Skip to main content
Channels logoLink to Channels
. 2015 Aug 20;9(6):352–359. doi: 10.1080/19336950.2015.1076597

Cardiac ion channels

Birgit T Priest 1,*, Jeff S McDermott 1
PMCID: PMC4850049  PMID: 26556552

Abstract

Ion channels are critical for all aspects of cardiac function, including rhythmicity and contractility. Consequently, ion channels are key targets for therapeutics aimed at cardiac pathophysiologies such as atrial fibrillation or angina. At the same time, off-target interactions of drugs with cardiac ion channels can be the cause of unwanted side effects. This manuscript aims to review the physiology and pharmacology of key cardiac ion channels. The intent is to highlight recent developments for therapeutic development, as well as elucidate potential mechanisms for drug-induced cardiac side effects, rather than present an in-depth review of each channel subtype.

Keywords: arrhythmia, cardiomyocyte, cardiac function, electrocardiogram, ion channel, pharmacology, vascular tone

Abbreviations

AF

Atrial Fibrillation

CiPA

Comprehensive in vitro Proarrhythmia Assay

VGCC

Voltage-gated calcium channel.

Introduction

A number of ion channels are expressed in the heart and in the vasculature and are the target for important therapeutics as well as for off-target cardiac side effects of drugs in development. In the following chapters, we aim to briefly discuss the key players, their expression profiles, physiological roles and pharmacology.

Sodium Channels

Voltage-gated sodium channels typically exist as a protein complex consisting of a large α subunit (∼260 kDa) with one or two β subunits. To date, 9 α subunits, Nav1.1–1.9, have been identified. The accessory β subunits (β1−β4) are not required to form functional channels but may affect trafficking and/or biophysical characteristics of the channel.1 The main sodium channel subtypes found in the heart are the tetrodotoxin (TTX) insensitive subtypes Nav1.5 and Nav1.8. Expression of TTX-sensitive sodium channels has been described,2 but the lack of cardiovascular effects of TTX in animals and human victims of accidental TTX poisoning suggests that these channels contribute little to normal cardiac function.3 Expression of Nav1 channels in smooth muscle cells of the vasculature has been reported,4 and the sodium channel activator veratridine induces contraction of several types of rodent blood vessels. The contractile effects of veratridine are partly mediated through actions on sympathetic nerves, but may also involve a direct effect on smooth muscle myocytes.5 However, the lack of effect of sodium channel blockers suggests that they do not contribute to vascular tone under physiological conditions.

Nav1.5

Nav1.5, encoded by the SCN5A gene, is known as the cardiac sodium channel. However, it is expressed to some degree in other excitable and nonexcitable tissues.6 With the exception of the sinoatrial (SA) node and atrioventricular (AV) node,7 Nav1.5 activation is responsible for action potential upstroke throughout the myocardium. This initial influx of Na+ provides the depolarization trigger for voltage-gated calcium channel activation, subsequent calcium dependent calcium release from the sarcoplasmic reticulum, and finally, contraction of the sarcomeres. In addition to their role in contraction, sodium channels are also the key driver of cardiac conduction. In the ventricles, the summation of the individual action potential upstrokes forms the ventricular depolarization wave responsible for the QRS complex in the electrocardiogram (Fig. 1). The atrial depolarization wave, reflected in the PR interval, while also sodium channel dependent, has a calcium channel dependent component as well. Conduction between adjacent cardiomyocytes involves gap junctions at the intercalated disc regions that express connexin proteins and high local concentrations of Nav1.5, and it is thought that local electrical fields may contribute to sodium channel activation in adjacent myocytes independent of gap junctional communication.8

Figure 1.

Figure 1.

The top panel shows the ventricular action potential and the currents contributing to each phase. The bottom panel depicts the features and intervals seen in a typical electrocardiogram (e.g., the PR interval is the time between the P wave and the R wave).

Human mutations in SCN5A have been linked to multiple perturbations in cardiac function: loss-of-function mutations are the cause of approximately 20% of Brugada syndrome cases9; whereas gain-of-function mutations cause long QT Syndrome type 3.10 SCN5A constitutive KO mice are embryonic lethal, with the SCN5A+/− heterozygotes displaying some of the conduction-related deficiencies seen in the human mutant population.11

Nav1.5 is the target of many common antiarrhythmic therapies. Based on the Vaughan-Williams schema, sodium channel blockers are grouped into Class I based on their propensity to decrease the upstroke velocity (Vmax or dV/dt) of ventricular cardiac action potentials. Sodium channel blockers are further subdivided based on their effects on the QRS interval and the effective refractory period (ERP). While all Class I antiarrhythmics have higher channel affinity at depolarized membrane potentials, their on- and off rates vary significantly. Compounds that prolong ERP with little effect on QRS (Class Ib) show fast on- and off-rates, whereas Class Ic compounds, which prolong QRS without major effects on ERP, were found to have slow kinetics.12 Class Ic drugs carry an increased risk of cardiac arrest, and consequently Nav1.5 inhibition and QRS prolongation are frequently studied endpoints in preclinical cardiac toxicology assays. The degree of Nav1.5 inhibition necessary to result in significant QRS prolongation is the subject of significant debate.13

Nav1.8

Like Nav1.5, Nav1.8 is a member of the TTX-resistant sodium channel family. As recently as 2010, Nav1.8 was thought to function exclusively in the peripheral nervous system. However, recently several genome-wide association studies have linked polymorphisms in SCN10A, the gene encoding Nav1.8, to prolongation of PR and QRS intervals.14,15 Subsequently, Nav1.8 expression was demonstrated in mouse and human cardiac myocytes and intracardiac neurons.16 In cardiac myocytes, block of Nav1.8 reduces late Na+ current and shortens action potential duration,17 while block of Nav1.8 in intracardiac neurons reduces action potential frequency.18 How these observations relate to cardiac conduction is not entirely clear. The picture is further complicated by the findings that SCN10A knock-out mice display a decreased PR interval, while pharmacological inhibition of Nav1.8 causes PR prolongation.19 The human polymorphisms occur in non-coding regions of SCN10A and their functional consequences are not known. Recent data suggest that at least one of these polymorphisms may affect transcriptional regulation of SCN5A and/or SCN10A.20

Calcium Channels

Voltage-gated calcium channels (VGCCs) are critical for all aspects of cardiovascular physiology. In mammalian species, 10 genes have been cloned that encode pore forming subunits of VGCCs.21 In native tissues, VGCCs exist as a complex of the pore-forming α-subunit, one of 4 distinct β-subunits, one of 4 α2δ-subunits and potentially one of 8 γ-subunits. The β-subunits and α2δ-subunits have profound effects on surface membrane expression and voltage dependence of gating, whereas much less is known about the role of the γ-subunits. Depending on cell type and subcellular location, VGCCs further co-assemble with a number of proteins, including proteins such as calmodulin and calcium-binding protein1 that regulate their activity, proteins that are involved in vesicle docking and neurotransmitter release and intracellular signaling molecules.22,23

L-type (Cav1.x) channels

Physiology

In adult cardiac myocytes, calcium influx through Cav1.2 is responsible for the majority of the inward current during the plateau phase of the cardiac action potential, and Cav1.2 is the dominant channel involved in excitation-contraction coupling. Calcium currents also contribute to the electrical properties of cardiomyocytes, and channel mutations are associated with different cardiac arrhythmias. In ventricular myocytes, Cav1.2 is the only VGCC, whereas both Cav1.2 and Cav1.3 are expressed in atrial myocytes, in the autonomously beating cells of the SA and AV node and in vascular smooth muscle cells. Currently, alternative splicing of VGCCs has attracted attention as a means of tissue specificity, and the dominant Cav1.2 variant was found to differ between smooth muscle and cardiac cells.24

Gain-of-function mutations in CACNA1C, the gene encoding Cav1.2, are associated with long-QT syndrome,25 as well as with the multisystem disorder known as Timothy syndrome. Timothy syndrome mutations are located in 2 mutually exclusive and tissue-specific exons. Mutations in both exons are associated with increased risk for ventricular arrhythmias and sudden cardiac death, a phenotype that is more pronounced if the mutation is found in the cardiac preferred exon.26. Loss-of-function mutations in CACNA1C, as well as mutations in the genes encoding the associated subunits β2 and α2δ, account for 12–13% of patients with Brugada syndrome, characterized by an elevated ST segment and increased risk for ventricular fibrillation and sudden cardiac death.27 In mice, global and cardiac-specific deletion of Cav1.2 is lethal, whereas smooth muscle specific knockout lowered blood pressure and reduced myogenic tone and contractility in isolated small-diameter arteries.28

A role for Cav1.3 in modulating heart rate has only recently been appreciated. Relative to other L-type calcium channels, Cav1.3 activates rapidly and at more hyperpolarized voltages; key attributes that support the role of Cav1.3 in pace making.

The first human channelopathy involving CACNA1D, the gene encoding Cav1.3, was identified in 2011.29 The loss-of-function mutation in an alternatively spliced exon is associated with congenital deafness. On further examination, affected individuals presented with bradycardia and impaired SA node function but normal QRS and QT intervals. In mice, global knock-out of Cav1.3 results in viable, fertile offspring presenting with deafness, bradycardia and SA and AV node dysfunction, much like their human counterparts. Recent evidence suggests that the C-terminus of Cav1.3 may function as a transcriptional regulator in atrial myocytes and modulate the expression of myosin light chain II and small conductance calcium-activated K+ channel.30

In addition to the classical role in conducting calcium currents, recent work highlights a metabotropic role of L-type calcium channels as signaling receptors activated by extracellular calcium binding.31 These findings can account for the rapid kinetics of excitation-contraction and excitation-secretion coupling; however sustained or repetitive contraction and neurotransmitter release requires calcium influx through VGCCs.

Pharmacology

L-type calcium channel blockers are used clinically to treat hypertension, angina, and/or atrial dysrhythmias. They are generally well tolerated, despite the wide-spread expression of Cav1 channels. Factors that affect the tolerability of L-type calcium channel blockers include the degree of voltage dependence and kinetics of channel block.

Based on their structures, calcium channel blockers can be classified as dihydropyridines (e.g., nifedipine, amlodipine), benzothiazepines (e.g., diltiazem) and phenylalkylamines (e.g., verapamil). The three classes of ligands bind to distinct sites located near the outer pore of the α-subunit. All L-type calcium channel blockers lower systolic and diastolic blood pressure by reducing peripheral vascular tone and improve blood supply to the heart by dilating coronary arteries.32 However, dihydropyridine (DHP) and non-dihydropyridine (non-DHP) calcium channel blockers differ in their effect on cardiac contractility and heart rate (Table 1). Non-DHP calcium channel blockers decrease contractility and heart rate; factors that decrease cardiac workload and are beneficial in patients suffering from angina, but are contra-indicated in patients with heart failure. In contrast, DHPs show functional selectivity for L-type channels in the vasculature, based in part on their more pronounced voltage dependence combined with the more depolarized resting membrane potential in the vasculature.33 DHPs typically do not affect contractility and differ in their effect on heart rate. Especially short-acting DHPs, associated with rapid fluctuation in blood pressure, can increase sympathetic activity, resulting in reflex tachycardia. This effect is largely mitigated in long-acting drugs that reach their steady-state plasma levels more slowly. Differing selectivity for Cav1.2 over Cav1.3, alternative splicing and different β-subunits may also contribute to the selectivity for cardiac versus vascular targets in vivo.33 Increased activity on Cav1.3 may be associated with less tachycardia, considering the role of Cav1.3 in heart rate modulation. DHP structure-activity relationships have been reviewed recently.34

Table 1.

Pharmacology of L-type Calcium Channels

  DHPs short-acting DHPs long-acting Non-DHP L-type blocker L-/N-type blocker L-/T-type blocker
Blood pressure
Vasodilation
Heart Rate
Myocardial Contractility
Sympathetic activity

Weak inhibition of calcium channels, with IC50s in the range of 5–50 μM, is found in many Medicinal Chemistry compounds. Most of these small molecules act in a voltage dependent manner and rarely result in adverse events. A decrease in systolic and diastolic blood pressure without effects in the electrocardiogram is suggestive of L-type calcium channel block. In the clinic, calcium channel blockers afford blood pressure lowering at free plasma concentrations close to the IC50.

N-type (Cav2.2) and T-type (Cav3.x) channels

The autonomic nervous system plays a key role in the control of heart rate. A number of voltage gated calcium channels are found in neurons of the autonomic nervous system. Of particular importance is the N-type calcium channel Cav2.2, because of its role in the release of norepinephrine from synaptic terminals in the sympathetic nervous system.35 Cav2.2 also plays a dominant role controlling the exercise pressor reflex through its expression in muscle afferents.36

T-type channels differ from other VGCCs by their lack of accessory subunits and by their activation at more hyperpolarized voltages. T-type channels are expressed widely and play a key role in modulating excitability and burst firing in many cell types. In the cardiovascular system, expression of Cav3.1 and Cav3.2 has been found in the SA and AV nodes and in arterial smooth muscle cells, where they play a role in pace-making and control of vascular tone, respectively.37,38 Interestingly, the endogenous vasodilator nitric oxide suppresses T-type calcium currents in aterial smooth mucle cells.39

No selective N-type or T-type small molecule blockers have been approved for clinical use. Recent developments suggest that mixed L-/N-type blockers (e.g., cilnidipine) and mixed L-/T-type (e.g. efonidipine, benidipine) may offer benefits over selective L-type blockers as anti-hypertensives.40 Mixed blockers are expected to lower blood pressure, while avoiding the reflex tachycardia caused by sympathetic activation. Indeed, cilnidipine has demonstrated efficacy in patients with morning hypertension, caused by increased sympathetic tone in the morning and associated with an elevated risk for heart attack and stroke in the early morning hours.41 Mixed L-/N-type should also avoid the exercise pressor reflex that can lead to an increased risk for heart attacks.

Future efforts in the development of calcium channel blockers are expected to focus on highly selective compounds, targeting a particular subtype or variant, and on mixed blockers demonstrating appropriate phenotypic responses and improved tolerability.

Voltage-Gated Potassium Channels

Several families of voltage-gated potassium channels are expressed in cardiac myocytes and together provide the majority of the outward current responsible for action potential repolarization. The activation and inactivation kinetics of each channel subtype determine their contribution to different phases of repolarization. In order of the repolarization phase they contribute to, the dominant Kv channel pore-forming subunits and corresponding currents are: Kv4.3 and Kv1.4/fast and slow components of the transient outward current (Ito), Kv1.5/ultra-rapid delayed-rectifier current (IKur), Kv11.1 aka hERG/rapid delayed-rectifier current (IKr), and Kv7.1 aka KvLQT1/slow delayed-rectifier current (IKs). In addition to the pore-forming subunit, cardiac myocytes express a variety of accessory proteins, including β-subunits, MinK, MinK-related proteins, potassium channel interacting proteins, and potassium channel accessory proteins.42,43 Inhibition of Kv channels generally leads to a depolarized action potential plateau and/or prolonged action potential duration.

The use of potassium channel blockers to treat arrhythmias has generally been disappointing because of their pro-arrhythmic potential. Especially, the association of reduced Kv11.1/hERG activity, either drug-related or caused by channel mutations, with an increased risk for the arrhythmia known as Torsades de pointes is well established.44 However two strategies have evolved to pharmacologically treat atrial fibrillation (AF), the most common form of arrhythmia. Both approaches involve selectively blocking atrial Kv channels, either by targeting Kv1.5 channels which are expressed in atrial but not ventricular myocytes, or through multichannel inhibitors with functional selectivity for the atrium.

Kv1.5

Kv1.5 is the gene product of KCNA5 and is responsible for the ultra-rapid delayed-rectifier current (IKur). IKur is expressed almost exclusively in atrial myocytes, making it a promising target in AF drug discovery.45 Block of Kv1.5 contributes to the mechanism of action of several agents 46; however no selective Kv1.5 blockers have been tested in patients suffering from AF. One compound, MK-0448, that is selective for Kv1.5 over other cardiac ion channels, showed efficacy in multiple preclinical models of AF and was examined in an invasive electrophysiological study in healthy volunteers.47 In this study, MK-0448 did not prolong the atrial refractory period. Follow-up studies in anesthetized dogs attributed the lack of effect to the high vagal tone in young, healthy subjects such as those participating in the study.47 Although vagal tone may be reduced in the patient population, the greater risk for episodes of AF associated with vagal stimulation does not bode well for this mechanism, and clinical development of MK-0448 was discontinued. However, a recent study with MK-0448 in human atrial tissue from sinus rhythm controls and from permanent AF patients showed that MK-0448 delayed the effective refractory period in atrial tissue from AF patients and not from control subjects. Therefore, the disappointing trial results with MK-0448 in healthy volunteers may not be predictive of eventual efficacy in an AF population.48

Multichannel inhibitors

Several compounds, that are relatively weak blockers of a number of cardiac ion channels, have entered development. Of these, intravenous formulations of vernakalant, AVE0118, and AZD7009 entered clinical development for acute conversion of AF to sinus rhythm (rhythm control). All three compounds inhibit Kv1.5, Kv4.3, Kir3.1/4 and block Nav1.5 at depolarized potentials. In addition, vernakalant blocks hERG, and AZD7009 blocks hERG and KvLQT1. All three compounds are effective in preclinical models of AF and relatively free of pro-arrhythmic risk.49 While development of AVE0118 and AZD7009 has been discontinued, vernakalant (Brinavess™) is approved in many countries for hospital use. Vernakalant is relatively ineffective in patients with long-standing AF, questioning its potential for preventative use.50

Voltage-dependent block of Nav1.5 is more effective in the atria because of their more depolarized membrane potential, especially during periods of tachycardia. It increases the atrial refractory period and protects against the pro-arrhythmic action associated with the block of ventricular Kv channels. Consequently, the combination of blocking potassium channels, especially those expressed in atrial myocyte and absent (Kv1.5, Kir3.1/4) or less prominent (Kv4.3) in ventricular myoctes, and blocking Nav1.5 confers functional atrial selectivity.51

Nature may have evolved its own multichannel modulator in that the sodium channel β1 subunit also associates with Kv4.3. Recently, mutations in β1 were linked to Brugada syndrome and were shown, in recombinant expression systems, to reduce Na+ current and increase Kv4.3 current.52

Inward Rectifying Potassium Channels

In contrast to voltage-gated potassium channels, inward rectifying potassium channels conduct current at hyperpolarized membrane potentials. The so-called weak inward rectifiers allow potassium flow across the entire voltage range; whereas in the strong inward rectifiers, outward potassium currents are largely blocked by voltage-dependent binding of intracellular Mg2+ and polyamines.53,54,55,56 Despite the greater capacity for passing inward currents, under physiological conditions inward rectifying potassium channels conduct outward potassium currents, hyperpolarizing the cell membrane. Seven families of inward rectifying potassium channels have been identified, and subunits of the Kir2, Kir3 and Kir6 families are expressed in the heart.57 The pharmacology of inward rectifying potassium channels is the subject of a recent review.58

Kir2.1

The mammalian Kir2 family has 5 strongly rectifying members (Kir2.1–2.4, Kir2.6). Kir2.1 is the key component of the current IK1 in ventricular myocytes and is a critical regulator of resting membrane potential 59-61; whereas Kir2.3 is the dominant Kir2 subunit in atrial myocytes.62 In the vasculature, Kir2.1 channels are found in smooth muscle cells of small diameter arteries, where they control vascular tone (Fig. 263). Block of Kir2 channels constricts these vessels, and indirect activation of Kir by adenosine causes vasodilation.64

Figure 2.

Figure 2.

The interplay between voltage-dependent calcium and potassium channels in vascular smooth muscle cells controls blood vessel diameter. Reproduced from.63 © Blackwell. Reproduced by permission of John Wiley and Sons. Permission to reuse must be obtained from the rightsholder.

Loss-of-function mutations in KCNJ2, the gene encoding Kir2.1, cause the long QT Syndrome known as Andersen-Tawil syndrome,65 associated with cardiac arrhythmias, periodic paralysis and physical abnormalities. The phospholipid PIP2 activates Kir2.1, and many mutations associated with Andersen-Tawil syndrome show reduced affinity of the channel for PIP2 rather than complete channel absence.66 Gain-of-function mutations have been identified in individuals presenting with short QT syndrome and families suffering from AF.67 Kir2.1 deficient mice die shortly after birth; whereas Kir2.2−/− mice are viable and display no gross abnormalities.68,69

Potent blockers of Kir2 channels are rare. Weak, off-target Kir2 activity has been found in a few drugs, including the breast cancer drug tamoxifen, the anti-histamine diphenhydramine and the anti-malarial agent chloroquine. Chloroquine appears to bind within the pore70; whereas tamoxifen is thought to inhibit Kir2 channels by interfering with PIP2 activation of the channel.71 MicroRNA-26 has been identified as an endogenous modulator of Kir2.1 expression and reduced expression of MicroRNA-26 may promote AF.72

Kir3.1/4 (IKACh)

Four mammalian Kir3 channels (Kir3.1–3.4) have been identified.57 All conduct strongly inward rectifying currents and are distinguished from other Kir families by being activated by Gi/o-coupled GPCRs via Gβγ. Kir3 channels can form homo- and hetero-tetrameric channels. The predominant subtype combinations are Kir3.1/3.2 in the CNS and Kir3.1/3.4 in cardiac tissue.57 The Kir3.1/3.4 heteromultimer gives rise to the current IKACh, named for its activation following acetylcholine binding to muscarinic receptors, and is predominantly expressed in the atria and SA and AV nodes, with little expression in the ventricles.73 In the SA node, IKACh is responsible for vagal control of heart rate.74 In atrial myocytes, IKACh contributes to repolarization and channel activation decreases action potential duration. IKACh inhibition prolongs the atrial refractory period without affecting the ventricles, making it an attractive target for drugs designed to treat AF.75

A loss-of-function mutation in KCNJ5, the gene encoding Kir3.4, has been reported as the cause of the long QT syndrome LQT13 in a single Chinese family. Given the lack of expression in ventricles, this finding is surprising and may in fact result from previously underestimated genetic ‘noise’.76 An association between early-onset AF and 2 common polymorphisms in KCNJ5 has been reported.77 However, since neither polymorphism results in an amino acid change, the implications of this finding are unclear. Mice with genetic ablation of Kir3.1 or Kir3.4 are viable and lack any IKACh current.78,74 Mice from both strains have normal resting heart rates, but significantly diminished vagal control of heart rate. Mice deficient in Kir3.4 were shown to be resistant to experimentally-induced AF.79

Potent inhibitors of Kir3 channels are rare, and the therapeutic potential of selective Kir3 inhibitors remains to be determined. A number of non-selective agents in development for AF, such as vernakalant, NIP-142 and AVE0118, include inhibition of IKACh in their mechanism of action.

Kir6

Only two Kir6 channels have been identified. They are unique in that they exist only in heteromeric complexes with accessory sulfonyl urea receptor (SUR) subunits. The heteromeric complex, formed by 4 Kir6 and 4 SUR subunits, gives rise to a current that is modulated by the intracellular concentrations of ATP and ADP and has been termed KATP.80 KATP channels are weak inwardly rectifying and closed by high intracellular ATP concentrations, found in cardiac myocytes under physiological conditions.81 Thus, channel open probability is typically low and can be regulated by other factors such as the intracellular concentration of MgADP and adenosine receptor activation.82

The sarcolemmal KATP channel is mainly formed from Kir6.2 and SUR2A in ventricular myocytes and from Kir6.2 and SUR1 in atrial myocytes; whereas the predominant vascular smooth muscle KATP channel consists of Kir6.1 and SUR2B subunits.83,80 A KATP current has also been reported in mitochondrial membranes. However, mitochondrial KATP is intact in mice lacking Kir6.1 or Kir6.2 and may represent a macro-molecular complex with Kir1.1 as the pore-forming subunit.84

KATP channels are important players in the cardioprotective mechanism known as ischemic preconditioning.82 The drop in intracellular ATP concentration associated with ischemia, activates KATP currents. Activation of KATP shortens action potential duration and reduces calcium influx. This may prevent cellular toxicity from calcium overload. In addition, activation of KATP causes vasodilation, increasing tissue perfusion and counteracting the ischemic event.85 This mechanism may also be important for the ability to adapt to increased oxygen demand during vigorous exercise.86 The effect of KATP on vascular tone is highlighted by the activation and inhibition of KATP by vasodilators and vasoconstrictors, respectively.64

Kir6.2 is encoded by KCNJ11. Loss-of-function and gain-of-function mutations in KCNJ11 have been identified. Afflicted individuals suffer from abnormalities in glucose handling, due to the expression of KCNJ11 in the pancreas, but no signs of cardiovascular pathophysiology have been reported.87 A gain-of-function variant of Kir6.1, encoded by KCNJ8, has been found in a small number of individuals showing symptoms of either Early Repolarization syndrome or Brugada syndrome; whereas loss-of-function mutations were found in 2 victims of sudden infant death syndrome.86 Recently, mutations in SUR2 have been linked to Cantu syndrome, characterized by multi-organ developmental abnormalities.88 These mutations are considered gain-of-function with regard to KATP, since they result in channels with reduced sensitivity to ATP inhibition. In mice, ablation and gain-of-function of Kir6.2 results mainly in metabolic deficiencies.89 Kir6.1 KO mice die within a few weeks from birth due to coronary artery contraction,89 whereas Kir6.1 gain-of-function animals are viable and have reduced blood pressure.

Unlike other Kir channel families, a rich pharmacology exists for KATP and has been reviewed extensively.90 KATP inhibitors have long been used in the treatment of type 2 diabetes; whereas KATP activators are used clinically to treat refractory hypertension and angina. All known KATP modulators bind to the SUR subunits.

Integrated Cardiac Safety Assessment

A few Drug Discovery efforts target specific ion channel subtypes to treat cardiovascular pathophysiologies; however, off-target interactions with cardiovascular ion channels affect many, if not most, drug development programs. Following the withdrawal of several drugs from the market, based on their risk for inducing the life-threatening ventricular arrhythmia Torsades de pointes, guidelines S7B and E14 were issued that govern the strategy to evaluate the risk for Torsades associated with a new chemical entity.91 These guidelines, issued in 2005, focus on the repolarizing potassium current IKr, the underlying channel hERG, and the heart rate-corrected QT interval. However, not all drugs that block hERG cause QT prolongation, and not all drugs that prolong the QT interval are proarrhythmic, presumably because of the complex interplay between multiple ion channels.92 While current guidelines have been successful at preventing new torsadogenic drugs from reaching the market,93 they may have inadvertently limited patient access to beneficial treatments.91 Efforts are underway to define a new cardiac safety paradigm referred to as CiPA (Comprehensive in vitro Proarrhythmia Assay) that will categorize drug candidates into low-, medium- and high-risk with regard to proarrhythmic potential. This initiative represents a collaborative effort involving multiple pharmaceutical, regulatory and contract research organizations and is a first of its kind. CiPA consists of 3 components: in vitro characterization of electrophysiological effects on 7 cardiac ion channels expressed recombinantly, in silico modeling of the impact on the human ventricular action potential, and characterization of electrophysiological effects on human stem cell-derived cardiomyocytes using multi-electrode arrays and voltage-sensitive dye.94 Expert working groups are tasked with developing and testing detailed protocols for each of the 3 components, which will be validated by testing 29 established drugs spanning the range from low to high proarrhythmic liability.95,96 Once established, CiPA may remove the need for expensive Thorough QT studies for most drug candidates.

Disclosure of Potential Conflicts of Interest

No potential conflicts of interest were disclosed.

References

  • 1.Isom LL. Sodium channel beta subunits: anything but auxiliary. Neuroscientist: Rev J Bringing Neurobiol, Neurol Psychiat 2001; 7:42-54; PMID:; http://dx.doi.org/ 10.1177/107385840100700108 [DOI] [PubMed] [Google Scholar]
  • 2.Maier SK, Westenbroek RE, McCormick KA, Curtis R, Scheuer T, Catterall WA. Distinct subcellular localization of different sodium channel alpha and beta subunits in single ventricular myocytes from mouse heart. Circulation 2004; 109:1421-7; PMID:; http://dx.doi.org/ 10.1161/01.CIR.0000121421.61896.24 [DOI] [PubMed] [Google Scholar]
  • 3.Zimmer T. Effects of tetrodotoxin on the mammalian cardiovascular system. Marine Drugs 2010; 8:741-62; PMID:; http://dx.doi.org/ 10.3390/md8030741 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Saleh S, Yeung SY, Prestwich S, Pucovsky V, Greenwood I. Electrophysiological and molecular identification of voltage-gated sodium channels in murine vascular myocytes. J Physiol 2005; 568:155-69; PMID:; http://dx.doi.org/ 10.1113/jphysiol.2005.090951 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Ho WS, Davis AJ, Chadha PS, Greenwood IA. Effective contractile response to voltage-gated Na+ channels revealed by a channel activator. Am J Physiol Cell Physiol 2013; 304:C739-47; PMID:; http://dx.doi.org/ 10.1152/ajpcell.00164.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Black JA, Waxman SG. Noncanonical roles of voltage-gated sodium channels. Neuron 2013; 80:280-91; PMID:; http://dx.doi.org/ 10.1016/j.neuron.2013.09.012 [DOI] [PubMed] [Google Scholar]
  • 7.Remme CA, Verkerk AO, Hoogaars WM, Aanhaanen WT, Scicluna BP, Annink C, van den Hoff MJ, Wilde AA, van Veen TA, Veldkamp MW, et al.. The cardiac sodium channel displays differential distribution in the conduction system and transmural heterogeneity in the murine ventricular myocardium. Basic Res Cardiol 2009; 104:511-22; PMID:; http://dx.doi.org/ 10.1007/s00395-009-0012-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Veeraraghavan R, Gourdie R, Poelzing S. Mechanisms of cardiac conduction: a history of revisions. Am J Physiol Heart Circ Physiol 2014; 306:H619-27; PMID:; http://dx.doi.org/ 10.1152/ajpheart.00760.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Sarquella-Brugada G, Campuzano O, Arbelo E, Brugada J, Brugada R. Brugada syndrome: clinical and genetic findings. Genet Med: Off J Am College Med Genet 2015; PMID: [DOI] [PubMed] [Google Scholar]
  • 10.Ruan Y, Liu N, Priori SG. Sodium channel mutations and arrhythmias. Nat Rev Cardiol 2009; 6:337-48; PMID:; http://dx.doi.org/ 10.1038/nrcardio.2009.44 [DOI] [PubMed] [Google Scholar]
  • 11.Papadatos GA, Wallerstein PM, Head CE, Ratcliff R, Brady PA, Benndorf K, Saumarez RC, Trezise AE, Huang CL, Vandenberg JI, et al.. Slowed conduction and ventricular tachycardia after targeted disruption of the cardiac sodium channel gene Scn5a. Proc Natl Acad Sci U S A 2002; 99:6210-5; PMID:; http://dx.doi.org/ 10.1073/pnas.082121299 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Vaughan Williams EM. Classifying antiarrhythmic actions: by facts or speculation. J Clin Pharmacol 1992; 32:964-77; PMID:; http://dx.doi.org/ 10.1002/j.1552-4604.1992.tb03797.x [DOI] [PubMed] [Google Scholar]
  • 13.Gintant GA, Gallacher DJ, Pugsley MK. The ‘overly-sensitive’ heart: sodium channel block and QRS interval prolongation. Brit J Pharmacol 2011; 164:254-9; PMID:; http://dx.doi.org/ 10.1111/j.1476-5381.2011.01433.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Chambers JC, Zhao J, Terracciano CM, Bezzina CR, Zhang W, Kaba R, Navaratnarajah M, Lotlikar A, Sehmi JS, Kooner MK, et al.. Genetic variation in SCN10A influences cardiac conduction. Nat Genet 2010; 42:149-52; PMID:; http://dx.doi.org/ 10.1038/ng.516 [DOI] [PubMed] [Google Scholar]
  • 15.Pfeufer A, van Noord C, Marciante KD, Arking DE, Larson MG, Smith AV, Tarasov KV, Muller M, Sotoodehnia N, Sinner MF, et al.. Genome-wide association study of PR interval. Nat Genet 2010; 42:153-9; PMID:; http://dx.doi.org/ 10.1038/ng.517 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Facer P, Punjabi PP, Abrari A, Kaba RA, Severs NJ, Chambers J, Kooner JS, Anand P. Localisation of SCN10A gene product Na(v)1.8 and novel pain-related ion channels in human heart. Int Heart J 2011; 52:146-52; PMID:; http://dx.doi.org/ 10.1536/ihj.52.146 [DOI] [PubMed] [Google Scholar]
  • 17.Yang T, Atack TC, Stroud DM, Zhang W, Hall L, Roden DM. Blocking Scn10a channels in heart reduces late sodium current and is antiarrhythmic. Circ Res 2012; 111:322-32; PMID:; http://dx.doi.org/ 10.1161/CIRCRESAHA.112.265173 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Verkerk AO, Remme CA, Schumacher CA, Scicluna BP, Wolswinkel R, de Jonge B, Bezzina CR, Veldkamp MW. Functional Nav1.8 channels in intracardiac neurons: the link between SCN10A and cardiac electrophysiology. Circ Res 2012; 111:333-43; PMID:; http://dx.doi.org/ 10.1161/CIRCRESAHA.112.274035 [DOI] [PubMed] [Google Scholar]
  • 19.London B. Whither art thou, SCN10A, and what art thou doing? Circ Res 2012; 111:268-70; PMID:; http://dx.doi.org/ 10.1161/CIRCRESAHA.112.275032 [DOI] [PubMed] [Google Scholar]
  • 20.van den Boogaard M, Barnett P, Christoffels VM. From GWAS to function: Genetic variation in sodium channel gene enhancer influences electrical patterning. Trends Cardiovasc Med 2013; 24:99-104; PMID:; http://dx.doi.org/ 10.1016/j.tcm.2013.09.001 [DOI] [PubMed] [Google Scholar]
  • 21.Catterall WA, Perez-Reyes E, Snutch TP, Striessnig J. International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharmacol Rev 2005; 57:411-25; PMID:; http://dx.doi.org/ 10.1124/pr.57.4.5 [DOI] [PubMed] [Google Scholar]
  • 22.Catterall WA. Signaling complexes of voltage-gated sodium and calcium channels. Neurosci Lett 2010; 486:107-16; PMID:; http://dx.doi.org/ 10.1016/j.neulet.2010.08.085 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Muller CS, Haupt A, Bildl W, Schindler J, Knaus HG, Meissner M, Rammner B, Striessnig J, Flockerzi V, Fakler B, et al.. Quantitative proteomics of the Cav2 channel nano-environments in the mammalian brain. Proc Natl Acad Sci U S A 2010; 107:14950-7; PMID:; http://dx.doi.org/ 10.1073/pnas.1005940107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Zhang SS, Shaw RM. Multilayered regulation of cardiac ion channels. Biochimica et Biophysica Acta 2013; 1833:876-85; PMID:; http://dx.doi.org/ 10.1016/j.bbamcr.2012.10.020 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Wemhoner K, Friedrich C, Stallmeyer B, Coffey AJ, Grace A, Zumhagen S, Seebohm G, Ortiz-Bonnin B, Rinne S, Sachse FB, et al.. Gain-of-function mutations in the calcium channel CACNA1C (Cav1.2) cause non-syndromic long-QT but not Timothy syndrome. J Mol Cell Cardiol 2015; 80:186-95; PMID:; http://dx.doi.org/ 10.1016/j.yjmcc.2015.01.002 [DOI] [PubMed] [Google Scholar]
  • 26.Splawski I, Timothy KW, Decher N, Kumar P, Sachse FB, Beggs AH, Sanguinetti MC, Keating MT. Severe arrhythmia disorder caused by cardiac L-type calcium channel mutations. Proc Natl Acad Sci U S A 2005; 102:8089-96; discussion 6-8; PMID:; http://dx.doi.org/ 10.1073/pnas.0502506102 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Hedley PL, Jorgensen P, Schlamowitz S, Moolman-Smook J, Kanters JK, Corfield VA, Christiansen M. The genetic basis of Brugada syndrome: a mutation update. Human Mutat 2009; 30:1256-66; PMID:; http://dx.doi.org/ 10.1002/humu.21066 [DOI] [PubMed] [Google Scholar]
  • 28.Moosmang S, Schulla V, Welling A, Feil R, Feil S, Wegener JW, Hofmann F, Klugbauer N. Dominant role of smooth muscle L-type calcium channel Cav1.2 for blood pressure regulation. EMBO J 2003; 22:6027-34; PMID:; http://dx.doi.org/ 10.1093/emboj/cdg583 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Baig SM, Koschak A, Lieb A, Gebhart M, Dafinger C, Nurnberg G, Ali A, Ahmad I, Sinnegger-Brauns MJ, Brandt N, et al.. Loss of Ca(v)1.3 (CACNA1D) function in a human channelopathy with bradycardia and congenital deafness. Nat Neurosci 2011; 14:77-84; PMID:; http://dx.doi.org/ 10.1038/nn.2694 [DOI] [PubMed] [Google Scholar]
  • 30.Lu L, Sirish P, Zhang Z, Woltz RL, Li N, Timofeyev V, Knowlton AA, Zhang XD, Yamoah EN, Chiamvimonvat N. Regulation of gene transcription by voltage-gated L-type calcium channel, Cav1.3. J Biol Chem 2015; 290:4663-76; PMID:; http://dx.doi.org/ 10.1074/jbc.M114.586883 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Atlas D. Voltage-gated calcium channels function as Ca-activated signaling receptors. Trends Biochem Sci 2014; PMID: [DOI] [PubMed] [Google Scholar]
  • 32.Eisenberg MJ, Brox A, Bestawros AN. Calcium channel blockers: an update. Am J Med 2004; 116:35-43; PMID:; http://dx.doi.org/ 10.1016/j.amjmed.2003.08.027 [DOI] [PubMed] [Google Scholar]
  • 33.Cataldi M, Bruno F. 1,4-dihydropyridines: the multiple personalities of a blockbuster drug family. Translat Med @ UniSa 2012; 4:12-26; PMID: [PMC free article] [PubMed] [Google Scholar]
  • 34.Locatelli A, Cosconati S, Micucci M, Leoni A, Marinelli L, Bedini A, Ioan P, Spampinato SM, Novellino E, Chiarini A, et al.. Ligand based approach to L-type calcium channel by imidazo[2,1-b]thiazole-1,4-dihydropyridines: from heart activity to brain affinity. J Med Chem 2013; 56:3866-77; PMID:; http://dx.doi.org/ 10.1021/jm301839q [DOI] [PubMed] [Google Scholar]
  • 35.Mori Y, Nishida M, Shimizu S, Ishii M, Yoshinaga T, Ino M, Sawada K, Niidome T. Ca(2+) channel alpha(1B) subunit (Ca(V) 2.2) knockout mouse reveals a predominant role of N-type channels in the sympathetic regulation of the circulatory system. Trends Cardiovasc Med 2002; 12:270-5; PMID:; http://dx.doi.org/ 10.1016/S1050-1738(02)00173-1 [DOI] [PubMed] [Google Scholar]
  • 36.Ramachandra R, Hassan B, McGrew SG, Dompor J, Farrag M, Ruiz-Velasco V, Elmslie KS. Identification of CaV channel types expressed in muscle afferent neurons. J Neurophysiol 2013; 110:1535-43; PMID:; http://dx.doi.org/ 10.1152/jn.00069.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Ono K, Iijima T. Pathophysiological significance of T-type Ca2+ channels: properties and functional roles of T-type Ca2+ channels in cardiac pacemaking. J Pharmacol Sci 2005; 99:197-204; PMID:; http://dx.doi.org/ 10.1254/jphs.FMJ05002X2 [DOI] [PubMed] [Google Scholar]
  • 38.Kuo IY, Howitt L, Sandow SL, McFarlane A, Hansen PB, Hill CE. Role of T-type channels in vasomotor function: team player or chameleon? Pflugers Archiv: Euro J Physiol 2014; 466:767-79; PMID:; http://dx.doi.org/ 10.1007/s00424-013-1430-x [DOI] [PubMed] [Google Scholar]
  • 39.Harraz OF, Brett SE, Welsh DG. Nitric oxide suppresses vascular voltage-gated T-type Ca2+ channels through cGMP/PKG signaling. Am J Physiol Heart Circ Physiol 2014; 306:H279-85; PMID:; http://dx.doi.org/ 10.1152/ajpheart.00743.2013 [DOI] [PubMed] [Google Scholar]
  • 40.Ogura C, Ono K, Miyamoto S, Ikai A, Mitani S, Sugimoto N, Tanaka S, Fujita M. L/T-type and L/N-type calcium-channel blockers attenuate cardiac sympathetic nerve activity in patients with hypertension. Blood Pressure 2012; 21:367-71; PMID:; http://dx.doi.org/ 10.3109/08037051.2012.694200 [DOI] [PubMed] [Google Scholar]
  • 41.Kario K, Ando S, Kido H, Nariyama J, Takiuchi S, Yagi T, Shimizu T, Eguchi K, Ohno M, Kinoshita O, et al.. The effects of the L/N-type calcium channel blocker (cilnidipine) on sympathetic hyperactive morning hypertension: results from ACHIEVE-ONE. J Clin Hypertens (Greenwich) 2013; 15:133-42; PMID:; http://dx.doi.org/ 10.1111/jch.12042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Aimond F, Kwak SP, Rhodes KJ, Nerbonne JM. Accessory Kvbeta1 subunits differentially modulate the functional expression of voltage-gated K+ channels in mouse ventricular myocytes. Circ Res 2005; 96:451-8; PMID:; http://dx.doi.org/ 10.1161/01.RES.0000156890.25876.63 [DOI] [PubMed] [Google Scholar]
  • 43.Abbott GW, Xu X, Roepke TK. Impact of ancillary subunits on ventricular repolarization. J Electrocardiol 2007; 40:S42-6; PMID:; http://dx.doi.org/ 10.1016/j.jelectrocard.2007.05.021 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.He FZ, McLeod HL, Zhang W. Current pharmacogenomic studies on hERG potassium channels. Trends Mol Med 2013; 19:227-38; PMID:; http://dx.doi.org/ 10.1016/j.molmed.2012.12.006 [DOI] [PubMed] [Google Scholar]
  • 45.Wang Z, Fermini B, Nattel S. Sustained depolarization-induced outward current in human atrial myocytes. Evidence for a novel delayed rectifier K+ current similar to Kv1.5 cloned channel currents. Circ Res 1993; 73:1061-76; PMID:; http://dx.doi.org/ 10.1161/01.RES.73.6.1061 [DOI] [PubMed] [Google Scholar]
  • 46.Ford JW, Milnes JT. New drugs targeting the cardiac ultra-rapid delayed-rectifier current (I Kur): rationale, pharmacology and evidence for potential therapeutic value. J Cardiovasc Pharmacol 2008; 52:105-20; PMID:; http://dx.doi.org/ 10.1097/FJC.0b013e3181719b0c [DOI] [PubMed] [Google Scholar]
  • 47.Pavri BB, Greenberg HE, Kraft WK, Lazarus N, Lynch JJ, Salata JJ, Bilodeau MT, Regan CP, Stump G, Fan L, et al.. MK-0448, a specific Kv1.5 inhibitor: safety, pharmacokinetics, and pharmacodynamic electrophysiology in experimental animal models and humans. Circ Arrhythmia Electrophysiol 2012; 5:1193-201; PMID:; http://dx.doi.org/ 10.1161/CIRCEP.111.969782 [DOI] [PubMed] [Google Scholar]
  • 48.Loose S, Mueller J, Wettwer E, Knaut M, Ford J, Milnes J, Ravens U. Effects of IKur blocker MK-0448 on human right atrial action potentials from patients in sinus rhythm and in permanent atrial fibrillation. Front Pharmacol 2014; 5:26; PMID:; http://dx.doi.org/ 10.3389/fphar.2014.00026 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Naccarelli GV, Wolbrette DL, Samii S, Banchs JE, Penny-Peterson E, Gonzalez MD. New antiarrhythmic treatment of atrial fibrillation. Exp Rev Cardiovasc Ther 2007; 5:707-14; PMID:; http://dx.doi.org/ 10.1586/14779072.5.4.707 [DOI] [PubMed] [Google Scholar]
  • 50.Tsuji Y, Dobrev D. Safety and efficacy of vernakalant for acute cardioversion of atrial fibrillation: an update. Vasc Health Risk Manage 2013; 9:165-75; PMID: [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Ravens U, Poulet C, Wettwer E, Knaut M. Atrial selectivity of antiarrhythmic drugs. J Physiol 2013; 591:4087-97; PMID:; http://dx.doi.org/ 10.1113/jphysiol.2013.256115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Hu D, Barajas-Martinez H, Medeiros-Domingo A, Crotti L, Veltmann C, Schimpf R, Urrutia J, Alday A, Casis O, Pfeiffer R, et al.. A novel rare variant in SCN1Bb linked to Brugada syndrome and SIDS by combined modulation of Na(v)1.5 and K(v)4.3 channel currents. Heart Rhythm: Off J Heart Rhythm Soc 2012; 9:760-9; PMID:; http://dx.doi.org/ 10.1016/j.hrthm.2011.12.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Vandenberg CA. Inward rectification of a potassium channel in cardiac ventricular cells depends on internal magnesium ions. Proc Natl Acad Sci U S A 1987; 84:2560-4; PMID:; http://dx.doi.org/ 10.1073/pnas.84.8.2560 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Matsuda H. Effects of external and internal K+ ions on magnesium block of inwardly rectifying K+ channels in guinea-pig heart cells. J Physiol 1991; 435:83-99; PMID:; http://dx.doi.org/ 10.1113/jphysiol.1991.sp018499 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Lopatin AN, Makhina EN, Nichols CG. Potassium channel block by cytoplasmic polyamines as the mechanism of intrinsic rectification. Nature 1994; 372:366-9; PMID:; http://dx.doi.org/ 10.1038/372366a0 [DOI] [PubMed] [Google Scholar]
  • 56.Shyng SL, Sha Q, Ferrigni T, Lopatin AN, Nichols CG. Depletion of intracellular polyamines relieves inward rectification of potassium channels. Proc Natl Acad Sci U S A 1996; 93:12014-9; PMID:; http://dx.doi.org/ 10.1073/pnas.93.21.12014 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Kubo Y, Adelman JP, Clapham DE, Jan LY, Karschin A, Kurachi Y, Lazdunski M, Nichols CG, Seino S, Vandenberg CA. International Union of Pharmacology. LIV. Nomenclature and molecular relationships of inwardly rectifying potassium channels. Pharmacol Rev 2005; 57:509-26; PMID:; http://dx.doi.org/ 10.1124/pr.57.4.11 [DOI] [PubMed] [Google Scholar]
  • 58.Bhave G, Lonergan D, Chauder BA, Denton JS. Small-molecule modulators of inward rectifier K+ channels: recent advances and future possibilities. Future Med Chem 2010; 2:757-74; PMID:; http://dx.doi.org/ 10.4155/fmc.10.179 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Lange PS, Er F, Gassanov N, Hoppe UC. Andersen mutations of KCNJ2 suppress the native inward rectifier current IK1 in a dominant-negative fashion. Cardiovasc Res 2003; 59:321-7; PMID:; http://dx.doi.org/ 10.1016/S0008-6363(03)00434-6 [DOI] [PubMed] [Google Scholar]
  • 60.Zobel C, Cho HC, Nguyen TT, Pekhletski R, Diaz RJ, Wilson GJ, Backx PH. Molecular dissection of the inward rectifier potassium current (IK1) in rabbit cardiomyocytes: evidence for heteromeric co-assembly of Kir2.1 and Kir2.2. J Physiol 2003; 550:365-72; PMID:; http://dx.doi.org/ 10.1113/jphysiol.2002.036400 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Tourneur Y. Action potential-like responses due to the inward rectifying potassium channel. J Membrane Biol 1986; 90:115-22; PMID:; http://dx.doi.org/ 10.1007/BF01869929 [DOI] [PubMed] [Google Scholar]
  • 62.Melnyk P, Zhang L, Shrier A, Nattel S. Differential distribution of Kir2.1 and Kir2.3 subunits in canine atrium and ventricle. Am J Physiol Heart Circul Physiol 2002; 283:H1123-33; PMID:; http://dx.doi.org/ 10.1152/ajpheart.00934.2001 [DOI] [PubMed] [Google Scholar]
  • 63.Jackson WF. Potassium channels and regulation of the microcirculation. Microcirculation 1998; 5:85-90; PMID:; http://dx.doi.org/ 10.1111/j.1549-8719.1998.tb00057.x [DOI] [PubMed] [Google Scholar]
  • 64.Park WS, Han J, Earm YE. Physiological role of inward rectifier K(+) channels in vascular smooth muscle cells. Pflugers Archiv: Euro J Physiol 2008; 457:137-47; PMID:; http://dx.doi.org/ 10.1007/s00424-008-0512-7 [DOI] [PubMed] [Google Scholar]
  • 65.Donaldson MR, Yoon G, Fu YH, Ptacek LJ. Andersen-Tawil syndrome: a model of clinical variability, pleiotropy, and genetic heterogeneity. Ann Med 2004; 36 Suppl 1:92-7; PMID:; http://dx.doi.org/ 10.1080/17431380410032490 [DOI] [PubMed] [Google Scholar]
  • 66.Donaldson MR, Jensen JL, Tristani-Firouzi M, Tawil R, Bendahhou S, Suarez WA, Cobo AM, Poza JJ, Behr E, Wagstaff J, et al.. PIP2 binding residues of Kir2.1 are common targets of mutations causing Andersen syndrome. Neurology 2003; 60:1811-6; PMID:; http://dx.doi.org/ 10.1212/01.WNL.0000072261.14060.47 [DOI] [PubMed] [Google Scholar]
  • 67.Deo M, Ruan Y, Pandit SV, Shah K, Berenfeld O, Blaufox A, Cerrone M, Noujaim SF, Denegri M, Jalife J, et al.. KCNJ2 mutation in short QT syndrome 3 results in atrial fibrillation and ventricular proarrhythmia. Proc Natl Acad Sci U S A 2013; 110:4291-6; PMID:; http://dx.doi.org/ 10.1073/pnas.1218154110 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Zaritsky JJ, Eckman DM, Wellman GC, Nelson MT, Schwarz TL. Targeted disruption of Kir2.1 and Kir2.2 genes reveals the essential role of the inwardly rectifying K(+) current in K(+)-mediated vasodilation. Circ Res 2000; 87:160-6; PMID:; http://dx.doi.org/ 10.1161/01.RES.87.2.160 [DOI] [PubMed] [Google Scholar]
  • 69.Zaritsky JJ, Redell JB, Tempel BL, Schwarz TL. The consequences of disrupting cardiac inwardly rectifying K(+) current (I(K1)) as revealed by the targeted deletion of the murine Kir2.1 and Kir2.2 genes. J Physiol 2001; 533:697-710; PMID:; http://dx.doi.org/ 10.1111/j.1469-7793.2001.t01-1-00697.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Noujaim SF, Stuckey JA, Ponce-Balbuena D, Ferrer-Villada T, Lopez-Izquierdo A, Pandit SV, Sanchez-Chapula JA, Jalife J. Structural bases for the different anti-fibrillatory effects of chloroquine and quinidine. Cardiovasc Res 2011; 89:862-9; PMID:; http://dx.doi.org/ 10.1093/cvr/cvr008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Ponce-Balbuena D, Lopez-Izquierdo A, Ferrer T, Rodriguez-Menchaca AA, Arechiga-Figueroa IA, Sanchez-Chapula JA. Tamoxifen inhibits inward rectifier K+ 2.x family of inward rectifier channels by interfering with phosphatidylinositol 4,5-bisphosphate-channel interactions. J Pharmacol Exp Ther 2009; 331:563-73; PMID:; http://dx.doi.org/ 10.1124/jpet.109.156075 [DOI] [PubMed] [Google Scholar]
  • 72.Luo X, Pan Z, Shan H, Xiao J, Sun X, Wang N, Lin H, Xiao L, Maguy A, Qi XY, et al.. MicroRNA-26 governs profibrillatory inward-rectifier potassium current changes in atrial fibrillation. J Clin Investigat 2013; 123:1939-51; PMID:; http://dx.doi.org/ 10.1172/JCI62185 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Krapivinsky G, Gordon EA, Wickman K, Velimirovic B, Krapivinsky L, Clapham DE. The G-protein-gated atrial K+ channel IKACh is a heteromultimer of two inwardly rectifying K(+)-channel proteins. Nature 1995; 374:135-41; PMID:; http://dx.doi.org/ 10.1038/374135a0 [DOI] [PubMed] [Google Scholar]
  • 74.Mesirca P, Marger L, Toyoda F, Rizzetto R, Audoubert M, Dubel S, Torrente AG, Difrancesco ML, Muller JC, Leoni AL, et al.. The G-protein-gated K+ channel, IKACh, is required for regulation of pacemaker activity and recovery of resting heart rate after sympathetic stimulation. J Gen Physiol 2013; 142:113-26; PMID:; http://dx.doi.org/ 10.1085/jgp.201310996 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Ehrlich JR. Inward rectifier potassium currents as a target for atrial fibrillation therapy. J Cardiovasc Pharmacol 2008; 52:129-35; PMID:; http://dx.doi.org/ 10.1097/FJC.0b013e31816c4325 [DOI] [PubMed] [Google Scholar]
  • 76.Giudicessi JR, Ackerman MJ. Genetic testing in heritable cardiac arrhythmia syndromes: differentiating pathogenic mutations from background genetic noise. Curr Opin Cardiol 2013; 28:63-71; PMID:; http://dx.doi.org/ 10.1097/HCO.0b013e32835b0a41 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Jabbari J, Olesen MS, Holst AG, Nielsen JB, Haunso S, Svendsen JH. Common polymorphisms in KCNJ5 [corrected] are associated with early-onset lone atrial fibrillation in Caucasians. Cardiology 2011; 118:116-20; PMID:; http://dx.doi.org/ 10.1159/000323840 [DOI] [PubMed] [Google Scholar]
  • 78.Bettahi I, Marker CL, Roman MI, Wickman K. Contribution of the Kir3.1 subunit to the muscarinic-gated atrial potassium channel IKACh. J Biol Chem 2002; 277:48282-8; PMID:; http://dx.doi.org/ 10.1074/jbc.M209599200 [DOI] [PubMed] [Google Scholar]
  • 79.Kovoor P, Wickman K, Maguire CT, Pu W, Gehrmann J, Berul CI, Clapham DE. Evaluation of the role of I(KACh) in atrial fibrillation using a mouse knockout model. J Am College of Cardiol 2001; 37:2136-43; PMID:; http://dx.doi.org/ 10.1016/S0735-1097(01)01304-3 [DOI] [PubMed] [Google Scholar]
  • 80.Babenko AP, Aguilar-Bryan L, Bryan J. A view of sur/KIR6.X, KATP channels. Ann Reviewf Physiol 1998; 60:667-87; PMID:; http://dx.doi.org/ 10.1146/annurev.physiol.60.1.667 [DOI] [PubMed] [Google Scholar]
  • 81.Ashcroft FM. Adenosine 5′-triphosphate-sensitive potassium channels. Annu Rev Neurosci 1988; 11:97-118; PMID:; http://dx.doi.org/ 10.1146/annurev.ne.11.030188.000525 [DOI] [PubMed] [Google Scholar]
  • 82.Liang BT. Direct preconditioning of cardiac ventricular myocytes via adenosine A1 receptor and KATP channel. Ame J Physiol 1996; 271:H1769-77; PMID: [DOI] [PubMed] [Google Scholar]
  • 83.Ashcroft FM, Gribble FM. Correlating structure and function in ATP-sensitive K+ channels. Trend Neurosci 1998; 21:288-94; PMID:; http://dx.doi.org/ 10.1016/S0166-2236(98)01225-9 [DOI] [PubMed] [Google Scholar]
  • 84.Wojtovich AP, Urciuoli WR, Chatterjee S, Fisher AB, Nehrke K, Brookes PS. Kir6.2 is not the mitochondrial KATP channel but is required for cardioprotection by ischemic preconditioning. Am J Physiol Heart Circ Physiol 2013; 304:H1439-45; PMID:; http://dx.doi.org/ 10.1152/ajpheart.00972.2012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Quayle JM, Nelson MT, Standen NB. ATP-sensitive and inwardly rectifying potassium channels in smooth muscle. Physiol Rev 1997; 77:1165-232; PMID: [DOI] [PubMed] [Google Scholar]
  • 86.Nakaya H. Role of ATP-sensitive K+ channels in cardiac arrhythmias. J Cardiovasc Pharmacol Ther 2014; 19:237-43; PMID:; http://dx.doi.org/ 10.1177/1074248413515078 [DOI] [PubMed] [Google Scholar]
  • 87.Nichols CG, Koster JC, Remedi MS. beta-cell hyperexcitability: from hyperinsulinism to diabetes. Diabetes, Obesity Metab 2007; 9 Suppl 2:81-8; PMID:; http://dx.doi.org/ 10.1111/j.1463-1326.2007.00778.x [DOI] [PubMed] [Google Scholar]
  • 88.Nichols CG, Singh GK, Grange DK. KATP channels and cardiovascular disease: suddenly a syndrome. Circ Res 2013; 112:1059-72; PMID:; http://dx.doi.org/ 10.1161/CIRCRESAHA.112.300514 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Seino S, Miki T. Gene targeting approach to clarification of ion channel function: studies of Kir6.x null mice. J Physiol 2004; 554:295-300; PMID:; http://dx.doi.org/ 10.1113/jphysiol.2003.047175 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Mannhold R. KATP channel openers: structure-activity relationships and therapeutic potential. Med Res Rev 2004; 24:213-66; PMID:; http://dx.doi.org/ 10.1002/med.10060 [DOI] [PubMed] [Google Scholar]
  • 91.Bass AS, Hombo T, Kasai C, Kinter LB, Valentin JP. A historical view and vision into the future of the field of safety pharmacology. Handbook Exp Pharmacol 2015; 229:3-45; PMID:; http://dx.doi.org/ 10.1007/978-3-662-46943-9_1 [DOI] [PubMed] [Google Scholar]
  • 92.Redfern WS, Carlsson L, Davis AS, Lynch WG, MacKenzie I, Palethorpe S, Siegl PK, Strang I, Sullivan AT, Wallis R, et al.. Relationships between preclinical cardiac electrophysiology, clinical QT interval prolongation and torsade de pointes for a broad range of drugs: evidence for a provisional safety margin in drug development. Cardiovasc Res 2003; 58:32-45; PMID:; http://dx.doi.org/ 10.1016/S0008-6363(02)00846-5 [DOI] [PubMed] [Google Scholar]
  • 93.Stockbridge N, Morganroth J, Shah RR, Garnett C. Dealing with global safety issues: was the response to QT-liability of non-cardiac drugs well coordinated? Drug safety 2013; 36:167-82; PMID:; http://dx.doi.org/ 10.1007/s40264-013-0016-z [DOI] [PubMed] [Google Scholar]
  • 94.Sager PT, Gintant G, Turner JR, Pettit S, Stockbridge N. Rechanneling the cardiac proarrhythmia safety paradigm: a meeting report from the cardiac safety research consortium. Am Heart J 2014; 167:292-300; PMID:; http://dx.doi.org/ 10.1016/j.ahj.2013.11.004 [DOI] [PubMed] [Google Scholar]
  • 95.Cavero I, Holzgrefe H. CiPA: Ongoing testing, future qualification procedures, and pending issues. J Pharmacol Toxicol Methods 2015; 76:27-37; PMID: [DOI] [PubMed] [Google Scholar]
  • 96.Fermini B, Hancox JC, Abi-Gerges N, Bridgland-Taylor M, Chaudhary KW, Colatsky T, Correll K, Crumb W, Damiano B, Erdemli G, et al.. A new perspective in the field of cardiac safety testing through the comprehensive in vitro proarrhythmia assay paradigm. J Biomol Screening 2015; PMID: [DOI] [PubMed] [Google Scholar]

Articles from Channels are provided here courtesy of Taylor & Francis

RESOURCES