I Kur /Kv1.5 channel blockers for the treatment of atrial fibrillation

18
Review 10.1517/13543780902762850 © 2009 Informa UK Ltd ISSN 1354-3784 399 All rights reserved: reproduction in whole or in part not permitted I Kur /Kv1.5 channel blockers for the treatment of atrial fibrillation Juan Tamargo , Ricardo Caballero, Ricardo Gómez & Eva Delpón Universidad Complutense, School of Medicine, Department of Pharmacology, Madrid, Spain Atrial fibrillation (AF) is the most common sustained arrhythmia. Anti-arrhythmic drugs remain the mainstay of therapy, but the available class I and III anti-arrhythmic drugs are only moderately effective in long- term restoring/maintaining sinus rhythm (SR) and can produce potentially fatal ventricular pro-arrhythmia. In an attempt to identify safer and more effective anti-arrhythmic drugs, drug discovery efforts have focused on ‘atrial selective drugs’ that target cardiac ion channel(s) that are exclusively or predominantly expressed in the atria. The ultra-rapid activating delayed rectifier K + current (I Kur ), carried by Kv1.5 channels, is a major repolarizing current in human atria, but seems to play no role in the ventricle. This finding offers the possibility of developing selective I Kur blockers to restore and maintain SR without a risk of ventricular pro-arrhythmia. Several I Kur blockers are now being developed but clinical data are still limited, so the precise role of these agents in the treatment of AF remains to be defined. In this review we analyze the possible advantages and disadvantages of the developmental I Kur blockers as they represent the first step for the development of potential atrial selective drugs for a more effective and safer treatment and prevention of AF. Keywords: atrial fibrillation, cardiac action potential, cardiac ion channels, I Kur , Kv1.5 channels, pharmacology Expert Opin. Investig. Drugs (2009) 18(4):399-416 1. Introduction Atrial fibrillation (AF) is the most common arrhythmia in clinical practice, with an estimated prevalence of 0.4 – 1% in the general population, increasing with age to 8% in patients older than 80 years [1]. AF impairs cardiac output, exercise tolerance and quality of life, increases stroke risk (five-fold), exacerbates heart failure, and is independently associated with a two-fold increase in mortality [1]. Overall, hospitalizations for AF have increased two- to three-fold in recent years, accounting for approximately one third of hospitalizations for cardiac rhythm disturbances [1,2]. Thus, AF represents a considerable financial burden on the healthcare system. Although AF can occur in the absence of underlying disease (idiopathic or lone AF), it is most commonly seen in patients with structural heart disease, including heart failure, arterial hypertension, coronary artery disease, cardiothoracic surgery, mitral valvular disease or thyrotoxicosis. Moreover, familial AF occurrence is more common than previously recognized, an observation supporting a genetic susceptibility for developing this dysrhythmia [1]. 1.1 Atrial fibrillation is often a self-perpetuating disease In many patients, initial paroxysms of AF (self-terminating) become more frequent and of longer duration to the point where AF becomes persistent (i.e., AF requires cardioversion to restore sinus rhythm; SR) and then permanent over time (AF cannot be converted to SR). This progression involves changes in the electrophysiological, structural and contractile properties of the atria (atrial remodeling) 1. Introduction 2. Atrial action potential 3. Treatment of atrial fibrillation 4. Structure of Kv1.5 channels encoding I Kur 5. The ultrarapid activating delayed rectifier K + current, I Kur 6. Electrophysiological effects of I Kur blockers 7. The future of new I Kur blockers (the two sides of the coin) 8. Expert opinion Expert Opin. Investig. Drugs Downloaded from informahealthcare.com by CNIO on 10/06/14 For personal use only.

Transcript of I Kur /Kv1.5 channel blockers for the treatment of atrial fibrillation

Review

10.1517/13543780902762850 © 2009 Informa UK Ltd ISSN 1354-3784 399All rights reserved: reproduction in whole or in part not permitted

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillationJuan Tamargo†, Ricardo Caballero, Ricardo Gómez & Eva Delpón†Universidad Complutense, School of Medicine, Department of Pharmacology, Madrid, Spain

Atrial fibrillation (AF) is the most common sustained arrhythmia. Anti-arrhythmic drugs remain the mainstay of therapy, but the available class I and III anti-arrhythmic drugs are only moderately effective in long-term restoring/maintaining sinus rhythm (SR) and can produce potentially fatal ventricular pro-arrhythmia. In an attempt to identify safer and more effective anti-arrhythmic drugs, drug discovery efforts have focused on ‘atrial selective drugs’ that target cardiac ion channel(s) that are exclusively or predominantly expressed in the atria. The ultra-rapid activating delayed rectifier K+ current (IKur), carried by Kv1.5 channels, is a major repolarizing current in human atria, but seems to play no role in the ventricle. This finding offers the possibility of developing selective IKur blockers to restore and maintain SR without a risk of ventricular pro-arrhythmia. Several IKur blockers are now being developed but clinical data are still limited, so the precise role of these agents in the treatment of AF remains to be defined. In this review we analyze the possible advantages and disadvantages of the developmental IKur blockers as they represent the first step for the development of potential atrial selective drugs for a more effective and safer treatment and prevention of AF.

Keywords: atrial fibrillation, cardiac action potential, cardiac ion channels, IKur, Kv1.5 channels, pharmacology

Expert Opin. Investig. Drugs (2009) 18(4):399-416

1. Introduction

Atrial fibrillation (AF) is the most common arrhythmia in clinical practice, with an estimated prevalence of 0.4 – 1% in the general population, increasing with age to 8% in patients older than 80 years [1]. AF impairs cardiac output, exercise tolerance and quality of life, increases stroke risk (five-fold), exacerbates heart failure, and is independently associated with a two-fold increase in mortality [1]. Overall, hospitalizations for AF have increased two- to three-fold in recent years, accounting for approximately one third of hospitalizations for cardiac rhythm disturbances [1,2]. Thus, AF represents a considerable financial burden on the healthcare system. Although AF can occur in the absence of underlying disease (idiopathic or lone AF), it is most commonly seen in patients with structural heart disease, including heart failure, arterial hypertension, coronary artery disease, cardiothoracic surgery, mitral valvular disease or thyrotoxicosis. Moreover, familial AF occurrence is more common than previously recognized, an observation supporting a genetic susceptibility for developing this dysrhythmia [1].

1.1 Atrialfibrillationisoftenaself-perpetuatingdiseaseIn many patients, initial paroxysms of AF (self-terminating) become more frequent and of longer duration to the point where AF becomes persistent (i.e., AF requires cardioversion to restore sinus rhythm; SR) and then permanent over time (AF cannot be converted to SR). This progression involves changes in the electrophysiological, structural and contractile properties of the atria (atrial remodeling)

1. Introduction

2. Atrial action potential

3. Treatment of atrial fibrillation

4. Structure of Kv1.5 channels

encoding IKur

5. The ultrarapid activating

delayed rectifier K+ current, IKur

6. Electrophysiological effects of

IKur blockers

7. The future of new IKur blockers

(the two sides of the coin)

8. Expert opinion

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillation

400 ExpertOpin.Investig.Drugs(2009) 18(4)

in a way that promotes the maintenance and recurrence of the arrhythmia. AF produces a rapid (within hours) and inhomogeneous shortening of the atrial action potential duration (APD) and effective refractory period (AERP) and a lack of adaptation of APD to fast heart rates [1,3-5]. Long-term AF also slows intra-atrial conduction, which together with the structural changes in the atrial wall facilitates AF perpetuation after changes in AERP are stabilized. Thus, atrial remodeling explains why paroxysmal AF tends to become persistent and/or permanent, chronic AF becomes resistant to anti-arrhythmic drugs, the higher success rate of cardioversion when AF is of short duration and the early recurrences of AF after cardioversion. The development of more effective anti-arrhythmic drugs should be based on a better understanding of the electrophysiological mechanisms underlying AF. AF is caused by a variety of mechanisms, including local ectopic activity and single- and multiple-circuit re-entry, which are not mutually exclusive and may coexist [1,3]. Over the last 50 years, the multiple wavelet re-entry hypothesis has been the most widely accepted conceptual model to explain AF [1,3]. Accordingly, AF is maintained by multiple continuously propagating waves of atrial re-entry (wavelets) separated by lines of conduction block, fixed (anatomical obstacles, branching, scars or areas of fibrosis) or functional (non-uniform recovery repolarization, regional stretch, gap-junction uncoupling), resulting in a short excitable gap between the fibrillation waves [3-5]. The number of wavelets that can coexist is determined by the atrial size and the wavelength (WL) (product of conduction velocity and AERP) of the atrial impulse. A large atrial mass with a short AERP and slow impulse conduction increases the number of wavelets, favor-ing the onset and maintenance of AF. The electrophysiologi-cal atrial remodeling shortens the atrial APD and refractoriness and slows conduction velocity in a heteroge-nous manner, providing a substrate for multiple-circuit re-entry. Thus, according to the multiple wavelet re-entry hypothesis, an effective way to prevent or terminate AF is to prolong the AERP by lengthening the atrial APD. This would increase the wavelength, thereby reducing the number of wavelets that can coexist in the atria to the point that AF fails to be sustained.

2. Atrialactionpotential

The human atrial action potential reflects a sequential activation and inactivation of ion channels that conduct depolarizing, inward (Na+ and Ca2+), and repolarizing, outward (K+), currents [6,7]. Figure 1 shows the ionic currents involved in the genesis of the five phases of the human atrial and ventricular action potential. The initial upstroke of the atrial cells (Phase 0) is due to the opening of Na+ channels. Most cardiac Na+ channels open transiently within milliseconds upon membrane depolarization, resulting in the peak transient current (INa), which determines excitation and conduction.

However, a small percentage of Na+ channels may continue to open and close spontaneously, carrying the so-called per-sistent or late Na+ current (INaL), which plays a major con-tribution to the shape and duration of the action potential plateau. The initial rapid repolarization (Phase 1) is a conse-quence of the rapid voltage-dependent closure (inactivation) of Na+ channels, and the activation of the transient outward current (Ito) and the ultrarapid component of the delayed rectifier K+ current (IKur). During Phase 2, inward depolar-izing currents through Na+ (slowly inactivated, INaL) and L-type Ca2+ channels (ICaL) are balanced by the different components of the delayed rectifier K+ current (IKur, rapid-IKr and slow-IKs). The net outward current during Phase 3 of repolarization is provided by the delayed rectifier K+ cur-rents, along with the inactivation of the ICaL. The strong inward rectifier K+ current (IK1) contributes to the terminal Phase 3 of repolarization and sets the resting potential (around − 80 mV) close to the K+ equilibrium potential (Phase 4). In atrial cells, acetylcholine released on vagal stimulation binds to muscarinic (M2) receptors and activates another inward rectifier K+ current (IKAch). Activation of IKAch hyper-polarizes the resting membrane potential and produces a non-uniform shortening of atrial APD and AERP, which decreases the WL of re-entrant circuits and promotes the stability of AF. The differences in action potential wave-forms between atrial and ventricular cells reflect differences in ion channel expression levels (Figure 1) [4,6,7]. The greater density of outward repolarizing currents (Ito and IKur) com-pared with inward currents can explain why the human atrial action potential is shorter and the plateau occurs at lower potentials than the ventricular action potential. Moreover, IKur and Ito are the main repolarizing currents in the human atria, while IKr and IKs are the main repolarizing currents in ventricular myocytes and Purkinje fibers. Most important, IKur is a major repolarizing current in human atria but is virtually absent in human ventricle [8-10]. In addition, atrial cells display a more gradual Phase III repo-larization, present a more depolarized resting membrane potential, possibly due to a lower density of IK1 [11] and a more negative diastolic threshold of excitation than that of ventricular cells.

3. Treatmentofatrialfibrillation

Although several non-pharmacological therapeutic approaches (i.e., surgery, catheter ablation, internal atrial defibrillators and pacing) have been developed, anti-arrhythmic drug therapy remains the mainstay of managing AF. The anti-arrhythmic treatment of AF involves two main strategies: to restore and maintain normal SR (rhythm control) with class I and III anti-arrhythmic drugs and/or to allow patients to remain in AF, but control the ventricular response (rate control) using atrio-ventricular nodal blocking drugs (beta-blockers, amiodarone, digitalis glycosides, or nondihydropyridine calcium channel antagonists) [1].

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

Tamargo,Caballero,Gómez&Delpón

ExpertOpin.Investig.Drugs(2009) 18(4) 401

3.1 RateversusrhythmcontrolWhile restoring and maintaining SR has several theoretical advantages over rate control, until recently there has been little evidence supporting the comparative advantages of either strategy. Five recent randomized clinical trials (PIAF, AFFIRM, RACE, STAF, HOT-CAFE) [12-16] compare rate with rhythm control in patients with recurrent AF or high risk of recurrences [1] (see [17] for a review). No significant differences were observed between both strategies in all-cause mortality, symptoms, quality of life, stroke or in any other point examined, although a non-significant trend for excess mortality in the rhythm control strategy was observed [1,17]. Moreover, rate control was associated with fewer hospitalizations and fewer adverse drug effects, being more cost-effective [17]. Thus, in appropriate patients with persistent AF or high risk of recurrence, a strategy of ventricular rate control, in combination with oral anti-coagulation, can be at least as effective as rhythm control in preventing clinical outcomes and may, in fact, prove to be superior. There are several explanations for these results [17]. One possible explanation is that the beneficial antiarrhythmic effect (e.g., maintenance of SR) on survival may be offset by the low efficacy of rhythm control to maintain SR (the percentage of patients in SR at the end of the study ranged between 38% and 63.5%) and the high frequency of serious adverse effects (pro-arrhythmia, bradycardia, organ toxicity) associated with anti-arrhythmic drug therapy, which, additionally, led to a higher rate of early drug discontinuation [12,14,18]. Additionally, premature discontinuation of oral anticoagulation in some patients who were believed

to be in SR could have contributed to an increased incidence of thromboembolic events. AF is common in patients with heart failure and is associated with an adverse outcome [1]. The AFFIRM trial suggests that patients without congestive heart failure fared significantly better with rate control [19]. Very recently, the AF-CHF (Atrial Fibrillation and Congestive Heart Failure) trial enrolled 1376 patients with systolic heart failure (left ventricular ejection fraction ≤ 35%, and New York Heart Association (NYHA) functional class II – III) and AF (two thirds had persistent AF) receiving optimal heart failure therapy and anticoagulation [20]. Over a follow-up period of 37 months, the study found no differences between rhythm- (amiodarone in 82% of patients) and rate-control strategies (β-blockers and digoxin) in reducing cardiovascular deaths or other secondary outcomes (death from any cause, worsening of heart failure, or stroke). However, the rate-control strategy eliminated the need for repeated cardioversion and reduced the rate of hospitalization. These findings suggest that rate control should be considered a primary approach for patients with AF and congestive heart failure. Nevertheless, there are several concerns with the study, including a high crossover rate (21%) between study groups and the suboptimal control of SR in the rhythm-control group (58% of patients had at least 1 recurrent episode of AF); at 12 months, significantly fewer patients received β-blockers in the rhythm-control group which could have influenced mortality; and, finally, during the follow-up, the percentage of patients in the rhythm-control group receiving amiodarone dropped to 73% at 36 months, which might have negated the potential benefits of this strategy.

A. B.

IK1, IKAch, IKATPINa

Ito1

IKur

ICaL

IKr, IKs

INaLPhase 0

Phase 1

Phase 2

Phase 3

Phase 4

Ito1

IKr

IKs

IK1, IKATP

Phase 0

Phase 1

Phase 2

Phase 3

Phase 4

ICaL

INa

INaL

Figure 1. Ionic currents involved in shaping of human atrial (A) and ventricular (B) action potentials. The initial upstroke (Phase 0) of the action potential is due to the activation of the inward Na+ current (INa). Repolarization is mainly due to outward efflux of K+ (blue arrows) that counterbalances the depolarizing effects of inward Ca2+ (ICaL) and Na+ (peak INa and late-INaL) influx (red arrows). The main voltage-dependent human K+ currents (Ito, IKur, IKr and IKs) are represented together with the inward rectifier currents (IK1, IKACh and IKATP). Some currents are exclusive to the atria (IKur) or are more important (Ito, IKACh) in the atria than in the ventricles and vice versa (IKr). See text for a proper explanation.

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillation

402 ExpertOpin.Investig.Drugs(2009) 18(4)

3.2 Rhythm-controlinAFClass IA and IC (Na+ channel blockers) and class III anti-arrhythmic drugs (which prolong atrial APD and AERP without slowing impulse conduction) have been widely used to restore and maintain normal SR for cardioversion of AF to SR [1]. Class IA and IC drugs prolong AERP by slowing the recovery of availability of Na+ channels, but they also exhibit other ancillary properties that contribute to their anti-arrhythmic action, including the blockade of several K+ currents [21]. Class III agents prolong atrial APD and AERP primarily by blocking one or several outward repolarizing K+ currents [22]. However, available drugs exhibit modest efficacy for maintaining SR (< 50% of patients remain in SR 1 year after cardioversion), produce potentially serious adverse effects and most of them are contraindicated in patients with several cardiovascular conditions commonly associated with AF, including heart failure, coronary artery disease and hypertension [1]. Moreover, these drugs prolong atrial APD via blockade of the IKr and/or IKs, which are expressed in both human atria and ventricle [1,3-5]; and, thus, they prolong ventricular repolarization and the corrected QT interval for heart rate (QTc) on the electrocardiogram (ECG) and increase the risk for ventricular tachyarrhythmias, such as torsades de pointes (TdP), which limits their utility in the treatment of AF.

In an attempt to identify safer and more effective drugs to terminate/prevent AF without the risk of ventricular pro-arrhythmia, drug discovery efforts have focused on ‘atrial selective drugs’, i.e., drugs that target cardiac ion channel(s) that are exclusively or predominantly expressed in the atria. Because IKur carried by Kv1.5 channels is present only in the atria, but not in the ventricles [6-10], IKur blockers would be expected to prolong human atrial APD and AERP, convert AF to SR and/or prevent recurrences of AF without undesired effects on impulse propagation, ventricular repolarization (QTc interval) or cardiac output and with a minimum risk of ventricular pro-arrhythmia.

In this review, we address the molecular basis and biophysical properties of IKur and the pharmacological properties, advantages and disadvantages of the new IKur blockers, as an example of atrial-selective drugs that might convert AF to SR and prevent recurrences of AF without the pro-arrhythmic risk of many currently available class I and III agents.

4. StructureofKv1.5channelsencodingIKur

Kv1.5 channels belong to the superfamily of voltage-gated K+ channels (Kv), formed by coassembly of four identical pore-forming α-subunits encoded by the KCNA5 gene (human chromosome 12p13.3) and four accessory β-subunits (Kvβ1.2, Kvβ1.3, and Kvβ2.1) that bind to the N-terminus of the α-subunit to form α4β4 complexes. Each α-subunit (613 amino acids) contains six transmembrane-spanning segments (S1 – S6) with cytoplasmic N-and C-terminal domains (Figure 2) [6,7,23-25]. The S1 – S4 module represents

the voltage sensor, which reacts to potential changes across the membrane and regulates the gating of the channel. The S4 segment contains positively charged residues (Arg or Lys) at every third position and serves as the principal voltage sensor. The highly conserved selectivity filter signature (Gly-Tyr-Gly) is located between S5 and S6 and is connected to S5 via the pore helix and a short loop located outside the membrane. The pore of the channel, through which K+ ions flow across the plasma membrane, is formed by S5, S6, the selectivity filter and the pore helix. The intracellular part of the pore is arranged in what has been dubbed an ‘inverted teepee’ cavity, with the selectivity filter located at its wide end [23]. The large water-filled pore cavity is lined with highly hydrophobic amino acids and has been identified as the binding site for IKur blockers. Kv1.5 α-subunits coassemble with Kvβ1.2 subunits to form the IKur in human atrium [6,7,26]. Immunohistochemical studies show that, in human atria, Kv1.5 channels are concentrated at the level of the intercalated disks (close to N-cadherin) and T-tubules, with some staining at the cell surface [27,28]. Although Kv1.5 channels are expressed both at the mRNA and protein levels in human ventricular tissues, functional IKur only exists in atrial cells [7,26,29].

5. TheultrarapidactivatingdelayedrectifierK+current,IKur

5.1 CurrentfeaturesandmolecularbasisIKur has been recorded in human [8,9,29], rat [30] and canine [31] atrial myocytes and in adult mouse ventricular myocytes (IK,slow1) [32]. The biophysical and pharmacological characteristics of the currents generated through Kv1.5 channels in heterologous expression systems are similar to the IKur recorded in human atria [7], and antisense-targeting Kv1.5 specifically suppresses IKur in human and rat atrial myocytes [29], which confirms that Kv1.5-based channels (10 – 14 pS) are the molecular basis for IKur. Kinetic differences of activation, inactivation and recovery from inactivation discriminate between various K+ currents. At 37°C, human atrial IKur presents a rapid activation (τ < 2 ms at 0 mV, midpoint (Vh) − 7 mV and slope factor (k) 6 – 12 mV) and outward rectification and a slow and incomplete inactivation during the time course of the action potential [6,8,9,33]. Recovery from inactivation proceeds slowly (time constants of 0.42 and 7.9 s at − 80 mV), which confers substantial rate-dependent properties on IKur over a physiologically relevant range of frequencies [33]. Coexpression of Kvβ1.2 subunits with Kv1.5 channels causes a hyperpolarizing shift in the voltage dependence of activation and inactivation gating and slows the rate of Kv1.5 channel deactivation [34]. It is widely accepted that IKur inactivates by a ‘type C’ mechanism as a result of a concerted constriction of the external mouth of the pore involving all four α-subunits [6,7,35]. C-type inactivation is thought to be voltage independent and occurs predominantly from the open state, is slowed by external tetraethylammonium (TEA) or elevation of extracellular K+

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

Tamargo,Caballero,Gómez&Delpón

ExpertOpin.Investig.Drugs(2009) 18(4) 403

concentration ([K+]o) or when the residue Arg485, located in the outer pore mouth, is substituted (Arg485Val) [36]. Binding of external K+ ions within the conducting pathway inhibits inac-tivation via a ‘foot-in-the-door’ mechanism that prevents con-striction of the outer pore mouth [36]. However, IKur does not present N-type inactivation because Kv1.5 channels do not have a long N-terminal domain [37].

5.2 PharmacologyofIKur

IKur is relatively insensitive to TEA, Ba2+ and selective IKr blockers (dofetilide, E-4031 and MK-499), but is highly sensitive to low concentrations of 4-aminopyridine (4-AP, ∼ 10 μM) [6,8,37]. Several anti-arrhythmic drugs (ambasilide, amiodarone, clofilium, dronedarone, flecainide, quinidine, propafenone, 5-hydroxy-propafenone, tedisamil) [6,37-40], local anesthetics (bupivacaine [41,42] and benzocaine [43,44]), angiotensin II type 1 receptor antagonists (losartan, candesartan, irbesartan) [45-47], spironolactone [47,48], the pacemaker-current inhibitors (zatebradine [49], ivabradine [50]), atorvastatin [51] and some calcium antagonists (verapamil, diltiazem, nifedipine) [39,52,53] inhibit IKur, in addition to their effects on other cardiac ionic channels. However, in the last decade, several pharmaceutical companies have developed selective IKur blockers in an attempt to prevent and treat AF (see [54-56] for a review). However, most of these agents can also inhibit cardiac Na+,

Ca2+ and a variety of K+ channels. The ion channel blocking profile and the chemical structures of some investigational IKur blockers is summarized in Table 1. In general, IKur blockers appear to act via a preferential interaction with Kv1.5 channels in their open state [6,57]. This assessment is based on the following findings.

They produce a rapid current decay upon depolariza-1. tion and upon repolarization slow current deactivation producing a crossover of tail current traces recorded in the absence and in the presence of the drug. These results indicate that the drugs enter into the channel cavity only after channel opening and inhibit IKur by binding to a receptor site located at this level, preventing normal closure of the inactivation gate (foot-in-the-door); thus, the drug must dissociate from their receptor site before the channel can close.They display a reduced affinity for channels in the 2. resting-closed (i.e., at a holding potential of – 80 mV) or inactivated state (i.e., at positive potentials).Kv1.5 channel block increases with the stimulation 3. frequency, which represents an advantageous property because channels would be most affected at fast action potential firing rates characteristics for AF and less affected in SR.

NH2 COOH(613)

Kv1.5 × 4

-

-

--

-

-

In

Out

Voltage sensorPore

+

+

++

Po

A.

α

α

α

α

β

ββ

SF

S5

S6

S5

S6

Po Po

V505I508V512V516

T479, T480

B.

Kvβ

S1 S2 S3 S4 S5 S6

β

Figure2.MolecularstructureoftheKv1.5channel(A)andviewoftheS5-S6domainoftwoα-subunitsoftheKv1.5channelshowingtheresiduesfacingtheporecavityinvolvedintheinhibitioninducedbyvernakalant,S9947andAVE0118(B).(A) Upper part: schematic representation of the structure of one Kv1.5 α-subunit with 6 membrane-spanning domains and the intracytoplasmatic accesory

β-subunits (Kvβ and KChiP). Lower part: α and accessory β-subunits coassemble as tetramers to form the functional channel. (B) Po: Pore helix;SF: Selectivity filter.

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillation

404 ExpertOpin.Investig.Drugs(2009) 18(4)

Tab

le1

.Eff

ects

of

I Ku

r/K

v1.5

blo

cker

so

nc

ard

iac

ion

cu

rren

ts.

Dru

gI K

ur

Kv1

.5I to

I Kr

I Ks

I K1

I KA

chI K

ATP

I Na

I CaL

Spec

ies

[ref

]

Aca

cetin

3.2

9.3

NS

at 3

04

NS

at 3

0N

S at

30

HEK

-293

, HA

M [8

0]

AV

E011

81.

1 –

1.3

1.1

– 6.

21.

8 –

3.4*

1010

% a

t 10

μM

4.5

10%

at

10

μM

22%

at

10

μM

CH

O, X

O, G

PAM

, GPV

MPA

M [8

1]

AV

E123

10.

9 –

1.1

3.6

3.3

– 5.

930

NS

at 3

0 μM

NS

at

30 μ

M8.

435

% a

t

30 μ

MN

S at

30

μM26

% a

t

30 μ

MG

PVM

, CH

O, G

PAM

, PA

M [8

2,83

]

AV

E329

50.

30.

30.

56%

at

10

μM

2.4

XO

, CH

O, P

AM

[81,

83]

AZD

7009

2724

*0.

619

34.

3 (a

t 10

Hz)

CH

O [8

4,85

]

C-9

356

4.4

3497

335

HEK

-293

[31]

Com

poun

d 9

4.2

100

7.8

340

Pate

nt W

O/2

006/

1363

05

DPO

-10.

03 –

0.0

80.

16N

S at

1 μ

M3%

at

3

μM24

% a

t 10

μM

15%

at

3

μMC

HO

, XO

, HA

M [8

6]

ICA

GEN

-41

0.16

– 1

.6Lt

k- , L9

29 C

ells

, XO

[66]

ISQ

-123

8 nM

> 3

0<

20%

at

10

μM

< 2

0% a

t

10 μ

M<

20%

at

10 μ

MC

HO

[87-

89]

NIP

-142

5.3

4.8

16.3

4440

% a

t

10 μ

M0.

640

50%

at

10

μM

HA

M, H

EK-2

93, X

O*

[90,

91]

S994

70.

070.

4220

% a

t

10 μ

M0%

at

10

μM

15%

at

10

μM

0C

HO

, XO

, RV

M,

HA

M [6

6,91

,92]

Ver

naka

lant

9 at

1 H

z0.

8 at

3 H

z13

3021

> 1

mM

1119

at

1 H

z22

0G

PVM

, HEK

-293

[65,

93-9

5]

XEN

-D01

010.

150.

24[5

5]

*hK

v4.3

/KC

hIP2

.2.

Dat

a ar

e ex

pres

sed

as IC

50 v

alue

s [c

once

ntra

tions

(μM

) pro

duci

ng t

he h

alf

max

imum

inhi

bitio

n].

CH

O: C

hine

se h

amst

er o

vary

cel

ls; G

PAM

: Gui

nea

pig

atria

l myo

cyte

s; G

PVM

: Gui

nea

pig

vent

ricul

ar m

yocy

tes;

HA

M: H

uman

atr

ial m

yocy

tes;

HEK

-293

: Hum

an e

mbr

yoni

c ki

dney

cel

ls; L

929:

Mou

se fi

brob

last

s; L

tk:

Mou

se fi

brob

last

s; P

AM

: Pig

atr

ial m

yocy

tes;

RV

M: R

at v

entr

icul

ar m

yocy

tes;

XO

: Ooc

ytes

of

Xen

opus

laev

is.

Aca

cetin

: 5,7

-dih

ydro

xy-2

-(4-

met

hoxy

-phe

nyl)-

4H-1

-ben

zopy

ran-

4-on

e.

AV

E011

8 (s

anofi

-ave

ntis

): 2′

-{[2

-(4-

met

hoxy

-phe

nyl)-

acet

ylam

ino]

-met

hyl}-

biph

enyl

-2-c

arbo

xylic

aci

d (2

-pyr

idin

-3-y

l-eth

yl)-

amid

e.

AZD

7009

: ter

t-Bu

tyl-2

-(7-

[(2S)

-3-(

4-cy

anop

heno

xy)-

2-hy

drox

ypro

pyl]-

9-ox

a-3,

7-di

azab

icyc

lo[3

.3.1

]non

-3-y

l)eth

ylca

rbam

ate.

DPO

-1(M

erck

and

Co.

): 2-

isop

ropy

l-5-m

ethy

lcyc

lohe

xyl)

diph

enyl

phos

phin

e ox

ide.

ICA

GEN

-4: N

-[3-

(4-e

thyl

-ben

cene

-sul

fony

lam

ino)

-2-h

ydro

xy-in

dan-

5-yl

]-3-

met

hoxy

-ben

zam

ide.

ISQ

-1 (M

erck

& C

o.):

3-[(d

imet

hyla

min

o)m

ethy

l]-6-

met

hoxy

-2-m

ethy

l-4-p

heny

lisoq

uino

lin-1

(2H

)-on

e.

NIP

-142

(Nis

san

Che

mic

al In

dust

ries)

: (3R

*, 4

S*)-

4-cy

clop

ropy

lam

ino-

3,4-

dihy

dro-

2,2-

dim

ethy

l-6-(

4-m

etho

xyph

enyl

acet

hyla

min

o)-7

-nitr

o-2H

-1-b

enzo

pyra

n-2-

ol h

ydro

chlo

ride.

S994

7 (s

anofi

-ave

ntis

): 2′

-(be

nzyl

oxyc

arbo

nyla

min

omet

hyl)-

biph

enyl

-2-c

arbo

xylic

aci

d 2,

4-di

fluor

o-be

nzyl

amid

e.

TAEA

(Mer

ck a

nd C

o.):

2-ph

enyl

-1,1

-dip

yrid

in-3

-yl-2

-pyr

rolid

in-1

-yl-e

than

ol.

Ver

naka

lant

: 3-P

yrro

lidin

ol, 1

-[(1

R,2R

)-2-

[2-(

3,4-

dim

etho

xyph

enyl

) eth

oxy]

cycl

ohex

yl]-

,hyd

roch

lorid

e, (3

R).

AV

E123

1 (s

anofi

-ave

ntis

), C

9356

(Car

diom

e/A

stel

las

Phar

ma)

and

XEN

-D01

01 (X

entio

n): s

truc

ture

indi

sclo

sed.

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

Tamargo,Caballero,Gómez&Delpón

ExpertOpin.Investig.Drugs(2009) 18(4) 405

However, in human atrial myocytes IKur amplitude is markedly reduced at rapid activation rates [58,59], so that it is possible that, during AF, IKur may be already so reduced that there is little current to block [60,61]. On the contrary, in a rat model, very rapid atrial pacing produces an immediate, transient and rate-dependent increase in Kv1.5 mRNA (peak effect at 2 h, but returns to baseline after 8 h) and a gradual decrease of Kv4.2 and Kv4.3 mRNA, while mRNA levels of Kv1.2, Kv1.4, Kv2.1, rERG, KvLQT1 and minK remain unchanged [62]. Thus, it is possible that the increased expression of the Kv1.5 channel protein might contribute, at least in part, to the rapid shortening of the AERP at the onset of AF. If so, it is possible that selective IKur blockers might antagonize the shortening of the atrial APD and AERP and prevent new episodes of AF, being more effective in preventing AF (because of the large amplitude of IKur at slow heart rates), than in terminating the arrhythmia [58].

5.3 MoleculardeterminantsofKv1.5channelblockThe Kv1.5 channel pore cavity has been identified as the binding site for open channel blockers. In fact, the potency of anti-arrhythmic drugs (quinidine) and local anesthetics (bupivacaine, benzocaine) to block Kv1.5 channels is affected by introducing specific mutations in residues located at the base of the pore helix (Thr479 and Thr-480) near the selectivity filter and in the S6 (Thr-507, Leu-510 and Val-514) [42,44,57]. Therefore, it was predicted that these latter residues are located facing towards the central cavity of the channel pore and identified as potential binding sites for these drugs. However, residues Thr-507, Leu-510 and Val-514 that determine the Kv1.5 channel blockade induced by irbesartan [46] and S0100176 [63] are not facing the channel pore. More recently, Ala-scanning mutagenesis of the pore domain, homology modeling of the Kv1.5 pore and docking simulation of drug-channel interaction confirm that two residues located in the base of the pore helix (Thr-479 and Thr-480) and several residues located in S6 facing the pore cavity are the potential sites for IKur block induced by AVE0118 (Val-505, Ile-508, Val-512 and Val-516), S0100176 (Val-505, Ile-508 and Val-512) and vernakalant (Val-505 and Ile-508) [63-65]. The positions of these residues are highly conserved among different cardiac voltage-dependent K+ channels (i.e., hERG or Kv11.1, KCNQ1 or Kv7.1, and Kv1.3), which explains why the development of selective IKur blockers is a difficult task and most of these compounds might also affect other K+ currents with affinities similar to that of Kv1.5 (Table 1) [63]. β-subunits can modulate drug-induced Kv channel blockade. Even when most of the IKur blockers seem to interact with a receptor site located in the channel cavity accessible from the cytoplasmic mouth of the pore, other alternative access routes for drugs to their binding site of Kv1.5 channels have been proposed, such as through a crevice between S5 and S6, in a similar way to that described for local anesthetics in voltage gated Na+ channels [51,66]. Furthermore, some evidence suggests that the external mouth of

the channel pore, formed by the P loop and adjacent S5 – S6 segments, is the binding site for some long-chain poly- unsaturated fatty acids (arachidonic and docosahexaenoic acids) [67] and for nifedipine [53]. The finding that R(+)-N-methyl-bupivacaine, a membrane-impermeant permanently charged drug, produces different effects on IKur when applied from the intra- or the extracellular side of the membrane, confirms the existence of an external binding site for bupivacaine-like local anesthetics [68].

6. ElectrophysiologicaleffectsofIKurblockers

Selective IKur blockers would be expected to prolong atrial APD and AERP, to decrease atrial vulnerability and reverse AF to SR without affecting ventricular repolarization (QTc interval) or increasing the risk of ventricular pro- arrhythmia. At low micromolar concentrations (10 – 50 μmol/L) 4-AP fully inhibits IKur without affecting other cur-rents [9,69,70], and thus it can be considered a selective IKur blocker. However, the reported effects of 4-AP on atrial APD90 are contradictory. Thus, 4-AP shortens atrial APD90 and ERP in isolated human atria [71] and in canine coro-nary-perfused right atrial preparations [60,72] and in human myocytes obtained from patients in SR (nonremodeled ‘healthy’ atria), these effects are accompanied by an increase in the height and a prolongation of the duration of the pla-teau phase (APD20) [73]. On the contrary, it has been reported that 4-AP prolongs APD90 in atrial myocytes iso-lated from canine atria [69,74], patients undergoing coronary bypass or cardiac surgery [9,75] and in anesthetized dogs, where 4-AP, at plasma concentrations between 25 and 50 μmol/L, significantly lengthens AERP without affecting ven-tricular refractoriness [70]. These contradictory results can be attributed to the different driving rates and experimental protocols (isolated atrial myocytes, isolated atria, in vivo vs. in vitro) and to the effects of 4-AP on autonomic tone [70,76]. Furthermore, in isolated myocytes from patients with chronic AF (‘remodeled atria’), 4-AP only slightly prolongs the APD90 [72,73]. AVE0118, at a concentration (6 μmol/L) at which Ito, IKAch and some IKr are also blocked (Table 1), reproduces the effects of 4-AP, so that it elevates the plateau potential and shortens the APD90 in atrial trabeculae in SR, but prolongs APD90 in AF [73]. It has been reported that both IKur density and Kv1.5 protein expression can be reduced by 50% in atrial cardiomyocytes from patients with valvular disease and persistent AF [77], a result that has not been confirmed in subsequent studies [75,78], probably reflecting differences in patient populations, AF duration or concomitant medications [4]. However, this does not imply that IKur blockers would not be effective in prolonging AERP in patients with AF. Using a mathematical model of the human atrial AP (Figure 3), it was predicted that inhibition of IKur by 90% leads to an increased activation of the ICaL, which shifts the plateau level to positive potentials at which IKr is rapidly activated, so that mid and late repolarization are

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillation

406 ExpertOpin.Investig.Drugs(2009) 18(4)

accelerated and thus, the APD remains unchanged during SR [79]. In accordance with the model predictions, in the presence of an IKur blocker (E-4031), 4-AP no longer shortens but prolongs atrial APD90 in SR. However, AF-induced electrical remodeling reduces ICaL and Ito and markedly shortens the atrial APD, so that inhibition of IKur increases the plateau potential over a range (below 0 mV) where the activation of other K+ repolarizing currents (IKr and IKs) is less pronounced, and so a prolongation of the APD90 is observed. These findings suggest that the shortening of atrial APD associated with AF increases the contribution of IKur to atrial repolarization even when the current can be downregulated [73,79]. Thus, selective IKur blockers may be more, rather than less, effective in prolonging APD in patients with AF than in controls.

6.1 EffectsofinvestigationalIKurblockersinanimalmodelsThe effects of some novel IKur blockers on cardiac ion currents recorded in native cells as well as in heterologous expression systems are summarized in Table 1. In anesthetized and/or conscious dogs [80-82,88-90,92,96-98], goats [82,89,99,100], pigs [81,82,92] and African green monkeys [88,89], novel IKur blockers produce a dose-dependent prolongation of atrial APD and ERP, an effect that is stronger on left atria (LA) (which has shorter refractoriness) than on right atria (RA), and decrease LA vulnerability to AF, without affecting intra-atrial conduction velocity, ventricular ERP or QTc intervals [55,90,92,97]. In fact, they are highly effective in terminating/preventing AF (induced by rapid atrial pacing, vagal stimulation, aconitine or heart failure) and/or atrial flutter (AFL; induced by intracaval and atrial free wall lesions, canine sterile pericarditis) without affecting ventricu-lar ERP or the QT interval [89-92,95,96,101-103] (see [55] for a review). By contrast, selective IKr blockers (sotalol, dofetilide, MK-499) and class III anti-arrhythmic drugs (azimilide, ibutilide) prolong RA ERP stronger than LA ERP and pro-long ventricular ERP and the QTc interval [88,89,92,97]. DPO-1 shows selectivity for hKv1.5/IKur over Ito (8-fold), IK1, IKr and IKs (20-fold) [86]. In human atrial myocytes, DPO-1 produces larger prolongation of the APD at the plateau level (APD20) than at 90% of repolarization (120% vs 25% at 0.2 μM), but this effect is not evident in human ventricular myocytes [87]. In addition, in African green monkeys, DPO-1 produces a significant prolongation of AERP without changes in atrial excitation threshold, ventricular refractory period, ECG intervals, heart rate, or blood pressure [104], and in a canine model of atrial flutter, DPO-1 terminates atrial arrhythmia at doses that increase AERP but have no effects on ventricular ERP or PR, QRS or QT intervals of the ECG [102]. Furthermore, in goats and pigs after 48 – 72 h of AF-induced electrical remodeling, IKur blockers (AVE0118, AVE1231) prolong AERP and AF wavelength in a dose-dependent manner, terminate AF and reduce AF inducibility; by contrast, dofetilide prolongs the

AERP in sinus rhythm, but not in remodeled atria [82,90,97,99,100]. These findings suggest that the effects of IKur blockers persist in electrically remodeled atria and support the selective inhibition of IKur as an approach to terminate AF. In isolated human atrial trabeculae from SR patients, 4-AP and AVE0118 shift the plateau to more positive values and increase contractile force [73,105], but this effect is not observed in ventricular trabeculae. Similarly, in goats with chronic AF, AVE0118 fully restores atrial contractility after cardioversion of AF to SR and this effect is accompanied by a significant prolongation of the atrial APD20 [106]. This inotropic effect is abolished by KBR7943, a blocker of the reverse mode Na+/Ca2+-exchange, but is independent of changes in peak ICaL [105]. Thus, the positive inotropic effect is atrial-specific and due to an increase in Ca2+ entry via reverse mode Na+/Ca2+ exchange. It can be hypothetized that this increase in atrial contractility may reduce the risk of thromboembolic events associated with atrial stunning following cardioversion [90,105,106].

6.2 MethodologicalconsiderationsAnimal models represent important tools to analyze the efficacy of new investigational IKur blockers, but because of the differences in the K+ current profiles responsible for atrial repolarization among different species, at this early stage of development caution should be exerted before extrapolating results observed in these models to patients with AF. Kv1.5 is the molecular correlate of IKur in human [8,9,29], dog [31], pig [107] and rat atria [30,108], but the molecular counterpart of IKur and its functional role in atrial repolarization in the rabbit [109], goat [110], guinea pig [55] and African green monkey [88,89] remains to be clarified. Atrial action potential is shorter and the degree of electrical remodeling is more pronounced in goats than in humans, so, in these two species, the relative contribution of IKur and Ito to repolarization and the effect of blockade of these currents on the shape of the action potential may differ [82,100]. Furthermore, atrial cellular electrophysiology in pigs and goats is incompletely explored. IKur is present in rat atrial and ventricular myocytes [7,30] and thus, IKur blockers prolong APD and ERP in both tissues [87,88,92]. Therefore, the rat model provides a measure of in vivo efficacy but no direct measurement of selectivity. Additionally, in rodents, Ito plays an important role in the repolarization [7], so they cannot differentiate Kv4.2/3 from Kv1.5 channel blockers [55]. Dog models are the most widely used to analyze the electrophysiological effects of the new IKur blockers against AF/AFL. Kv1.5 mRNA and protein are the molecular counterparts of IKur in canine atria and ventricle [31,74], so IKur blockers can prolong ventricular APD90 and ERP as as the QTc interval. In fact, DPO-1 prolongs canine atrial and ventricular APD [74,104], whereas it has no effect on the APD in human ventricular myocytes [86]. Another important difference is that, in

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

Tamargo,Caballero,Gómez&Delpón

ExpertOpin.Investig.Drugs(2009) 18(4) 407

contrast to chronic human AF, in a canine model of chronic AF, IKur remains unchanged [111]. Finally, a potential effect of IKur blockade on in vivo canine ventricular repolarization might be obscured by pentobarbital, which blocks several K+ repolarizing currents, and prolongs canine ventricular APD50 and APD90 at concentrations achieved during general anesthesia [112]. This may explain, at least in part, the lack of effect of DPO-1 on canine ventricular repolarization in dogs anesthetized with pentobarbital [102].

6.3 ClinicaltrialsOnly a few new IKur blockers have been studied in clinical trials. Two drugs from the Icagen/Bristol-Myers Squibb program progressed to Phase I trials, but were discontinued [55,113]. AVE0118 was studied in a Phase II trial, but the results have not been published. Because it undergoes rapid first-pass hepatic metabolism and presents a short half-life in dogs and pigs (0.2 – 0.4 h) [55,97,98], AVE0118 has been replaced by the orally active drug

NAP

NAP

to 9 pA/pF

to 9 pA/pF

Ica, L

Ito

Ikur

Ik

AFAP

AFAPAFAP + Ikur inhibition

AFAP + Ikur inhibition

0

5

-5

5

0

0 200 400 600-5

0

-90

V (

mV

)I (

pA

/pF

)I (

pA

/pF

)

∆APD-60 = 0 ms∆APD-60 = + 25 ms

NAP + Ikur inhibition

NAP + Ikur inhibition

t (ms)

0 200 400 600

t (ms)

Figure3.EffectofIKurinhibition(90%reductioninmaximalconductance)onthenormal(NAP)andatrialfibrillation(AF)-modifiedhumanatrialactionpotentials(AFAP).The top panels show the control action potential morphologies and the action potentials after IKur inhibition. The net change in action potential duration is given.

The middle and lower panels show the evolution of several membrane currents in the absence or presence of IKur. Ito: Transient outward K+ current; IK: Delayed

rectifier K+ current; IKur: Ultrarapid delayed rectifier current; ICaL: L-type calcium current. Taken from Courtemanche etal. [79] (with permission).

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillation

408 ExpertOpin.Investig.Drugs(2009) 18(4)

AVE1231, which is currently in Phase I [55]. A Phase I trial with XEN-D0101 was announced in 2005, but the results of this study have not been reported. AZD7009 was studied in a Phase II study in 122 patients with persistent AF or AFL of 48 h-90 days that were randomized to placebo or intravenous (IV) AZD7009 (3-h infusion). AZD7009 was given in doses to obtain target pseudo-steady-state plasma levels between 0.25 and 2.5 μmol/L after 30 min of infusion [114]. At the three highest doses (1.5, 2 and 2.5 μmol/L), AZD7009 converted AF to SR in 45, 64 and 70% of patients after a mean time of 62, 55 and 26 min, respectively, whereas no placebo-treated patients converted. Conversion to SR occurred in 78% of patients with AF ≤ 1 week and in 52% of those with a duration > 1 week; SR was maintained for 24 h in 21 of 22 patients. At plasma concentrations ≥ 0.75 μmol/L, AZD7009 increased the QTc interval (40 ms to 80 ms) and a number of outliers with QTc values > 550 ms were seen, particularly after conversion to SR. Clinical adverse effects (dizziness, bradycardia, hypotension and nausea) were significantly more common in the AZD7009 group than in the placebo group, and one patient exhibited a non-sustained, polymorphic ventricular tachycardia of eight beats with TdP-like features after AZD7009 infusion. The potential of AZD7009 to prolong the QTc interval and possibly cause TdP clearly limits its clinical development.

Vernakalant (RSD1235) is a mixed Na+ (INa and INaL) and atria-preferred K+ channel (IKur, Ito and IKACh) blocker under development for the acute cardioversion of AF to SR [56,65,93-96]. Vernakalant produces a voltage- and frequency-dependent open channel INa block, increasing in depolarized tissues (from an IC50 of 107 μM at -120 mV to 31 μM at -60 mV) and at high heart rates (from an IC50 of 2 μM at 1 Hz, to an IC50 of 0.5 μM at 10 Hz), so under these con-ditions vernakalant cannot be considered as a selective IKur blocker [93-95]. In animal models of AF, the overall rate of conversion with vernakalant is close to 52% within 90 min of infusion (placebo conversion rate 3.8%), but the drug has no effect on PR or QTc intervals [56,115]. The efficacy and safety of oral and IV vernakalant have been analyzed in seven clinical trials. The characteristics of these trials are summarized in Table 2. In the Phase II dose-step Conversion of Rapid Atrial Fibrillation (CRAFT) trial, patients were randomized to vernakalant or placebo [116]. Vernakalant (2 mg/kg followed by 3 mg/kg) shows significant differences over placebo in termination of AF (61 vs 5%), proportion of patients in SR at 30 min (56 vs 5%) and 1 h post-infu-sion (53 vs 5%) and median time to conversion to SR (14 vs 162 min), while no changes in QTc and QRS intervals were observed. In the Atrial Arrhythmia Conversion Trials (ACT) I and III, patients were randomized to vernakalant or placebo. Vernakalant produces a higher conversion rate to SR within 90 min than placebo in patients with AF lasting 3 h to 7 days (52% vs 3.8%) or 3 h to 45 days (37% vs 3%), than in patients in AF for

8 – 45 days (8% vs 0%), the mean time of conversion being 11 min and 8 min in the ACT I and ACT III trials, respectively [115,117-119]. However, in these two trials, less than 8% of the patients with AFL were converted to SR by vernakalant. The ACT II trial compared the efficacy and safety of the IV vernakalant and placebo in patients who developed AF or AFL following coronary artery bypass graft and/or valve replacement surgery [120]. A higher percentage of patients treated with vernakalant converted to SR within 90 min compared with placebo-treated patients; the mean time to conversion was 12 min and 75% converted on the first dose. The open-label ACT IV trial evaluated the safety of vernakalant in patients with symptomatic AF sustained for > 3 h but ≤ 45 days duration [121]. The overall conversion rates to SR for patients with recent-onset AF (< 7 days) was 51% within a median time of 14 min, while conversion rate for those with AF of 8 days or longer was < 10%. These findings confirm that vernakalant is effective and safe to convert recent AF to SR. Intravenous vernakalant has a very short half-life (2.9 – 3.4 h), is metabolized primarily by the cytochrome P450 CYP2D6 and presents a mean steady-state volume of distribution of 88.5 L for extensive metabolizers (111.5 L for poor metabolizers) [115,122]. The most common side effects were dysgeusia, paresthesia, nausea, cough, pruritus, sneezing, bradycardia and hypotension, but no death or cases of TdP or ventricular fibrillation were reported [115,119,122]. In three randomized studies, 39 (5.3%) of 737 patients treated with vernakalant developed any ventricular arrhythmia within 2 h after infusion and further 69 (9.4%) had an event between 2 and 24 h after infusion compared with 20 (6.3%) and 41 (13%) patients in the placebo arm [122]. Discontinuation rate was reported to be 3.8%. However, the safety of vernakalant in patients with heart failure (they were excluded from trials), bradicardia, coronary artery disease or QTc prolongation (> 440 ms) has yet to be determined. An oral formulation of vernakalant is under development for the long-term maintenance of SR after cardioversion. A Phase Ib study confirmed that the drug is orally bioavailable (78%), reaching peak plasma concentration after 1 – 2 h [122]. In a Phase IIa trial in patients with AF lasting 3 – 180 days, vernakalant (300 mg b.i.d.) added to conventional therapy was superior to placebo in maintaining SR postcardioversion (57% vs 39%) [123]. In this study, the most common adverse events were bradycardia and first-degree AV block, but no TdP was observed. An ongoing dose-ranging Phase IIb study analyzes the safety and tolerability, pharmacokinetics and efficacy of vernakalant (150, 300 or 500 mg b.i.d.) in 735 patients at risk of recurrent AF. Patients in AF were electrically cardioverted and then, randomized to vernakalant or placebo for the next 90 days. Very recently, Cardiome announced the results of the interim analysis of 446 patients, indicating that the median time to recurrence of AF was greater than 90 days in patients treated with vernakalant, 500 mg b.i.d., compared with 39 days in those of the placebo group [124].

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

Tamargo,Caballero,Gómez&Delpón

ExpertOpin.Investig.Drugs(2009) 18(4) 409

Tab

le2

.Clin

ical

dev

elo

pm

ent

of

vern

akal

ant.

Stu

dy

[ref

]N

Phas

ed

esig

nPa

tien

tch

arac

teri

stic

sD

ose

(m

g/k

g)

Prim

ary

end

po

int

Ou

tco

me

vs

pla

ceb

oC

om

men

ts

CRA

FT [9

5]56

II, R

, DB,

PC

AF

3-72

h0.

5 +

1 IV

2 +

3 IV

CSR

dur

ing

infu

sion

or

with

in 3

0 m

inC

SR: 6

1 vs

5%

(P <

0.0

1)M

TC: 1

4 vs

162

min

AC

T I [

98]

336

60A

FLIII

P, R

, MC

, PC

AF

3 h-

45 d

3 h

– 7

d, n

=220

8 –

45 d

, n=1

16

3 m

g/kg

for

10

min

plu

s 2

mg/

kg a

fter

15

min

IV

CSR

with

in 9

0 m

in

CSR

: 52

vs 4

%(P

< 0

.01)

Hig

hest

con

vers

ion

A

F ≤

72 h

(78%

)*M

TC: 1

1 m

in

AC

T II

[99]

150

AF

3 -7

2 h

afte

r

card

iac

surg

ery

CSR

: 47

vs 1

4%M

TC: 1

2 m

in

AC

T III

[98]

262

23A

FLA

F 3

h-45

d3

h –

7 d,

n=1

708–

45

d, n

=116

CSR

: 51.

2 vs

3.6

%(P

< 0

.01)

Hig

hest

con

vers

ion

A

F ≤

72 h

(71%

)*M

TC: 8

min

AC

T IV

[100

]16

7III

Ope

nA

F 3

h-45

d3

h –

7 d,

n=1

708

– 45

d, n

=69

CSR

: 51%

MTC

: 14

min

Prev

entio

n [1

01]

221

IIa, P

CPe

rsis

tent

AF

post

card

iove

rsio

nO

ral 3

00 o

r 60

0 b.

i.d.

SR a

t 1

mon

th61

vs

43%

Diff

eren

ce a

t 30

0 m

g b.

i.d.*

Prev

entio

n73

5IIb

, R, D

B, P

CO

ral 1

50, 3

00 o

r 50

0 b.

i.d.

SR a

t 3

mon

ths

52 v

s 39

%(P

< 0

.01)

MTR

on

500

mg

90 v

s 39

day

s

*p <

0.0

5

AFL

: Atr

ial fl

utte

r; C

SR: C

onve

rsio

n to

sin

us r

hyth

m; d

: Day

s; D

B: D

oubl

e-bl

ind;

h: H

ours

; MC

: Mul

ticen

ter;

MTC

: Mea

n tim

e to

con

vers

ion;

MTR

: Mea

n tim

e of

rec

urre

nce;

P: P

rosp

ectiv

e; P

C: P

lace

bo-c

ontr

oled

;

R: R

ando

miz

ed.

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillation

410 ExpertOpin.Investig.Drugs(2009) 18(4)

7. ThefutureofnewIKurblockers(thetwosidesofthecoin)

Data from experimental models and clinical evidence in AF patients confirm that some new IKur blockers prolong atrial APD and AERP and increase atrial contractility at concentrations at which these drugs have almost no effect on intracardiac conduction and ventricular repolarization (QTc interval). Thus, IKur blockers have the potential to become a new therapeutic approach for cardioversion of AF to SR and prevent recurrences. However, most of the new drugs developed as selective IKur blockers display mixed ion channel activity, blocking other cardiac K+ (Ito, IKr and IKAch) Ca2+ and Na+ currents with affinities comparable with that of IKur, i.e., they produce a non-selective IKur blockade (Table 1). Only S9947, ISQ-1, DPO-1 and C9356 might be regarded as selective IKur blockers because their blocking activity on this current clearly dominates the activity on the currents (IKr or IKs) that determine ventricular repolarization [54,55]. The lack of atrial selectivity may explain why the clinical development of some IKur blockers has been withheld due to effects on ventricular repolarization. However, it should be mentioned that the effects of these four drugs on several important ionic currents are unknown (see Table 1), the effects of C9356 have been studied only in HEK-cells and the electrophysiological effects of ISQ-1, DPO-1 and S9947 have been studied in dogs and pigs anesthetized with pentobarbital [88,89,92,104]. Therefore, it is uncertain whether these drugs can be considered as selective IKur blockers. The question is whether the limited selectivity of IKur blockers is beneficial or detrimental in the treatment of AF. Since several different ionic currents contribute to the human atrial AP, agents that affect more than one of these currents would probably be much more effective than a blocker of just one current [54]. Chronic AF leads to an increase of the agonist-independent, constitutively active IKACh channels that may contribute to the shortening of atrial APD [125]. Thus, probably the most attractive combination for the development of atrio-selective drugs might be a mixed blockade of IKur and constitutive IKACh, because both currents are present in atrial cells with little or no activity in ventricular myocytes [4-6]. In some patients AF can be triggered by ectopic beats originating from rapidly firing atrial foci mainly located within or near the pulmonary veins [1]. Furthermore, in certain cases, the maintenance of AF may depend upon the periodic activity of a small number of sustained, high-frequency, functional re-entrant sources (rotors) localized mainly in the LA [126]. A recent simulation study showed that specific blockade of IKur and/or Ito (but not of IKr or IKs) can terminate rotor activity [127]. The underlying mechanism is related to a maximal prolongation of atrial APD at the plateau phase, rather than at the terminal phase of repolarization. This leads to random tip meander and wavebreak, resulting in rotor termination. But despite these suggestive findings, the therapeutic role of IKur blockers

on atrial ectopic and/rotor activity is unknown. Furthermore, important electrical heterogeneities have been found in canine isolated arterially perfused RA, characterized by regional (epicardial vs endocardial) and transmural distribution of repolarization [72], and Ito/IKur block with low concentrations of 4-AP inhomogeneously abbreviates APD leading to a canine perfused RA leading to a reduction in dispersion of repolarization. Interestingly, several IKur blockers (AVE0118, C9356, NIP-142) prolong human APD at the plateau level, while the APD90 values remain unaltered [31,73,86,105] and inhibit Ito (Table 1), the Ito/IKur blocking ratio ranging between 1.6 and 8. Thus, inhibition of Ito may contribute to the atrio-selective effects of IKur blockers and the termination of AF under certain circumstances. Recent evidence suggests the existence of atrial-selective INa blockers that specifically or predominantly depress Na+ channels in atria versus ventricles [128] and it has been proposed that atrial-selective Na+ channel blockers may represent a new strategy in the treatment of AF. There are major differences in action potential characteristics (Figure 1) and biophysical properties of Na+ channels of atrial and ventricular myocytes, so that the voltage-dependence of inactivation is shifted to more negative potentials, the resting membrane potential is less negative and recovery of INa from inactivation is slower in atria than in ventricles [128,129]. Therefore, a larger fraction of atrial versus ventricular Na+ channels would be inactivated at a given take-off potential. At fast stimulation rates (e.g., during AF), the take-off potential is progressively depolarized and decreases the diastolic time for drug unblock from the channels. Both effects increase the fraction of inactivated Na+ channels in the atria, which results in a greater inhibition of INa, and the development of use-dependent postrepolarization refractoriness (PRR) [21,128,130]. Therefore, an IKur blocker with a high affinity for the inactivated state of the Na+ channel may be more effective in blocking atrial than ventricular Na+ channels, leading to an atrial-selective prolongation of AERP at rapid activation rates.

The selectivity of IKur blockers to prolong atrial versus ventricular ERP can be explained by the presence of IKur in atria but not in ventricles. However, in healthy non-remodeled atria, IKur blockers shorten atrial APD90 and prolong AERP; for example, they prolong the ERP/APD90 ratio and PRR. This prolongation of the ERP/APD90 ratio is a typical prop-erty of Na+ channel blockers, but not of K+ channel block-ers, which suggests that the prolongation of AERP can be related to the blockade of INa. Some IKur blockers (AZD7009 and vernakalant) inhibit the INa (Table 1) and slow conduction velocity and increase diastolic threshold of excitation in atria but not in ventricles [56,94,101,131]. In isolated canine atria, AZD7009 concentration-dependently reduces the Vmax (an indirect index of INa) and increases the ERP through APD90 lengthening and PRR [131]. By contrast, the increase in ventricular ERP is small and ventricular conduction time is unaffected by AZD7009 [101,131]. Thus, effects on both repolarization and INa may contribute to the atrial-selective

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

Tamargo,Caballero,Gómez&Delpón

ExpertOpin.Investig.Drugs(2009) 18(4) 411

prolongation of AERP and the anti-arrhythmic effects of AZD7009, particularly at rapid activation rates. A similar mechanism has been proposed to explain the atrial-selective effects of some Na+ channel blockers [128,130] and AVE0118 [61]. From the clinical point of view, only a few investigational IKur blockers have been studied in clinical trials, and there are no comparative studies with class I and III anti-arrhythmic drugs. At present, AVE1231 and XEN-D0101 are in Phase I and only the multichannel blocker vernakalant has progressed to Phase III. Additionally, the pharmacodynamic and pharmacokinetic properties and the long-term safety profile of most of the new IKur blockers are presently unknown. Because Kv1.5 channels are also present in other tissues (i.e., brain, macrophages and vascular, intestinal, and airway smooth muscle), the potential effects of these drugs on non-cardiac tissues should also be carefully addressed. Therefore, at present the clinical efficacy and safety (pro-arrhythmia) and the precise role of IKur blockers to restore and/or maintain SR in patients with AF remains to be defined. However, if selective IKur blockers are found effective and safe in long-term prospective clinical trials, they will represent the first step in the development of atrial-specific anti-arrhythmic drugs that may circumvent the risk of ventricular pro-arrhythmia and the negative inotropism observed with class I and III anti-arrhythmic drugs.

Therefore, the final question is whether a selective IKur blocker can suppress AF. It has been hypothesized [130] that a selective IKur blocker would not be able of effectively suppress AF, because IKur is markedly reduced at fast stimulation rates (e.g., during AF), so that in patients with AF IKur may already be greatly reduced without much left to block [58,59]. However, and even when in remodeled atria, a selective IKur blocker would slightly prolong atrial APD90 and AERP, when IKur block is combined with the inhibition of INa and Ito (and perhaps with IKACh) it can suppress AF/AFL both in experimental models [99,101,130] and in clinical studies [114,116,119]. Unfortunately, we do not have any ‘pure’ IKur blocker, so the hypothesis that a selective IKur blocker can (or cannot) suppress AF cannot be tested.

7.1 CritiquetoIkurblockadeasatherapeuticapproachTwo findings two findings raise the possibility that blocking of IKur can provide the substrate for development of AF. First, loss-of-function mutations of KCNA5 gene located at the N-terminus (A115V), S1 (A251T), S1-S2 linker (P307S, P310L), and C-terminus (P532L and R578K) have been reported in three ethnic groups with idiopathic-lone AF [132,133]. Kv1.5 currents generated by the KCN5A variants are indistinguishable from those generated by wild-type (WT) channels, but C-terminus variants are strikingly less sensitive to blocking by quinidine and propafenone (IC50 = 8.4 μM for WT vs 133 μM for the variants). Thus, mutations may also contribute to variability in drug-induced blockade. Another mutation (E375X) found in a patient with lone AF truncates the Kv1.5 channel protein, eliminating

S4 – S6, the pore region and the C-terminus [134]. Heterologously expressed E375X hKv1.5 channels fail to generate IKur and exert a dominant-negative effect on WT current. Exposure of the proband to isoproterenol induces irregular atrial discharges degenerating into AF. In human atrial myocytes, 4-AP (50 μM) prolongs atrial APD50 and APD90 values, and this effect is associated with early after-depolarizations and triggered activity upon isoproterenol challenge [134]. These findings indicate that defective Kv1.5 channels may predispose the atria to electrical instability and arrhythmogenesis. Second, it is well known that a shortening of AERP can provide the substrate for the development of AF and selective IKur blockers shorten APD90 in healthy human trabeculae [73,79]. Moreover, in isolated canine coronary-perfused right atrial preparations in which AF could not be induced with programmed stimulation, low concentrations of 4-AP (10 – 25 μM, that selectively inhibit IKur) shorten atrial APD90 and AERP, and now non-sustained episodes of AF could be induced in a third of the preparations [60]. However, in acutely remodeled atria (following the administration of acetylcholine or ischemia/reperfusion plus isoproterenol), 4-AP produces almost no prolongation of the APD90, and only at concen-trations ≥ 50 μM it prolongs atrial APD90 and AERP and exerts a mild anti-arrhythmic effect. From this data, it was proposed that blocking of IKur can provide the substrate for the development of AF in healthy canine atria [60]. Even when these results should be interpreted with caution because they were obtained in atrial preparations lacking a number of principal electrical and structural changes observed in chronically remodeled atria, they raise the question of whether IKur blockade may be pro-arrhythmic in healthy atria through the shortening of atrial APD and ERP.

8. Expertopinion

Atrial fibrillation, the most common cardiac arrhythmia, is associated with significant morbidity and mortality and impairment of quality of life. Despite significant advances in the non-pharmacologic treatment of AF, anti-arrhythmic drug therapy remains the mainstay of managing AF for most patients. However, class I and III anti-arrhythmic drugs commonly used to restore and maintain SR exhibit suboptimal efficacy (less than 50% of patients remain in SR 1 year after a successful cardioversion), produce serious adverse effects (including pro-arrhythmia, negative inotropy and extracardiac organ toxicity) that may offset the benefits of SR, and are contraindicated in patients with specific cardiovascular conditions commonly associated with AF, including heart failure, coronary artery disease and structural heart diseases. In an attempt to develop new anti-arrhythmic agents, safer and more effective than those presently used, in the last decade drug discovery efforts have focused on ‘atrial-selective drugs’ that target cardiac ion channel(s) that are exclusively or predominantly expressed in the atria. IKur

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillation

412 ExpertOpin.Investig.Drugs(2009) 18(4)

is a major repolarizing current in human atria, but not in the ventricles, so that blocking of IKur is thought to be a promising target for atrial-specific therapy of AF. Hence, selective inhibition of IKur would be expected to prolong human atrial APD and refractoriness, and to restore and maintain SR with a minimum risk of ventricular pro- arrhythmia. In the last decade, several IKur blockers have been developed and are now being tested in experimental models. However, only a few drugs can be regarded as selective IKur blockers, while the majority of the new compounds block a variety of cardiac K+ channels that determine ven-tricular repolarization as well as Na+ channels with affinities comparable with that of IKur. Moreover, only a few investiga-tional IKur blockers have been studied in clinical trial, and there are no comparative studies with class I and III anti-arrhythmic drugs. At present, AVE1231 and XEN-D0101 are in Phase I

and only the multichannel blocker vernakalant has progressed to Phase III. Therefore, at present, adequately powered, pro-spective, randomized, controlled trials are needed to assess the long-term clinical efficacy and safety of these newly developed agents to restore and/or maintain SR in patients with AF. Nevertheless, these drugs represent the first step in the development of atrial-specific, anti-arrhythmic drugs that may circumvent the risk of ventricular pro-arrhythmia and the negative inotropism observed with the class I and III anti-arrhythmic drugs available at present.

Declarationofinterest

This work was supported by CICYT (SAF2005 – 04609 and SAF2008-04903), FIS-FEDER (PI080665), Red HERACLES (RD06/0009) and Lilly Foundation Grants.

BibliographyPapers of special note have been highlighted as either of interest (•) or of considerable interest (••) to readers.

1. Fuster V, Ryden LE, Cannom DS, et al. ACC/AHA/ESC 2006 Guidelines for the Management of Patients with Atrial Fibrillation. A Report of the American College of Cardiology/American Heart Association Task Force on Practice Guidelines and the European Society of Cardiology Committee for Practice Guidelines developed in collaboration with the European Heart Rhythm Association and the Heart Rhythm Society. Circulation 2006;114:e257-354

2. Wattigney WA, Mensah GA, Croft JB. Increasing trends in hospitalization for atrial fibrillation in the United States, 1985 through 1999: implications for primary prevention. Circulation 2003;108:711-16

3. Nattel S. New ideas about atrial fibrillation 50 years on. Nature 2002;415:219-26

4. Nattel S, Maguy A, Le Bouter S, et al. Arrhythmogenic ion-channel remodeling in the heart: heart failure, myocardial infarction, and atrial fibrillation. Physiol Rev 2007;87:425-56

5. Tamargo J, Caballero R, Delpón E. Pharmacological approaches in the treatment of atrial fibrillation. Curr Med Chem 2004;11:13-28

6. Tamargo J, Caballero R, Gómez R, et al. Pharmacology of cardiac potassium channels. Cardiovasc Res 2004;62:9-33

7. Nerbonne JM, Kass RS. Molecular physiology of cardiac repolarization. Physiol Rev 2005;85:1205-53

8. Fedida D, Wible B, Wang Z, et al. Identity of a novel delayed rectifier current from human heart with a cloned K+ channel current. Circ Res 1993;73:210-16

9. 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;73:1061-06

10. Amor GJ, Wettwer E, Metzger F, et al. Differences between outward currents of human atrial and subepicardial ventricular myocytes. J Physiol 1996;491:31-50

11. Golod DA, Kumar R, Joyner RW. Determinants of action potential initiation in isolated rabbit atrial and ventricular myocytes. Am J Physiol 1998;274:H1902-13

12. Hohnloser SH, Kuck KH, Lilienthal J. Rhythm or rate control in atrial fibrillation–pharmacologic intervention in atrial fibrillation (PIAF): a randomised trial. Lancet 2000;356:1789-94

13. Atrial Fibrillation Follow-up Investigation of Rhythm Management (AFFIRM) Investigators. A comparison of rate control and rhythm control in patients with atrial fibrillation. N Engl J Med 2002;347:1825-33

14. Van Gelder IC, Hagens VE, Bosker HA, et al. for the RACE Investigators. A comparison of rate control and rhythm control in patients with recurrent persistent atrial fibrillation. N Engl J Med 2002;347:1834-40

15. Carlsson J, Miketic S, Windeler J, et al. Randomized trial of rate-control versus

rhythm-control in persistent atrial fibrillation: the Strategies of Treatment of Atrial Fibrillation (STAF) study. J Am Coll Cardiol 2003;41:1690-6

16. Opolski G, Torbicki A, Kosior DA, et al. Rate control vs rhythm control in patients with nonvalvular atrial fibrillation: the results of the Polish How to Treat Chronic Atrial Fibrillation (HOT CAFE) study. Chest 2004;126:476-86

17. de Denus S, Sanoski CA, Carlsson J, et al. Rate vs rhythm control in patients with atrial fibrillation: a meta-analysis. Arch Intern Med 2005;165:258-62

18. Steinberg JS, Sadaniantz A, Kron J, et al. Analysis of cause-specific mortality in the Atrial Fibrillation Follow-up Investigation of Rhythm Management (AFFIRM) study. Circulation 2004;109:1973-80

19. Curtis AB, Gersh BJ, Corley SD, et al.; for the AFFIRM Investigators. Clinical factors that influence response to treatment strategies in atrial fibrillation: the Atrial Fibrillation Follow-up Investigation of Rhythm Management (AFFIRM) study. Am Heart J 2005;149:645-9

20. Roy D, Talajic M, Nattel S, et al. Atrial Fibrillation and Congestive Heart Failure Investigators. Rhythm control versus rate control for atrial fibrillation and heart failure. N Engl J Med 2008;358:2667-77

21. Tamargo J, Valenzuela C, Delpón E. New insights into the pharmacology of sodium channel blockers. Eur Heart J 1992;13(F Suppl):2-13

22. Nattel S, Singh BN. Evolution, Mechanisms, and classification of

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

Tamargo,Caballero,Gómez&Delpón

ExpertOpin.Investig.Drugs(2009) 18(4) 413

antiarrhythmic drugs: focus on class III actions. Am J Cardiol 1999;84:11R-19R

23. Doyle D, Morais Cabral J, Pfuetzner RA, et al. The structure of the potassium channel: molecular basis of K+ conduction and selectivity. Science 1998;280:69-77

24. Jiang Y, Le A, Chen J, et al. X-ray structure of a voltage-dependent K+ channel. Nature 2003;423:33-41

25. Long SB, Campbell EB, Mackinnon R. Crystal structure of a mammalian voltage-dependent Shaker family K+ channel. Science 2005;309:897-903

26. Martens JR, Kwak YG, Tamkun MM. Modulation of Kv channel α/β subunit interactions. Trends Cardiovasc Med 1999;9:253-8

27. Mays DJ, Foose JM, Philipson LH, et al. Localization of the Kv1.5 K+ channel protein in explanted cardiac tissue. J Clin Invest 1995;96:282-92

28. Eldstrom J, Van Wagoner DR, Moore ED, et al. Localization of Kv1.5 channels in rat and canine myocyte sarcolemma. FEBS Lett 2006;580:6039-46

29. Feng J, Wible B, Li GR, et al. Antisense oligonucleotides directed against Kv1.5 mRNA specifically inhibit ultrarapid delayed rectifier K+ current in cultured adult human atrial myocytes. Circ Res 1997;80:572-9

30. Boyle WA, Nerbonne JM. Two functionally distinct 4-aminopyridine-sensitive outward K+ currents in rat atrial myocytes. J Gen Physiol 1992;100:1041-67

31. Fedida D, Eldstrom J, Hesketh JC, et al. Kv1.5 is an important component of repolarizing K+ current in canine atrial myocytes. Circ Res 2003;93:744-51

32. London B, Guo W, Pan XH, et al. Targeted replacement of KV1.5 in the mouse leads to loss of the 4-aminopyridine-sensitive component of IK,slow and resistance to drug-induced qt prolongation. Circ Res 2001;88:940-6

33. Feng J, Yue L, Wang Z, et al. Ionic mechanisms of regional action potential heterogeneity in the canine right atrium. Circ Res 1998;83:541-51

34. De Biasi M, Wang Z, Accili E, et al. Open channel block of human heart hKv1.5 by the beta-subunit hKvβ1.2. Am J Physiol 1997;272:H2932-41

35. Fedida D, Maruoka ND, Lin S. Modulation of slow inactivation in human cardiac Kv1.5 channels by extra- and

intracellular permeant cations. J Physiol 1999;515:315-29

36. López-Barneo J, Hoshi T, Heinemann SH, et al. Effects of external cations and mutations in the pore region of C-type inactivation Shaker potassium channels. Recept Channels 1993;1:61-71

37. Snyders D, Tamkun MM, Bennett PB. A rapidly activating and slowly inactivating potassium channel cloned from human heart. J Gen Physiol 1993;101:513-43

38. Wang Z, Fermini B, Nattel S. Effects of flecainide, quinidine, and 4-aminopyridine on transient outward and ultrarapid delayed rectifier currents in human atrial myocytes. J Pharmacol Exp Ther 1995;272:184-96

39. Yue L, Feng JL, Wang Z, et al. Effects of ambasilide, quinidine, flecainide and verapamil on ultra-rapid delayed rectifier potassium currents in canine atrial myocytes. Cardiovasc Res 2000;46:151-61

40. Doggrell SA, Hancox JC. Dronedarone: an amiodarone analogue. Expert Opin Investig Drugs 2004;13:415-26

41. Valenzuela C, Delpón E, Tamkun MM, et al. Stereoselective block of a human cardiac potassium channel (Kv1.5) by bupivacaine enantiomers. Biophys J 1995;69:418-27

42. Franqueza L, Longobardo M, Vicente J, et al. Molecular determinants of stereoselective bupivacaine block of hKv1.5 channels. Circ Res 1997;81:1053-64

43. Delpón E, Caballero R, Valenzuela C, et al. Benzocaine enhances and inhibits the K+ current through a human cardiac cloned channel (Kv1.5). Cardiovasc Res 1999;42:510-20

44. Caballero R, Moreno I, González T, et al. Putative binding sites for benzocaine on a human cardiac cloned channel (Kv1.5). Cardiovasc Res 2002;56:104-17

45. Caballero R, Delpón E, Valenzuela C, et al. Losartan and its metabolite E3174 modify cardiac delayed rectifier K+ currents. Circulation 2000;101:1199-205

46. Moreno I, Caballero R, González T, et al. Effects of irbesartan on cloned potassium channels involved in human cardiac repolarization. J Pharmacol Exp Ther 2003;304:862-73

47. Delpón E, Caballero R, Gómez R, et al. Angiotensin II, angiotensin II antagonists and spironolactone and their modulation of

cardiac repolarization. Trends Pharmacol Sci 2005;26:155-61

48. Gómez R, Núñez L, Caballero R, et al. Spironolactone and its main metabolite canrenoic acid block hKv1.5, Kv4.3 and Kv7.1 + minK channels. Br J Pharmacol 2005;146:146-61

49. Valenzuela C, Delpón E, Franqueza L, et al. Class III antiarrhythmic effects of zatebradine. Time-, state-, use-, and voltage-dependent block of hKv1.5 channels. Circulation 1996;94:562-70

50. Delpón E, Valenzuela C, Pérez O, et al. Mechanisms of block of a human cloned potassium channel by the enantiomers of a new bradycardic agent: S-16257-2 and S-16260-2. Br J Pharmacol 1996;117:1293-301

51. Vaquero M, Caballero R, Gómez R, et al. Effects of atorvastatin and simvastatin on atrial plateau currentss. J Mol Cell Cardiol 2007;42:931-45

52. Caballero R, Gómez R, Nuñez L, et al. Diltiazem inhibits hKv1.5 and Kv4.3 currents at therapeutic concentrations. Cardiovasc Res 2004;64:457-66

53. Zhang X, Anderson JW, Fedida D. Characterization of nifedipine block of the human heart delayed rectifier, hKv1.5. J Pharmacol Exp Ther 1997;281:1247-56

54. Brendel J, Peukert S. Blockers of the Kv1.5 channel for the treatment of atrial arrhythmias. Curr Med Chem Cardiovasc Hematol Agents 2003;1:273-87

55. Ford JW, Milnes JT. New drugs targeting the cardiac ultra-rapid delayed-rectifier current (IKur): rationale, pharmacology and evidence for potential therapeutic value. J Cardiovasc Pharmacol 2008;52:105-20

56. Fedida D. Vernakalant (RSD1235): a novel, atrial-selective antifibrillatory agent. Expert Opin Investig Drugs 2007;16:519-32

57. Yeola SW, Rich TC, Uebele VN, et al. Molecular analysis of a binding site for quinidine in a human cardiac delayed rectifier K+ channel. Role of S6 in antiarrhythmic drug binding. Circ Res 1996;78:1105-14

58. Feng J, Xu D, Wang Z, et al. Ultrarapid delayed rectifier current inactivation in human atrial myocytes: properties and consequences. Am J Physiol 1998;275:H1717-25

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillation

414 ExpertOpin.Investig.Drugs(2009) 18(4)

59. Ehrlich JR, Ocholla H, Ziemek D, et al. Characterisation of human cardiac Kv1.5 inhibition by the novel atrial-selective antiarrhythmic compound AVE1231. J Cardiovasc Pharmacol 2008;51:380-7

60. Burashnikov A, Antzelevitch C. Can inhibition of IKur promote atrial fibrillation? Heart Rhythm 2008;5:1304-9

61. Burashnikov A, Antzelevitch C. How do atrial-selective drugs differ from atrial antiarrhythmic drugs currently used in the treatment of atrial fibrillation? J Atrial Fibrillation 2008;1:98-107

62. Yamashita T, Murakawa Y, Hayami N, et al. Short-term effects of rapid pacing on mRNA level of voltage-dependent K+ channels in rat atrium: electrical remodeling in paroxysmal atrial tachycardia. Circulation 2000;101:2007-14

63. Dechner N, Pirard B, Bundis F, et al. Molecular basis for Kv1.5 channel block: conservation of drug binding sites among voltage-gated K+ channels. J Biol Chem 2004;279:394-400

64. Decher N, Kumar P, González T, et al. Binding site of a novel Kv1.5 blocker: a ‘foot in the door’ against atrial fibrillation. Mol Pharmacol 2006;70:1204-11

65. Eldstrom J, Wang Z, Xu H, et al. The molecular basis of high-affinity binding of the antiarrhythmic compound vernakalant (RSD1235) to Kv1.5 channels. Mol Pharmacol 2007;72:1522-34

66. Strutz-Seebohm N, Gutcher I, Decher N, et al. Comparison of potent Kv1.5 potassium channel inhibitors reveals the molecular basis for blocking kinetics and binding mode. Cell Physiol Biochem 2007;20:791-800

67. Honoré E, Barhanin J, Attali B, et al. External blockade of the major cardiac delayed-rectifier K+ channel (Kv1.5) by polyunsaturated fatty acids. Proc Natl Acad Sci USA 1994;91:1937-41

68. Longobardo M, González T, Navarro-Polanco R, et al. Effects of a quaternary bupivacaine derivative on delayed rectifier K+ currents. Br J Pharmacol 2000;130:391-401

69. Yue L, Feng J, Li GR, et al. Characterization of an ultrarapid delayed rectifier potassium channel involved in canine atrial repolarization. J Physiol 1996;496:647-62

70. Nattel S, Matthews C, De Blasio E, et al. Dose-dependence of 4-aminopyridine plasma concentrations and

electrophysiological effects in dogs: potential relevance to ionic mechanisms in vivo. Circulation 2000;101:1179-84

71. Escande D, Loisance D, Planche C, et al. Age-related changes of action potential plateau shape in isolated human atrial fibres. Am J Physiol 1985;249:H843-50

72. Burashnikov A, Mannava S, Antzelevitch C. Transmembrane action potential heterogeneity in the canine isolated arterially-perfused atrium: effect of IKr and Ito/IKur block. Am J Physiol 2004;286:H2393-400

73. Wettwer E, Hala O, Christ T, et al. Role of IKur in controlling action potential shape and contractility in the human atrium: influence of chronic atrial fibrillation. Circulation 2004;110:2299-306

74. Sridhar A, Cunha DN, Lacombe VA, et al. The plateau outward current in canine ventricle, sensitive to 4-aminopyridine, is a constitutive contributor to ventricular repolarization. Br J Pharmacol 2007;152:870-9

75. Workman AJ, Kane KA, Rankin AC. The contribution of ionic currents to changes in refractoriness of human atrial myocytes associated with chronic atrial fibrillation. Cardiovasc Res 2001;52:226-35

76. Bowman WC, Marshall RJ, Rodger IW, et al. Actions of 4-aminopyridine on the cardiovascular systems of anesthetized cats and dogs. Br J Anaesth 1981;53:555-65

77. Van Wagoner DR, Pond AL, McCarthy PM, et al. Outward K+ current densities and Kv1.5 expression are reduced in chronic human atrial fibrillation. Circ Res 1997;80:772-81

78. Bosch RF, Zeng X, Grammer JB, et al. Ionic mechanisms of electrical remodeling in human atrial fibrillation. Cardiovasc Res 1999;44:121-31

79. Courtemanche M, Ramírez RJ, Nattel S. Ionic targets for drug therapy and atrial fibrillation-induced electrical remodeling: insights from a mathematical model. Cardiovasc Res 1999;42:477-89

80. Li GR, Wang HB, Qin GW, et al. Acacetin, a natural flavone, selectively inhibits human atrial repolarization potassium currents and prevents atrial fibrilation in dogs. Circulation 2008;117:2449-457

81. Gögelein H, Brendel J, Steinmeyer K, et al. Effects of the atrial antiarrhythmic drug AVE0118 on cardiac ion channels. Naunyn Schmiedebergs Arch Pharmacol 2004;370:183-92

82. Wirth KJ, Brendel J, Stenmeyer K, et al. In vitro and in vivo effect of the atrial selective antiarrhythmic compound AVE1231. J Cardiovasc Pharmacol 2007;49:197-206

83. Gögelein H, Brendel J, Bleich M, et al. Novel atrial K channel blockers AVE3295 and AVE1231 for prevention of recurrence of atrial fibrillation. Heart Rhythm 2004;1:S103

84. Persson F, Carlsson L, Duker G, et al. Blocking characteristics of hERG, hNav1.5 and hKvLQT1/hminK after administration of the novel antiarrhythmic compound AZD7009. J Cardiovasc Electrophysiol 2005;16:329-41

85. Persson F, Carlsson L, Duker G, et al. Blocking characteristics of hKv1.5 and hKv4.3/hKChIP2.2 after administration of the novel antiarrhythmic compound AZD7009. J Cardiovasc Pharmacol 2005;46:7-17

86. Lagrutta A, Wang J, Fermini B, et al. Novel, potent inhibitors of human Kv1.5 K+ channels and ultrarapidly activating delayed rectifier potassium current. J Pharmacol Exp Ther 2006;317:1054-63

87. Trotter BW, Nanda KK, Kett NR, et al. Design and synthesis of novel isoquinoline-3-nitriles as orally bioavailable Kv1.5 antagonists for the treatment of atrial fibrillation. J Med Chem 2006;49:6954-7

88. Regan CP, Stump GL, Wallace AA, et al. In vivo cardiac electrophysiologic and antiarrhythmic effects of an isoquinoline IKur blocker, ISQ-1, in rat, dog, and nonhuman primate. J Cardiovasc Pharmacol 2007;49:236-45

89. Regan CP, Kiss L, Stump GL, et al. Atrial antifibrillatory effects of structurally distinct IKur blockers 3-[(dimethylamino)methyl]-6-methoxy-2-methyl-4-phenylisoquinolin-1(2H)-one and 2-phenyl-1,1-dipyridin-3-yl-2-pyrrolidin-1-yl-ethanol in dogs with underlying heart failure. J Pharmacol Exp Ther 2008;324:322-30

90. Tanaka H, Hashimoto NA. Multiple ion channel blocker, NIP-142, for the treatment of atrial fibrillation. Cardiovasc Drug Rev 2007;25:342-56

91. Bachmann A, Gutcher I, Kopp K, et al. Characterization of a novel Kv1.5 channel blocker in Xenopus oocytes, CHO cells, human and rat cardiomyocytes.

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

Tamargo,Caballero,Gómez&Delpón

ExpertOpin.Investig.Drugs(2009) 18(4) 415

Naunyn Schmiedeberg’s Arch Pharmacol 2001;364:472-8

92. Knobloch K, Brendel J, Peukert S, et al. Electrophysiological and antiarrhythmic effects of the novel Kur channel blockers, S9947 and S20951, on left vs. right pig atrium in vivo in comparison with the IKr blockers dofetilide, azimilide, d,l-sotalol and ibutilide. Naunyn Schmiedebergs Arch Pharmacol 2002;366:482-7

93. Beatch GN, Lin SP, Hesketh C, et al. Electrophysiological mechanism of RSD1235, a new atrial fibrillation converting drug. Circulation 2003;108:IV-85

94. Fedida D, Orth PM, Chen JY, et al. The mechanism of atrial antiarrhythmic action of RSD1235. J Cardiovasc Electrophysiol 2005;16:1227-38

95. Orth PM, Hesketh JC, Mak CK, et al. RSD1235 blocks late INa and suppresses early afterdepolarizations and torsades de pointes induced by class III agents. Cardiovasc Res 2006;70:486-96

96. Nagasawa H, Fujiki A, FujiKura N, et al. Effects of a novel class III antiarrhythmic agent, NIP-142, on canine atrial fibrillation and flutter. Circ J 2002;66:185-91

97. Wirth KJ, Paehler T, Rosenstein B, et al. Atrial effects of the novel K+-channel-blocker AVE0118 in anesthetized pigs. Cardiovasc Res 2003;60:298-306

98. Oros A, Volders PG, Beekman JD, et al. Atrial-specific drug AVE0118 is free of torsades de pointes in anaesthetized dogs with chronic complete atrioventricular block. Heart Rhythm 2006;3:1339-45

99. Blaauw Y, Gögelein H, Tieleman RG, et al. ‘Early’ class III drugs for the treatment of atrial fibrillation: efficacy and atrial selectivity of AVE0118 in remodelled atria of the goat. Circulation 2004;110:1717-24

100. Linz DK, Afkham F, Itter G, et al. Effect of at arial electrical remodeling on the efficacy of antiarrhythmic drugs: comparison of amiodarone with IKr-and Ito/IKur-blockade in vivo. J Cardiovasc Electrophysiol 2007;18:1313-20

101. Goldstein RN, Khrestian C, Carlsson L, et al. AZD7009: a new antiarrhythmic drug with predominant effects on the atria effectively terminates and prevents reinduction of atrial fibrillation and flutter in the sterile pericarditis model. J Cardiovasc Electrophysiol 2004;15:1444-50

102. Stump GL, Wallace AA, Regan CP, et al. In vivo antiarrhythmic and cardiac electrophysiologic effects of a novel diphenylphosphine oxide IKur blocker (2-isopropyl-5-methylcyclohexyl) diphenylphosphine oxide. J Pharmacol Exp Ther 2005;315:1362-7

103. Rivard L, Shiroshita-Takeshita A, Maltais C, et al. Electrophysiological and atrial antiarrhythmic effects of a novel IKur/Kv1.5 blocker in dogs. Heart Rhythm 2005;2:S180

104. Regan CP, Wallace AA, Cresswell HK, et al. In vivo cardiac electrophysiologic effects of a novel diphenylphosphine oxide IKur blocker, (2-Isopropyl-5-methylcyclohexyl) diphenylphosphine oxide, in rat and nonhuman primate. J Pharmacol Exp Ther 2006;316:727-32

105. Schotten U, De Haan S, Verheule S, et al. Blockade of atrial-specific K+-currents increases atrial but not ventricular contractility by enhancing reverse mode Na+/Ca+-exchange. Cardiovasc Res 2007;73:37-47

106. De Haan S, Greiser M, Harks E, et al. AVE0118, blocker of the transient outward current (Ito) and ultrarapid delayed rectifier current (IKur), fully restores atrial contractility after cardioversion of atrial fibrillation in the goat. Circulation 2006;114:1234-42

107. Ehrlich JR, Hoche C, Coutu P, et al. Properties of a time-dependent potassium current in pig atrium: evidence for a role of Kv1.5 in repolarization. J Pharmacol Exp Ther 2006;319:898-906

108. Eldstrom J, Van Wagoner DR, Moore ED, et al. Localization of Kv1.5 channels in rat and canine myocyte sarcolemma. FEBS Lett 2006;580:6039-46

109. Saski Y, Ishii K, Nunoki K, et al. The voltage-dependent K+ channel (Kv1.5) cloned from rabbit heart and facilitation of inactivation of the delayed rectifier current by the rat β-unit. FEBS Lett 1995;372:20-4

110. Van der Velden HMW, Van der Zee L, Wijffels MC, et al. Atrial fibrillation in the goat induces changes in monophasic action potential and mRNA expression of ion channels involved in repolarization. J Cardiovasc Electrophysiol 2000;11:1262-9

111. Yue L, Feng J, Gaspo R, et al. Ionic remodeling underlying action potential changes in a canine model of atrial fibrillation. Circ Res 1997;81:512-25

112. Nattel S, Wang ZG, Matthews C. Direct electrophysiological actions of pentobarbital at concentrations achieved during general anesthesia. Am J Physiol 1990;259:H1743-51

113. Yang Q, Wang X, Du L, et al. strategies for atrial fibrillation therapy: focusing on IKur potassium channel. Expert Opin Ther Pat 2007;17:1443-56

114. Crijns HJ, Van Gelder IC, Walfridsson H, et al. Safe and effective conversion of persistent atrial fibrillation to sinus rhythm by intravenous AZD7009. Heart Rhythm 2006;3:1321-31

115. Naccarelli GV, Wolbrette DL, Samii S, et al. Vernakalant–a promising therapy for conversion of recent-onset atrial fibrillation. Expert Opin Investig Drugs 2008;17:805-10

116. Roy D, Rowe BH, Stiell IG, et al. A randomized, controlled trial of RSD1235, a novel anti-arrhythmic agent, in the treatment of recent onset atrial fibrillation. J Am Coll Cardiol 2004;44:2355-61

117. Torp-Pedersen C, Roy D, Pratt C, et al. Efficacy and safety of vernakalant (RSD1235 injection) in the treatment of atrial fibrillation: combined analysis of two hase III trials. Eur Heart J 2006;27(Suppl):887

118. Roy D, Pratt C, Camm AJ, et al. RSD1235 effectively converts atrial fibrillation to sinus rhythm independent of background use of oral rate- or rhythm-control medications. Circulation 2006;114(Suppl):II-790

119. Roy D, Pratt CM, Torp-Pedersen C, et al. Vernakalant hydrochloride for rapid conversion of atrial fibrillation: a phase 3, randomized, placebo-controlled trial. Circulation 2008;117:1518-25

120. Kowey PR, Roy D, Pratt CM, et al. Efficacy and safety of vernakalant hydrochloride injection for the treatment of atrial fibrillation after valvular or coronary artery bypass surgery. Circulation 2007;116(Suppl):II-636

121. Stiell I, Roy D, Pratt C, et al. Vernakalant hydrochloride injection (RSD 1235) effectively converts acute atrial fibrillation to sinus rhythm independent of background use of oral rate-or rhythm-control medications. Ann Emerg Med 2006;48(Suppl):46-7

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.

IKur/Kv1.5channelblockersforthetreatmentofatrialfibrillation

416 ExpertOpin.Investig.Drugs(2009) 18(4)

122. Cheng JW. Vernakalant in the management of atrial fibrillation. Ann Pharmacother 2008;42:533-42

123. Torp-Pedersen C, Raev D, Georgiev P, et al. Vernakalant (oral) investigators. P2538: oral vernakalant (RSD1235-SR) prevents recurrence of atrial fibrillation cardioversion. Eur Heart J 2007;28(Suppl):401

124. Available from: www.cardiome.com/wordpress/?p =443

125. Dobrev D, Friedrich A, Voigt N, et al. The G protein-gated potassium current IK,(Ach) is constitutively active in patients with chronic atrial fibrillation. Circulation 2005;112:3697-706

126. Mandapati R, Skanes A, Chen J, et al. Stable microreentrant sources as a mechanism of atrial fibrillation in the isolated sheep heart. Circulation 2000;101:194-9

127. Pandit SV, Berenfeld O, Anumonwo JM, et al. Ionic determinants of functional reentry in a 2-D model of human atrial cells during simulated chronic atrial fibrillation. Biophys J 2005;88:3806-21

128. Burashnikov A, Di Diego JM, Zygmunt AC, et al. Atrium-selective sodium channel block as a strategy for suppression of atrial fibrillation: differences in sodium channel inactivation between atria and ventricles and the role of ranolazine. Circulation 2007;116:1449-57

129. Li GR, Lau CP, Shrier A. Heterogeneity of sodium current in atrial vs epicardial ventricular myocytes of adult guinea pig hearts. J Mol Cell Cardiol 2002;34:1185-94

130. Burashnikov A, Antzelevitch C. Atrial-selective sodium channel blockers: do they exist? J Cardiovasc Pharmacol 2008;52:121-8

131. Carlsson L, Chartier D, Nattel S. Characterization of the in vivo and in vitro electrophysiological effects of the novel antiarrhythmic agent AZD7009 in atrial and ventricular tissue of the dog. J Cardiovasc Pharmacol 2006;47:123-32

132. Simard C, Drolet B, Yang P, et al. Polymorphism screening in the cardiac K+ channel gene KCNA5. Clin Pharmacol Ther 2005;77:138-44

133. Drolet B, Simard C, Mizoue L, Roden DM. Human cardiac potassium channel DNA polymorphism modulates acces to drug-binding site and causes drug resistance. J Clin Invest 2005;115:2209-13

134. Olson TM, Alekseev AE, Liu X, et al. Kv1.5 channelopathy due to KCNA5 loss-of-function mutation causes human atrial fibrillation. Hum Mol Genet 2006;15:2185-91

AffiliationJuan Tamargo†, Ricardo Caballero, Ricardo Gómez & Eva Delpón †Author for correspondenceUniversidad Complutense, School of Medicine, Department of Pharmacology, 28040 Madrid, Spain Tel: +34 91 3941472; Fax: +34 91 3941470; E-mail: [email protected]

Exp

ert O

pin.

Inv

estig

. Dru

gs D

ownl

oade

d fr

om in

form

ahea

lthca

re.c

om b

y C

NIO

on

10/0

6/14

For

pers

onal

use

onl

y.