Review Open Access
Copyright ©The Author(s) 2016. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Cardiol. Feb 26, 2016; 8(2): 180-191
Published online Feb 26, 2016. doi: 10.4330/wjc.v8.i2.180
Genetic testing in congenital heart disease: A clinical approach
Marie A Chaix, Paul Khairy, Montreal Heart Institute Adult Congenital Center, Université de Montréal, Montreal, Quebec H1T 1C8, Canada
Gregor Andelfinger, Sainte-Justine University Hospital, Université de Montréal, Montreal, Quebec H1T 1C8, Canada
Author contributions: Chaix MA, Andelfinger G and Khairy P fulfill ICMJE authorship criteria based on substantial contributions to conception and design, acquisition of data, or analysis and interpretation of data; drafting the article or revising it critically for important intellectual content; and final approval of the version to be published.
Supported by A Canada research chair in electrophysiology and adult congenital heart disease (Paul Khairy).
Conflict-of-interest statement: The authors declare no conflicts of interest regarding this manuscript.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Dr. Paul Khairy, Montreal Heart Institute Adult Congenital Center, Université de Montréal, 5000 Belanger St. E, Montreal, Quebec H1T 1C8, Canada. paul.khairy@umontreal.ca
Telephone: +1-514-3763330 Fax: +1-514-5932551
Received: June 8, 2015
Peer-review started: June 10, 2015
First decision: September 18, 2015
Revised: October 16, 2015
Accepted: December 9, 2015
Article in press: December 11, 2015
Published online: February 26, 2016
Processing time: 259 Days and 2.8 Hours

Abstract

Congenital heart disease (CHD) is the most common type of birth defect. Traditionally, a polygenic model defined by the interaction of multiple genes and environmental factors was hypothesized to account for different forms of CHD. It is now understood that the contribution of genetics to CHD extends beyond a single unified paradigm. For example, monogenic models and chromosomal abnormalities have been associated with various syndromic and non-syndromic forms of CHD. In such instances, genetic investigation and testing may potentially play an important role in clinical care. A family tree with a detailed phenotypic description serves as the initial screening tool to identify potentially inherited defects and to guide further genetic investigation. The selection of a genetic test is contingent upon the particular diagnostic hypothesis generated by clinical examination. Genetic investigation in CHD may carry the potential to improve prognosis by yielding valuable information with regards to personalized medical care, confidence in the clinical diagnosis, and/or targeted patient follow-up. Moreover, genetic assessment may serve as a tool to predict recurrence risk, define the pattern of inheritance within a family, and evaluate the need for further family screening. In some circumstances, prenatal or preimplantation genetic screening could identify fetuses or embryos at high risk for CHD. Although genetics may appear to constitute a highly specialized sector of cardiology, basic knowledge regarding inheritance patterns, recurrence risks, and available screening and diagnostic tools, including their strengths and limitations, could assist the treating physician in providing sound counsel.

Key Words: Congenital heart disease, Genetics, Genetic screening, Genetic testing

Core tip: Monogenic models and chromosomal abnormalities have been associated with syndromic and non-syndromic forms of congenital heart disease (CHD), paving the way for genetic investigation and testing to shoulder an important role in patient management. Herein, we present an overview of the role of genetics in CHD, propose various clinical scenarios in which genetic testing may be appropriate, and discuss practical implications with regards to when and how to order genetic tests. Summary tables are provided regarding the various genes implicated in syndromic and non-syndromic forms of CHD and recurrence risks in siblings and offspring.



INTRODUCTION

Congenital heart disease (CHD) afflicts 2 to 3 children per 100 live births[1,2]. It is the most common type of birth defect and encompasses a wide range of malformations. The spectrum of severity ranges from insignificant and even self-resolving lesions, such as ventricular septal defects that spontaneously close, to highly complex and multiorgan manifestations that are incompatible with natural survival. While much progress has been made regarding the management of children and adults with CHD, a greater understanding of underlying etiologies could potentially lead to further advances in preventive care and therapeutic strategies[3].

The complexity and heterogeneity of CHD has traditionally been attributed to multifactorial etiologies arising from interactions between multiple genes and environmental factors (so-called “polygenic model”)[4]. Early investigations into environmental factors spawned recommendations for maternal multivitamin supplementation containing folic acid to reduce risks of developing CHD[5-7]. Other implicated maternal factors include pregestational diabetes, pollakiuria, febrile illnesses, rubeola, influenza, alcohol consummation, cigarette smoking, and teratogenic pharmacological agents such as thalidomide, warfarin, angiotensin converting enzyme inhibitors, and certain anticonvulsant and anti-inflammatory drugs[8].

Technological advances have permitted the confirmation of clinically suspected monogenetic subtypes of CHD, with dominant or recessive inheritance patterns. However, some forms of CHD could not be explained by a polygenic model[9], with much higher recurrence risks in first-degree relatives than predicted[3,10]. Chromosomal abnormalities have been associated with cardiac defects, particularly in the setting of syndromic phenotypes (e.g., trisomy 21, DiGeorge, and Williams-Beuren syndromes). In so-called multiplex families with several affected members, identified candidate genes have been consistent with monogenetic models with Mendelian inheritance. Furthermore, the rate of CHD increases with consanguinity, as described in Arabic countries[11].

The fact that monogenic and chromosomal abnormality models account for a substantial proportion of CHD enhances the potential value of genetic investigation and testing[12]. Genetics carries the potential to unravel etiological mysteries that underpin CHD, provide pathophysiological insights, assist in risk assessment, inform clinical management, and counsel families regarding future offspring. The focus of this review is on the genetics of structural CHD, as opposed to other disease categories such as inherited channelopathies. We review known implicated genes and chromosomal abnormalities, discussed when and how to perform genetic testing, and shared our perspective regarding clinical applications.

GENETICS IN STRUCTURAL CONGENITAL HEART DISEASES

Approximately 30% of patients diagnosed with CHD have syndromic phenotypes with extracardiac manifestations. The influence of genetics is well established for chromosomal aneuploidies such as Down, Turner, and DiGeorge syndromes. Other syndromes are linked to a mutation or deletion in one gene, such as Noonan, Alagille, and Holt-Oram syndromes[3]. For the 70% of CHD cases that are non-syndromic, new genes with Mendelian inheritance (dominant or recessive) have been identified, particularly in families with several affected members. Table 1 summarizes current knowledge regarding genetic etiology for several forms of CHD with syndromic or non-syndromic phenotypes.

Table 1 Genes described in syndromic and non-syndromic forms of congenital heart disease.
Syndromic: Syndrome name Non syndromic: Gene implicatedPhenotype associated with structural heart diseaseSyndromic: Chromosomal aneuploidy, microdeletion or gene/locus/inheritance Non syndromic: Locus/inheritanceRef.
Atrioventricular septal defect (AVSD)
Down syndrome1MR, facial dysmorphiaTrisomy 21[71-73]
Edward syndromeIUGR, facial dysmorphia, clenched fingersTrisomy 18[74]
Patau syndromeCleft lip and palate, microphthalmia, polydactylyTrisomy 13[74]
Holt-Oram syndromePreaxial limb defects, absent or dysmorphic thumbs, cardiac conduction diseaseTBX5/12q24.1/AD[75]
Noonan syndromeHypertrophic cardiomyopathy, short stature, broad neck, unusual chest shape, facial dysmorphia, developmental delayPTPN11/12q24/AD, de novo; SOS1/2p21/AD, de novo; KRAS /12p12.1/AD, de novo[76]
Ellis-van Creveld syndromeCommon atrium, polydactyly, deformity of upper lip, dwarfism with narrow thorax, ASVD partial to completeEVC and EVC2/4p16/AR[77]
Locus 1p31-p21AVSD partial to completeGene no yet found/AD[78]
CRELD1Partial AVSD, heterotaxy syndrome3p25/AD[79-81]
GATA4Family with ASD, VSD and one member with AVSD8p23.1/AD, de novo[82]
Atrial septal defect (ASD)
Holt Oram syndromeSee AVSD aboveSee above AVSD[83,84]
Noonan syndromeSee AVSD aboveSee above AVSD[85,86]
Ellis-van Creveld syndromeSee AVSD aboveSee above AVSD[77]
Cardiofaciocutaneous syndromeHypertrophic cardiomyopathy, facial dysmorphia, skin abnormalities: keratosis pilaris, neviMAP2K1/15q22.31/AD, de novo; MAP2K2 /19p13.3/AD, de novo; KRAS/7q34/AD, de novo; BRAF/12p12.1/AD, de novo[87]
Cri du ChatSound of cry similar to cat’s cry, facial dysmorphia, MRCTNND2/5p15.2/de novo[88]
NK2X-5+/- Atrioventricular block5q35.1/AD[89]
GATA4+/- Pulmonary stenosis8p23.1/AD, de novo[82,90]
MYH614q11.2/AD[13]
TBX207p14.2/AD[14,91]
Ventricular septal defect (VSD)
Holt Oram syndromeSee AVSD aboveSee AVSD above[92]
Ellis-van Creveld syndromeSee AVSD aboveSee AVSD above[77]
Cri du chatSee AVSD aboveSee AVSD above[88]
Down syndromeSee AVSD aboveSee AVSD above[93,94]
Edward/Patau syndromeSee AVSD aboveSee AVSD above[95,96]
DiGeorge syndromeFacial dysmorphia, speech delay, learning delay, psychiatric disorder, cleft palate, immune deficiency, hypoplastic/aplastic thymus, hypocalcaemia.Deletion 22q11.21/de novo, AD[97]
NK2X-5Atrioventricular block5q35.1/AD[89]
GATA48p23.1/AD, de novo[82,98]
Ebstein anomaly
Down syndromeSee AVSD aboveSee AVSD above[99]
NKX2-55q35.1/AD[100]
Pulmonary stenosis
Noonan syndrome1See ASVD aboveSee AVSD above[101,102]
Costello syndromeHypertrophic cardiomyopathy, MR, loose skin, facial dysmorphia large mouthHRAS/11p15.5/AD[103]
Leopard syndromeLentigines, short stature, hearing loss, similar to Noonan syndromePTPN11/12q24/AD, de novo; RAF1/3p25.2 /AD, de novo; BRAF/7q34/AD, de novo[104]
Alagille syndromePulmonary branch stenosis, bile duct paucity, cholestasis, facial dysmorphia, deep-set eyes, butterfly vertebraeJAG1/20p12/AD; NOTCH2/1p12/AD; Deletion/20p12/AD[105,106]
Cardiofaciocutaneous syndromeSee ASD aboveSee ASD above[107]
GATA4+/- Atrial septal defect8p23.1/AD, de novo[90, 108]
Aortic valve stenosis
Turner syndrome1Female, webbed neck, widely spaced nipples, short stature, streaked ovariesMonosomy X or mosaics (45,X/46,XX)[109]
Noonan syndromeSee above AVSDSee above AVSD[76]
NOTCH19q34.3/AD[16, 110]
SMAD615q22.31/?[111]
Supravalvular aortic stenosis
Williams-Beuren syndrome1Elfin facies, cocktail personality, hypercalcaemia, developmental delay, thyroid disorder, renal and connective tissue abnormalities.Deletion/7q11.23/de novo, AD[112]
Aortic coarctation
Turner syndrome1See aortic valve stenosis aboveSee aortic valve stenosis above[109]
Down/Edward/Patau syndromeSee AVSD aboveSee AVSD above[113]
NOTCH19q34.3/AD[110]
Bicuspid aortic valve
Turner syndromeSee aortic valve stenosis aboveSee aortic valve stenosis above[109]
Anderson syndromeLong QT syndrome, ventricular arrhythmias, sudden cardiac death, facial dysmorphia, short statureKCNJ2/17q24.3/AD[114]
NOTCH19q34.3/AD[16,110]
SMAD615q22.31/?[111]
Tetralogy of Fallot
DiGeorge syndrome1See VSD aboveSee VSD above[97]
Alagille syndromeSee pulmonary stenosis aboveSee pulmonary stenosis above[105,106]
Cat-Eye syndromeDysmorphic ears, microphthalmia, anal atresia, renal abnormalities, coloboma, cleft palateDuplication/22q11/de novo[115]
NKX2.55q35.1/AD[100]
GATA48p23.1/AD, de novo[116]
NOTCH19q34.3/AD[16]
FOG28q23.1/?[117,118]
Truncus arteriosus
DiGeorge syndromeSee VSD aboveSee VSD above[97]
Hypoplastic left heart syndrome
NKX2.55q35.1/AD[119]
NOTCH19q34.3/AD[16,110]

Genes etiologically linked to CHD directly impact embryologic development. For example, defects in genes responsible for the embryonic formation of the atrial septum (e.g., MYH6, TBX20) can result in atrial septal defects (ASD)[13,14]. In addition to their function in embryologic cardiac development, implicated genes may also play a role in heart regulation throughout life[15]. The critical purpose of these genes, which are primarily transcription factors, explains the possibility of dominant heritability. A mutation that modifies the protein function in one of these genes may have a major effect on cardiac development and regulation. Furthermore, interactions between transcription factors explain the diverse consequences associated with individual mutations. For example, NKX2.5 mutations may result in ASD, atrioventricular block, ventricular septal defect (VSD), Ebstein anomaly, and tetrology of fallot (TOF). GATA4, a transcription factor, has been associated with ASD, VSD, and pulmonary stenosis. TBX1 has been implicated in TOF, patent ductus arteriosus, and interrupted aortic arch; and TBX20 in ASD, VSD, valve defects, and impaired chamber growth. In addition to these transcription factors, other genes with varied roles have been implicated in CHD, such as MYH6, which codes for an alpha myosin heavy chain (ASD) and Notch 1, which is implicated in valve formation (bicuspid aortic valve and aortic stenosis)[15,16].

WHEN AND HOW TO PERFORM A GENETIC INVESTIGATION?
When to consider genetic testing

The first clinical situation to consider genetic testing in CHD is the presence of a syndromic phenotype. A comprehensive clinical examination is paramount in recognzing extracardiac involvement. Common physical findings include facial dysmorphia (eye, ear, mouth, nose abnormalities), limb dysmorphia (atrophy, length reduction), hand and feet dysmorphia (polydactyly, short fingers, clinodactyly), and other skeletal abnormalities such as scoliosis[17]. Growth delays may be identified by monitoring height and weight and neurological status must be assessed to diagnose mental impairment and learning disabilities. Other organs must be screened to exclude associated gastrointestinal, urologic, and genital defects. Thus, a thorough investigation often involves a multidisciplinary approach including a neurologist, ophthalmologist, otolaryngologist, gastrointestinal specialist, and orthopedic surgeon. Additional paraclinical testing may be guided by the clinical examination: radiography, abdominal ultrasound, cerebral imaging, and laboratory testing (liver and renal function, and others depending on the clinical examination). While investigating newborns in the intensive care unit can be particularly difficult, it is important to identify defects that may benefit from early surgical intervention. Variable expressivity adds a further layer of complexity justifying a broader screening approach. For example, it has been recommended to screen all children with supra-valvular aortic stenosis or pulmonary stenosis for Williams-Beuren syndrome and those with an interrupted aortic arch, truncus arteriosus, TOF, VSD with aortic arch anomaly, isolated aortic arch anomaly, or discontinuous branch pulmonary arteries for DiGeorge syndrome[17]. In general, genetic consultation is recommended when a probable syndromic phenotype is identified.

The second clinical situation to consider genetic testing is in the context of a multiplex family, i.e., a family in which a person diagnosed with CHD has an afflicted first- or second-degree relative. A comprehensive clinical investigation includes a detailed assesment of past medical, surgical, and family histories. A family history can point to a genetically transmitted disease and is important in understanding inheritance patterns (autosomal recessive, dominant, X-linked, and mitochondrial), penetrance, and expressivity of genetic variations. While some have advocated exhaustive family history questionnaires[18], basic themes include screening for cardiac diseases within families, particular phenotypes such as dysmorphias, aborted pregnancies, other birth defects, infertility, and early deaths. Importantly, in some families with CHD, different phenotypes may be expressed such as a bicuspid aortic valve in one family member and hypoplastic left heart syndrome in another. The origin of all four grandparents may provide relevant information, such as the potential for consanguinity. If positive elements are detected, a detailed family tree should be performed that includes each proband’s first- and second-degree relatives. The family tree may be further expanded, depending of which side of the family has diseased members. Supportive documents, such as surgical and autopsy reports, should be sought. It is also important to update family pedigrees to include new events over time.

In summary, the phenotypic description associated with the family tree is an essential tool in guiding further genetic investigation. Identification of a clinical feature related to an established syndrome associated with CHD should prompt syndrome-specific investigation. Wider scale screening is recommended on the basis of variable expressivity for syndromes such as Williams-Beuren and DiGeorge. For non-syndromic CHD, the family tree may orient the clinician towards a genetic etiology and a specific pattern of inheritance. Nevertheless, in the majority of cases, there are no known karyotype abnormalities to investigate. Currently, genetic testing of known cardiac candidate genes is not routinely recommended in the clinical setting. However, genetic testing of multiplex families in the context of research studies may identify novel mutations in known genes or entirely new causal genes.

Choosing a genetic test

An individualized approach to genetic testing begins with the diagnostic hypotheses elicited by a thorough clinical assessment. In general, chromosomal abnormalities represent changes in the structure or number of chromosomes and are diagnosed by cytogenetic methods. The standard metaphase karyotype analysis detects numerical and structural chromosomal aberrations with a resolution of 5 megabases. It is indicated to search for such anomalies as trisomy 21, 18, 13, or monosomy X. Fluorescence in situ hybridization (FISH) is a method to detect deletion or duplication of specific regions of DNA using targeted probes. It provides a higher resolution than karyotype and is the predominant technique used to identify Williams-Beuren, DiGeorge, and Alagille syndromes. Subtelomere FISH analyses, while less commonly used today, provide a high resolution to detect abnormalities in subtelomere (i.e., DNA segments between telomeric caps and chromatin) and telomere (i.e., regions of repetive nucleotide sequences at each end of a chromatid) DNA regions[17]. Subtelomeric anomalies have been reported in patients with a syndromic phenotype associated with facial dysmorphia and mental retardation combined with CHD such as VSD, ASD, pulmonary stenosis, and right sided aortic arch[19,20].

Array-based comparative genomic hybridization (aCGH) is used to detect unbalanced structural and numerical chromosomal abnormalities with a resolution inferior to 5 megabases, such as copy number variants (CNV), i.e., number of copies of a particular gene that deviate from normal (two for autosomes, one X chromosome for males (XY), and two X chromosomes for females (XX). This molecular karyotype provides rapid identification of duplications/deletions, unbalanced translocations, and aneuploidies. This method analyzes the entire genome and compares it to controls, in contrast to FISH techniques that target specific DNA regions. It may be particularly useful when a probable chromosomal syndrome is identified but the karyotype is normal and there is no known specific region to test[21]. Furthermore, this method is of additional value in detecting CNVs such as in screening for DiGeorge syndrome when the karyotype and 22q11 microdeletion analyses by FISH are unrevealing[20]. Cytogenetic testing has been recommended for all children with CHD associated with mental retardation, developmental delay, dysmorphic features, or other organ involvement and for establishing a prenatal diagnosis when CHD is identified by fetal echocardiography[17]. Most CNV studies in CHD report 10%-25% of abnormal findings across the disease spectrum.

Gene mutations represent a second category of genetic abnormalities. Mutations can affect the coding portion of a gene, a case in which interpretation is usually straightforward. They can also affect the non-coding portion of the genome, in which case they are more difficult to interpret. With the advent of NextGeneration sequencing technologies, large gene panels, which specifically target genes that are known or suspected to play a role in cardiac biology, can be more readily screened than previously possible by Sanger sequencing[22,23]. This approach affords a high quality diagnosis. Gene sequencing can be helpful in conditions such Noonan syndrome, Alagille syndrome with a normal FISH analysis, Holt-Oram syndrome, and several other diseases.

Interpretation of a genetic test

When a genetic variation is diagnosed, the clinician must determine its relation to the phenotype. Although genetic variants are identified with increasing frequency by high throughput sequencing, not all variants are pathogenic[22,23]. Determination of pathogenic potential is based on the following three questions: (1) Has this genetic variant already been described in association with the particular phenotype? (2) Is the genetic variant predicted to alter gene function or regulation, gene coding, or the gene splice site, and does it occur in an evolutionarily conserved nucleotide? and (3) Does the genetic variant segregate with the affected family members and not unaffected members or controls? This assessment is not foolproof. For example, genetic variants may be identified in unaffected family members because of variable penetrance and expressivity. Each genetic result must, therefore, be placed in context of the clinical and family evaluation.

Genetic counseling

Genetic counseling is important before and after genetic testing[24]. Prior to testing, the patient or guarantor should be informed of the risks of a negative result arising from the fact that all genes implicated in a given phenotype have not been identified. Second, the pathogenic potential of a genetic variant may be difficult to determine. Third, if a genetic familial disorder is identified, the patient is responsible for informing the family. After genetic testing, counseling is important to review the results, explain the genetic variant, and discuss implications with the patient and family[25].

OBJECTIVES OF GENETIC TESTING IN CLINICAL PRACTICE
Confidence in the diagnosis

Objectives of genetic testing may vary according to the clinical scenario. One objective is to establish confidence in the diagnosis. An accurate diagnosis could allow the clinician to explain causes and mechanisms of disease, provide more precise prognostic information, and elucidate implications for future offspring. Genetic counseling is of paramount importance in relaying such information[26].

Appropriate management

Non-cardiac organ involvement: An accurate diagnosis could alert the clinician to the possibility of associated non-cardiac organ involvement. Down, Patau, Edward, DiGeorge, Turner, Williams-Beuren, Noonan, and Alagille syndromes all involve extracardiac abnormalities[27].

Craniofacial anomalies have been associated with endocardial cushion defect, truncus arteriosus, and aortic arch anomalies; respiratory disease with endocardial cushion defect and pulmonary valve disease; genitourinary malformations with septal defects, pulmonary valve disease, aortic valve disease, and truncus arteriosus; and situs inversus with heterotaxy and endocardial cushion defect[27].

Establishing a genetic diagnosis could help orient clinical and paraclinical investigations and subspecialty referral for all potential organs involved. Unrecognized and untreated interactions between various organ pathologies could worsen the cardiac prognosis. Identification of a genetic syndrome may also prove useful in the event of an emergency, when a frequent complication associated with a given syndrome occurs. Moreover, recognition of a syndrome provides a more defined guide for follow-up, including surveillance and screening for reported complications.

Other associated cardiac complications: In addition to the genetic origins of CHD, genetic variations can modulate the propensity to develop associated cardiac complications, such as arrhythmias[28] and heart failure[29,30]. Transcription factors play a key role in the formation of cardiac structures and maintenance of cardiac function and, conversely, their dysregulation can have multifaceted manifestations. For example, in the setting of an ASD, those with an NKX2.5 syndrome are more likely to develop atrioventricular block and progressive ventricular dysfunction[28]. Interestingly, patients with NKX2.5 mutations can also develop dilated cardiomyopathy[31]. TBX5, a gene implicated in Holt-Oram syndrome (septation defects, atrioventricular node disease, and upper limb defects) also modulates diastolic function[32]. Genes implicated in RASopathy syndromes responsible for Noonan, Leopard, cardiofaciocutaneous, and Costello syndromes are also responsible for cardiac hypertrophy in later development[33]. Thus, the genetic environment could modulate the prognosis of various forms of CHD, help to elucidate risks of developing conduction defects and systolic and diastolic dysfunction, and a provide a basis to adapt follow-up accordingly.

Overlap of CHD with muscular heart disease: Structural CHD and cardiomyopathy may be modulated by the same mutations that give rise to varied phenotypes within the same family. For example, some family members with a TBX20 mutation may have an underlying ASD, VSD, or mitral valve disease or may present exclusively with pulmonary hypertension or cardiomyopathy[14]. Mutations in MYH6 (alpha-cardiac myosin heavy chain) are associated with various forms of CHD but also dilated and hypertrophic cardiomyopathy[34]. Moreover, mutations in MYH7 have been reported in patients with Ebstein anomaly and left ventricular noncompaction[35,36]. Some family members may have CHD whereas others could develop progressive cardiomyopathy or electrophysiologic disorders. Thus, if a mutation is discovered in a family with a discordant phenotype, clinical screening and genetic testing can identify seemingly phenotypically normal individuals who are at risk of developing cardiomyopathy or electrophysiologic manifestations.

Prognosis: In addition to the prognostic implications of genetic factors discussed above, certain gene defects have been associated with post-operative survival and long-term outcomes. For example, endothelin-1 G5665T has been associated with transplant-free survival in patients with single ventricles, primarily hypoplastic left heart syndrome[37]. This variant is linked to increased vascular reactivity and hypertension. Similarly, in a study of genetic variants involved in vascular response and oxidative stress, two major alleles of two single nucleotide polymorphisms (SNPs; i.e., VEGFA rs833069 and SOD2 rs2758331) were associated with worse transplant-free survival in patients with non-syndromic CHD[38]. The higher number of copies of deleterious alleles, the worse the prognosis[38]. Genotype has also been associated with early postoperative outcomes. For example, in patients with TOF, 22q11.2 deletion (DiGeorge syndrome) predicts a longer cardiopulmonary bypass time and a greater length of stay in intensive care[39]. While several explanations have been proposed, potential factors include a higher prevalence of aortopulmonary shunts and respiratory problems prior to surgical repair in patients with 22q11.2 deletion, resulting in longer mechanical ventilatory support. Conceivably, a SNP profile may one day prove to be of value in pre-operative risk assessment.

Therapeutic potential: Ultimately, the holy grail of genetically diagnosing CHD is to provide targeted curative therapy. While such interventions are currently beyond our reach, provocative studies support its potential. For example, a knock-out model of Wnt2 in null mutants results in a phenotype resembling complete atrioventricular septal defect[40,41]. The phenotype could be rescued in vivo by pharmacological activation of Wnt signalling.

Genetics and recurrence risk

With a Mendelian pattern of inheritance, recurrence risks are 50% and 25% for autosomally dominant and recessive genes, respectively. However, variable penetrance complicates these predictions, even for syndromic CHD. In the majority of cases with CHD, difficulties in estimating recurrence risks are compounded by the absence of a clear genetic diagnosis[42,43]. Estimates are, therefore, largely based on a detailed family tree and the published literature[18].

In patients with atrial septal defects, the recurrence risk has been estimated to be 3% in first-degree relatives, although a dominant inheritance pattern has been described in some families. A CHD recurrence risk of 1.2% was reported for first-degree relatives with an isolated septal defect[10]. For probands with atrioventricular septal defects, the prevalence of any CHD in a family member appears to be in the order of 12%-15% overall, 1%-2% of parents, 2%-4% of siblings, and 10%-14% of offspring[44-46]. Risk of recurrence is greater if the mother rather than the father has the atrioventricular septal defect (i.e., 14% vs 10%). Nevertheless, exact figures remain debated with some studies reporting considerably lower risks[10]. In TOF, the recurrence risk has been estimated to be 2.5%-3% overall, with a phenotype that is often concordant[44,47]. However, the recurrence risk in offspring is higher when the mother is affected[9]. Moreover, some families without a 22q11 deletion syndrome have been suspected of having a recessive inheritance pattern[48]. In complete transposition of the great arteries, a very low recurrence risk has been described with no offspring affected in a British collaborative study, suggesting a sporadic model[9]. Other studies have reported a recurrence risk of 1.8% in siblings[49] and 2.7% in first-degree relatives (siblings and parents)[50], which includes varied forms of CHD such as aortic valve stenosis and double outlet right ventricle[50]. In patients with congenitally corrected transposition of the great arteries, a 5.2% recurrence risk was reported in siblings, with concordant and discordant phenotypes, including complete transposition of the great arteries, suggesting that some genes may be common to both types of transposition[51].

Left-sided obstructive lesions (e.g., aortic coarctation, hypoplastic left heart syndrome, aortic stenosis, bicuspid aortic valve, and hypoplastic aortic arch) may segregate within families, suggesting a common genetic basis[52,53]. Overall recurrence risks have ranged from 1.8% to 3.2% of siblings, 3% of offspring of affected fathers, and 8% to 13% of offspring of affected mothers[52]. However, much higher recurrence risks have been described in certain geographic locations, such as 37% of first-degree relatives in Texas[52]. Moreover, some defects appear to have higher recurrence risks, such as aortic coarctation (13% of siblings)[50], hypoplastic left heart syndrome (31% of siblings)[50], and bicuspid aortic valves (> 10% of siblings)[54,55]. However, considerably lower recurrence risks for left-sided obstructive lesions in first degree relatives have also been reported (e.g., 0.79% with a relative risk of 12.9)[10].

As noted by the examples above, estimating recurrence risks is an imperfect science. Empiric estimates consider the mathematical prediction of recurrence in a polygenic model of inheritance combined with the type of CHD, current knowledge base, and relationship to proband. As a general rule of thumb, recurrence risks are in the order of 1% to 6% for siblings of affected probands with unaffected parents and increase to approximately 10% when two siblings are affected. Recurrence risks in offspring are greater than siblings, higher if the proband is the mother[3], and generally higher for left-sided obstructive lesions (8%-10%). A recent population-based study from Denmark challenges these statistics and provides far lower estimates for first-degree relatives than previously reported, as summarized in Table 2[10]. These disparate results could be explained, in part, by differences in the study designs and methodologies employed, and underscore the difficulties in accurately quantifying recurrence risks. Estimating recurrence risks must consider an in depth analysis of the family history to identify specific patterns of inheritance. If the pedigree is not informative and estimates are based on a polygenic model of inheritance, limitations of empiric estimates should be discussed with the patients, including the possibility of under- or overestimation. The notion of concordant or discordant recurrent phenotypes should also be conveyed. Overall, exact concordance is low for left-sided obstructive defects (26%), intermediate for outflow tract defects (37%), and higher for septal defects (48%)[43]. Conceptually, CHD may be grouped into constellations of malformations such as septal defects, conotroncal anomalies, and left-sided obstructive lesions that share implicated genes, although such a concept is not universally supported[56].

Table 2 Recurrence risks for non-syndromic congenital heart disease in first-degree relatives.
Type of non-syndromic CHDRecurrence risk of same CHD in first-degree relatives (%)Recurrence risk of discordant CHD in first-degree relatives (%)Recurrence risk of any CHD in first-degree relatives (%)
ASVD1.102.23.30
ASD0.882.43.28
VSD0.671.92.57
ASD and VSD0.242.22.44
Conotruncal defect11.302.43.70
Right ventricular outflow tract obstruction21.703.04.70
Left sided obstructions30.792.43.19
Assessing family members

A strong case has been made for screening first-degree relatives of patients with left-sided obstructive lesions and bicuspid aortic valves. As previously noted, recurrent phenotypes in first-degree relatives are relatively common and frequently discordant such that a bicuspid aortic valve, aortic coarctation, and/or aortic dilation may be identified in asymptomatic family members. Echocardiographic screening has been recommended for first-degree relatives of patients with bicuspid aortic valve or supra-aortic stenosis, since a physical examination alone lacks sensitivity[57]. The rationale for family screening is that early detection may help avert complications related to aortic dilatation (e.g., 6-fold higher risk of aortic dissection), aortic stenosis, aortic insufficiency, endocarditis, and aortic coarctation (e.g., arterial hypertension). Early detection may lead to lifestyle recommendations (e.g., limit isometric exercises), enhanced monitoring (e.g., for progressive aortic dilatation), or preventive surgery (e.g., prior to aortic dissection). Age at screening remains controversial. It should generally be proposed to adults if not previously performed during childhood.

At present, systematic screening of first-degree relatives is not recommended for other forms of non-syndromic CHD. However, fetal echocardiographic screening is indicated if either parent is afflicted with any form of CHD. It should be performed in a specialized center at 18-20 wk of gestation[58]. Early detection of complex CHD can drastically improve outcomes by planning delivery in a specialized (level 3) tertiary care center with appropriate monitoring and early catheter-based or surgical interventions when indicated[58-60]. Furthermore, prenatal diagnosis may lead to a parental decision to terminate the pregnancy.

Finally, identification of a specific mutation in a multiplex family with CHD may allow for targeted screening of additional family members. While there is no clear-cut indication for genetic screening to identify CHD in family members with structurally normal hearts, there may be a rationale to screen seemingly normal family members for entities that include CHD as one aspect of a multiple constellation phenotype.

Prenatal diagnosis

The impact of a prenatal diagnosis of CHD on the pregnancy termination rate varies by region. For example, reported pregnancy termination rates for severe CHD identified by prenatal screening were 45% in the Netherlands[58], 49% in Boston, MA[61], and 86% in Switzerland[62]. In a study from France, factors associated with pregnancy termination included severity of CHD, gestational age at diagnosis, presence of chromosomal abnormalities, and parental ethnicity[63].

Fetal genetic screening for CHD is also possible, including genome-wide high-resolution SNP arrays to identify CNVs[64] and competitive genomic hybridization to detect submicroscopic chromosomal aberrations[65]. A prenatal diagnostic test can be performed after chorionic villus sampling before 14 wk of gestation. Thus far, such testing has been limited to specific disease entities such as trisomy 21, 18, and 13, cystic fibrosis, and microdeletion syndromes (e.g., DiGeorge). It could also be performed for any severe monogenetic disease if the result could influence the decision to terminate pregnancy[66]. Preimplantation diagnostic testing could be proposed in selected cases, particularly for women with a history of multiple therapeutic abortions. It has already been used for Holt Oram and Marfan syndromes[67]. Beyond syndromes such as trisomy 21, 18 or 13, prenatal or preimplantation genetic screening remains controversial. Ethical dilemmas may arise as a result of uncertainties in interpreting tests, potential for false positives, and the inability to predict disease severity, penetrance and expressivity of a mutation, and concordant or discordant phenotypes.

Limitation of genetics in CHD

Despite the fact that CHD is the most common birth defect, the genetic etiology remains unknown in the majority of cases, with slower progress than for other forms of heart disease such as inherited arrhythmia syndromes and hypertrophic and dilated cardiomyopathy. Genetic studies in CHD were traditionally restricted to multiplex families with strong phenotypic penetrance, which represent the minority of cases[40]. The relatively low familial recurrence risk is not fully understood but may be due, in part, to de novo mutations, incomplete penetrance, and other etiological factors such as environmental influences. Patterns of inheritance may be difficult to sort out in the presence of environmental interactions, age-dependent or incomplete penetrance, and variable expressivity. In addition, genetic analysis based on individual families requires a large number of members or consanguinity[11]. Moreover, mutations may involve non-exonic DNA, such as regulatory regions, the functional validation of which is more difficult and resource consuming. Establishing genotype-phenotype correlations may be further complicated by mutations that are rare and unique to individual families[2]. In fact, most CHD mutations identified to date appear to be private or do not recur. Despite these numerous limitations, genetics has and will hopefully continue to provide insights into the etiology of CHD, embryonic heart development, potential therapeutic targets, risk assessment, and patterns of inheritance.

Future perspective

Objetives of genetic testing for clinical reasons differ from research goals. From a clinical perspective, a genetic test should be directly relevant to a patient by serving the purpose of establishing or confirming a diagnosis, providing prognostic information, informing therapeutic decisions, and/or assisting with family planning. In contrast, genetic testing for research purposes may provide pathophysiological insights into a disease entity and identify potential therapeutic targets, thereby carrying the potential to impact care at a longer-term horizon. Nevertheless, genetic results derived from research studies are generally communicated to the clinical team and may directly contribute to the care of a given family[68]. It is important, therefore, for the clinical team to be well versed in the domain in order to effectively communicate with the patient, explain results, and establish an appropriate surveillance plan. In parallel, genetic testing within clinical laboratories may discover new mutations in known genes and novel implicated genes, particulary when modern technologies that sequence a broad array of genes are applied. In the future, therefore, enhanced partnerships between clinical and research teams could maximize the potential for progress. Resources available for research, including highly qualified personnel, informatics infrastructures, laboratory equipment, novel platforms, and more rapid time to analyses could complement the clinical laboratory setting in enhancing clinical care. Greater integration between clinical and research teams could also contribute to ensuring that discoveries are progressive and clinically meaningful, with direct applications to patient care. Along these lines, the multicenter prospective “CHD GENES” study was initiated in December 2010 to explore relationships between genetic factors, clinical features, and outcomes in patients with CHD[2].

CONCLUSION

Despite major inroads over the past few decades in genetics related to CHD, the majority of patients with CHD are without a genetic diagnosis such that the etiology of their CHD remains incompletely understood. In this article, we discussed the multifaceted implications of genetics in CHD including the potential for personalized care, confidence in the clinical diagnosis, prognostic implications, early identification of non-cardiac organ involvement and associated complications, and tailoring clinical follow-up. Genetic testing could also provide valuable information in predicting recurrence risk, defining the pattern of inheritance, screening family members, and family planning. Various methodologies are available to diagnose chromosomal abnormalities and gene mutations. The challenge lies in first identifying potential genetic etiologies, selecting the appropriate test, and interpreting the test within the context of available knowledge. Collaboration between clinicians and genetics researchers offers the best opportunity for progress in clinical care and innovative breakthroughs[69,70]. Much remains to be discovered in tapping the potential of genetics in CHD.

Footnotes

P- Reviewer: Kettering K, Najafi M, Said SAM, Ueda H S- Editor: Qiu S L- Editor: A E- Editor: Lu YJ

References
1.  Lambrechts D, Devriendt K, Driscoll DA, Goldmuntz E, Gewillig M, Vlietinck R, Collen D, Carmeliet P. Low expression VEGF haplotype increases the risk for tetralogy of Fallot: a family based association study. J Med Genet. 2005;42:519-522.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 50]  [Cited by in F6Publishing: 54]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
2.  Gelb B, Brueckner M, Chung W, Goldmuntz E, Kaltman J, Kaski JP, Kim R, Kline J, Mercer-Rosa L, Porter G. The Congenital Heart Disease Genetic Network Study: rationale, design, and early results. Circ Res. 2013;112:698-706.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 104]  [Cited by in F6Publishing: 115]  [Article Influence: 10.5]  [Reference Citation Analysis (0)]
3.  van der Bom T, Zomer AC, Zwinderman AH, Meijboom FJ, Bouma BJ, Mulder BJ. The changing epidemiology of congenital heart disease. Nat Rev Cardiol. 2011;8:50-60.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 391]  [Cited by in F6Publishing: 462]  [Article Influence: 33.0]  [Reference Citation Analysis (0)]
4.  Nora JJ. Multifactorial inheritance hypothesis for the etiology of congenital heart diseases. The genetic-environmental interaction. Circulation. 1968;38:604-617.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 213]  [Cited by in F6Publishing: 227]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
5.  Czeizel AE. Periconceptional folic acid containing multivitamin supplementation. Eur J Obstet Gynecol Reprod Biol. 1998;78:151-161.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 107]  [Cited by in F6Publishing: 122]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
6.  Botto LD, Mulinare J, Erickson JD. Occurrence of congenital heart defects in relation to maternal mulitivitamin use. Am J Epidemiol. 2000;151:878-884.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 187]  [Cited by in F6Publishing: 197]  [Article Influence: 8.2]  [Reference Citation Analysis (0)]
7.  Scanlon KS, Ferencz C, Loffredo CA, Wilson PD, Correa-Villaseñor A, Khoury MJ, Willett WC. Preconceptional folate intake and malformations of the cardiac outflow tract. Baltimore-Washington Infant Study Group. Epidemiology. 1998;9:95-98.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 96]  [Cited by in F6Publishing: 95]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
8.  Jenkins KJ, Correa A, Feinstein JA, Botto L, Britt AE, Daniels SR, Elixson M, Warnes CA, Webb CL. Noninherited risk factors and congenital cardiovascular defects: current knowledge: a scientific statement from the American Heart Association Council on Cardiovascular Disease in the Young: endorsed by the American Academy of Pediatrics. Circulation. 2007;115:2995-3014.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 538]  [Cited by in F6Publishing: 539]  [Article Influence: 31.7]  [Reference Citation Analysis (0)]
9.  Burn J, Brennan P, Little J, Holloway S, Coffey R, Somerville J, Dennis NR, Allan L, Arnold R, Deanfield JE. Recurrence risks in offspring of adults with major heart defects: results from first cohort of British collaborative study. Lancet. 1998;351:311-316.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 265]  [Cited by in F6Publishing: 226]  [Article Influence: 8.7]  [Reference Citation Analysis (0)]
10.  Øyen N, Poulsen G, Boyd HA, Wohlfahrt J, Jensen PK, Melbye M. Recurrence of congenital heart defects in families. Circulation. 2009;120:295-301.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 244]  [Cited by in F6Publishing: 242]  [Article Influence: 16.1]  [Reference Citation Analysis (0)]
11.  Aburawi EH, Aburawi HE, Bagnall KM, Bhuiyan ZA. Molecular insight into heart development and congenital heart disease: An update review from the Arab countries. Trends Cardiovasc Med. 2015;25:291-301.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 13]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
12.  Grosse SD, Khoury MJ. What is the clinical utility of genetic testing? Genet Med. 2006;8:448-450.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  Ching YH, Ghosh TK, Cross SJ, Packham EA, Honeyman L, Loughna S, Robinson TE, Dearlove AM, Ribas G, Bonser AJ. Mutation in myosin heavy chain 6 causes atrial septal defect. Nat Genet. 2005;37:423-428.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 196]  [Cited by in F6Publishing: 189]  [Article Influence: 9.9]  [Reference Citation Analysis (0)]
14.  Kirk EP, Sunde M, Costa MW, Rankin SA, Wolstein O, Castro ML, Butler TL, Hyun C, Guo G, Otway R. Mutations in cardiac T-box factor gene TBX20 are associated with diverse cardiac pathologies, including defects of septation and valvulogenesis and cardiomyopathy. Am J Hum Genet. 2007;81:280-291.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 257]  [Cited by in F6Publishing: 255]  [Article Influence: 15.0]  [Reference Citation Analysis (0)]
15.  Bruneau BG. The developmental genetics of congenital heart disease. Nature. 2008;451:943-948.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 531]  [Cited by in F6Publishing: 535]  [Article Influence: 33.4]  [Reference Citation Analysis (0)]
16.  Garg V, Muth AN, Ransom JF, Schluterman MK, Barnes R, King IN, Grossfeld PD, Srivastava D. Mutations in NOTCH1 cause aortic valve disease. Nature. 2005;437:270-274.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1040]  [Cited by in F6Publishing: 1001]  [Article Influence: 52.7]  [Reference Citation Analysis (0)]
17.  Pierpont ME, Basson CT, Benson DW, Gelb BD, Giglia TM, Goldmuntz E, McGee G, Sable CA, Srivastava D, Webb CL. Genetic basis for congenital heart defects: current knowledge: a scientific statement from the American Heart Association Congenital Cardiac Defects Committee, Council on Cardiovascular Disease in the Young: endorsed by the American Academy of Pediatrics. Circulation. 2007;115:3015-3038.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 595]  [Cited by in F6Publishing: 554]  [Article Influence: 32.6]  [Reference Citation Analysis (0)]
18.  Hinton RB. The family history: reemergence of an established tool. Crit Care Nurs Clin North Am. 2008;20:149-158, v.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 25]  [Cited by in F6Publishing: 26]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
19.  Baker E, Hinton L, Callen DF, Altree M, Dobbie A, Eyre HJ, Sutherland GR, Thompson E, Thompson P, Woollatt E. Study of 250 children with idiopathic mental retardation reveals nine cryptic and diverse subtelomeric chromosome anomalies. Am J Med Genet. 2002;107:285-293.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 89]  [Cited by in F6Publishing: 97]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
20.  Anderlid BM, Schoumans J, Annerén G, Sahlén S, Kyllerman M, Vujic M, Hagberg B, Blennow E, Nordenskjöld M. Subtelomeric rearrangements detected in patients with idiopathic mental retardation. Am J Med Genet. 2002;107:275-284.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 94]  [Cited by in F6Publishing: 104]  [Article Influence: 4.7]  [Reference Citation Analysis (0)]
21.  van Trier DC, Feenstra I, Bot P, de Leeuw N, Draaisma JM. Cardiac anomalies in individuals with the 18q deletion syndrome; report of a child with Ebstein anomaly and review of the literature. Eur J Med Genet. 2013;56:426-431.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 24]  [Cited by in F6Publishing: 25]  [Article Influence: 2.3]  [Reference Citation Analysis (0)]
22.  Rehm HL. Disease-targeted sequencing: a cornerstone in the clinic. Nat Rev Genet. 2013;14:295-300.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 286]  [Cited by in F6Publishing: 278]  [Article Influence: 25.3]  [Reference Citation Analysis (0)]
23.  Teekakirikul P, Kelly MA, Rehm HL, Lakdawala NK, Funke BH. Inherited cardiomyopathies: molecular genetics and clinical genetic testing in the postgenomic era. J Mol Diagn. 2013;15:158-170.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 132]  [Cited by in F6Publishing: 145]  [Article Influence: 12.1]  [Reference Citation Analysis (0)]
24.  Charron P, Arad M, Arbustini E, Basso C, Bilinska Z, Elliott P, Helio T, Keren A, McKenna WJ, Monserrat L. Genetic counselling and testing in cardiomyopathies: a position statement of the European Society of Cardiology Working Group on Myocardial and Pericardial Diseases. Eur Heart J. 2010;31:2715-2726.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 310]  [Cited by in F6Publishing: 339]  [Article Influence: 24.2]  [Reference Citation Analysis (0)]
25.  Resta R, Biesecker BB, Bennett RL, Blum S, Hahn SE, Strecker MN, Williams JL. A new definition of Genetic Counseling: National Society of Genetic Counselors’ Task Force report. J Genet Couns. 2006;15:77-83.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 520]  [Cited by in F6Publishing: 484]  [Article Influence: 26.9]  [Reference Citation Analysis (0)]
26.  Bates BR, Templeton A, Achter PJ, Harris TM, Condit CM. What does “a gene for heart disease” mean? A focus group study of public understandings of genetic risk factors. Am J Med Genet A. 2003;119A:156-161.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 55]  [Cited by in F6Publishing: 69]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
27.  Egbe A, Lee S, Ho D, Uppu S, Srivastava S. Prevalence of congenital anomalies in newborns with congenital heart disease diagnosis. Ann Pediatr Cardiol. 2014;7:86-91.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 36]  [Cited by in F6Publishing: 41]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
28.  Pashmforoush M, Lu JT, Chen H, Amand TS, Kondo R, Pradervand S, Evans SM, Clark B, Feramisco JR, Giles W. Nkx2-5 pathways and congenital heart disease; loss of ventricular myocyte lineage specification leads to progressive cardiomyopathy and complete heart block. Cell. 2004;117:373-386.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 329]  [Cited by in F6Publishing: 308]  [Article Influence: 15.4]  [Reference Citation Analysis (0)]
29.  Shaddy RE, Webb G. Applying heart failure guidelines to adult congenital heart disease patients. Expert Rev Cardiovasc Ther. 2008;6:165-174.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 41]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
30.  Warnes CA. Adult congenital heart disease importance of the right ventricle. J Am Coll Cardiol. 2009;54:1903-1910.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 102]  [Cited by in F6Publishing: 99]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
31.  Costa MW, Guo G, Wolstein O, Vale M, Castro ML, Wang L, Otway R, Riek P, Cochrane N, Furtado M. Functional characterization of a novel mutation in NKX2-5 associated with congenital heart disease and adult-onset cardiomyopathy. Circ Cardiovasc Genet. 2013;6:238-247.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 63]  [Cited by in F6Publishing: 71]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
32.  Zhu Y, Gramolini AO, Walsh MA, Zhou YQ, Slorach C, Friedberg MK, Takeuchi JK, Sun H, Henkelman RM, Backx PH. Tbx5-dependent pathway regulating diastolic function in congenital heart disease. Proc Natl Acad Sci USA. 2008;105:5519-5524.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 54]  [Cited by in F6Publishing: 55]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
33.  Hickey EJ, Mehta R, Elmi M, Asoh K, McCrindle BW, Williams WG, Manlhiot C, Benson L. Survival implications: hypertrophic cardiomyopathy in Noonan syndrome. Congenit Heart Dis. 2011;6:41-47.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 47]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
34.  Granados-Riveron JT, Ghosh TK, Pope M, Bu’Lock F, Thornborough C, Eason J, Kirk EP, Fatkin D, Feneley MP, Harvey RP. Alpha-cardiac myosin heavy chain (MYH6) mutations affecting myofibril formation are associated with congenital heart defects. Hum Mol Genet. 2010;19:4007-4016.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 102]  [Cited by in F6Publishing: 110]  [Article Influence: 7.9]  [Reference Citation Analysis (0)]
35.  Postma AV, van Engelen K, van de Meerakker J, Rahman T, Probst S, Baars MJ, Bauer U, Pickardt T, Sperling SR, Berger F. Mutations in the sarcomere gene MYH7 in Ebstein anomaly. Circ Cardiovasc Genet. 2011;4:43-50.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 120]  [Cited by in F6Publishing: 112]  [Article Influence: 8.0]  [Reference Citation Analysis (0)]
36.  Vermeer AM, van Engelen K, Postma AV, Baars MJ, Christiaans I, De Haij S, Klaassen S, Mulder BJ, Keavney B. Ebstein anomaly associated with left ventricular noncompaction: an autosomal dominant condition that can be caused by mutations in MYH7. Am J Med Genet C Semin Med Genet. 2013;163C:178-184.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 35]  [Article Influence: 3.2]  [Reference Citation Analysis (0)]
37.  Kirshbom PM, Mahle WT, Joyner RW, Leong T, Wilson M, Kogon BE, Kanter KR, Bouzyk MM. The endothelin-1 G5665T polymorphism impacts transplant-free survival for single ventricle patients. J Thorac Cardiovasc Surg. 2008;136:117-122.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 13]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
38.  Kim DS, Kim JH, Burt AA, Crosslin DR, Burnham N, McDonald-McGinn DM, Zackai EH, Nicolson SC, Spray TL, Stanaway IB. Patient genotypes impact survival after surgery for isolated congenital heart disease. Ann Thorac Surg. 2014;98:104-110; discussion 110-111.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 21]  [Cited by in F6Publishing: 19]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
39.  Mercer-Rosa L, Pinto N, Yang W, Tanel R, Goldmuntz E. 22q11.2 Deletion syndrome is associated with perioperative outcome in tetralogy of Fallot. J Thorac Cardiovasc Surg. 2013;146:868-873.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 60]  [Cited by in F6Publishing: 65]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
40.  Barnett P, Postma AV. Genetics of congenital heart disease: Beyond half-measures. Trends Cardiovasc Med. 2015;25:302-304.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 4]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
41.  Tian Y, Yuan L, Goss AM, Wang T, Yang J, Lepore JJ, Zhou D, Schwartz RJ, Patel V, Cohen ED. Characterization and in vivo pharmacological rescue of a Wnt2-Gata6 pathway required for cardiac inflow tract development. Dev Cell. 2010;18:275-287.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 100]  [Cited by in F6Publishing: 103]  [Article Influence: 7.4]  [Reference Citation Analysis (0)]
42.  Whittemore R, Hobbins JC, Engle MA. Pregnancy and its outcome in women with and without surgical treatment of congenital heart disease. Am J Cardiol. 1982;50:641-651.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 248]  [Cited by in F6Publishing: 263]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
43.  Gill HK, Splitt M, Sharland GK, Simpson JM. Patterns of recurrence of congenital heart disease: an analysis of 6,640 consecutive pregnancies evaluated by detailed fetal echocardiography. J Am Coll Cardiol. 2003;42:923-929.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 139]  [Cited by in F6Publishing: 108]  [Article Influence: 5.1]  [Reference Citation Analysis (0)]
44.  Calcagni G, Digilio MC, Sarkozy A, Dallapiccola B, Marino B. Familial recurrence of congenital heart disease: an overview and review of the literature. Eur J Pediatr. 2007;166:111-116.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 63]  [Cited by in F6Publishing: 64]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
45.  Digilio MC, Marino B, Cicini MP, Giannotti A, Formigari R, Dallapiccola B. Risk of congenital heart defects in relatives of patients with atrioventricular canal. Am J Dis Child. 1993;147:1295-1297.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5]  [Cited by in F6Publishing: 13]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
46.  Emanuel R, Nichols J, Anders JM, Moores EC, Somerville J. Atrioventricular defects--a study of 92 families. Br Heart J. 1968;30:645-653.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 26]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
47.  Digilio MC, Marino B, Giannotti A, Toscano A, Dallapiccola B. Recurrence risk figures for isolated tetralogy of Fallot after screening for 22q11 microdeletion. J Med Genet. 1997;34:188-190.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 38]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
48.  Miller ME, Smith DW. Conotruncal malformation complex: examples of possible monogenic inheritance. Pediatrics. 1979;63:890-893.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 35]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
49.  Digilio MC, Casey B, Toscano A, Calabrò R, Pacileo G, Marasini M, Banaudi E, Giannotti A, Dallapiccola B, Marino B. Complete transposition of the great arteries: patterns of congenital heart disease in familial precurrence. Circulation. 2001;104:2809-2814.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 82]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
50.  Loffredo CA, Chokkalingam A, Sill AM, Boughman JA, Clark EB, Scheel J, Brenner JI. Prevalence of congenital cardiovascular malformations among relatives of infants with hypoplastic left heart, coarctation of the aorta, and d-transposition of the great arteries. Am J Med Genet A. 2004;124A:225-230.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 86]  [Cited by in F6Publishing: 92]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
51.  Piacentini G, Digilio MC, Capolino R, Zorzi AD, Toscano A, Sarkozy A, D’Agostino R, Marasini M, Russo MG, Dallapiccola B. Familial recurrence of heart defects in subjects with congenitally corrected transposition of the great arteries. Am J Med Genet A. 2005;137:176-180.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 32]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
52.  McBride KL, Pignatelli R, Lewin M, Ho T, Fernbach S, Menesses A, Lam W, Leal SM, Kaplan N, Schliekelman P. Inheritance analysis of congenital left ventricular outflow tract obstruction malformations: Segregation, multiplex relative risk, and heritability. Am J Med Genet A. 2005;134A:180-186.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 163]  [Cited by in F6Publishing: 146]  [Article Influence: 7.7]  [Reference Citation Analysis (0)]
53.  Wessels MW, Berger RM, Frohn-Mulder IM, Roos-Hesselink JW, Hoogeboom JJ, Mancini GS, Bartelings MM, Krijger Rd, Wladimiroff JW, Niermeijer MF. Autosomal dominant inheritance of left ventricular outflow tract obstruction. Am J Med Genet A. 2005;134A:171-179.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 58]  [Cited by in F6Publishing: 62]  [Article Influence: 3.3]  [Reference Citation Analysis (0)]
54.  Hales AR, Mahle WT. Echocardiography screening of siblings of children with bicuspid aortic valve. Pediatrics. 2014;133:e1212-e1217.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 27]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
55.  Cripe L, Andelfinger G, Martin LJ, Shooner K, Benson DW. Bicuspid aortic valve is heritable. J Am Coll Cardiol. 2004;44:138-143.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 435]  [Cited by in F6Publishing: 401]  [Article Influence: 20.1]  [Reference Citation Analysis (0)]
56.  Oyen N, Poulsen G, Wohlfahrt J, Boyd HA, Jensen PK, Melbye M. Recurrence of discordant congenital heart defects in families. Circ Cardiovasc Genet. 2010;3:122-128.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 51]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
57.  Nistri S, Basso C, Marzari C, Mormino P, Thiene G. Frequency of bicuspid aortic valve in young male conscripts by echocardiogram. Am J Cardiol. 2005;96:718-721.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 118]  [Cited by in F6Publishing: 116]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
58.  van Velzen C, Clur S, Rijlaarsdam M, Bax C, Pajkrt E, Heymans M, Bekker M, Hruda J, de Groot C, Blom N. Prenatal detection of congenital heart disease-results of a national screening programme. BJOG. 2015;123: 400-407.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 118]  [Cited by in F6Publishing: 127]  [Article Influence: 14.1]  [Reference Citation Analysis (0)]
59.  Bonnet D, Coltri A, Butera G, Fermont L, Le Bidois J, Kachaner J, Sidi D. Detection of transposition of the great arteries in fetuses reduces neonatal morbidity and mortality. Circulation. 1999;99:916-918.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 519]  [Cited by in F6Publishing: 464]  [Article Influence: 18.6]  [Reference Citation Analysis (0)]
60.  Mahle WT, Clancy RR, McGaurn SP, Goin JE, Clark BJ. Impact of prenatal diagnosis on survival and early neurologic morbidity in neonates with the hypoplastic left heart syndrome. Pediatrics. 2001;107:1277-1282.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 257]  [Cited by in F6Publishing: 249]  [Article Influence: 10.8]  [Reference Citation Analysis (0)]
61.  Beroukhim RS, Gauvreau K, Benavidez OJ, Baird CW, LaFranchi T, Tworetzky W. Perinatal outcome after prenatal diagnosis of single-ventricle cardiac defects. Ultrasound Obstet Gynecol. 2015;45:657-663.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 24]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
62.  Rossier MC, Mivelaz Y, Addor MC, Sekarski N, Meijboom EJ, Vial Y. Evaluation of prenatal diagnosis of congenital heart disease in a regional controlled case study. Swiss Med Wkly. 2014;144:w14068.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 6]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
63.  Chenni N, Lacroze V, Pouet C, Fraisse A, Kreitmann B, Gamerre M, Boubli L, D’Ercole C. Fetal heart disease and interruption of pregnancy: factors influencing the parental decision-making process. Prenat Diagn. 2012;32:168-172.  [PubMed]  [DOI]  [Cited in This Article: ]
64.  Liao C, Li R, Fu F, Xie G, Zhang Y, Pan M, Li J, Li D. Prenatal diagnosis of congenital heart defect by genome-wide high-resolution SNP array. Prenat Diagn. 2014;34:858-863.  [PubMed]  [DOI]  [Cited in This Article: ]
65.  Yan Y, Wu Q, Zhang L, Wang X, Dan S, Deng D, Sun L, Yao L, Ma Y, Wang L. Detection of submicroscopic chromosomal aberrations by array-based comparative genomic hybridization in fetuses with congenital heart disease. Ultrasound Obstet Gynecol. 2014;43:404-412.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 42]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
66.  McDermott DA, Basson CT, Hatcher CJ. Genetics of cardiac septation defects and their pre-implantation diagnosis. Methods Mol Med. 2006;126:19-42.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
67.  He J, McDermott DA, Song Y, Gilbert F, Kligman I, Basson CT. Preimplantation genetic diagnosis of human congenital heart malformation and Holt-Oram syndrome. Am J Med Genet A. 2004;126A:93-98.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 36]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
68.  Shalowitz DI, Miller FG. Communicating the results of clinical research to participants: attitudes, practices, and future directions. PLoS Med. 2008;5:e91.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 152]  [Cited by in F6Publishing: 165]  [Article Influence: 10.3]  [Reference Citation Analysis (0)]
69.  McInerney JD. Genetics education for health professionals: a context. J Genet Couns. 2008;17:145-151.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 24]  [Cited by in F6Publishing: 25]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
70.  Guttmacher AE, Porteous ME, McInerney JD. Educating health-care professionals about genetics and genomics. Nat Rev Genet. 2007;8:151-157.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 201]  [Cited by in F6Publishing: 229]  [Article Influence: 13.5]  [Reference Citation Analysis (0)]
71.  Down JL. Observations on an ethnic classification of idiots. 1866. Ment Retard. 1995;33:54-56.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 24]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
72.  Marino B, Vairo U, Corno A, Nava S, Guccione P, Calabrò R, Marcelletti C. Atrioventricular canal in Down syndrome. Prevalence of associated cardiac malformations compared with patients without Down syndrome. Am J Dis Child. 1990;144:1120-1122.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 95]  [Cited by in F6Publishing: 102]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
73.  Park SC, Mathews RA, Zuberbuhler JR, Rowe RD, Neches WH, Lenox CC. Down syndrome with congenital heart malformation. Am J Dis Child. 1977;131:29-33.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 17]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
74.  Cesko I, Hajdú J, Marton T, Tóth-Pál E, Papp C, Papp Z. [Fetal atrioventricular septal defect associated with Patau and Edwards syndromes, as well as trisomy 22]. Orv Hetil. 1998;139:1087-1089.  [PubMed]  [DOI]  [Cited in This Article: ]
75.  Kumar A, Van Mierop LH, Epstein ML. Pathogenetic implications of muscular ventricular septal defect in Holt-Oram syndrome. Am J Cardiol. 1994;73:993-995.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 5]  [Cited by in F6Publishing: 7]  [Article Influence: 0.2]  [Reference Citation Analysis (0)]
76.  Marino B, Digilio MC, Toscano A, Giannotti A, Dallapiccola B. Congenital heart diseases in children with Noonan syndrome: An expanded cardiac spectrum with high prevalence of atrioventricular canal. J Pediatr. 1999;135:703-706.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 164]  [Cited by in F6Publishing: 175]  [Article Influence: 7.0]  [Reference Citation Analysis (0)]
77.  Digilio MC, Marino B, Ammirati A, Borzaga U, Giannotti A, Dallapiccola B. Cardiac malformations in patients with oral-facial-skeletal syndromes: clinical similarities with heterotaxia. Am J Med Genet. 1999;84:350-356.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 1]  [Reference Citation Analysis (0)]
78.  Sheffield VC, Pierpont ME, Nishimura D, Beck JS, Burns TL, Berg MA, Stone EM, Patil SR, Lauer RM. Identification of a complex congenital heart defect susceptibility locus by using DNA pooling and shared segment analysis. Hum Mol Genet. 1997;6:117-121.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 69]  [Cited by in F6Publishing: 72]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
79.  Robinson SW, Morris CD, Goldmuntz E, Reller MD, Jones MA, Steiner RD, Maslen CL. Missense mutations in CRELD1 are associated with cardiac atrioventricular septal defects. Am J Hum Genet. 2003;72:1047-1052.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 143]  [Cited by in F6Publishing: 125]  [Article Influence: 6.0]  [Reference Citation Analysis (0)]
80.  Guo Y, Shen J, Yuan L, Li F, Wang J, Sun K. Novel CRELD1 gene mutations in patients with atrioventricular septal defect. World J Pediatr. 2010;6:348-352.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 20]  [Cited by in F6Publishing: 23]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
81.  Maslen CL, Babcock D, Robinson SW, Bean LJ, Dooley KJ, Willour VL, Sherman SL. CRELD1 mutations contribute to the occurrence of cardiac atrioventricular septal defects in Down syndrome. Am J Med Genet A. 2006;140:2501-2505.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 61]  [Cited by in F6Publishing: 62]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
82.  Garg V, Kathiriya IS, Barnes R, Schluterman MK, King IN, Butler CA, Rothrock CR, Eapen RS, Hirayama-Yamada K, Joo K. GATA4 mutations cause human congenital heart defects and reveal an interaction with TBX5. Nature. 2003;424:443-447.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 908]  [Cited by in F6Publishing: 858]  [Article Influence: 40.9]  [Reference Citation Analysis (0)]
83.  Starke H, Schimke RN, Dunn M. Upper-limb cardiovascular syndrome. A family study. Am J Cardiol. 1967;19:588-592.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 15]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
84.  Ito M, Misawa T, Fujino M, Ito S, Fukumoto T. A family of Holt-Oram syndrome. Jpn Heart J. 1975;16:480-487.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1]  [Cited by in F6Publishing: 2]  [Article Influence: 0.0]  [Reference Citation Analysis (0)]
85.  Allanson JE. Noonan syndrome. J Med Genet. 1987;24:9-13.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 275]  [Cited by in F6Publishing: 291]  [Article Influence: 7.9]  [Reference Citation Analysis (0)]
86.  Sharland M, Burch M, McKenna WM, Paton MA. A clinical study of Noonan syndrome. Arch Dis Child. 1992;67:178-183.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 312]  [Cited by in F6Publishing: 268]  [Article Influence: 8.4]  [Reference Citation Analysis (0)]
87.  Lynch JI, Perry LW, Takakuwa T, Scott LP. Congenital heart disease and chondroectodermal dysplasia. Report of two cases, one in a Negro. Am J Dis Child. 1968;115:80-87.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 24]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
88.  Rodríguez-Caballero A, Torres-Lagares D, Rodríguez-Pérez A, Serrera-Figallo MA, Hernández-Guisado JM, Machuca-Portillo G. Cri du chat syndrome: a critical review. Med Oral Patol Oral Cir Bucal. 2010;15:e473-e478.  [PubMed]  [DOI]  [Cited in This Article: ]
89.  Schott JJ, Benson DW, Basson CT, Pease W, Silberbach GM, Moak JP, Maron BJ, Seidman CE, Seidman JG. Congenital heart disease caused by mutations in the transcription factor NKX2-5. Science. 1998;281:108-111.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 996]  [Cited by in F6Publishing: 872]  [Article Influence: 33.5]  [Reference Citation Analysis (0)]
90.  Chen Y, Mao J, Sun Y, Zhang Q, Cheng HB, Yan WH, Choy KW, Li H. A novel mutation of GATA4 in a familial atrial septal defect. Clin Chim Acta. 2010;411:1741-1745.  [PubMed]  [DOI]  [Cited in This Article: ]
91.  Posch MG, Gramlich M, Sunde M, Schmitt KR, Lee SH, Richter S, Kersten A, Perrot A, Panek AN, Al Khatib IH. A gain-of-function TBX20 mutation causes congenital atrial septal defects, patent foramen ovale and cardiac valve defects. J Med Genet. 2010;47:230-235.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 88]  [Cited by in F6Publishing: 92]  [Article Influence: 6.1]  [Reference Citation Analysis (0)]
92.  Bennhagen RG, Menahem S. Holt-Oram syndrome and multiple ventricular septal defects: an association suggesting a possible genetic marker? Cardiol Young. 1998;8:128-130.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 7]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
93.  Marino B, Papa M, Guccione P, Corno A, Marasini M, Calabrò R. Ventricular septal defect in Down syndrome. Anatomic types and associated malformations. Am J Dis Child. 1990;144:544-545.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 24]  [Cited by in F6Publishing: 27]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
94.  Marino B, Corno A, Guccione P, Marcelletti C. Ventricular septal defect and Down’s syndrome. Lancet. 1991;337:245-246.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 12]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
95.  Van Praagh S, Truman T, Firpo A, Bano-Rodrigo A, Fried R, McManus B, Engle MA, Van Praagh R. Cardiac malformations in trisomy-18: a study of 41 postmortem cases. J Am Coll Cardiol. 1989;13:1586-1597.  [PubMed]  [DOI]  [Cited in This Article: ]
96.  Lauritsen JG. Aetiology of spontaneous abortion. A cytogenetic and epidemiological study of 288 abortuses and their parents. Acta Obstet Gynecol Scand Suppl. 1976;52:1-29.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 124]  [Cited by in F6Publishing: 125]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
97.  McDonald-McGinn DM, Tonnesen MK, Laufer-Cahana A, Finucane B, Driscoll DA, Emanuel BS, Zackai EH. Phenotype of the 22q11.2 deletion in individuals identified through an affected relative: cast a wide FISHing net! Genet Med. 2001;3:23-29.  [PubMed]  [DOI]  [Cited in This Article: ]
98.  Wang J, Fang M, Liu XY, Xin YF, Liu ZM, Chen XZ, Wang XZ, Fang WY, Liu X, Yang YQ. A novel GATA4 mutation responsible for congenital ventricular septal defects. Int J Mol Med. 2011;28:557-564.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2]  [Cited by in F6Publishing: 16]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
99.  Pepeta L, Clur SA. Ebstein’s anomaly and Down’s syndrome. Cardiovasc J Afr. 2013;24:382-384.  [PubMed]  [DOI]  [Cited in This Article: ]
100.  Benson DW, Silberbach GM, Kavanaugh-McHugh A, Cottrill C, Zhang Y, Riggs S, Smalls O, Johnson MC, Watson MS, Seidman JG. Mutations in the cardiac transcription factor NKX2.5 affect diverse cardiac developmental pathways. J Clin Invest. 1999;104:1567-1573.  [PubMed]  [DOI]  [Cited in This Article: ]
101.  Pernot C, Marçon F, Worms AM, Cloez JL, Gilgenkrantz S, Marois L. [Cardiovascular dysplasia in Noonan’s syndrome. Apropos of 64 cases]. Arch Mal Coeur Vaiss. 1987;80:434-443.  [PubMed]  [DOI]  [Cited in This Article: ]
102.  Lin AE. Noonan syndrome. J Med Genet. 1988;25:64-65.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 9]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
103.  Lin AE, Alexander ME, Colan SD, Kerr B, Rauen KA, Noonan J, Baffa J, Hopkins E, Sol-Church K, Limongelli G. Clinical, pathological, and molecular analyses of cardiovascular abnormalities in Costello syndrome: a Ras/MAPK pathway syndrome. Am J Med Genet A. 2011;155A:486-507.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 85]  [Cited by in F6Publishing: 90]  [Article Influence: 6.9]  [Reference Citation Analysis (0)]
104.  Lin AE, Grossfeld PD, Hamilton RM, Smoot L, Gripp KW, Proud V, Weksberg R, Wheeler P, Picker J, Irons M. Further delineation of cardiac abnormalities in Costello syndrome. Am J Med Genet. 2002;111:115-129.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 85]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
105.  Emerick KM, Rand EB, Goldmuntz E, Krantz ID, Spinner NB, Piccoli DA. Features of Alagille syndrome in 92 patients: frequency and relation to prognosis. Hepatology. 1999;29:822-829.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 475]  [Cited by in F6Publishing: 373]  [Article Influence: 14.9]  [Reference Citation Analysis (0)]
106.  Harrison LW, Auld DS, Vallee BL. Intramolecular arsanilazotyrosine-248-Zn complex of carboxypeptidase A: a monitor of catalytic events. Proc Natl Acad Sci USA. 1975;72:3930-3933.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 210]  [Cited by in F6Publishing: 166]  [Article Influence: 7.5]  [Reference Citation Analysis (0)]
107.  Niihori T, Aoki Y, Narumi Y, Neri G, Cavé H, Verloes A, Okamoto N, Hennekam RC, Gillessen-Kaesbach G, Wieczorek D. Germline KRAS and BRAF mutations in cardio-facio-cutaneous syndrome. Nat Genet. 2006;38:294-296.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 411]  [Cited by in F6Publishing: 407]  [Article Influence: 22.6]  [Reference Citation Analysis (0)]
108.  Rajagopal SK, Ma Q, Obler D, Shen J, Manichaikul A, Tomita-Mitchell A, Boardman K, Briggs C, Garg V, Srivastava D. Spectrum of heart disease associated with murine and human GATA4 mutation. J Mol Cell Cardiol. 2007;43:677-685.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 180]  [Cited by in F6Publishing: 174]  [Article Influence: 10.2]  [Reference Citation Analysis (0)]
109.  Elsheikh M, Dunger DB, Conway GS, Wass JA. Turner’s syndrome in adulthood. Endocr Rev. 2002;23:120-140.  [PubMed]  [DOI]  [Cited in This Article: ]
110.  McBride KL, Riley MF, Zender GA, Fitzgerald-Butt SM, Towbin JA, Belmont JW, Cole SE. NOTCH1 mutations in individuals with left ventricular outflow tract malformations reduce ligand-induced signaling. Hum Mol Genet. 2008;17:2886-2893.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 146]  [Cited by in F6Publishing: 153]  [Article Influence: 9.6]  [Reference Citation Analysis (0)]
111.  Tan HL, Glen E, Töpf A, Hall D, O’Sullivan JJ, Sneddon L, Wren C, Avery P, Lewis RJ, ten Dijke P. Nonsynonymous variants in the SMAD6 gene predispose to congenital cardiovascular malformation. Hum Mutat. 2012;33:720-727.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 83]  [Article Influence: 6.9]  [Reference Citation Analysis (0)]
112.  Collins RT, Kaplan P, Somes GW, Rome JJ. Long-term outcomes of patients with cardiovascular abnormalities and williams syndrome. Am J Cardiol. 2010;105:874-878.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 88]  [Cited by in F6Publishing: 97]  [Article Influence: 6.9]  [Reference Citation Analysis (0)]
113.  Balderston SM, Shaffer EM, Washington RL, Sondheimer HM. Congenital polyvalvular disease in trisomy 18: echocardiographic diagnosis. Pediatr Cardiol. 1990;11:138-142.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 36]  [Article Influence: 1.1]  [Reference Citation Analysis (0)]
114.  Andelfinger G, Tapper AR, Welch RC, Vanoye CG, George AL, Benson DW. KCNJ2 mutation results in Andersen syndrome with sex-specific cardiac and skeletal muscle phenotypes. Am J Hum Genet. 2002;71:663-668.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 189]  [Cited by in F6Publishing: 157]  [Article Influence: 7.1]  [Reference Citation Analysis (0)]
115.  Cory CC, Jamison DL. The cat eye syndrome. Arch Ophthalmol. 1974;92:259-262.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 9]  [Article Influence: 0.2]  [Reference Citation Analysis (0)]
116.  Zhang W, Li X, Shen A, Jiao W, Guan X, Li Z. GATA4 mutations in 486 Chinese patients with congenital heart disease. Eur J Med Genet. 2002;51:527-535.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 71]  [Cited by in F6Publishing: 63]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
117.  Pizzuti A, Sarkozy A, Newton AL, Conti E, Flex E, Digilio MC, Amati F, Gianni D, Tandoi C, Marino B. Mutations of ZFPM2/FOG2 gene in sporadic cases of tetralogy of Fallot. Hum Mutat. 2003;22:372-377.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 100]  [Cited by in F6Publishing: 103]  [Article Influence: 5.2]  [Reference Citation Analysis (0)]
118.  De Luca A, Sarkozy A, Ferese R, Consoli F, Lepri F, Dentici ML, Vergara P, De Zorzi A, Versacci P, Digilio MC. New mutations in ZFPM2/FOG2 gene in tetralogy of Fallot and double outlet right ventricle. Clin Genet. 2011;80:184-190.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 54]  [Cited by in F6Publishing: 63]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
119.  McElhinney DB, Geiger E, Blinder J, Benson DW, Goldmuntz E. NKX2.5 mutations in patients with congenital heart disease. J Am Coll Cardiol. 2003;42:1650-1655.  [PubMed]  [DOI]  [Cited in This Article: ]