12, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33]. In mouse atrial and ventricular myocytes, quantification of SK1 and SK3 transcripts showed a higher level of SK1 expression in atria versus ventricles, while SK3 is expressed at a simar level in atria and ventricles [12]. Apamin-sensitive Ca2+-activated K+ currents (SK currents) have been detected in the pulmonary veins (PV), Bachmann’s bundle (BB), sinoatrial node (SAN) myocytes, and mouse atrioventricular nodes (AVN) [34, 35, 36]. SK channels have also been found expressed intracellularly in the ventricular mitochondrial membranes of guinea pigs, rats and humans [37, 38]. Indeed, transcripts of both SK2 and SK3 isoforms are found in the inner mitochondrial membrane (IMM), but not SK1 [38]. Notably, although there are fewer SK channels in the ventricles, under certain pathological conditions such as HF and chronic myocardial infarction (MI), SK currents in ventricles are upregulated [39, 40]. The results are summarized in Table 1.

SpeciesTissue/cardiomyocytemRNAProteinIon channel recordingSK channelsReference
MouseAtrial and ventricular myocytesHigher level of expression of SK1 in atria versus ventricles; SK3 is expressed at a simar level in bothProtein expression of SK1 and low level of SK3SK current is more prominent in atria than in ventriclesSK1, SK3[12]
Anatomical atrioventricular nodes (AVN)N/ASK2 channel protein detected in AVN cellsSK current recorded in AVN cellsSK2[36]
SAN cellsPresence of the transcripts of SK1, SK2 and SK3All three isoforms are present and are preferentially distributed along the Z line, especially SK3SK currents recordedSK1, SK2, SK3[41]
RabbitPulmonary vein (PV) and Bachmann’s bundle (BB)SK2 and SK3 mRNA detected in both PV and BBPresence of SK2 and SK3 protein in both PV and BBSK current recorded in both PV and BBSK2, SK3[6]
VentriclesN/AN/ASK currents are not significant in normal ventricles but are upregulated in failing ventriclesN/A[39]
Pulmonary vein (PV) and sinoatrial node (SAN) myocyteN/AN/ASAN exhibits greater SK currents than PVN/A[35]
VentriclesN/ASK2 are more abundantly present in the Purkinje cells (PCs) than in the ventricular myocytesApamin prolonged APD in PCs. SK current density was larger in PCs than in ventricular myocytesSK2[42]
Guinea pigVentricular mitochondriaN/ASK2 and SK3 channel proteins were present in IMMSK current recordedSK2, SK3[16]
HumanRight and left atrial appendagesmRNAs of SK1, SK2 and SK3 were detectedSK1, SK2 and SK3 were expressed in human atrial myocytesSK current detectedSK1, SK2, SK3[27]
Atrial and ventricular tissueSK1 had a higher level of expression in atria compared to ventricle, while the expression level of SK2 and SK3 were not different in atria vs. ventricleN/AN/ASK1, SK2, SK3[25]
Atrial tissuesN/ASK3 polypeptide level was unchanged pre- and post-cardioplegic arrestN/ASK3[19]
Right atrial tissuesN/AInsignificant decrease in SK3 protein level in diabetic myocardium vs. nondiabeticsDiabetes significantly reduced SK currentsSK3[20]
Right atrial tissuesN/ANeither cardioplegic arrest nor diabetes resulted in a significant change of SK3 protein levelN/ASK3[21]
Human atrial tissue, coronary arterioles and coronary artery endothelial cellsNo significant difference in SK3 mRNA levels in diabetic atrial tissues and endothelial cells vs. nondiabeticsNo significant difference in SK3 protein levels in diabetic atrial tissues and endothelial cells vs. nondiabeticsN/ASK3[33]
Multiple speciesMouse atrial and ventricular and human atrial myocytesPresence of SK2 mRNA in human and mouse cardiac myocytesPresence of SK2 channelSK current was more prominent in atria versus ventriclesSK2[14]
Guinea pig, rat and human ventricular mitochondriamRNAs of SK2 and SK3, but not SK1, are present in guinea pig ventricular myocytes. SK3 mRNAs are present in human ventricular tissues.SK3 is expressed in human and guinea pig ventricular IMMN/ASK3[38]

Table 1.

Functional expression of SK channels in the heart.

Advertisement

3. Physiological functions of SK channels in the heart

In neuronal cells, SK channels contribute to the slow afterhyperpolarization following an action potential [43]. The physiologic functions of SK channels in the heart were not clear until Xu et al. first reported their role in the repolarization of cardiac myocytes. Apamin-treated cardiac cells show significantly longer action potential durations (APD) due to slower repolarization compared to control groups, indicating SK channel involvement in cardiac repolarization. These effects are more prominent in atrial versus ventricular cells, likely due to the higher level of expression of SK channels in atria [14]. In another study, the AVN in SK2-null mutant mice show decreased firing frequency and longer APD compared with controls, while mice overexpressing SK2 channels show shorter APD [36]. Apamin application results in decreased action potential firing frequency akin to that seen in SK2-null mutants [36].

Consistently, overexpression of SK3 channels also results in significantly shorter APD, suggesting a similar role for different isoforms of SK channels in the repolarization process in cardiac cells [44]. The prolongation of APD resulting from apamin has also been recorded in human right and left atrial appendages, rabbit PV and SAN, and mouse SAN cells [27, 35, 41]. These studies consistently show that SK channels contribute to repolarization and shorten the action potentials in normal hearts, especially in the atria where they are more densely expressed than in the ventricles. In SAN, SK blockade also leads to significant depolarization of the maximal diastolic potential (MDP) and a decrease in the diastolic depolarization slope [41]. Notably, in PVs with intact endothelium, SK channels contribute to hyperpolarization and vessel dilation [35]. Figure 1 shows the cardiac action potential including the QT interval affected by SK channels.

Figure 1.

Cardiac action potential. P = atrial depolarization; QRS = ventricular depolarization; T = ventricular repolarization. The significance of the U wave is largely unknown, although may represent Purkinje fiber repolarization. The QT interval indicated by the dark bar between “Q” and “T”. SK channels have been shown to influence the duration of the QT interval of cardiomyocyte action potentials. Diagram courtesy of ECGpedia.org (https://en.ecgpedia.org/wiki/Action_potential) [45].

Meanwhile, with respect to mitochondria, DCBE, an SK channel opener, reduces mitochondrial injury when given before cardiac ischemia, suggesting that SK channel opening may protect the heart and mitochondria against ischemia-reperfusion (IR) injury. SK channels have also been found to reduce oxidative stress by reducing mitochondrial Ca2+ overload [38].

Advertisement

4. Signaling pathway

The Ca2+ that activates SK channels comes from various intracellular and extracellular sources. In hippocampal neurons, SK channels are coupled to the LTCC using microdomains of submembrane calcium, and Ca2+ influx through LTCC activates SK channels [46]. Similar co-localization of LTCC and SK channels can be found in cardiac myocytes, where SK2 channels couple with LTCCs through α-actinin2. This suggests that, in cardiac myocytes, SK2 channels are activated by local subsarcolemmal Ca2+ that enters the cell through LTCCs [47]. Besides LTCCs, the SR also plays a role in providing the Ca2+ needed for activating SK channels. In mouse cardiac myocytes, the type 2 ryanodine receptor (RyR2) mediates intracellular Ca2+ release from the SR, which activates SK channels [48]. It is possible that in cardiac myocytes, SK channels could be activated by either Ca2+ influx through voltage-gated Ca2+ channels or through release of intracellular Ca2+ from SR stores.

SK channels are gated by intracellular Ca2+ through interactions between the channel α-subunits and calmodulin (CaM) [49]. The intracellular C-terminal domain of the SK channels immediately adjacent to the sixth transmembrane segment (S6) consists of four α-helices, named A, B, C and D, that are critical to the channel gating. CaM binds to channel regions A-D constitutively, whereas CaM binding to regions B-C and B-D is Ca2+ dependent. This indicates that regions B-C and B-D are involved in the Ca2+ gating mechanism [49]. Ca2+ binds to the EF hands in the N-lobe of CaM, which leads to conformational changes and opening of SK channels [50].

Besides Ca2+-activated channel gating, CaM is also critical in the regulation of cell surface expression of the SK channels, independent of the binding of Ca2+ [51]. Several cytoskeletal proteins, including α-actinin2, filamin A, and MLC2, are important in SK2 channel trafficking [52, 53, 54]. Notably, cell membrane localization of SK2 channels is Ca2+-dependent when the channels are co-expressed with α-actinin2. An increase in intracellular Ca2+ such as in AF is predicted to increase the expression of SK2 channels and lead to shortened APD and maintenance of the arrhythmias [54].

Casein kinase 2 (CK2) and protein phosphatase 2A (PP2A) are critical components of the SK channels that regulate the Ca2+ sensitivity of the channels by phosphorylating or dephosphorylating CaM. CK2 decreases the Ca2+ sensitivity of closed SK channels, while PP2A increases the Ca2+ sensitivity of open SK channels [9]. Increased expression of PP2A and decreased co-localization of CK2 with SK2 may be the underlying mechanism of increased sensitivity of apamin-sensitive K+ current to intracellular Ca2+ in HF, as shown in a volume-overload HF rat model [55].

Various mechanisms of SK channel regulation may have clinical significance with respect to SK channel pathology seen in certain disease states. For example, upregulation of SK channels in ventricular myocytes in cardiac hypertrophy has been shown to result from phosphorylation by both calcium/calmodulin-dependent protein kinase II (CaMKII) and protein kinase A (PKA) [28, 56]. In addition, microRNA (miRNA) also plays a role in cardiac SK channels regulation. MicroRNA 499 (miR-499) is upregulated in atrial tissue from patients with permanent AF, resulting in increased binding to the KCNN3 gene and downregulation of SK2 channel expression [26].

Advertisement

5. SK channels in heart disease

5.1 Atrial fibrillation

5.1.1 Overview

Atrial fibrillation (AF), a condition characterized by rapid and disorganized atrial activation, is the most common cardiac arrhythmia with a prevalence of 1–2% among the general population [57]. A variety of risk factors have been linked to the development of AF, such as age, male sex, obesity, hypertension, heart failure, structural heart disease (valvulopathy, CHF, MI), diabetes, hyperthyroidism, and family history [58]. Of these factors, age and sex carry the highest risk: males have a 1.5–2× greater chance of developing AF than females, and individuals between ages 40–55 carry a lifetime risk of 22–26% [59].

AF is classified into several categories based on the duration of the AF episode(s). These include paroxysmal AF, persistent AF, long-standing persistent AF, and permanent AF [60]. The term “lone AF” remains a diagnosis of exclusion and has been used to describe AF in younger patients without a prior history of structural heart disease or cardiovascular risk factors [61]. The specific molecular and electrophysiological mechanisms that underly AF are highly complex and remain poorly understood. Nonetheless, most current conceptual frameworks of the pathogenesis of AF involve a combination of structural remodeling, electrical remodeling, and autonomic remodeling that generate atrial reentry circuits, rotors (localized electrical spiral waves), and ectopic impulse generation [57].

Regarding electrical remodeling, a variety of altered ionic currents can be seen in AF, including increased activity of LTCC, increased activity of inward rectifier potassium channels (KiR), and decreased function of gap junctions [62, 63, 64, 65]. Excessive LTCC activity may trigger excessive cardiomyocyte SR Ca2+ release via RyR activation, leading to hyperexcitability. Excessive KiR activity may alter atrial myocyte resting potential and phase 3 activation, resulting in reduced atrial refractoriness. Gap junction defects may slow atrial conduction velocity, which when combined with atrial myocyte hyperexcitability and decreased refractoriness favors reentry, ectopic foci, and initiation of AF.

Structural remodeling in AF largely involves fibrosis and atrial dilation. Atrial fibrosis may be due to several factors such as aging, myocardial infarction, volume overload, or aberrant renin-angiotensin-aldosterone system (RAAS) activity [66, 67, 68, 69, 70]. Proliferation of fibroblasts and extracellular matrix in fibrotic atria may create barriers to electrical conduction that interfere with cardiomyocyte electrical coupling, creating conduction abnormalities and variable action potential durations that predispose to ectopic activity and reentrant circuits [71, 72]. In addition, larger atria have increased odds of developing reentrant circuits [73]. Finally, autonomic remodeling in AF involves increased sympathetic or parasympathetic tone in the atria that influences atrial tachypacing [74, 75, 76, 77].

5.1.2 SK channels and AF heritability

The heritability of AF has been widely studied over the past two decades, and lone AF appears to have a greater heritable component than structural AF [78]. Recent genome-wide association studies (GWAS) have identified several single nucleotide polymorphisms (SNPs) that are associated with increased risk of AF. Some examples include rs2200733 on chromosome 4q25, rs2106261 on chromosome 16q22, rs10824026 on chromosome 10q22, and rs7394190 on chromosome 10q22 [61, 78, 79]. Overall, GWAS studies of AF atrial tissue have found mutations in genes coding for ion channels, gap junction connexin proteins, nuclear membrane components, calcium homeostasis, and cardio-genesis [80].

Potassium channel gene mutations are of special interest in AF because of the important role of potassium channels in maintaining the resting membrane potential and facilitating repolarization after generation of the action potential. For example, gain of function mutations have been found in KCNQ1 (the alpha subunit of the inward rectifier cardiac potassium channel Kv7.1), KCNH2 (voltage gated potassium channel Kv11.1), KCND3 (alpha subunit of the voltage-gated potassium channel Kv4.3), KCNJ2 (inward rectifying potassium channel Kir2.1), and KCNA5 (voltage-gated potassium channel Kv1.5) [81, 82, 83, 84, 85]. All of these changes lead to increased inward potassium currents in atrial myocytes, which may shorten APD, QT intervals, and the effective refractory periods. In 2010, Ellinor et al. in their GWAS discovered a novel AF susceptibility locus on chromosome 1q21 among Caucasian individuals. The most significant SNP at this locus, rs13376333, can be found in the intron between the first and second exon of KCNN3, which encodes the SK3 channel [61]. Rs13376333 is strongly associated with lone AF, bearing an odds ratio of 1.56 (P = 6.3 × 10−12). The association of rs13376333 with more typical forms of AF is also significant, albeit weaker (OR = 1.13, P = 0.006) [61].

Similar results have been reported by several other labs using genomic data from identical or different ethnic groups. Indeed, Chang et al. report significant associations between rs13376333 and risk of lone (OR 3.02) and structural (OR 2.18) AF among a Taiwanese cohort, with even stronger odds ratios than those reported by Ellinor et al. [86]. Among a Han Chinese cohort, Luo et al. found that the frequency of rs13376333 at KCNN3 was significantly higher in lone AF than in controls, although there were no significant differences in rs13376333 frequency between total AF patients and controls [87]. Curiously, an earlier study by Li et al. failed to produce a significant association between rs13376333 and AF among Han Chinese AF individuals [88]. It is possible that different ethnic demographic groups may exhibit different risk propensities with respect to SNPs, which would imply contribution of some form of gene-environment interplay that either diminishes or enhances genetic effects. Given the complexity of AF as a disease process, this is a very likely scenario and requires further investigations involving larger sample sizes and different demographic groups. Besides rs13376333, another SNP at KCNN3, rs1131820, has also been associated with increased risk of lone AF in individuals of Danish ethnicity, with an OR of 2.85 [89].

In addition to KCNN3, a weak association of AF with KCNN2 has also been discovered in a Han Chinese cohort, at the SNP rs13184658 [90]. How variations in KCNN alleles specifically affect the risk of AF remains unclear. A recent study revealed that rs13376333 is associated with increased mRNA expression of KCNN3 in human atrial tissue [91]. In addition, given that most other potassium channel mutations observed in AF are gain of function mutations, perhaps the SK channel SNPs follow the same trend. Further study of the KCNN3 locus could help reveal pathways underlying the association between KCNN3 and AF, with the potential of developing novel treatments of AF that target KCNN3.

5.1.3 SK expression and electrophysiology in AF

At the mRNA and protein levels, there is mixed evidence about whether SK channels are over or under-expressed in atrial remodeling in human and animal models of AF. Ozgen et al. first reported that SK channels are associated with initiation of atrial remodeling. Using a burst-paced rabbit atrium, they showed that SK2 mRNA and protein levels were upregulated in the region of the left atrium where the pulmonary veins empty; this suggests that SK channels have a role in burst-pacing induced APD shortening [34]. Likewise, Qi et al. reported upregulation of SK1 and SK2 protein induced by atrial tachypacing in dog PVs and left atrial cells; SK2 mRNA expression was also increased, although no significant changes in SK1 mRNA were observed, suggesting that overexpression of SK1 may be the result of posttranslational modifications or altered membrane trafficking of SK1 channels [23]. In dopamine tautomerase-deficient mice induced to exhibit AF by apamin administration, Tsai et al. observed increased protein and mRNA levels of SK1 and SK3 in mouse right atrial tissue [92].

These results seem to contradict other studies that reported decreased expression of SK channels in human and animal AF. Darkow et al. found decreased expression of KCNN2 (SK2) mRNA in atrial tissue of patients with AF when compared with healthy controls [25]. Similarly, transcripts of KCNN1–3 were downregulated in paroxysmal AF and chronic AF patients at a similar level [93]. Rahm et al. also examined KCNN mRNA expression in a pig model of atrial tachypacing-induced AF with reduced left ventricular function and found decreased expression of KCNN2 and KCNN3 with normal levels of KCNN1 [93]. Furthermore, Fan et al. report decreased mRNA and protein levels of SK1, SK2, and SK3 in atrial appendage tissue from humans with chronic AF [32]. Another study of atrial tissue from chronic AF patients by Yu et al. replicated these results with respect to SK1 and SK2 mRNA and protein expression, although no differences were observed in SK3 expression at either level [27]. Finally, Skisbye et al. observed reduced SK2 and SK3 expression in chronic AF human atrial tissue, with no changes in SK1 expression [7]. The research findings of the up/down regulation of SK channels in AF were summarized in Table 2.

SpeciesModelUp/down regulation of SK channelsReference
RabbitBurst-paced rabbit atriumSK2 mRNA and protein levels were upregulated in the left atrium[34]
DogAtrial tachypacing in dog pulmonary veins and left atrial cellsUpregulation of SK1 and SK2 protein; upregulated SK2 mRNA expression[23]
MouseInduced AF in dopamine tautomerase-deficient miceIncreased protein and mRNA levels of SK1 and SK3 in right atrial tissue[92]
PigAtrial tachypacing-induced AF with reduced left ventricular functionDecreased expression of KCNN2 and KCNN3 with normal levels of KCNN1[93]
HumanPatients with AFDecreased expression of SK2 mRNA in atrial tissue[25]
Paroxysmal AF and chronic AF patientsTranscripts of KCNN1–3 were downregulated[93]
Atrial appendage tissue from chronic AF patientsDecreased mRNA and protein levels of SK1, SK2, and SK3[32]
Atrial tissue from chronic AF patientsDecreased mRNA and protein levels of SK1, SK2; no difference in SK3 expression levels[27]
Atrial tissue from chronic AF patientsReduced SK2 and SK3 expression; no changes in SK1 expression[7]

Table 2.

Up/down regulation of SK channels in atrial fibrillation (AF).

Ozgen et al. first discovered that burst pacing induced increased SK2 channel trafficking to the cell membrane and increased SK currents in rabbit pulmonary veins [34]. In their dog model, Qi et al. found that atrial tachypacing enhanced SK currents and single-channel open probabilities [23]. In contrast, Yu et al. reported decreased SK currents in right and left atrial appendage tissue from patients with chronic AF, alongside decreased SK1 and SK2 mRNA and protein levels (discussed earlier) [27]. Another possible mechanism of increased SK currents in AF is enhanced activity of CaMKII, which is required for SK channel calcium-dependent activation. CaMKII exhibited significantly increased expression alongside, increased intracellular calcium levels in human AF tissue studied by Fan et al. [32].

Approached from a different angle, a mouse model of SK3 overexpression also showed considerable shortening of APD and increased SK channel currents in atrial myocytes [44]. Likewise, mice engineered to overexpress SK2 channels displayed significant atrioventricular nodal dysfunction, manifesting as increased firing frequency, and shortening of spontaneous action potential, while SK2 ablation eliminated these effects [36]. Furthermore, SK2 knockout mouse models studied by Li et al. exhibited the opposite effects: significant prolongation of atrial myocyte APD among homozygous and heterozygous mutants, with homozygous mutants having an increased susceptibility to AF [94].

Inconsistencies regarding observed SK channel expression and activity in AF across different studies might be due to a variety of factors including the specific method of AF induction, differences in atrial tissue characteristics among different species of animal models, patient population demographic differences, and different stages or durations of AF. It is also possible that SK channels are initially upregulated in AF before being downregulated due to atrial remodeling. There are several reported mechanisms by which SK channels are regulated in AF, including histone deacetylase related epigenetic mechanisms [95], miRNA [26], and CaMKII [32].

5.1.4 Pharmacologic modulation of SK channels in AF

Pharmacologic studies of SK channel modulation provide additional insights into the role of SK channels in AF, although the results are complicated. First, the IK antagonist HMR1556 preserved hemodynamic stability in pigs induced by atrial burst pacing to express persistent AF [96]. At time of sacrifice, HMR1556 treated pigs also exhibited significantly higher left ventricular ejection fraction than untreated pigs, along with significantly longer right atrial APD [96]. Next, application of the SK channel inhibitor apamin increased spontaneous action potential generation in isolated rabbit PVs while decreasing spontaneous activity and prolonging APD in sinoatrial nodal myocytes [35]. However, in isolated canine left atrial tissue, apamin treatment significantly increased APD heterogeneity and proved to be pro-arrythmogenic [24].

The SK channel inhibitor NS8593 increased APD and effective refractory in right atrial appendage tissue of patients with AF [7]. Identical results were found by Qi et al. in a dog model of AF [23]. Haugaard et al. and Burashnikov et al. examined the effects of NS8593, along with another SK channel inhibitor (UCL1684), in human and equine atrial myocytes, and verified the ability of both inhibitors to reduce AF inducibility or terminate induced AF [97, 98]. However, conflicting results were found by Fenner et al. in their horse model of tachypacing-induced persistent AF [99]. There, delayed right atrial conduction after NS8593 treatment actually increased AF complexity through increased anisotropy and electrical dissociation [99]. Meanwhile, the left atrium exhibited no change at all in AF complexity, and neither left nor right atrium ultimately resulted in cardioversion [99].

Moving on, treatment with SK2 inhibitor AP30663 in atrially tachypaced live AF pigs resulted in conversion to sinus rhythm, increased right atrial effective refractory periods, and prevented reinduction of AF [100]. Additional studies using whole-cell and inside-out patch clamp recordings of guinea-pig heart tissue confirmed a right-shift of the calcium-activation curve of SK2 channels in the presence of AP30663, with concentration-dependent prolongation of atrial refractoriness and minor effects of QT prolongation [91].

Finally, Saljic et al. tested the use of an antisense oligonucleotide GapmeR in rats and showed that GapmeR downregulates SK3 protein expression in the heart and provides protection against AF [101]. Though targeting the expression level of SK channel seems promising, Darkow et al., showed that that SK3 was upregulated in ventricular tissue in heart failure patients, suggesting that SK channels are not likely to be an atria-selective target as previously expected [25].

In a recent study, Gatta et al. conducted a detailed examination of the effects of SK channel inhibitor AP14145 on goat hearts induced to AF after 30 days of burst-pacing stimulation delivered by pericardial electrodes implanted above the left atria [102]. The authors found that AP14145 produced dose-dependent prolongation of AF cycle length and increased the effective refractory period of atrial impulses [102]. Interestingly, atrial conduction velocity in AF following AP14145 treatment remained unchanged until the final seconds before AF termination, where sudden organization of fibrillatory conduction occurred prior to AF cardioversion [102].

Most animal models of AF discussed up to this point involve AF induction via simple electrode-delivered burst-pacing to atrial tissue. However, other approaches also exist. For example, Yan et al. studied an atrial stretch-induced rabbit AF model [103]. As discussed earlier, atrial enlargement increases the risk of AF by shortening atrial effective refractory periods. Hence the authors of this study placed an inflatable balloon into their rabbit heart left atria to mechanically dilate atria to various sizes. Sustained AF was induced by brief delivery of burst pacing to dilated atria that produced rapid irregular atrial rhythms. For the experimental group, the SK inhibitor ICA was applied to rabbit hearts before atrial stretch and burst pacing. Final analyses showed that ICA pretreatment significantly attenuated stretch-induced atrial effective refractory period and reduced overall AF inducibility and duration when compared with untreated hearts [103]. Another alternative approach was taken by Celotto et al., who focused on the role of autonomic dysfunction in AF pathogenesis [104]. Using human atrial cell and tissue models, the authors induced AF via high dose acetylcholine administration, which shortened atrial APD [104]. SK channel blockade was able to partially revert APD shortening due to acetylcholine, while a combination of SK blockade and the adrenergic agonist isoproterenol were able to completely reverse APD shortening back to pre-AF baseline [104].

How SK channel inhibitors compare to current mainstay antiarrhythmic medications is another important question with significant clinical implications. Two studies by Kirchhoff et al. provides some insights into this issue [105, 106].

In one study, Kirchhoff et al. explored the effect of combining SK channel inhibition via ICA and voltage-gated sodium channel inhibition on AF in atrial-burst pacing induced AF guinea pigs [105]. Ultimately, AF combining ICA with normally subefficacious concentrations of flecainide or ranolazine (sodium channel blockers) reduced AF duration [105].

Next, Kirchhoff et al. examined the effect of the SK inhibitor ICA on AF in an atrial burst-pacing guinea pig model when used alongside amiodarone or dofetilide, two major class III antiarrhythmics. The authors found that combining ICA with either dofetilite or amiodarone reduced AF duration at normally sub-optimal concentrations of all three drugs if used individually [106]. In addition, ICA combined with a standard therapeutic dose of dofetilide prevented QT prolongation that is often seen with dofetilide monotherapy [106]. Overall, both studies suggest that combining SK channel inhibitors with traditional antiarrhythmics may provide a useful synergistic benefit that allows for reducing doses of traditional antiarrhythmics. This in turn may help mitigate against adverse effects of high-dose antiarrhythmics (e.g. long QT, VA).

Note that when all available evidence (pharmacologic, omic, and knockout studies) is considered together, over and underactivity of SK channels both appear to increase likelihood of developing AF. Perhaps discrepancies between different studies are the result of differences in experimental techniques or different species. Alternatively, a two-fold mechanism may drive SK pathology in AF. On one hand, gradually increasing potassium currents due to increased insertion or presence of SK channel in cell membranes may contribute to the observed accelerated actional potential refractory periods and steady progression of APD shortening. On the other hand, action potential prolongation due to diminished potassium currents in the absence of SK channel activity may increase likelihood of generating early afterdepolarizations that may evolve into ectopic foci or reentry circuits. If correct, this proposal implies that seeking to modulate SK channel activity in AF is not a simple matter of complete pharmacologic blockade or complete potentiation. Rather, the objective would be restoring balance towards a homeostatic level of SK channel activity. Further research is required to define the precise pathophysiology of altered SK channel activity in AF to better guide development and protocols for new therapeutic tactics.

5.2 SK channels in ventricular arrhythmias and heart failure

Ventricular arrhythmias (VA) are abnormal heart rhythms that include premature ventricular complexes (PVC), non-sustained ventricular tachycardia (VT), accelerated idioventricular rhythm, and sustained VT or ventricular fibrillation (VF). VT and VF in particular may cause sudden cardiac death (SCD) [107]. Ventricular arrhythmias can be seen in patients with structurally normal hearts, but malignant ventricular arrhythmias usually occur in patients with underlying structural heart disease including HF, ischemic cardiomyopathy, and nonischemic cardiomyopathy [107, 108].

In the normal heart, SK1 channels are expressed at a higher level in atria versus in ventricles, while SK2 and SK3 are expressed at a similar level in both chambers [14, 25]. Normally, apamin-sensitive currents are more prominent in atria than ventricles [12, 14]. However, under certain pathological conditions such as HF and chronic MI, SK currents are upregulated in ventricles, suggesting that SK channels play an important role in ventricular repolarization and VA in pathologic hearts [39, 40].

In healthy ventricles, apamin does not alter APD [14, 22]. Under pathological conditions such as HF, chronic MI, cardiac hypertrophy, and hypokalemia, SK channel blockers prolong APD in ventricles as shown in human and animal models [29, 39, 40, 105, 109]. In acute MI, however, the results are mixed. Some studies show that SK channel blockade prolongs APD in rat acute MI models [110, 111], but a recent study using a porcine model of acute MI showed no significant effect of APD alteration by SK blockade [112]. In the studies that showed prolonged APD by SK blockade in pathologic ventricles, the effects could be either antiarrhythmic or proarrhythmic, probably due to different baseline heart rhythms in the specific animal models used for the studies [18].

Antiarrhythmic effects of SK blockade may occur through attenuation of APD shortening and reduction of repolarization heterogeneity in pathologic hearts [112]. Indeed, Chua et al. were the first to show proarrhythmic effects of SK channels in the ventricles. They showed that HF heterogeneously increased the SK channel’s sensitivity to intracellular Ca2+ and upregulated SK currents, which led to APD shortening and recurrent spontaneous VF in a rabbit model of tachycardia-induced HF [39]. In this scenario, the rapid heart rate in heart failure caused APD shortening, and excessive APD shortening was arrhythmogenic. APD shortening led to increased intracellular Ca2+, which activated SK currents and further shortened APD, resulting in late phase 3 early afterdepolarization and recurrent spontaneous VF [39]. The antiarrhythmic effects of SK blockers in the ventricles have also been demonstrated in human HF [29], in rabbits with chronic MI [40], in rats with acute MI [110, 111], in rats with cardiac hypertrophy [113, 114], and in hypokalemic guinea pig heart [100].

Chen et al. reported that SK channels are proarrhythmic and play a role in inducing J wave syndrome (JWS). They showed that concurrent activation of SK currents and inhibition of Na+ currents shortened APD and induced JWS and SVF in rabbit hearts. SK channel blockade in this rabbit model was antiarrhythmic—JWS was reduced and SVF was abolished, suggesting that SK current activation contributed to the development JWS and SVF in rabbit ventricles [109]. A recent study showed that colocalization of LTCC and SK channels in ventricular myocytes activates SK currents, which then promote phase 2 reentry and T-wave alternans, leading to JWS and VA [115].

Although SK blockers have potential antiarrhythmic benefits in the management of ventricular arrhythmias, blocking SK channels may carry significant proarrhythmic risk in patients with underlying heart disease such as HF, MI, and cardiac hypertrophy, based on animal model studies [116, 117, 118]. Blocking SK channels might reduce the repolarization reserve in patients, trigger early after depolarizations (EADs), increase risk of developing torsade’s de pointes (TdP) and induce fatal arrhythmia [18, 112, 116]. In addition, SK blockers might be proarrhythmic in hypokalemic hearts. Chan et al. reported that hypokalemia activates SK channels, shortening APD and maintaining the repolarization reserve at late activation sites. In their rabbit model of hypokalemic ventricles, apamin was proarrhythmic by prolonging APD at late activation sites and inducing VF [119].

In addition, Wan et al. showed that SK blockade might even interfere with ventricular automaticity in normal ventricles [120]. Although SK channels do not participate in repolarization in healthy ventricles, SK currents and SK2 protein are prominent in Purkinje cells in normal rabbit ventricles [42]. Wan et al. reported that apamin accelerated ventricular escape rhythms from the Purkinje fibers, enhanced ventricular automaticity and led to VT in normal rabbit ventricles [120].

To summarize, the heterogenous activation of SK channels is proarrhythmic and contributes to the development of ventricular arrythmia in diseased hearts. SK blockers such as apamin have some antiarrhythmic benefits, but also carry significant proarrhythmic risks, thus limiting the practical use of SK blockers for managing ventricular arrythmia. Furthermore, SK channels are widely expressed in the human body including in the nervous system, so blocking SK channels may have undesired off-target effects. Alternatively, drugs that target the signaling pathway of SK channels might have antiarrhythmic effects without directly blocking the channel. Given that increased intracellular Ca2+ triggers the upregulation of SK channel in pathologic ventricles, drugs that affect the interactions of SK channel and intracellular Ca2+ might have antiarrhythmic or proarrhythmic effects. For example, β-blockers, a known treatment for ventricular arrhythmias, downregulate SK1 and SK3 expression, the SK channel’s sensitivity to Ca2+, and the SK current density as shown in a volume-overload rat model [121]. Kamada et al. showed that β- adrenoreceptor stimulation activated SK channels via CaMKII activity in hypertrophied rat hearts, which might contribute to the antiarrhythmic effects of β-blockers [122]. Finally, recent studies showed that sex differences existed with respect to ventricular SK channel activation in response to autonomic stimulation [109, 123], which might have important clinical implications for drug efficacy and safety.

5.3 SK channels and ischemia/reperfusion

During cell ischemia and hypoxia, the altered redox state leads to excess reactive oxygen species (ROS) production which overwhelms the ROS scavenger system, causing mitochondrial Ca2+ overload that results in cell apoptosis and necrosis [124]. Given that mitochondria are important in ROS production, ion channels present on the mitochondrial membranes could contribute to the regulation of homeostasis by regulating ROS production during ischemia [124]. In 2013, Stowe et al. discovered that SK channels were located in the guinea pig cardiac IMM, and that SK channel opening had a protective effect during ischemia [37]. In their study, the SK and IK channel opener DCEB resulted in decreased infarct size, reduced superoxide (O2) and mitochondrial Ca2+ levels, and more normal NADH and FAD levels. The protective effects were reduced when TBAP, a dismutator of O2 was added, suggesting that the benefits of channel openers were related to ROS production [37].

Later on, in a separate study, Stowe et al. also showed that SK channel opening improved contractility and reduced infarct size during ischemia/reperfusion (IR) [125]. In their guinea pig heart model of global IR injury, the SK channel opener DCEB improved contractile function, while SK antagonists worsened contractility and increased infarct size. Furthermore, in cardiac mitochondria after IR, combined SK channel and large conductance Ca2+-activated K+ (BK) channel agonists improved respiratory control index and Ca2+ retention capacity, while the combined antagonists worsened Ca2+ retention capacity [125]. Once again, these results show that SK channel plays a role in regulating homeostasis of mitochondria and reducing cell damage during IR.

In 2017, Yang et al., showed that SK3 channels were located in the mitochondria of guinea pig, rat and human ventricular myocytes. They reported that SK channel agonists were protective against IR injury while SK antagonists worsened IR injury. Overexpression of SK3.1 specifically increased Ca2+-activated K+ uptake in mouse atrial tumor cells and protected the cells from hypoxia/reoxygenation injury. Consistently, silencing SK3.1 channel expression exacerbated cell injury and death [38]. Hence the authors conclude that the protective effect of SK channels during IR suggests their role in reducing oxidative stress resulting from mitochondrial Ca2+ overload [38].

In addition, in hypertrophic hearts, mitochondrial SK channels also appear to have protective benefits by decreasing mitochondrial ROS production as shown in a rat model [126]. Kim et al., reported that SK channel enhancers reversed the oxidation of RyRs, improved RyR function and stabilized SR Ca2+ release, leading to the protective effects of SK channels in hypertrophic heart [126]. To summarize, mitochondrial SK channel have important protective effects during cardiac cell ischemia, hypoxia and hypertrophy by regulating Ca2+ homeostasis and ROS production in the mitochondria.

Advertisement

6. Conclusion

It is increasingly evident that SK channels have important roles in myocardial physiology and pathophysiology. While different studies report different expression levels of various SK channel isoforms in atrial vs. ventricular tissue, all SK channel isoforms present in the heart are necessary for promoting atrial and ventricular repolarization following myocardial impulse generation. The high calcium sensitivity of SK channels, mediated by channel-bound calmodulin and modulated by important regulators such CK2 and PP2A, renders them crucial for feedback control of myocardial contractility and prevention of runaway excitation. In the context of arrythmias, SK channel polymorphisms confer increased risk of atrial fibrillation, and both hyper- and hypoactivity of SK channels along with aberrant SK channel expression likely contribute to automaticity, re-entry circuits, and ectopic pacemaker activity that drives atrial and ventricular tachyarrythmias; these effects are exacerbated in the context of underlying heart disease, such as congestive heart failure. Furthermore, SK channel activity appears to have a protective effect in mitigating oxidative stress during ischemia, with particular significance given to myocardial mitochondrial SK channels.

Pharmacologic studies of SK channel inhibition or activation show promise for treating many animal models of arrythmias and ischemia-reperfusion, although results are still not consistent across different models and different protocols; hence additional research will be required prior to clinical trials of SK channel antagonists in humans. In the future, more research on SK channel pathology using human atrial or ventricular tissue will also be necessary to complement and verify cellular/molecular findings from animal models. Likewise, the specific mechanisms behind altered SK channel expression in many of the diseases discussed in this chapter remain murky and must be elucidated to better characterize specific aspects of pathology at play.

References

  1. 1. Grant AO. Cardiac ion channels. Circulation. Arrhythmia and Electrophysiology. 2009;2(2):185-194
  2. 2. Amin AS, Tan HL, Wilde AAM. Cardiac ion channels in health and disease. Heart Rhythm. 2010;7(1):117-126
  3. 3. Rahm A-K, Lugenbiel P, Schweizer PA, Katus HA, Thomas D. Role of ion channels in heart failure and channelopathies. Biophysical Reviews. 2018;10(4):1097-1106
  4. 4. Wei X, Yohannan S, Richards JR. Physiology, cardiac repolarization dispersion and reserve. In: StatPearls [Internet]. Treasure Island (FL): StatPearls Publishing; 2022. Available from: http://www.ncbi.nlm.nih.gov/books/NBK537194/
  5. 5. Hancock JM, Weatherall KL, Choisy SC, James AF, Hancox JC, Marrion NV. Selective activation of heteromeric SK channels contributes to action potential repolarization in mouse atrial myocytes. Heart Rhythm. 2015;12(5):1003-1015
  6. 6. Gu M, Zhu Y, Yin X, Zhang D-M. Small-conductance Ca2+−activated K+ channels: Insights into their roles in cardiovascular disease. Experimental & Molecular Medicine. 2018;50(4):1-7
  7. 7. Skibsbye L, Poulet C, Diness JG, Bentzen BH, Yuan L, Kappert U, et al. Small-conductance calcium-activated potassium (SK) channels contribute to action potential repolarization in human atria. Cardiovascular Research. 2014;103(1):156-167
  8. 8. Weisbrod D. Small and intermediate calcium activated potassium channels in the heart: Role and strategies in the treatment of cardiovascular diseases. Frontiers in Physiology. 2020;11:590534
  9. 9. Adelman JP, Maylie J, Sah P. Small-conductance Ca2+−activated K+ channels: Form and function. Annual Review of Physiology. 2012;74:245-269
  10. 10. Köhler M, Hirschberg B, Bond CT, Kinzie JM, Marrion NV, Maylie J, et al. Small-conductance, calcium-activated potassium channels from mammalian brain. Science. 1996;273(5282):1709-1714
  11. 11. Stocker M. Ca(2+)-activated K+ channels: Molecular determinants and function of the SK family. Nature Reviews. Neuroscience. 2004;5(10):758-770
  12. 12. Tuteja D, Xu D, Timofeyev V, Lu L, Sharma D, Zhang Z, et al. Differential expression of small-conductance Ca2+−activated K+ channels SK1, SK2, and SK3 in mouse atrial and ventricular myocytes. American Journal of Physiology. Heart and Circulatory Physiology. 2005;289(6):H2714-H2723
  13. 13. Wei AD, Gutman GA, Aldrich R, Chandy KG, Grissmer S, Wulff H. International Union of Pharmacology. LII. Nomenclature and molecular relationships of calcium-activated potassium channels. Pharmacological Reviews. 2005;57(4):463-472
  14. 14. Xu Y, Tuteja D, Zhang Z, Xu D, Zhang Y, Rodriguez J, et al. Molecular identification and functional roles of a Ca(2+)-activated K+ channel in human and mouse hearts. The Journal of Biological Chemistry. 2003;278(49):49085-49094
  15. 15. Xu Y, Zhang Z, Timofeyev V, Sharma D, Xu D, Tuteja D, et al. The effects of intracellular Ca2+ on cardiac K+ channel expression and activity: Novel insights from genetically altered mice. The Journal of Physiology. 2005;562(Pt 3):745-758
  16. 16. Zhang X-D, Lieu DK, Chiamvimonvat N. Small-conductance Ca2+−activated K+ channels and cardiac arrhythmias. Heart Rhythm. 2015;12(8):1845-1851
  17. 17. Zhang X-D, Thai PN, Lieu DK, Chiamvimonvat N. Cardiac small-conductance calcium-activated potassium channels in health and disease. Pflügers Archiv. 2021;473(3):477-489
  18. 18. Chang P-C, Chen P-S. SK channels and ventricular arrhythmias in heart failure. Trends in Cardiovascular Medicine. 2015;25(6):508-514
  19. 19. Feng J, Liu Y, Clements RT, Sodha NR, Khabbaz KR, Senthilnathan V, et al. Calcium-activated potassium channels contribute to human coronary microvascular dysfunction after cardioplegic arrest. Circulation. 2008;118(Suppl. 14):S46-S51
  20. 20. Liu Y, Xie A, Singh AK, Ehsan A, Choudhary G, Dudley S, et al. Inactivation of endothelial small/intermediate conductance of calcium-activated potassium channels contributes to coronary arteriolar dysfunction in diabetic patients. Journal of the American Heart Association. 2015;4(8):e002062
  21. 21. Liu Y, Cole V, Lawandy I, Ehsan A, Sellke FW, Feng J. Decreased coronary arteriolar response to KCa channel opener after cardioplegic arrest in diabetic patients. Molecular and Cellular Biochemistry. 2018;445(1):187-194
  22. 22. Nagy N, Szuts V, Horváth Z, Seprényi G, Farkas AS, Acsai K, et al. Does small-conductance calcium-activated potassium channel contribute to cardiac repolarization? Journal of Molecular and Cellular Cardiology. 2009;47(5):656-663
  23. 23. Qi X-Y, Diness JG, Brundel BJJM, Zhou X-B, Naud P, Wu C-T, et al. Role of small-conductance calcium-activated potassium channels in atrial electrophysiology and fibrillation in the dog. Circulation. 2014;129(4):430-440
  24. 24. Hsueh C-H, Chang P-C, Hsieh Y-C, Reher T, Chen P-S, Lin S-F. Proarrhythmic effect of blocking the small conductance calcium activated potassium channel in isolated canine left atrium. Heart Rhythm. 2013;10(6):891-898
  25. 25. Darkow E, Nguyen TT, Stolina M, Kari FA, Schmidt C, Wiedmann F, et al. Small conductance Ca2 +−activated K+ (Sk) channel mRNA expression in human atrial and ventricular tissue: Comparison between donor, atrial fibrillation and heart failure tissue. Frontiers in Physiology. 2021;12:650964
  26. 26. Ling T-Y, Wang X-L, Chai Q, Lau T-W, Koestler CM, Park SJ, et al. Regulation of the SK3 channel by microRNA-499--potential role in atrial fibrillation. Heart Rhythm. 2013;10(7):1001-1009
  27. 27. Yu T, Deng C, Wu R, Guo H, Zheng S, Yu X, et al. Decreased expression of small-conductance Ca2+−activated K+ channels SK1 and SK2 in human chronic atrial fibrillation. Life Sciences. 2012;90(5–6):219-227
  28. 28. Hamilton S, Polina I, Terentyeva R, Bronk P, Kim TY, Roder K, et al. PKA phosphorylation underlies functional recruitment of sarcolemmal SK2 channels in ventricular myocytes from hypertrophic hearts. The Journal of Physiology. 2020;598(14):2847-2873
  29. 29. Chang P-C, Turker I, Lopshire JC, Masroor S, Nguyen B-L, Tao W, et al. Heterogeneous upregulation of apamin-sensitive potassium currents in failing human ventricles. Journal of the American Heart Association. 2013;2(1):e004713
  30. 30. Diness JG, Sørensen US, Nissen JD, Al-Shahib B, Jespersen T, Grunnet M, et al. Inhibition of small-conductance Ca2+-activated k + channels terminates and protects against atrial fibrillation. Circulation. Arrhythmia and Electrophysiology. 2010;3(4):380-390
  31. 31. Mahida S, Mills RW, Tucker NR, Simonson B, Macri V, Lemoine MD, et al. Overexpression of KCNN3 results in sudden cardiac death. Cardiovascular Research. 2014;101(2):326-334
  32. 32. Fan X, Yu Y, Lan H, Ou X, Yang L, Li T, et al. Ca2+/calmodulin-dependent protein kinase ii (Camkii) increases small-conductance ca2+-activated k+ current in patients with chronic atrial fibrillation. Medical Science Monitor. 2018;24:3011-3023
  33. 33. Liu Y, Kabakov AY, Xie A, Shi G, Singh AK, Sodha NR, et al. Metabolic regulation of endothelial SK channels and human coronary microvascular function. International Journal of Cardiology. 2020;312:1-9
  34. 34. Ozgen N, Dun W, Sosunov E, Anyukhovsky E, Hirose M, Duffy H, et al. Early electrical remodeling in rabbit pulmonary vein results from trafficking of intracellular SK2 channels to membrane sites. Cardiovascular Research. 2007;75(4):758-769
  35. 35. Chen W-T, Chen Y-C, Lu Y-Y, Kao Y-H, Huang J-H, Lin Y-K, et al. Apamin modulates electrophysiological characteristics of the pulmonary vein and the sinoatrial node. European Journal of Clinical Investigation. 2013;43(9):957-963
  36. 36. Zhang Q, Timofeyev V, Lu L, Li N, Singapuri A, Long MK, et al. Functional roles of a Ca2+-activated K+ channel in atrioventricular nodes. Circulation Research. 2008;102(4):465-471
  37. 37. Stowe DF, Gadicherla AK, Zhou Y, Aldakkak M, Cheng Q, Kwok W-M, et al. Protection against cardiac injury by small Ca2+−sensitive K+ channels identified in Guinea pig cardiac inner mitochondrial membrane. Biochimica et Biophysica Acta (BBA)—Biomembranes. 2013;1828(2):427-442
  38. 38. Yang M, Camara AKS, Aldakkak M, Kwok W-M, Stowe DF. Identity and function of a cardiac mitochondrial small conductance Ca2+ −activated K+ channel splice variant. Biochimica et Biophysica Acta (BBA)—Bioenergetics. 2017;1858(6):442-458
  39. 39. Chua S-K, Chang P-C, Maruyama M, Turker I, Shinohara T, Shen MJ, et al. Small-conductance calcium-activated potassium channel and recurrent ventricular fibrillation in failing rabbit ventricles. Circulation Research. 2011;108(8):971-979
  40. 40. Lee YS, Chang P-C, Hsueh C-H, Maruyama M, Park HW, Rhee K-S, et al. Apamin-sensitive calcium-activated potassium currents in rabbit ventricles with chronic myocardial infarction: SK currents in chronic mi. Journal of Cardiovascular Electrophysiology. 2013;24(10):1144-1153
  41. 41. Torrente AG, Zhang R, Wang H, Zaini A, Kim B, Yue X, et al. Contribution of small conductance K+ channels to sinoatrial node pacemaker activity: Insights from atrial-specific Na+/Ca2+ exchange knockout mice: Role of small conductance K + channels in sinus node activity. The Journal of Physiology. 2017;595(12):3847-3865
  42. 42. Reher TA, Wang Z, Hsueh C-H, Chang P-C, Pan Z, Kumar M, et al. Small-conductance calcium-activated potassium current in normal rabbit cardiac purkinje cells. Journal of the American Heart Association. 2017;6(6):e005471
  43. 43. Bond CT. Small conductance Ca2+−activated K+ channel knock-out mice reveal the identity of calcium-dependent afterhyperpolarization currents. Journal of Neuroscience. 2004;24(23):5301-5306
  44. 44. Zhang X-D, Timofeyev V, Li N, Myers RE, Zhang D-M, Singapuri A, et al. Critical roles of a small conductance Ca2+−activated K+ channel (Sk3) in the repolarization process of atrial myocytes. Cardiovascular Research. 2014;101(2):317-325
  45. 45. Action potential—en.ecgpedia.org. n.d. Available from: https://en.ecgpedia.org/wiki/Action_potential [Accessed: February 18, 2022]
  46. 46. Marrion NV, Tavalin SJ. Selective activation of Ca2+-activated K+ channels by co-localized Ca2+ channels in hippocampal neurons. Nature. 1998;395(6705):900-905
  47. 47. Lu L, Zhang Q, Timofeyev V, Zhang Z, Young JN, Shin H-S, et al. Molecular coupling of a Ca2+-activated K+ channel to L-type Ca2+ channels via α-actinin2. Circulation Research. 2007;100(1):112-120
  48. 48. Mu Y, Zhao W, Duan P, Chen Y, Zhao W, Wang Q, et al. Ryr2 modulates a Ca2+−activated K+ current in mouse cardiac myocytes. Attali B, editor. PLoS One. 2014;9(4):e94905
  49. 49. Xia X-M, Fakler B, Rivard A, Wayman G, Johnson-Pais T, Keen JE, et al. Mechanism of calcium gating in small-conductance calcium-activated potassium channels. Nature. 1998;395(6701):503-507
  50. 50. Schumacher MA, Rivard AF, Bächinger HP, Adelman JP. Structure of the gating domain of a Ca2+-activated K+ channel complexed with Ca2+/calmodulin. Nature. 2001;410(6832):1120-1124
  51. 51. Maylie J, Bond CT, Herson PS, Lee W-S, Adelman JP. Small conductance Ca2+-activated K+ channels and calmodulin: SK channels and calmodulin. The Journal of Physiology. 2004;554(2):255-261
  52. 52. Lu L, Timofeyev V, Li N, Rafizadeh S, Singapuri A, Harris TR, et al. Actinin2 cytoskeletal protein is required for the functional membrane localization of a Ca2+-activated K+ channel (SK2 channel). Proceedings of the National Academy of Sciences. 2009;106(43):18402-18407
  53. 53. Lu L, Sirish P, Zhang Z, Woltz RL, Li N, Timofeyev V, et al. Regulation of gene transcription by voltage-gated l-type calcium channel, cav1. 3. Journal of Biological Chemistry. 2015;290(8):4663-4676
  54. 54. Rafizadeh S, Zhang Z, Woltz RL, Kim HJ, Myers RE, Lu L, et al. Functional interaction with filamin A and intracellular Ca2+ enhance the surface membrane expression of a small-conductance Ca2+-activated K+ (Sk2) channel. Proceedings of the National Academy of Sciences of the United States of America. 2014;111(27):9989-9994
  55. 55. Yang D, Wang T, Ni Y, Song B, Ning F, Hu P, et al. Apamin-sensitive K+ current upregulation in volume-overload heart failure is associated with the decreased interaction of CK2 with SK2. Journal of Membrane Biology. 2015;248(6):1181-1189
  56. 56. Mizukami K, Yokoshiki H, Mitsuyama H, Watanabe M, Tenma T, Takada S, et al. Small-conductance Ca2+-activated K + current is upregulated via the phosphorylation of CaMKII in cardiac hypertrophy from spontaneously hypertensive rats. American Journal of Physiology-Heart and Circulatory Physiology. 2015;309(6):H1066-H1074
  57. 57. Pellman J, Sheikh F. Atrial fibrillation: Mechanisms, therapeutics, and future directions. Comprehensive Physiology. 2015;5(2):649-665
  58. 58. Brandes A, Smit MD, Nguyen BO, Rienstra M, Van Gelder IC. Risk factor management in atrial fibrillation. Arrhythmia & Electrophysiology Review. 2018;7(2):118-127
  59. 59. Andrade J, Khairy P, Dobrev D, Nattel S. The clinical profile and pathophysiology of atrial fibrillation: Relationships among clinical features, epidemiology, and mechanisms. Circulation Research. 2014;114(9):1453-1468
  60. 60. January CT, Wann LS, Alpert JS, Calkins H, Cigarroa JE, Cleveland JC, et al. 2014 aha/acc/hrs guideline for the management of patients with atrial fibrillation. Journal of the American College of Cardiology. 2014;64(21):e1-e76
  61. 61. Ellinor PT, Lunetta KL, Glazer NL, Pfeufer A, Alonso A, Chung MK, et al. Common variants in KCNN3 are associated with lone atrial fibrillation. Nature Genetics. 2010;42(3):240-244
  62. 62. Nattel S, Dobrev D. The multidimensional role of calcium in atrial fibrillation pathophysiology: Mechanistic insights and therapeutic opportunities. European Heart Journal. 2012;33(15):1870-1877
  63. 63. Pandit SV, Berenfeld O, Anumonwo JMB, Zaritski RM, Kneller J, Nattel S, et al. Ionic determinants of functional reentry in a 2-d model of human atrial cells during simulated chronic atrial fibrillation. Biophysical Journal. 2005;88(6):3806-3821
  64. 64. Voigt N, Maguy A, Yeh Y-H, Qi X, Ravens U, Dobrev D, et al. Changes in IK,ACh single-channel activity with atrial tachycardia remodelling in canine atrial cardiomyocytes. Cardiovascular Research. 2007;77(1):35-43
  65. 65. Igarashi T, Finet JE, Takeuchi A, Fujino Y, Strom M, Greener ID, et al. Connexin gene transfer preserves conduction velocity and prevents atrial fibrillation. Circulation. 2012;125(2):216-225
  66. 66. Dun W, Boyden PA. Aged atria: Electrical remodeling conducive to atrial fibrillation. Journal of Interventional Cardiac Electrophysiology. 2009;25(1):9-18
  67. 67. Nishida K, Qi XY, Wakili R, Comtois P, Chartier D, Harada M, et al. Mechanisms of atrial tachyarrhythmias associated with coronary artery occlusion in a chronic canine model. Circulation. 2011;123(2):137-146
  68. 68. Deroubaix E, Folliguet T, Rücker-Martin C, Dinanian S, Boixel C, Validire P, et al. Moderate and chronic hemodynamic overload of sheep atria induces reversible cellular electrophysiologic abnormalities and atrial vulnerability. Journal of the American College of Cardiology. 2004;44(9):1918-1926
  69. 69. Kim S, Iwao H. Molecular and cellular mechanisms of angiotensin II-mediated cardiovascular and renal diseases. Pharmacological Reviews. 2000;52(1):11-34
  70. 70. Tsai C-T, Chiang F-T, Tseng C-D, Hwang J-J, Kuo K-T, Wu C-K, et al. Increased expression of mineralocorticoid receptor in human atrial fibrillation and a cellular model of atrial fibrillation. Journal of the American College of Cardiology. 2010;55(8):758-770
  71. 71. Burstein B, Comtois P, Michael G, Nishida K, Villeneuve L, Yeh Y-H, et al. Changes in connexin expression and the atrial fibrillation substrate in congestive heart failure. Circulation Research. 2009;105(12):1213-1222
  72. 72. Zlochiver S, Muñoz V, Vikstrom KL, Taffet SM, Berenfeld O, Jalife J. Electrotonic myofibroblast-to-myocyte coupling increases propensity to reentrant arrhythmias in two-dimensional cardiac monolayers. Biophysical Journal. 2008;95(9):4469-4480
  73. 73. Zou R, Kneller J, Leon LJ, Nattel S. Substrate size as a determinant of fibrillatory activity maintenance in a mathematical model of canine atrium. American Journal of Physiology-Heart and Circulatory Physiology. 2005;289(3):H1002-H1012
  74. 74. Sharifov OF, Fedorov VV, Beloshapko GG, Glukhov AV, Yushmanova AV, Rosenshtraukh LV. Roles of adrenergic and cholinergic stimulation in spontaneous atrial fibrillation in dogs. Journal of the American College of Cardiology. 2004;43(3):483-490
  75. 75. Chang C-M, Wu T-J, Zhou S, Doshi RN, Lee M-H, Ohara T, et al. Nerve sprouting and sympathetic hyperinnervation in a canine model of atrial fibrillation produced by prolonged right atrial pacing. Circulation. 2001;103(1):22-25
  76. 76. Shen MJ, Chang H-C, Park H-W, George Akingba A, Chang P-C, Zhang Z, et al. Low-level vagus nerve stimulation upregulates small conductance calcium-activated potassium channels in the stellate ganglion. Heart Rhythm. 2013;10(6):910-915
  77. 77. Yu L, Scherlag BJ, Li S, Sheng X, Lu Z, Nakagawa H, et al. Low-level vagosympathetic nerve stimulation inhibits atrial fibrillation inducibility: Direct evidence by neural recordings from intrinsic cardiac ganglia. Journal of Cardiovascular Electrophysiology. 2011;22(4):455-463
  78. 78. Lubitz SA, Ozcan C, Magnani JW, Kääb S, Benjamin EJ, Ellinor PT. Genetics of atrial fibrillation: Implications for future research directions and personalized medicine. Circulation. Arrhythmia and Electrophysiology. 2010;3(3):291-299
  79. 79. Xu C, Zhang R, Xia Y, Xiong L, Yang W, Wang P. Annotation of susceptibility SNPs associated with atrial fibrillation. Aging. 2020;12(17):16981-16998
  80. 80. Feghaly J, Zakka P, London B, MacRae CA, Refaat MM. Genetics of atrial fibrillation. Journal of the American Heart Association. 2018;7(20):e009884. DOI: 10.1161/JAHA.118.009884
  81. 81. Chen S, Zhang L, Bryant R, Vincent G, Flippin M, Lee J, et al. kcnq1 mutations in patients with a family history of lethal cardiac arrhythmias and sudden death: Genotype-phenotype correlations in patients with kcnq1 mutations. Clinical Genetics. 2003;63(4):273-282
  82. 82. Sinner MF, Pfeufer A, Akyol M, Beckmann B-M, Hinterseer M, Wacker A, et al. The non-synonymous coding IKr-channel variant KCNH2-K897T is associated with atrial fibrillation: Results from a systematic candidate gene-based analysis of KCNH2 (Herg). European Heart Journal. 2008;29(7):907-914
  83. 83. Olesen MS, Refsgaard L, Holst AG, Larsen AP, Grubb S, Haunsø S, et al. A novel KCND3 gain-of-function mutation associated with early-onset of persistent lone atrial fibrillation. Cardiovascular Research. 2013;98(3):488-495
  84. 84. Xia M, Jin Q, Bendahhou S, He Y, Larroque M-M, Chen Y, et al. A Kir2.1 gain-of-function mutation underlies familial atrial fibrillation. Biochemical and Biophysical Research Communications. 2005;332(4):1012-1019
  85. 85. Olson TM, Alekseev AE, Liu XK, Park S, Zingman LV, Bienengraeber M, et al. Kv1.5 channelopathy due to KCNA5 loss-of-function mutation causes human atrial fibrillation. Human Molecular Genetics. 2006;15(14):2185-2191
  86. 86. Chang S-H, Chang S-N, Hwang J-J, Chiang F-T, Tseng C-D, Lee J-K, et al. Significant association of rs13376333 in kcnn3 on chromosome 1q21 with atrial fibrillation in a Taiwanese population. Circulation Journal. 2012;76(1):184-188
  87. 87. Luo Z, Yan C, Zhang W, Shen X, Zheng W, Chen F, et al. Association between SNP rs13376333 and rs1131820 in the KCNN3 gene and atrial fibrillation in the Chinese Han population. Clinical Chemistry and Laboratory Medicine. 2014;52(12):1867-1873
  88. 88. Li C, Wang F, Yang Y, Fu F, Xu C, Shi L, et al. Significant association of SNP rs2106261 in the ZFHX3 gene with atrial fibrillation in a Chinese Han GeneID population. Human Genetics. 2011;129(3):239-246
  89. 89. Olesen MS, Jabbari J, Holst AG, Nielsen JB, Steinbrüchel DA, Jespersen T, et al. Screening of KCNN3 in patients with early-onset lone atrial fibrillation. EP Europace. 2011;13(7):963-967
  90. 90. Yu C-C, Chia-Ti T, Chen P-L, Wu C-K, Chiu F-C, Chiang F-T, et al. KCNN2 polymorphisms and cardiac tachyarrhythmias. Medicine. 2016;95(29):e4312
  91. 91. Bentzen BH, Bomholtz SH, Simó-Vicens R, Folkersen L, Abildgaard L, Speerschneider T, et al. Mechanisms of action of the kca2-negative modulator ap30663, a novel compound in development for treatment of atrial fibrillation in man. Frontiers in Pharmacology. 2020;11:610
  92. 92. Tsai W-C, Chan Y-H, Hsueh C-H, Everett TH, Chang P-C, Choi E-K, et al. Small conductance calcium-activated potassium current and the mechanism of atrial arrhythmia in mice with dysfunctional melanocyte-like cells. Heart Rhythm. 2016;13(7):1527-1535
  93. 93. Rahm A-K, Gramlich D, Wieder T, Müller ME, Schoeffel A, El Tahry FA, et al. Trigger-specific remodeling of kca2 potassium channels in models of atrial fibrillation. Pharmacogenomics and Personalized Medicine. 2021;14:579-590
  94. 94. Li N, Timofeyev V, Tuteja D, Xu D, Lu L, Zhang Q, et al. Ablation of a Ca 2+ −activated K + channel (SK2 channel) results in action potential prolongation in atrial myocytes and atrial fibrillation: Ablation of SK2 channel results in action potential prolongation in atrial myocytes. The Journal of Physiology. 2009;587(5):1087-1100
  95. 95. Rahm A-K, Wieder T, Gramlich D, Müller ME, Wunsch MN, El Tahry FA, et al. HDAC2-dependent remodeling of KCa2.2 (Kcnn2) and KCa2.3 (Kcnn3) K+ channels in atrial fibrillation with concomitant heart failure. Life Sciences. 2021;266:118892
  96. 96. Bauer A, Koch M, Kraft P, Becker R, Kelemen K, Voss F, et al. The new selective IKs? Blocking agent HMR 1556 restores sinus rhythm and prevents heart failure in pigs with persistent atrial fibrillation. Basic Research in Cardiology. 2005;100(3):270-278
  97. 97. Haugaard MM, Hesselkilde EZ, Pehrson S, Carstensen H, Flethøj M, Præstegaard KF, et al. Pharmacologic inhibition of small-conductance calcium-activated potassium (Sk) channels by NS8593 reveals atrial antiarrhythmic potential in horses. Heart Rhythm. 2015;12(4):825-835
  98. 98. Burashnikov A, Barajas-Martinez H, Hu D, Robinson VM, Grunnet M, Antzelevitch C. The small conductance calcium-activated potassium channel inhibitors ns8593 and ucl1684 prevent the development of atrial fibrillation through atrial-selective inhibition of sodium channel activity. Journal of Cardiovascular Pharmacology. 2020;76(2):164-172
  99. 99. Fenner MF, Gatta G, Sattler S, Kuiper M, Hesselkilde EM, Adler DMT, et al. Inhibition of small-conductance calcium-activated potassium current (Ik, ca) leads to differential atrial electrophysiological effects in a horse model of persistent atrial fibrillation. Frontiers in Physiology. 2021;12:614483
  100. 100. Diness JG, Abildgaard L, Bomholtz SH, Skarsfeldt MA, Edvardsson N, Sørensen US, et al. Inhibition of kca2 channels decreased the risk of ventricular arrhythmia in the Guinea pig heart during induced hypokalemia. Frontiers in Pharmacology. 2020;11:749
  101. 101. Saljic A, Soattin L, Trachsel DS, Boddum K, Jespersen T. In vivo knockdown of SK3 channels using antisense oligonucleotides protects against atrial fibrillation in rats. Journal of Molecular and Cellular Cardiology. 2020;147:18-26
  102. 102. Gatta G, Sobota V, Citerni C, Diness JG, Sørensen US, Jespersen T, et al. Effective termination of atrial fibrillation by SK channel inhibition is associated with a sudden organization of fibrillatory conduction. EP Europace. 2021;23(11):1847-1859
  103. 103. Yan Y, Skarsfeldt MA, Diness JG, Bentzen BH. Small conductance calcium activated K+ channel inhibitor decreases stretch induced vulnerability to atrial fibrillation. IJC Heart & Vasculature. 2021;37:100898
  104. 104. Celotto C, Sanchez C, Mountris KA, Laguna P, Pueyo E. Sk channel block and adrenergic stimulation counteract acetylcholine-induced arrhythmogenic effects in human atria. Annual International Conference of the IEEE Engineering in Medicine and Biology Society. 2020;2020:2303-2306
  105. 105. Kirchhoff JE, Goldin Diness J, Sheykhzade M, Grunnet M, Jespersen T. Synergistic antiarrhythmic effect of combining inhibition of Ca2+-activated K+ (Sk) channels and voltage-gated Na+ channels in an isolated heart model of atrial fibrillation. Heart Rhythm. 2015;12(2):409-418
  106. 106. Kirchhoff JE, Diness JG, Abildgaard L, Sheykhzade M, Grunnet M, Jespersen T. Antiarrhythmic effect of the Ca2+-activated K+ (Sk) channel inhibitor ICA combined with either amiodarone or dofetilide in an isolated heart model of atrial fibrillation. Pflügers Archiv—European Journal of Physiology. 2016;468(11–12):1853-1863
  107. 107. Bradfield JS, Boyle NG, Shivkumar K. Ventricular arrhythmias. In: Fuster V, Harrington RA, Narula J, Eapen ZJ, editors. Hurst’s The Heart [Internet]. 14th ed. New York, NY: McGraw-Hill Education; 2017. Available from: accessmedicine.mhmedical.com/content.aspx?aid=1161718616
  108. 108. Hynes BJ, Luck JC, Wolbrette DL, Boehmer J, Naccarelli GV. Arrhythmias in patients with heart failure. Current Treatment Options in Cardiovascular Medicine. 2002;4(6):467-485
  109. 109. Chen M, Xu D-Z, Wu AZ, Guo S, Wan J, Yin D, et al. Concomitant SK current activation and sodium current inhibition cause J wave syndrome. JCI Insight. 2018;3(22):e122329
  110. 110. Gui L, Bao Z, Jia Y, Qin X, (Jack) Cheng Z, Zhu J, et al. Ventricular tachyarrhythmias in rats with acute myocardial infarction involves activation of small-conductance Ca2+-activated K + channels. American Journal of Physiology-Heart and Circulatory Physiology. 2013;304(1):H118-H130
  111. 111. Hundahl LA, Sattler SM, Skibsbye L, Diness JG, Tfelt-Hansen J, Jespersen T. Pharmacological blockade of small conductance Ca2+-activated K+ channels by ICA reduces arrhythmic load in rats with acute myocardial infarction. Pflügers Archiv—European Journal of Physiology. 2017;469(5–6):739-750
  112. 112. Lubberding AF, Sattler SM, Grunnet M, Sørensen US, Tfelt-Hansen J, Jespersen T. Arrhythmia development during inhibition of small-conductance calcium-activated potassium channels in acute myocardial infarction in a porcine model. EP Europace. 2019;21(10):1584-1593
  113. 113. Tenma T, Mitsuyama H, Watanabe M, Kakutani N, Otsuka Y, Mizukami K, et al. Small-conductance Ca2+−activated K+ channel activation deteriorates hypoxic ventricular arrhythmias via CaMKII in cardiac hypertrophy. American Journal of Physiology. Heart and Circulatory Physiology. 2018;315(2):H262-H272
  114. 114. Takahashi M, Yokoshiki H, Mitsuyama H, Watanabe M, Temma T, Kamada R, et al. SK channel blockade prevents hypoxia-induced ventricular arrhythmias through inhibition of Ca 2+/voltage uncoupling in hypertrophied hearts. American Journal of Physiology-Heart and Circulatory Physiology. 2021;320(4):H1456-H1469
  115. 115. Landaw J, Zhang Z, Song Z, Liu MB, Olcese R, Chen P-S, et al. Small-conductance Ca2+−activated K+ channels promote J-wave syndrome and phase 2 reentry. Heart Rhythm. 2020;17(9):1582-1590
  116. 116. Chang P-C, Hsieh Y-C, Hsueh C-H, Weiss JN, Lin S-F, Chen P-S. Apamin induces early afterdepolarizations and torsades de pointes ventricular arrhythmia from failing rabbit ventricles exhibiting secondary rises in intracellular calcium. Heart Rhythm. 2013;10(10):1516-1524
  117. 117. Bonilla IM, Long VP, Vargas-Pinto P, Wright P, Belevych A, Lou Q, et al. Calcium-activated potassium current modulates ventricular repolarization in chronic heart failure. Rodriguez B, editor. PLoS ONE. 2014;9(10):e108824
  118. 118. Kim TY, Terentyeva R, Roder KH, Li W, Liu M, Greener I, et al. SK channel enhancers attenuate Ca2+-dependent arrhythmia in hypertrophic hearts by regulating mito-ROS-dependent oxidation and activity of RyR. Cardiovascular Research. 2017;113(3):343–353. DOI: 10.1093/cvr/cvx005
  119. 119. Chan Y-H, Tsai W-C, Ko J-S, Yin D, Chang P-C, Rubart M, et al. Small-conductance calcium-activated potassium current is activated during hypokalemia and masks short-term cardiac memory induced by ventricular pacing. Circulation. 2015;132(15):1377-1386
  120. 120. Wan J, Chen M, Wang Z, Everett TH, Rubart-von der Lohe M, Shen C, et al. Small-conductance calcium-activated potassium current modulates the ventricular escape rhythm in normal rabbit hearts. Heart Rhythm. 2019;16(4):615-623
  121. 121. Ni Y, Wang T, Zhuo X, Song B, Zhang J, Wei F, et al. Bisoprolol reversed small conductance calcium-activated potassium channel (Sk) remodeling in a volume-overload rat model. Molecular and Cellular Biochemistry. 2013;384(1–2):95-103
  122. 122. Kamada R, Yokoshiki H, Mitsuyama H, Watanabe M, Mizukami K, Tenma T, et al. Arrhythmogenic β-adrenergic signaling in cardiac hypertrophy: The role of small-conductance calcium-activated potassium channels via activation of CaMKII. European Journal of Pharmacology. 2019;844:110-117
  123. 123. Fei Y-D, Chen M, Guo S, Ueoka A, Chen Z, Rubart-von der Lohe M, et al. Simultaneous activation of the small conductance calcium-activated potassium current by acetylcholine and inhibition of sodium current by ajmaline cause J-wave syndrome in Langendorff-perfused rabbit ventricles. Heart Rhythm. 2021;18(1):98-108
  124. 124. Stowe DF, Camara AKS. Mitochondrial reactive oxygen species production in excitable cells: Modulators of mitochondrial and cell function. Antioxidants & Redox Signaling. 2009;11(6):1373-1414
  125. 125. Stowe DF, Yang M, Heisner JS, Camara AKS. Endogenous and agonist-induced opening of mitochondrial big versus small Ca2+-sensitive K+ channels on cardiac cell and mitochondrial protection. Journal of Cardiovascular Pharmacology. 2017;70(5):314-328
  126. 126. Kim TY, Terentyeva R, Roder KHF, Li W, Liu M, Greener I, et al. SK channel enhancers attenuate Ca2+−dependent arrhythmia in hypertrophic hearts by regulating Mito-ROS-dependent oxidation and activity of RyR. Cardiovascular Research. 2017;113(3):343-353

Written By

Katherine Zhong, Shawn Kant, Frank Sellke and Jun Feng

Submitted: 06 February 2022 Reviewed: 02 March 2022 Published: 13 April 2022