Birth defects contribute to ∼0.3% of global infant mortality in the first month of life, and congenital heart disease (CHD) is the most common birth defect among newborns worldwide. Despite the significant impact on human health, most treatments available for this heterogenous group of disorders are palliative at best. For this reason, the complex process of cardiogenesis, governed by multiple interlinked and dose-dependent pathways, is well investigated. Tissue, animal and, more recently, computerized models of the developing heart have facilitated important discoveries that are helping us to understand the genetic, epigenetic and mechanobiological contributors to CHD aetiology. In this Review, we discuss the strengths and limitations of different models of normal and abnormal cardiogenesis, ranging from single-cell systems and 3D cardiac organoids, to small and large animals and organ-level computational models. These investigative tools have revealed a diversity of pathogenic mechanisms that contribute to CHD, including genetic pathways, epigenetic regulators and shear wall stresses, paving the way for new strategies for screening and non-surgical treatment of CHD. As we discuss in this Review, one of the most-valuable advances in recent years has been the creation of highly personalized platforms with which to study individual diseases in clinically relevant settings.

Congenital heart disease (CHD) refers to a heterogeneous collection of structural abnormalities of the heart or the great vessels. These are the most-common birth defect in newborns, affecting >1 million live births per annum globally and causing 10% of stillbirths, where moderate and severe forms affect 6–20 per thousand live-births (Hoffman and Kaplan, 2002). CHDs are a major cause of infant mortality and morbidity in the developed world (Fahed et al., 2013; Jin et al., 2017; Yang et al., 2006; Zaidi and Brueckner, 2017), accounting for ∼40% of infant deaths in North America (Agopian et al., 2017). Affected newborns, especially those with cyanotic CHD (Box 1, Glossary), may require palliative surgery during infancy, and affected children sometimes require subsequent invasive procedures and eventual heart transplants owing to cardiac failure caused by progressive ventricular dysfunction (Bacha et al., 2008; Yagi et al., 2018). Nevertheless, most CHD patients now are expected to survive, including children with complex cardiac defects (GB_2017_Congenital_Heart_Disease_Collaborators, 2020). This increased survival is attributable to innovative advances in imaging techniques, minimally invasive surgeries, improved clinical surveillance and translational research, all of which have dramatically improved the clinical management of CHD.

Box 1. Glossary

Anisotropic hyperelastic constitutive law: a constitutive law is a mathematical description of the mechanical properties of matter. The hyperelastic model is a form of constitutive law that captures elastic properties well. Anisotropic refers to a material that has different mechanical properties (such as stiffness) in different directions, and is opposed to isotropic, which refers to material that has the same properties in any direction.

Coarctation of the aorta(CoA): narrowing of the aorta between upper and lower body branches, typically just after the arch of the aorta.

Aortic isthmus: the section of the aorta between its connection to the left subclavian artery and the ductus arteriosus.

Aortic valvuloplasty: a surgical procedure to expand the stenosis at the aortic valve by using a percutaneous balloon catheter.

Cardia bifida: an embryonic malformation incompatible with life, denoting a ‘double’ or ‘split’ heart in which heart tube formation produces two heart cylinders instead of one, arresting normal development and circulation. It arises from improper cell migration or differentiation and is often genetically driven.

Cardiac neural crest: multipotent embryonic cells that undergo epithelial-mesenchymal transition and migrate to the heart via pharyngeal arches 3, 4 and 6. The cardiac neural crest complex aids cardiac outflow tract and aortic arch modeling.

Chromatin quantitative trait loci (chQTL): DNA loci that encode a quantitative phenotypic trait of an organism, which is identified in the genome by the association of certain molecular signatures (e.g. single-nucleotide polymorphisms) with the observed physical characteristic(s). These loci underlie characteristics that can be described as continuous, e.g. a severity spectrum of disease phenotype, which may be polygenic and under environmental influence, rather than being a discrete (e.g. present/absent) trait.

Conotruncal defects: this group of CHD affects the cardiac outflow tracts (pulmonary artery and aorta) and is frequently associated with the 22q11.2 deletion (DiGeorge syndrome). Examples include pulmonary atresia, double-outlet right ventricle, double-outlet left ventricle, transposition of the great arteries and tetralogy of Fallot.

Critical regulators of organ development: proteins that fulfil fundamental roles in cardiac and other organ development, an example of which are the NKX-homeodomain factors (see below).

Cyanotic congenital heart disease: a group of CHD that cause chronically low blood oxygenation, leading to a blue (cyanosed) appearance of the lips, fingernails and skin.

Double-outlet left ventricle (DOLV)/double-outlet right ventricle (DORV): rare cardiac malformation in which both great arteries originate from the morphological left/right ventricle.

Doppler optical coherence tomography: non-invasive imaging technique that uses backscattered light and Doppler waveforms to achieve high-resolution tomographic images of biological tissues. It is particularly effective for in vivo blood flow imaging.

Ductus venosus (DV): a vein that shunts a portion of umbilical vein blood flow directly to the inferior vena cava in a fetus.

E- and A-waves: two peaks in the cardiac ventricular inflow velocity waveforms during diastole, when the heart muscle relaxes. E-wave corresponds to first wave of inflow caused by the passive relaxation of the ventricle, whereas A-wave corresponds to the second wave of inflow caused by active contraction of the atrium.

Fetoscopy: surgical procedure that uses a small-calibre laparoscope to obtain minimally invasive access to the fetus, amniotic cavity, umbilical cord and the fetal side of the placenta.

First and second heart fields: the heart fields arise from the embryonic mesoderm; the primary (first) heart field forms the initial heart tube, left ventricle and left and right atria, whereas the second heart field contributes to the right ventricle, both atria and the outflow tracts.

Fontan surgery: surgical repair of single-ventricle congenital heart malformations that leads to systemic flow of venous blood to the lungs without passing through a ventricle.

Heterotaxy syndrome: a series of congenital defects caused by disruption of left-right laterality, i.e. the programmed rotation of organs so that they reach their final location within the body cavity, resulting in disordered arrangement of thoracic and abdominal viscera in relation to the body axis.

His–Purkinje system: fibers located in the atrial and ventricular myocardium that form the electrical conduction pathway of the heart. The bundle of His transmits signals received from the sinoatrial and atrioventricular nodes to the Purkinje fibers. The latter split in the interventricular septum to supply both ventricles with electrical signals that initiate contractions.

Mitral and tricuspid valves: atrioventricular valves control the direction of blood flow from the atrium into the ventricle. The mitral valve is in the left heart and the tricuspid valve in the right heart.

Navier–Stokes equations: a set of equations that describe the physics of fluid flow, relating fluid flow patterns to fluid forces, derived from the conservation law principles of mass, momentum and energy.

Nkx5-2: homeodomain transcription factors, which exhibit a characteristic folded structure that allows them to bind target DNA to regulate gene expression. These transcription factors are important for cell differentiation during the early embryonic period and, thus, essential in cardiogenesis. Mutations cause disorders of normal heart development.

Proepicardium: transient extracardiac tissue that appears in the embryonic period and aids cardiogenesis in vertebrates. It arises from the lateral plate mesoderm. Mesothelial cells cross from the proepicardium to attach to the myocardium, eventually forming the epicardium.

Pulmonary atresia: a condition in which the pulmonary valve develops incorrectly, prohibiting blood flow between the right ventricle into the lungs.

Sarcomere: a unit of striated muscle, whose long fibrous protein filaments slide past each other during muscle contraction or relaxation, composed of actin (thin filament, bound to the Z lines that define the sarcomere unit) and myosin (thick filament, bound to adenosine triphosphate providing energy for action).

Short-hairpin RNAs (shRNAs): artificial RNA molecules that comprise a tight hairpin turn to silence endogenous messenger RNA translation via the RNA-induced silencing complex that binds to and cleaves mRNA or suppresses its translation. Expression of shRNA in the target cell is accomplished by using viral vectors or plasmids.

Slit-Robo: a cellular signaling complex involved in axon development and branching, as well as neuronal cell migration. The complex comprises the secreted protein Slit and its transmembrane receptor Roundabout (Robo).

Spatio-temporal image correlation (STIC): a mode of ultrasound imaging providing 4D B-mode images (3D over time), enabled by image-correlation processing and a slow motorized out-of-plane sweep of 2D scanning elements during the scan.

Tetralogy of Fallot (TOF): a common cyanotic CHD, characterized by pulmonary stenosis/right ventricular outflow tract obstruction (RVOTO), ventricular septal defect (VSD), overriding aorta and hypertrophy of the right ventricle (RVH).

Transposition of the great arteries: rare heart defect due to a positional swap of the pulmonary artery and the aorta. Transposition of the great arteries: discordant ventriculo-arterial connection – the right ventricle is connected to the aorta (instead of pulmonary artery), and left ventricle to pulmonary artery (instead of aorta).

CoA, coarctation of the aorta; ECM, extracellular matrix; EMT, epithelial to mesenchymal transition; HLHS, hypoplastic left heart syndrome; LPA, left pulmonary artery; TOF, tetralogy of Fallot.

CHD is caused by abnormal cardiogenesis, a complex developmental process (Fig. 1) governed by multiple interlinked and dose-dependent pathways (Witman et al., 2020). Although 80–85% of CHDs are caused by multifactorial or unidentified causes, large epidemiological and molecular studies have identified monogenic (3–5%) or chromosomal (8–10%) anomalies, copy number variants (3–25%) and environmental causes (2%), such as maternal diabetes, smoking or alcohol use, in ∼20–30% of patients (Cowan and Ware, 2015; Hoffman and Kaplan, 2002; Kuciene and Dulskiene, 2008; Meberg et al., 2007; van der Bom et al., 2011). The genetics of CHD are also complex; a single candidate gene or genetic variant can produce a spectrum of heart malformations and may even occur in phenotypically normal humans (Table 1). Variance in genetic penetrance also occurs within affected families, resulting in a range of CHD phenotypes. These events increase the difficulty of identifying and characterizing the genetic risk factors for CHD, and emphasize the importance of understanding cardiogenesis at the molecular level (Brown et al., 2004; Farr et al., 2018).

Fig. 1.

Cardiac development in the human embryo. This schematic shows the embryonic development of the human heart through first and second heart field (HF) formation, heart tube formation and pumping, looping, neural crest migration and septation, resulting in a fully developed heart at the end of gestation. Boxed areas highlight structures typically affected by congenital malformations (solid lines: atrial and ventricular septal defects; dotted lines: hypoplastic ventricles; dashed lines: aortic and pulmonary valve defects and defects of the great vessels, such as transposition).

Fig. 1.

Cardiac development in the human embryo. This schematic shows the embryonic development of the human heart through first and second heart field (HF) formation, heart tube formation and pumping, looping, neural crest migration and septation, resulting in a fully developed heart at the end of gestation. Boxed areas highlight structures typically affected by congenital malformations (solid lines: atrial and ventricular septal defects; dotted lines: hypoplastic ventricles; dashed lines: aortic and pulmonary valve defects and defects of the great vessels, such as transposition).

Table 1.

Genes and loci commonly associated with syndromic congenital heart disease

Genes and loci commonly associated with syndromic congenital heart disease
Genes and loci commonly associated with syndromic congenital heart disease

There remains much to learn about the underlying causative mechanisms of CHD, which we are only now beginning to understand through tissue and computational modeling. As a result, existing clinical therapies remain palliative at best as they do not address the underlying mechanisms and causes of CHD. Among the most-influential genetic factors that contribute to CHD are multiple under- or over-expressed genes, incomplete penetrance of variants, aberrant epigenetic controls of gene expression and mechanobiological effects on the developing heart structures.

Here, we review the tissue and computational models that have enabled researchers to uncover several associated and causative factors implicated in various CHD. We aim to provide scientists and clinicians with an overview of the abnormal cardiogenic pathways uncovered by studying single cells, organoids, animal models and, intriguingly, in silico computational models of normal and abnormal human fetal hearts. We discuss how these have led to a better understanding of the impact of genetic or physical insults has on the developing fetal heart, particularly for readers interested in CHD research. The incidence of CHD in adults is increasing steadily with advances in cardiac interventions (Pierpont et al., 2018), which have reduced CHD-related infant mortality, highlighting the pressing need for molecular therapies that address actual causes rather than physical consequences and that can be implemented early in development, such as in the young infant or even the developing fetus.

An overview of CHD

As CHD comprise a heterogeneous group of cardiac malformations, the key clinical manifestation depends on the type of CHD. Classification schemes have evolved substantially since Maude Abbott generated The Atlas of Congenital Cardiac Disease more than 100 years ago (Abbott, 1936). Broadly speaking, CHD can be classified, based on morphology and hemodynamics, into cyanotic and acyanotic CHD (Fig. 2). Cyanosis is a bluish discoloration of the skin and mucous membrane, resulting from reduced oxygen saturation of the circulating blood. Patients with cyanotic CHD have mix of deoxygenated and oxygenated blood, with an overall reduction of circulating oxyhemoglobin.

Fig. 2.

Classification of CHD into acyanotic and cyanotic CHD, and the status of pulmonary blood flow (PBF).

Fig. 2.

Classification of CHD into acyanotic and cyanotic CHD, and the status of pulmonary blood flow (PBF).

Tetralogy of Fallot (TOF) is a common cyanotic CHD characterized by pulmonary stenosis, right ventricular outflow tract obstruction (RVOTO), ventricular septal defect (VSD), overriding aorta and right ventricular hypertrophy (Lahm et al., 2015). Owing to the RVOTO, deoxygenated blood in the right ventricle shunts across the VSD into the left ventricle and out through the aorta to enter systemic circulation, resulting in reduction of systemic oxygen saturation and, consequently, cyanosis. Another example of cyanotic CHD is hypoplastic left heart syndrome (HLHS), a critical CHD characterized by underdevelopment of the left-sided structures of the heart, including the ascending aorta, left ventricle and aortic and mitral valves (Box 1), often culminating in childhood death (Barron et al., 2009; Tchervenkov et al., 2006). HLHS results from diminished ventricular flow and disruption of genetic regulatory networks governing left ventricular chamber development (deAlmeida et al., 2007; Hinton et al., 2007; Iascone et al., 2012).

Common examples of acyanotic CHD include atrial septal defect (ASD) and VSD. In ASD, there is a hole in the septum, the wall that divides the left and right atrium; in VSD, the defect is in the septum between the left and right ventricle. In both defects, oxygenated blood shunts from the oxygen-rich (left) chambers to the oxygen-poor (right) chambers; therefore, these lesions are also termed left-to-right shunts. As the mixing is from the oxygenated to the deoxygenated chambers, there is no reduction in the oxygen saturation in the systemic circulation and, thus, no cyanosis.

The etiology of CHD is multifactorial, with only 30% of cases traced to a known cause (Pierpont et al., 2018). Causes of CHD are divided into genetic (Table 1) and non-genetic. Non-genetic etiologies of CHD such as environmental teratogens, infectious agents, and maternal diabetes mellitus have been widely studied (Fahed et al., 2013; Zhu et al., 2009) but there is still much to be uncovered regarding the genetics and epigenetics of CHD (Fahed et al., 2013; Zhu et al., 2009). Epidemiology points to a strong genetic contribution regarding CHD, including a higher concordance of CHD among monozygotic compared to dizygotic twins, and an elevated recurrence risk of related forms of CHD among siblings (Lyu et al., 2018; Pediatric Cardiac Genomics et al., 2013; Zaidi and Brueckner, 2017). By contrast, there is no family history of CHD for a large proportion of particularly severely affected patients (Burn et al., 1998), which implies that de novo genetic events significantly contribute to the etiology of CHD, including chromosomal abnormalities, point mutations and structural and copy number variants (Zaidi et al., 2013). It is now estimated that ≤35% of CHD cases, with or without extracardiac anomalies, can be attributed to genetic factors, including copy number variation (Breckpot et al., 2010; Goldmuntz et al., 2011; Lalani et al., 2013; Richards et al., 2008; Syrmou et al., 2013; Thienpont et al., 2007), chromosomal abnormalities (Roos-Hesselink et al., 2005; van der Bom et al., 2011) or a single gene mutation (van der Bom et al., 2011).

In vitro modeling of CHD

Evidence accumulated over the past decade demonstrates the power of human induced pluripotent stem cell (iPSC)-based disease modeling as investigative tools for biomedical research and regenerative medicine. iPSCs may, to various degrees, recapitulate the pathophysiological and clinical manifestations of human disease. Additionally, they can be derived from any individual, thereby, providing the ability of unlimited supply of cells bearing clinically relevant phenotypes, presenting unprecedented opportunities to evaluate the etiology and progression of a variety of inherited cardiac diseases.

iPSC-derived models of CHD

The controlled differentiation of iPSCs into cardiomyocytes provides a unique platform with which to establish personalized models of CHD that feature patient-specific factors and phenotypes. As originally reported in 2007, overexpression of the transcription factors Oct3/4 (officially known as POU5F1), Sox2, Klf4 and Myc (also known as Yamanaka factors) can induce human fibroblasts to become pluripotent (Takahashi et al., 2007). Since this ground-breaking study was first published, a plethora of reprogramming cocktails have been designed to convert various somatic cells to pluripotency (Lee et al., 2010; Liao et al., 2008; Maherali et al., 2008; Okita et al., 2013). The key advantages of iPSCs are (1) derivation from human somatic cells, which in some jurisdictions would more readily receive ethical approval for generation and use (Bilic and Izpisua Belmonte, 2012); and (2) preservation of the donor's genomic and epigenetic profile to study precise genetic mutations (Parrotta et al., 2020). Here, we review the use and potential applications of iPSC-derived cardiac models for investigating the pathophysiology of complex CHD and designing novel treatment strategies.

For successful generation of CHD models from iPSCs, it is crucially important to adopt protocols that can induce, expand and purify functional cardiomyocytes. Substantial progress has been made in recent years towards understanding the regulatory mechanisms of cardiac differentiation, and in developing robust strategies to generate iPSC-derived cardiomyocytes efficiently and reproducibly. These differentiation protocols often recapitulate the multi-stage developmental processes of mesoderm commitment and cardiac specification; some involve spontaneous differentiation from embryoid body systems (Kattman et al., 2011; Mummery et al., 2012), whereas others employ adherent monolayer culture systems (Kadari et al., 2015) and chemically defined serum-free cultures (Burridge et al., 2014; van den Berg et al., 2016). These different protocols, nevertheless, typically feature a core set of regulatory networks and signaling pathways that play essential roles in establishing the cardiovascular system, including bone morphogenic proteins (BMPs), fibroblast growth factors (FGFs) and WNT signaling factors (Kattman et al., 2011; Lian et al., 2012; Tirosh-Finkel et al., 2010; Zhao et al., 2019). The iPSC-derived cardiomyocytes generated according to these different protocols typically beat spontaneously, express channels and sarcomeric components, and exhibit Ca2+ transients and action potentials (Karakikes et al., 2015; Ma et al., 2011; Zhang et al., 2009; Zwi et al., 2009).

Hypoplastic left heart syndrome

As discussed earlier, HLHS (Figs 1 and 3A) is a common and serious congenital heart defect that often results in childhood death (Barron et al., 2009; Tchervenkov et al., 2006). Cardiomyocytes have been produced from iPSC lines derived from patients with HLHS, albeit with lesser efficiency than iPSC production from normal donors (Jiang et al., 2014; Kobayashi et al., 2014). Cardiomyocytes generated from HLHS-derived iPSCs showed developmental and functional defects, and more-primitive cardiac phenotypes in gene expression studies compared to control-iPSC lines and embryonic stem cells. The cells demonstrated reduced expression of several key cardiac mesoderm marker genes, including those of mesoderm posterior BHLH transcription factor 1 (MESP1), cardiac troponin T2 (TNNT2) and connexin 43 (GJA1), which mediates electromechanical transduction and cardiomyocyte alignment, and delayed expression of cardiac progenitor marker gene GATA binding protein 4 (GATA4) during cardiac differentiation. HLHS-derived cardiomyocytes also showed accelerated decrease of Ca2+ and generated Ca2+ transients in the presence of caffeine, indicating dysfunctional ryanodine receptors, huge ion channels that mediate Ca2+ release from the sarco/endoplasmic reticulum into the cytoplasm (Jiang et al., 2014; Kobayashi et al., 2014). These findings indicate that HLHS patient-derived cardiomyocytes undergo impaired cardiac lineage differentiation, involving cardiac mesoderm formation, cardiac progenitor maturation and cell commitment to atrial or ventricular phenotypes. A new mechanism for HLHS development has also been discovered in studies that used iPSC, and potentially involves a variant of myosin heavy chain, cardiac muscle alpha isoform (MYH6), i.e. MYH6-R443P, which results in an arginine to proline change in MYH6 (Kim et al., 2020). In this study, iPSC-derived cardiomyocytes were generated from a HLHS-affected family comprising one affected proband and one affected and one unaffected parent. Resulting cardiomyocytes showed dysmorphic sarcomeres (Box 1) that caused decreased atrial contractility and compensatory increased expression of MYH7. When the ‘phenotypically normal’ parental iPSCs were edited by using CRISPR/Cas9 to contain the MYH6-R443P variant, the resulting cardiomyocytes reproduced the HLHS phenotype of the proband. By contrast, correction of the MYH6-R443P variant to wild-type MYH6 by using gene editing in iPSCs of the proband rescued cardiomyogenic differentiation, contractility and velocity, and sarcomere organization. This is the first report describing atrial sarcomere disorganization in cardiac tissues from HLHS patients with MYH6 variants to lead to decreased atrial contractility and, consequently, hypoplastic left ventricular development. Although not unexpected – since MYH6 is more abundantly expressed in atria than in ventricles – these findings demonstrate the feasibility of using HLHS-derived iPSC-cardiomyocytes as in vitro tools to test drugs that target atrial contractility to alleviate HLHS deficiencies, or to design strategies that alter developmental pathways in order to minimize or even prevent development of MYH6-variant-associated HLHS.

Fig. 3.

In vitro modeling of CHD. (A) Patient-derived fibroblasts or other somatic cells are induced into pluripotent stem cells (iPSC) by addition of the Yamanaka factors POU5F1, SOX2, KLF4 and MYC. The resulting iPSCs are subsequently reprogrammed in differentiation medium containing differentiation and growth factors, such as human BMP4, activin, Wnt/β-catenin and FGF to yield cardiomyocytes bearing the gene defects of interest; i.e. reduced gene expression (HLHS), mutations (TOF, septal defects). (B) The resulting individual iPSCs can be used to study the cellular and molecular signatures in 2D cell structures relevant in a number of CHDs or can be further cultured in 3D to form cardiac organoids to study other environmental interactions. TNNT2, cardiac troponin 2; WT1, WT1 transcription factor (Wilms tumor 1); TJP, tight junction proteins 1, 2 and 3; THY1, Thy-1 cell surface antigen, VIM, vimentin; PECAM1, platelet and endothelial cell adhesion molecule 1; NFATC1, nuclear factor of activated T cells 1 (cytoplasmic).

Fig. 3.

In vitro modeling of CHD. (A) Patient-derived fibroblasts or other somatic cells are induced into pluripotent stem cells (iPSC) by addition of the Yamanaka factors POU5F1, SOX2, KLF4 and MYC. The resulting iPSCs are subsequently reprogrammed in differentiation medium containing differentiation and growth factors, such as human BMP4, activin, Wnt/β-catenin and FGF to yield cardiomyocytes bearing the gene defects of interest; i.e. reduced gene expression (HLHS), mutations (TOF, septal defects). (B) The resulting individual iPSCs can be used to study the cellular and molecular signatures in 2D cell structures relevant in a number of CHDs or can be further cultured in 3D to form cardiac organoids to study other environmental interactions. TNNT2, cardiac troponin 2; WT1, WT1 transcription factor (Wilms tumor 1); TJP, tight junction proteins 1, 2 and 3; THY1, Thy-1 cell surface antigen, VIM, vimentin; PECAM1, platelet and endothelial cell adhesion molecule 1; NFATC1, nuclear factor of activated T cells 1 (cytoplasmic).

Tetralogy of Fallot

As described above, TOF (see Figs 1 and 3A) causes a severe cyanotic form of CHD (Lahm et al., 2015). In 2020, Grunert et al. generated iPSCs from two well-characterized TOF patients and differentiated them into cardiomyocytes by using standard methods (Grunert et al., 2020). The authors described using TOF-derived iPSC-cardiomyocytes for the first time as an in vitro tool to study the transcriptional landscape of TOF. Some of their findings included disease-causing mutations in tumor protein p53 (TP53), which is important for smooth muscle cell migration, fibulin 2 (FBLN2), which regulates atrioventricular valvular-septal morphogenesis, and in dishevelled associated activator of morphogenesis 2 (DAAM2), which mediates sarcomere assembly and myocardial maturation (Grunert et al., 2020). In a second study, Han et al. generated iPSCs from a child diagnosed with TOF, who carried a mutation in T-box transcription factor 1 (TBX1) (Han et al., 2020). TBX1 is known to be involved in cardiac outflow tract formation, demonstrating the contribution of the TBX1 mutation to TOF-associated conotruncal defects (Box 1) (Hasten et al., 2018). Such patient-specific iPSC lines are useful tools for proving that a genetic mutation causes the observed phenotypes and for investigating gene therapy options for TOF. For example, homozygous correction of the c.455G>A mutation in the DAND5 gene has been achieved by using CRISPR/Cas9 (Inácio et al., 2020). DAND5 is involved in the NODAL signaling pathway that controls early embryo patterning, mesoderm and endoderm formation, and the specification of left-right asymmetry (Barnes and Black, 2016).

Septal defects

VSD and ASD account for almost 50% of all CHD (Hoffman, 1995). In mice, deletion of Gata4 results in cardia bifida (Box 1) and early fetal death (Misra et al., 2012). In humans, the heterozygous missense mutation G296S in the cardiac transcription factor GATA4 causes ASD, VSD and pulmonary valve stenosis (Garg et al., 2003). In vitro investigations demonstrated that this GATA4-G296S mutation reduces DNA-binding affinity and transcriptional activity of the resulting GATA4 mutant protein, and abolishes its interaction with T-box transcription factor 5 (TBX5), which is important for normal heart formation (Ang et al., 2016; Garg et al., 2003). Cardiomyocytes derived from iPSCs of patients harboring GATA4 mutations also show disruption of interactions between transcriptional proteins; an example is the loss of TBX5 recruitment to the cardiac super-enhancers (elements that are associated with genes for heart development and muscle contraction). This leads to impaired cardiomyocyte contractility, Ca2+ handling and metabolic activity (Ang et al., 2016). Taken together, these studies confirm the crucial role of GATA4 and TBX5 in maintaining the robust cardiac gene programming necessary for the prevention of septal defects in both murine and human cardiovascular systems (CVSs).

Organoids and engineered 3D models

Despite the importance of traditional 2D models, in particular their cost-effectiveness, reproducibility and complementarity to animal models, their ability to recapitulate in vivo cardiac architecture and simulate cardiac tissue complexity and organization is limited. Cell-cell interactions are restricted to side-by-side contact. This limitation, potentially, affects cellular morphology, survival and proliferation, which in turn restricts the study of disease mechanisms in such 2D settings. To circumvent the intrinsic limitations of conventional 2D cell cultures, human cardiac organoids have been generated using 3D culture systems, which can more closely recapitulate endogenous cardiomyocyte tissues by incorporating key elements of the cardiac system, including supporting scaffolds and external stimuli (Fig. 3B). Thus, cardiac organoids are, potentially, useful for investigating human heart development and the etiology of CHD, as they facilitate physiologically relevant experiments that would be ethically or technically impossible to perform in vivo. 3D cardiac organoids also provide more-robust biological models for studying tissue regeneration and performing drug toxicity screens.

Studies have shown that human iPSC-derived cardiomyocytes cultured in 3D undergo greater maturation, exhibit improved adult cardiomyocyte physiology, and show increased contractility and electrical function relative to cardiomyocytes differentiated and cultured in 2D (Correia et al., 2018; Karbassi et al., 2020; Lemoine et al., 2017). 3D cardiac organoids derived from human iPSCs also enabled modeling of early heart field development (Andersen et al., 2018) and aided identification of signaling pathways involved in the development of both first and second heart fields (Box 1) (Kelly et al., 2014), furthering our understanding of heart field evolution and chamber-specific cardiac pathogenesis (Andersen et al., 2018).

Cardiac organoids can be generated using a highly efficient, scalable and reproducible method (Israeli et al., 2020 preprint), and can be differentiated from several human iPSC lines by using stepwise chemical modulation of the Wnt signaling pathway under completely defined culture conditions. The resulting organoids start to contract around day 6 of culture and grow up to 1mm in diameter by day 15 post differentiation. Cardiac organoids consist of multiple cell types, and can exhibit interconnected internal chambers and morphological complexity similar to that of the human fetal heart (Israeli et al., 2020 preprint). The main cardiac-related lineages contained within these organoids, as exhibited on confocal microscopy and cell marker analyses, include cardiomyocytes (TNNT2), epicardial cells (WT1, TJP), cardiac fibroblasts (THY1, VIM), endothelial cells (PECAM1) and endocardial cells (NFATC1). Provided in parenthesis are the marker genes generally used to isolate or identify the cells of interest (Israeli et al., 2020 preprint).

The ability to evaluate contractile function and Ca2+ handling of cardiomyocytes is highly relevant for studying the most common cause of cardiac death in CHD patients – chronic heart failure. To recapitulate more closely the endogenous architecture and mechanics of cardiomyocytes, 3D-engineered heart tissues (EHT) have been developed. 3D EHTs are generally constructed using both cardiomyocytes and non-cardiomyocytes, such as fibroblasts and endothelial cells, mixed with biodegradable matrices that can be polymerized into sheets, cylinders or rings. With cyclic strain and electrical pacing in 3D culture, these cells first remodel the extracellular matrix (ECM), over time forming a structure resembling the intact myocardium with uniformly aligned cardiomyocytes that form gap junctions. The EHTs generally remain viable in culture for 4–6 weeks without overgrowth of non-cardiomyocytes or loss of contractile performance; various testing apparatuses can be used to measure Ca2+ transients, action potentials and twitch forces (Hirt et al., 2014). EHTs are also constructed by casting hydrogel-containing cardiomyocytes into moulds maintained under a mechanically defined load. Subsequently, cardiomyocytes in the hydrogel remodel and align relative to the mechanical load to form coherently beating syncytia (Lee et al., 2017). This configuration allows the analysis of contractile force under stable conditions that simulate the cardiac microenvironment, i.e. 3D heart-like muscle strips that contract under auxotonic conditions (Mannhardt et al., 2016). Mature 3D EHTs have been fabricated from iPSCs in a custom-designed bioreactor that generates electromechanical stimulation (Ronaldson-Bouchard et al., 2019), an attractive system with great potential for translational research, enabling assessment of cardiac contractility, and facilitating investigation of various genetic mutations and molecular mechanisms of CHD-related heart failure. Recently, by using single-cell RNA sequencing (scRNAseq), 2D and 3D cardiac tissues served as powerful tools with which to understand the transcriptomic basis and molecular regulation of iPSC-cardiomyocytes that carry disease-associated genes, further expanding the repertoire of CHD research (Lam et al., 2020; Ni et al., 2020).

Challenges

Patient-derived iPSCs and iPSC-derived cardiomyocytes have well-established advantages as models of cardiac disease, given that they recapitulate the pathophysiological characteristics of a broad range of CHDs (Ebert et al., 2012; Parrotta et al., 2020). However, it is uncertain how closely these in vitro iPSCs and iPSC-derived cardiomyocytes recapitulate the disease phenotype of the donor, particularly its severity. In vitro differentiated cardiomyocytes have, typically, immature structures and function (Karakikes et al., 2015; Veerman et al., 2015). Their gene expression profiles resemble those of fetal cardiomyocytes, with smaller resting membrane potentials and upstroke velocities, relatively fewer mitochondria, and shorter and more disorganized sarcomeres relative to those of adult cardiomyocytes (Gherghiceanu et al., 2011; Karakikes et al., 2015; Koivumäki et al., 2018; Veerman et al., 2015; Yang et al., 2014). Incomplete pluripotency reprogramming and errors that occur during reprogramming can also alter phenotype and function. Atypical methylation patterns and genetic mutations in iPSCs are both associated with variants in parental somatic cells, with reprogramming or with the duration of in vitro culture (Yoshihara et al., 2017). Additionally, the various differentiation protocols used can result in iPSC-derived cardiomyocytes, with heterogenous subtypes resembling atrial, ventricular and nodal cells (Burridge et al., 2015). With varying ratios of cardiomyocyte subtypes, the comparability of readouts, like contractile force and electrophysiological activity, may be compromised (Streckfuss-Bömeke et al., 2013; van den Heuvel et al., 2014). Another limitation of iPSC-generated disease models is their inability to fully recapitulate a clinical condition in vitro, often due to absence of the complicated interplay among multiple genetic and environmental factors, including age, sex and ethnicity. It is important not to overlook these relative disadvantages when interpreting data from iPSC-derived cardiomyocytes. This discussion, reviewed in detail elsewhere (Doss and Sachinidis, 2019) highlights the importance of ongoing strategies in developing authentic health and disease models, and their complementary bioassays for the functional testing of drugs and other therapeutic interventions.

In vivo modeling of CHD

Animal models, large and small, have been instrumental in the discovery of the multigenic nature of CHD and its complex cardiac morphologies (Kraft and de Andrade, 2003). Large, conserved networks of genes control early cardiogenesis and developing cardiac function, and many are highly conserved in lower-order animals that can be genetically manipulated, bred and studied with relative ease (Bier and Bodmer, 2004). With a range of widely available molecular techniques, such as morpholino-oligonucleotides (MOs) and gene editing techniques (reviewed by Kaltenbrun et al., 2011; Zhang and Liu, 2015), candidate genes can be knocked down, knocked out or knocked in, enabling studies of their causality and supporting genetic screens. Higher-order animals have, as yet, not been used to model CHD due to the technical difficulties, time and expense involved in breeding, for example, transgenic sheep or non-human primates through assisted reproductive programmes, and due to the lack of naturally occurring structural anomalies that resemble human CHD (Moran et al., 2015; Proudfoot et al., 2015). Nevertheless, large animals are valuable as surgical models of outflow track obstruction, complex vascular configurations that mimic TOF, and aortic or pulmonary stenosis – both of which can be produced in utero (Table 2).

Table 2.

Animal models of CHD used in mechanistic and therapeutic studies

Animal models of CHD used in mechanistic and therapeutic studies
Animal models of CHD used in mechanistic and therapeutic studies

The main advantage of using lower-order animal species to model CHD are their low cost, fast reproduction times and genetic tractability, made by possible by the fact that the genetic and epigenetic elements that contribute to human CHD are highly conserved in fly, zebrafish, frog, chicken and mouse (Bakkers, 2011; Bier and Bodmer, 2004; Gut et al., 2017; Kooij et al., 2014; Sparrow et al., 2000). These models were used to great effect to investigate the impact of genes and gene dosing (via transgenesis), and of silencing (via MOs) on the aetiology and pathology of CHD. They were also used to investigate new candidate genes that had been detected through whole-genome sequencing of patient samples and then introduced into animal genomes by using precise molecular tools (Kaltenbrun et al., 2011). Together, these models helped to uncover the diverse genetic and non-genetic mechanisms that regulate cardiovascular physiological and pathological processes, ECM remodeling and mechanobiology (Table 2). In this section, we discuss the advantages and limitations of each animal model of CHD, as well as the genetic screening and corrective strategies that rely on a certainty of causation – which can be difficult to prove in human patients.

Zebrafish

Although the zebrafish (Danio rerio) heart develops into a simple, single ventricular structure, its cardiac function is similar to that of the human CVS. In addition, heart development in zebrafish relies on the same proteins as does heart development in humans (Farr et al., 2018; Khodiyar et al., 2013) and can also be altered by certain environmental toxins (reviewed by Brown et al., 2016). The heart rate and electrophysiological cardiac behavior of zebrafish are similar to these parameters in human (Genge et al., 2016; Sedmera et al., 2003). Genetic pathways influential in cardiogenesis are also highly conserved between humans and zebrafish, and specific molecular and cellular tools are available to investigate their function, including MOs, gene-editing tools and cardiac-specific fluorescent transgenes that can be used to investigate cardiovascular pathology in individual heart structures (Gut et al., 2017). Zebrafish embryos are also fertilized externally and are transparent, facilitating their real-time live imaging using high-speed fluorescent microscopy. However, the simple configuration of zebrafish hearts makes studying the structurally complex CHD abnormalities challenging, as the phenotypes caused by CHD-associated genetic mutations might be quite different compared with their human counterparts (Poon and Brand, 2013). Other species-specific differences include the absence of a His-Purkinje system (Box 1) (Sedmera et al., 2003), and genetic redundancy, which can hamper gene dosage studies (Peng, 2019). In addition, there is the challenge of accurately measuring cardiac volume and output, which are often prone to bias, from 2D zebrafish images (Akerberg et al., 2019). Given these limitations, zebrafish is best used to screen for novel drug therapies rather than to assess genetic, cell or surgical therapies, which limits the translational applicability of this model.

Frog

The African clawed frog (Xenopus laevis) has contributed greatly to our understanding of vertebrate development and gene function, with the availability of established molecular tools with which to investigate, by gene knock-in or knockdown, the functions of candidate genes (Kaltenbrun et al., 2011). Like the human heart, the Xenopus linear heart tube structure undergoes looping and multi-chamber formation, and demonstrates atrial and ventricular excitation wave patterns that are similar to His-Purkinje conduction in humans (Sedmera et al., 2003). Cardiac function is not required for embryo development; thus, molecular manipulation of the heart can be conducted in vivo (Mohun et al., 2000), including the termination of important signaling pathways. Like the zebrafish, 3D models of Xenopus cardiac stages can be constructed by using serial sectioning and microscopy imaging. In vivo, longitudinal 4D imaging can also be performed; for example, by using Doppler optical coherence tomography (Box 1), to study cardiac structure, conduction and blood flow, due to the large size and ease of manipulation of this model (Jenkins et al., 2012; Kaltenbrun et al., 2011; Mohun et al., 2000). MOs or transgenesis can both be used to investigate the roles of individual genes in normal cardiogenesis, including promoter–enhancer analyses, sequence mapping and the investigation of complex gene regulatory networks. However, the ability to perform gene targeting in Xenopus embryos to study cardiogenic proteins is hampered by low-level transgene expression and mosaicism in cardiogenic tissues that make dose-dependent effects difficult to appreciate, by insufficient sensitivity of standard in situ hybridization to detect low-level protein expression in cardiogenic tissues and the high mortality of gene-targeted embryos, as a high expression of transgenes can interfere with the normal expression of cardiogenic markers (Cleaver et al., 1996; Fu et al., 1998; Horb and Thomsen, 1999). Aberrant phenotypes may only occur with overexpression of transgenic proteins, thus restricting this translational utility of this model. Nonetheless, the Xenopus model has been informative for studying complex CHD (Table 2), its usefulness derived from elucidating the role of genes that are crucial in early cardiogenesis, i.e. the tinman homologs Nkx2-3 and Nkx2-5 9 (Box 1), and in causation of septal defects, i.e. transcription factor Tbx5, and heterotaxy syndrome (see Box 1), i.e. Zinc finger of the cerebellum protein Zic3 (Fu et al., 1998; Horb and Thomsen, 1999; Kitaguchi et al., 2000).

Fly

Drosophila melanogaster is an extremely versatile and genetically tractable model, with readily available molecular tools that can be used to gain new insights into cardiogenesis and heart physiology (Camacho et al., 2016; Rotstein and Paululat, 2016; Vogler and Bodmer, 2015). Unlike the closed circulatory system that comes with high hydrostatic pressure in humans, the developing Drosophila heart is a simple structure, comprising a heart tube that pumps haemolymph in an open circulatory system at low hydrostatic pressure and starts beating upon completion of embryogenesis (Rotstein and Paululat, 2016). This simple, in vivo model provides the means to evaluate evolutionarily conserved genetic control mechanisms and developmental processes (Bier and Bodmer, 2004). There is general functional conservation of genes, cardiac proteins, ECM components (e.g. collagens, laminins and integrins) and cardiac precursors between flies and mammals (Bryantsev and Cripps, 2009; Hughes and Jacobs, 2017; Kooij et al., 2014; Reim and Frasch, 2005). Indeed, ∼30% of human disease-causing genes with homologs in Drosophila have the same function in humans and flies (Bier and Bodmer, 2004; Reiter et al., 2001). Tools have also been developed to perform large-scale genetic screens to identify disease-causing mutations, such as the novel overexpression activity (NOVA) screen (Guichard et al., 2002). Because the Drosophila embryo does not rely on the circulatory function of the heart during development, physical, cellular and subcellular manipulations to its cardiac function can be performed during embryogenesis to assess their downstream effects in vivo (Sénatore et al., 2010), a feat that cannot be achieved in higher-order vertebrates. This has advanced our understanding of the crucial roles that certain signaling pathways play in cardiogenesis, including the function of the Slit-Robo complex (Box 1), ECM components involved in cardiomyocyte contractility (Nishimura et al., 2014; Vogler and Bodmer, 2015) and of the highly conserved homeobox protein Nkx2-5 (Tanaka et al., 2001). Mutations in genes that encode components of such pathways are associated with septal and valvular defects, and with abnormal venous connections to the heart, including TOF and atrioventricular conduction delays (Bartlett et al., 2007; Schott et al., 1998). Live imaging of Drosophila cardiac development can also be enhanced by using fluorescent reporter transgenes (Morin et al., 2001). Thus, the simple genetic and cardiac system of Drosophila is an advantage when attempting to comprehend the multi-gene interactions that contribute to CHD. Limited by its simple structure, Drosophila cannot be used for observing morphological defects that arise from cardiac looping, septation or conduction events (Bier and Bodmer, 2004).

Chick

In the chick (Gallus gallus domesticus), cardiac progenitors, migrating cardiac neural crest cells and a proepicardium (Box 1) develop to give rise to a primary heart tube, which then undergoes looping to form chambers, outflow tracks and a conduction system (Martinsen, 2005). In addition to the analysis of genetic mutations, chick embryos are useful for studying the mechanobiology of the CVS and the consequences of hemodynamic perturbations, as they are amenable to simple surgical manipulations, including cardiac neural crest ablation and outflow track obstruction; they can also be imaged in ovo by using optical coherence tomography (Midgett et al., 2017). Our current understanding of how faulty cardiac neural crest cell migration and abnormal endocardial cushion development contribute to cardiac outflow tract malformations, aorta and pulmonary trunk mispositioning, and myocardial dysfunction, derive from studies in the chick (Hutson and Kirby, 2007; Kirby et al., 1985, 1983; Yelbuz et al., 2002). Models of outflow track obstruction created by ligation of the left pulmonary artery or the left atrial appendage in the chick have been used to elucidate the relationship between hemodynamic flow, vessel wall shear stress, chamber and valve formation and myocardial function, as summarized in Table 2 (Johnson Kameny et al., 2019; Tobita and Keller, 2000). The chick also lends itself to the study of signaling pathways involved in cardiogenesis, being more readily genetically manipulable than mice. Mutations studied in the chick embryo are summarized in Table 2. In terms of genome manipulation, the chick is lagging behind the mouse due to technical difficulties with introducing tools for gain- or loss-of-function (Gammill and Krull, 2011; Sauka-Spengler and Barembaum, 2008). The recent technical ability to transiently knockdown endogenous protein, primarily by using short-hairpin RNAs (shRNAs) expressed from plasmids to interfere with RNA expression (continuous shRNAs production, inexpensive) or antisense MO (effect diluted with cell divisions, expensive), both introduced by electroporation, offers expanded opportunities to effectively study functions of specific genes (Dai et al., 2005; Eisen and Smith, 2008). Although not as genetically tractable as the mouse – in terms of simulating syndromic CHD – the chick is still a valuable model for structural cardiac diseases (Suzuki et al., 2009; Tutarel et al., 2005); however, it is important to note that certain cardiac events differ between chicks and humans, including the development of the septum secundium and the pharyngeal arch artery system. As such, it might not always be possible to faithfully recapitulate in chicks the abnormal cardiogenesis present in human CHD patients (Martinsen, 2005).

Mouse

In the mouse (Mus musculus), the heart begins as a linear tube, which then loops and forms a four-chambered structure, and then undergoes further development, including septation and vessel formation (Andersen et al., 2014; Moorman and Christoffels, 2003). Numerous human disease models have been generated in mice by introducing or disrupting genes of interest in mouse embryos. In the case of CHD, the resulting abnormal phenotypes have provided many new insights into the molecular mechanisms of human cardiac disease, as summarized in Table 2 (Kloesel et al., 2016). The significant anatomical and developmental similarities that exist between mice and humans, i.e. cardiac looping, valvuloseptal formation, myocardial compaction and trabeculation (Desgrange et al., 2018; Ivanovitch et al., 2017; Le Garrec et al., 2013; Zaffran et al., 2004), allow clinically relevant data to be derived from some mouse models (Wessels and Sedmera, 2003). In addition, genetic knockout mouse models of cardiac neural crest mutations recapitulate the phenotypes of physical neural crest ablation in the chick (Hutson and Kirby, 2003, 2007).

Mice have been used to develop a number of CHD models, in which mechanistic pathways can be analyzed at cellular and molecular levels. The effective utilization of complementary advanced genetics, e.g. whole exome sequencing, bulk transcriptomics and scRNAseq, in transgenic mice that facilitate the study of morphological sequelae have enabled scientists to examine essential signaling and regulatory networks in cardiogenesis, uncovering stage-specific disturbances related to CHD (Li et al., 2014a,b). These powerful tools provide personalized models of genetic mutations that were identified in individuals with CHD, and help to understand gene functions and regulatory networks. Six significant examples are given in the following: (1) Adamts19 knockout mice exhibit similar valvular disease as affected patients, thereby demonstrating its role in the perturbation of the WNT–ADAMTS19–KLF2 axis that is crucial for valvular development (Wünnemann et al., 2020). (2) The importance of semaphorin (Sema)/plexin signaling in cardiomyopathies was demonstrated by knocking down Plxnd1, a class-3 Sema receptor in endothelial cells, and caused excessive myocardial trabeculation and impaired compaction associated with Notch1 overexpression (Sandireddy et al., 2019). (3) Cited2 mutations that perturb transcription factors and cardiopoietic gene expression in knockout mice, result in disrupted cardiac differentiation and septal defects (Santos et al., 2019). A defective transcription factor HAND1 in left ventricle cardiomyocytes, leads to disruption of regulatory pathways that control cardiac hypertrophy, cardiomyocyte proliferation, balanced development of cardiac cell lineages, septation, conduction and trabeculation (de Soysa et al., 2019; Firulli et al., 2020; Han et al., 2019). (5) Abberant endothelial-to-mesenchymal transition disrupts outflow tract formation (Liu et al., 2019). (6) Notch pathway disruption and abnormal cardiac cell behavior cause abnormal trabeculation, defective compaction and cardiomyopathy (Del Monte-Nieto et al., 2018).

Mice have also been used to assess the impact of environmental causes of CHD and to elucidate pathways of action at single-cell level. Examples include studies using teratogenic drugs, such as valproic acid, which is commonly used to treat epilepsy and inhibits histone deacetylases, thus, affecting transcription factors involved in cardiac structure development and cardiomyocyte differentiation, giving rise to myocardial disorganization and contractile dysfunction (Philbrook et al., 2019). Intrauterine hypoxia, e.g. during uteroplacental insufficiency that causes fetal growth restriction, reduces cellular proliferation in the second heart field due to downregulated FGF signaling, thereby causing global downregulation of protein synthesis of, for example, Nkx2-5 and hypoxia-inducible factor 1α, as well as perturbation of cardiac outflow tract formation, alignment or elongation, resulting in VSD, overriding aorta, double-outlet right ventricle (DORV) or transposition of great arteries (Box 1) (Moreau et al., 2019; Shi et al., 2016). Persistent hyperglycemia of gestational diabetes can lead to (i) embryopathy through enhanced production of reactive oxygen or reactive nitrogen species, thereby causing endothelial stress (Kumar et al., 2007, 2008; Moazzen et al., 2014; Wentzel et al., 2008); (ii) transcriptional alterations of genes crucial for cardiac development, e.g. Vegfa, Bmp4, sonic hedgehog homolog, thereby leading to increased apoptosis in the heart (Moazzen et al., 2014; Wentzel et al., 2008); (iii) inhibition of Wnt and Notch signaling (Wang et al., 2015); (iv) other cardinal transcriptional pathways, summarized in (Basu and Garg, 2018); (v) hyperglycemia-driven inhibition of cardiac stem cell differentiation and cardiomyocyte maturation (Nakano et al., 2017; Yang et al., 2016), and (vi) overstimulation of myocardial growth (Gordon et al., 2015).

Yet, mouse models are not without limitations. The murine CVS, for example, differs in important aspects from that of humans, i.e. regarding its substantially higher heart rate and distinct myocardial Na+/Ca2+ exchange mechanisms (Ottolia et al., 2013). Developing mouse knockout models is costly, as it requires assisted fertility and embryo manipulation. In addition, severe complex heart diseases are often perinatal lethal, leaving researchers with mice that express milder forms of disease (Bier and Bodmer, 2004). Despite these limitations, translational CHD research has benefited greatly from transgenic mice, and this model continues to be crucial in the search for molecular therapies to target aberrant genetic pathways.

Sheep

The heart of sheep (Ovis aries) shares significant anatomical and physiological homologies with the human heart, providing a model with which to understand normal and pathological cardiogenesis at structural, cellular and molecular levels (Fishman et al., 1978; Weil et al., 1993). The size, gestation and physiology of the sheep fetus also enhances the value of this model, facilitating the study of in utero and perinatal effects of complex CHD (Johnson et al., 2014; Valdeomillos et al., 2019). CHD models in sheep are all surgically generated and used to investigate abnormal cardiac mechanobiology (summarized in Table 2). As twinning is common in sheep, this system naturally provides non-treated controls that can be studied in tandem.

Surgically generated models of CHD in sheep have, particularly, advanced our understanding of the developmental pathophysiology of outflow track obstruction. For example, they have revealed maladaptive changes that result in uncompensated right ventricular hypertrophy (RVH) and progress to heart failure (Montgomery et al., 1998), differences in miRNA expression profiles in compensated and uncompensated RVH, as well as the effects of failed repair (Johnson Kameny et al., 2019; Kameny et al., 2018; Van Puyvelde et al., 2019).

The main limitations of sheep as a species in which to model CHD is their lack of genetic tractability. Transgenic sheep would be prohibitively expensive to produce and, thus, to prove causation between candidate genes and observed morphology would be impossible. Nevertheless, surgical interventions can be trialed and tailored to the disorder studied, and sheep are valuable models to evaluate short- and long-term outcomes of intrauterine or early neonatal therapies (Schmidt et al., 2008). The sheep model also remains highly useful to study the mechanical sequelae of outflow tract obstruction, particularly because, as we cover in the next section, mechanobiology contributes greatly to our understanding of non-genetic non-syndromic CHD.

In silico models of CHD

A recent novel approach to CHD research is the development of organ-level computational models to study the biomechanics and physiological functions of the prenatal CVS (Dewan et al., 2017; Trusty et al., 2018; Wiputra et al., 2018). This approach has been extrapolated from computational tools used to study adult CVS diseases (Meoli et al., 2015; Morris et al., 2016). The motivation for studying prenatal CVS biomechanics is based on evidence that cardiac biomechanics are crucial in the multifactorial aetiology of human CHD, highlighted in animal models (Table 2). For example, outflow obstruction that alters haemodynamic wall shear stresses in chick and zebrafish embryos leads to cardiac underdevelopment and left ventricular hypertrophy (LVH) (Hove et al., 2003; Tobita and Keller, 2000). These changes occur through altered expression of ET1, NOS3, KLF2 and PIEZO1, all of which are flow-responsive and mechanosensitive proteins, and crucially important for trabeculation and valve development (Duchemin et al., 2019; Faucherre et al., 2020; Groenendijk et al., 2007; Rasouli et al., 2018). Further, in silent heart (sih/−) zebrafish embryos, which carry a tnnt2 mutation, absence of mechanical stimuli from cardiac contractions prevented the formation of atrioventricular cardiac cushions (Bartman et al., 2004). Together, these data demonstrate conclusively the importance of biomechanics in cardiogenesis and the need to appreciate the mechanobiology of normal heart development.

The emergence of fetoscopy (Box 1) and ultrasound-guided interventions to treat CHD patients in utero demonstrates the potential benefit of prenatal biomechanical therapies to alter the course of cardiac morphological development, and mitigate the risks of structural abnormalities at birth (Freud et al., 2014; Gardiner, 2019; Tulzer et al., 2018). These interventions typically take place during the late-second to early-third trimesters, when a detailed 2D echocardiographic (2DE) study of structural abnormalities is feasible and when ultrasound-guided invasive procedures can be performed in anticipation of correcting the developmental path towards malformation. To give an example of such approach, in fetal critical aortic stenosis with evolving HLHS, the LV grows normally until aortic outflow obstruction occurs mid-gestation, causing abnormal ventricular pressures, contractile motions and, eventually HLHS – a potentially fatal condition that can only be treated by transplantation (Makikallio et al., 2006; Saraf et al., 2019). In utero aortic valvuloplasty (Box 1) can be performed in affected fetuses to relieve outflow obstruction, and has increased biventricular cardiac function at birth from 26% (without intervention) to 59% (Friedman et al., 2018; Makikallio et al., 2006).

It is, thus, important to understand the biomechanical characteristics of the fetal heart, and their effect on cardiac morphological and functional development. A good way to achieve this is by using computational modeling, which can assess current biomechanical characteristics and predict the effects of any changes in developing cardiac structures. Computational models can also capture fetal heart growth and remodeling in response to biomechanical and haemodynamic alterations. These models can be broadly categorized into models of blood flow models (Courchaine and Rugonyi, 2018) and tissue mechanics (Dewan et al., 2017). Fig. 4 provides a schematic of the modeling process. Blood flow modeling is performed by using computational fluid dynamics (CFD) simulations, which involve computational reconstructions of the blood volume of interest, dividing it into many small elements (meshing), followed by iteratively solving the Navier–Stokes governing equations (Box 1) of fluid motion for each of these elements, in order to derive a description of flow and flow forces within the volume (Bluestein, 2017). CFD modeling of the human fetal heart has been performed in normal healthy environments, and in congenital heart malformations, and is explained as follows. 4D clinical ultrasound fetal heart images were obtained using spatio–temporal image correlation (STIC; Box 1) from structurally normal human fetuses between 20–32 weeks of gestation. These images were used to extract data on ventricular anatomy and ventricular wall motion, which were then used to generate fluid dynamic simulations (Lai et al., 2016; Wiputra et al., 2016). Collectively, Lai et al., and Wiputra et al. described the same flow patterns for the left and right ventricles – a pair of vortex rings emanating from the mitral or tricupsid inlet during diastole, corresponding to E- and A-waves (Box 1) – the primary mechanisms by which flow shear stresses are imposed on the endocardial wall. The fetal right ventricle (RV) has been reported to contract with a wave-like motion from the sinus to the infundibulum; computational modeling indicates that this is to conserve energy required for blood ejection (Wiputra et al., 2017). The same approach was used to model the flow in hearts of fetuses with TOF between 21 and 32 weeks of gestation (Wiputra et al., 2018). Wiputra and colleagues observed that TOF hearts need to work harder to eject blood against the outlet obstruction (pulmonary valve stenosis), and the consequential high-volume flow through the RV caused higher wall shear stresses and chaotic flow patterns. In these fetal hearts, VSD – which is part of TOF – created a connection between both ventricles to equilibrate their pressures and to prevent higher pressures in either ventricle (Wiputra et al., 2018). These alterations are considered influential on the developing morphology of the fetal heart.

Fig. 4.

Computational simulation of the human fetal heart. Summary of the steps involved in in-silico modeling based on biomechanics derived from clinical ultrasound images. From left to right: Images are first processed to reconstruct the anatomy of the relevant cardiac structures. Motion tracking and modeling, as well as digital reconstruction can be carried out to accurately extract and mathematically model their motion. This is followed by finite element modeling of myocardial mechanics to understand myocardial stresses and strains, and to evaluate cardiac function under hypothetical conditions. Computational fluid dynamics can be simulated to reveal details of flow patterns and forces. Clinical and computational reconstruction images captured and developed by C.H.Y.

Fig. 4.

Computational simulation of the human fetal heart. Summary of the steps involved in in-silico modeling based on biomechanics derived from clinical ultrasound images. From left to right: Images are first processed to reconstruct the anatomy of the relevant cardiac structures. Motion tracking and modeling, as well as digital reconstruction can be carried out to accurately extract and mathematically model their motion. This is followed by finite element modeling of myocardial mechanics to understand myocardial stresses and strains, and to evaluate cardiac function under hypothetical conditions. Computational fluid dynamics can be simulated to reveal details of flow patterns and forces. Clinical and computational reconstruction images captured and developed by C.H.Y.

Computational modeling of blood flow dynamics in fetal blood vessels have also been conducted to better understand flow patterns and forces in fetal coarctation of the aorta (CoA; Box 1) (Chen et al., 2017). 3D STIC fetal echocardiography images of the aortic and ductal arches of normal human fetuses at 32 weeks of gestation were obtained and computational models with reduced aortic isthmus (Box 1) dimensions created to emulate CoA. Chen and colleagues demonstrated that increased narrowing of the isthmus modifies vortical and helical flow patterns, causing substantial increase in velocity and wall shear stresses, and reduction in descending aortic pressure. Given that prenatal diagnosis of CoA can be challenging (Buyens et al., 2012; Familiari et al., 2017), such computational modeling may prove useful to enhance diagnostic accuracy of related haemodynamic changes, which, in turn, may influence patient morbidity and survival (Franklin et al., 2002). In another study, computational modeling of blood flow in the ductus venosus (DV; Box 1) in chromosomally or structurally abnormal fetuses was performed using 3D computational simulation and Doppler measurements, and alterations in the characteristic DV flow patterns and pressures were observed (Rezaee and Hassani, 2016).

In addition to 3D flow simulations, fetal circulation has also been modeled using a one-dimensional approach, known as Windkessel or lumped-parameter modeling (Shi et al., 2011). In this approach, a network of electronic circuitry is used as an analogue for vascular flow resistance and compliances. This approach removes the complexity of modeling any one organ, allowing the entire circulatory system to be included in the model, which is useful for testing systemic responses to pathophysiological changes in CHD. This approach has been used to model fetal circulation in pregnancies complicated by intrauterine growth restrictions, in order to understand reactive changes to cerebral and placental vascular resistance (Garcia-Canadilla et al., 2014; van den Wijngaard et al., 2006). Pennati et al. have also developed a carefully considered Windkessel fetal circulation model based on in vivo measurements, and have recommended scaling factors that can be applied to this model at various gestational ages (Pennati and Fumero, 2000).

Tissue mechanics modeling of the fetal CVS has been achieved mainly in the form of finite element modeling (FEM) of myocardial biomechanics. The FEM method divides the solid structure of interest into many small elements, and iteratively solves the governing equations of momentum and deformational mechanics within these elements, to derive a prediction of the deformational and stress characteristics of the object of interest (Lluch et al., 2019). The technique has originally been developed by civil and aeronautical engineers to study structural stability and aircraft structural integrity respectively, and was later imported into biomedical engineering to study myocardial and vascular biomechanics (Costopoulos et al., 2017; Shavik et al., 2018). In one FEM modeling study of the human fetal heart, anatomic dimensions were extracted from 2DE studies of normal fetuses between 20 and 37 weeks of gestation. Myocardial stiffness and active contractions have been mathematically described with an anisotropic hyperelastic constitutive law (Box 1), and myofiber orientations measured from published histology were adopted in the model (Peña et al., 2010). The authors concluded that fetal myocardial active tension increases significantly between 20 and 37 weeks of gestation, which was attributed to myocyte enlargement, differentiation and proliferation. A recent FEM modeling study of human fetal hearts between 22 and 32 weeks of gestation provided insights into the pathophysiology of fetal aortic stenosis with evolving HLHS (Ong et al., 2020). The study calculated the pressures, myocardial strain and stresses, and blood flow velocities at the heart valves at various degrees of stenosis severity, endocardial fibroelastosis, LV wall thickening and loss of contractility. The authors found that changes to myocardial stiffness due to fibroelastosis do not impede cardiac function. By using clinical measurements as comparisons, they further concluded that LV hypertrophy and reduced contractility are likely to manifest after aortic stenosis. FEM modeling has also demonstrated the ability to model the morphological growth of normal and HLHS fetal hearts. These studies accurately replicated the published measurements of left ventricular volumes at progressive gestational ages for both normal and diseased groups, reporting that reduced ventricular filling, chamber shape deformation and myocardial stresses influence the morphological alterations that result in HLHS (Dewan et al., 2017; Ohayon et al., 2002).

These computational modeling techniques have found much success when applied to investigating postnatal and adult-onset disease, designing treatments, and assisting clinical decision-making. For example, in patients with single-ventricle defects scheduled for Fontan surgery (Box 1), computational simulations of blood flow have been used to assess preoperative flow conditions and to predict the outcomes of various surgical options (Trusty et al., 2018). Fluid-structure interaction simulations have also been deployed to evaluate transcatheter aortic valve replacement, to identify potentials to optimize device design (Wu et al., 2016), whereas FEM simulations have been used to evaluate the effectiveness of novel procedures, such as leaflet laceration after valve-in-valve procedures (Yaakobovich et al., 2020). Application of the same techniques to fetal CVS and CHD research can have similar translational potential. For example, computational modeling to predict the outcomes of intrauterine surgical interventions to treat CHD may enhance patient selection and intervention planning, to minimize risks and to maximize the potential benefits.

Understanding epigenetic contributions to CHD

Despite the numerous CHD syndromes caused by single-gene variants (Table 1), specific causal mutations are recognized in only 20–30% of sporadic CHD (Mitchell et al., 2007; Pierpont et al., 2007), as revealed by whole-exome and -genome sequencing efforts (Homsy et al., 2015; Jin et al., 2017; Zaidi et al., 2013). Gene mutations account for only 35% of CHD. Importantly, many of these genes encode epigenetic pathway components, such as histone-modifying enzymes, leading to investigations of molecular epigenetics or pathways that regulate gene expression programs in CHD. Non-coding genomic regulatory loci, marked by epigenetic profiles and deregulated gene expression programs, are also likely to play important roles in the pathophysiology of CHD and its sequelae (Jarrell et al., 2019). Genetic variants at such non-coding regulatory sequences might influence the presentation of complex disorders by influencing chromatin accessibility, histone modifications, and downstream consequences of chromatin remodeling and gene expression (Table 3).

Table 3.

Epigenetic modifiers in heart development

Epigenetic modifiers in heart development
Epigenetic modifiers in heart development

Cardiac epigenomic maps have been generated for human and rodent hearts (Preissl et al., 2015; VanOudenhove et al., 2020), but only little is known about how specific histone modifications contribute to individual CHD phenotypes. Our understanding of epigenetic enhancer-promoter regulation of cardiogenesis has recently advanced with the development of genome-wide association studies (GWAS) and chromatin quantitative trait loci (chQTL; Box 1) analyses, elucidated by the epigenetic profiling of specific tissues, and by chromosome conformation capture (3C) and its derivatives (4C, Hi-C and HiChIP) that profile the spatial interaction of loci within and between chromosomes. Epigenetic factors contribute to phenotypic variation and are valuable for prioritizing variants for disease association (Kirk et al., 2006; Koopmann et al., 2014; Sati and Cavalli, 2017). For example, GWAS have associated non-coding, single nucleotide polymorphisms (SNPs) with various forms of CHD, although understanding the molecular functions of these intergenic SNPs remains quite challenging (Theis et al., 2015; Wijnands et al., 2017). One study has reported that a non-coding enhancer SNP disrupts binding of the transcription factor Tfap2a in the zebrafish heart, perturbing the expression of the atrial fibrillation-associated gene pitx2c through long-range interaction (Ye et al., 2016). Mutations in and the deletion of genes that encode histone modifying enzymes (e.g. EZH2, EP300) also produce deleterious effects on cardiovascular development in mice (Chen et al., 2009, 2012; He et al., 2012). However, to identify the disease regulatory variants with GWAS alone remains a challenge owing to the large cohort sizes needed to achieve statistical power, to disease heterogeneity and the need to assign cis-regulatory functions for implicated candidate loci (Ferrero, 2018). Other challenges in assessing epigenomic influences include modifier genes that are difficult to identify and characterize, candidate genes or sequence variants that are associated with a variety of heart malformations, and genes and variants that display incomplete penetrance, and may even be found in normal phenotypes (Farr et al., 2018).

A key aim of CHD research is to develop personalized therapies for prenatally diagnosed CHD, which can correct a known genetic mutation through gene-editing or gene addition, or that can correct abnormal biomechanics through minimally invasive surgical techniques. The fetal approach is ideal because it enables clinicians to arrest abnormal cardiac development before end-organ damage can occur, thereby producing better clinical outcomes (Kumar and Lodge, 2019).

A key challenge remains the further development of safe and effective therapies by using clinically relevant, large animal models. These efforts are supported by advances in gene modification technologies, which should enable large animal models to be genetically engineered such they carry common genetic mutations identified in CHD patients. We expect these technologies to be developed in tandem with molecular therapies to target the at-risk fetus in utero, aiming to correct cardiogenesis as early as possible in order to arrest pathology before it results in ventricular hypertrophy, hypotrophy or heart failure. Additionally, correcting CHD should be considered through ethical lenses, so that a potential novel treatment can be assessed not only on its ability to tackle a scientific challenge and rectify the disease, but also on its benefits to the patient and avoidance of harm.

Funding

This work was supported by the National Medical Research Council, Singapore (NMRC/CSA-INV/0012/2016 to C.N.Z.M.).

Abbott
,
M.
(
1936
).
Dr. Maude Abbott's Atlas of Congenital Cardiac Diseases
.
Abu-Farha
,
M.
,
Lambert
,
J. P.
,
Al-Madhoun
,
A. S.
,
Elisma
,
F.
,
Skerjanc
,
I. S.
and
Figeys
,
D.
(
2008
).
The tale of two domains: proteomics and genomics analysis of SMYD2, a new histone methyltransferase
.
Mol. Cell. Proteomics
7
,
560
-
572
.
Adams
,
M. E.
,
Hurd
,
E. A.
,
Beyer
,
L. A.
,
Swiderski
,
D. L.
,
Raphael
,
Y.
and
Martin
,
D. M.
(
2007
).
Defects in vestibular sensory epithelia and innervation in mice with loss of Chd7 function: implications for human CHARGE syndrome
.
J. Comp. Neurol.
504
,
519
-
532
.
Agopian
,
A. J.
,
Goldmuntz
,
E.
,
Hakonarson
,
H.
,
Sewda
,
A.
,
Taylor
,
D.
and
Mitchell
,
L. E.
(
2017
).
Genome-wide association studies and meta-analyses for congenital heart defects
.
Circ Cardiovasc Genet
10
,
e001449
.
Ahmad
,
F.
,
Banerjee
,
S. K.
,
Lage
,
M. L.
,
Huang
,
X. N.
,
Smith
,
S. H.
,
Saba
,
S.
,
Rager
,
J.
,
Conner
,
D. A.
,
Janczewski
,
A. M.
,
Tobita
,
K.
et al. 
(
2008
).
The role of cardiac troponin T quantity and function in cardiac development and dilated cardiomyopathy
.
PLoS ONE
3
,
e2642
.
Akerberg
,
A. A.
,
Burns
,
C. E.
,
Burns
,
C. G.
and
Nguyen
,
C.
(
2019
).
Deep learning enables automated volumetric assessments of cardiac function in zebrafish
.
Dis. Model. Mech.
12
.
Andersen
,
T. A.
,
Troelsen
,
K. d. L. L.
and
Larsen
,
L. A.
(
2014
).
Of mice and men: molecular genetics of congenital heart disease
.
Cell. Mol. Life Sci.
71
,
1327
-
1352
.
Andersen
,
P.
,
Tampakakis
,
E.
,
Jimenez
,
D. V.
,
Kannan
,
S.
,
Miyamoto
,
M.
,
Shin
,
H. K.
,
Saberi
,
A.
,
Murphy
,
S.
,
Sulistio
,
E.
,
Chelko
,
S. P.
et al. 
(
2018
).
Precardiac organoids form two heart fields via Bmp/Wnt signaling
.
Nat. Commun.
9
,
3140
.
Ang
,
Y. S.
,
Rivas
,
R. N.
,
Ribeiro
,
A. J. S.
,
Srivas
,
R.
,
Rivera
,
J.
,
Stone
,
N. R.
,
Pratt
,
K.
,
Mohamed
,
T. M. A.
,
Fu
,
J. D.
,
Spencer
,
C. I.
et al. 
(
2016
).
Disease model of GATA4 mutation reveals transcription factor cooperativity in human cardiogenesis
.
Cell
167
,
1734
-
1749.e22
.
Asim
,
A.
,
Agarwal
,
S.
,
Panigrahi
,
I.
,
Saiyed
,
N.
and
Bakshi
,
S.
(
2017
).
MTHFR promoter hypermethylation may lead to congenital heart defects in Down syndrome
.
Intractable Rare Dis. Res.
6
,
295
-
298
.
Ataliotis
,
P.
,
Ivins
,
S.
,
Mohun
,
T. J.
and
Scambler
,
P. J.
(
2005
).
XTbx1 is a transcriptional activator involved in head and pharyngeal arch development in Xenopus laevis
.
Dev. Dyn.
232
,
979
-
991
.
Azhar
,
M.
,
Wang
,
P. Y.
,
Frugier
,
T.
,
Koishi
,
K.
,
Deng
,
C.
,
Noakes
,
P. G.
and
McLennan
,
I. S.
(
2010
).
Myocardial deletion of Smad4 using a novel α skeletal muscle actin Cre recombinase transgenic mouse causes misalignment of the cardiac outflow tract
.
Int. J. Biol. Sci.
6
,
546
-
555
.
Bacha
,
E. A.
,
Cooper
,
D.
,
Thiagarajan
,
R.
,
Franklin
,
R. C.
,
Krogmann
,
O.
,
Deal
,
B.
,
Mavroudis
,
C.
,
Shukla
,
A.
,
Yeh
,
T.
,
Barach
,
P.
et al. 
(
2008
).
Cardiac complications associated with the treatment of patients with congenital cardiac disease: consensus definitions from the Multi-Societal Database Committee for Pediatric and Congenital Heart Disease
.
Cardiol Young 18 Suppl
2
,
196
-
201
.
Bakkers
,
J.
(
2011
).
Zebrafish as a model to study cardiac development and human cardiac disease
.
Cardiovasc. Res.
91
,
279
-
288
.
Baldini
,
A.
(
2004
).
DiGeorge syndrome: an update
.
Curr. Opin. Cardiol.
19
,
201
-
204
.
Barnes
,
R. M.
and
Black
,
B. L.
(
2016
).
Nodal signaling and congenital heart defects
. In
Etiology and Morphogenesis of Congenital Heart Disease: From Gene Function and Cellular Interaction to Morphology
(ed.
T.
Nakanishi
,
R. R.
Markwald
,
H. S.
Baldwin
,
B. B.
Keller
,
D.
Srivastava
and
H.
Yamagishi
), pp.
183
-
192
.
Tokyo
.
Barnett
,
J. V.
and
Desgrosellier
,
J. S.
(
2003
).
Early events in valvulogenesis: a signaling perspective
.
Birth Defects Res. C Embryo Today
69
,
58
-
72
.
Barron
,
D. J.
,
Kilby
,
M. D.
,
Davies
,
B.
,
Wright
,
J. G.
,
Jones
,
T. J.
and
Brawn
,
W. J.
(
2009
).
Hypoplastic left heart syndrome
.
Lancet
374
,
551
-
564
.
Bartlett
,
H. L.
,
Sutherland
,
L.
,
Kolker
,
S. J.
,
Welp
,
C.
,
Tajchman
,
U.
,
Desmarais
,
V.
and
Weeks
,
D. L.
(
2007
).
Transient early embryonic expression of Nkx2-5 mutations linked to congenital heart defects in human causes heart defects in Xenopus laevis
.
Dev. Dyn.
236
,
2475
-
2484
.
Bartman
,
T.
,
Walsh
,
E. C.
,
Wen
,
K.-K.
,
McKane
,
M.
,
Ren
,
J.
,
Alexander
,
J.
,
Rubenstein
,
P. A.
and
Stainier
,
D. Y.
(
2004
).
Early myocardial function affects endocardial cushion development in zebrafish
.
PLoS Biol.
2
,
e129
.
Basu
,
M.
and
Garg
,
V.
(
2018
).
Maternal hyperglycemia and fetal cardiac development: Clinical impact and underlying mechanisms
.
Birth Defects Res.
110
,
1504
-
1516
.
Beppu
,
H.
,
Malhotra
,
R.
,
Beppu
,
Y.
,
Lepore
,
J. J.
,
Parmacek
,
M. S.
and
Bloch
,
K. D.
(
2009
).
BMP type II receptor regulates positioning of outflow tract and remodeling of atrioventricular cushion during cardiogenesis
.
Dev. Biol.
331
,
167
-
175
.
Bier
,
E.
and
Bodmer
,
R.
(
2004
).
Drosophila, an emerging model for cardiac disease
.
Gene
342
,
1
-
11
.
Bilic
,
J.
and
Izpisua Belmonte
,
J. C.
(
2012
).
Concise review: Induced pluripotent stem cells versus embryonic stem cells: close enough or yet too far apart?
Stem Cells
30
,
33
-
41
.
Bluestein
,
D.
(
2017
).
Utilizing Computational Fluid Dynamics in cardiovascular engineering and medicine—What you need to know. Its translation to the clinic/bedside
.
Artif. Organs
41
,
117
.
Bokinni
,
Y.
(
2012
).
Kabuki syndrome revisited
.
J. Hum. Genet.
57
,
223
-
227
.
Bosman
,
E. A.
,
Penn
,
A. C.
,
Ambrose
,
J. C.
,
Kettleborough
,
R.
,
Stemple
,
D. L.
and
Steel
,
K. P.
(
2005
).
Multiple mutations in mouse Chd7 provide models for CHARGE syndrome
.
Hum. Mol. Genet.
14
,
3463
-
3476
.
Breckpot
,
J.
,
Thienpont
,
B.
,
Peeters
,
H.
,
de Ravel
,
T.
,
Singer
,
A.
,
Rayyan
,
M.
,
Allegaert
,
K.
,
Vanhole
,
C.
,
Eyskens
,
B.
,
Vermeesch
,
J. R.
et al. 
(
2010
).
Array comparative genomic hybridization as a diagnostic tool for syndromic heart defects
.
J. Pediatr.
156
,
810
-
817
,
817.e1
-
817.e4
.
Brown
,
C. B.
,
Boyer
,
A. S.
,
Runyan
,
R. B.
and
Barnett
,
J. V.
(
1999
).
Requirement of type III TGF-beta receptor for endocardial cell transformation in the heart
.
Science
283
,
2080
-
2082
.
Brown
,
C. B.
,
Wenning
,
J. M.
,
Lu
,
M. M.
,
Epstein
,
D. J.
,
Meyers
,
E. N.
and
Epstein
,
J. A.
(
2004
).
Cre-mediated excision of Fgf8 in the Tbx1 expression domain reveals a critical role for Fgf8 in cardiovascular development in the mouse
.
Dev. Biol.
267
,
190
-
202
.
Brown
,
D. R.
,
Samsa
,
L. A.
,
Qian
,
L.
and
Liu
,
J.
(
2016
).
Advances in the study of heart development and disease using Zebrafish
.
J. Cardiovas. Dev. Dis.
3
,
13
.
Bruneau
,
B. G.
,
Nemer
,
G.
,
Schmitt
,
J. P.
,
Charron
,
F.
,
Robitaille
,
L.
,
Caron
,
S.
,
Conner
,
D. A.
,
Gessler
,
M.
,
Nemer
,
M.
,
Seidman
,
C. E.
et al. 
(
2001
).
A murine model of Holt-Oram syndrome defines roles of the T-box transcription factor Tbx5 in cardiogenesis and disease
.
Cell
106
,
709
-
721
.
Bryantsev
,
A. L.
and
Cripps
,
R. M.
(
2009
).
Cardiac gene regulatory networks in Drosophila
.
Biochim. Biophys. Acta
1789
,
343
-
353
.
Burn
,
J.
,
Brennan
,
P.
,
Little
,
J.
,
Holloway
,
S.
,
Coffey
,
R.
,
Somerville
,
J.
,
Dennis
,
N. R.
,
Allan
,
L.
,
Arnold
,
R.
,
Deanfield
,
J. E.
et al. 
(
1998
).
Recurrence risks in offspring of adults with major heart defects: results from first cohort of British collaborative study
.
Lancet
351
,
311
-
316
.
Burridge
,
P. W.
,
Matsa
,
E.
,
Shukla
,
P.
,
Lin
,
Z. C.
,
Churko
,
J. M.
,
Ebert
,
A. D.
,
Lan
,
F.
,
Diecke
,
S.
,
Huber
,
B.
,
Mordwinkin
,
N. M.
et al. 
(
2014
).
Chemically defined generation of human cardiomyocytes
.
Nat. Methods
11
,
855
-
860
.
Burridge
,
P. W.
,
Sharma
,
A.
and
Wu
,
J. C.
(
2015
).
Genetic and epigenetic regulation of human cardiac reprogramming and differentiation in regenerative medicine
.
Annu. Rev. Genet.
49
,
461
-
484
.
Buyens
,
A.
,
Gyselaers
,
W.
,
Coumans
,
A.
,
Al Nasiry
,
S.
,
Willekes
,
C.
,
Boshoff
,
D.
,
Frijns
,
J.-P.
and
Witters
,
I.
(
2012
).
Difficult prenatal diagnosis: fetal coarctation
.
Facts Views Vis. ObGyn
4
,
230
.
Cai
,
C. L.
,
Zhou
,
W.
,
Yang
,
L.
,
Bu
,
L.
,
Qyang
,
Y.
,
Zhang
,
X.
,
Li
,
X.
,
Rosenfeld
,
M. G.
,
Chen
,
J.
and
Evans
,
S.
(
2005
).
T-box genes coordinate regional rates of proliferation and regional specification during cardiogenesis
.
Development
132
,
2475
-
2487
.
Camacho
,
P.
,
Fan
,
H.
,
Liu
,
Z.
and
He
,
J. Q.
(
2016
).
Large mammalian animal models of heart disease
.
J. Cardiovasc. Dev. Dis.
3
,
30
.
Camenisch
,
T. D.
,
Spicer
,
A. P.
,
Brehm-Gibson
,
T.
,
Biesterfeldt
,
J.
,
Augustine
,
M. L.
,
Calabro
,
A.
, Jr.
,
Kubalak
,
S.
,
Klewer
,
S. E.
and
McDonald
,
J. A.
(
2000
).
Disruption of hyaluronan synthase-2 abrogates normal cardiac morphogenesis and hyaluronan-mediated transformation of epithelium to mesenchyme
.
J. Clin. Invest.
106
,
349
-
360
.
Chen
,
B.
,
Bronson
,
R. T.
,
Klaman
,
L. D.
,
Hampton
,
T. G.
,
Wang
,
J. F.
,
Green
,
P. J.
,
Magnuson
,
T.
,
Douglas
,
P. S.
,
Morgan
,
J. P.
and
Neel
,
B. G.
(
2000
).
Mice mutant for Egfr and Shp2 have defective cardiac semilunar valvulogenesis
.
Nat. Genet.
24
,
296
-
299
.
Chen
,
G.
,
Zhu
,
J.
,
Lv
,
T.
,
Wu
,
G.
,
Sun
,
H.
,
Huang
,
X.
and
Tian
,
J.
(
2009
).
Spatiotemporal expression of histone acetyltransferases, p300 and CBP, in developing embryonic hearts
.
J. Biomed. Sci.
16
,
24
.
Chen
,
L.
,
Ma
,
Y.
,
Kim
,
E. Y.
,
Yu
,
W.
,
Schwartz
,
R. J.
,
Qian
,
L.
and
Wang
,
J.
(
2012
).
Conditional ablation of Ezh2 in murine hearts reveals its essential roles in endocardial cushion formation, cardiomyocyte proliferation and survival
.
PLoS ONE
7
,
e31005
.
Chen
,
Z.
,
Zhou
,
Y.
,
Wang
,
J.
,
Liu
,
X.
,
Ge
,
S.
and
He
,
Y.
(
2017
).
Modeling of coarctation of aorta in human fetuses using 3D/4D fetal echocardiography and computational fluid dynamics
.
Echocardiography
34
,
1858
-
1866
.
Clark
,
K. L.
,
Yutzey
,
K. E.
and
Benson
,
D. W.
(
2006
).
Transcription factors and congenital heart defects
.
Annu. Rev. Physiol.
68
,
97
-
121
.
Cleaver
,
O. B.
,
Patterson
,
K. D.
and
Krieg
,
P. A.
(
1996
).
Overexpression of the tinman-related genes XNkx-2.5 and XNkx-2.3 in Xenopus embryos results in myocardial hyperplasia
.
Development
122
,
3549
-
3556
.
Conway
,
S. J.
,
Henderson
,
D. J.
,
Kirby
,
M. L.
,
Anderson
,
R. H.
and
Copp
,
A. J.
(
1997
).
Development of a lethal congenital heart defect in the splotch (Pax3) mutant mouse
.
Cardiovasc. Res.
36
,
163
-
173
.
Correia
,
C.
,
Koshkin
,
A.
,
Duarte
,
P.
,
Hu
,
D.
,
Carido
,
M.
,
Sebastiao
,
M. J.
,
Gomes-Alves
,
P.
,
Elliott
,
D. A.
,
Domian
,
I. J.
,
Teixeira
,
A. P.
et al. 
(
2018
).
3D aggregate culture improves metabolic maturation of human pluripotent stem cell derived cardiomyocytes
.
Biotechnol. Bioeng.
115
,
630
-
644
.
Costopoulos
,
C.
,
Huang
,
Y.
,
Brown
,
A. J.
,
Calvert
,
P. A.
,
Hoole
,
S. P.
,
West
,
N. E.
,
Gillard
,
J. H.
,
Teng
,
Z.
and
Bennett
,
M. R.
(
2017
).
Plaque rupture in coronary atherosclerosis is associated with increased plaque structural stress
.
JACC: Cardiovascular Imaging
10
,
1472
-
1483
.
Courchaine
,
K.
and
Rugonyi
,
S.
(
2018
).
Quantifying blood flow dynamics during cardiac development: demystifying computational methods
.
Philos. Trans. R. Soc. Lond. B Biol. Sci.
373
.
Cowan
,
J. R.
and
Ware
,
S. M.
(
2015
).
Genetics and genetic testing in congenital heart disease
.
Clin. Perinatol.
42
,
373
-
393
.
Dai
,
F.
,
Yusuf
,
F.
,
Farjah
,
G. H.
and
Brand-Saberi
,
B.
(
2005
).
RNAi-induced targeted silencing of developmental control genes during chicken embryogenesis
.
Dev. Biol.
285
,
80
-
90
.
deAlmeida
,
A.
,
McQuinn
,
T.
and
Sedmera
,
D.
(
2007
).
Increased ventricular preload is compensated by myocyte proliferation in normal and hypoplastic fetal chick left ventricle
.
Circ. Res.
100
,
1363
-
1370
.
de Soysa
,
T. Y.
,
Ranade
,
S. S.
,
Okawa
,
S.
,
Ravichandran
,
S.
,
Huang
,
Y.
,
Salunga
,
H. T.
,
Schricker
,
A.
,
Del Sol
,
A.
,
Gifford
,
C. A.
and
Srivastava
,
D.
(
2019
).
Single-cell analysis of cardiogenesis reveals basis for organ-level developmental defects
.
Nature
572
,
120
-
124
.
DeLaughter
,
D. M.
,
Saint-Jean
,
L.
,
Baldwin
,
H. S.
and
Barnett
,
J. V.
(
2011
).
What chick and mouse models have taught us about the role of the endocardium in congenital heart disease
.
Birth Defects Res. A Clin. Mol. Teratol.
91
,
511
-
525
.
Delgado-Olguín
,
P.
,
Huang
,
Y.
,
Li
,
X.
,
Christodoulou
,
D.
,
Seidman
,
C. E.
,
Seidman
,
J. G.
,
Tarakhovsky
,
A.
and
Bruneau
,
B. G.
(
2012
).
Epigenetic repression of cardiac progenitor gene expression by Ezh2 is required for postnatal cardiac homeostasis
.
Nat. Genet.
44
,
343
-
347
.
Del Monte-Nieto
,
G.
,
Ramialison
,
M.
,
Adam
,
A. A. S.
,
Wu
,
B.
,
Aharonov
,
A.
,
D'Uva
,
G.
,
Bourke
,
L. M.
,
Pitulescu
,
M. E.
,
Chen
,
H.
,
de la Pompa
,
J. L.
et al. 
(
2018
).
Control of cardiac jelly dynamics by NOTCH1 and NRG1 defines the building plan for trabeculation
.
Nature
557
,
439
-
445
.
Desgrange
,
A.
,
Le Garrec
,
J. F.
and
Meilhac
,
S. M.
(
2018
).
Left-right asymmetry in heart development and disease: forming the right loop
.
Development
145
.
Dewan
,
S.
,
Krishnamurthy
,
A.
,
Kole
,
D.
,
Conca
,
G.
,
Kerckhoffs
,
R.
,
Puchalski
,
M. D.
,
Omens
,
J. H.
,
Sun
,
H.
,
Nigam
,
V.
and
McCulloch
,
A. D.
(
2017
).
Model of human fetal growth in hypoplastic left heart syndrome: reduced ventricular growth due to decreased ventricular filling and altered shape
.
Front. Pediatr.
5
,
25
.
Dobosz
,
A.
,
Grabowska
,
A.
and
Bik-Multanowski
,
M.
(
2019
).
Hypermethylation of NRG1 gene correlates with the presence of heart defects in Down's syndrome
.
J. Genet.
98
.
Donovan
,
J.
,
Kordylewska
,
A.
,
Jan
,
Y. N.
and
Utset
,
M. F.
(
2002
).
Tetralogy of fallot and other congenital heart defects in Hey2 mutant mice
.
Curr. Biol.
12
,
1605
-
1610
.
Doss
,
M. X.
and
Sachinidis
,
A.
(
2019
).
Current challenges of iPSC-based disease modeling and therapeutic implications
.
Cells
8
.
Duchemin
,
A.-L.
,
Vignes
,
H.
and
Vermot
,
J.
(
2019
).
Mechanically activated piezo channels modulate outflow tract valve development through the Yap1 and Klf2-Notch signaling axis
.
Elife
8
,
e44706
.
Duffy
,
J. B.
(
2002
).
GAL4 system in Drosophila: a fly geneticist's Swiss army knife
.
Genesis
34
,
1
-
15
.
Ebert
,
A. D.
,
Liang
,
P.
and
Wu
,
J. C.
(
2012
).
Induced pluripotent stem cells as a disease modeling and drug screening platform
.
J. Cardiovasc. Pharmacol.
60
,
408
-
416
.
Eghtesady
,
P.
,
Michelfelder
,
E.
,
Altaye
,
M.
,
Ballard
,
E.
,
Hirsh
,
R.
and
Beekman
,
R. H.
III
(
2007
).
Revisiting animal models of aortic stenosis in the early gestation fetus
.
Ann. Thorac. Surg.
83
,
631
-
639
.
Eisen
,
J. S.
and
Smith
,
J. C.
(
2008
).
Controlling morpholino experiments: don't stop making antisense
.
Development
135
,
1735
-
1743
.
England
,
J.
,
Pang
,
K. L.
,
Parnall
,
M.
,
Haig
,
M. I.
and
Loughna
,
S.
(
2016
).
Cardiac troponin T is necessary for normal development in the embryonic chick heart
.
J. Anat.
229
,
436
-
449
.
Epstein
,
D. J.
,
Vekemans
,
M.
and
Gros
P.
(
1991
).
Splotch (Sp2H), a mutation affecting development of the mouse neural tube, shows a deletion within the paired homeodomain of Pax-3. Cell. 67, 767-774.
Epstein
,
J. A.
,
Li
,
J.
,
Lang
,
D.
,
Chen
,
F.
,
Brown
,
C. B.
,
Jin
,
F.
,
Lu
,
M. M.
,
Thomas
,
M.
,
Liu
,
E.
,
Wessels
,
A.
et al. 
(
2000
).
Migration of cardiac neural crest cells in Splotch embryos
.
Development
127
,
1869
-
1878
.
Fahed
,
A. C.
,
Gelb
,
B. D.
,
Seidman
,
J. G.
and
Seidman
,
C. E.
(
2013
).
Genetics of congenital heart disease: the glass half empty
.
Circ. Res.
112
,
707
-
720
.
Familiari
,
A.
,
Morlando
,
M.
,
Khalil
,
A.
,
Sonesson
,
S.-E.
,
Scala
,
C.
,
Rizzo
,
G.
,
Del Sordo
,
G.
,
Vassallo
,
C.
,
Elena Flacco
,
M.
and
Manzoli
,
L.
(
2017
).
Risk factors for coarctation of the aorta on prenatal ultrasound: a systematic review and meta-analysis
.
Circulation
135
,
772
-
785
.
Farr
,
G. H.
, III
,
Imani
,
K.
,
Pouv
,
D.
and
Maves
,
L.
(
2018
).
Functional testing of a human PBX3 variant in zebrafish reveals a potential modifier role in congenital heart defects
.
Dis. Model. Mech.
11
,
dmm035972
.
Faucherre
,
A.
,
ou Maati
,
H. M.
,
Nasr
,
N.
,
Pinard
,
A.
,
Theron
,
A.
,
Odelin
,
G.
,
Desvignes
,
J.-P.
,
Salgado
,
D.
,
Collod-Béroud
,
G.
and
Avierinos
,
J.-F.
(
2020
).
Piezo1 and outflow tract and aortic valve development
.
J. Mol. Cell. Cardiol.
143
,
51
-
62
.
Feng
,
Q.
,
Song
,
W.
,
Lu
,
X.
,
Hamilton
,
J. A.
,
Lei
,
M.
,
Peng
,
T.
and
Yee
,
S. P.
(
2002
).
Development of heart failure and congenital septal defects in mice lacking endothelial nitric oxide synthase
.
Circulation
106
,
873
-
879
.
Ferrero
,
E.
(
2018
).
Using regulatory genomics data to interpret the function of disease variants and prioritise genes from expression studies
.
F1000Res
7
,
121
.
Firulli
,
B. A.
,
George
,
R. M.
,
Harkin
,
J.
,
Toolan
,
K. P.
,
Gao
,
H.
,
Liu
,
Y.
,
Zhang
,
W.
,
Field
,
L. J.
,
Liu
,
Y.
,
Shou
,
W.
et al. 
(
2020
).
HAND1 loss-of-function within the embryonic myocardium reveals survivable congenital cardiac defects and adult heart failure
.
Cardiovasc. Res.
116
,
605
-
618
.
Fishman
,
N. H.
,
Hof
,
R. B.
,
Rudolph
,
A. M.
and
Heymann
,
M. A.
(
1978
).
Models of congenital heart disease in fetal lambs
.
Circulation
58
,
354
-
364
.
Franklin
,
O.
,
Burch
,
M.
,
Manning
,
N.
,
Sleeman
,
K.
,
Gould
,
S.
and
Archer
,
N.
(
2002
).
Prenatal diagnosis of coarctation of the aorta improves survival and reduces morbidity
.
Heart
87
,
67
-
69
.
Freud
,
L. R.
,
McElhinney
,
D. B.
,
Marshall
,
A. C.
,
Marx
,
G. R.
,
Friedman
,
K. G.
,
del Nido
,
P. J.
,
Emani
,
S. M.
,
Lafranchi
,
T.
,
Silva
,
V.
and
Wilkins-Haug
,
L. E.
(
2014
).
Fetal aortic valvuloplasty for evolving hypoplastic left heart syndrome: postnatal outcomes of the first 100 patients
.
Circulation
130
,
638
-
645
.
Friedman
,
K. G.
,
Sleeper
,
L. A.
,
Freud
,
L. R.
,
Marshall
,
A. C.
,
Godfrey
,
M. E.
,
Drogosz
,
M.
,
Lafranchi
,
T.
,
Benson
,
C. B.
,
Wilkins-Haug
,
L. E.
and
Tworetzky
,
W.
(
2018
).
Improved technical success, postnatal outcome and refined predictors of outcome for fetal aortic valvuloplasty
.
Ultrasound Obstet. Gynecol.
52
,
212
-
220
.
Fu
,
Y.
,
Yan
,
W.
,
Mohun
,
T. J.
and
Evans
,
S. M.
(
1998
).
Vertebrate tinman homologues XNkx2-3 and XNkx2-5 are required for heart formation in a functionally redundant manner
.
Development
125
,
4439
-
4449
.
Fukiishi
,
Y.
and
Morriss-Kay
,
G. M.
(
1992
).
Migration of cranial neural crest cells to the pharyngeal arches and heart in rat embryos
.
Cell Tissue Res.
268
,
1
-
8
.
Furtado
,
M. B.
,
Wilmanns
,
J. C.
,
Chandran
,
A.
,
Perera
,
J.
,
Hon
,
O.
,
Biben
,
C.
,
Willow
,
T. J.
,
Nim
,
H. T.
,
Kaur
,
G.
,
Simonds
,
S.
et al. 
(
2017
).
Point mutations in murine Nkx2-5 phenocopy human congenital heart disease and induce pathogenic Wnt signaling
.
JCI Insight
2
,
e88271
.
Gage
,
P. J.
,
Suh
,
H.
and
Camper
,
S. A.
(
1999
).
Dosage requirement of Pitx2 for development of multiple organs
.
Development
126
,
4643
-
4651
.
Gammill
,
L. S.
and
Krull
,
C. E.
(
2011
).
Embryological and genetic manipulation of chick development
. In
Vertebrate Embryogenesis: Embryological, Cellular, and Genetic Methods
(ed.
F. J.
Pelegri
), Vol. 770, pp.
119
-
137
.
Totowa, NJ
:
Humana Press
.
Gao
,
S. J.
,
Zhang
,
G. F.
and
Zhang
,
R. P.
(
2016
).
High CpG island methylation of p16 gene and loss of p16 protein expression associate with the development and progression of tetralogy of Fallot
.
J. Genet.
95
,
831
-
837
.
Gao
,
X.
,
Yang
,
L.
,
Luo
,
H.
,
Tan
,
F.
,
Ma
,
X.
and
Lu
,
C.
(
2018
).
A rare Rs139365823 polymorphism in Pre-miR-138 is associated with risk of congenital heart disease in a chinese population
.
DNA Cell Biol.
37
,
109
-
116
.
Garcia-Canadilla
,
P.
,
Rudenick
,
P. A.
,
Crispi
,
F.
,
Cruz-Lemini
,
M.
,
Palau
,
G.
,
Camara
,
O.
,
Gratacos
,
E.
and
Bijens
,
B. H.
(
2014
).
A computational model of the fetal circulation to quantify blood redistribution in intrauterine growth restriction
.
PLoS Comput. Biol.
10
,
e1003667
.
Gardiner
,
H. M.
(
2019
).
In utero intervention for severe congenital heart disease
.
Best Pract. Res. Clin. Obstet. Gynaecol.
58
,
42
-
54
.
Garg
,
V.
,
Kathiriya
,
I. S.
,
Barnes
,
R.
,
Schluterman
,
M. K.
,
King
,
I. N.
,
Butler
,
C. A.
,
Rothrock
,
C. R.
,
Eapen
,
R. S.
,
Hirayama-Yamada
,
K.
,
Joo
,
K.
et al. 
(
2003
).
GATA4 mutations cause human congenital heart defects and reveal an interaction with TBX5
.
Nature
424
,
443
-
447
.
Garrity
,
D. M.
,
Childs
,
S.
and
Fishman
,
M. C.
(
2002
).
The heartstrings mutation in zebrafish causes heart/fin Tbx5 deficiency syndrome
.
Development
129
,
4635
-
4645
.
GB_2017_Congenital_Heart_Disease_Collaborators
(
2020
).
Global, regional, and national burden of congenital heart disease, 1990-2017: a systematic analysis for the Global Burden of Disease Study 2017
.
Lancet Child Adolesc Health
4
,
185
-
200
.
Genge
,
C. E.
,
Lin
,
E.
,
Lee
,
L.
,
Sheng
,
X.
,
Rayani
,
K.
,
Gunawan
,
M.
,
Stevens
,
C. M.
,
Li
,
A. Y.
,
Talab
,
S. S.
,
Claydon
,
T. W.
et al. 
(
2016
).
The Zebrafish Heart as a Model of Mammalian Cardiac Function
.
Rev. Physiol. Biochem. Pharmacol.
171
,
99
-
136
.
Gessert
,
S.
and
Kuhl
,
M.
(
2009
).
Comparative gene expression analysis and fate mapping studies suggest an early segregation of cardiogenic lineages in Xenopus laevis
.
Dev. Biol.
334
,
395
-
408
.
Gherghiceanu
,
M.
,
Barad
,
L.
,
Novak
,
A.
,
Reiter
,
I.
,
Itskovitz-Eldor
,
J.
,
Binah
,
O.
and
Popescu
,
L. M.
(
2011
).
Cardiomyocytes derived from human embryonic and induced pluripotent stem cells: comparative ultrastructure
.
J. Cell. Mol. Med.
15
,
2539
-
2551
.
Glaser
,
S.
,
Schaft
,
J.
,
Lubitz
,
S.
,
Vintersten
,
K.
,
van der Hoeven
,
F.
,
Tufteland
,
K. R.
,
Aasland
,
R.
,
Anastassiadis
,
K.
,
Ang
,
S. L.
and
Stewart
,
A. F.
(
2006
).
Multiple epigenetic maintenance factors implicated by the loss of Mll2 in mouse development
.
Development
133
,
1423
-
1432
.
Glaser
,
S.
,
Lubitz
,
S.
,
Loveland
,
K. L.
,
Ohbo
,
K.
,
Robb
,
L.
,
Schwenk
,
F.
,
Seibler
,
J.
,
Roellig
,
D.
,
Kranz
,
A.
,
Anastassiadis
,
K.
et al. 
(
2009
).
The histone 3 lysine 4 methyltransferase, Mll2, is only required briefly in development and spermatogenesis
.
Epigenetics Chromatin
2
,
5
.
Goldmuntz
,
E.
,
Paluru
,
P.
,
Glessner
,
J.
,
Hakonarson
,
H.
,
Biegel
,
J. A.
,
White
,
P. S.
,
Gai
,
X.
and
Shaikh
,
T. H.
(
2011
).
Microdeletions and microduplications in patients with congenital heart disease and multiple congenital anomalies
.
Congenit Heart Dis.
6
,
592
-
602
.
Gong
,
J.
,
Sheng
,
W.
,
Ma
,
D.
,
Huang
,
G.
and
Liu
,
F.
(
2019
).
DNA methylation status of TBX20 in patients with tetralogy of Fallot
.
BMC Med Genomics
12
,
75
.
Goodrich
,
L. V.
,
Milenkovic
,
L.
,
Higgins
,
K. M.
and
Scott
,
M. P.
(
1997
).
Altered neural cell fates and medulloblastoma in mouse patched mutants. Science 277, 1109 -1113.
Gordon
,
E. E.
,
Reinking
,
B. E.
,
Hu
,
S.
,
Yao
,
J.
,
Kua
,
K. L.
,
Younes
,
A. K.
,
Wang
,
C.
,
Segar
,
J. L.
and
Norris
,
A. W.
(
2015
).
Maternal Hyperglycemia Directly and Rapidly Induces Cardiac Septal Overgrowth in Fetal Rats
.
J. Diabetes Res.
2015
,
479565
.
Groenendijk
,
B. C.
,
Hierck
,
B. P.
,
Vrolijk
,
J.
,
Baiker
,
M.
,
Pourquie
,
M. J.
,
Gittenberger-de Groot
,
A. C.
and
Poelmann
,
R. E.
(
2005
).
Changes in shear stress–related gene expression after experimentally altered venous return in the chicken embryo
.
Circ. Res.
96
,
1291
-
1298
.
Groenendijk
,
B. C.
,
Van der Heiden
,
K.
,
Hierck
,
B. P.
and
Poelmann
,
R. E.
(
2007
).
The role of shear stress on ET-1, KLF2, and NOS-3 expression in the developing cardiovascular system of chicken embryos in a venous ligation model
.
Physiology (Bethesda)
22
,
380
-
389
.
Grunert
,
M.
,
Dorn
,
C.
,
Cui
,
H.
,
Dunkel
,
I.
,
Schulz
,
K.
,
Schoenhals
,
S.
,
Sun
,
W.
,
Berger
,
F.
,
Chen
,
W.
and
Sperling
,
S. R.
(
2016
).
Comparative DNA methylation and gene expression analysis identifies novel genes for structural congenital heart diseases
.
Cardiovasc. Res.
112
,
464
-
477
.
Grunert
,
M.
,
Appelt
,
S.
,
Schonhals
,
S.
,
Mika
,
K.
,
Cui
,
H.
,
Cooper
,
A.
,
Cyganek
,
L.
,
Guan
,
K.
and
Sperling
,
S. R.
(
2020
).
Induced pluripotent stem cells of patients with Tetralogy of Fallot reveal transcriptional alterations in cardiomyocyte differentiation
.
Sci. Rep.
10
,
10921
.
Guichard
,
A.
,
Srinivasan
,
S.
,
Zimm
,
G.
and
Bier
,
E.
(
2002
).
A screen for dominant mutations applied to components in the Drosophila EGF-R pathway
.
Proc. Natl. Acad. Sci. U.S.A.
99
,
3752
-
3757
.
Guo
,
T.
,
Chung
,
J. H.
,
Wang
,
T.
,
McDonald-McGinn
,
D. M.
,
Kates
,
W. R.
,
Hawuła
,
W.
,
Coleman
,
K.
,
Zackai
,
E.
,
Emanuel
,
B. S.
and
Morrow
,
B. E.
(
2015
).
Histone modifier genes alter conotruncal heart phenotypes in 22q11.2 deletion syndrome
.
Am. J. Hum. Genet.
97
,
869
-
877
.
Gut
,
P.
,
Reischauer
,
S.
,
Stainier
,
D. Y. R.
and
Arnaout
,
R.
(
2017
).
Little fish, big data: zebrafish as a model for cardiovascular and metabolic disease
.
Physiol. Rev.
97
,
889
-
938
.
Hamamoto
,
R.
,
Furukawa
,
Y.
,
Morita
,
M.
,
Iimura
,
Y.
,
Silva
,
F. P.
,
Li
,
M.
,
Yagyu
,
R.
and
Nakamura
,
Y.
(
2004
).
SMYD3 encodes a histone methyltransferase involved in the proliferation of cancer cells
.
Nat. Cell Biol.
6
,
731
-
740
.
Han
,
X.
,
Zhang
,
J.
,
Liu
,
Y.
,
Fan
,
X.
,
Ai
,
S.
,
Luo
,
Y.
,
Li
,
X.
,
Jin
,
H.
,
Luo
,
S.
,
Zheng
,
H.
et al. 
(
2019
).
The lncRNA Hand2os1/Uph locus orchestrates heart development through regulation of precise expression of Hand2
.
Development
146
.
Han
,
S.
,
Zhang
,
Y. Y.
,
Meng
,
M. Y.
,
Hou
,
Z. L.
,
Meng
,
P.
,
Zhao
,
Y. Y.
,
Gao
,
H.
,
Tang
,
J.
,
Liu
,
Z.
,
Yang
,
L. L.
et al. 
(
2020
).
Generation of human iPSC line from a patient with Tetralogy of Fallot, YAHKMUi001-A, carrying a mutation in TBX1 gene
.
Stem Cell Res.
42
,
101687
.
Hasten
,
E.
,
McDonald-McGinn
,
D. M.
,
Crowley
,
T. B.
,
Zackai
,
E.
,
Emanuel
,
B. S.
,
Morrow
,
B. E.
and
Racedo
,
S. E.
(
2018
).
Dysregulation of TBX1 dosage in the anterior heart field results in congenital heart disease resembling the 22q11.2 duplication syndrome
.
Hum. Mol. Genet.
27
,
1847
-
1857
.
Hatano
,
S.
,
Kimata
,
K.
,
Hiraiwa
,
N.
,
Kusakabe
,
M.
,
Isogai
,
Z.
,
Adachi
,
E.
,
Shinomura
,
T.
and
Watanabe
,
H.
(
2012
).
Versican/PG-M is essential for ventricular septal formation subsequent to cardiac atrioventricular cushion development
.
Glycobiology
22
,
1268
-
1277
.
Haworth
,
K. E.
,
Kotecha
,
S.
,
Mohun
,
T. J.
and
Latinkic
,
B. V.
(
2008
).
GATA4 and GATA5 are essential for heart and liver development in Xenopus embryos
.
BMC Dev. Biol.
8
,
74
.
He
,
A.
,
Ma
,
Q.
,
Cao
,
J.
,
von Gise
,
A.
,
Zhou
,
P.
,
Xie
,
H.
,
Zhang
,
B.
,
Hsing
,
M.
,
Christodoulou
,
D. C.
,
Cahan
,
P.
et al. 
(
2012
).
Polycomb repressive complex 2 regulates normal development of the mouse heart
.
Circ. Res.
110
,
406
-
415
.
Hinton
,
R. B.
, Jr.
,
Martin
,
L. J.
,
Tabangin
,
M. E.
,
Mazwi
,
M. L.
,
Cripe
,
L. H.
and
Benson
,
D. W.
(
2007
).
Hypoplastic left heart syndrome is heritable
.
J. Am. Coll. Cardiol.
50
,
1590
-
1595
.
Hirschhorn
,
K.
,
Cooper
,
H. L.
and
Firschein
,
I. L.
(
1965
).
Deletion of short arms of chromosome 4-5 in a child with defects of midline fusion
.
Humangenetik
1
,
479
-
482
.
Hirt
,
M. N.
,
Boeddinghaus
,
J.
,
Mitchell
,
A.
,
Schaaf
,
S.
,
Bornchen
,
C.
,
Muller
,
C.
,
Schulz
,
H.
,
Hubner
,
N.
,
Stenzig
,
J.
,
Stoehr
,
A.
et al. 
(
2014
).
Functional improvement and maturation of rat and human engineered heart tissue by chronic electrical stimulation
.
J. Mol. Cell. Cardiol.
74
,
151
-
161
.
Hoffman
,
J. I.
(
1995
).
Incidence of congenital heart disease: I. Postnatal incidence
.
Pediatr. Cardiol.
16
,
103
-
113
.
Hoffman
,
J. I.
and
Kaplan
,
S.
(
2002
).
The incidence of congenital heart disease
.
J. Am. Coll. Cardiol.
39
,
1890
-
1900
.
Hogers
,
B.
,
DeRuiter
,
M.
,
Gittenberger-de Groot
,
A.
and
Poelmann
,
R.
(
1997
).
Unilateral vitelline vein ligation alters intracardiac blood flow patterns and morphogenesis in the chick embryo
.
Circ. Res.
80
,
473
-
481
.
Homsy
,
J.
,
Zaidi
,
S.
,
Shen
,
Y.
,
Ware
,
J. S.
,
Samocha
,
K. E.
,
Karczewski
,
K. J.
,
DePalma
,
S. R.
,
McKean
,
D.
,
Wakimoto
,
H.
,
Gorham
,
J.
et al. 
(
2015
).
De novo mutations in congenital heart disease with neurodevelopmental and other congenital anomalies
.
Science
350
,
1262
-
1266
.
Horb
,
M. E.
and
Thomsen
,
G. H.
(
1999
).
Tbx5 is essential for heart development
.
Development
126
,
1739
-
1751
.
Hou
,
Y. S.
,
Wang
,
J. Z.
,
Shi
,
S.
,
Han
,
Y.
,
Zhang
,
Y.
,
Zhi
,
J. X.
,
Xu
,
C.
,
Li
,
F. F.
,
Wang
,
G. Y.
and
Liu
,
S. L.
(
2020
).
Identification of epigenetic factor KAT2B gene variants for possible roles in congenital heart diseases
.
Biosci. Rep.
40
.
Hove
,
J. R.
,
Koster
,
R. W.
,
Forouhar
,
A. S.
,
Acevedo-Bolton
,
G.
,
Fraser
,
S. E.
and
Gharib
,
M.
(
2003
).
Intracardiac fluid forces are an essential epigenetic factor for embryonic cardiogenesis
.
Nature
421
,
172
-
177
.
Hu
,
N.
,
Christensen
,
D. A.
,
Agrawal
,
A. K.
,
Beaumont
,
C.
,
Clark
,
E. B.
and
Hawkins
,
J. A.
(
2009
).
Dependence of aortic arch morphogenesis on intracardiac blood flow in the left atrial ligated chick embryo
.
Anat. Rec. (Hoboken)
292
,
652
-
660
.
Huang
,
G. Y.
,
Cooper
,
E. S.
,
Waldo
,
K.
,
Kirby
,
M. L.
,
Gilula
,
N. B.
and
Lo
,
C. W.
(
1998
).
Gap junction-mediated cell-cell communication modulates mouse neural crest migration
.
J. Cell Biol.
143
,
1725
-
1734
.
Huang
,
J.
,
Li
,
X.
,
Li
,
H.
,
Su
,
Z.
,
Wang
,
J.
and
Zhang
,
H.
(
2015
).
Down-regulation of microRNA-184 contributes to the development of cyanotic congenital heart diseases
.
Int. J. Clin. Exp. Pathol.
8
,
14221
-
14227
.
Hughes
,
C. J. R.
and
Jacobs
,
J. R.
(
2017
).
Dissecting the role of the extracellular matrix in heart disease: lessons from the drosophila genetic model
.
Vet Sci
4
,
24
.
Hutson
,
M. R.
and
Kirby
,
M. L.
(
2003
).
Neural crest and cardiovascular development: a 20-year perspective
.
Birth Defects Res. C Embryo Today
69
,
2
-
13
.
Hutson
,
M. R.
and
Kirby
,
M. L.
(
2007
).
Model systems for the study of heart development and disease. Cardiac neural crest and conotruncal malformations
.
Semin. Cell Dev. Biol.
18
,
101
-
110
.
Iascone
,
M.
,
Ciccone
,
R.
,
Galletti
,
L.
,
Marchetti
,
D.
,
Seddio
,
F.
,
Lincesso
,
A. R.
,
Pezzoli
,
L.
,
Vetro
,
A.
,
Barachetti
,
D.
,
Boni
,
L.
et al. 
(
2012
).
Identification of de novo mutations and rare variants in hypoplastic left heart syndrome
.
Clin. Genet.
81
,
542
-
554
.
Inácio
,
J. M.
,
Almeida
,
M.
,
Cristo
,
F.
and
Belo
,
J. A.
(
2020
).
Generation of a gene-corrected human induced pluripotent stem cell line derived from a patient with laterality defects and congenital heart anomalies with a c.455G>A alteration in DAND5
.
Stem Cell Res
42
,
101677
.
Ishii
,
Y.
,
Garriock
,
R. J.
,
Navetta
,
A. M.
,
Coughlin
,
L. E.
and
Mikawa
,
T.
(
2010
).
BMP signals promote proepicardial protrusion necessary for recruitment of coronary vessel and epicardial progenitors to the heart
.
Dev. Cell
19
,
307
-
316
.
Israeli
,
Y.
,
Gabalski
,
M.
,
Ball
,
K.
,
Wasserman
,
A.
,
Zou
,
J.
,
Ni
,
G.
,
Zhou
,
C.
and
Aguirre
,
A.
(
2020
).
Generation of heart organoids modeling early human cardiac development under defined conditions
.
bioRxiv
2020.06.25.171611
.
Ivanovitch
,
K.
,
Temiño
,
S.
and
Torres
,
M.
(
2017
).
Live imaging of heart tube development in mouse reveals alternating phases of cardiac differentiation and morphogenesis
.
Elife
6
,
e30668
.
Jarrell
,
D. K.
,
Lennon
,
M. L.
and
Jacot
,
J. G.
(
2019
).
Epigenetics and mechanobiology in heart development and congenital heart disease
.
Diseases (Basel, Switzerland)
7
,
52
.
Jenkins
,
M. W.
,
Watanabe
,
M.
and
Rollins
,
A. M.
(
2012
).
Longitudinal imaging of heart development with optical coherence tomography
.
IEEE J. Sel. Top. Quantum Electron.
18
,
1166
-
1175
.
Jiang
,
Y.
,
Habibollah
,
S.
,
Tilgner
,
K.
,
Collin
,
J.
,
Barta
,
T.
,
Al-Aama
,
J. Y.
,
Tesarov
,
L.
,
Hussain
,
R.
,
Trafford
,
A. W.
,
Kirkwood
,
G.
et al. 
(
2014
).
An induced pluripotent stem cell model of hypoplastic left heart syndrome (HLHS) reveals multiple expression and functional differences in HLHS-derived cardiac myocytes
.
Stem Cells Transl. Med.
3
,
416
-
423
.
Jiao
,
K.
,
Kulessa
,
H.
,
Tompkins
,
K.
,
Zhou
,
Y.
,
Batts
,
L.
,
Baldwin
,
H. S.
and
Hogan
,
B. L.
(
2003
).
An essential role of Bmp4 in the atrioventricular septation of the mouse heart
.
Genes Dev.
17
,
2362
-
2367
.
Jin
,
S. C.
,
Homsy
,
J.
,
Zaidi
,
S.
,
Lu
,
Q.
,
Morton
,
S.
,
DePalma
,
S. R.
,
Zeng
,
X.
,
Qi
,
H.
,
Chang
,
W.
,
Sierant
,
M. C.
et al. 
(
2017
).
Contribution of rare inherited and de novo variants in 2,871 congenital heart disease probands
.
Nat. Genet.
49
,
1593
-
1601
.
Johnson Kameny
,
R.
,
Datar
,
S. A.
,
Boehme
,
J. B.
,
Morris
,
C.
,
Zhu
,
T.
,
Goudy
,
B. D.
,
Johnson
,
E. G.
,
Galambos
,
C.
,
Raff
,
G. W.
,
Sun
,
X.
et al. 
(
2019
).
Ovine models of congenital heart disease and the consequences of hemodynamic alterations for pulmonary artery remodeling
.
Am. J. Respir. Cell Mol. Biol.
60
,
503
-
514
.
Johnson
,
R. C.
,
Datar
,
S. A.
,
Oishi
,
P. E.
,
Bennett
,
S.
,
Maki
,
J.
,
Sun
,
C.
,
Johengen
,
M.
,
He
,
Y.
,
Raff
,
G. W.
,
Redington
,
A. N.
et al. 
(
2014
).
Adaptive right ventricular performance in response to acutely increased afterload in a lamb model of congenital heart disease: evidence for enhanced Anrep effect
.
Am. J. Physiol. Heart Circ. Physiol.
306
,
H1222
-
H1230
.
Junno
,
J.
,
Bruun
,
E.
,
Gutierrez
,
J. H.
,
Erkinaro
,
T.
,
Haapsamo
,
M.
,
Acharya
,
G.
and
Räsänen
,
J.
(
2013
).
Fetal sheep left ventricle is more sensitive than right ventricle to progressively worsening hypoxemia and acidemia
.
Eur. J. Obstet. Gynecol. Reprod. Biol.
167
,
137
-
141
.
Kadari
,
A.
,
Mekala
,
S.
,
Wagner
,
N.
,
Malan
,
D.
,
Koth
,
J.
,
Doll
,
K.
,
Stappert
,
L.
,
Eckert
,
D.
,
Peitz
,
M.
,
Matthes
,
J.
et al. 
(
2015
).
Robust Generation of cardiomyocytes from human iPS cells requires precise modulation of BMP and WNT signaling
.
Stem Cell Rev Rep
11
,
560
-
569
.
Kaltenbrun
,
E.
,
Tandon
,
P.
,
Amin
,
N. M.
,
Waldron
,
L.
,
Showell
,
C.
and
Conlon
,
F. L.
(
2011
).
Xenopus: an emerging model for studying congenital heart disease
.
Birth Defects Res. A Clin. Mol. Teratol.
91
,
495
-
510
.
Kameny
,
R. J.
,
He
,
Y.
,
Morris
,
C.
,
Sun
,
C.
,
Johengen
,
M.
,
Gong
,
W.
,
Raff
,
G. W.
,
Datar
,
S. A.
,
Oishi
,
P. E.
and
Fineman
,
J. R.
(
2015
).
Right ventricular nitric oxide signaling in an ovine model of congenital heart disease: a preserved fetal phenotype
.
Am. J. Physiol. Heart Circ. Physiol.
309
,
H157
-
H165
.
Kameny
,
R. J.
,
He
,
Y.
,
Zhu
,
T.
,
Gong
,
W.
,
Raff
,
G. W.
,
Chapin
,
C. J.
,
Datar
,
S. A.
,
Boehme
,
J. T.
,
Hata
,
A.
and
Fineman
,
J. R.
(
2018
).
Analysis of the microRNA signature driving adaptive right ventricular hypertrophy in an ovine model of congenital heart disease
.
Am. J. Physiol. Heart Circ. Physiol.
315
,
H847
-
h854
.
Karakikes
,
I.
,
Ameen
,
M.
,
Termglinchan
,
V.
and
Wu
,
J. C.
(
2015
).
Human induced pluripotent stem cell-derived cardiomyocytes: insights into molecular, cellular, and functional phenotypes
.
Circ. Res.
117
,
80
-
88
.
Karbassi
,
E.
,
Fenix
,
A.
,
Marchiano
,
S.
,
Muraoka
,
N.
,
Nakamura
,
K.
,
Yang
,
X.
and
Murry
,
C. E.
(
2020
).
Cardiomyocyte maturation: advances in knowledge and implications for regenerative medicine
.
Nat Rev Cardiol
17
,
341
-
359
.
Kattman
,
S. J.
,
Witty
,
A. D.
,
Gagliardi
,
M.
,
Dubois
,
N. C.
,
Niapour
,
M.
,
Hotta
,
A.
,
Ellis
,
J.
and
Keller
,
G.
(
2011
).
Stage-specific optimization of activin/nodal and BMP signaling promotes cardiac differentiation of mouse and human pluripotent stem cell lines
.
Cell Stem Cell
8
,
228
-
240
.
Kelly
,
R. G.
,
Buckingham
,
M. E.
and
Moorman
,
A. F.
(
2014
).
Heart fields and cardiac morphogenesis
.
Cold Spring Harb Perspect Med
4
,
a015750
.
Kern
,
C. B.
,
Wessels
,
A.
,
McGarity
,
J.
,
Dixon
,
L. J.
,
Alston
,
E.
,
Argraves
,
W. S.
,
Geeting
,
D.
,
Nelson
,
C. M.
,
Menick
,
D. R.
and
Apte
,
S. S.
(
2010
).
Reduced versican cleavage due to Adamts9 haploinsufficiency is associated with cardiac and aortic anomalies
.
Matrix Biol.
29
,
304
-
316
.
Keyes
,
W. M.
,
Logan
,
C.
,
Parker
,
E.
and
Sanders
,
E. J.
(
2003
).
Expression and function of bone morphogenetic proteins in the development of the embryonic endocardial cushions
.
Anat Embryol (Berl)
207
,
135
-
147
.
Khodiyar
,
V. K.
,
Howe
,
D.
,
Talmud
,
P. J.
,
Breckenridge
,
R.
and
Lovering
,
R. C.
(
2013
).
From zebrafish heart jogging genes to mouse and human orthologs: using Gene Ontology to investigate mammalian heart development
.
F1000Res
2
,
242
.
Kim
,
R. Y.
,
Robertson
,
E. J.
and
Solloway
,
M. J.
(
2001
).
Bmp6 and Bmp7 are required for cushion formation and septation in the developing mouse heart
.
Dev. Biol.
235
,
449
-
466
.
Kim
,
M. S.
,
Fleres
,
B.
,
Lovett
,
J.
,
Anfinson
,
M.
,
Samudrala
,
S. S. K.
,
Kelly
,
L. J.
,
Teigen
,
L. E.
,
Cavanaugh
,
M.
,
Marquez
,
M.
,
Geurts
,
A. M.
et al. 
(
2020
).
Contractility of induced pluripotent stem cell-cardiomyocytes with an MYH6 head domain variant associated with hypoplastic left heart syndrome
.
Frontiers in cell and developmental biology
8
,
440
.
Kirby
,
M. L.
,
Aronstam
,
R. S.
and
Buccafusco
,
J. J.
(
1985
).
Changes in cholinergic parameters associated with failure of conotruncal septation in embryonic chick hearts after neural crest ablation
.
Circ. Res.
56
,
392
-
401
.
Kirby
,
M. L.
,
Gale
,
T. F.
and
Stewart
,
D. E.
(
1983
).
Neural crest cells contribute to normal aorticopulmonary septation
.
Science
220
,
1059
-
1061
.
Kirk
,
E. P.
,
Hyun
,
C.
,
Thomson
,
P. C.
,
Lai
,
D.
,
Castro
,
M. L.
,
Biben
,
C.
,
Buckley
,
M. F.
,
Martin
,
I. C.
,
Moran
,
C.
and
Harvey
,
R. P.
(
2006
).
Quantitative trait loci modifying cardiac atrial septal morphology and risk of patent foramen ovale in the mouse
.
Circ. Res.
98
,
651
-
658
.
Kitaguchi
,
T.
,
Nagai
,
T.
,
Nakata
,
K.
,
Aruga
,
J.
and
Mikoshiba
,
K.
(
2000
).
Zic3 is involved in the left-right specification of the Xenopus embryo
.
Development
127
,
4787
-
4795
.
Kloesel
,
B.
,
DiNardo
,
J. A.
and
Body
,
S. C.
(
2016
).
Cardiac Embryology and Molecular Mechanisms of Congenital Heart Disease: A Primer for Anesthesiologists
.
Anesth. Analg.
123
,
551
-
569
.
Kobayashi
,
J.
,
Yoshida
,
M.
,
Tarui
,
S.
,
Hirata
,
M.
,
Nagai
,
Y.
,
Kasahara
,
S.
,
Naruse
,
K.
,
Ito
,
H.
,
Sano
,
S.
and
Oh
,
H.
(
2014
).
Directed differentiation of patient-specific induced pluripotent stem cells identifies the transcriptional repression and epigenetic modification of NKX2-5, HAND1, and NOTCH1 in hypoplastic left heart syndrome
.
PLoS One
9
,
e102796
.
Koivumäki
,
J. T.
,
Naumenko
,
N.
,
Tuomainen
,
T.
,
Takalo
,
J.
,
Oksanen
,
M.
,
Puttonen
,
K. A.
,
Lehtonen
,
Š.
,
Kuusisto
,
J.
,
Laakso
,
M.
,
Koistinaho
,
J.
et al. 
(
2018
).
Structural immaturity of human iPSC-derived cardiomyocytes: in silico investigation of effects on function and disease modeling
.
Front Physiol
9
,
80
.
Kooij
,
V.
,
Venkatraman
,
V.
,
Tra
,
J.
,
Kirk
,
J. A.
,
Rowell
,
J.
,
Blice-Baum
,
A.
,
Cammarato
,
A.
and
Van Eyk
,
J. E.
(
2014
).
Sizing up models of heart failure: proteomics from flies to humans
.
Proteomics Clin. Appl.
8
,
653
-
664
.
Koopmann
,
T. T.
,
Adriaens
,
M. E.
,
Moerland
,
P. D.
,
Marsman
,
R. F.
,
Westerveld
,
M. L.
,
Lal
,
S.
,
Zhang
,
T.
,
Simmons
,
C. Q.
,
Baczko
,
I.
,
dos Remedios
,
C.
et al. 
(
2014
).
Genome-wide identification of expression quantitative trait loci (eQTLs) in human heart
.
PLoS ONE
9
,
e97380
.
Kosho
,
T.
,
Okamoto
,
N.
,
Ohashi
,
H.
,
Tsurusaki
,
Y.
,
Imai
,
Y.
,
Hibi-Ko
,
Y.
,
Kawame
,
H.
,
Homma
,
T.
,
Tanabe
,
S.
,
Kato
,
M.
et al. 
(
2013
).
Clinical correlations of mutations affecting six components of the SWI/SNF complex: detailed description of 21 patients and a review of the literature
.
Am. J. Med. Genet. A
161a
,
1221
-
1237
.
Kowalski
,
W. J.
,
Teslovich
,
N. C.
,
Menon
,
P. G.
,
Tinney
,
J. P.
,
Keller
,
B. B.
and
Pekkan
,
K.
(
2014
).
Left atrial ligation alters intracardiac flow patterns and the biomechanical landscape in the chick embryo
.
Dev. Dyn.
243
,
652
-
662
.
Kraft
,
P.
and
de Andrade
,
M.
(
2003
).
Group 6: Pleiotropy and multivariate analysis
.
Genet. Epidemiol.
25
Suppl. 1
,
S50
-
S56
.
Krenz
,
M.
,
Yutzey
,
K. E.
and
Robbins
,
J.
(
2005
).
Noonan syndrome mutation Q79R in Shp2 increases proliferation of valve primordia mesenchymal cells via extracellular signal-regulated kinase 1/2 signaling
.
Circ. Res.
97
,
813
-
820
.
Kuciene
,
R.
and
Dulskiene
,
V.
(
2008
).
Selected environmental risk factors and congenital heart defects
.
Medicina (Kaunas)
44
,
827
-
832
.
Kumar
,
S.
and
Lodge
,
J.
(
2019
).
Prenatal therapy for fetal cardiac disorders
.
J. Matern. Fetal. Neonatal. Med.
32
,
3871
-
3881
.
Kumar
,
S. D.
,
Dheen
,
S. T.
and
Tay
,
S. S.
(
2007
).
Maternal diabetes induces congenital heart defects in mice by altering the expression of genes involved in cardiovascular development
.
Cardiovasc. Diabetol.
6
,
34
.
Kumar
,
S. D.
,
Yong
,
S. K.
,
Dheen
,
S. T.
,
Bay
,
B. H.
and
Tay
,
S. S.
(
2008
).
Cardiac malformations are associated with altered expression of vascular endothelial growth factor and endothelial nitric oxide synthase genes in embryos of diabetic mice
.
Exp. Biol. Med. (Maywood)
233
,
1421
-
1432
.
Kume
,
T.
,
Jiang
,
H.
,
Topczewska
,
J. M.
and
Hogan
,
B. L.
(
2001
).
The murine winged helix transcription factors, Foxc1 and Foxc2, are both required for cardiovascular development and somitogenesis
.
Genes Dev.
15
,
2470
-
2482
.
LaHaye
,
S.
,
Majumdar
,
U.
,
Yasuhara
,
J.
,
Koenig
,
S. N.
,
Matos-Nieves
,
A.
,
Kumar
,
R.
and
Garg
,
V.
(
2019
).
Developmental origins for semilunar valve stenosis identified in mice harboring congenital heart disease-associated <em>GATA4</em> mutation
.
Disease Models &amp; Mechanisms
12
,
dmm036764
.
Lahm
,
H.
,
Schon
,
P.
,
Doppler
,
S.
,
Dressen
,
M.
,
Cleuziou
,
J.
,
Deutsch
,
M. A.
,
Ewert
,
P.
,
Lange
,
R.
and
Krane
,
M.
(
2015
).
Tetralogy of fallot and hypoplastic left heart syndrome - complex clinical phenotypes meet complex genetic networks
.
Curr. Genomics
16
,
141
-
158
.
Lai
,
C. Q.
,
Lim
,
G. L.
,
Jamil
,
M.
,
Mattar
,
C. N.
,
Biswas
,
A.
and
Yap
,
C. H.
(
2016
).
Fluid mechanics of blood flow in human fetal left ventricles based on patient-specific 4D ultrasound scans
.
Biomech. Model. Mechanobiol.
15
,
1159
-
1172
.
Lalani
,
S. R.
,
Shaw
,
C.
,
Wang
,
X.
,
Patel
,
A.
,
Patterson
,
L. W.
,
Kolodziejska
,
K.
,
Szafranski
,
P.
,
Ou
,
Z.
,
Tian
,
Q.
,
Kang
,
S. H.
et al. 
(
2013
).
Rare DNA copy number variants in cardiovascular malformations with extracardiac abnormalities
.
Eur. J. Hum. Genet.
21
,
173
-
181
.
Lam
,
Y. Y.
,
Keung
,
W.
,
Chan
,
C. H.
,
Geng
,
L.
,
Wong
,
N.
,
Brenière-Letuffe
,
D.
,
Li
,
R. A.
and
Cheung
,
Y. F.
(
2020
).
Single-cell transcriptomics of engineered cardiac tissues from patient-specific induced pluripotent stem cell-derived cardiomyocytes reveals abnormal developmental trajectory and intrinsic contractile defects in hypoplastic right heart syndrome
.
J Am Heart Assoc
9
,
e016528
.
Layman
,
W. S.
,
McEwen
,
D. P.
,
Beyer
,
L. A.
,
Lalani
,
S. R.
,
Fernbach
,
S. D.
,
Oh
,
E.
,
Swaroop
,
A.
,
Hegg
,
C. C.
,
Raphael
,
Y.
,
Martens
,
J. R.
et al. 
(
2009
).
Defects in neural stem cell proliferation and olfaction in Chd7 deficient mice indicate a mechanism for hyposmia in human CHARGE syndrome
.
Hum. Mol. Genet.
18
,
1909
-
1923
.
Le Garrec
,
J.-F.
,
Ragni
,
C. V.
,
Pop
,
S.
,
Dufour
,
A.
,
Olivo-Marin
,
J.-C.
,
Buckingham
,
M. E.
and
Meilhac
,
S. M.
(
2013
).
Quantitative analysis of polarity in 3D reveals local cell coordination in the embryonic mouse heart
.
Development
140
,
395
-
404
.
Lee
,
S.
,
Lee
,
J. W.
and
Lee
,
S. K.
(
2012
).
UTX, a histone H3-lysine 27 demethylase, acts as a critical switch to activate the cardiac developmental program
.
Dev. Cell
22
,
25
-
37
.
Lee
,
S.
,
Serpooshan
,
V.
,
Tong
,
X.
,
Venkatraman
,
S.
,
Lee
,
M.
,
Lee
,
J.
,
Chirikian
,
O.
,
Wu
,
J. C.
,
Wu
,
S. M.
and
Yang
,
F.
(
2017
).
Contractile force generation by 3D hiPSC-derived cardiac tissues is enhanced by rapid establishment of cellular interconnection in matrix with muscle-mimicking stiffness
.
Biomaterials
131
,
111
-
120
.
Lee
,
T. H.
,
Song
,
S. H.
,
Kim
,
K. L.
,
Yi
,
J. Y.
,
Shin
,
G. H.
,
Kim
,
J. Y.
,
Kim
,
J.
,
Han
,
Y. M.
,
Lee
,
S. H.
,
Lee
,
S. H.
et al. 
(
2010
).
Functional recapitulation of smooth muscle cells via induced pluripotent stem cells from human aortic smooth muscle cells
.
Circ. Res.
106
,
120
-
128
.
Lemoine
,
M. D.
,
Mannhardt
,
I.
,
Breckwoldt
,
K.
,
Prondzynski
,
M.
,
Flenner
,
F.
,
Ulmer
,
B.
,
Hirt
,
M. N.
,
Neuber
,
C.
,
Horvath
,
A.
,
Kloth
,
B.
et al. 
(
2017
).
Human iPSC-derived cardiomyocytes cultured in 3D engineered heart tissue show physiological upstroke velocity and sodium current density
.
Sci. Rep.
7
,
5464
.
León
,
L. E.
,
Benavides
,
F.
,
Espinoza
,
K.
,
Vial
,
C.
,
Alvarez
,
P.
,
Palomares
,
M.
,
Lay-Son
,
G.
,
Miranda
,
M.
and
Repetto
,
G. M.
(
2017
).
Partial microduplication in the histone acetyltransferase complex member KANSL1 is associated with congenital heart defects in 22q11.2 microdeletion syndrome patients
.
Sci. Rep.
7
,
1795
.
Lepic
,
E.
,
Burger
,
D.
,
Lu
,
X.
,
Song
,
W.
and
Feng
,
Q.
(
2006
).
Lack of endothelial nitric oxide synthase decreases cardiomyocyte proliferation and delays cardiac maturation
.
Am. J. Physiol. Cell Physiol.
291
,
C1240
-
C1246
.
Li
,
J.
,
Cao
,
Y.
,
Ma
,
X. J.
,
Wang
,
H. J.
,
Zhang
,
J.
,
Luo
,
X.
,
Chen
,
W.
,
Wu
,
Y.
,
Meng
,
Y.
,
Zhang
,
J.
et al. 
(
2013
).
Roles of miR-1-1 and miR-181c in ventricular septal defects
.
Int. J. Cardiol.
168
,
1441
-
1446
.
Li
,
D.
,
Ji
,
L.
,
Liu
,
L.
,
Liu
,
Y.
,
Hou
,
H.
,
Yu
,
K.
,
Sun
,
Q.
and
Zhao
,
Z.
(
2014a
).
Characterization of circulating microRNA expression in patients with a ventricular septal defect
.
PLoS ONE
9
,
e106318
.
Li
,
X.
,
Martinez-Fernandez
,
A.
,
Hartjes
,
K. A.
,
Kocher
,
J. P.
,
Olson
,
T. M.
,
Terzic
,
A.
and
Nelson
,
T. J.
(
2014b
).
Transcriptional atlas of cardiogenesis maps congenital heart disease interactome
.
Physiol. Genomics
46
,
482
-
495
.
Li
,
N.
,
Subrahmanyan
,
L.
,
Smith
,
E.
,
Yu
,
X.
,
Zaidi
,
S.
,
Choi
,
M.
,
Mane
,
S.
,
Nelson-Williams
,
C.
,
Behjati
,
M.
,
Kazemi
,
M.
et al. 
(
2016
).
Mutations in the histone modifier PRDM6 are associated with isolated nonsyndromic patent ductus arteriosus
.
Am. J. Hum. Genet.
99
,
1000
.
Lian
,
X.
,
Hsiao
,
C.
,
Wilson
,
G.
,
Zhu
,
K.
,
Hazeltine
,
L. B.
,
Azarin
,
S. M.
,
Raval
,
K. K.
,
Zhang
,
J.
,
Kamp
,
T. J.
and
Palecek
,
S. P.
(
2012
).
Robust cardiomyocyte differentiation from human pluripotent stem cells via temporal modulation of canonical Wnt signaling
.
Proc. Natl. Acad. Sci. USA
109
,
E1848
-
E1857
.
Liao
,
J.
,
Wu
,
Z.
,
Wang
,
Y.
,
Cheng
,
L.
,
Cui
,
C.
,
Gao
,
Y.
,
Chen
,
T.
,
Rao
,
L.
,
Chen
,
S.
,
Jia
,
N.
et al. 
(
2008
).
Enhanced efficiency of generating induced pluripotent stem (iPS) cells from human somatic cells by a combination of six transcription factors
.
Cell Res.
18
,
600
-
603
.
Lin
,
C. R.
,
Kioussi
,
C.
,
O'Connell
,
S.
,
Briata
,
P.
,
Szeto
,
D.
,
Liu
,
F.
,
Izpisúa-Belmonte
,
J. C.
and
Rosenfeld
,
M. G.
(
1999
).
Pitx2 regulates lung asymmetry, cardiac positioning and pituitary and tooth morphogenesis
.
Nature
401
,
279
-
282
.
Lindsay
,
E. A.
,
Botta
,
A.
,
Jurecic
,
V.
,
Carattini-Rivera
,
S.
,
Cheah
,
Y.-C.
,
Rosenblatt
,
H. M.
,
Bradley
,
A.
and
Baldini
,
A.
(
1999
).
Congenital heart disease in mice deficient for the DiGeorge syndrome region
.
Nature
401
,
379
-
383
.
Liu
,
Y.
and
Feng
,
Q.
(
2012
).
NOing the heart: role of nitric oxide synthase-3 in heart development
.
Differentiation
84
,
54
-
61
.
Liu
,
J.
and
Stainier
,
D. Y.
(
2010
).
Tbx5 and Bmp signaling are essential for proepicardium specification in zebrafish
.
Circ. Res.
106
,
1818
-
1828
.
Liu
,
W.
,
Selever
,
J.
,
Wang
,
D.
,
Lu
,
M. F.
,
Moses
,
K. A.
,
Schwartz
,
R. J.
and
Martin
,
J. F.
(
2004
).
Bmp4 signaling is required for outflow-tract septation and branchial-arch artery remodeling
.
Proc. Natl. Acad. Sci. USA
101
,
4489
-
4494
.
Liu
,
X.
,
Chen
,
W.
,
Li
,
W.
,
Li
,
Y.
,
Priest
,
J. R.
,
Zhou
,
B.
,
Wang
,
J.
and
Zhou
,
Z.
(
2019
).
Single-Cell RNA-Seq of the developing cardiac outflow tract reveals convergent development of the vascular smooth muscle cells
.
Cell Rep
28
,
1346
-
1361.e4
.
Lluch
,
È.
,
De Craene
,
M.
,
Bijnens
,
B.
,
Sermesant
,
M.
,
Noailly
,
J.
,
Camara
,
O.
and
Morales
,
H. G.
(
2019
).
Breaking the state of the heart: meshless model for cardiac mechanics
.
Biomech. Model. Mechanobiol.
18
,
1549
-
1561
.
Lo
,
C. W.
,
Waldo
,
K. L.
and
Kirby
,
M. L.
(
1999
).
Gap junction communication and the modulation of cardiac neural crest cells
.
Trends Cardiovasc. Med.
9
,
63
-
69
.
Lu
,
M. F.
,
Pressman
,
C.
,
Dyer
,
R.
,
Johnson
,
R. L.
and
Martin
,
J. F.
(
1999
).
Function of Rieger syndrome gene in left-right asymmetry and craniofacial development
.
Nature
401
,
276
-
278
.
Luna-Zurita
,
L.
,
Prados
,
B.
,
Grego-Bessa
,
J.
,
Luxán
,
G.
,
del Monte
,
G.
,
Benguría
,
A.
,
Adams
,
R. H.
,
Pérez-Pomares
,
J. M.
and
de la Pompa
,
J. L.
(
2010
).
Integration of a Notch-dependent mesenchymal gene program and Bmp2-driven cell invasiveness regulates murine cardiac valve formation
.
J. Clin. Invest.
120
,
3493
-
3507
.
Lyu
,
G.
,
Zhang
,
C.
,
Ling
,
T.
,
Liu
,
R.
,
Zong
,
L.
,
Guan
,
Y.
,
Huang
,
X.
,
Sun
,
L.
,
Zhang
,
L.
,
Li
,
C.
et al. 
(
2018
).
Genome and epigenome analysis of monozygotic twins discordant for congenital heart disease
.
BMC Genomics
19
,
428
.
Ma
,
L.
,
Lu
,
M. F.
,
Schwartz
,
R. J.
and
Martin
,
J. F.
(
2005
).
Bmp2 is essential for cardiac cushion epithelial-mesenchymal transition and myocardial patterning
.
Development
132
,
5601
-
5611
.
Ma
,
J.
,
Guo
,
L.
,
Fiene
,
S. J.
,
Anson
,
B. D.
,
Thomson
,
J. A.
,
Kamp
,
T. J.
,
Kolaja
,
K. L.
,
Swanson
,
B. J.
and
January
,
C. T.
(
2011
).
High purity human-induced pluripotent stem cell-derived cardiomyocytes: electrophysiological properties of action potentials and ionic currents
.
Am. J. Physiol. Heart Circ. Physiol.
301
,
H2006
-
H2017
.
Maherali
,
N.
,
Ahfeldt
,
T.
,
Rigamonti
,
A.
,
Utikal
,
J.
,
Cowan
,
C.
and
Hochedlinger
,
K.
(
2008
).
A high-efficiency system for the generation and study of human induced pluripotent stem cells
.
Cell Stem Cell
3
,
340
-
345
.
Makikallio
,
K.
,
McElhinney
,
D. B.
,
Levine
,
J. C.
,
Marx
,
G. R.
,
Colan
,
S. D.
,
Marshall
,
A. C.
,
Lock
,
J. E.
,
Marcus
,
E. N.
and
Tworetzky
,
W.
(
2006
).
Fetal aortic valve stenosis and the evolution of hypoplastic left heart syndrome: patient selection for fetal intervention
.
Circulation
113
,
1401
-
1405
.
Mannhardt
,
I.
,
Breckwoldt
,
K.
,
Letuffe-Brenière
,
D.
,
Schaaf
,
S.
,
Schulz
,
H.
,
Neuber
,
C.
,
Benzin
,
A.
,
Werner
,
T.
,
Eder
,
A.
,
Schulze
,
T.
et al. 
(
2016
).
Human Engineered Heart Tissue: Analysis of Contractile Force
.
Stem Cell Reports
7
,
29
-
42
.
Marguerie
,
A.
,
Bajolle
,
F.
,
Zaffran
,
S.
,
Brown
,
N. A.
,
Dickson
,
C.
,
Buckingham
,
M. E.
and
Kelly
,
R. G.
(
2006
).
Congenital heart defects in Fgfr2-IIIb and Fgf10 mutant mice
.
Cardiovasc. Res.
71
,
50
-
60
.
Marques
,
S. R.
and
Yelon
,
D.
(
2009
).
Differential requirement for BMP signaling in atrial and ventricular lineages establishes cardiac chamber proportionality
.
Dev. Biol.
328
,
472
-
482
.
Martinsen
,
B. J.
(
2005
).
Reference guide to the stages of chick heart embryology
.
Dev. Dyn.
233
,
1217
-
1237
.
Meberg
,
A.
,
Hals
,
J.
and
Thaulow
,
E.
(
2007
).
Congenital heart defects--chromosomal anomalies, syndromes and extracardiac malformations
.
Acta Paediatr.
96
,
1142
-
1145
.
Meoli
,
A.
,
Cutrì
,
E.
,
Krishnamurthy
,
A.
,
Dubini
,
G.
,
Migliavacca
,
F.
,
Hsia
,
T.-Y.
,
Pennati
,
G.
,
Group
,
M. o. C. H. A.
,
Taylor
,
A.
and
Giardini
,
A.
(
2015
).
A multiscale model for the study of cardiac biomechanics in single-ventricle surgeries: a clinical case
.
Interface focus
5
,
20140079
.
Midgett
,
M.
,
Lopez
,
C. S.
,
David
,
L.
,
Maloyan
,
A.
and
Rugonyi
,
S.
(
2017
).
Increased hemodynamic load in early embryonic stages alters endocardial to mesenchymal transition
.
Front Physiol
8
,
56
.
Misra
,
C.
,
Sachan
,
N.
,
McNally
,
C. R.
,
Koenig
,
S. N.
,
Nichols
,
H. A.
,
Guggilam
,
A.
,
Lucchesi
,
P. A.
,
Pu
,
W. T.
,
Srivastava
,
D.
and
Garg
,
V.
(
2012
).
Congenital heart disease-causing Gata4 mutation displays functional deficits in vivo
.
PLoS Genet.
8
,
e1002690
.
Mitchell
,
M. E.
,
Sander
,
T. L.
,
Klinkner
,
D. B.
and
Tomita-Mitchell
,
A.
(
2007
).
The molecular basis of congenital heart disease
.
Semin. Thorac. Cardiovasc. Surg.
19
,
228
-
237
.
Miura
,
G. I.
and
Yelon
,
D.
(
2011
).
A guide to analysis of cardiac phenotypes in the zebrafish embryo
.
Methods Cell Biol.
101
,
161
-
180
.
Moazzen
,
H.
,
Lu
,
X.
,
Ma
,
N. L.
,
Velenosi
,
T. J.
,
Urquhart
,
B. L.
,
Wisse
,
L. J.
,
Gittenberger-de Groot
,
A. C.
and
Feng
,
Q.
(
2014
).
N-Acetylcysteine prevents congenital heart defects induced by pregestational diabetes
.
Cardiovasc. Diabetol.
13
,
46
.
Mohun
,
T. J.
,
Leong
,
L. M.
,
Weninger
,
W. J.
and
Sparrow
,
D. B.
(
2000
).
The morphology of heart development in Xenopus laevis
.
Dev. Biol.
218
,
74
-
88
.
Montgomery
,
M. O.
,
Jiao
,
Y.
,
Phillips
,
S. J.
,
Singh
,
G.
,
Xu
,
J.
,
Balsara
,
R.
and
Litvin
,
J.
(
1998
).
Alterations in sheep fetal right ventricular tissue with induced hemodynamic pressure overload
.
Basic Res. Cardiol.
93
,
192
-
200
.
Moorman
,
A. F.
and
Christoffels
,
V. M.
(
2003
).
Cardiac chamber formation: development, genes, and evolution
.
Physiol. Rev.
83
,
1223
-
1267
.
Moran
,
S.
,
Chi
,
T.
,
Prucha
,
M. S.
,
Ahn
,
K. S.
,
Connor-Stroud
,
F.
,
Jean
,
S.
,
Gould
,
K.
and
Chan
,
A. W.
(
2015
).
Germline transmission in transgenic Huntington's disease monkeys
.
Theriogenology
84
,
277
-
285
.
Moreau
,
J. L. M.
,
Kesteven
,
S.
,
Martin
,
E.
,
Lau
,
K. S.
,
Yam
,
M. X.
,
O'Reilly
,
V. C.
,
Del Monte-Nieto
,
G.
,
Baldini
,
A.
,
Feneley
,
M. P.
,
Moon
,
A. M.
et al. 
(
2019
).
Gene-environment interaction impacts on heart development and embryo survival
.
Development
146
,
Morin
,
X.
,
Daneman
,
R.
,
Zavortink
,
M.
and
Chia
,
W.
(
2001
).
A protein trap strategy to detect GFP-tagged proteins expressed from their endogenous loci in Drosophila
.
Proc. Natl. Acad. Sci. USA
98
,
15050
-
15055
.
Morris
,
P. D.
,
Narracott
,
A.
,
von Tengg-Kobligk
,
H.
,
Soto
,
D. A. S.
,
Hsiao
,
S.
,
Lungu
,
A.
,
Evans
,
P.
,
Bressloff
,
N. W.
,
Lawford
,
P. V.
and
Hose
,
D. R.
(
2016
).
Computational fluid dynamics modelling in cardiovascular medicine
.
Heart
102
,
18
-
28
.
Morrison-Graham
,
K.
,
Schatteman
,
G. C.
,
Bork
,
T.
,
Bowen-Pope
,
D. F.
and
Weston
,
J. A.
(
1992
).
A PDGF receptor mutation in the mouse (Patch) perturbs the development of a non-neuronal subset of neural crest-derived cells
.
Development
115
,
133
-
142
.
Mummery
,
C. L.
,
Zhang
,
J.
,
Ng
,
E. S.
,
Elliott
,
D. A.
,
Elefanty
,
A. G.
and
Kamp
,
T. J.
(
2012
).
Differentiation of human embryonic stem cells and induced pluripotent stem cells to cardiomyocytes: a methods overview
.
Circ. Res.
111
,
344
-
358
.
Nadeau
,
M.
,
Georges
,
R. O.
,
Laforest
,
B.
,
Yamak
,
A.
,
Lefebvre
,
C.
,
Beauregard
,
J.
,
Paradis
,
P.
,
Bruneau
,
B. G.
,
Andelfinger
,
G.
and
Nemer
,
M.
(
2010
).
An endocardial pathway involving Tbx5, Gata4, and Nos3 required for atrial septum formation
.
Proc. Natl. Acad. Sci. USA
107
,
19356
-
19361
.
Nakano
,
H.
,
Minami
,
I.
,
Braas
,
D.
,
Pappoe
,
H.
,
Wu
,
X.
,
Sagadevan
,
A.
,
Vergnes
,
L.
,
Fu
,
K.
,
Morselli
,
M.
,
Dunham
,
C.
et al. 
(
2017
).
Glucose inhibits cardiac muscle maturation through nucleotide biosynthesis
.
Elife
6
.
Nguyen
,
A. T.
,
Xiao
,
B.
,
Neppl
,
R. L.
,
Kallin
,
E. M.
,
Li
,
J.
,
Chen
,
T.
,
Wang
,
D. Z.
,
Xiao
,
X.
and
Zhang
,
Y.
(
2011
).
DOT1L regulates dystrophin expression and is critical for cardiac function
.
Genes Dev.
25
,
263
-
274
.
Ni
,
X.
,
Xu
,
K.
,
Zhao
,
Y.
,
Li
,
J.
,
Wang
,
L.
,
Yu
,
F.
and
Li
,
G.
(
2020
).
Single-cell analysis reveals the purification and maturation effects of glucose starvation in hiPSC-CMs
.
Biochem. Biophys. Res. Commun.
534
,
367
-
373
.
Nimura
,
K.
,
Ura
,
K.
,
Shiratori
,
H.
,
Ikawa
,
M.
,
Okabe
,
M.
,
Schwartz
,
R. J.
and
Kaneda
,
Y.
(
2009
).
A histone H3 lysine 36 trimethyltransferase links Nkx2-5 to Wolf-Hirschhorn syndrome
.
Nature
460
,
287
-
291
.
Nishimura
,
M.
,
Kumsta
,
C.
,
Kaushik
,
G.
,
Diop
,
S. B.
,
Ding
,
Y.
,
Bisharat-Kernizan
,
J.
,
Catan
,
H.
,
Cammarato
,
A.
,
Ross
,
R. S.
,
Engler
,
A. J.
et al. 
(
2014
).
A dual role for integrin-linked kinase and β1-integrin in modulating cardiac aging
.
Aging Cell
13
,
431
-
440
.
Ohayon
,
J.
,
Cai
,
H.
,
Jouk
,
P.-S.
,
Usson
,
Y.
and
Azancot
,
A.
(
2002
).
A model of the structural and functional development of the normal human fetal left ventricle based on a global growth law
.
Computer Methods in Biomech. Biomed. Engin.
5
,
113
-
126
.
Okita
,
K.
,
Yamakawa
,
T.
,
Matsumura
,
Y.
,
Sato
,
Y.
,
Amano
,
N.
,
Watanabe
,
A.
,
Goshima
,
N.
and
Yamanaka
,
S.
(
2013
).
An efficient nonviral method to generate integration-free human-induced pluripotent stem cells from cord blood and peripheral blood cells
.
Stem Cells
31
,
458
-
466
.
Ong
,
C. W.
,
Ren
,
M.
,
Wiputra
,
H.
,
Mojumder
,
J.
,
Chan
,
W. X.
,
Tulzer
,
A.
,
Tulzer
,
G.
,
Buist
,
M. L.
,
Mattar
,
C. N. Z.
and
Lee
,
L. C.
(
2020
).
Biomechanics of human fetal hearts with critical aortic stenosis
.
Ann. Biomed. Eng.
1
-
16
.
Ottolia
,
M.
,
Torres
,
N.
,
Bridge
,
J. H. B.
,
Philipson
,
K. D.
and
Goldhaber
,
J. I.
(
2013
).
Na/Ca exchange and contraction of the heart
.
J. Mol. Cell. Cardiol.
61
,
28
-
33
.
Parrotta
,
E. I.
,
Lucchino
,
V.
,
Scaramuzzino
,
L.
,
Scalise
,
S.
and
Cuda
,
G.
(
2020
).
Modeling cardiac disease mechanisms using induced pluripotent stem cell-derived cardiomyocytes: progress, promises and challenges
.
Int. J. Mol. Sci.
21
.
Pediatric Cardiac Genomics
,
C.
,
Gelb
,
B.
,
Brueckner
,
M.
,
Chung
,
W.
,
Goldmuntz
,
E.
,
Kaltman
,
J.
,
Kaski
,
J. P.
,
Kim
,
R.
,
Kline
,
J.
,
Mercer-Rosa
,
L.
et al. 
(
2013
).
The congenital heart disease genetic network study: rationale, design, and early results
.
Circ. Res.
112
,
698
-
706
.
Peña
,
E.
,
Tracqui
,
P.
,
Azancot
,
A.
,
Doblare
,
M.
and
Ohayon
,
J.
(
2010
).
Unraveling changes in myocardial contractility during human fetal growth: a finite element analysis based on in vivo ultrasound measurements
.
Ann. Biomed. Eng.
38
,
2702
-
2715
.
Peng
,
J.
(
2019
).
Gene redundancy and gene compensation: an updated view
.
J. Genet. Genomics
46
,
329
-
333
.
Pennati
,
G.
and
Fumero
,
R.
(
2000
).
Scaling approach to study the changes through the gestation of human fetal cardiac and circulatory behaviors
.
Ann. Biomed. Eng.
28
,
442
-
452
.
Pesevski
,
Z.
,
Kvasilova
,
A.
,
Stopkova
,
T.
,
Nanka
,
O.
,
Drobna Krejci
,
E.
,
Buffinton
,
C.
,
Kockova
,
R.
,
Eckhardt
,
A.
and
Sedmera
,
D.
(
2018
).
Endocardial fibroelastosis is secondary to hemodynamic alterations in the chick embryonic model of hypoplastic left heart syndrome
.
Dev. Dyn.
247
,
509
-
520
.
Peterkin
,
T.
,
Gibson
,
A.
and
Patient
,
R.
(
2007
).
Redundancy and evolution of GATA factor requirements in development of the myocardium
.
Dev. Biol.
311
,
623
-
635
.
Philbrook
,
N. A.
,
Nikolovska
,
A.
,
Maciver
,
R. D.
,
Belanger
,
C. L.
and
Winn
,
L. M.
(
2019
).
Characterizing the effects of in utero exposure to valproic acid on murine fetal heart development
.
Birth Defects Res
111
,
1551
-
1560
.
Pierpont
,
M. E.
,
Basson
,
C. T.
,
Benson
,
D. W.
, Jr
,
Gelb
,
B. D.
,
Giglia
,
T. M.
,
Goldmuntz
,
E.
,
McGee
,
G.
,
Sable
,
C. A.
,
Srivastava
,
D.
,
Webb
,
C. L.
et al. 
(
2007
).
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
115
,
3015
-
3038
.
Pierpont
,
M. E.
,
Brueckner
,
M.
,
Chung
,
W. K.
,
Garg
,
V.
,
Lacro
,
R. V.
,
McGuire
,
A. L.
,
Mital
,
S.
,
Priest
,
J. R.
,
Pu
,
W. T.
,
Roberts
,
A.
et al. 
(
2018
).
Genetic basis for congenital heart disease: revisited: a scientific statement from the american heart association
.
Circulation
138
,
e653
-
e711
.
Poon
,
K. L.
and
Brand
,
T.
(
2013
).
The zebrafish model system in cardiovascular research: a tiny fish with mighty prospects
.
Global cardiology science & practice
2013
,
9
-
28
.
Potts
,
J. D.
,
Dagle
,
J. M.
,
Walder
,
J. A.
,
Weeks
,
D. L.
and
Runyan
,
R. B.
(
1991
).
Epithelial-mesenchymal transformation of embryonic cardiac endothelial cells is inhibited by a modified antisense oligodeoxynucleotide to transforming growth factor beta 3
.
Proc. Natl. Acad. Sci. USA
88
,
1516
-
1520
.
Preissl
,
S.
,
Schwaderer
,
M.
,
Raulf
,
A.
,
Hesse
,
M.
,
Grüning
,
B. A.
,
Köbele
,
C.
,
Backofen
,
R.
,
Fleischmann
,
B. K.
,
Hein
,
L.
and
Gilsbach
,
R.
(
2015
).
Deciphering the epigenetic code of cardiac myocyte transcription
.
Circ. Res.
117
,
413
-
423
.
Proudfoot
,
C.
,
Carlson
,
D. F.
,
Huddart
,
R.
,
Long
,
C. R.
,
Pryor
,
J. H.
,
King
,
T. J.
,
Lillico
,
S. G.
,
Mileham
,
A. J.
,
McLaren
,
D. G.
,
Whitelaw
,
C. B.
et al. 
(
2015
).
Genome edited sheep and cattle
.
Transgenic Res.
24
,
147
-
153
.
Qi
,
X.
,
Yang
,
G.
,
Yang
,
L.
,
Lan
,
Y.
,
Weng
,
T.
,
Wang
,
J.
,
Wu
,
Z.
,
Xu
,
J.
,
Gao
,
X.
and
Yang
,
X.
(
2007
).
Essential role of Smad4 in maintaining cardiomyocyte proliferation during murine embryonic heart development
.
Dev. Biol.
311
,
136
-
146
.
Qian
,
Y.
,
Xiao
,
D.
,
Guo
,
X.
,
Chen
,
H.
,
Hao
,
L.
,
Ma
,
X.
,
Huang
,
G.
,
Ma
,
D.
and
Wang
,
H.
(
2017
).
Hypomethylation and decreased expression of BRG1 in the myocardium of patients with congenital heart disease
.
Birth Defects Res
109
,
1183
-
1195
.
Rasouli
,
S. J.
,
El-Brolosy
,
M.
,
Tsedeke
,
A. T.
,
Bensimon-Brito
,
A.
,
Ghanbari
,
P.
,
Maischein
,
H.-M.
,
Kuenne
,
C.
and
Stainier
,
D. Y.
(
2018
).
The flow responsive transcription factor Klf2 is required for myocardial wall integrity by modulating Fgf signaling
.
Elife
7
,
e38889
.
Reim
,
I.
and
Frasch
,
M.
(
2005
).
The Dorsocross T-box genes are key components of the regulatory network controlling early cardiogenesis in Drosophila
.
Development
132
,
4911
-
4925
.
Reiter
,
L. T.
,
Potocki
,
L.
,
Chien
,
S.
,
Gribskov
,
M.
and
Bier
,
E.
(
2001
).
A systematic analysis of human disease-associated gene sequences in Drosophila melanogaster
.
Genome Res.
11
,
1114
-
1125
.
Rezaee
,
T.
and
Hassani
,
K.
(
2016
).
Numerical investigation of the haemodynamics in the human fetal umbilical vein/ductus venosus based on the experimental data
.
Biosci. Rep.
36
,
e00384
.
Richards
,
A. A.
,
Santos
,
L. J.
,
Nichols
,
H. A.
,
Crider
,
B. P.
,
Elder
,
F. F.
,
Hauser
,
N. S.
,
Zinn
,
A. R.
and
Garg
,
V.
(
2008
).
Cryptic chromosomal abnormalities identified in children with congenital heart disease
.
Pediatr. Res.
64
,
358
-
363
.
Rivera-Feliciano
,
J.
and
Tabin
,
C. J.
(
2006
).
Bmp2 instructs cardiac progenitors to form the heart-valve-inducing field
.
Dev. Biol.
295
,
580
-
588
.
Robson
,
A.
,
Makova
,
S. Z.
,
Barish
,
S.
,
Zaidi
,
S.
,
Mehta
,
S.
,
Drozd
,
J.
,
Jin
,
S. C.
,
Gelb
,
B. D.
,
Seidman
,
C. E.
,
Chung
,
W. K.
et al. 
(
2019
).
Histone H2B monoubiquitination regulates heart development via epigenetic control of cilia motility
.
Proc. Natl. Acad. Sci. USA
116
,
14049
-
14054
.
Ronaldson-Bouchard
,
K.
,
Yeager
,
K.
,
Teles
,
D.
,
Chen
,
T.
,
Ma
,
S.
,
Song
,
L.
,
Morikawa
,
K.
,
Wobma
,
H. M.
,
Vasciaveo
,
A.
,
Ruiz
,
E. C.
et al. 
(
2019
).
Engineering of human cardiac muscle electromechanically matured to an adult-like phenotype
.
Nat. Protoc.
14
,
2781
-
2817
.
Roos-Hesselink
,
J. W.
,
Kerstjens-Frederikse
,
W. S.
,
Meijboom
,
F. J.
and
Pieper
,
P. G.
(
2005
).
Inheritance of congenital heart disease
.
Neth Heart J
13
,
88
-
91
.
Rotstein
,
B.
and
Paululat
,
A.
(
2016
).
On the morphology of the Drosophila heart
.
Journal of Cardiovascular Development and Disease
3
,
15
.
Roubertie
,
F.
,
Eschalier
,
R.
,
Zemmoura
,
A.
,
Thambo
,
J. B.
,
Rooryck
,
C.
,
Labrousse
,
L.
,
Ploux
,
S.
,
Ritter
,
P.
,
Haïssaguerre
,
M.
,
Dos Santos
,
P.
et al. 
(
2016
).
Cardiac contractility modulation in a model of repaired tetralogy of fallot: a sheep model
.
Pediatr. Cardiol.
37
,
826
-
833
.
Sandireddy
,
R.
,
Cibi
,
D. M.
,
Gupta
,
P.
,
Singh
,
A.
,
Tee
,
N.
,
Uemura
,
A.
,
Epstein
,
J. A.
and
Singh
,
M. K.
(
2019
).
Semaphorin 3E/PlexinD1 signaling is required for cardiac ventricular compaction
.
JCI Insight
4
,
e125908
.
Santos
,
J. M. A.
,
Mendes-Silva
,
L.
,
Afonso
,
V.
,
Martins
,
G.
,
Machado
,
R. S. R.
,
Lopes
,
J. A.
,
Cancela
,
L.
,
Futschik
,
M. E.
,
Sachinidis
,
A.
,
Gavaia
,
P.
et al. 
(
2019
).
Exogenous WNT5A and WNT11 proteins rescue CITED2 dysfunction in mouse embryonic stem cells and zebrafish morphants
.
Cell Death Dis
10
,
582
.
Saraf
,
A.
,
Book
,
W. M.
,
Nelson
,
T. J.
and
Xu
,
C.
(
2019
).
Hypoplastic left heart syndrome: from bedside to bench and back
.
J. Mol. Cell. Cardiol.
135
,
109
-
118
.
Sati
,
S.
and
Cavalli
,
G.
(
2017
).
Chromosome conformation capture technologies and their impact in understanding genome function
.
Chromosoma
126
,
33
-
44
.
Sato
,
M.
and
Yost
,
H. J.
(
2003
).
Cardiac neural crest contributes to cardiomyogenesis in zebrafish
.
Dev. Biol.
257
,
127
-
139
.
Sauka-Spengler
,
T.
and
Barembaum
,
M.
(
2008
).
Gain- and loss-of-function approaches in the chick embryo
.
Methods Cell Biol.
87
,
237
-
256
.
Schlueter
,
J.
,
Männer
,
J.
and
Brand
,
T.
(
2006
).
BMP is an important regulator of proepicardial identity in the chick embryo
.
Dev. Biol.
295
,
546
-
558
.
Schmidt
,
M.
,
Jaeggi
,
E.
,
Ryan
,
G.
,
Hyldebrandt
,
J.
,
Lilly
,
J.
,
Peirone
,
A.
,
Benson
,
L.
and
Chaturvedi
,
R.
(
2008
).
Percutaneous ultrasound-guided stenting of the atrial septum in fetal sheep
.
Ultrasound in Obstet. Gynecol.
32
,
923
-
928
.
Schott
,
J. J.
,
Benson
,
D. W.
,
Basson
,
C. T.
,
Pease
,
W.
,
Silberbach
,
G. M.
,
Moak
,
J. P.
,
Maron
,
B. J.
,
Seidman
,
C. E.
and
Seidman
,
J. G.
(
1998
).
Congenital heart disease caused by mutations in the transcription factor NKX2-5
.
Science
281
,
108
-
111
.
Sedmera
,
D.
,
Reckova
,
M.
,
deAlmeida
,
A.
,
Sedmerova
,
M.
,
Biermann
,
M.
,
Volejnik
,
J.
,
Sarre
,
A.
,
Raddatz
,
E.
,
McCarthy
,
R. A.
,
Gourdie
,
R. G.
et al. 
(
2003
).
Functional and morphological evidence for a ventricular conduction system in zebrafish and Xenopus hearts
.
Am. J. Physiol. Heart Circ. Physiol.
284
,
H1152
-
H1160
.
Sénatore
,
S.
,
Rami Reddy
,
V.
,
Sémériva
,
M.
,
Perrin
,
L.
and
Lalevée
,
N.
(
2010
).
Response to mechanical stress is mediated by the TRPA channel painless in the Drosophila heart
.
PLoS Genet.
6
,
e1001088
.
Seo
,
S.
and
Kume
,
T.
(
2006
).
Forkhead transcription factors, Foxc1 and Foxc2, are required for the morphogenesis of the cardiac outflow tract
.
Dev. Biol.
296
,
421
-
436
.
Serra-Juhé
,
C.
,
Cuscó
,
I.
,
Homs
,
A.
,
Flores
,
R.
,
Torán
,
N.
and
Pérez-Jurado
,
L. A.
(
2015
).
DNA methylation abnormalities in congenital heart disease
.
Epigenetics
10
,
167
-
177
.
Shavik
,
S. M.
,
Jiang
,
Z.
,
Baek
,
S.
and
Lee
,
L. C.
(
2018
).
High spatial resolution multi-organ finite element modeling of ventricular-arterial coupling
.
Front. Physiol.
9
,
119
.
Sheng
,
W.
,
Qian
,
Y.
,
Wang
,
H.
,
Ma
,
X.
,
Zhang
,
P.
,
Chen
,
L.
,
Ma
,
D.
and
Huang
,
G.
(
2013a
).
Association between mRNA levels of DNMT1, DNMT3A, DNMT3B, MBD2 and LINE-1 methylation status in infants with tetralogy of Fallot
.
Int. J. Mol. Med.
32
,
694
-
702
.
Sheng
,
W.
,
Qian
,
Y.
,
Wang
,
H.
,
Ma
,
X.
,
Zhang
,
P.
,
Diao
,
L.
,
An
,
Q.
,
Chen
,
L.
,
Ma
,
D.
and
Huang
,
G.
(
2013b
).
DNA methylation status of NKX2-5, GATA4 and HAND1 in patients with tetralogy of fallot
.
BMC Med. Genomics
6
,
46
.
Sheng
,
W.
,
Wang
,
H.
,
Ma
,
X.
,
Qian
,
Y.
,
Zhang
,
P.
,
Wu
,
Y.
,
Zheng
,
F.
,
Chen
,
L.
,
Huang
,
G.
and
Ma
,
D.
(
2012
).
LINE-1 methylation status and its association with tetralogy of fallot in infants
.
BMC Med. Genomics
5
,
20
.
Sheng
,
W.
,
Qian
,
Y.
,
Zhang
,
P.
,
Wu
,
Y.
,
Wang
,
H.
,
Ma
,
X.
,
Chen
,
L.
,
Ma
,
D.
and
Huang
,
G.
(
2014
).
Association of promoter methylation statuses of congenital heart defect candidate genes with Tetralogy of Fallot
.
J. Transl. Med.
12
,
31
.
Sheng
,
W.
,
Chen
,
L.
,
Wang
,
H.
,
Ma
,
X.
,
Ma
,
D.
and
Huang
,
G.
(
2016
).
CpG island shore methylation of ZFPM2 is identified in tetralogy of fallot samples
.
Pediatr. Res.
80
,
151
-
158
.
Shi
,
Y.
,
Lawford
,
P.
and
Hose
,
R.
(
2011
).
Review of zero-D and 1-D models of blood flow in the cardiovascular system
.
Biomed. Eng. Online
10
,
33
.
Shi
,
H.
,
O'Reilly
,
V. C.
,
Moreau
,
J. L.
,
Bewes
,
T. R.
,
Yam
,
M. X.
,
Chapman
,
B. E.
,
Grieve
,
S. M.
,
Stocker
,
R.
,
Graham
,
R. M.
,
Chapman
,
G.
et al. 
(
2016
).
Gestational stress induces the unfolded protein response, resulting in heart defects
.
Development
143
,
2561
-
2572
.
Siebert
,
J. R.
,
Graham
,
J. M.
, Jr
and
MacDonald
,
C.
(
1985
).
Pathologic features of the CHARGE association: support for involvement of the neural crest
.
Teratology
31
,
331
-
336
.
Singh
,
M. K.
,
Christoffels
,
V. M.
,
Dias
,
J. M.
,
Trowe
,
M. O.
,
Petry
,
M.
,
Schuster-Gossler
,
K.
,
Bürger
,
A.
,
Ericson
,
J.
and
Kispert
,
A.
(
2005
).
Tbx20 is essential for cardiac chamber differentiation and repression of Tbx2
.
Development
132
,
2697
-
2707
.
Smith
,
S. J.
,
Ataliotis
,
P.
,
Kotecha
,
S.
,
Towers
,
N.
,
Sparrow
,
D. B.
and
Mohun
,
T. J.
(
2005
).
The MLC1v gene provides a transgenic marker of myocardium formation within developing chambers of the Xenopus heart
.
Dev. Dyn.
232
,
1003
-
1012
.
Solloway
,
M. J.
and
Robertson
,
E. J.
(
1999
).
Early embryonic lethality in Bmp5;Bmp7 double mutant mice suggests functional redundancy within the 60A subgroup
.
Development
126
,
1753
-
1768
.
Song
,
L.
,
Yan
,
W.
,
Chen
,
X.
,
Deng
,
C. X.
,
Wang
,
Q.
and
Jiao
,
K.
(
2007
).
Myocardial smad4 is essential for cardiogenesis in mouse embryos
.
Circ. Res.
101
,
277
-
285
.
Sparrow
,
D. B.
,
Cai
,
C.
,
Kotecha
,
S.
,
Latinkic
,
B.
,
Cooper
,
B.
,
Towers
,
N.
,
Evans
,
S. M.
and
Mohun
,
T. J.
(
2000
).
Regulation of the tinman homologues in Xenopus embryos
.
Dev. Biol.
227
,
65
-
79
.
Stankunas
,
K.
,
Hang
,
C. T.
,
Tsun
,
Z. Y.
,
Chen
,
H.
,
Lee
,
N. V.
,
Wu
,
J. I.
,
Shang
,
C.
,
Bayle
,
J. H.
,
Shou
,
W.
,
Iruela-Arispe
,
M. L.
et al. 
(
2008
).
Endocardial Brg1 represses ADAMTS1 to maintain the microenvironment for myocardial morphogenesis
.
Dev. Cell
14
,
298
-
311
.
Stennard
,
F. A.
and
Harvey
,
R. P.
(
2005
).
T-box transcription factors and their roles in regulatory hierarchies in the developing heart
.
Development
132
,
4897
-
4910
.
Streckfuss-Bömeke
,
K.
,
Wolf
,
F.
,
Azizian
,
A.
,
Stauske
,
M.
,
Tiburcy
,
M.
,
Wagner
,
S.
,
Hübscher
,
D.
,
Dressel
,
R.
,
Chen
,
S.
,
Jende
,
J.
et al. 
(
2013
).
Comparative study of human-induced pluripotent stem cells derived from bone marrow cells, hair keratinocytes, and skin fibroblasts
.
Eur. Heart J.
34
,
2618
-
2629
.
Su
,
D.
,
Li
,
Q.
,
Guan
,
L.
,
Gao
,
X.
,
Zhang
,
H.
,
Dandan
,
E.
,
Zhang
,
L.
and
Ma
,
X.
(
2013
).
Down-regulation of EBAF in the heart with ventricular septal defects and its regulation by histone acetyltransferase p300 and transcription factors smad2 and cited2
.
Biochim. Biophys. Acta
1832
,
2145
-
2152
.
Suzuki
,
Y.
,
Yeung
,
A. C.
and
Ikeno
,
F.
(
2009
).
The pre-clinical animal model in the translational research of interventional cardiology
.
JACC Cardiovasc Interv
2
,
373
-
383
.
Svensson
,
E. C.
,
Huggins
,
G. S.
,
Dardik
,
F. B.
,
Polk
,
C. E.
and
Leiden
,
J. M.
(
2000
).
A functionally conserved N-terminal domain of the friend of GATA-2 (FOG-2) protein represses GATA4-dependent transcription
.
J. Biol. Chem.
275
,
20762
-
20769
.
Syrmou
,
A.
,
Tzetis
,
M.
,
Fryssira
,
H.
,
Kosma
,
K.
,
Oikonomakis
,
V.
,
Giannikou
,
K.
,
Makrythanasis
,
P.
,
Kitsiou-Tzeli
,
S.
and
Kanavakis
,
E.
(
2013
).
Array comparative genomic hybridization as a clinical diagnostic tool in syndromic and nonsyndromic congenital heart disease
.
Pediatr. Res.
73
,
772
-
776
.
Takahashi
,
K.
,
Tanabe
,
K.
,
Ohnuki
,
M.
,
Narita
,
M.
,
Ichisaka
,
T.
,
Tomoda
,
K.
and
Yamanaka
,
S.
(
2007
).
Induction of pluripotent stem cells from adult human fibroblasts by defined factors
.
Cell
131
,
861
-
872
.
Takeuchi
,
D.
,
Nakanishi
,
T.
,
Tomimatsu
,
H.
and
Nakazawa
,
M.
(
2006
).
Evaluation of right ventricular performance long after the atrial switch operation for transposition of the great arteries using the Doppler Tei index
.
Pediatr. Cardiol.
27
,
78
-
83
.
Tanaka
,
M.
,
Schinke
,
M.
,
Liao
,
H. S.
,
Yamasaki
,
N.
and
Izumo
,
S.
(
2001
).
Nkx2.5 and Nkx2.6, homologs of Drosophila tinman, are required for development of the pharynx
.
Mol. Cell. Biol.
21
,
4391
-
4398
.
Tchervenkov
,
C. I.
,
Jacobs
,
J. P.
,
Weinberg
,
P. M.
,
Aiello
,
V. D.
,
Beland
,
M. J.
,
Colan
,
S. D.
,
Elliott
,
M. J.
,
Franklin
,
R. C.
,
Gaynor
,
J. W.
,
Krogmann
,
O. N.
et al. 
(
2006
).
The nomenclature, definition and classification of hypoplastic left heart syndrome
.
Cardiol. Young
16
,
339
-
368
.
Tessadori
,
F.
,
Roessler
,
H. I.
,
Savelberg
,
S. M. C.
,
Chocron
,
S.
,
Kamel
,
S. M.
,
Duran
,
K. J.
,
van Haelst
,
M. M.
,
van Haaften
,
G.
and
Bakkers
,
J.
(
2018
).
Effective CRISPR/Cas9-based nucleotide editing in zebrafish to model human genetic cardiovascular disorders
.
Disease models & mechanisms
11
,
dmm035469
.
Tevosian
,
S. G.
,
Deconinck
,
A. E.
,
Tanaka
,
M.
,
Schinke
,
M.
,
Litovsky
,
S. H.
,
Izumo
,
S.
,
Fujiwara
,
Y.
and
Orkin
,
S. H.
(
2000
).
FOG-2, a cofactor for GATA transcription factors, is essential for heart morphogenesis and development of coronary vessels from epicardium
.
Cell
101
,
729
-
739
.
Theis
,
J. L.
,
Hrstka
,
S. C.
,
Evans
,
J. M.
,
O'Byrne
,
M. M.
,
de Andrade
,
M.
,
O'Leary
,
P. W.
,
Nelson
,
T. J.
and
Olson
,
T. M.
(
2015
).
Compound heterozygous NOTCH1 mutations underlie impaired cardiogenesis in a patient with hypoplastic left heart syndrome
.
Hum. Genet.
134
,
1003
-
1011
.
Thienpont
,
B.
,
Mertens
,
L.
,
de Ravel
,
T.
,
Eyskens
,
B.
,
Boshoff
,
D.
,
Maas
,
N.
,
Fryns
,
J. P.
,
Gewillig
,
M.
,
Vermeesch
,
J. R.
and
Devriendt
,
K.
(
2007
).
Submicroscopic chromosomal imbalances detected by array-CGH are a frequent cause of congenital heart defects in selected patients
.
Eur. Heart J.
28
,
2778
-
2784
.
Tirosh-Finkel
,
L.
,
Zeisel
,
A.
,
Brodt-Ivenshitz
,
M.
,
Shamai
,
A.
,
Yao
,
Z.
,
Seger
,
R.
,
Domany
,
E.
and
Tzahor
,
E.
(
2010
).
BMP-mediated inhibition of FGF signaling promotes cardiomyocyte differentiation of anterior heart field progenitors
.
Development
137
,
2989
-
3000
.
Tobita
,
K.
and
Keller
,
B. B.
(
2000
).
Right and left ventricular wall deformation patterns in normal and left heart hypoplasia chick embryos
.
Am. J. Physiol. Heart Circ. Physiol.
279
,
H959
-
H969
.
Trusty
,
P. M.
,
Slesnick
,
T. C.
,
Wei
,
Z. A.
,
Rossignac
,
J.
,
Kanter
,
K. R.
,
Fogel
,
M. A.
and
Yoganathan
,
A. P.
(
2018
).
Fontan Surgical Planning: Previous Accomplishments, Current Challenges, and Future Directions
.
J. Cardiovasc. Transl. Res.
11
,
133
-
144
.
Tu
,
C. T.
,
Yang
,
T. C.
and
Tsai
,
H. J.
(
2009
).
Nkx2.7 and Nkx2.5 function redundantly and are required for cardiac morphogenesis of zebrafish embryos
.
PLoS ONE
4
,
e4249
.
Tulzer
,
A.
,
Arzt
,
W.
,
Gitter
,
R.
,
Prandstetter
,
C.
,
Grohmann
,
E.
,
Mair
,
R.
and
Tulzer
,
G.
(
2018
).
Immediate effects and outcome of in-utero pulmonary valvuloplasty in fetuses with pulmonary atresia with intact ventricular septum or critical pulmonary stenosis
.
Ultrasound Obstet. Gynecol.
52
,
230
-
237
.
Tutarel
,
O.
,
Norozi
,
K.
,
Hornung
,
O.
,
Orhan
,
G.
,
Wübbolt-Lehmann
,
P.
,
Wessel
,
A.
and
Yelbuz
,
T. M.
(
2005
).
Images in cardiovascular medicine. Cardiac failure in the chick embryo resembles heart failure in humans
.
Circulation
112
,
e352
-
e353
.
Valdeomillos
,
E.
,
Jalal
,
Z.
,
Metras
,
A.
,
Roubertie
,
F.
,
Benoist
,
D.
,
Bernus
,
O.
,
Haïssaguerre
,
M.
,
Bordachar
,
P.
,
Iriart
,
X.
and
Thambo
,
J. B.
(
2019
).
Animal models of repaired tetralogy of fallot: current applications and future perspectives
.
Can. J. Cardiol.
35
,
1762
-
1771
.
van den Berg
,
C. W.
,
Elliott
,
D. A.
,
Braam
,
S. R.
,
Mummery
,
C. L.
and
Davis
,
R. P.
(
2016
).
Differentiation of human pluripotent stem cells to cardiomyocytes under defined conditions
. In
Patient-Specific Induced Pluripotent Stem Cell Models: Generation and Characterization
(ed.
A.
Nagy
and
K.
Turksen
), vol. 1353, pp.
163
-
180
.
New York, NY
:
Springer New York
.
van den Heuvel
,
N. H.
,
van Veen
,
T. A.
,
Lim
,
B.
and
Jonsson
,
M. K.
(
2014
).
Lessons from the heart: mirroring electrophysiological characteristics during cardiac development to in vitro differentiation of stem cell derived cardiomyocytes
.
J. Mol. Cell. Cardiol.
67
,
12
-
25
.
van den Wijngaard
,
J. P.
,
Westerhof
,
B. E.
,
Faber
,
D. J.
,
Ramsay
,
M. M.
,
Westerhof
,
N.
and
Van Gemert
,
M. J.
(
2006
).
Abnormal arterial flows by a distributed model of the fetal circulation
.
Am. J. Physiol. Regul. Integr. Comp. Physiol.
291
,
R1222
-
R1233
.
van der Bom
,
T.
,
Zomer
,
A. C.
,
Zwinderman
,
A. H.
,
Meijboom
,
F. J.
,
Bouma
,
B. J.
and
Mulder
,
B. J. M.
(
2011
).
The changing epidemiology of congenital heart disease
.
Nat. Rev. Cardiol.
8
,
50
-
60
.
Van Puyvelde
,
J.
,
Rega
,
F.
,
Minami
,
T.
,
Claus
,
P.
,
Cools
,
B.
,
Gewillig
,
M.
and
Meyns
,
B.
(
2019
).
Creation of the Fontan circulation in sheep: a survival model
.
Interact. Cardiovasc. Thorac. Surg.
29
,
15
-
21
.
VanOudenhove
,
J.
,
Yankee
,
T.
,
Wilderman
,
A.
and
Cotney
,
J.
(
2020
).
Epigenomic and transcriptomic dynamics during human heart organogenesis
.
Circ. Res.
127
,
e184
-
e209
.
Veerman
,
C. C.
,
Kosmidis
,
G.
,
Mummery
,
C. L.
,
Casini
,
S.
,
Verkerk
,
A. O.
and
Bellin
,
M.
(
2015
).
Immaturity of human stem-cell-derived cardiomyocytes in culture: fatal flaw or soluble problem?
Stem Cells Dev.
24
,
1035
-
1052
.
Vogler
,
G.
and
Bodmer
,
R.
(
2015
).
Cellular mechanisms of Drosophila heart morphogenesis
.
J. Cardiovasc. Dev. Dis.
2
,
2
-
16
.
von Elten
,
K.
,
Sawyer
,
T.
,
Lentz-Kapua
,
S.
,
Kanis
,
A.
and
Studer
,
M.
(
2013
).
A case of Wolf-Hirschhorn syndrome and hypoplastic left heart syndrome
.
Pediatr. Cardiol.
34
,
1244
-
1246
.
Wang
,
F.
,
Fisher
,
S. A.
,
Zhong
,
J.
,
Wu
,
Y.
and
Yang
,
P.
(
2015
).
Superoxide Dismutase 1 In vivo Ameliorates Maternal Diabetes Mellitus-Induced Apoptosis and Heart Defects Through Restoration of Impaired Wnt Signaling
.
Circ. Cardiovasc. Genet.
8
,
665
-
676
.
Wang
,
J.
,
Greene
,
S. B.
and
Martin
,
J. F.
(
2011
).
BMP signaling in congenital heart disease: new developments and future directions
.
Birth Defects Res. A Clin. Mol. Teratol.
91
,
441
-
448
.
Wang
,
J.
,
Xu
,
N.
,
Feng
,
X.
,
Hou
,
N.
,
Zhang
,
J.
,
Cheng
,
X.
,
Chen
,
Y.
,
Zhang
,
Y.
and
Yang
,
X.
(
2005
).
Targeted disruption of Smad4 in cardiomyocytes results in cardiac hypertrophy and heart failure
.
Circ. Res.
97
,
821
-
828
.
Wang
,
L.
,
Li
,
Z.
,
Song
,
X.
,
Liu
,
L.
,
Su
,
G.
and
Cui
,
Y.
(
2016a
).
Bioinformatic Analysis of Genes and MicroRNAs Associated With Atrioventricular Septal Defect in Down Syndrome Patients
.
Int. Heart J.
57
,
490
-
495
.
Wang
,
Y.
,
Du
,
X.
,
Zhou
,
Z.
,
Jiang
,
J.
,
Zhang
,
Z.
,
Ye
,
L.
and
Hong
,
H.
(
2016b
).
A gain-of-function ACTC1 3'UTR mutation that introduces a miR-139-5p target site may be associated with a dominant familial atrial septal defect
.
Sci. Rep.
6
,
25404
.
Weil
,
S. R.
,
Russo
,
P. A.
,
Heckman
,
J. L.
,
Balsara
,
R. K.
,
Pasiecki
,
V.
and
Dunn
,
J. M.
(
1993
).
Pressure-volume relationship of the fetal lamb heart
.
Ann. Thorac. Surg.
55
,
470
-
475
.
Wentzel
,
P.
,
Gäreskog
,
M.
and
Eriksson
,
U. J.
(
2008
).
Decreased cardiac glutathione peroxidase levels and enhanced mandibular apoptosis in malformed embryos of diabetic rats
.
Diabetes
57
,
3344
-
3352
.
Wessels
,
A.
and
Sedmera
,
D.
(
2003
).
Developmental anatomy of the heart: a tale of mice and man
.
Physiol. Genomics
15
,
165
-
176
.
Wijnands
,
K. P.
,
Chen
,
J.
,
Liang
,
L.
,
Verbiest
,
M. M.
,
Lin
,
X.
,
Helbing
,
W. A.
,
Gittenberger-de Groot
,
A. C.
,
van der Spek
,
P. J.
,
Uitterlinden
,
A. G.
and
Steegers-Theunissen
,
R. P.
(
2017
).
Genome-wide methylation analysis identifies novel CpG loci for perimembranous ventricular septal defects in human
.
Epigenomics
9
,
241
-
251
.
Winnier
,
G. E.
,
Kume
,
T.
,
Deng
,
K.
,
Rogers
,
R.
,
Bundy
,
J.
,
Raines
,
C.
,
Walter
,
M. A.
,
Hogan
,
B. L.
and
Conway
,
S. J.
(
1999
).
Roles for the winged helix transcription factors MF1 and MFH1 in cardiovascular development revealed by nonallelic noncomplementation of null alleles
.
Dev. Biol.
213
,
418
-
431
.
Wiputra
,
H.
,
Lai
,
C. Q.
,
Lim
,
G. L.
,
Heng
,
J. J.
,
Guo
,
L.
,
Soomar
,
S. M.
,
Leo
,
H. L.
,
Biwas
,
A.
,
Mattar
,
C. N.
and
Yap
,
C. H.
(
2016
).
Fluid mechanics of human fetal right ventricles from image-based computational fluid dynamics using 4D clinical ultrasound scans
.
Am. J. Physiol. Heart Circ. Physiol.
311
,
H1498
-
H1508
.
Wiputra
,
H.
,
Lim
,
G. L.
,
Chua
,
K. C.
,
Nivetha
,
R.
,
Soomar
,
S. M.
,
Biwas
,
A.
,
Mattar
,
C. N. Z.
,
Leo
,
H. L.
and
Yap
,
C. H.
(
2017
).
Peristaltic-like motion of the human fetal right ventricle and its effects on fluid dynamics and energy dynamics
.
Ann. Biomed. Eng.
45
,
2335
-
2347
.
Wiputra
,
H.
,
Chen
,
C. K.
,
Talbi
,
E.
,
Lim
,
G. L.
,
Soomar
,
S. M.
,
Biswas
,
A.
,
Mattar
,
C. N. Z.
,
Bark
,
D.
,
Leo
,
H. L.
and
Yap
,
C. H.
(
2018
).
Human fetal hearts with tetralogy of Fallot have altered fluid dynamics and forces
.
Am. J. Physiol. Heart Circ. Physiol.
315
,
H1649
-
H1659
.
Witman
,
N.
,
Zhou
,
C.
,
Grote Beverborg
,
N.
,
Sahara
,
M.
and
Chien
,
K. R.
(
2020
).
Cardiac progenitors and paracrine mediators in cardiogenesis and heart regeneration
.
Semin. Cell Dev. Biol.
100
,
29
-
51
.
Wu
,
Y.
,
Ma
,
X. J.
,
Wang
,
H. J.
,
Li
,
W. C.
,
Chen
,
L.
,
Ma
,
D.
and
Huang
,
G. Y.
(
2014
).
Expression of Cx43-related microRNAs in patients with tetralogy of Fallot
.
World J. Pediatr.
10
,
138
-
144
.
Wu
,
W.
,
Pott
,
D.
,
Mazza
,
B.
,
Sironi
,
T.
,
Dordoni
,
E.
,
Chiastra
,
C.
,
Petrini
,
L.
,
Pennati
,
G.
,
Dubini
,
G.
and
Steinseifer
,
U.
(
2016
).
Fluid–structure interaction model of a percutaneous aortic valve: comparison with an in vitro test and feasibility study in a patient-specific case
.
Ann. Biomed. Eng.
44
,
590
-
603
.
Wünnemann
,
F.
,
Ta-Shma
,
A.
,
Preuss
,
C.
,
Leclerc
,
S.
,
van Vliet
,
P. P.
,
Oneglia
,
A.
,
Thibeault
,
M.
,
Nordquist
,
E.
,
Lincoln
,
J.
,
Scharfenberg
,
F.
et al. 
(
2020
).
Loss of ADAMTS19 causes progressive non-syndromic heart valve disease
.
Nat. Genet.
52
,
40
-
47
.
Xu
,
J.
,
Hu
,
Z.
,
Xu
,
Z.
,
Gu
,
H.
,
Yi
,
L.
,
Cao
,
H.
,
Chen
,
J.
,
Tian
,
T.
,
Liang
,
J.
,
Lin
,
Y.
et al. 
(
2009
).
Functional variant in microRNA-196a2 contributes to the susceptibility of congenital heart disease in a Chinese population
.
Hum. Mutat.
30
,
1231
-
1236
.
Yaakobovich
,
H.
,
Mayo
,
R. P.
,
Zaretsky
,
U.
,
Finkelstein
,
A.
and
Marom
,
G.
(
2020
).
Numerical models of valve-in-valve implantation: effect of intentional leaflet laceration on the anchorage
.
Biomech. Model. Mechanobiol.
19
,
415
-
426
.
Yagi
,
H.
,
Liu
,
X.
,
Gabriel
,
G. C.
,
Wu
,
Y.
,
Peterson
,
K.
,
Murray
,
S. A.
,
Aronow
,
B. J.
,
Martin
,
L. J.
,
Benson
,
D. W.
and
Lo
,
C. W.
(
2018
).
The Genetic Landscape of Hypoplastic Left Heart Syndrome
.
Pediatr. Cardiol.
39
,
1069
-
1081
.
Yang
,
Q.
,
Chen
,
H.
,
Correa
,
A.
,
Devine
,
O.
,
Mathews
,
T. J.
and
Honein
,
M. A.
(
2006
).
Racial differences in infant mortality attributable to birth defects in the United States, 1989-2002
.
Birth Defects Res. A Clin. Mol. Teratol.
76
,
706
-
713
.
Yang
,
X.
,
Pabon
,
L.
and
Murry
,
C. E.
(
2014
).
Engineering adolescence: maturation of human pluripotent stem cell-derived cardiomyocytes
.
Circ. Res.
114
,
511
-
523
.
Yang
,
P.
,
Chen
,
X.
,
Kaushal
,
S.
,
Reece
,
E. A.
and
Yang
,
P.
(
2016
).
High glucose suppresses embryonic stem cell differentiation into cardiomyocytes : High glucose inhibits ES cell cardiogenesis
.
Stem Cell Res. Ther.
7
,
187
.
Yang
,
X.
,
Kong
,
Q.
,
Li
,
Z.
,
Xu
,
M.
,
Cai
,
Z.
and
Zhao
,
C.
(
2018
).
Association between the promoter methylation of the TBX20 gene and tetralogy of fallot
.
Scand. Cardiovasc. J.
52
,
287
-
291
.
Ye
,
J.
,
Tucker
,
N. R.
,
Weng
,
L. C.
,
Clauss
,
S.
,
Lubitz
,
S. A.
and
Ellinor
,
P. T.
(
2016
).
A Functional Variant Associated with Atrial Fibrillation Regulates PITX2c Expression through TFAP2a
.
Am. J. Hum. Genet.
99
,
1281
-
1291
.
Yelbuz
,
T. M.
,
Waldo
,
K. L.
,
Kumiski
,
D. H.
,
Stadt
,
H. A.
,
Wolfe
,
R. R.
,
Leatherbury
,
L.
and
Kirby
,
M. L.
(
2002
).
Shortened outflow tract leads to altered cardiac looping after neural crest ablation
.
Circulation
106
,
504
-
510
.
Yoshihara
,
M.
,
Hayashizaki
,
Y.
and
Murakawa
,
Y.
(
2017
).
Genomic instability of iPSCs: challenges towards their clinical applications
.
Stem Cell Rev. Rep.
13
,
7
-
16
.
Zaffran
,
S.
,
Kelly
,
R. G.
,
Meilhac
,
S. M.
,
Buckingham
,
M. E.
and
Brown
,
N. A.
(
2004
).
Right ventricular myocardium derives from the anterior heart field
.
Circ. Res.
95
,
261
-
268
.
Zaidi
,
S.
and
Brueckner
,
M.
(
2017
).
Genetics and genomics of congenital heart disease
.
Circ. Res.
120
,
923
-
940
.
Zaidi
,
S.
,
Choi
,
M.
,
Wakimoto
,
H.
,
Ma
,
L.
,
Jiang
,
J.
,
Overton
,
J. D.
,
Romano-Adesman
,
A.
,
Bjornson
,
R. D.
,
Breitbart
,
R. E.
,
Brown
,
K. K.
et al. 
(
2013
).
De novo mutations in histone-modifying genes in congenital heart disease
.
Nature
498
,
220
-
223
.
Zhang
,
Q.-J.
and
Liu
,
Z.-P.
(
2015
).
Histone methylations in heart development, congenital and adult heart diseases
.
Epigenomics
7
,
321
-
330
.
Zhang
,
H.
and
Bradley
,
A.
(
1996
).
Mice deficient for BMP2 are nonviable and have defects in amnion/chorion and cardiac development
.
Development
122
,
2977
-
2986
.
Zhang
,
J.
,
Wilson
,
G. F.
,
Soerens
,
A. G.
,
Koonce
,
C. H.
,
Yu
,
J.
,
Palecek
,
S. P.
,
Thomson
,
J. A.
and
Kamp
,
T. J.
(
2009
).
Functional cardiomyocytes derived from human induced pluripotent stem cells
.
Circ. Res.
104
,
e30
-
e41
.
Zhao
,
M.
,
Tang
,
Y.
,
Zhou
,
Y.
and
Zhang
,
J.
(
2019
).
Deciphering role of Wnt signalling in cardiac mesoderm and cardiomyocyte differentiation from human iPSCs: Four-dimensional control of Wnt pathway for hiPSC-CMs differentiation
.
Sci. Rep.
9
,
19389
.
Zhu
,
H.
,
Kartiko
,
S.
and
Finnell
,
R. H.
(
2009
).
Importance of gene-environment interactions in the etiology of selected birth defects
.
Clin. Genet.
75
,
409
-
423
.
Zhu
,
C.
,
Yu
,
Z. B.
,
Chen
,
X. H.
,
Ji
,
C. B.
,
Qian
,
L. M.
and
Han
,
S. P.
(
2011a
).
DNA hypermethylation of the NOX5 gene in fetal ventricular septal defect
.
Exp. Ther. Med.
2
,
1011
-
1015
.
Zhu
,
C.
,
Yu
,
Z. B.
,
Chen
,
X. H.
,
Pan
,
Y.
,
Dong
,
X. Y.
,
Qian
,
L. M.
and
Han
,
S. P.
(
2011b
).
Screening for differential methylation status in fetal myocardial tissue samples with ventricular septal defects by promoter methylation microarrays
.
Mol. Med. Rep.
4
,
137
-
143
.
Zwi
,
L.
,
Caspi
,
O.
,
Arbel
,
G.
,
Huber
,
I.
,
Gepstein
,
A.
,
Park
,
I. H.
and
Gepstein
,
L.
(
2009
).
Cardiomyocyte differentiation of human induced pluripotent stem cells
.
Circulation
120
,
1513
-
1523
.

Competing interests

The authors declare no competing or financial interests.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution and reproduction in any medium provided that the original work is properly attributed.