Skip to main content
Erschienen in: Journal of Interventional Cardiac Electrophysiology 2/2015

Open Access 01.08.2015 | REVIEWS

The genetic basis for inherited forms of sinoatrial dysfunction and atrioventricular node dysfunction

verfasst von: Raffaella Milanesi, Annalisa Bucchi, Mirko Baruscotti

Erschienen in: Journal of Interventional Cardiac Electrophysiology | Ausgabe 2/2015

Abstract

The sinoatrial node (SAN) and the atrioventricular node (AVN) are the anatomical and functional regions of the heart which play critical roles in the generation and conduction of the electrical impulse. Their functions are ensured by peculiar structural cytological properties and specific collections of ion channels. Impairment of SAN and AVN activity is generally acquired,but in some cases familial inheritance has been established and therefore a genetic cause is involved. In recent years, combined efforts of clinical practice and experimental basic science studies have identified and characterized several causative gene mutations associated with the nodal syndromes. Channelopathies, i.e., diseases associated with defective ion channels, remain the major cause of genetically determined nodal arrhythmias; however, it is becoming increasingly evident that mutations in other classes of regulatory and structural proteins also have profound pathophysiological roles. In this review, we will present some aspects of the genetic identification of the molecular mechanism underlying both SAN and AVN dysfunctions with a particular focus on mutations of the Na, pacemaker (HCN), and Ca channels. Genetic defects in regulatory proteins and calcium-handling proteins will be also considered. In conclusion, the identification of the genetic defects associated with familial nodal dysfunction is an essential step for implementing an appropriate therapeutic treatment.

1 Introduction

The sinoatrial node (SAN) and the atrioventricular node (AVN), together with the His–Purkinje fibers, form the cardiac conduction system (CCS). The SAN is the primary rhythm generator of the heart due to the spontaneous electrical activity of its pacemaker cells, while the AVN ensures, after imposing an appropriate frequency-dependent delay, the propagation of the impulse from the atria to the ventricles. At a mature stage, the cells of the CCS, and particularly those of the SAN and AVN, have significant cytological and functional differences when compared to those of the working myocardium: the nodal cells are smaller, have a less-defined sarcomeric organization, and a lower cell-to-cell electrical coupling [1]. Moreover, nodal cells spontaneously generate action potentials due to a specific set of ion channels [2]. Any alteration of the delicate balance between the structural organization and the electrical profile of these cells can thus lead to dysfunctions of impulse generation and conduction. SAN and AVN dysfunctions include a large number of pathological conditions: in most cases they are acquired; but in some cases, they are inherited, and therefore a genetic cause is involved [1, 36]. Genetic transmission of a disease is said to be monogenic when the defect is localized in a single gene and to be multigenic when the mutations are present in two or more genes. Despite this simple distinction, there is ample evidence that the phenotypic expression of gene defects is highly complex since incomplete penetrance and high variability are often present. Incomplete penetrance, for example, occurs when an autosomal dominant monogenic disease trait is present in genetically related siblings and only some of them are clinically affected. The presence of variability instead indicates that affected individuals may present a large phenotypic heterogeneity that spans from mild to severe symptoms.
Genetic diseases are commonly thought to have an early onset in life, but this concept does not always hold true since physiological age-dependent molecular and structural remodeling of a given tissue/organ may magnify the pathological impact of a defective gene in the adult or in the elderly. In this regard, it should be considered that the SAN is composed of a relatively low number of cells, and it is known to undergo substantial structural and electrical modifications during ageing, with a reduction in the nodal area and possibly with an increase of fibrotic tissue [1]. Thus, it is conceivable that these structural and electrical age-dependent modifications could in principle set the stage for the late onset of genetic diseases which are not present in young individuals.
The identification of defective genes causing SAN and AVN dysfunctions is starting to emerge from multiple approaches such as genome-wide association studies, the candidate gene approach, and the whole-genome or whole-exome next-generation sequencing. In addition to these genetic studies carried out on humans, a source of valuable information comes from the use of transgenic animal models.

2 Anatomy and physiology of the sinoatrial and atrioventricular nodes

The SAN is located in the posterior wall of the right atrium, in the intercaval region adjacent to the atrial muscle of the crista terminalis, extending from the superior to near the inferior vena cava, and is organized as a mesh-like structure of sparsely organized myocytes embedded in a dense supporting connective matrix [1, 7] (Fig. 1a). The cellular architecture of the SAN is not uniform since, moving from the center to the periphery, there is a smooth transition from smaller primary pacemaker cells to cells that progressively assume more atrial-like features [1]. Highly specialized central SAN myocytes generate rhythmic pacemaker action potentials that are then orderly conveyed to the rest of the myocardium. It is worth mentioning that when rates are evaluated at the single-cell level, cells from the periphery of the node beat faster than those from the center of the node; however, in in vivo conditions, the cells from the periphery have a slower rate due to the hyperpolarizing influence of the surrounding atrial tissue [8].
The main feature of SAN cell autorhythmicity is the presence between consecutive action potentials of a slow diastolic depolarization (pacemaker phase) which sets the cardiac rate. Among the molecular mechanisms that are at work to support this phase, primary roles are played by the pacemaker f-channels (membrane clock) [2, 9] and by intracellular Ca2+ release (Ca2+-clock) [10]. Molecular studies have shown that specific patterns of mRNA/protein expression can be used as markers of SAN cells. For example, SAN cells are rich in pacemaker “f” (HCN4) channels and Tbx3 factor but lack connexin 43 (Cx43) and atrial natriuretic peptide (ANP) (Fig. 1b, c) [1, 7]. Opposite features are indeed typical of working atrial cells.
The AVN is a complex structure located at the base of the atrial septum at the apex of the triangle of Kock and is the sole anatomical site that allows the electrical continuity (connection) between the atria and the ventricles. The histological organization of the AVN is highly heterogeneous, and this structural complexity (compact node, central fibrous body, penetrating bundle, transitional areas) underlies the existence of two routes for electrical conduction: the fast and the slow pathways [1]. In the presence of pathological atrial/SAN activity, the AVN is endowed with two additional roles: during atrial tachyarrhythmias, it acts as a low-pass filter that protects the ventricles from overwhelming and deadly excitations; during marked sinus bradycardia or arrest, the AVN becomes the pacing unit of the ventricular tissue. Several studies have investigated the different cellular morphologies and action potential shapes of AVN cells. Despite the large heterogeneity, the two most representative cell types of the AVN are those presenting ovoid and rod shapes [11] (Fig. 2a). Both types can be autorhythmic even though the action potential profiles and the underlying pool of ion channels are different. In particular, the action potentials of ovoid cells closely resemble those of sinoatrial cells and thus have a higher pacing rate, while rod cells are more atrial-like and have a slower rate (Fig. 2b). Differences in automaticity are paralleled by the finding that the pacemaker I f current is robustly expressed in ovoid cells and less so in rod cells (Fig. 2c).

3 Pathology and genetics of SAN and AVN dysfunction

The term SAN dysfunction (or sick sinus syndrome, SSS) is commonly used to identify various pathological conditions related to the inability of the SAN to generate heart rates that are appropriate for the physiologic needs of an individual. Various cardiac disorders such as inappropriate sinus bradycardia and tachycardia, sinus arrest, sinus-exit block, alternating periods of bradycardia and tachycardia, and chronotropic incompetence are recognized as manifestations of sinus dysfunction [12, 13]. Since SSS patients often present with exercise intolerance, presyncope, or syncope, a correct diagnosis represents an important indication for pacemaker implantation. Often, SAN dysfunction is the consequence of structural degeneration or remodeling process due to either pathological conditions such as atrial fibrillation, heart failure, and infarction or the ageing process. In addition, idiopathic degeneration of the SAN with familial inheritance has long been established. Of note is the clinical observation that while congenital idiopathic sinus dysfunction often progresses to atrial standstill, this progression is not typical of the acquired forms [12].
AVN dysfunction is a set of diseases which occur when there is a partial or total block of impulse conduction through the AVN: a condition known as atrioventricular block (AVB). Atrioventricular conduction diseases can be classified on the basis of the degree of block and particularly on the characteristics of the ECG PR interval. AVB most often occurs in association with conduction defects in the His–Purkinje system, a condition that is typical of progressive and non-progressive cardiac conduction disease. In rare cases, a congenital progressive AVB can occur in the absence of concomitant additional conduction defects [14].
Starting from the 1990s, the combined efforts of clinical practice and experimental basic science studies have identified and characterized several causative gene mutations associated with nodal syndromes. These studies have clearly established the concept that one mutation can often cause multiple clinical phenotypes (diseases) and conversely that one disease can often be associated with different types of mutations. This concept thus explains the large overlap and divergence of symptoms that can be observed within one individual, within related siblings, and among unrelated individuals carrying the same mutation/s (genotype).
In the following sections, we will discuss the genetic aspects of the SAN and AVN dysfunction considering both human gene mutations and data from transgenic animal models; information derived from the animal models can indeed be helpful in guiding genetic testing in humans.

4 Na+ channels

4.1 Cardiac Nav1.5 channels

The SCN5A gene is located on the short arm of chromosome 3 and encodes the α-subunit of the cardiac sodium channel (Nav1.5). In the working myocardium, Nav1.5 currents ensure the fast depolarization (phase 0) of the action potential, and mutant channels are associated with the Brugada syndrome (BrS) and the LQT3 syndrome [5]. In cardiac SAN cells, the presence of Nav1.5 channels has long been a debated issue. There is now evidence that in most of the animal models investigated, Nav1.5 channels are either absent or scarce in the center of the node, while they are robustly expressed in the periphery where they functionally contribute to impulse conduction [8, 1518]. In human primary SAN cells, protein and mRNA detection experiments confirmed that Nav1.5 channel signal is small or absent (Fig. 3a) [21]; however, Verkerk et al. [22] recorded a Na+ current from single human SAN cells and, based on the biophysical features, proposed an Nav1.5 origin [22]. This hypothesis contrasts with the molecular data provided by Chandler et al. [21]: a possible explanation for this discrepancy is that the human SAN cells used for patch-clamp experiments [22] were peripheral pacemaker cells. In addition, as will be further discussed in a following section of this review, it is possible that at least part of this current could flow through non-Nav1.5 Na+ channel isoforms that are also present in SAN cells.
Despite the lack of Nav1.5 channels in primary pacemaker cells, there is ample evidence that many SCN5A mutations are associated with inherited SAN dysfunction manifestations such as bradycardia and sinus-exit block and that symptomatic and asymptomatic sinus bradycardia is also often observed both in patients affected by Brugada or LQT3 syndrome (Table 1; Fig. 3b, c).
Table 1
Mutations of the Nav1.5 channel (SCN5A) associated with nodal and conduction dysfunctions. The functional aspects of all mutations presented in the table have been investigated in heterologous expression studies with the patch-clamp technique. Symptoms listed were identified in the proband and/or in siblings carrying the mutation. Although several mutations displayed biophysical features compatible with both loss of and gain of function; when possible only the major phenotypic impact is listed
Type of mutation, ↑ G-O-F, ↓ L-O-F
Mutation
Impulse generation and conduction dysfunctions
Additional phenotypes
Ref. #
Q55X
Sinus pauses, I° AVB
BrS, intraventricular conduction delay
[23]
R121W
SSS, PCCD
Atrial flutter, VT
[24]
Compound heterozygosity
W156X
R255W
II° AVB, severe CCD, degenerative changes in the conduction system
VT of unknown origin
[25]
E161K
SB, sinus arrest or exit block, combinations of sinoatrial and atrioventricular conduction disturbances
BrS, paroxysmal atrial tachyarrhythmias
[26]
T187I
SB
BrS
[20]
L212P
Atrial standstill
 
[27, 28]
T220I
SB, AVB
DCM, AF, BBB
[12, 29]
G298S
Progressive AVB
 
[30]
D356N
III° AVB
BrS
[20]
R376C
Sinus block, AVB
AF
[31]
R376H
AV-conduction disturbances
BrS, SD
[31, 32]
V411M
II° AVB
LQT3
[33]
T512I
H558R
I°, II° AVB
Type 2 conduction system, (Purkinje) block
[34]
G514C
Bradycardia (SB or suppressed conduction through atrial tissues in the vicinity of the sinus node), I° AVB
Slow conduction throughout the atria and ventricles
[35]
R878C
SB, sinus pauses, slow SAN conduction, I° AVB block
Slow intraventricular conduction
[36]
↑,↓
A1180V
AVB
AF, DCM
[37]
D1275N
Sinus node dysfunction, SB, AVB
BBB, AF, atrial flutter, DCM, CHF
[28, 29]
P1298L
I° AVB
 
[12]
G1408R
I° AVB, CCD
BrS, prolonged QRS
[12, 38]
Compound heterozygosity
P1298L
G1408R
SSS
 
[12]
W1421X
SSS, SB, I° AVB, CCD
BBB, SD
[28, 39]
1493delK
SB, I°AVB, CCD
SD
[40]
↑,↓
ΔK1500
Sinus pauses, I° AVB
LQT3, BrS, BBB, SD
[41]
delKPQ1505-1507
PCCD, SB
LQT3, BrS
[42]
delQKP1507-1509
SB, sinus arrhythmia
LQT3
[43]
K1578fs/52
SSS, sinus arrest, I° AVB,
BrS, SD
[20]
D1595N
Progressive AVB
 
[30]
Compound heterozygosity
delF1617
R1632H
Bradycardia, atrial standstill
Prolonged QRS and prolonged His-ventricle conduction time
[12]
R1623X
Sinus arrest, I° AVB
BrS
[12, 20]
Compound heterozygosity
R1623X
T220I
Absent P waves, sinus pauses
 
[12]
R1632H
I° AVB
 
[12]
S1710L
I° AVB
Ventricular fibrillation
[44]
V1763M
Fetal bradycardia, postnatal II° AVB
VT, LQT3
[45]
V1777M
II° AVB
LQT3
[46]
E1784K
SB, sinus pauses
LQT3, SD
[47]
D1790G
Sinus arrest
LQT3
[48]
↑,↓
1795insD
SB, sinus pauses
LQT3, BrS, SD
[49, 50]
L1821fs/10
SB, sinus pauses, CCD, II° AVB
VT, BBB, atrial flutter
[51]
↑,↓
R1860Gfs*12
SB, sinus pauses, I° AVB
AF, atrial flutter
[52]
IVS22_2T_C
AVB, PCCD
BBB
[53, 54]
5280delG
I °AVB, CCD
BBB
[53, 55]
gain of function (G-O-F), loss of function (L-O-F), AVB atrioventricular block, BrS Brugada syndrome, SSS sick sinus syndrome, PCCD progressive cardiac conduction disease, CCD cardiac conduction disease, VT ventricular tachycardia, SB sinus bradycardia, DCM dilated cardiomyopathy, AF atrial fibrillation, BBB bundle branch block, SD sudden death, LQT3 type 3 long QT syndrome, CHF congestive heart failure
Sinus bradycardia could occur as a consequence of either a slower diastolic depolarization or an increased duration of the action potential. In vitro characterization of biophysical defects of Na+ channels associated with cardiac rate slowing indicates that loss-of-function mutations (typical of Brugada syndrome) are responsible for a slower diastolic depolarization, while gain-of-function mutations (typical of LQT3 syndrome) are responsible for the longer duration of the action potential due to the presence of a non-inactivating sodium current component [49, 56]. In both cases, the overall cycle length increases.
Butters et al. [57] have investigated the effect of loss-of-function Nav1.5 channel mutations by developing a mathematical model of the electrical activity of both isolated SAN cells and SAN-atrium two-dimensional preparation. This study has confirmed that at the single cell level only peripheral cells are affected, but in the SAN-atrium simulation the model reproduced a slowing of both SAN rate and impulse conduction leading to SAN exit block or arrest. This study thus provides the rationale that supports the observation that mutations of the hNav1.5 channel are associated with clinical features typical of SAN dysfunction. A thorough investigation of the presence and role of mutant Nav1.5 currents in native human SAN myocytes is obviously difficult; therefore, the use of pacemaker cells derived from induced pluripotent stem cell (IPSc) obtained from patients with SAN dysfunction may represent an important future research tool.
To our knowledge, SCN5A mutations associated with pure sinus tachy-arrhythmias have not yet been identified. Whether this is still a missing piece of data or is an additional indication that Nav1.5 currents are not primarily involved with impulse generation is still an open question.
The functional integrity of Na+ channels is also critical for the physiology of the AVN. The first identification of an autosomal dominant SCN5A mutation responsible for AVB was accomplished by investigating the genetic basis of two familial forms of progressive (Lev-Lenegre syndrome) and non-progressive cardiac conduction diseases [53]. After this initial finding, several other mutations have been identified (Table 1). As shown in Table 1, SCN5A mutations associated with AVN dysfunctions have been observed alone or in combination with Brugada and/or LQT3 syndromes. The mechanisms by which loss-of-function and gain-of-function mutations functionally converge toward a common pathological outcome in the AVN (AVB) may be similar to the one described above for SAN bradycardic dysfunction. However, this hypothesis might be an oversimplification given the higher complexity of the physiology of the AVN conduction.
We can thus conclude that the biophysical properties of SCN5A mutant channels associated with nodal dysfunction have been largely identified. However, when the focus is moved from the bench to the bedside, a large overlapping of phenotypes [5], together with incomplete penetrance and variability, often complicate the possibility to adopt a personalized (mutation-tailored) clinical treatment.

4.2 Other Na+ channel isoforms and auxiliary β-subunits

The first identification of a non-Nav1.5 isoform was carried out in neonatal rabbit SAN cells, where a Nav1.1 tetrodotoxin (TTX)-sensitive current was shown to be selectively expressed during the first month after birth and to significantly contribute to the pacemaker activity [15, 58, 59]. Several other studies have further investigated the presence of Nav1.1 and of other isoforms in the adult mammalian SAN, and there is now ample evidence that Nav1.1 (SCN1A), Nav1.2 (SCN2A), Nav1.3 (SCN3A), and Nav1.6 (SCN8A) channels are present in SAN cells, although substantial differences in the expression profiles exist among different species [18, 60, 61].
In the human SAN, the presence of detectable albeit low quantity of Nav1.1, Nav1.2, and Nav1.4 (SCN4A) channels was confirmed by Chandler and colleagues [21]. Based on this evidence, it is therefore possible that the INa current originally recorded in the human SAN [22] may be at least partly associated with these isoforms.
The presence of non-Nav1.5 channel isoforms has also been investigated in the AVN of several species, and positive identification has been accomplished for Nav1.1, Nav1.2, Nav1.3, and Nav1.7 (SCN9A) [60, 62]. In humans, Greener et al. [63] have confirmed the presence of Nav1.1, Nav1.3, Nav1.6, and Nav1.7, but their quantitative expression is about tenfold lower than the Nav1.5 channels.
Currently, there is no evidence of mutations of non-Nav1.5 isoforms associated with SAN or AVN dysfunction in humans. However, it is interesting to consider that in a mouse model, the conditional and cardiac-specific heterozygous knockout of the SCN1A gene causes a reduction of heart rate, an increase in the PR interval, and an increase in the heart rate variability [64]. An additional confirmation of the pathophysiological roles of Nav1.1 channels comes from the evidence that in a rat model of heart failure (HF), a profound decline of Nav1.1 (−60.7 %) and Nav1.6 (−47.4 %) isoforms likely contributes to the impairment of SAN function [61]. It remains an open question whether a remodeling of Nav1.1 channels also occurs in human HF patients where a depressed SAN function is often observed [65].
We can thus suggest that non-cardiac (non-Nav1.5) Na+ channels should also be investigated when looking for the genetic basis of nodal dysfunction.
Na+ channel accessory β-subunits are important modulators of the pore-forming α-subunits, and therefore their functional integrity is also extremely relevant in order to ensure proper Na+ currents. Loss-of-function mutations of the Navβ1 (SCN1B) protein have indeed been reported to be associated with the Brugada syndrome and/or conduction disease [66]. Further indication supporting a pathological role of Navβ1 defects comes from a Navβ1 null mouse which exhibits both cardiac (prolonged QT and bradycardia) and neurological disorders [67].

5 HCN4 channels

Pacemaker f-channels belong to the hyperpolarization-activated cyclic nucleotide-gated (HCN) channel family and have long been recognized for their primary role in generating and modulating the automaticity of SAN cells [9, 68]. HCN4 is the principal isoform (~80 % of the total I f current) expressed in the conductive tissue and in particular in the SAN, and screening analysis performed on arrhythmic patients has identified several HCN4 mutations associated with symptomatic and non-symptomatic alterations of the sinus rhythm (Table 2; Fig. 4a).
Table 2
Mutations of the pacemaker (HCN4) channels associated with nodal dysfunctions. The functional aspects of all mutations presented in the table have been investigated in heterologous expression studies with the patch-clamp technique. Symptoms listed were identified in the proband and/or in siblings carrying the mutation
Type of mutation, ↑ G-O-F, ↓ L-O-F
Mutation
Impulse generation and conduction dysfunctions
Additional phenotypes
Ref. #
A414G
SB
LVNC, AF, atrial standstill
[69]
G480R
SB
 
[70]
Y481H
SB
LVNC, AF, degeneration of the mitral valve; polymorphic ventricular, extrasystoles during exercise
[69]
G482R
SB, I°AVB, impaired chronotropic capacity
LVNC, intermittent ectopic atrial rhythms, mitral valve prolapse, out-of-hospital cardiac arrest
[69, 71]
A485V
SB
Out-of-hospital cardiac arrest during extreme exercise, inducible AF, paroxysmal AF
[72]
K530N
SB, sinus pauses, tachycardia-bradycardia syndrome
Persistent AF
[73]
D553N
SB
QT prolongation; torsade de pointes, cardiac arrest for 40 s followed by polymorphic VT
[74]
573X
SB, chronotropic incompetence during exercise
Intermittent AF
[75]
S672R
SB
 
[76]
695X
SB, episodes of distinctive sinus arrhythmia linked to adrenergic stress
LVNC, susceptibility to atrial and ventricular premature beats, mitral valve prolapse
[71, 77]
SB sinus bradycardia, LVNC left ventricular noncompaction cardiomyopathy, AF atrial fibrillation, AVB atrioventricular block, VT ventricular tachycardia
All the mutations identified thus far are of the loss-of-function type, and their resulting phenotypes are compatible with a reduced contribution of the pacemaker current to the diastolic depolarization of SAN cells in resting conditions.
Recently, it has been demonstrated that selective knockout of HCN4 channels in adult transgenic mouse causes sinus bradycardia and AVB that progresses from PQ prolongation to complete heart block (Fig. 4b) [78]. While the sinus bradycardia was obviously predictable, the presence of the AVB was completely unexpected. The evidence that a reduction of the pacemaker current (of about 75 % both in SAN and AVN cells) interferes with the conduction of the impulse through the AVN is not of immediate interpretation. Indeed these data challenge the current view that the sole role of the HCN4 current is to ensure subsidiary pacemaker AVN activity when the primary impulse from the SAN is pathologically impaired. Although it is only speculative, a possible interpretation of the knockout data is that proper SAN impulse propagation throughout the AVN requires AVN cells to be partially depolarized by their own spontaneous diastolic phase; if this spontaneous background depolarization is altered, then also impulse propagation through the AVN is impaired.
Of extreme clinical interest is the association of atrial fibrillation (AF) with the presence of spontaneous arrhythmogenic activity localized in proximal part of the pulmonary veins (PVs) near the left atrium. The identification of this region as an important therapeutic target for the treatment of AF has guided the seminal work of several groups which have demonstrated that focal radioablation of the tissue around the PVs often successfully resolves AF. Investigation of the cytological organization of the pulmonary sleeves in tissue samples from human AF patients has confirmed the presence of cells similar to those of the conduction tissue [79]. This information, together with the established evidence that in small mammals the I f current is highly expressed in this region, raises the question whether familial forms of AF arising from the pulmonary vein sleeves are related to a pathological alteration of the I f expression [80].

6 Ca2+ channels and Ca2+ handling proteins

6.1 Ca2+ channels

Studies on animal models have shown that voltage-dependent L- and T-type Ca2+ currents and intracellular Ca2+ oscillation provide an important contribution to support the pacemaker activity of sinoatrial node cells. In central nodal cells, Ca2+ currents (L-type: Cav1.2 and Cav1.3 channels; T-type: Cav3.1 channels) indeed provide a depolarizing force that both supports the second part of the diastolic depolarization and ensures the upstroke of the action potential [81, 82].
Mangoni et al. have shown that transgenic Cav3.1−/− mice, lacking the T-type calcium currents, present with a moderate reduction of the intrinsic heart rate and have a first-degree AVB [83]. No evidence of mutations affecting Cav3.1 channels has been reported in humans. However, a possible pathological association between loss-of-function of T-type channels and bradycardia and atrioventricular block is found in pediatric patients affected by congenital heart block [84] caused by IgG-induced inhibition of calcium T- and L-type channels in children of mothers affected by an autoimmune disease of rheumatic origin [85].
In 2011, the first homozygous loss-of-function mutation (1208_1209insGGG) of the human L-type Cav1.3 channel was identified in patients affected by the following symptoms: severe to profound deafness, resting bradycardia (32–52 bpm), increased heart rate variability, altered atrioventricular conduction, and junctional escape rhythm; SAN arrest and exit block were also observed [86] (Fig. 5). This complex disorder was termed SAN dysfunction and deafness syndrome, and the presence of cardiac as well as hearing impairment is reminiscent of the phenotype observed in Cav1.3 knockout mice which also show SAN and AVN dysfunction as well as deafness [87, 88].

6.2 Ca2+ handling proteins

In working myocytes, the intracellular control of the Ca2+ concentration is a fundamental aspect of the excitation–contraction coupling process. Intracellular Ca2+ oscillations also represent an important mechanism that, together with plasma membrane ion currents, governs the automaticity of spontaneously beating SAN cells. It is therefore expected that mutations of Ca2+ handling proteins may have important pathological consequences both in working myocytes and in myocytes of the conduction tissue.
Calsequestrin (CASQ2) is a sarcoplasmic Ca2+-binding protein, and ryanodine receptor type 2 (RYR2) is a calcium channel of the sarcoplasmic membrane. Loss-of-function mutations of CASQ2 and RyR2 have been associated with catecholaminergic polymorphic ventricular tachycardia (CPVT), and CPVT patients often present with SAN bradycardia [89, 90].
The Na+/Ca2+ exchanger (NCX1) is an important effector of the Ca2+-clock mechanism since it provides a depolarizing current that contributes to the pacemaker activity of SAN cells [10]. At present, there are no reports on mutations of NCX1 associated with nodal dysfunction; however, recent studies in mice have shown that partial ablation (70–80 % knockout) of NCX1 does not affect both in vivo and ex vivo basal heart rates but severely impairs the β-adrenergic-induced response, while its complete knockout fully eliminates the SAN pacemaker activity [91, 92]. Given the substantial difference between the two phenotypes, it can be hypothesized that in the full knockout model, the severe impairment of SAN cell function can be associated both to the lack of electrogenic activity of the NCX1 pump and to substantial impairment of multiple additional pathways associated with the altered Ca2+ homeostasis.

7 Popdc proteins and twin-pore potassium (K2P) channels

The Popeye domain-containing gene (Popdc) family is a class of membrane proteins largely present in the heart with an evolutionary conserved structural element (the Popeye domain), which binds to the cellular second messenger cAMP. Interestingly, there is a structural similarity between the Popeye domain and the cyclic nucleotide binding domains of the cardiac pacemaker (HCN) channels and of the cAMP-dependent protein kinase (PKA) [93]. In mice, Popdc1 and Popdc2 isoforms are highly expressed in the entire cardiac conduction system, and their knockout induces the appearance of severe sinus node dysfunction, but only in the presence of stress-induced conditions and in an age-dependent manner [94]. The histological analysis of these arrhythmic mutant mice demonstrates a loss of HCN4 positive pacemaker cells [94]. Also, functional knockout of Popdc2 in the zebrafish model also induces severe baseline brady-arrhythmias and AVB [95].
One hypothesis that can explain how the Popdc pathway modulates the cardiac automaticity is based on the observation that Popdc proteins are functional modulators of TREK-1 channels which are members of the twin-pore potassium ion (K2P) channel family [94], and K2P channels are known to control the excitability, the resting membrane potential, and the repolarization phase of cardiomyocytes [96]. Although TREK1 channels appear to be absent from the human SAN, other members of this family such as TWIK1 and TASK1 have been identified [21]. Interestingly, studies of SAN remodelling associated with heart failure (HF) in a rat model have shown that TWIK1, TWIK2, and TASK1 are upregulated during HF and could thus represent a mechanism contributing to the decreased intrinsic heart rate associated with HF [97]. Also indicative is the finding that in Drosophila a K2P channel (ORK1) directly contributes to setting the duration of the diastolic depolarization [98]. Additional proposed mechanisms of the action of Popdc proteins include (a) possible interactions with other ion channels such as HCN or Na channels, (b) a cAMP-dependent role as transcriptional regulators since Popdc proteins have been found also within cell nucleus, and (c) interaction with caveolin-3 since loss of Popdc1 impairs both the caveolar size and the correct trafficking of proteins normally associated with caveolae [99, 100].
Whether these genes are also functionally relevant in the cardiac settings of human nodal diseases remains to be determined.

8 Lamin, emerin, nesprin

Lamin A/C, emerin, and nesprin proteins are part of an intricate protein network of muscle cells that connect the nuclear envelope to the cell nucleo- and cytoskeleton. Interestingly, lamins are considered to be the ancestors of all intermediate filaments. Mutations affecting the correct functions of these genes have been linked to skeletal muscle (Emery–Dreifuss muscular dystrophy) and cardiac diseases including conduction dysfunction [6]. A clear understanding of the functional role and pathological aspects of these proteins is still lacking; however, it may be that the integrity of this protein network is necessary for proper tissue-dependent chromatin dynamics and gene expression [101].

9 Ankyrin-B and caveolin-3

Ankyrins are a class of proteins ubiquitously expressed in excitable and non-excitable cells whose function is to connect integral membrane proteins to the cytoskeleton. Far from being purely passive structural elements, ankyrins contribute to maintaining the proper spatial and functional organization of ion channels, transporters (Na+/K+ exchanger, Na+/K+ ATPase), and Ca2+ handling proteins of the sarcoplasmic reticulum such as inositol 1,4,5, trisphosphate (InsP3) receptor, and ryanodine receptor (RyR2) [102]. It is thus not surprising that loss-of-function mutations of ankyrin-B, an isoform largely expressed in the heart, profoundly alter the proper homeostatic control of Ca2+ and Na+ ions, thus causing a critical substrate for the occurrence of mechano-electrical instability. This condition can then precipitate to a pathological manifestation known as Ankyrin-B syndrome. Symptoms associated with this syndrome include sinus arrhythmias, prolonged QT, torsade de pointes, and atrial fibrillation [102, 103]. Sinus node dysfunction is observed in a relevant fraction (up to 75 %) of individuals affected by the Ankyrin-B syndrome, while QT prolongation is rarer and only occurs in the most severe clinical manifestations [103105]. Of relevance is also the evidence that a distinct ankyrin protein (ANK-G) selectively interacts with Nav1.5 channels (and with Nav.12, Nav1.6), and patients presenting a mutation in the Nav1.5 channel that disrupt this interaction present with Brugada syndrome [102].
Caveolin-3 (CAV3) are integral membrane proteins abundantly present in cardiac and skeletal myocytes where they contribute to the formation of caveolae which are Ω-shaped invaginations of the membrane. Caveolar organization of the membrane favors the co-localization of ion channels, transporters, receptors, and several other proteins in order to ensure optimal spatial and time confinement of local signal transduction. In the heart, CAV3 proteins are virtually present in all myocytes and particularly in SAN cells [106]. Genetic defects of caveolin-3 have been associated both with LQT9 and with sudden infant death syndromes, and the underlying mechanism responsible is the large increase of the late component (up to fivefold) of the Nav1.5 sodium current which is caused by these mutations [107, 108]. Interestingly, patients carrying the T78M mutations also displayed marked sinus bradycardia. The proposed mechanism for this T78M-dependent bradycardia is a modification of the Nav1.5 current which is expressed in peripheral SAN cells. Since Cav3 is also largely present in the center of the SAN, and it functionally interacts with several other ion channels, among which the HCN channels [109], it is also possible that the sinus bradycardia observed in patients carrying the T78M mutation might be caused by the disruption of these additional interactions with other channels.

10 Protein kinase PRKAG2

The 5′ AMP-activated protein kinase (AMPK) is a metabolic sensor that couples cardiac cell physiology to the metabolic state of the heart and is activated by AMP and inhibited by ATP. If the energetic content of the cells becomes too low, the AMPK will counteract this condition by limiting ATP consumption and by activating ATP-producing pathways. Mutations of the PRKAG2 protein are often associated with ventricular pre-excitation (typical of the Wolff–Parkinson–White syndrome) and with AVB and sinus bradycardia [3, 6, 110]. Cardiac hypertrophy of working and conductive myocytes due to an exaggerated glycogen deposition is also typically observed. At present, a convincing molecular explanation for the clinical phenotypes is still missing. While it is possible that the structural modifications due to the hypertrophic state can lead to electrical remodeling and thus to conduction disorders, it cannot be excluded that mutant PRKAG2 proteins may play a role in controlling the expression and modulation of ion channel of conductive cardiomyocytes.

11 Conclusions

The SAN and the AVN together with the His–Purkinje system ensure the appropriate generation and delivery of the electrical impulse to the working myocardium. The physiological functions of these strategic regions can occur only if all the structural and electrical features of these cells are not impaired by acquired or inherited dysfunctions. In recent years, combined efforts of clinical practice and experimental basic science studies have identified and characterized several causative gene mutations associated with the nodal syndromes. One concept that emerges strongly from these studies is that even a single mutation can induce several phenotypical manifestations, and therefore overlapping syndromes are often simultaneously present. The presence of incomplete penetrance and variability often further increases this complexity. Taken together, these concepts explain the clinical observation that “pure nodal dysfunctions”, i.e., dysfunctions of the SAN and AVN in the absence of any additional cardiac disorders, are rare. Despite this complexity, the genetic identification of mutations associated with inherited nodal dysfunctions is becoming a valuable clinical element that may be utilized in order to optimize the clinical therapeutic approach for each patient.

Acknowledgments

The authors declare that they have no conflict of interest.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution License which permits any use, distribution, and reproduction in any medium, provided the original author(s) and the source are credited.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Literatur
1.
Zurück zum Zitat Dobrzynski, H., Anderson, R. H., Atkinson, A., Borbas, Z., D'Souza, A., Fraser, J. F., et al. (2013). Structure, function and clinical relevance of the cardiac conduction system, including the atrioventricular ring and outflow tract tissues. Pharmacology and Therapeutics, 139(2), 260–288.PubMed Dobrzynski, H., Anderson, R. H., Atkinson, A., Borbas, Z., D'Souza, A., Fraser, J. F., et al. (2013). Structure, function and clinical relevance of the cardiac conduction system, including the atrioventricular ring and outflow tract tissues. Pharmacology and Therapeutics, 139(2), 260–288.PubMed
2.
Zurück zum Zitat Baruscotti, M., Barbuti, A., & Bucchi, A. (2010). The cardiac pacemaker current. Journal of Molecular and Cellular Cardiology, 48(1), 55–64.PubMed Baruscotti, M., Barbuti, A., & Bucchi, A. (2010). The cardiac pacemaker current. Journal of Molecular and Cellular Cardiology, 48(1), 55–64.PubMed
3.
Zurück zum Zitat Beinart, R., Ruskin, J., & Milan, D. (2010). The genetics of conduction disease. Heart Failure Clinics, 6(2), 201–214.PubMed Beinart, R., Ruskin, J., & Milan, D. (2010). The genetics of conduction disease. Heart Failure Clinics, 6(2), 201–214.PubMed
4.
Zurück zum Zitat Benson, D. W. (2004). Genetics of atrioventricular conduction disease in humans. The Anatomical Record Part A: Discoveries in Molecular, Cellular, and Evolutionary Biology, 280(2), 934–939. Benson, D. W. (2004). Genetics of atrioventricular conduction disease in humans. The Anatomical Record Part A: Discoveries in Molecular, Cellular, and Evolutionary Biology, 280(2), 934–939.
5.
Zurück zum Zitat Remme, C. A. (2013). Cardiac sodium channelopathy associated with SCN5A mutations: electrophysiological, molecular and genetic aspects. The Journal of Physiology, 591(Pt 17), 4099–4116.PubMedCentralPubMed Remme, C. A. (2013). Cardiac sodium channelopathy associated with SCN5A mutations: electrophysiological, molecular and genetic aspects. The Journal of Physiology, 591(Pt 17), 4099–4116.PubMedCentralPubMed
6.
Zurück zum Zitat Wolf, C. M., & Berul, C. I. (2006). Inherited conduction system abnormalities—one group of diseases, many genes. Journal of Cardiovascular Electrophysiology, 17(4), 446–455.PubMed Wolf, C. M., & Berul, C. I. (2006). Inherited conduction system abnormalities—one group of diseases, many genes. Journal of Cardiovascular Electrophysiology, 17(4), 446–455.PubMed
7.
Zurück zum Zitat Brioschi, C., Micheloni, S., Tellez, J. O., Pisoni, G., Longhi, R., Moroni, P., et al. (2009). Distribution of the pacemaker HCN4 channel mRNA and protein in the rabbit sinoatrial node. Journal of Molecular and Cellular Cardiology, 47(2), 221–227.PubMed Brioschi, C., Micheloni, S., Tellez, J. O., Pisoni, G., Longhi, R., Moroni, P., et al. (2009). Distribution of the pacemaker HCN4 channel mRNA and protein in the rabbit sinoatrial node. Journal of Molecular and Cellular Cardiology, 47(2), 221–227.PubMed
8.
Zurück zum Zitat Honjo, H., Boyett, M. R., Kodama, I., & Toyama, J. (1996). Correlation between electrical activity and the size of rabbit sino-atrial node cells. The Journal of Physiology, 496(Pt 3), 795–808.PubMedCentralPubMed Honjo, H., Boyett, M. R., Kodama, I., & Toyama, J. (1996). Correlation between electrical activity and the size of rabbit sino-atrial node cells. The Journal of Physiology, 496(Pt 3), 795–808.PubMedCentralPubMed
9.
Zurück zum Zitat DiFrancesco, D. (2010). The role of the funny current in pacemaker activity. Circulation Research, 106(3), 434–446.PubMed DiFrancesco, D. (2010). The role of the funny current in pacemaker activity. Circulation Research, 106(3), 434–446.PubMed
10.
Zurück zum Zitat Lakatta, E. G., & DiFrancesco, D. (2009). What keeps us ticking: a funny current, a calcium clock, or both? Journal of Molecular and Cellular Cardiology, 47(2), 157–170.PubMed Lakatta, E. G., & DiFrancesco, D. (2009). What keeps us ticking: a funny current, a calcium clock, or both? Journal of Molecular and Cellular Cardiology, 47(2), 157–170.PubMed
11.
Zurück zum Zitat Munk, A. A., Adjemian, R. A., Zhao, J., Ogbaghebriel, A., & Shrier, A. (1996). Electrophysiological properties of morphologically distinct cells isolated from the rabbit atrioventricular node. The Journal of Physiology, 493(Pt 3), 801–818.PubMedCentralPubMed Munk, A. A., Adjemian, R. A., Zhao, J., Ogbaghebriel, A., & Shrier, A. (1996). Electrophysiological properties of morphologically distinct cells isolated from the rabbit atrioventricular node. The Journal of Physiology, 493(Pt 3), 801–818.PubMedCentralPubMed
12.
Zurück zum Zitat Benson, D. W., Wang, D. W., Dyment, M., Knilans, T. K., Fish, F. A., Strieper, M. J., et al. (2003). Congenital sick sinus syndrome caused by recessive mutations in the cardiac sodium channel gene (SCN5A). The Journal of Clinical Investigation, 112(7), 1019–1028.PubMedCentralPubMed Benson, D. W., Wang, D. W., Dyment, M., Knilans, T. K., Fish, F. A., Strieper, M. J., et al. (2003). Congenital sick sinus syndrome caused by recessive mutations in the cardiac sodium channel gene (SCN5A). The Journal of Clinical Investigation, 112(7), 1019–1028.PubMedCentralPubMed
13.
Zurück zum Zitat Mangrum, J. M., & DiMarco, J. P. (2000). The evaluation and management of bradycardia. The New England Journal of Medicine, 342(10), 703–709.PubMed Mangrum, J. M., & DiMarco, J. P. (2000). The evaluation and management of bradycardia. The New England Journal of Medicine, 342(10), 703–709.PubMed
14.
Zurück zum Zitat Baruteau, A. E., Behaghel, A., Fouchard, S., Mabo, P., Schott, J. J., Dina, C., et al. (2012). Parental electrocardiographic screening identifies a high degree of inheritance for congenital and childhood nonimmune isolated atrioventricular block. Circulation, 126(12), 1469–1477.PubMed Baruteau, A. E., Behaghel, A., Fouchard, S., Mabo, P., Schott, J. J., Dina, C., et al. (2012). Parental electrocardiographic screening identifies a high degree of inheritance for congenital and childhood nonimmune isolated atrioventricular block. Circulation, 126(12), 1469–1477.PubMed
15.
Zurück zum Zitat Baruscotti, M., DiFrancesco, D., & Robinson, R. B. (1996). A TTX-sensitive inward sodium current contributes to spontaneous activity in newborn rabbit sino-atrial node cells. The Journal of Physiology, 492(Pt 1), 21–30.PubMedCentralPubMed Baruscotti, M., DiFrancesco, D., & Robinson, R. B. (1996). A TTX-sensitive inward sodium current contributes to spontaneous activity in newborn rabbit sino-atrial node cells. The Journal of Physiology, 492(Pt 1), 21–30.PubMedCentralPubMed
16.
Zurück zum Zitat Kodama, I., Nikmaram, M. R., Boyett, M. R., Suzuki, R., Honjo, H., & Owen, J. M. (1997). Regional differences in the role of the Ca2+ and Na+ currents in pacemaker activity in the sinoatrial node. The American Journal of Physiology, 272(6 Pt 2), H2793–2806.PubMed Kodama, I., Nikmaram, M. R., Boyett, M. R., Suzuki, R., Honjo, H., & Owen, J. M. (1997). Regional differences in the role of the Ca2+ and Na+ currents in pacemaker activity in the sinoatrial node. The American Journal of Physiology, 272(6 Pt 2), H2793–2806.PubMed
17.
Zurück zum Zitat Remme, C. A., Verkerk, A. O., Hoogaars, W. M., Aanhaanen, W. T., Scicluna, B. P., Annink, C., et al. (2009). The cardiac sodium channel displays differential distribution in the conduction system and transmural heterogeneity in the murine ventricular myocardium. Basic Research in Cardiology, 104(5), 511–522.PubMedCentralPubMed Remme, C. A., Verkerk, A. O., Hoogaars, W. M., Aanhaanen, W. T., Scicluna, B. P., Annink, C., et al. (2009). The cardiac sodium channel displays differential distribution in the conduction system and transmural heterogeneity in the murine ventricular myocardium. Basic Research in Cardiology, 104(5), 511–522.PubMedCentralPubMed
18.
Zurück zum Zitat Maier, S. K., Westenbroek, R. E., Yamanushi, T. T., Dobrzynski, H., Boyett, M. R., Catterall, W. A., et al. (2003). An unexpected requirement for brain-type sodium channels for control of heart rate in the mouse sinoatrial node. Proceedings of the National Academy of Sciences of the United States of America, 100(6), 3507–3512.PubMedCentralPubMed Maier, S. K., Westenbroek, R. E., Yamanushi, T. T., Dobrzynski, H., Boyett, M. R., Catterall, W. A., et al. (2003). An unexpected requirement for brain-type sodium channels for control of heart rate in the mouse sinoatrial node. Proceedings of the National Academy of Sciences of the United States of America, 100(6), 3507–3512.PubMedCentralPubMed
19.
Zurück zum Zitat Dobrzynski, H., Boyett, M. R., & Anderson, R. H. (2007). New insights into pacemaker activity: promoting understanding of sick sinus syndrome. Circulation, 115(14), 1921–1932.PubMed Dobrzynski, H., Boyett, M. R., & Anderson, R. H. (2007). New insights into pacemaker activity: promoting understanding of sick sinus syndrome. Circulation, 115(14), 1921–1932.PubMed
20.
Zurück zum Zitat Makiyama, T., Akao, M., Tsuji, K., Doi, T., Ohno, S., Takenaka, K., et al. (2005). High risk for bradyarrhythmic complications in patients with Brugada syndrome caused by SCN5A gene mutations. Journal of the American College of Cardiology, 46(11), 2100–2106.PubMed Makiyama, T., Akao, M., Tsuji, K., Doi, T., Ohno, S., Takenaka, K., et al. (2005). High risk for bradyarrhythmic complications in patients with Brugada syndrome caused by SCN5A gene mutations. Journal of the American College of Cardiology, 46(11), 2100–2106.PubMed
21.
Zurück zum Zitat Chandler, N. J., Greener, I. D., Tellez, J. O., Inada, S., Musa, H., Molenaar, P., et al. (2009). Molecular architecture of the human sinus node: insights into the function of the cardiac pacemaker. Circulation, 119(12), 1562–1575.PubMed Chandler, N. J., Greener, I. D., Tellez, J. O., Inada, S., Musa, H., Molenaar, P., et al. (2009). Molecular architecture of the human sinus node: insights into the function of the cardiac pacemaker. Circulation, 119(12), 1562–1575.PubMed
22.
Zurück zum Zitat Verkerk, A. O., Wilders, R., van Borren, M. M., & Tan, H. L. (2009). Is sodium current present in human sinoatrial node cells? International Journal of Biological Sciences, 5(2), 201–204.PubMedCentralPubMed Verkerk, A. O., Wilders, R., van Borren, M. M., & Tan, H. L. (2009). Is sodium current present in human sinoatrial node cells? International Journal of Biological Sciences, 5(2), 201–204.PubMedCentralPubMed
23.
Zurück zum Zitat Makita, N., Sumitomo, N., Watanabe, I., & Tsutsui, H. (2007). Novel SCN5A mutation (Q55X) associated with age-dependent expression of Brugada syndrome presenting as neurally mediated syncope. Heart Rhythm, 4(4), 516–519.PubMed Makita, N., Sumitomo, N., Watanabe, I., & Tsutsui, H. (2007). Novel SCN5A mutation (Q55X) associated with age-dependent expression of Brugada syndrome presenting as neurally mediated syncope. Heart Rhythm, 4(4), 516–519.PubMed
24.
Zurück zum Zitat Holst, A. G., Liang, B., Jespersen, T., Bundgaard, H., Haunso, S., Svendsen, J. H., et al. (2010). Sick sinus syndrome, progressive cardiac conduction disease, atrial flutter and ventricular tachycardia caused by a novel SCN5A mutation. Cardiology, 115(4), 311–316.PubMed Holst, A. G., Liang, B., Jespersen, T., Bundgaard, H., Haunso, S., Svendsen, J. H., et al. (2010). Sick sinus syndrome, progressive cardiac conduction disease, atrial flutter and ventricular tachycardia caused by a novel SCN5A mutation. Cardiology, 115(4), 311–316.PubMed
25.
Zurück zum Zitat Bezzina, C. R., Rook, M. B., Groenewegen, W. A., Herfst, L. J., van der Wal, A. C., Lam, J., et al. (2003). Compound heterozygosity for mutations (W156X and R225W) in SCN5A associated with severe cardiac conduction disturbances and degenerative changes in the conduction system. Circulation Research, 92(2), 159–168.PubMed Bezzina, C. R., Rook, M. B., Groenewegen, W. A., Herfst, L. J., van der Wal, A. C., Lam, J., et al. (2003). Compound heterozygosity for mutations (W156X and R225W) in SCN5A associated with severe cardiac conduction disturbances and degenerative changes in the conduction system. Circulation Research, 92(2), 159–168.PubMed
26.
Zurück zum Zitat Smits, J. P., Koopmann, T. T., Wilders, R., Veldkamp, M. W., Opthof, T., Bhuiyan, Z. A., et al. (2005). A mutation in the human cardiac sodium channel (E161K) contributes to sick sinus syndrome, conduction disease and Brugada syndrome in two families. Journal of Molecular and Cellular Cardiology, 38(6), 969–981.PubMed Smits, J. P., Koopmann, T. T., Wilders, R., Veldkamp, M. W., Opthof, T., Bhuiyan, Z. A., et al. (2005). A mutation in the human cardiac sodium channel (E161K) contributes to sick sinus syndrome, conduction disease and Brugada syndrome in two families. Journal of Molecular and Cellular Cardiology, 38(6), 969–981.PubMed
27.
Zurück zum Zitat Makita, N., Sasaki, K., Groenewegen, W. A., Yokota, T., Yokoshiki, H., Murakami, T., et al. (2005). Congenital atrial standstill associated with coinheritance of a novel SCN5A mutation and connexin 40 polymorphisms. Heart Rhythm, 2(10), 1128–1134.PubMed Makita, N., Sasaki, K., Groenewegen, W. A., Yokota, T., Yokoshiki, H., Murakami, T., et al. (2005). Congenital atrial standstill associated with coinheritance of a novel SCN5A mutation and connexin 40 polymorphisms. Heart Rhythm, 2(10), 1128–1134.PubMed
28.
Zurück zum Zitat Gui, J., Wang, T., Jones, R. P., Trump, D., Zimmer, T., & Lei, M. (2010). Multiple loss-of-function mechanisms contribute to SCN5A-related familial sick sinus syndrome. PLoS One, 5(6), e10985.PubMedCentralPubMed Gui, J., Wang, T., Jones, R. P., Trump, D., Zimmer, T., & Lei, M. (2010). Multiple loss-of-function mechanisms contribute to SCN5A-related familial sick sinus syndrome. PLoS One, 5(6), e10985.PubMedCentralPubMed
29.
Zurück zum Zitat Olson, T. M., Michels, V. V., Ballew, J. D., Reyna, S. P., Karst, M. L., Herron, K. J., et al. (2005). Sodium channel mutations and susceptibility to heart failure and atrial fibrillation. JAMA, 293(4), 447–454.PubMedCentralPubMed Olson, T. M., Michels, V. V., Ballew, J. D., Reyna, S. P., Karst, M. L., Herron, K. J., et al. (2005). Sodium channel mutations and susceptibility to heart failure and atrial fibrillation. JAMA, 293(4), 447–454.PubMedCentralPubMed
30.
Zurück zum Zitat Wang, D. W., Viswanathan, P. C., Balser, J. R., George, A. L., Jr., & Benson, D. W. (2002). Clinical, genetic, and biophysical characterization of SCN5A mutations associated with atrioventricular conduction block. Circulation, 105(3), 341–346.PubMed Wang, D. W., Viswanathan, P. C., Balser, J. R., George, A. L., Jr., & Benson, D. W. (2002). Clinical, genetic, and biophysical characterization of SCN5A mutations associated with atrioventricular conduction block. Circulation, 105(3), 341–346.PubMed
31.
Zurück zum Zitat Detta, N., Frisso, G., Limongelli, G., Marzullo, M., Calabro, R., & Salvatore, F. (2014). Genetic analysis in a family affected by sick sinus syndrome may reduce the sudden death risk in a young aspiring competitive athlete. International Journal of Cardiology, 170(3), e63–65.PubMed Detta, N., Frisso, G., Limongelli, G., Marzullo, M., Calabro, R., & Salvatore, F. (2014). Genetic analysis in a family affected by sick sinus syndrome may reduce the sudden death risk in a young aspiring competitive athlete. International Journal of Cardiology, 170(3), e63–65.PubMed
32.
Zurück zum Zitat Rossenbacker, T., Carroll, S. J., Liu, H., Kuiperi, C., de Ravel, T. J., Devriendt, K., et al. (2004). Novel pore mutation in SCN5A manifests as a spectrum of phenotypes ranging from atrial flutter, conduction disease, and Brugada syndrome to sudden cardiac death. Heart Rhythm, 1(5), 610–615.PubMed Rossenbacker, T., Carroll, S. J., Liu, H., Kuiperi, C., de Ravel, T. J., Devriendt, K., et al. (2004). Novel pore mutation in SCN5A manifests as a spectrum of phenotypes ranging from atrial flutter, conduction disease, and Brugada syndrome to sudden cardiac death. Heart Rhythm, 1(5), 610–615.PubMed
33.
Zurück zum Zitat Horne, A. J., Eldstrom, J., Sanatani, S., & Fedida, D. (2011). A novel mechanism for LQT3 with 2:1 block: a pore-lining mutation in Nav1.5 significantly affects voltage-dependence of activation. Heart Rhythm, 8(5), 770–777.PubMed Horne, A. J., Eldstrom, J., Sanatani, S., & Fedida, D. (2011). A novel mechanism for LQT3 with 2:1 block: a pore-lining mutation in Nav1.5 significantly affects voltage-dependence of activation. Heart Rhythm, 8(5), 770–777.PubMed
34.
Zurück zum Zitat Viswanathan, P. C., Benson, D. W., & Balser, J. R. (2003). A common SCN5A polymorphism modulates the biophysical effects of an SCN5A mutation. The Journal of Clinical Investigation, 111(3), 341–346.PubMedCentralPubMed Viswanathan, P. C., Benson, D. W., & Balser, J. R. (2003). A common SCN5A polymorphism modulates the biophysical effects of an SCN5A mutation. The Journal of Clinical Investigation, 111(3), 341–346.PubMedCentralPubMed
35.
Zurück zum Zitat Tan, H. L., Bink-Boelkens, M. T., Bezzina, C. R., Viswanathan, P. C., Beaufort-Krol, G. C., van Tintelen, P. J., et al. (2001). A sodium-channel mutation causes isolated cardiac conduction disease. Nature, 409(6823), 1043–1047.PubMed Tan, H. L., Bink-Boelkens, M. T., Bezzina, C. R., Viswanathan, P. C., Beaufort-Krol, G. C., van Tintelen, P. J., et al. (2001). A sodium-channel mutation causes isolated cardiac conduction disease. Nature, 409(6823), 1043–1047.PubMed
36.
Zurück zum Zitat Zhang, Y., Wang, T., Ma, A., Zhou, X., Gui, J., Wan, H., et al. (2008). Correlations between clinical and physiological consequences of the novel mutation R878C in a highly conserved pore residue in the cardiac Na+ channel. Acta Physiologica (Oxford, England), 194(4), 311–323. Zhang, Y., Wang, T., Ma, A., Zhou, X., Gui, J., Wan, H., et al. (2008). Correlations between clinical and physiological consequences of the novel mutation R878C in a highly conserved pore residue in the cardiac Na+ channel. Acta Physiologica (Oxford, England), 194(4), 311–323.
37.
Zurück zum Zitat Ge, J., Sun, A., Paajanen, V., Wang, S., Su, C., Yang, Z., et al. (2008). Molecular and clinical characterization of a novel SCN5A mutation associated with atrioventricular block and dilated cardiomyopathy. Circulation. Arrhythmia and Electrophysiology, 1(2), 83–92.PubMed Ge, J., Sun, A., Paajanen, V., Wang, S., Su, C., Yang, Z., et al. (2008). Molecular and clinical characterization of a novel SCN5A mutation associated with atrioventricular block and dilated cardiomyopathy. Circulation. Arrhythmia and Electrophysiology, 1(2), 83–92.PubMed
38.
Zurück zum Zitat Kyndt, F., Probst, V., Potet, F., Demolombe, S., Chevallier, J. C., Baro, I., et al. (2001). Novel SCN5A mutation leading either to isolated cardiac conduction defect or Brugada syndrome in a large French family. Circulation, 104(25), 3081–3086.PubMed Kyndt, F., Probst, V., Potet, F., Demolombe, S., Chevallier, J. C., Baro, I., et al. (2001). Novel SCN5A mutation leading either to isolated cardiac conduction defect or Brugada syndrome in a large French family. Circulation, 104(25), 3081–3086.PubMed
39.
Zurück zum Zitat Niu, D. M., Hwang, B., Hwang, H. W., Wang, N. H., Wu, J. Y., Lee, P. C., et al. (2006). A common SCN5A polymorphism attenuates a severe cardiac phenotype caused by a nonsense SCN5A mutation in a Chinese family with an inherited cardiac conduction defect. Journal of Medical Genetics, 43(10), 817–821.PubMedCentralPubMed Niu, D. M., Hwang, B., Hwang, H. W., Wang, N. H., Wu, J. Y., Lee, P. C., et al. (2006). A common SCN5A polymorphism attenuates a severe cardiac phenotype caused by a nonsense SCN5A mutation in a Chinese family with an inherited cardiac conduction defect. Journal of Medical Genetics, 43(10), 817–821.PubMedCentralPubMed
40.
Zurück zum Zitat Zumhagen, S., Veldkamp, M. W., Stallmeyer, B., Baartscheer, A., Eckardt, L., Paul, M., et al. (2013). A heterozygous deletion mutation in the cardiac sodium channel gene SCN5A with loss- and gain-of-function characteristics manifests as isolated conduction disease, without signs of Brugada or long QT syndrome. PLoS One, 8(6), e67963.PubMedCentralPubMed Zumhagen, S., Veldkamp, M. W., Stallmeyer, B., Baartscheer, A., Eckardt, L., Paul, M., et al. (2013). A heterozygous deletion mutation in the cardiac sodium channel gene SCN5A with loss- and gain-of-function characteristics manifests as isolated conduction disease, without signs of Brugada or long QT syndrome. PLoS One, 8(6), e67963.PubMedCentralPubMed
41.
Zurück zum Zitat Grant, A. O., Carboni, M. P., Neplioueva, V., Starmer, C. F., Memmi, M., Napolitano, C., et al. (2002). Long QT syndrome, Brugada syndrome, and conduction system disease are linked to a single sodium channel mutation. The Journal of Clinical Investigation, 110(8), 1201–1209.PubMedCentralPubMed Grant, A. O., Carboni, M. P., Neplioueva, V., Starmer, C. F., Memmi, M., Napolitano, C., et al. (2002). Long QT syndrome, Brugada syndrome, and conduction system disease are linked to a single sodium channel mutation. The Journal of Clinical Investigation, 110(8), 1201–1209.PubMedCentralPubMed
42.
Zurück zum Zitat Bennett, P. B., Yazawa, K., Makita, N., & George, A. L., Jr. (1995). Molecular mechanism for an inherited cardiac arrhythmia. Nature, 376(6542), 683–685.PubMed Bennett, P. B., Yazawa, K., Makita, N., & George, A. L., Jr. (1995). Molecular mechanism for an inherited cardiac arrhythmia. Nature, 376(6542), 683–685.PubMed
43.
Zurück zum Zitat Keller, D. I., Acharfi, S., Delacretaz, E., Benammar, N., Rotter, M., Pfammatter, J. P., et al. (2003). A novel mutation in SCN5A, delQKP 1507–1509, causing long QT syndrome: role of Q1507 residue in sodium channel inactivation. Journal of Molecular and Cellular Cardiology, 35(12), 1513–1521.PubMed Keller, D. I., Acharfi, S., Delacretaz, E., Benammar, N., Rotter, M., Pfammatter, J. P., et al. (2003). A novel mutation in SCN5A, delQKP 1507–1509, causing long QT syndrome: role of Q1507 residue in sodium channel inactivation. Journal of Molecular and Cellular Cardiology, 35(12), 1513–1521.PubMed
44.
Zurück zum Zitat Akai, J., Makita, N., Sakurada, H., Shirai, N., Ueda, K., Kitabatake, A., et al. (2000). A novel SCN5A mutation associated with idiopathic ventricular fibrillation without typical ECG findings of Brugada syndrome. FEBS Letters, 479(1–2), 29–34.PubMed Akai, J., Makita, N., Sakurada, H., Shirai, N., Ueda, K., Kitabatake, A., et al. (2000). A novel SCN5A mutation associated with idiopathic ventricular fibrillation without typical ECG findings of Brugada syndrome. FEBS Letters, 479(1–2), 29–34.PubMed
45.
Zurück zum Zitat Chang, C. C., Acharfi, S., Wu, M. H., Chiang, F. T., Wang, J. K., Sung, T. C., et al. (2004). A novel SCN5A mutation manifests as a malignant form of long QT syndrome with perinatal onset of tachycardia/bradycardia. Cardiovascular Research, 64(2), 268–278.PubMed Chang, C. C., Acharfi, S., Wu, M. H., Chiang, F. T., Wang, J. K., Sung, T. C., et al. (2004). A novel SCN5A mutation manifests as a malignant form of long QT syndrome with perinatal onset of tachycardia/bradycardia. Cardiovascular Research, 64(2), 268–278.PubMed
46.
Zurück zum Zitat Lupoglazoff, J. M., Cheav, T., Baroudi, G., Berthet, M., Denjoy, I., Cauchemez, B., et al. (2001). Homozygous SCN5A mutation in long-QT syndrome with functional two-to-one atrioventricular block. Circulation Research, 89(2), E16–21.PubMed Lupoglazoff, J. M., Cheav, T., Baroudi, G., Berthet, M., Denjoy, I., Cauchemez, B., et al. (2001). Homozygous SCN5A mutation in long-QT syndrome with functional two-to-one atrioventricular block. Circulation Research, 89(2), E16–21.PubMed
47.
Zurück zum Zitat Wei, J., Wang, D. W., Alings, M., Fish, F., Wathen, M., Roden, D. M., et al. (1999). Congenital long-QT syndrome caused by a novel mutation in a conserved acidic domain of the cardiac Na+ channel. Circulation, 99(24), 3165–3171.PubMed Wei, J., Wang, D. W., Alings, M., Fish, F., Wathen, M., Roden, D. M., et al. (1999). Congenital long-QT syndrome caused by a novel mutation in a conserved acidic domain of the cardiac Na+ channel. Circulation, 99(24), 3165–3171.PubMed
48.
Zurück zum Zitat An, R. H., Wang, X. L., Kerem, B., Benhorin, J., Medina, A., Goldmit, M., et al. (1998). Novel LQT-3 mutation affects Na+ channel activity through interactions between alpha- and beta1-subunits. Circulation Research, 83(2), 141–146.PubMed An, R. H., Wang, X. L., Kerem, B., Benhorin, J., Medina, A., Goldmit, M., et al. (1998). Novel LQT-3 mutation affects Na+ channel activity through interactions between alpha- and beta1-subunits. Circulation Research, 83(2), 141–146.PubMed
49.
Zurück zum Zitat Veldkamp, M. W., Wilders, R., Baartscheer, A., Zegers, J. G., Bezzina, C. R., & Wilde, A. A. (2003). Contribution of sodium channel mutations to bradycardia and sinus node dysfunction in LQT3 families. Circulation Research, 92(9), 976–983.PubMed Veldkamp, M. W., Wilders, R., Baartscheer, A., Zegers, J. G., Bezzina, C. R., & Wilde, A. A. (2003). Contribution of sodium channel mutations to bradycardia and sinus node dysfunction in LQT3 families. Circulation Research, 92(9), 976–983.PubMed
50.
Zurück zum Zitat Bezzina, C., Veldkamp, M. W., van Den Berg, M. P., Postma, A. V., Rook, M. B., Viersma, J. W., et al. (1999). A single Na(+) channel mutation causing both long-QT and Brugada syndromes. Circulation Research, 85(12), 1206–1213.PubMed Bezzina, C., Veldkamp, M. W., van Den Berg, M. P., Postma, A. V., Rook, M. B., Viersma, J. W., et al. (1999). A single Na(+) channel mutation causing both long-QT and Brugada syndromes. Circulation Research, 85(12), 1206–1213.PubMed
51.
Zurück zum Zitat Tan, B. H., Iturralde-Torres, P., Medeiros-Domingo, A., Nava, S., Tester, D. J., Valdivia, C. R., et al. (2007). A novel C-terminal truncation SCN5A mutation from a patient with sick sinus syndrome, conduction disorder and ventricular tachycardia. Cardiovascular Research, 76(3), 409–417.PubMedCentralPubMed Tan, B. H., Iturralde-Torres, P., Medeiros-Domingo, A., Nava, S., Tester, D. J., Valdivia, C. R., et al. (2007). A novel C-terminal truncation SCN5A mutation from a patient with sick sinus syndrome, conduction disorder and ventricular tachycardia. Cardiovascular Research, 76(3), 409–417.PubMedCentralPubMed
52.
Zurück zum Zitat Ziyadeh-Isleem, A., Clatot, J., Duchatelet, S., Gandjbakhch, E., Denjoy, I., Hidden-Lucet, F., et al. (2014). A truncating SCN5A mutation combined with genetic variability causes sick sinus syndrome and early atrial fibrillation. Heart Rhythm, 11(6), 1015–1023.PubMedCentralPubMed Ziyadeh-Isleem, A., Clatot, J., Duchatelet, S., Gandjbakhch, E., Denjoy, I., Hidden-Lucet, F., et al. (2014). A truncating SCN5A mutation combined with genetic variability causes sick sinus syndrome and early atrial fibrillation. Heart Rhythm, 11(6), 1015–1023.PubMedCentralPubMed
53.
Zurück zum Zitat Schott, J. J., Alshinawi, C., Kyndt, F., Probst, V., Hoorntje, T. M., Hulsbeek, M., et al. (1999). Cardiac conduction defects associate with mutations in SCN5A. Nature Genetics, 23(1), 20–21.PubMed Schott, J. J., Alshinawi, C., Kyndt, F., Probst, V., Hoorntje, T. M., Hulsbeek, M., et al. (1999). Cardiac conduction defects associate with mutations in SCN5A. Nature Genetics, 23(1), 20–21.PubMed
54.
Zurück zum Zitat Probst, V., Kyndt, F., Potet, F., Trochu, J. N., Mialet, G., Demolombe, S., et al. (2003). Haploinsufficiency in combination with aging causes SCN5A-linked hereditary Lenegre disease. Journal of the American College of Cardiology, 41(4), 643–652.PubMed Probst, V., Kyndt, F., Potet, F., Trochu, J. N., Mialet, G., Demolombe, S., et al. (2003). Haploinsufficiency in combination with aging causes SCN5A-linked hereditary Lenegre disease. Journal of the American College of Cardiology, 41(4), 643–652.PubMed
55.
Zurück zum Zitat Herfst, L. J., Potet, F., Bezzina, C. R., Groenewegen, W. A., Le Marec, H., Hoorntje, T. M., et al. (2003). Na+ channel mutation leading to loss of function and non-progressive cardiac conduction defects. Journal of Molecular and Cellular Cardiology, 35(5), 549–557.PubMed Herfst, L. J., Potet, F., Bezzina, C. R., Groenewegen, W. A., Le Marec, H., Hoorntje, T. M., et al. (2003). Na+ channel mutation leading to loss of function and non-progressive cardiac conduction defects. Journal of Molecular and Cellular Cardiology, 35(5), 549–557.PubMed
56.
Zurück zum Zitat Lei, M., Huang, C. L., & Zhang, Y. (2008). Genetic Na+ channelopathies and sinus node dysfunction. Progress in Biophysics and Molecular Biology, 98(2–3), 171–178.PubMed Lei, M., Huang, C. L., & Zhang, Y. (2008). Genetic Na+ channelopathies and sinus node dysfunction. Progress in Biophysics and Molecular Biology, 98(2–3), 171–178.PubMed
57.
Zurück zum Zitat Butters, T. D., Aslanidi, O. V., Inada, S., Boyett, M. R., Hancox, J. C., Lei, M., et al. (2010). Mechanistic links between Na+ channel (SCN5A) mutations and impaired cardiac pacemaking in sick sinus syndrome. Circulation Research, 107(1), 126–137.PubMedCentralPubMed Butters, T. D., Aslanidi, O. V., Inada, S., Boyett, M. R., Hancox, J. C., Lei, M., et al. (2010). Mechanistic links between Na+ channel (SCN5A) mutations and impaired cardiac pacemaking in sick sinus syndrome. Circulation Research, 107(1), 126–137.PubMedCentralPubMed
58.
Zurück zum Zitat Baruscotti, M., Westenbroek, R., Catterall, W. A., DiFrancesco, D., & Robinson, R. B. (1997). The newborn rabbit sino-atrial node expresses a neuronal type I-like Na+ channel. The Journal of Physiology, 498(Pt 3), 641–648.PubMedCentralPubMed Baruscotti, M., Westenbroek, R., Catterall, W. A., DiFrancesco, D., & Robinson, R. B. (1997). The newborn rabbit sino-atrial node expresses a neuronal type I-like Na+ channel. The Journal of Physiology, 498(Pt 3), 641–648.PubMedCentralPubMed
59.
Zurück zum Zitat Baruscotti, M., DiFrancesco, D., & Robinson, R. B. (2000). Na(+) current contribution to the diastolic depolarization in newborn rabbit SA node cells. American Journal of Physiology - Heart and Circulatory Physiology, 279(5), H2303–2309.PubMed Baruscotti, M., DiFrancesco, D., & Robinson, R. B. (2000). Na(+) current contribution to the diastolic depolarization in newborn rabbit SA node cells. American Journal of Physiology - Heart and Circulatory Physiology, 279(5), H2303–2309.PubMed
60.
Zurück zum Zitat Haufe, V., Cordeiro, J. M., Zimmer, T., Wu, Y. S., Schiccitano, S., Benndorf, K., et al. (2005). Contribution of neuronal sodium channels to the cardiac fast sodium current INa is greater in dog heart Purkinje fibers than in ventricles. Cardiovascular Research, 65(1), 117–127.PubMed Haufe, V., Cordeiro, J. M., Zimmer, T., Wu, Y. S., Schiccitano, S., Benndorf, K., et al. (2005). Contribution of neuronal sodium channels to the cardiac fast sodium current INa is greater in dog heart Purkinje fibers than in ventricles. Cardiovascular Research, 65(1), 117–127.PubMed
61.
Zurück zum Zitat Du, Y., Huang, X., Wang, T., Han, K., Zhang, J., Xi, Y., et al. (2007). Downregulation of neuronal sodium channel subunits Nav1.1 and Nav1.6 in the sinoatrial node from volume-overloaded heart failure rat. Pflügers Archiv, 454(3), 451–459.PubMed Du, Y., Huang, X., Wang, T., Han, K., Zhang, J., Xi, Y., et al. (2007). Downregulation of neuronal sodium channel subunits Nav1.1 and Nav1.6 in the sinoatrial node from volume-overloaded heart failure rat. Pflügers Archiv, 454(3), 451–459.PubMed
62.
Zurück zum Zitat Marionneau, C., Couette, B., Liu, J., Li, H., Mangoni, M. E., Nargeot, J., et al. (2005). Specific pattern of ionic channel gene expression associated with pacemaker activity in the mouse heart. The Journal of Physiology, 562(Pt 1), 223–234.PubMedCentralPubMed Marionneau, C., Couette, B., Liu, J., Li, H., Mangoni, M. E., Nargeot, J., et al. (2005). Specific pattern of ionic channel gene expression associated with pacemaker activity in the mouse heart. The Journal of Physiology, 562(Pt 1), 223–234.PubMedCentralPubMed
63.
Zurück zum Zitat Greener, I. D., Monfredi, O., Inada, S., Chandler, N. J., Tellez, J. O., Atkinson, A., et al. (2011). Molecular architecture of the human specialised atrioventricular conduction axis. Journal of Molecular and Cellular Cardiology, 50(4), 642–651.PubMed Greener, I. D., Monfredi, O., Inada, S., Chandler, N. J., Tellez, J. O., Atkinson, A., et al. (2011). Molecular architecture of the human specialised atrioventricular conduction axis. Journal of Molecular and Cellular Cardiology, 50(4), 642–651.PubMed
64.
Zurück zum Zitat Kalume, F., Westenbroek, R. E., Cheah, C. S., Yu, F. H., Oakley, J. C., Scheuer, T., et al. (2013). Sudden unexpected death in a mouse model of Dravet syndrome. The Journal of Clinical Investigation, 123(4), 1798–1808.PubMedCentralPubMed Kalume, F., Westenbroek, R. E., Cheah, C. S., Yu, F. H., Oakley, J. C., Scheuer, T., et al. (2013). Sudden unexpected death in a mouse model of Dravet syndrome. The Journal of Clinical Investigation, 123(4), 1798–1808.PubMedCentralPubMed
65.
Zurück zum Zitat Sanders, P., Kistler, P. M., Morton, J. B., Spence, S. J., & Kalman, J. M. (2004). Remodeling of sinus node function in patients with congestive heart failure: reduction in sinus node reserve. Circulation, 110(8), 897–903.PubMed Sanders, P., Kistler, P. M., Morton, J. B., Spence, S. J., & Kalman, J. M. (2004). Remodeling of sinus node function in patients with congestive heart failure: reduction in sinus node reserve. Circulation, 110(8), 897–903.PubMed
66.
Zurück zum Zitat Watanabe, H., Koopmann, T. T., Le Scouarnec, S., Yang, T., Ingram, C. R., Schott, J. J., et al. (2008). Sodium channel beta1 subunit mutations associated with Brugada syndrome and cardiac conduction disease in humans. The Journal of Clinical Investigation, 118(6), 2260–2268.PubMedCentralPubMed Watanabe, H., Koopmann, T. T., Le Scouarnec, S., Yang, T., Ingram, C. R., Schott, J. J., et al. (2008). Sodium channel beta1 subunit mutations associated with Brugada syndrome and cardiac conduction disease in humans. The Journal of Clinical Investigation, 118(6), 2260–2268.PubMedCentralPubMed
67.
Zurück zum Zitat Lopez-Santiago, L. F., Meadows, L. S., Ernst, S. J., Chen, C., Malhotra, J. D., McEwen, D. P., et al. (2007). Sodium channel Scn1b null mice exhibit prolonged QT and RR intervals. Journal of Molecular and Cellular Cardiology, 43(5), 636–647.PubMedCentralPubMed Lopez-Santiago, L. F., Meadows, L. S., Ernst, S. J., Chen, C., Malhotra, J. D., McEwen, D. P., et al. (2007). Sodium channel Scn1b null mice exhibit prolonged QT and RR intervals. Journal of Molecular and Cellular Cardiology, 43(5), 636–647.PubMedCentralPubMed
68.
Zurück zum Zitat Verkerk, A. O., Wilders, R., van Borren, M. M., Peters, R. J., Broekhuis, E., Lam, K., et al. (2007). Pacemaker current (I(f)) in the human sinoatrial node. European Heart Journal, 28(20), 2472–2478.PubMed Verkerk, A. O., Wilders, R., van Borren, M. M., Peters, R. J., Broekhuis, E., Lam, K., et al. (2007). Pacemaker current (I(f)) in the human sinoatrial node. European Heart Journal, 28(20), 2472–2478.PubMed
69.
Zurück zum Zitat Milano, A., Vermeer, A. M., Lodder, E. M., Barc, J., Verkerk, A. O., Postma, A. V., et al. (2014). HCN4 mutations in multiple families with bradycardia and left ventricular noncompaction cardiomyopathy. Journal of the American College of Cardiology, 64(8), 745–756.PubMed Milano, A., Vermeer, A. M., Lodder, E. M., Barc, J., Verkerk, A. O., Postma, A. V., et al. (2014). HCN4 mutations in multiple families with bradycardia and left ventricular noncompaction cardiomyopathy. Journal of the American College of Cardiology, 64(8), 745–756.PubMed
70.
Zurück zum Zitat Nof, E., Luria, D., Brass, D., Marek, D., Lahat, H., Reznik-Wolf, H., et al. (2007). Point mutation in the HCN4 cardiac ion channel pore affecting synthesis, trafficking, and functional expression is associated with familial asymptomatic sinus bradycardia. Circulation, 116(5), 463–470.PubMed Nof, E., Luria, D., Brass, D., Marek, D., Lahat, H., Reznik-Wolf, H., et al. (2007). Point mutation in the HCN4 cardiac ion channel pore affecting synthesis, trafficking, and functional expression is associated with familial asymptomatic sinus bradycardia. Circulation, 116(5), 463–470.PubMed
71.
Zurück zum Zitat Schweizer, P. A., Schroter, J., Greiner, S., Haas, J., Yampolsky, P., Mereles, D., et al. (2014). The symptom complex of familial sinus node dysfunction and myocardial noncompaction is associated with mutations in the HCN4 channel. Journal of the American College of Cardiology, 64(8), 757–767.PubMed Schweizer, P. A., Schroter, J., Greiner, S., Haas, J., Yampolsky, P., Mereles, D., et al. (2014). The symptom complex of familial sinus node dysfunction and myocardial noncompaction is associated with mutations in the HCN4 channel. Journal of the American College of Cardiology, 64(8), 757–767.PubMed
72.
Zurück zum Zitat Laish-Farkash, A., Glikson, M., Brass, D., Marek-Yagel, D., Pras, E., Dascal, N., et al. (2010). A novel mutation in the HCN4 gene causes symptomatic sinus bradycardia in Moroccan Jews. Journal of Cardiovascular Electrophysiology, 21(12), 1365–1372.PubMedCentralPubMed Laish-Farkash, A., Glikson, M., Brass, D., Marek-Yagel, D., Pras, E., Dascal, N., et al. (2010). A novel mutation in the HCN4 gene causes symptomatic sinus bradycardia in Moroccan Jews. Journal of Cardiovascular Electrophysiology, 21(12), 1365–1372.PubMedCentralPubMed
73.
Zurück zum Zitat Duhme, N., Schweizer, P. A., Thomas, D., Becker, R., Schroter, J., Barends, T. R., et al. (2013). Altered HCN4 channel C-linker interaction is associated with familial tachycardia-bradycardia syndrome and atrial fibrillation. European Heart Journal, 34(35), 2768–2775.PubMed Duhme, N., Schweizer, P. A., Thomas, D., Becker, R., Schroter, J., Barends, T. R., et al. (2013). Altered HCN4 channel C-linker interaction is associated with familial tachycardia-bradycardia syndrome and atrial fibrillation. European Heart Journal, 34(35), 2768–2775.PubMed
74.
Zurück zum Zitat Ueda, K., Nakamura, K., Hayashi, T., Inagaki, N., Takahashi, M., Arimura, T., et al. (2004). Functional characterization of a trafficking-defective HCN4 mutation, D553N, associated with cardiac arrhythmia. The Journal of Biological Chemistry, 279(26), 27194–27198.PubMed Ueda, K., Nakamura, K., Hayashi, T., Inagaki, N., Takahashi, M., Arimura, T., et al. (2004). Functional characterization of a trafficking-defective HCN4 mutation, D553N, associated with cardiac arrhythmia. The Journal of Biological Chemistry, 279(26), 27194–27198.PubMed
75.
Zurück zum Zitat Schulze-Bahr, E., Neu, A., Friederich, P., Kaupp, U. B., Breithardt, G., Pongs, O., et al. (2003). Pacemaker channel dysfunction in a patient with sinus node disease. The Journal of Clinical Investigation, 111(10), 1537–1545.PubMedCentralPubMed Schulze-Bahr, E., Neu, A., Friederich, P., Kaupp, U. B., Breithardt, G., Pongs, O., et al. (2003). Pacemaker channel dysfunction in a patient with sinus node disease. The Journal of Clinical Investigation, 111(10), 1537–1545.PubMedCentralPubMed
76.
Zurück zum Zitat Milanesi, R., Baruscotti, M., Gnecchi-Ruscone, T., & DiFrancesco, D. (2006). Familial sinus bradycardia associated with a mutation in the cardiac pacemaker channel. The New England Journal of Medicine, 354(2), 151–157.PubMed Milanesi, R., Baruscotti, M., Gnecchi-Ruscone, T., & DiFrancesco, D. (2006). Familial sinus bradycardia associated with a mutation in the cardiac pacemaker channel. The New England Journal of Medicine, 354(2), 151–157.PubMed
77.
Zurück zum Zitat Schweizer, P. A., Duhme, N., Thomas, D., Becker, R., Zehelein, J., Draguhn, A., et al. (2010). cAMP sensitivity of HCN pacemaker channels determines basal heart rate but is not critical for autonomic rate control. Circulation. Arrhythmia and Electrophysiology, 3(5), 542–552.PubMed Schweizer, P. A., Duhme, N., Thomas, D., Becker, R., Zehelein, J., Draguhn, A., et al. (2010). cAMP sensitivity of HCN pacemaker channels determines basal heart rate but is not critical for autonomic rate control. Circulation. Arrhythmia and Electrophysiology, 3(5), 542–552.PubMed
78.
Zurück zum Zitat Baruscotti, M., Bucchi, A., Viscomi, C., Mandelli, G., Consalez, G., Gnecchi-Rusconi, T., et al. (2011). Deep bradycardia and heart block caused by inducible cardiac-specific knockout of the pacemaker channel gene Hcn4. Proceedings of the National Academy of Sciences of the United States of America, 108(4), 1705–1710.PubMedCentralPubMed Baruscotti, M., Bucchi, A., Viscomi, C., Mandelli, G., Consalez, G., Gnecchi-Rusconi, T., et al. (2011). Deep bradycardia and heart block caused by inducible cardiac-specific knockout of the pacemaker channel gene Hcn4. Proceedings of the National Academy of Sciences of the United States of America, 108(4), 1705–1710.PubMedCentralPubMed
79.
Zurück zum Zitat Perez-Lugones, A., McMahon, J. T., Ratliff, N. B., Saliba, W. I., Schweikert, R. A., Marrouche, N. F., et al. (2003). Evidence of specialized conduction cells in human pulmonary veins of patients with atrial fibrillation. Journal of Cardiovascular Electrophysiology, 14(8), 803–809.PubMed Perez-Lugones, A., McMahon, J. T., Ratliff, N. B., Saliba, W. I., Schweikert, R. A., Marrouche, N. F., et al. (2003). Evidence of specialized conduction cells in human pulmonary veins of patients with atrial fibrillation. Journal of Cardiovascular Electrophysiology, 14(8), 803–809.PubMed
80.
Zurück zum Zitat Suenari, K., Cheng, C. C., Chen, Y. C., Lin, Y. K., Nakano, Y., Kihara, Y., et al. (2012). Effects of ivabradine on the pulmonary vein electrical activity and modulation of pacemaker currents and calcium homeostasis. Journal of Cardiovascular Electrophysiology, 23(2), 200–206.PubMed Suenari, K., Cheng, C. C., Chen, Y. C., Lin, Y. K., Nakano, Y., Kihara, Y., et al. (2012). Effects of ivabradine on the pulmonary vein electrical activity and modulation of pacemaker currents and calcium homeostasis. Journal of Cardiovascular Electrophysiology, 23(2), 200–206.PubMed
81.
Zurück zum Zitat Mangoni, M. E., & Nargeot, J. (2008). Genesis and regulation of the heart automaticity. Physiological Reviews, 88(3), 919–982.PubMed Mangoni, M. E., & Nargeot, J. (2008). Genesis and regulation of the heart automaticity. Physiological Reviews, 88(3), 919–982.PubMed
82.
Zurück zum Zitat Christel, C. J., Cardona, N., Mesirca, P., Herrmann, S., Hofmann, F., Striessnig, J., et al. (2012). Distinct localization and modulation of Cav1.2 and Cav1.3 L-type Ca2+ channels in mouse sinoatrial node. The Journal of Physiology, 590(Pt 24), 6327–6342.PubMedCentralPubMed Christel, C. J., Cardona, N., Mesirca, P., Herrmann, S., Hofmann, F., Striessnig, J., et al. (2012). Distinct localization and modulation of Cav1.2 and Cav1.3 L-type Ca2+ channels in mouse sinoatrial node. The Journal of Physiology, 590(Pt 24), 6327–6342.PubMedCentralPubMed
83.
Zurück zum Zitat Mangoni, M. E., Traboulsie, A., Leoni, A. L., Couette, B., Marger, L., Le Quang, K., et al. (2006). Bradycardia and slowing of the atrioventricular conduction in mice lacking CaV3.1/alpha1G T-type calcium channels. Circulation Research, 98(11), 1422–1430.PubMed Mangoni, M. E., Traboulsie, A., Leoni, A. L., Couette, B., Marger, L., Le Quang, K., et al. (2006). Bradycardia and slowing of the atrioventricular conduction in mice lacking CaV3.1/alpha1G T-type calcium channels. Circulation Research, 98(11), 1422–1430.PubMed
84.
Zurück zum Zitat Strandberg, L. S., Cui, X., Rath, A., Liu, J., Silverman, E. D., Liu, X., et al. (2013). Congenital heart block maternal sera autoantibodies target an extracellular epitope on the alpha1G T-type calcium channel in human fetal hearts. PLoS One, 8(9), e72668.PubMedCentralPubMed Strandberg, L. S., Cui, X., Rath, A., Liu, J., Silverman, E. D., Liu, X., et al. (2013). Congenital heart block maternal sera autoantibodies target an extracellular epitope on the alpha1G T-type calcium channel in human fetal hearts. PLoS One, 8(9), e72668.PubMedCentralPubMed
85.
Zurück zum Zitat Hu, K., Qu, Y., Yue, Y., & Boutjdir, M. (2004). Functional basis of sinus bradycardia in congenital heart block. Circulation Research, 94(4), e32–38.PubMed Hu, K., Qu, Y., Yue, Y., & Boutjdir, M. (2004). Functional basis of sinus bradycardia in congenital heart block. Circulation Research, 94(4), e32–38.PubMed
86.
Zurück zum Zitat Baig, S. M., Koschak, A., Lieb, A., Gebhart, M., Dafinger, C., Nurnberg, G., et al. (2011). Loss of Ca(v)1.3 (CACNA1D) function in a human channelopathy with bradycardia and congenital deafness. Nature Neuroscience, 14(1), 77–84.PubMed Baig, S. M., Koschak, A., Lieb, A., Gebhart, M., Dafinger, C., Nurnberg, G., et al. (2011). Loss of Ca(v)1.3 (CACNA1D) function in a human channelopathy with bradycardia and congenital deafness. Nature Neuroscience, 14(1), 77–84.PubMed
87.
Zurück zum Zitat Platzer, J., Engel, J., Schrott-Fischer, A., Stephan, K., Bova, S., Chen, H., et al. (2000). Congenital deafness and sinoatrial node dysfunction in mice lacking class D L-type Ca2+ channels. Cell, 102(1), 89–97.PubMed Platzer, J., Engel, J., Schrott-Fischer, A., Stephan, K., Bova, S., Chen, H., et al. (2000). Congenital deafness and sinoatrial node dysfunction in mice lacking class D L-type Ca2+ channels. Cell, 102(1), 89–97.PubMed
88.
Zurück zum Zitat Mangoni, M. E., Couette, B., Bourinet, E., Platzer, J., Reimer, D., Striessnig, J., et al. (2003). Functional role of L-type Cav1.3 Ca2+ channels in cardiac pacemaker activity. Proceedings of the National Academy of Sciences of the United States of America, 100(9), 5543–5548.PubMedCentralPubMed Mangoni, M. E., Couette, B., Bourinet, E., Platzer, J., Reimer, D., Striessnig, J., et al. (2003). Functional role of L-type Cav1.3 Ca2+ channels in cardiac pacemaker activity. Proceedings of the National Academy of Sciences of the United States of America, 100(9), 5543–5548.PubMedCentralPubMed
89.
Zurück zum Zitat Lahat, H., Eldar, M., Levy-Nissenbaum, E., Bahan, T., Friedman, E., Khoury, A., et al. (2001). Autosomal recessive catecholamine- or exercise-induced polymorphic ventricular tachycardia: clinical features and assignment of the disease gene to chromosome 1p13-21. Circulation, 103(23), 2822–2827.PubMed Lahat, H., Eldar, M., Levy-Nissenbaum, E., Bahan, T., Friedman, E., Khoury, A., et al. (2001). Autosomal recessive catecholamine- or exercise-induced polymorphic ventricular tachycardia: clinical features and assignment of the disease gene to chromosome 1p13-21. Circulation, 103(23), 2822–2827.PubMed
90.
Zurück zum Zitat Sumitomo, N., Sakurada, H., Taniguchi, K., Matsumura, M., Abe, O., Miyashita, M., et al. (2007). Association of atrial arrhythmia and sinus node dysfunction in patients with catecholaminergic polymorphic ventricular tachycardia. Circulation Journal, 71(10), 1606–1609.PubMed Sumitomo, N., Sakurada, H., Taniguchi, K., Matsumura, M., Abe, O., Miyashita, M., et al. (2007). Association of atrial arrhythmia and sinus node dysfunction in patients with catecholaminergic polymorphic ventricular tachycardia. Circulation Journal, 71(10), 1606–1609.PubMed
91.
Zurück zum Zitat Gao, Z., Rasmussen, T. P., Li, Y., Kutschke, W., Koval, O. M., Wu, Y., et al. (2013). Genetic inhibition of Na+–Ca2+ exchanger current disables fight or flight sinoatrial node activity without affecting resting heart rate. Circulation Research, 112(2), 309–317.PubMedCentralPubMed Gao, Z., Rasmussen, T. P., Li, Y., Kutschke, W., Koval, O. M., Wu, Y., et al. (2013). Genetic inhibition of Na+–Ca2+ exchanger current disables fight or flight sinoatrial node activity without affecting resting heart rate. Circulation Research, 112(2), 309–317.PubMedCentralPubMed
92.
Zurück zum Zitat Groenke, S., Larson, E. D., Alber, S., Zhang, R., Lamp, S. T., Ren, X., et al. (2013). Complete atrial-specific knockout of sodium-calcium exchange eliminates sinoatrial node pacemaker activity. PLoS One, 8(11), e81633.PubMedCentralPubMed Groenke, S., Larson, E. D., Alber, S., Zhang, R., Lamp, S. T., Ren, X., et al. (2013). Complete atrial-specific knockout of sodium-calcium exchange eliminates sinoatrial node pacemaker activity. PLoS One, 8(11), e81633.PubMedCentralPubMed
93.
Zurück zum Zitat Simrick, S., Schindler, R. F., Poon, K. L., & Brand, T. (2013). Popeye domain-containing proteins and stress-mediated modulation of cardiac pacemaking. Trends in Cardiovascular Medicine, 23(7), 257–263.PubMed Simrick, S., Schindler, R. F., Poon, K. L., & Brand, T. (2013). Popeye domain-containing proteins and stress-mediated modulation of cardiac pacemaking. Trends in Cardiovascular Medicine, 23(7), 257–263.PubMed
94.
Zurück zum Zitat Froese, A., Breher, S. S., Waldeyer, C., Schindler, R. F., Nikolaev, V. O., Rinne, S., et al. (2012). Popeye domain containing proteins are essential for stress-mediated modulation of cardiac pacemaking in mice. The Journal of Clinical Investigation, 122(3), 1119–1130.PubMedCentralPubMed Froese, A., Breher, S. S., Waldeyer, C., Schindler, R. F., Nikolaev, V. O., Rinne, S., et al. (2012). Popeye domain containing proteins are essential for stress-mediated modulation of cardiac pacemaking in mice. The Journal of Clinical Investigation, 122(3), 1119–1130.PubMedCentralPubMed
95.
Zurück zum Zitat Kirchmaier, B. C., Poon, K. L., Schwerte, T., Huisken, J., Winkler, C., Jungblut, B., et al. (2012). The Popeye domain containing 2 (popdc2) gene in zebrafish is required for heart and skeletal muscle development. Developmental Biology, 363(2), 438–450.PubMedCentralPubMed Kirchmaier, B. C., Poon, K. L., Schwerte, T., Huisken, J., Winkler, C., Jungblut, B., et al. (2012). The Popeye domain containing 2 (popdc2) gene in zebrafish is required for heart and skeletal muscle development. Developmental Biology, 363(2), 438–450.PubMedCentralPubMed
96.
Zurück zum Zitat Szûts, V., Ötvös, F., Dézsi, L., Vágvölgyi, C., Szalontai, B., Dobrzynski, H., et al. (2012). What have we learned from two-pore potassium channels? Their molecular configuration and function in the human heart. Acta Biologica Szegediensis, 56(2), 93–107. Szûts, V., Ötvös, F., Dézsi, L., Vágvölgyi, C., Szalontai, B., Dobrzynski, H., et al. (2012). What have we learned from two-pore potassium channels? Their molecular configuration and function in the human heart. Acta Biologica Szegediensis, 56(2), 93–107.
97.
Zurück zum Zitat Yanni, J., Tellez, J. O., Maczewski, M., Mackiewicz, U., Beresewicz, A., Billeter, R., et al. (2011). Changes in ion channel gene expression underlying heart failure-induced sinoatrial node dysfunction. Circulation. Heart Failure, 4(4), 496–508.PubMed Yanni, J., Tellez, J. O., Maczewski, M., Mackiewicz, U., Beresewicz, A., Billeter, R., et al. (2011). Changes in ion channel gene expression underlying heart failure-induced sinoatrial node dysfunction. Circulation. Heart Failure, 4(4), 496–508.PubMed
98.
Zurück zum Zitat Lalevee, N., Monier, B., Senatore, S., Perrin, L., & Semeriva, M. (2006). Control of cardiac rhythm by ORK1, a Drosophila two-pore domain potassium channel. Current Biology, 16(15), 1502–1508.PubMed Lalevee, N., Monier, B., Senatore, S., Perrin, L., & Semeriva, M. (2006). Control of cardiac rhythm by ORK1, a Drosophila two-pore domain potassium channel. Current Biology, 16(15), 1502–1508.PubMed
99.
Zurück zum Zitat Schindler, R. F., Poon, K. L., Simrick, S., & Brand, T. (2012). The Popeye domain containing genes: essential elements in heart rate control. Cardiovasc Diagn Ther, 2(4), 308–319.PubMedCentralPubMed Schindler, R. F., Poon, K. L., Simrick, S., & Brand, T. (2012). The Popeye domain containing genes: essential elements in heart rate control. Cardiovasc Diagn Ther, 2(4), 308–319.PubMedCentralPubMed
100.
Zurück zum Zitat Brand, T., Simrick, S. L., Poon, K. L., & Schindler, R. F. (2014). The cAMP-binding Popdc proteins have a redundant function in the heart. Biochemical Society Transactions, 42(2), 295–301.PubMedCentralPubMed Brand, T., Simrick, S. L., Poon, K. L., & Schindler, R. F. (2014). The cAMP-binding Popdc proteins have a redundant function in the heart. Biochemical Society Transactions, 42(2), 295–301.PubMedCentralPubMed
101.
Zurück zum Zitat Camozzi, D., Capanni, C., Cenni, V., Mattioli, E., Columbaro, M., Squarzoni, S., et al. (2014). Diverse lamin-dependent mechanisms interact to control chromatin dynamics: focus on laminopathies. Nucleus, 5(5). Camozzi, D., Capanni, C., Cenni, V., Mattioli, E., Columbaro, M., Squarzoni, S., et al. (2014). Diverse lamin-dependent mechanisms interact to control chromatin dynamics: focus on laminopathies. Nucleus, 5(5).
102.
Zurück zum Zitat Mohler, P. J., & Bennett, V. (2005). Ankyrin-based cardiac arrhythmias: a new class of channelopathies due to loss of cellular targeting. Current Opinion in Cardiology, 20(3), 189–193.PubMed Mohler, P. J., & Bennett, V. (2005). Ankyrin-based cardiac arrhythmias: a new class of channelopathies due to loss of cellular targeting. Current Opinion in Cardiology, 20(3), 189–193.PubMed
103.
Zurück zum Zitat Robaei, D., Ford, T., & Ooi, S. Y. (2014). Ankyrin-b syndrome: a case of sinus node dysfunction, atrial fibrillation and prolonged QT in a young adult. Heart Lung Circulatory Robaei, D., Ford, T., & Ooi, S. Y. (2014). Ankyrin-b syndrome: a case of sinus node dysfunction, atrial fibrillation and prolonged QT in a young adult. Heart Lung Circulatory
104.
Zurück zum Zitat Mohler, P. J., Le Scouarnec, S., Denjoy, I., Lowe, J. S., Guicheney, P., Caron, L., et al. (2007). Defining the cellular phenotype of “ankyrin-B syndrome” variants: human ANK2 variants associated with clinical phenotypes display a spectrum of activities in cardiomyocytes. Circulation, 115(4), 432–441.PubMed Mohler, P. J., Le Scouarnec, S., Denjoy, I., Lowe, J. S., Guicheney, P., Caron, L., et al. (2007). Defining the cellular phenotype of “ankyrin-B syndrome” variants: human ANK2 variants associated with clinical phenotypes display a spectrum of activities in cardiomyocytes. Circulation, 115(4), 432–441.PubMed
105.
Zurück zum Zitat Ackerman, M. J., & Mohler, P. J. (2010). Defining a new paradigm for human arrhythmia syndromes: phenotypic manifestations of gene mutations in ion channel- and transporter-associated proteins. Circulation Research, 107(4), 457–465.PubMedCentralPubMed Ackerman, M. J., & Mohler, P. J. (2010). Defining a new paradigm for human arrhythmia syndromes: phenotypic manifestations of gene mutations in ion channel- and transporter-associated proteins. Circulation Research, 107(4), 457–465.PubMedCentralPubMed
106.
Zurück zum Zitat Barbuti, A., Terragni, B., Brioschi, C., & DiFrancesco, D. (2007). Localization of f-channels to caveolae mediates specific beta2-adrenergic receptor modulation of rate in sinoatrial myocytes. Journal of Molecular and Cellular Cardiology, 42(1), 71–78.PubMed Barbuti, A., Terragni, B., Brioschi, C., & DiFrancesco, D. (2007). Localization of f-channels to caveolae mediates specific beta2-adrenergic receptor modulation of rate in sinoatrial myocytes. Journal of Molecular and Cellular Cardiology, 42(1), 71–78.PubMed
107.
Zurück zum Zitat Vatta, M., Ackerman, M. J., Ye, B., Makielski, J. C., Ughanze, E. E., Taylor, E. W., et al. (2006). Mutant caveolin-3 induces persistent late sodium current and is associated with long-QT syndrome. Circulation, 114(20), 2104–2112.PubMed Vatta, M., Ackerman, M. J., Ye, B., Makielski, J. C., Ughanze, E. E., Taylor, E. W., et al. (2006). Mutant caveolin-3 induces persistent late sodium current and is associated with long-QT syndrome. Circulation, 114(20), 2104–2112.PubMed
108.
Zurück zum Zitat Cronk, L. B., Ye, B., Kaku, T., Tester, D. J., Vatta, M., Makielski, J. C., et al. (2007). Novel mechanism for sudden infant death syndrome: persistent late sodium current secondary to mutations in caveolin-3. Heart Rhythm, 4(2), 161–166.PubMedCentralPubMed Cronk, L. B., Ye, B., Kaku, T., Tester, D. J., Vatta, M., Makielski, J. C., et al. (2007). Novel mechanism for sudden infant death syndrome: persistent late sodium current secondary to mutations in caveolin-3. Heart Rhythm, 4(2), 161–166.PubMedCentralPubMed
109.
Zurück zum Zitat Barbuti, A., Scavone, A., Mazzocchi, N., Terragni, B., Baruscotti, M., & Difrancesco, D. (2012). A caveolin-binding domain in the HCN4 channels mediates functional interaction with caveolin proteins. Journal of Molecular and Cellular Cardiology, 53(2), 187–195.PubMed Barbuti, A., Scavone, A., Mazzocchi, N., Terragni, B., Baruscotti, M., & Difrancesco, D. (2012). A caveolin-binding domain in the HCN4 channels mediates functional interaction with caveolin proteins. Journal of Molecular and Cellular Cardiology, 53(2), 187–195.PubMed
110.
Zurück zum Zitat Liu, Y., Bai, R., Wang, L., Zhang, C., Zhao, R., Wan, D., et al. (2013). Identification of a novel de novo mutation associated with PRKAG2 cardiac syndrome and early onset of heart failure. PLoS One, 8(5), e64603.PubMedCentralPubMed Liu, Y., Bai, R., Wang, L., Zhang, C., Zhao, R., Wan, D., et al. (2013). Identification of a novel de novo mutation associated with PRKAG2 cardiac syndrome and early onset of heart failure. PLoS One, 8(5), e64603.PubMedCentralPubMed
Metadaten
Titel
The genetic basis for inherited forms of sinoatrial dysfunction and atrioventricular node dysfunction
verfasst von
Raffaella Milanesi
Annalisa Bucchi
Mirko Baruscotti
Publikationsdatum
01.08.2015
Verlag
Springer US
Erschienen in
Journal of Interventional Cardiac Electrophysiology / Ausgabe 2/2015
Print ISSN: 1383-875X
Elektronische ISSN: 1572-8595
DOI
https://doi.org/10.1007/s10840-015-9998-z

Weitere Artikel der Ausgabe 2/2015

Journal of Interventional Cardiac Electrophysiology 2/2015 Zur Ausgabe

Schadet Ärger den Gefäßen?

14.05.2024 Arteriosklerose Nachrichten

In einer Studie aus New York wirkte sich Ärger kurzfristig deutlich negativ auf die Endothelfunktion gesunder Probanden aus. Möglicherweise hat dies Einfluss auf die kardiovaskuläre Gesundheit.

Intervallfasten zur Regeneration des Herzmuskels?

14.05.2024 Herzinfarkt Nachrichten

Die Nahrungsaufnahme auf wenige Stunden am Tag zu beschränken, hat möglicherweise einen günstigen Einfluss auf die Prognose nach akutem ST-Hebungsinfarkt. Darauf deutet eine Studie an der Uniklinik in Halle an der Saale hin.

Shunt-Therapie bei Herzinsuffizienz: Kein Anzug, der allen passt

13.05.2024 Chronische Herzinsuffizienz Nachrichten

Die Anlage eines interatrialen Shunts zur Reduktion des linksatrialen Drucks ist ein neuer Therapieansatz bei Herzinsuffizienz. Viele Patienten sprechen darauf an, andere jedoch nicht. 

Vorsicht, erhöhte Blutungsgefahr nach PCI!

10.05.2024 Koronare Herzerkrankung Nachrichten

Nach PCI besteht ein erhöhtes Blutungsrisiko, wenn die Behandelten eine verminderte linksventrikuläre Ejektionsfraktion aufweisen. Das Risiko ist umso höher, je stärker die Pumpfunktion eingeschränkt ist.

Update Kardiologie

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.