Feeds:
Posts
Comments

Posts Tagged ‘David S. Park’

Genetics of Conduction Disease: Atrioventricular (AV) Conduction Disease (block): Gene Mutations – Transcription, Excitability, and Energy Homeostasis

Curator: Aviva Lev-Ari, PhD, RN

UPDATED on 6/13/2013

with a CASE of  Anti-Ro Antibodies and Reversible Atrioventricular Block

N Engl J Med 2013; 368:2335-2337 June 13, 2013 DOI: 10.1056/NEJMc1300484

As an Introduction to the Genetics of Conduction Disease, we selected the following article which represents the MOST comprehensive review of the Human Cardiac Conduction System presented to date:

Circulation.2011; 123: 904-915 doi: 10.1161/​CIRCULATIONAHA.110.942284

The Cardiac Conduction System

  1. David S. Park, MD, PhD;
  2. Glenn I. Fishman, MD

+Author Affiliations


  1. From the Leon H. Charney Division of Cardiology, New York University School of Medicine, New York, NY.
  1. Correspondence to Glenn I. Fishman, MD, Leon H. Charney Division of Cardiology, New York University School of Medicine, 522 First Ave, Smilow 801, New York, NY 10016. E-mail glenn.fishman@med.nyu.edu

Key Words:

The human heart beats 2.5 billion times during a normal lifespan, a feat accomplished by cells of the cardiac conduction system (CCS). The functional components of the CCS can be broadly divided into the impulse-generating nodes and the impulse-propagating His-Purkinje system. Human diseases of the conduction system have been identified that alter impulse generation, impulse propagation, or both. CCS dysfunction is primarily due to acquired conditions such as myocardial ischemia/infarct, age-related degeneration, procedural complications, and drug toxicity. Inherited forms of CCS disease are rare, but each new mutation provides invaluable insight into the molecular mechanisms governing CCS development and function. Applying a multidisciplinary approach, which includes human genetic screening, biophysical analysis, and transgenic mouse technology, has yielded a broad array of gene families involved in maintaining normal CCS physiology (Figure 1). In this review, we discuss gene families that have been implicated in human CCS diseases of rhythm, conduction block, accessory conduction, and development (Table). We also investigate evolving therapeutic strategies that may serve as adjuvant or replacement therapy to current implantable pacemakers.

Figure 1.

View larger version:

Figure 1.

Cardiac conduction system cell. Genes identified in human cardiac conduction system disease are highlighted.

Table.

Genetic Basis of Conduction System Disease

Diseases of Automaticity

The human sinoatrial node (SAN) is a crescent-shaped, intramural structure with its head located subepicardially at the junction of the right atrium and the superior vena cava and its tail extending 10 to 20 mm along the crista terminalis.26 The SAN has complex 3-dimensional tissue architecture with central and peripheral components made up of distinct ion channel and gap junction expression profiles.27 Central and peripheral cells have different action potential characteristics and conduction properties (Figure 2).27Experimental and computational models have demonstrated that SAN heterogeneity is necessary to maintain normal automaticity and impulse conduction.28,,30

Figure 2.

Figure 2.

Electrophysiological heterogeneity of the sinoatrial node (SAN). The central SAN, the site of dominant pacemaking, is electronically insulated from the hyperpolarizing atrial myocardium through the differential expression of connexins and ion channels. Peripheral SAN cells are electrophysiologically intermediate between central cells and atrial cardiomyocytes. SR indicates sarcoplasmic reticulum.

Pacemaker automaticity is due to spontaneous diastolic depolarization of phase 4, which depolarizes the membrane to threshold potential generating rhythmic action potentials. The current paradigm of SAN automaticity has been modeled as 2 clocks that function in concert, the “membrane voltage clock” and the “calcium clock.” The membrane voltage clock is produced by the net disequilibrium between the decay of outward potassium currents (IK) and the activation of inward currents that include, but are not limited to, background sodium-sensitive current (Ib Na), L- and T-type calcium currents (ICa,L,ICa,T), sustained inward (Ist) current, and hyperpolarization-activated current (If) (Figure 2).27,31,,33

The subsarcolemmal calcium clock contributes to SAN diastolic depolarization through the spontaneous, rhythmic release of Ca2+ from the sarcoplasmic reticulum (SR) via the ryanodine type 2 receptor (RYR2).34 The local intracellular calcium (Cai) elevations drive the sodium-calcium exchange current (INCX) to substitute 1 intracellular Ca2+ for 3 extracellular Na+. The net gain in positive charge results in membrane depolarization.35The elevation of intracellular Ca2+ occurs in the latter third of diastolic depolarization and is sensitive to β-adrenergic stimulation.36

Human mutations affecting the voltage clock

  • (SCN5A and HCN4),

  • calcium clock (RYR2 and CASQ2), or both mechanisms

  • (ANKB) have been identified that negatively affect sinus node function.37,38

Diseases of Conduction BlockConduction block can occur at any level of the CCS and can manifest as sinoatrial exit block, atrioventricular block, infra-Hisian block, or bundle branch block. Impaired conduction can be caused by ion channel defects that alter action potential shape or by defective coupling between cardiomyocytes. Inherited defects in cardiac conduction have been linked to mutations in SCN5A and SCN1B (both affect phase 0) and KCNJ2 (affects phase 3 and 4). 

The cardiac sodium channel consists of the pore-forming α-subunit (encoded by SCN5A) and a modulatory β-subunit (encoded by SCN1B). The α-subunit contains a voltage sensor that allows for rapid activation in response to membrane depolarization. After depolarization, the sodium channel undergoes a period of inactivation, in which it is refractory to further impulses. SCN5A requires membrane repolarization to relieve the inactivated state. The inward rectifier potassium channel, Kir2.1, encoded by KCNJ2, maintains the resting membrane potential. Therefore, proper functioning of Nav1.5 and Kir2.1 is necessary for normal cardiac excitability.

SCN5A

Progressive cardiac conduction defect, or Lev-Lenègre disease, is characterized by age-related, fibrosclerotic degeneration of the His-Purkinje system.6 Impulse propagation through the proximal ventricular conduction system progressively declines, resulting in bundle branch blocks and eventually complete atrioventricular block. An inherited form of Lev-Lenègre disease is associated with loss of function mutations in SCN5A and can exist alone or as overlap syndromes with Brugada or long QT syndrome 3.6 Inherited progressive cardiac conduction defect is associated with a high risk of complete atrioventricular block and Stoke-Adams syncope without ventricular dysrhythmia.7 Schott et al8 identified a mutation in SCN5A that cosegregates with Lenègre disease in a large French family. Affected individuals had variable degrees of conduction block requiring pacemaker implantation in 4 family members because of syncope or complete heart block. Linkage analysis and candidate gene sequencing identified a T>C substitution at position +2 of the donor splice site of intron 22 (IVS22+2 T>C), which results in a mutant lacking the voltage-sensitive segment.8 Functional analysis demonstrated no transient inward sodium current in response to depolarization, consistent with a loss-of-function mutation.6

SCN1B

The majority of patients with Brugada and conduction disease do not have SCN5Amutations. Therefore, modifiers of Nav1.5 expression or function have become the target of candidate gene sequencing approaches. Watanabe et al9 identified SCN1B mutations in 3 families with conduction disease with or without Brugada syndrome. Coexpression of mutant β-subunits with Nav1.5 resulted in diminished sodium current.

KCNJ2

Mutations in KCNJ2 have been found in a rare autosomal dominant condition called Andersen-Tawil syndrome, characterized by periodic paralysis, dysmorphic features, polymorphic ventricular tachycardia, and cardiac conduction disease.10,11 ECG evaluation of 96 patients with Andersen-Tawil syndrome from 33 unrelated kindreds revealed conduction defects at multiple levels from the atrioventricular node to the distal conduction system.55 Cardiomyocytes expressing a dominant-negative subunit of Kir2.1 exhibited a 95% reduction in IK1, resulting in significant action potential prolongation. Mouse models of Andersen-Tawil syndrome exhibited a slower heart rate and significant slowing of conduction.56,57

Diseases of Accessory Conduction

Wolff-Parkinson-White (WPW) syndrome is characterized by preexcitation of ventricular myocardium via an accessory pathway (bundle of Kent) that bypasses the normal slow conduction through the atrioventricular node. Ventricular preexcitation is common, with a disease prevalence of 1.5 to 3 per 1000 people.22,58 Histological evaluation of Kent bundles resected from human subjects displayed features of typical ventricular myocytes with expression of connexin 43 (Cx43).59 The expression of high-conductance gap junctions in bypass tracts enables them to preexcite ventricular myocardium, manifesting as a short PR and a slurred QRS complex, or “delta wave,” on the ECG. The vast majority of WPW cases are sporadic, and the underlying mechanism remains unknown; however, rare inherited forms have been reported. Vidaillet et al60 determined that 3.4% of probands with WPW had 1 or more first-degree relatives with accessory conduction.

PRKAG2

A familial form of WPW with an autosomal dominant mode of transmission was identified in 2 families. Thirty-one affected individuals had evidence of preexcitation and cardiac hypertrophy. A missense mutation in PRKAG2 was identified that results in a constitutively active form of the γ2 regulatory subunit of AMP-activated protein kinase.22,23 Under normal conditions, AMP-activated protein kinase responds to energy-depleted states by increasing glucose uptake and promoting glycolysis. Transgenic mice expressing a heart-restricted, constitutively active mutant, PRKAG2N488I, recapitulated the human WPW phenotype of cardiac hypertrophy, preexcitation, and conduction defects. The predominant histological finding was ventricular myocyte engorgement with glycogen-laden vacuoles. The disruption of the annulus fibrosus by vacuolated ventricular myocytes resulted in the preexcitation phenotype.61 Using a mouse model of reversible glycogen-storage defect, Wolf et al62 demonstrated that the cardiomyopathy and CCS degeneration seen in PRKAG2N488I mice were reversible processes.

BMP2

Lalani et al24 reported a novel WPW syndrome associated with microdeletion of the bone morphogenetic protein-2 (Bmp2) region within 20p12.3 that is characterized by variable cognitive deficits and dysmorphic features. The BMPs are members of the transforming growth factor-β superfamily and play a critical role in cardiac development. Mice with cardiac deletion of BMP receptor type Ia (Bmpr1a) were embryonic lethal before E18.5 because of abnormal development of trabecular and compact myocardium, interventricular septum, and endocardial cushion.63 More restricted deletion of Bmpr1a in the atrioventricular canal resulted in defective atrioventricular valve formation and maturation defects in the annulus fibrosus, resulting in preexcitation.64,65

 

Diseases of CCS Development

Congenital heart disease is the most common form of birth defect, affecting 1% to 2% of live births.66 Arrhythmias may result from defective CCS specification/patterning, malformation or displacement of the conduction system, altered hemodynamics, prolonged hypoxic states, or postoperative sequelae.67,68 Developmentally, the conduction system derives from myocardial precursor cells within the fetal heart.69,,71The process by which conduction cells are specified or recruited into a “conduction” versus “working myocyte” lineage is determined by the regional expression of transcription factors.69,,74 The main transcription factors identified in human CCS development are the T-box and homeobox factors.

TBX5

Holt-Oram syndrome is an autosomal dominant condition characterized by preaxial radial ray limb deformities (defects of the radius, carpal bones, and/or thumbs) and cardiac septation defects. The septal defects are typically ostium secundum atrial septal defects, muscular ventricular septal defects, and atrioventricular canal defects. Patients with Holt-Oram syndrome manifest variable degrees of CCS dysfunction, such as sinus bradycardia and atrioventricular block, even in the absence of overt structural heart disease. In 1997, Basson et al18 screened 2 families with Holt-Oram syndrome and identified mutations in the T-box transcription factor, TBX5. The T-box transcription factors can function as transcriptional activators or repressors and are known to be critical regulators of cardiac specification and differentiation. Seven TBX family members are expressed in the developing heart, 3 of which (TBX1, TBX5, TBX20) have been linked to human congenital heart disease.75

Mice deficient in Tbx5 were embryonic lethal at E10.5 because of arrested development of the atria and left ventricle. Tbx5+/− mice recapitulated the upper limb and cardiac manifestations of human Holt-Oram syndrome, including the conduction abnormalities.72,76 Significant maturation defects in the atrioventricular canal and ventricular conduction system were present.76 Moskowitz et al76 demonstrated thatTbx5+/− mice have maturation failure of the atrioventricular canal manifesting as persistent atrioventricular rings around the tricuspid and mitral valves. Patterning defects were noted in the His bundle and bundle branches, including complete absence of right bundle branch formation in some cases. Expression of CCS-enriched markers, such as atrial natriuretic factor and Cx40, were found to be significantly downregulated, implicating TBX5 as a transcriptional activator of these genes. TBX5 and the homeobox transcription factor NKX2-5 were found to act synergistically in upregulating atrial natriuretic factor and Cx40 expression.76

Conduction Disease Associated With Neuromuscular Disorders

Neuromuscular disorders represent a diverse collection of diseases that commonly present with cardiac involvement. Mutations have been identified in genes involved in the cytoskeleton, nuclear envelope, and mitochondrial function. Cardiac involvement typically manifests as dilated or hypertrophic cardiomyopathy, atrioventricular conduction defects, and atrial and ventricular dysrhythmias.82

EMD and LMNA

Mutations affecting the nuclear envelope have been associated with significant CCS dysfunction. The inner membrane of the nuclear envelope is a highly organized structure, composed of integral membrane proteins and nuclear cytoskeletal proteins that function together in higher-order chromatin structure and transcriptional regulation. The lamins (A, B, and C) are an integral part of an intermediate filament network that imparts structural rigidity to the inner nuclear membrane. Emerin, a member of the nuclear lamina-associated protein family, putatively mediates anchoring of chromatin to the cytoskeleton. Mutations in emerin (EMD) or lamin A/C (LMNA) result in X-linked Emery-Dreifuss muscular dystrophy and autosomal dominant Emery-Dreifuss muscular dystrophy,20respectively. Individuals with Emery-Dreifuss muscular dystrophy develop progressive skeletal muscle weakness in the first decade of life and cardiac involvement (dilated cardiomyopathy and atrioventricular block) in the second decade.82,83

Arimura et al84 engineered a mouse model of autosomal dominant Emery-Dreifuss muscular dystrophy by knocking-in an Lmna missense mutation (H222P) previously identified from a family with typical autosomal dominant Emery-Dreifuss muscular dystrophy. The mouse model faithfully recapitulated the human disease with LmnaH222P/H222P mice exhibiting locomotive defects, dilated cardiomyopathy, and CCS dysfunction. Telemetric evaluation of the mutant mice revealed PR prolongation and QRS complex widening. A similar CCS defect was seen in mice haploinsufficient in the Lmna gene. Lmna+/− mice exhibited sinus bradycardia with variable degrees of atrioventricular block. Histological evaluation of these mice revealed nuclear deformation and apoptosis in atrioventricular node cells.85 Another engineered mouse line expressing LmnaN195K, known to cause autosomal dominant dilated cardiomyopathy with conduction disease in humans,86 exhibited high-grade atrioventricular block and complete heart block. Biochemical evaluation revealed reduced expression and mislocalization of Cx40 and Cx43 in mutant atrial tissue.87 Desmin staining also revealed structural defects of the sarcomere and intercalated discs.87

Genome-wide expression profiling of Lmna H222P mouse hearts revealed significant increases in mitogen-activated protein kinase (MAPK) signaling pathways.88Hyperactivation of MAPK pathways is associated with cardiomyopathy and CCS dysfunction. A significant increase of the activated forms of 2 MAPKs, JNK and ERK1/2, was noted in mutant hearts that predated the onset of overt or molecularly defined cardiomyopathy.88 Treatment of Lmna H222P mice with an inhibitor of ERK phosphorylation abrogated the increase in biomarkers of cardiomyopathy and restored ejection fraction to normal levels. These findings directly link MAPK hyperactivation to the cardiomyopathic phenotype in Lmna H222P mice.89

On the basis of the phenotypes of these mouse models, lamin A/C appears to maintain the functional integrity of the CCS in 2 ways: (1) protection of the nucleus against mechanical stress and (2) maintenance of proper chromatin organization to ensure accurate gene expression, such as in connexin expression and MAPK signaling pathways.83

Future Directions

Linkage analysis with positional cloning has been a highly effective means of identifying gene mutations within kindreds of monogenic disease. More than 1000 genes have been identified with this approach, including those in this review. With the sequencing of the human genome, the promise of identifying genetic causes of complex, multifactorial diseases is becoming more of a reality. One major step in this direction was the development of genome-wide association studies.94

The genome-wide association study is a test of association between a disease and genetic markers that span the entire genome. The technique relies on the fact that variance at one locus predicts with high probability variance of an adjacent locus because of linkage disequilibrium. In other words, there is nonrandom cosegregation of a series of genetic markers that are close together in the genome. This cluster of linked markers is known as a haplotype. The first study of haplotype structure within 4 populations (Yoruban, Northern/Western Europeans, Chinese, and Japanese) was published in Naturein 2005 by the International HapMap Consortium. Their work reported that individual genetic markers (single nucleotide polymorphisms) predict adjacent markers typically with a resolution of ≈30 000 bp. Considering that the human genome is ≈3×109 bp, they projected that <500 000 single nucleotide polymorphisms would be needed to survey the entire genome for all common genetic variants.94,95

Genome-wide association studies have now been used to identify genetic variants that influence ECG parameters in different populations. Intermediate parameters, such as heart rate or PR interval, were used as surrogate markers of disease for 2 reasons: (1) They have an association with cardiovascular morbidity and atrial fibrillation, and (2) they have tighter associations with gene variants than the actual disease. Holm et al96reported several genome-wide associations using a cutoff P value <1.6×10−9. One locus harboring MYH6 was associated with heart rate, 4 loci (TBX5SCN10ACAV1, andARHGAP24) were associated with PR interval, and 4 loci (TBX5SCN10A6p21, and10q21) were associated with QRS duration. They went on to test these associations with individuals manifesting different arrhythmias in an Icelandic and Norwegian population. Correlations were found between atrial fibrillation and TBX5 and CAV1 (P=4.0×10−5 andP=0.00032, respectively), between advanced atrioventricular block and TBX5 (P=0.0067), and between pacemaker implantation and SCN10A (P=0.0029).

Similar loci were identified by 2 additional independent genome-wide association studies in a European population and an Indian Asian population. Pfeufer et al97 reported 9 loci that were highly associated with PR interval (P<5×10−8) from a meta-analysis of the CHARGE Consortium with >28 000 European subjects. One locus had associations with 2 sodium channels (SCN10A and SCN5A), and 6 loci were near genes involved in cardiac development (CAV1-CAV2NKX2-5SOX5WNT11MEIS1and TBX5-TBX3). Of these,SCN10ASCN5ACAV1-CAV2NKX2-5, and SOX5 were found to be associated with atrial fibrillation. Chambers et al98 also reported the association between SCN10A and PR interval in 6543 Indian Asians. Physiological testing of Scn10a-deficient mice revealed shortened PR intervals in knockout mice with no significant difference in all other ECG and echocardiographic parameters.

The discovery of novel gene families associated with human conduction and arrhythmic diseases with the use of genome-wide association studies is well under way. Identification of SCN10A by 3 independent groups studying different populations confirms the fidelity of this approach. Further experiments confirming the significance of these associations will need to be performed. In addition to identifying novel gene targets, this technique will also aid in the discovery of new associations with noncoding regions in which new epigenetic modifiers and transcriptional/translational regulators, such as microRNAs, will be identified.

Therapeutic Strategies

The current standard of care for symptomatic bradycardia due to conduction system disease is the implantation of an electronic pacemaker. Despite their success, electronic pacemakers have limitations, which include lead complications, finite battery life, potential for infection, lack of autonomic responsiveness, and size restriction in younger patients. These limitations have spurred on the development of biological pacemakers, the premise of which is to restore pacemaking activity with the use of viral-based or stem cell–based gene delivery systems.99 The identification and characterization of genes involved in generating pacemaker currents have allowed biological pacemaker technology to become a reality.

The restoration of sinus pacing rates can be achieved by modulating inward and outward currents to establish or increase the slope of diastolic depolarization in cardiac tissue. Increasing inward currents and/or decreasing outward currents increase the slope of diastolic depolarization and therefore the pacing rate. Genes that have been investigated or are under current investigation include the following: (1) β2-adrenergic receptor,100,101(2) dominant-negative Kir2.1 mutants,102 (3) adenylate cyclase type VI (ACVI),103,104and (4) HCN channels.105 The β2-adrenergic receptor and adenylate cyclase type VI both increase cAMP levels, leading to activation of endogenous HCN channels and calcium clock mechanisms. Although initial animal models using the β2-adrenergic receptor showed promise with transient increases in heart rate, the potential for proarrhythmia and the inability of this approach to establish de novo pacemaker activity limited its efficacy.101

Another approach focused on modifying ionic currents that convert working myocardial cells, which have relatively stable diastolic potentials, into cells with phase 4 diastolic depolarization. It was postulated that atrial and ventricular myocytes have the potential for automaticity, but that hyperpolarizing currents, such as IK1, prevent diastolic depolarization by stabilizing the resting membrane potential. Miake et al102 confirmed this hypothesis when they demonstrated that adenoviral delivery of a dominant-negative Kir2.1 construct into the left ventricle of guinea pigs resulted in conversion of quiescent myocytes into pacemaker cells. Unfortunately, significant action potential prolongation limited the clinical utility of this treatment strategy.102

Rosen and colleagues105,106 demonstrated that automaticity could be induced in quiescent myocardium with the use of heterologous expression of HCN channels that produce the pacemaker current If. Qu and Plotnikov et al demonstrated that stable autonomous rhythms could be generated when adenovirus encoding HCN2 was injected into the left atrium105 or left bundle branch106 of a canine heart. To bypass the limitations of viral-based systems, such as host immune response, several groups reported the successful use of cell-based delivery systems. Plotnikov et al107 reported the successful implantation of human mesenchymal stem cells expressing HCN2 in the left ventricle of a canine model of atrioventricular block. Dogs maintained stable ectopic pacemaker activity for >6 weeks without the use of immunosuppression.107 Human mesenchymal stem cells electronically couple to host myocardium through gap junctions; therefore, conditions with significant gap junction remodeling may affect the efficacy of this method.

Although standalone biological pacemakers may be far into the future, adjuvant biological pacemakers may find real-world utility for current deficiencies of electronic pacemakers, such as limited battery life and device infections. For example, biological preparations used in conjunction with device therapy may be used to extend battery life, decreasing the frequency of generator changes. Transient injectable pacemakers may also function as bridge therapy after lead extraction of an infected device. The need for adjuvant biological pacemakers is clear, but continued refinement of gene- and cell-based delivery systems will be necessary to make this technology a reality.99

Conclusion

Although rare, inherited arrhythmias have become an invaluable tool in identifying the genetic determinants of CCS function. Each new mutation enhances our understanding and appreciation of the biochemical and structural complexity needed for cardiac impulse generation and propagation. This methodology is hampered, however, by the relative scarcity of inherited conditions affecting the CCS. The addition of genome-wide association studies has broadened this search for novel genes beyond rare familial afflictions to include common, multifactorial conditions. It is hoped that this exciting new frontier will bring to light the complex interplay of genes and genetic/epigenetic modifiers that influence the prevalence of common diseases. These genetic screens will ultimately yield a bevy of new gene targets for pharmaceutical or gene-based therapeutics of the future.

Sources of Funding

Studies in the Fishman laboratory are supported by National Institutes of Health grants HL64757, HL081336, and HL82727 and a New York State STEM award (to Dr Fishman) and a Heart Rhythm Foundation Fellowship (to Dr Park).

Genetics of Atrioventricular Conduction Disease in Humans.

Benson DW.

Source

Division of Cardiology, ML7042, Children’s Hospital Medical Center, 3333 Burnet Avenue, Cincinnati, OH 45229, USA. woody.benson@cchmc.org

Abstract

Atrioventricular (AV) conduction disease (block) describes impairment of the electrical continuity between the atria and ventricles. Classification of AV block has utilized biophysical characteristics, usually the extent (first, second, or third degree) and site of block (above or below His bundle recording site). The genetic significance of this classification is unknown. In young patients, AV block may result from injury or be the major cardiac manifestation of neuromuscular disease. However, in some cases, AV block has unknown or idiopathic cause. In such cases, familial clustering has been noted and published pedigrees show autosomal dominant inheritance; associated heart disease is common (e.g., congenital heart malformation, cardiomyopathy). The latter finding is not surprising given the common origin of working myocytes and specialized conduction system elements. Using genetic models incorporating reduced penetrance (disease absence in some individuals with disease gene), variable expressivity (individuals with disease gene have different phenotypes), and genetic heterogeneity (similar phenotypes, different genetic cause), molecular genetic causes of AV block are being identified. Mutations identified in genes with diverse functions (transcription, excitability, and energy homeostasis) for the first time provide the means to assess risk and offer insight into the molecular basis of this important clinical condition previously defined only by biophysical characteristics.

http://www.ncbi.nlm.nih.gov/pubmed/15372490

Additional Studies on Genetic Congenital AV Block

1) 12738236
Na+ channel mutation leading to loss of function and non-progressive cardiac conduction defects.
BACKGROUND: We previously described a Dutch family in which congenital cardiac conduction disorder has clinically been identified. The ECG of the index patient showed a first-degree AV block associated with extensive ventricular conduction delay. Sequencing of the SCN5A locus coding for the human cardiac Na+ channel revealed a single nucleotide deletion at position 5280, resulting in a frame-shift in the sequence coding for the pore region of domain IV and a premature stop codon at the C-terminus. METHODS AND RESULTS: Wild type and mutant Na+ channel proteins were expressed in Xenopus laevis oocytes and in mammalian cells. Voltage clamp experiments demonstrated the presence of fast activating and inactivating inward currents in cells expressing the wild type channel alone or in combination with the beta1 subinut (SCN1B). In contrast, cells expressing the mutant channels did not show any activation of inward current with or without the beta1 subunit. Culturing transfected cells at 25 degrees C did not restore the Na+ channel activity of the mutant protein. Transient expression of WT and mutant Na+ channels in the form of GFP fusion proteins in COS-7 cells indicated protein expression in the cytosol. But in contrast to WT channels were not associated with the plasma membrane. CONCLUSIONS: The SCN5A/5280delG mutation results in the translation into non-function channel proteins that do not reach the plasma membrane. This could explain the cardiac conduction defects in patients carrying the mutation.
2) 12956334
The genetic origin of atrioventricular conduction disturbance in humans.
Atrioventricular (AV) conduction disturbance (block) describes impairment of the electrical continuity between the atria and ventricles. Clinical classification of AV block has utilized biophysical characteristics, usually the extent (1st, 2nd, 3rd degree) and site of block (above or below His bundle recording site). The genetic significance of this classification is not known. In some casesAV block occurrence is associated with intrauterine exposure to maternal antibody (anti-Ro, anti-La), and other cases are associated with injury (e.g. surgery). Based on familial clustering of idiopathic AV block, a genetic cause has also been suspected. Published pedigrees show autosomal dominant inheritance, and associated heart disease is common (e.g. congenital heart malformation, cardiomyopathy, etc.). The latter finding is not unexpected given the common origin of working myocytes and elements of the specialized conduction system. Using genetic models incorporating reduced penetrance (presence of disease genotype in absence of phenotype), variable expressivity (presence of a disease genotype with variable phenotypes) and genetic heterogeneity (similar phenotypes, different disease genotypes), molecular genetic causes of AV block are being identified. These findings are significant as they provide insight into the molecular basis of a clinical condition previously defined only by biophysical characteristics.
3) 15372490
Genetics of atrioventricular conduction disease in humans.
Atrioventricular (AV) conduction disease (block) describes impairment of the electrical continuity between the atria and ventricles. Classification of AV block has utilized biophysical characteristics, usually the extent (first, second, or third degree) and site of block (above or below His bundle recording site). The genetic significance of this classification is unknown. In young patients, AV block may result from injury or be the major cardiac manifestation of neuromuscular disease. However, in some cases, AV blockhas unknown or idiopathic cause. In such cases, familial clustering has been noted and published pedigrees show autosomal dominant inheritance; associated heart disease is common (e.g., congenital heart malformation, cardiomyopathy). The latter finding is not surprising given the common origin of working myocytes and specialized conduction system elements. Using genetic models incorporating reduced penetrance (disease absence in some individuals with disease gene), variable expressivity (individuals with disease gene have different phenotypes), and genetic heterogeneity (similar phenotypes, different genetic cause), molecular genetic causes of AV block are being identified. Mutations identified in genes with diverse functions (transcription, excitability, and energy homeostasis) for the first time provide the means to assess risk and offer insight into the molecular basis of this important clinical condition previously defined only by biophysical characteristics.

SOURCE:

Anti-Ro Antibodies and Reversible Atrioventricular Block

N Engl J Med 2013; 368:2335-2337 June 13, 2013DOI: 10.1056/NEJMc1300484

To the Editor:

Transplacental transfer of anti-Ro antibodies is a well-known cause of conduction defects and permanent atrioventricular block in newborns.1 In adults, conduction disturbances related to these antibodies are rare.2

We report a case of a 26-year-old woman with no history of this condition who was admitted to the hospital through the emergency department after having several syncopal episodes. Electrocardiography (ECG) performed while the patient was at rest showed complete atrioventricular block and ventricular escape rhythm associated with left bundle-branch block (Figure 1AFIGURE 1Electrocardiographic Findings.). Laboratory evaluation included a positive test for antinuclear antibodies (with the HEp-2 cell substrate) at a titer of 1:320, with a speckled pattern and specificity for extractable nuclear antigens, including antibodies against Ro52 confirmed by means of immunoblot and enzyme-linked immunosorbent assays (first measurement of antibodies, 1.2 U per milliliter). No clinical manifestations of rheumatologic disease were present. Other causes of reversible atrioventricular block were ruled out. The patient had no history of cardiac surgery, ablation procedures, or drug use. There was no evidence of infiltrative diseases (e.g., sarcoidosis or amyloidosis) or myocardial ischemia, nor was there clinical suspicion of infectious diseases that cause conduction disturbances (e.g., Lyme disease or Chagas’ disease). Levels of electrolytes and thyrotropin were normal. Transthoracic echocardiography and magnetic resonance imaging were unremarkable.

During the first 4 days after admission, the patient had varying degrees of atrioventricular block. An electrophysiological study showed a mildly prolonged HV interval of 62 msec during sinus rhythm (normal values, 35 to 55 msec) and a pathologic response to atrial pacing, with atrioventricular block occurring after the deflection of the bundle of His during continuous stimulation at a fixed cycle length of 490 msec (Figure 1B). Intravenous methylprednisolone was initiated at a dose of 1 mg per kilogram of body weight per day, and 1:1 atrioventricular conduction was subsequently maintained on surface ECG. A second electrophysiological study during treatment showed normal atrioventricular conduction.

Maintenance immunosuppressive therapy with azathioprine (at a dose of 100 mg daily) and methylprednisolone (at a dose of 4 mg daily) was initiated and continued for 12 months. Serial anti-Ro (SS-A) levels fluctuated during follow-up and became negative after 1 year. Because of the uncertainty of the outcome, a backup pacemaker was implanted. The patient remained completely asymptomatic for 12 months with sustained normal atrioventricular conduction.

In this case of atrioventricular block in an adult patient with positive anti-Ro antibodies, we used electrophysiological testing to localize the conduction defect below the atrioventricular node. This finding, together with left bundle-branch block detected on ECG, suggests specific involvement of the Purkinje fibers. The pathogenesis of cardiac conduction disturbances in adults with positive anti-Ro (SS-A) antibodies remains unclear.3 Experimental studies suggest that anti-Ro antibodies interact with calcium channels and cause reversible inhibition of calcium currents. In fetal hearts, the internalization of these channels leads to apoptosis and fibrosis of the conduction tissue. The presence of a fully developed sarcoplasmic reticulum and the apparent lack of antibody-induced apoptosis in adult cardiomyocytes may explain the differential susceptibility of adult hearts to anti-Ro antibodies2 and, conceivably, the reversibility of the conduction disease in such persons.

Irene Santos-Pardo, M.D.
Melania Martínez-Morillo, M.D.
Roger Villuendas, M.D.
Antoni Bayes-Genis, M.D., Ph.D.
Hospital Universitari Germans Trias i Pujol, Badalona, Spain
abayes.germanstrias@gencat.cat

REFERENCES

1

Chameides L, Truex RC, Vetter V, Rashkind WJ, Galioto FM Jr, Noonan JA. Association of maternal systemic lupus erythematosus with congenital complete heart block. N Engl J Med 1977;297:1204-1207
Full Text | Web of Science | Medline

Lazzerini PE, Capecchi PL, Laghi-Pasini F. Anti-Ro/SSA antibodies and cardiac arrhythmias in the adult: facts and hypotheses. Scand J Immunol 2010;72:213-222
CrossRef | Web of Science | Medline

Costedoat-Chalumeau N, Georgin-la-Vialle S, Amoura Z, Piette J-C. Anti-SSA/Ro and anti-SSB/La antibody-mediated congenital heart block. Lupus 2005;14:660-664
CrossRef | Web of Science | Medline

SOURCE

http://www.nejm.org/doi/full/10.1056/NEJMc1300484?query=TOC

New Research on the Genetics of Conduction Disease

2010  
Heart failure clinics

  

conduction diseases (CD) include defects in impulse generation and conduction. Patients with CD may manifest a wide range of clinical presentations, from asymptomatic to potentially life-threatening arrhythmias. The pathophysiologic mechanisms underlying CD are diverse and may have implications for diagnosis, treatment, and prognosis. Known causes of functional CD include cardiac ion channelopathies or defects in modifying proteins, such as cytoskeletal proteins. Progress in molecular biology and genetics along with development of animal models has increased the understanding of the molecular mechanisms of these disorders. This article discusses the genetic basis for CD and its clinical implications.
(Beinart et al. 2010)
Beinart R, Ruskin J, et al. (2010). The genetics of conduction disease. Heart Fail Clin 6 (2): 201-14.
PMID: 20347788  DOI: 10.1016/j.hfc.2009.11.006  PII: S1551-7136(09)00108-1
2012  
PLoS genetics

  

(Curran and Mohler 2012)
Curran J and Mohler PJ (2012). Defining the Pathways Underlying the Prolonged PR Interval in Atrioventricular Conduction Disease. PLoS Genet. 8 (12): e1003154.
PMID: 23236297  DOI: 10.1371/journal.pgen.1003154  PII: PGENETICS-D-12-02668
2003  
BMC medical genetics

  

BACKGROUND: Mutations in the gene encoding the nuclear membrane protein lamin A/C have been associated with at least 7 distinct diseases including autosomal dominant dilated cardiomyopathy withconduction system disease, autosomal dominant and recessive Emery Dreifuss Muscular Dystrophy, limb girdle muscular dystrophy type 1B, autosomal recessive type 2 Charcot Marie Tooth, mandibuloacral dysplasia, familial partial lipodystrophy and Hutchinson-Gilford progeria.METHODS: We used mutation detection to evaluate the lamin A/C gene in a 45 year-old woman with familial dilated cardiomyopathy and conduction system disease whose family has been well characterized for this phenotype 1.RESULTS: DNA from the proband was analyzed, and a novel 2 base-pair deletion c.908_909delCT in LMNA was identified.CONCLUSIONS: Mutations in the gene encoding lamin A/C can lead to significant cardiac conductionsystem disease that can be successfully treated with pacemakers and/or defibrillators. Genetic screening can help assess risk for arrhythmia and need for device implantation.
(MacLeod et al. 2003)
MacLeod HM, Culley MR, et al. (2003). Lamin A/C truncation in dilated cardiomyopathy with conduction disease. BMC Med. Genet. 4: 4.
PMID: 12854972  DOI: 10.1186/1471-2350-4-4
2012  
Heart (British Cardiac Society)

  

(MacRae 2012)
MacRae CA (2012). Pattern recognition: combining informatics and genetics to re-evaluate conduction disease. Heart 98 (17): 1263-4.
PMID: 22875820  DOI: 10.1136/heartjnl-2012-302408  PII: heartjnl-2012-302408
2004  
The anatomical record. Part A, Discoveries in molecular, cellular, and evolutionary biology

  

Atrioventricular (AV) conduction disease (block) describes impairment of the electrical continuity between the atria and ventricles. Classification of AV block has utilized biophysical characteristics, usually the extent (first, second, or third degree) and site of block (above or below His bundle recording site). The genetic significance of this classification is unknown. In young patients, AV block may result from injury or be the major cardiac manifestation of neuromuscular disease. However, in some cases, AV block has unknown or idiopathic cause. In such cases, familial clustering has been noted and published pedigrees show autosomal dominant inheritance; associated heart disease is common (e.g., congenital heart malformation, cardiomyopathy). The latter finding is not surprising given the common origin of working myocytes and specialized conduction system elements. Using genetic models incorporating reduced penetrance (disease absence in some individuals with diseasegene), variable expressivity (individuals with disease gene have different phenotypes), and genetic heterogeneity (similar phenotypes, different genetic cause), molecular genetic causes of AV block are being identified. Mutations identified in genes with diverse functions (transcription, excitability, and energy homeostasis) for the first time provide the means to assess risk and offer insight into the molecular basis of this important clinical condition previously defined only by biophysical characteristics.
(Benson 2004) – ORIGINAL FIRST PAPER on the Subject
Benson DW (2004). Genetics of atrioventricular conduction disease in humans. Anat Rec A Discov Mol Cell Evol Biol 280 (2): 934-9.
PMID: 15372490  DOI: 10.1002/ar.a.20099
SOURCE:
SOURCE for REFERENCES Listed below:
  1. 1.
    1. Benson DW,
    2. Wang DW,
    3. Dyment M,
    4. Knilans TK,
    5. Fish FA,
    6. Strieper MJ,
    7. Rhodes TH,
    8. George AL Jr.

    . Congenital sick sinus syndrome caused by recessive mutations in the cardiac sodium channel gene (SCN5A). J Clin Invest. 2003;112:1019–1028.

  2. 2.
    1. Makiyama T,
    2. Akao M,
    3. Tsuji K,
    4. Doi T,
    5. Ohno S,
    6. Takenaka K,
    7. Kobori A,
    8. Ninomiya T,
    9. Yoshida H,
    10. Takano M,
    11. Makita N,
    12. Yanagisawa F,
    13. Higashi Y,
    14. Takeyama Y,
    15. Kita T,
    16. HorieM

    . High risk for bradyarrhythmic complications in patients with Brugada syndrome caused by SCN5A gene mutations. J Am Coll Cardiol. 2005;46:2100–2106.

  3. 3.
    1. Groenewegen WA,
    2. Firouzi M,
    3. Bezzina CR,
    4. Vliex S,
    5. van Langen IM,
    6. Sandkuijl L,
    7. Smits JP,
    8. Hulsbeek M,
    9. Rook MB,
    10. Jongsma HJ,
    11. Wilde AA

    . A cardiac sodium channel mutation cosegregates with a rare connexin40 genotype in familial atrial standstill. Circ Res. 2003;92:14–22.

  4. 4.
    1. Veldkamp MW,
    2. Wilders R,
    3. Baartscheer A,
    4. Zegers JG,
    5. Bezzina CR,
    6. Wilde AA

    .Contribution of sodium channel mutations to bradycardia and sinus node dysfunction in LQT3 families. Circ Res. 2003;92:976–983.

  5. 5.
    1. Probst V,
    2. Allouis M,
    3. Sacher F,
    4. Pattier S,
    5. Babuty D,
    6. Mabo P,
    7. Mansourati J,
    8. VictorJ,
    9. Nguyen JM,
    10. Schott JJ,
    11. Boisseau P,
    12. Escande D,
    13. Le Marec H

    . Progressive cardiac conduction defect is the prevailing phenotype in carriers of a Brugada syndrome SCN5A mutation. J Cardiovasc Electrophysiol. 2006;17:270–275.

  6. 6.
    1. Probst V,
    2. Kyndt F,
    3. Potet F,
    4. Trochu JN,
    5. Mialet G,
    6. Demolombe S,
    7. Schott JJ,
    8. Baro I,
    9. Escande D,
    10. Le Marec H

    . Haploinsufficiency in combination with aging causes SCN5A-linked hereditary Lenegre disease. J Am Coll Cardiol. 2003;41:643–652.

  7. 7.
    1. Kyndt F,
    2. Probst V,
    3. Potet F,
    4. Demolombe S,
    5. Chevallier JC,
    6. Baro I,
    7. Moisan JP,
    8. Boisseau P,
    9. Schott JJ,
    10. Escande D,
    11. Le Marec H

    . Novel SCN5A mutation leading either to isolated cardiac conduction defect or Brugada syndrome in a large French family.Circulation. 2001;104:3081–3086.

  8. 8.
    1. Schott JJ,
    2. Alshinawi C,
    3. Kyndt F,
    4. Probst V,
    5. Hoorntje TM,
    6. Hulsbeek M,
    7. Wilde AA,
    8. Escande D,
    9. Mannens MM,
    10. Le Marec H

    . Cardiac conduction defects associate with mutations in SCN5A. Nat Genet. 1999;23:20–21.

  9. 9.
    1. Watanabe H,
    2. Koopmann TT,
    3. Le Scouarnec S,
    4. Yang T,
    5. Ingram CR,
    6. Schott JJ,
    7. Demolombe S,
    8. Probst V,
    9. Anselme F,
    10. Escande D,
    11. Wiesfeld AC,
    12. Pfeufer A,
    13. Kaab S,
    14. Wichmann HE,
    15. Hasdemir C,
    16. Aizawa Y,
    17. Wilde AA,
    18. Roden DM,
    19. Bezzina CR

    . Sodium channel beta1 subunit mutations associated with Brugada syndrome and cardiac conduction disease in humans. J Clin Invest. 2008;118:2260–2268.

  10. 10.
    1. Andelfinger G,
    2. Tapper AR,
    3. Welch RC,
    4. Vanoye CG,
    5. George AL Jr.,
    6. Benson DW

    .KCNJ2 mutation results in Andersen syndrome with sex-specific cardiac and skeletal muscle phenotypes. Am J Hum Genet. 2002;71:663–668.

  11. 11.
    1. Plaster NM,
    2. Tawil R,
    3. Tristani-Firouzi M,
    4. Canun S,
    5. Bendahhou S,
    6. Tsunoda A,
    7. Donaldson MR,
    8. Iannaccone ST,
    9. Brunt E,
    10. Barohn R,
    11. Clark J,
    12. Deymeer F,
    13. George AL Jr.,
    14. Fish FA,
    15. Hahn A,
    16. Nitu A,
    17. Ozdemir C,
    18. Serdaroglu P,
    19. Subramony SH,
    20. Wolfe G,
    21. Fu YH,
    22. Ptacek LJ

    . Mutations in Kir2.1 cause the developmental and episodic electrical phenotypes of Andersen’s syndrome. Cell. 2001;105:511–519.

  12. 12.
    1. Schulze-Bahr E,
    2. Neu A,
    3. Friederich P,
    4. Kaupp UB,
    5. Breithardt G,
    6. Pongs O,
    7. IsbrandtD

    . Pacemaker channel dysfunction in a patient with sinus node disease. J Clin Invest.2003;111:1537–1545.

  13. 13.
    1. Milanesi R,
    2. Baruscotti M,
    3. Gnecchi-Ruscone T,
    4. DiFrancesco D

    . Familial sinus bradycardia associated with a mutation in the cardiac pacemaker channel. N Engl J Med.2006;354:151–157.

  14. 14.
    1. Nof E,
    2. Luria D,
    3. Brass D,
    4. Marek D,
    5. Lahat H,
    6. Reznik-Wolf H,
    7. Pras E,
    8. Dascal N,
    9. Eldar M,
    10. Glikson M

    . Point mutation in the HCN4 cardiac ion channel pore affecting synthesis, trafficking, and functional expression is associated with familial asymptomatic sinus bradycardia. Circulation. 2007;116:463–470.

  15. 15.
    1. Bhuiyan ZA,
    2. van den Berg MP,
    3. van Tintelen JP,
    4. Bink-Boelkens MT,
    5. Wiesfeld AC,
    6. Alders M,
    7. Postma AV,
    8. van Langen I,
    9. Mannens MM,
    10. Wilde AA

    . Expanding spectrum of human RYR2-related disease: new electrocardiographic, structural, and genetic features.Circulation. 2007;116:1569–1576.

  16. 16.
    1. Postma AV,
    2. Denjoy I,
    3. Kamblock J,
    4. Alders M,
    5. Lupoglazoff JM,
    6. Vaksmann G,
    7. Dubosq-Bidot L,
    8. Sebillon P,
    9. Mannens MM,
    10. Guicheney P,
    11. Wilde AA

    . Catecholaminergic polymorphic ventricular tachycardia: RYR2 mutations, bradycardia, and follow up of the patients. J Med Genet. 2005;42:863–870.

  17. 17.
    1. Postma AV,
    2. Denjoy I,
    3. Hoorntje TM,
    4. Lupoglazoff JM,
    5. Da Costa A,
    6. Sebillon P,
    7. Mannens MM,
    8. Wilde AA,
    9. Guicheney P

    . Absence of calsequestrin 2 causes severe forms of catecholaminergic polymorphic ventricular tachycardia. Circ Res. 2002;91;e21–e26.

  18. 18.
    1. Basson CT,
    2. Bachinsky DR,
    3. Lin RC,
    4. Levi T,
    5. Elkins JA,
    6. Soults J,
    7. Grayzel D,
    8. Kroumpouzou E,
    9. Traill TA,
    10. Leblanc-Straceski J,
    11. Renault B,
    12. Kucherlapati R,
    13. Seidman JG,
    14. Seidman CE

    . Mutations in human TBX5 [corrected] cause limb and cardiac malformation in Holt-Oram syndrome. Nat Genet. 1997;15:30–35.

  19. 19.
    1. Schott JJ,
    2. Benson DW,
    3. Basson CT,
    4. Pease W,
    5. Silberbach GM,
    6. Moak JP,
    7. MaronBJ,
    8. Seidman CE,
    9. Seidman JG

    . Congenital heart disease caused by mutations in the transcription factor NKX2-5. Science. 1998;281:108–111.

  20. 20.
    1. Bonne G,
    2. Di Barletta MR,
    3. Varnous S,
    4. Becane HM,
    5. Hammouda EH,
    6. Merlini L,
    7. Muntoni F,
    8. Greenberg CR,
    9. Gary F,
    10. Urtizberea JA,
    11. Duboc D,
    12. Fardeau M,
    13. Toniolo D,
    14. Schwartz K

    . Mutations in the gene encoding lamin A/C cause autosomal dominant Emery-Dreifuss muscular dystrophy. Nat Genet. 1999;21:285–288.

  21. 21.
    1. Le Scouarnec S,
    2. Bhasin N,
    3. Vieyres C,
    4. Hund TJ,
    5. Cunha SR,
    6. Koval O,
    7. MarionneauC,
    8. Chen B,
    9. Wu Y,
    10. Demolombe S,
    11. Song LS,
    12. Le Marec H,
    13. Probst V,
    14. Schott JJ,
    15. AndersonME,
    16. Mohler PJ

    . Dysfunction in ankyrin-B-dependent ion channel and transporter targeting causes human sinus node disease. Proc Natl Acad Sci U S A. 2008;105:15617–15622.

  22. 22.
    1. Gollob MH,
    2. Green MS,
    3. Tang AS,
    4. Gollob T,
    5. Karibe A,
    6. Ali Hassan AS,
    7. Ahmad F,
    8. Lozado R,
    9. Shah G,
    10. Fananapazir L,
    11. Bachinski LL,
    12. Roberts R

    . Identification of a gene responsible for familial Wolff-Parkinson-White syndrome. N Engl J Med.2001;344:1823–1831.

  23. 23.
    1. Arad M,
    2. Benson DW,
    3. Perez-Atayde AR,
    4. McKenna WJ,
    5. Sparks EA,
    6. Kanter RJ,
    7. McGarry K,
    8. Seidman JG,
    9. Seidman CE

    . Constitutively active AMP kinase mutations cause glycogen storage disease mimicking hypertrophic cardiomyopathy. J Clin Invest.2002;109:357–362.

  24. 24.
    1. Lalani SR,
    2. Thakuria JV,
    3. Cox GF,
    4. Wang X,
    5. Bi W,
    6. Bray MS,
    7. Shaw C,
    8. Cheung SW,
    9. Chinault AC,
    10. Boggs BA,
    11. Ou Z,
    12. Brundage EK,
    13. Lupski JR,
    14. Gentile J,
    15. Waisbren S,
    16. PursleyA,
    17. Ma L,
    18. Khajavi M,
    19. Zapata G,
    20. Friedman R,
    21. Kim JJ,
    22. Towbin JA,
    23. Stankiewicz P,
    24. Schnittger S,
    25. Hansmann I,
    26. Ai T,
    27. Sood S,
    28. Wehrens XH,
    29. Martin JF,
    30. Belmont JW,
    31. PotockiL

    . 20p12.3 Microdeletion predisposes to Wolff-Parkinson-White syndrome with variable neurocognitive deficits. J Med Genet. 2009;46:168–175.

  25. 25.
    1. Mahadevan M,
    2. Tsilfidis C,
    3. Sabourin L,
    4. Shutler G,
    5. Amemiya C,
    6. Jansen G,
    7. NevilleC,
    8. Narang M,
    9. Barcelo J,
    10. O’Hoy K,
    11. Leblond S,
    12. Earle-Macdonald J,
    13. De Jong PJ,
    14. WieringaB,
    15. Korneluk RG

    . Myotonic dystrophy mutation: an unstable CTG repeat in the 3′ untranslated region of the gene. Science. 1992;255:1253–1255.

  26. 26.
    1. Anderson KR,
    2. Ho SY,
    3. Anderson RH

    . Location and vascular supply of sinus node in human heart. Br Heart J. 1979;41:28–32.

  27. 27.
    1. Lei M,
    2. Zhang H,
    3. Grace AA,
    4. Huang CL

    . SCN5A and sinoatrial node pacemaker function. Cardiovasc Res. 2007;74:356–365.

  28. 28.
    1. Lei M,
    2. Jones SA,
    3. Liu J,
    4. Lancaster MK,
    5. Fung SS,
    6. Dobrzynski H,
    7. Camelliti P,
    8. MaierSK,
    9. Noble D,
    10. Boyett MR

    . Requirement of neuronal- and cardiac-type sodium channels for murine sinoatrial node pacemaking. J Physiol. 2004;559:835–848.

  29. 29.
    1. Zhang H,
    2. Holden AV,
    3. Kodama I,
    4. Honjo H,
    5. Lei M,
    6. Varghese T,
    7. Boyett MR

    .Mathematical models of action potentials in the periphery and center of the rabbit sinoatrial node. Am J Physiol. 2000;279;H397–H421.

  30. 30.
    1. Dobrzynski H,
    2. Li J,
    3. Tellez J,
    4. Greener ID,
    5. Nikolski VP,
    6. Wright SE,
    7. Parson SH,
    8. Jones SA,
    9. Lancaster MK,
    10. Yamamoto M,
    11. Honjo H,
    12. Takagishi Y,
    13. Kodama I,
    14. Efimov IR,
    15. Billeter R,
    16. Boyett MR

    . Computer three-dimensional reconstruction of the sinoatrial node.Circulation. 2005;111:846–854.

  31. 31.
    1. Liu J,
    2. Noble PJ,
    3. Xiao G,
    4. Abdelrahman M,
    5. Dobrzynski H,
    6. Boyett MR,
    7. Lei M,
    8. NobleD

    . Role of pacemaking current in cardiac nodes: insights from a comparative study of sinoatrial node and atrioventricular node. Prog Biophys Mol Biol. 2008;96:294–304.

  32. 32.
    1. Liu J,
    2. Dobrzynski H,
    3. Yanni J,
    4. Boyett MR,
    5. Lei M

    . Organisation of the mouse sinoatrial node: structure and expression of HCN channels. Cardiovasc Res.2007;73:729–738.

  33. 33.
    1. Chen PS,
    2. Joung B,
    3. Shinohara T,
    4. Das M,
    5. Chen Z,
    6. Lin SF

    . The initiation of the heart beat. Circ J. 2010;74:221–225.

  34. 34.
    1. Huser J,
    2. Blatter LA,
    3. Lipsius SL

    . Intracellular Ca2+ release contributes to automaticity in cat atrial pacemaker cells. J Physiol. 2000;524();415–422.

  35. 35.
    1. Bogdanov KY,
    2. Vinogradova TM,
    3. Lakatta EG

    . Sinoatrial nodal cell ryanodine receptor and Na(+)-Ca(2+) exchanger: molecular partners in pacemaker regulation. Circ Res. 2001;88:1254–1258.

  36. 36.
    1. Vinogradova TM,
    2. Bogdanov KY,
    3. Lakatta EG

    . β-Adrenergic stimulation modulates ryanodine receptor Ca(2+) release during diastolic depolarization to accelerate pacemaker activity in rabbit sinoatrial nodal cells. Circ Res. 2002;90:73–79.

  37. 37.
    1. Lakatta EG,
    2. Maltsev VA,
    3. Vinogradova TM

    . A coupled SYSTEM of intracellular Ca2+ clocks and surface membrane voltage clocks controls the timekeeping mechanism of the heart’s pacemaker. Circ Res. 2010;106:659–673.

  38. 38.
    1. Joung B,
    2. Ogawa M,
    3. Lin SF,
    4. Chen PS

    . The calcium and voltage clocks in sinoatrial node automaticity. Korean Circ J. 2009;39:217–222.

  39. 39.
    1. Ruan Y,
    2. Liu N,
    3. Priori SG

    . Sodium channel mutations and arrhythmias. Nat Rev Cardiol. 2009;6:337–348.

  40. 40.
    1. Smits JP,
    2. Koopmann TT,
    3. Wilders R,
    4. Veldkamp MW,
    5. Opthof T,
    6. Bhuiyan ZA,
    7. Mannens MM,
    8. Balser JR,
    9. Tan HL,
    10. Bezzina CR,
    11. Wilde AA

    . A mutation in the human cardiac sodium channel (E161K) contributes to sick sinus syndrome, conduction disease and Brugada syndrome in two families. J Mol Cell Cardiol. 2005;38:969–981.

  41. 41.
    1. Butters TD,
    2. Aslanidi OV,
    3. Inada S,
    4. Boyett MR,
    5. Hancox JC,
    6. Lei M,
    7. Zhang H

    .Mechanistic links between Na+ channel (SCN5A) mutations and impaired cardiac pacemaking in sick sinus syndrome. Circ Res. 2010;107:126–137.

  42. 42.
    1. Verheijck EE,
    2. van Kempen MJ,
    3. Veereschild M,
    4. Lurvink J,
    5. Jongsma HJ,
    6. BoumanLN

    . Electrophysiological features of the mouse sinoatrial node in relation to connexin distribution. Cardiovasc Res. 2001;52:40–50.

  43. 43.
    1. Fedorov VV,
    2. Schuessler RB,
    3. Hemphill M,
    4. Ambrosi CM,
    5. Chang R,
    6. Voloshina AS,
    7. Brown K,
    8. Hucker WJ,
    9. Efimov IR

    . Structural and functional evidence for discrete exit pathways that connect the canine sinoatrial node and atria. Circ Res. 2009;104:915–923.

  44. 44.
    1. Herrmann S,
    2. Stieber J,
    3. Ludwig A

    . Pathophysiology of HCN channels. Pflugers Arch. 2007;454:517–522.

  45. 45.
    1. Biel M,
    2. Wahl-Schott C,
    3. Michalakis S,
    4. Zong X

    . Hyperpolarization-activated cation channels: from genes to function. Physiol Rev. 2009;89:847–885.

  46. 46.
    1. Stieber J,
    2. Herrmann S,
    3. Feil S,
    4. Loster J,
    5. Feil R,
    6. Biel M,
    7. Hofmann F,
    8. Ludwig A

    . The hyperpolarization-activated channel HCN4 is required for the generation of pacemaker action potentials in the embryonic heart. Proc Natl Acad Sci U S A.2003;100:15235–15240.

  47. 47.
    1. Ludwig A,
    2. Herrmann S,
    3. Hoesl E,
    4. Stieber J

    . Mouse models for studying pacemaker channel function and sinus node arrhythmia. Prog Biophys Mol Biol. 2008;98:179–185.

  48. 48.
    1. Herrmann S,
    2. Stieber J,
    3. Stockl G,
    4. Hofmann F,
    5. Ludwig A

    . HCN4 provides a ‘depolarization reserve’ and is not required for heart rate acceleration in mice. EMBO J.2007;26:4423–4432.

  49. 49.
    1. Hoesl E,
    2. Stieber J,
    3. Herrmann S,
    4. Feil S,
    5. Tybl E,
    6. Hofmann F,
    7. Feil R,
    8. Ludwig A

    .Tamoxifen-inducible gene deletion in the cardiac conduction system. J Mol Cell Cardiol.2008;45:62–69.

  50. 50.
    1. Rubenstein DS,
    2. Lipsius SL

    . Mechanisms of automaticity in subsidiary pacemakers from cat right atrium. Circ Res. 1989;64:648–657.

  51. 51.
    1. Ju YK,
    2. Allen DG

    . How does beta-adrenergic stimulation increase the heart rate? The role of intracellular Ca2+ release in amphibian pacemaker cells. J Physiol.1999;516();793–804.

  52. 52.
    1. Rizzi N,
    2. Liu N,
    3. Napolitano C,
    4. Nori A,
    5. Turcato F,
    6. Colombi B,
    7. Bicciato S,
    8. Arcelli D,
    9. Spedito A,
    10. Scelsi M,
    11. Villani L,
    12. Esposito G,
    13. Boncompagni S,
    14. Protasi F,
    15. Volpe P,
    16. PrioriSG

    . Unexpected structural and functional consequences of the R33Q homozygous mutation in cardiac calsequestrin: a complex arrhythmogenic cascade in a knock in mouse model. Circ Res. 2008;103:298–306.

  53. 53.
    1. Liu N,
    2. Colombi B,
    3. Memmi M,
    4. Zissimopoulos S,
    5. Rizzi N,
    6. Negri S,
    7. Imbriani M,
    8. Napolitano C,
    9. Lai FA,
    10. Priori SG

    . Arrhythmogenesis in catecholaminergic polymorphic ventricular tachycardia: insights from a RyR2 R4496C knock-in mouse model. Circ Res.2006;99:292–298.

  54. 54.
    1. Mohler PJ,
    2. Schott JJ,
    3. Gramolini AO,
    4. Dilly KW,
    5. Guatimosim S,
    6. duBell WH,
    7. SongLS,
    8. Haurogne K,
    9. Kyndt F,
    10. Ali ME,
    11. Rogers TB,
    12. Lederer WJ,
    13. Escande D,
    14. Le Marec H,
    15. Bennett V

    . Ankyrin-B mutation causes type 4 long-QT cardiac arrhythmia and sudden cardiac death. Nature. 2003;421:634–639.

  55. 55.
    1. Zhang L,
    2. Benson DW,
    3. Tristani-Firouzi M,
    4. Ptacek LJ,
    5. Tawil R,
    6. Schwartz PJ,
    7. George AL,
    8. Horie M,
    9. Andelfinger G,
    10. Snow GL,
    11. Fu YH,
    12. Ackerman MJ,
    13. Vincent GM

    .Electrocardiographic features in Andersen-Tawil syndrome patients with KCNJ2 mutations: characteristic T-U-wave patterns predict the KCNJ2 genotype. Circulation.2005;111:2720–2726.

  56. 56.
    1. Zaritsky JJ,
    2. Redell JB,
    3. Tempel BL,
    4. Schwarz TL

    . The consequences of disrupting cardiac inwardly rectifying K(+) current (I(K1)) as revealed by the targeted deletion of the murine Kir2.1 and Kir2.2 genes. J Physiol. 2001;533:697–710.

  57. 57.
    1. McLerie M,
    2. Lopatin AN

    . Dominant-negative suppression of I(K1) in the mouse heart leads to altered cardiac excitability. J Mol Cell Cardiol. 2003;35:367–378.

  58. 58.
    1. Light PE

    . Familial Wolff-Parkinson-White syndrome: a disease of glycogen storage or ion channel dysfunction? J Cardiovasc Electrophysiol. 2006;17():S158–S161.

  59. 59.
    1. Peters NS,
    2. Rowland E,
    3. Bennett JG,
    4. Green CR,
    5. Anderson RH,
    6. Severs NJ

    . The Wolff-Parkinson-White syndrome: the cellular substrate for conduction in the accessory atrioventricular pathway. Eur Heart J. 1994;15:981–987.

  60. 60.
    1. Vidaillet HJ Jr.,
    2. Pressley JC,
    3. Henke E,
    4. Harrell FE Jr.,
    5. German LD

    . Familial occurrence of accessory atrioventricular pathways (preexcitation syndrome). N Engl J Med. 1987;317:65–69.

  61. 61.
    1. Arad M,
    2. Moskowitz IP,
    3. Patel VV,
    4. Ahmad F,
    5. Perez-Atayde AR,
    6. Sawyer DB,
    7. WalterM,
    8. Li GH,
    9. Burgon PG,
    10. Maguire CT,
    11. Stapleton D,
    12. Schmitt JP,
    13. Guo XX,
    14. Pizard A,
    15. Kupershmidt S,
    16. Roden DM,
    17. Berul CI,
    18. Seidman CE,
    19. Seidman JG

    . Transgenic mice overexpressing mutant PRKAG2 define the cause of Wolff-Parkinson-White syndrome in glycogen storage cardiomyopathy. Circulation. 2003;107:2850–2856.

  62. 62.
    1. Wolf CM,
    2. Arad M,
    3. Ahmad F,
    4. Sanbe A,
    5. Bernstein SA,
    6. Toka O,
    7. Konno T,
    8. Morley G,
    9. Robbins J,
    10. Seidman JG,
    11. Seidman CE,
    12. Berul CI

    . Reversibility of PRKAG2 glycogen-storage cardiomyopathy and electrophysiological manifestations. Circulation.2008;117:144–154.

  63. 63.
    1. Gaussin V,
    2. Van de Putte T,
    3. Mishina Y,
    4. Hanks MC,
    5. Zwijsen A,
    6. Huylebroeck D,
    7. Behringer RR,
    8. Schneider MD

    . Endocardial cushion and myocardial defects after cardiac myocyte-specific conditional deletion of the bone morphogenetic protein receptor ALK3.Proc Natl Acad Sci U S A. 2002;99:2878–2883.

  64. 64.
    1. Stroud DM,
    2. Gaussin V,
    3. Burch JB,
    4. Yu C,
    5. Mishina Y,
    6. Schneider MD,
    7. Fishman GI,
    8. Morley GE

    . Abnormal conduction and morphology in the atrioventricular node of mice with atrioventricular canal targeted deletion of Alk3/Bmpr1a receptor. Circulation.2007;116:2535–2543.

  65. 65.
    1. Gaussin V,
    2. Morley GE,
    3. Cox L,
    4. Zwijsen A,
    5. Vance KM,
    6. Emile L,
    7. Tian Y,
    8. Liu J,
    9. HongC,
    10. Myers D,
    11. Conway SJ,
    12. Depre C,
    13. Mishina Y,
    14. Behringer RR,
    15. Hanks MC,
    16. Schneider MD,
    17. Huylebroeck D,
    18. Fishman GI,
    19. Burch JB,
    20. Vatner SF

    . Alk3/Bmpr1a receptor is required for development of the atrioventricular canal into valves and annulus fibrosus. Circ Res.2005;97:219–226.

  66. 66.
    1. Hoffman JI,
    2. Kaplan S

    . The incidence of congenital heart disease. J Am Coll Cardiol. 2002;39:1890–1900.

  67. 67.
    1. Walsh SR,
    2. Tang T,
    3. Wijewardena C,
    4. Yarham SI,
    5. Boyle JR,
    6. Gaunt ME

    .Postoperative arrhythmias in general surgical patients. Ann R Coll Surg Engl.2007;89:91–95.

  68. 68.
    1. Walsh EP,
    2. Cecchin F

    . Arrhythmias in adult patients with congenital heart disease.Circulation. 2007;115:534–545.

  69. 69.
    1. Gourdie RG,
    2. Mima T,
    3. Thompson RP,
    4. Mikawa T

    . Terminal diversification of the myocyte lineage generates Purkinje fibers of the cardiac conduction system.Development. 1995;121:1423–1431.

  70. 70.
    1. Cheng G,
    2. Litchenberg WH,
    3. Cole GJ,
    4. Mikawa T,
    5. Thompson RP,
    6. Gourdie RG

    .Development of the cardiac conduction system involves recruitment within a multipotent cardiomyogenic lineage. Development. 1999;126:5041–5049.

  71. 71.
    1. Moorman AF,
    2. Christoffels VM

    . Development of the cardiac conduction system: a matter of chamber development. Novartis Found Symp. 2003;250:25–34;discussion 34–43, 276–279.

  72. 72.
    1. Bruneau BG,
    2. Nemer G,
    3. Schmitt JP,
    4. Charron F,
    5. Robitaille L,
    6. Caron S,
    7. Conner DA,
    8. Gessler M,
    9. Nemer M,
    10. Seidman CE,
    11. Seidman JG

    . A murine model of Holt-Oram syndrome defines roles of the T-box transcription factor Tbx5 in cardiogenesis and disease. Cell. 2001;106:709–721.

  73. 73.
    1. Christoffels VM,
    2. Smits GJ,
    3. Kispert A,
    4. Moorman AF

    . Development of the pacemaker tissues of the heart. Circ Res. 2010;106:240–254.

  74. 74.
    1. Moskowitz IP,
    2. Kim JB,
    3. Moore ML,
    4. Wolf CM,
    5. Peterson MA,
    6. Shendure J,
    7. NobregaMA,
    8. Yokota Y,
    9. Berul C,
    10. Izumo S,
    11. Seidman JG,
    12. Seidman CE

    . A molecular pathway including Id2, Tbx5, and Nkx2-5 required for cardiac conduction system development.Cell. 2007;129:1365–1376.

  75. 75.
    1. Stennard FA,
    2. Harvey RP

    . T-box transcription factors and their roles in regulatory hierarchies in the developing heart. Development. 2005;132:4897–4910.

  76. 76.
    1. Moskowitz IP,
    2. Pizard A,
    3. Patel VV,
    4. Bruneau BG,
    5. Kim JB,
    6. Kupershmidt S,
    7. RodenD,
    8. Berul CI,
    9. Seidman CE,
    10. Seidman JG

    . The T-box transcription factor Tbx5 is required for the patterning and maturation of the murine cardiac conduction system. Development.2004;131:4107–4116.

  77. 77.
    1. Ranganayakulu G,
    2. Elliott DA,
    3. Harvey RP,
    4. Olson EN

    . Divergent roles for NK-2 class homeobox genes in cardiogenesis in flies and mice. Development.1998;125:3037–3048.

  78. 78.
    1. Tanaka M,
    2. Chen Z,
    3. Bartunkova S,
    4. Yamasaki N,
    5. Izumo S

    . The cardiac homeobox gene Csx/Nkx2.5 lies genetically upstream of multiple genes essential for heart development. Development. 1999;126:1269–1280.

  79. 79.
    1. Jay PY,
    2. Harris BS,
    3. Maguire CT,
    4. Buerger A,
    5. Wakimoto H,
    6. Tanaka M,
    7. KupershmidtS,
    8. Roden DM,
    9. Schultheiss TM,
    10. O’Brien TX,
    11. Gourdie RG,
    12. Berul CI,
    13. Izumo S

    . Nkx2-5 mutation causes anatomic hypoplasia of the cardiac conduction system. J Clin Invest.2004;113:1130–1137.

  80. 80.
    1. Wakimoto H,
    2. Kasahara H,
    3. Maguire CT,
    4. Izumo S,
    5. Berul CI

    . Developmentally modulated cardiac conduction failure in transgenic mice with fetal or postnatal overexpression of DNA nonbinding mutant Nkx2.5. J Cardiovasc Electrophysiol.2002;13:682–688.

  81. 81.
    1. Kasahara H,
    2. Wakimoto H,
    3. Liu M,
    4. Maguire CT,
    5. Converso KL,
    6. Shioi T,
    7. Huang WY,
    8. Manning WJ,
    9. Paul D,
    10. Lawitts J,
    11. Berul CI,
    12. Izumo S

    . Progressive atrioventricular conduction defects and heart failure in mice expressing a mutant Csx/Nkx2.5 homeoprotein. J Clin Invest. 2001;108:189–201.

  82. 82.
    1. Hsu DT

    . Cardiac manifestations of neuromuscular disorders in children. Paediatr Respir Rev. 2010;11:35–38.

  83. 83.
    1. Holaska JM

    . Emerin and the nuclear lamina in muscle and cardiac disease. Circ Res. 2008;103:16–23.

  84. 84.
    1. Arimura T,
    2. Helbling-Leclerc A,
    3. Massart C,
    4. Varnous S,
    5. Niel F,
    6. Lacene E,
    7. FromesY,
    8. Toussaint M,
    9. Mura AM,
    10. Keller DI,
    11. Amthor H,
    12. Isnard R,
    13. Malissen M,
    14. Schwartz K,
    15. Bonne G

    . Mouse model carrying H222P-Lmna mutation develops muscular dystrophy and dilated cardiomyopathy similar to human striated muscle laminopathies. Hum Mol Genet. 2005;14:155–169.

  85. 85.
    1. Wolf CM,
    2. Wang L,
    3. Alcalai R,
    4. Pizard A,
    5. Burgon PG,
    6. Ahmad F,
    7. Sherwood M,
    8. Branco DM,
    9. Wakimoto H,
    10. Fishman GI,
    11. See V,
    12. Stewart CL,
    13. Conner DA,
    14. Berul CI,
    15. Seidman CE,
    16. Seidman JG

    . Lamin A/C haploinsufficiency causes dilated cardiomyopathy and apoptosis-triggered cardiac conduction system disease. J Mol Cell Cardiol.2008;44:293–303.

  86. 86.
    1. Fatkin D,
    2. MacRae C,
    3. Sasaki T,
    4. Wolff MR,
    5. Porcu M,
    6. Frenneaux M,
    7. Atherton J,
    8. Vidaillet HJ Jr.,
    9. Spudich S,
    10. De Girolami U,
    11. Seidman JG,
    12. Seidman C,
    13. Muntoni F,
    14. MuehleG,
    15. Johnson W,
    16. McDonough B

    . Missense mutations in the rod domain of the lamin A/C gene as causes of dilated cardiomyopathy and conduction-system disease. N Engl J Med. 1999;341:1715–1724.

  87. 87.
    1. Mounkes LC,
    2. Kozlov SV,
    3. Rottman JN,
    4. Stewart CL

    . Expression of an LMNA-N195K variant of A-type lamins results in cardiac conduction defects and death in mice.Hum Mol Genet. 2005;14:2167–2180.

  88. 88.
    1. Muchir A,
    2. Pavlidis P,
    3. Decostre V,
    4. Herron AJ,
    5. Arimura T,
    6. Bonne G,
    7. Worman HJ

    .Activation of MAPK pathways links LMNA mutations to cardiomyopathy in Emery-Dreifuss muscular dystrophy. J Clin Invest. 2007;117:1282–1293.

  89. 89.
    1. Muchir A,
    2. Shan J,
    3. Bonne G,
    4. Lehnart SE,
    5. Worman HJ

    . Inhibition of extracellular signal-regulated kinase signaling to prevent cardiomyopathy caused by mutation in the gene encoding A-type lamins. Hum Mol Genet. 2009;18:241–247.

  90. 90.
    1. Morgenlander JC,
    2. Nohria V,
    3. Saba Z

    . EKG abnormalities in pediatric patients with myotonic dystrophy. Pediatr Neurol. 1993;9:124–126.

  91. 91.
    1. Berul CI,
    2. Maguire CT,
    3. Aronovitz MJ,
    4. Greenwood J,
    5. Miller C,
    6. Gehrmann J,
    7. Housman D,
    8. Mendelsohn ME,
    9. Reddy S

    . DMPK dosage alterations result in atrioventricular conduction abnormalities in a mouse myotonic dystrophy model. J Clin Invest. 1999;103;R1–R7.

  92. 92.
    1. Wakimoto H,
    2. Maguire CT,
    3. Sherwood MC,
    4. Vargas MM,
    5. Sarkar PS,
    6. Han J,
    7. ReddyS,
    8. Berul CI

    . Characterization of cardiac conduction system abnormalities in mice with targeted disruption of Six5 gene. J Interv Card Electrophysiol. 2002;7:127–135.

  93. 93.
    1. Mahadevan MS,
    2. Yadava RS,
    3. Yu Q,
    4. Balijepalli S,
    5. Frenzel-McCardell CD,
    6. BourneTD,
    7. Phillips LH

    . Reversible model of RNA toxicity and cardiac conduction defects in myotonic dystrophy. Nat Genet. 2006;38:1066–1070.

  94. 94.
    1. Hardy J,
    2. Singleton A

    . Genomewide association studies and human disease. N Engl J Med. 2009;360:1759–1768.

  95. 95.
    A haplotype map of the human genome. Nature. 2005;437:1299–1320.
  96. 96.
    1. Holm H,
    2. Gudbjartsson DF,
    3. Arnar DO,
    4. Thorleifsson G,
    5. Thorgeirsson G,
    6. Stefansdottir H,
    7. Gudjonsson SA,
    8. Jonasdottir A,
    9. Mathiesen EB,
    10. Njolstad I,
    11. Nyrnes A,
    12. Wilsgaard T,
    13. Hald EM,
    14. Hveem K,
    15. Stoltenberg C,
    16. Lochen ML,
    17. Kong A,
    18. Thorsteinsdottir U,
    19. Stefansson K

    . Several common variants modulate heart rate, PR interval and QRS duration. Nat Genet. 2010;42:117–122.

  97. 97.
    1. Pfeufer A,
    2. van Noord C,
    3. Marciante KD,
    4. Arking DE,
    5. Larson MG,
    6. Smith AV,
    7. Tarasov KV,
    8. Muller M,
    9. Sotoodehnia N,
    10. Sinner MF,
    11. Verwoert GC,
    12. Li M,
    13. Kao WH,
    14. KottgenA,
    15. Coresh J,
    16. Bis JC,
    17. Psaty BM,
    18. Rice K,
    19. Rotter JI,
    20. Rivadeneira F,
    21. Hofman A,
    22. Kors JA,
    23. Stricker BH,
    24. Uitterlinden AG,
    25. van Duijn CM,
    26. Beckmann BM,
    27. Sauter W,
    28. Gieger C,
    29. LubitzSA,
    30. Newton-Cheh C,
    31. Wang TJ,
    32. Magnani JW,
    33. Schnabel RB,
    34. Chung MK,
    35. Barnard J,
    36. SmithJD,
    37. Van Wagoner DR,
    38. Vasan RS,
    39. Aspelund T,
    40. Eiriksdottir G,
    41. Harris TB,
    42. Launer LJ,
    43. Najjar SS,
    44. Lakatta E,
    45. Schlessinger D,
    46. Uda M,
    47. Abecasis GR,
    48. Muller-Myhsok B,
    49. EhretGB,
    50. Boerwinkle E,
    51. Chakravarti A,
    52. Soliman EZ,
    53. Lunetta KL,
    54. Perz S,
    55. Wichmann HE,
    56. Meitinger T,
    57. Levy D,
    58. Gudnason V,
    59. Ellinor PT,
    60. Sanna S,
    61. Kaab S,
    62. Witteman JC,
    63. Alonso A,
    64. Benjamin EJ,
    65. Heckbert SR

    . Genome-wide association study of PR interval. Nat Genet.2010;42:153–159.

  98. 98.
    1. Chambers JC,
    2. Zhao J,
    3. Terracciano CM,
    4. Bezzina CR,
    5. Zhang W,
    6. Kaba R,
    7. Navaratnarajah M,
    8. Lotlikar A,
    9. Sehmi JS,
    10. Kooner MK,
    11. Deng G,
    12. Siedlecka U,
    13. ParasramkaS,
    14. El-Hamamsy I,
    15. Wass MN,
    16. Dekker LR,
    17. de Jong JS,
    18. Sternberg MJ,
    19. McKenna W,
    20. Severs NJ,
    21. de Silva R,
    22. Wilde AA,
    23. Anand P,
    24. Yacoub M,
    25. Scott J,
    26. Elliott P,
    27. Wood JN,
    28. Kooner JS

    . Genetic variation in SCN10A influences cardiac conduction. Nat Genet.2010;42:149–152.

  99. 99.
    1. Cho HC,
    2. Marban E

    . Biological therapies for cardiac arrhythmias: can genes and cells replace drugs and devices? Circ Res. 2010;106:674–685.

  100. 100.
    1. Edelberg JM,
    2. Aird WC,
    3. Rosenberg RD

    . Enhancement of murine cardiac chronotropy by the molecular transfer of the human beta(2) adrenergic receptor cDNA. J Clin Invest. 1998;101:337–343.

  101. 101.
    1. Edelberg JM,
    2. Huang DT,
    3. Josephson ME,
    4. Rosenberg RD

    . Molecular enhancement of porcine cardiac chronotropy. Heart. 2001;86:559–562.

  102. 102.
    1. Miake J,
    2. Marban E,
    3. Nuss HB

    . Gene therapy: biological pacemaker created by gene transfer. Nature. 2002;419:132–133.

  103. 103.
    1. Ruhparwar A,
    2. Kallenbach K,
    3. Klein G,
    4. Bara C,
    5. Ghodsizad A,
    6. Sigg DC,
    7. Karck M,
    8. Haverich A,
    9. Niehaus M

    . Adenylate-cyclase VI transforms ventricular cardiomyocytes into biological pacemaker cells. Tissue Eng Part A. 2010;16:1867–1872.

  104. 104.
    1. Sastry A,
    2. Arnold E,
    3. Gurji H,
    4. Iwasa A,
    5. Bui H,
    6. Hassankhani A,
    7. Patel HH,
    8. Feramisco JR,
    9. Roth DM,
    10. Lai NC,
    11. Hammond HK,
    12. Narayan SM

    . Cardiac-directed expression of adenylyl cyclase VI facilitates atrioventricular nodal conduction. J Am Coll Cardiol. 2006;48:559–565.

  105. 105.
    1. Qu JH,
    2. Plotnikov AN,
    3. Danilo P,
    4. Shlapakova I,
    5. Cohen IS,
    6. Robinson RB,
    7. RosenMR

    . Expression and function of a biological pacemaker in canine heart. Circulation.2003;107:1106–1109.

  106. 106.
    1. Plotnikov AN,
    2. Sosunov EA,
    3. Qu JH,
    4. Shlapakova IN,
    5. Anyukhovsky EP,
    6. Liu LL,
    7. Janse MJ,
    8. Brink PR,
    9. Cohen IS,
    10. Robinson RB,
    11. Danilo P,
    12. Rosen MR

    . Biological pacemaker implanted in canine left bundle branch provides ventricular escape rhythms that have physiologically acceptable rates. Circulation. 2004;109:506–512.

  107. 107.
    1. Plotnikov AN,
    2. Shlapakova I,
    3. Szabolcs MJ,
    4. Danilo P,
    5. Lorell BH,
    6. Potapova IA,
    7. LuZJ,
    8. Rosen AB,
    9. Mathias RT,
    10. Brink PR,
    11. Robinson RB,
    12. Cohen IS,
    13. Rosen MR

    . Xenografted adult human mesenchymal stem cells provide a platform for sustained biological pacemaker function in canine heart. Circulation. 2007;116:706–713.

Other related articles published on this Open Access Online Scientific Journal, include the following:

Aviva Lev-Ari, PhD, RN and Larry H. Bernstein, MD, FCAP

Read Full Post »