Open access peer-reviewed chapter

Animal Models of Cardiomyopathies

Written By

Enkhsaikhan Purevjav

Submitted: 26 March 2019 Reviewed: 06 August 2019 Published: 15 October 2019

DOI: 10.5772/intechopen.89033

From the Edited Volume

Animal Models in Medicine and Biology

Edited by Eva Tvrdá and Sarat Chandra Yenisetti

Chapter metrics overview

1,283 Chapter Downloads

View Full Metrics

Abstract

Cardiomyopathies are a heterogeneous group of disorders of heart muscle that ultimately result in congestive heart failure (CHF). Rapid progress in genetics as well as in molecular and cellular biology over the past three decades has greatly improved the understanding of pathogenic signaling pathways in inherited cardiomyopathies. This chapter will focus on animal models of different clinical forms of human cardiomyopathies with their summaries of triggered key molecules, and signaling pathways will be described.

Keywords

  • cardiomyopathy
  • heart failure
  • genetic mutation

1. From genetic abnormality to cardiomyopathy phenotype

It’s widely accepted that inherited cardiomyopathies are a group of heterogeneous diseases of heart muscle resulting from genetic alterations in cardiac myocytes, the chief contractile cell type in the heart [1]. The genes encoding proteins that build muscle cytoskeleton and contractile apparatus are responsible for a cardiomyopathy phenotype with distinctive morpho-/histological cardiac remodeling [2]. Further, disruption of particular genetic and protein networks and pathways may intersect with other intracellular and intercellular pathways and disturbances in molecular signaling. Apoptosis, necrosis, autophagy, and metabolic and arrhythmogenic fluxes—which may present as the sole features or as overlapping signs of decompensated cardiac homeostasis—result in definitive forms of cardiac remodeling including fibrosis, cardiomyocyte hypertrophy, and atrophy. Typically, molecular signaling activates associated compensatory responses and cooperates with other modifiers such as genetic modifiers and environment, stress, or toxicity related that, in turn, may or may not influence the final cardiomyopathy phenotype. Alterations in cellular morphology and size, gene expression patterns, and metabolic shifts in cardiomyocytes initially compensate and maintain cardiac function in the subtle, preclinical stages of cardiomyopathy. Thus, inherited forms of cardiomyopathy, irrespective of the specific genetic or morpho-/clinical condition, may or may not present signs of a failing heart. Five types of inherited cardiomyopathies are distinguished based on clinical features: dilated cardiomyopathy (DCM), hypertrophic cardiomyopathy (HCM), restrictive cardiomyopathy (RCM), arrhythmogenic ventricular cardiomyopathies (ACM), and left ventricular noncompaction cardiomyopathy (LVNC) [3] as demonstrated in Table 1. DCM is characterized by left ventricular (LV) dilation and systolic dysfunction; HCM is characterized by LV hypertrophy with diastolic dysfunction; and RCM is accompanied by increased stiffness of the myocardium and dilated atria due to diastolic dysfunction without significant hypertrophy [4]. Frequent and often life-threatening arrhythmias and associated sudden cardiac death and progressive heart failure are the main hallmarks of ACMs [5], while myocardial hypertrabeculation, intertrabecular recesses, and thin compact LV wall are the characteristics of LVNC [6]. Sustained maladaptive remodeling due to pathologic genetic insult results in the development of decompensated cardiomyopathy when the failing heart is unable to keep up with the hemodynamic demands at all levels, from the molecule to the whole organism. When compensatory mechanisms fail, additional neuroendocrine signaling and other pathways are activated on an organ and whole organism level, leading to CHF. Cellular and molecular level alterations of end-stage cardiomyopathy and CHF respond to irreversible cardiac remodeling with significant changes in membrane ion currents and intracellular Ca2+ metabolism, fibrosis, hypertrophic or atrophic remodeling, and cell death. Cardiac function is significantly depressed with depleted contractile force development and slowed relaxation [7].

Table 1.

Clinical types of inherited cardiomyopathy and specific hallmarks of different types of cardiomyopathy.

Advertisement

2. Animal models of human cardiomyopathies

Translational comparative animal research is of considerable value in inherited cardiomyopathies, because animal models enable to explore and investigate the cellular and molecular pathology originating from the initial genetic assault but also may closely recapitulate the effects of cardiac remodeling culminating into a specific cardiomyopathy type seen in humans. Animal models carrying human gene mutations may not present clinical phenotypic signs of cardiomyopathy resembling the human disease until adulthood, supporting a temporal mechanism by which chronically altered cellular responses and cardiac remodeling lead to the clinically relevant phenotype.

2.1 Naturally occurring animal models of cardiomyopathy

Naturally occurring cardiomyopathy among small and large animals is commonly observed in canine and feline species [8, 9]. HCM is a common disease in pet cats, affecting 10–15% of the pet cat population [10], while DCM is more typical in dogs [11]. The similarity to human HCM or DCM, the rapid progression of disease, and the defined and readily determined endpoints of feline HCM or in canine DCM make them excellent natural models that are genotypically and phenotypically similar to human heart muscle disease [12]. The Maine Coon and Ragdoll cats are particularly valuable models of HCM associated with myosin binding protein C (MyBP-C) mutations and even higher disease incidence compared to the overall feline population [13, 14]. In canine, mutations in genes such as dystrophin (DYST) in German Shorthaired Pointers [15], desmin (DES) and α-actinin in the Doberman [16, 17], titin-cap (TCAP) in Irish Wolfhounds [18], and striatin in Boxers [19] were reported to be associated with DCM. In addition, many naturally occurring porcine HCM and DCM have been described offering the useful models for translational research [20, 21, 22].

2.2 Genetically engineered animal models of cardiomyopathy

Experimentally, numerous small and large animal models including fruit fly, fish, rodents, rabbit, canine, pig, and other species have been developed to discover pathogenetic mechanisms involved in cardiomyopathy in the research field [23, 24, 25]. Characterization of the mechanisms of cardiomyopathies using the study of animal models is challenging owing to the complexity of disease-causing mechanisms and modulators of pathology [25]. Moreover, animal models are successfully used for genome-wide screening, assessing of cardiac phenotypes and disease symptoms, genotype-phenotype association studies, and drug discovery and development assays. The accessibility of transgenic (TG), knockout (KO) and knock-in (KI) murine models has, however, been one of the most successful approaches for studying genetic cardiomyopathies [26]. With recent advances in CRISPR/Cas9 technology, researchers are able to achieve more effective and precise genome editing because of its simplicity, design, and efficiency over other traditional methods for genetic editing such as transgenesis and homologous recombination targeting techniques [27, 28, 29].

The lowest species that has typically been used for cardiomyopathy research is Drosophila melanogaster as a tool to study various developmental biological processes and mechanisms underlying congenital defects and inherited heart diseases [30, 31]. The Drosophila heart looks as a primitive linear tube similar to embryonic heart tube in vertebrates, and many heart development, function, and aging regulatory genes and networks such as NK-2, MEF2, GATA, Tbx, and Hand have been evolutionarily conserved. The conserved development of the heart in simple model organisms and vertebrates provides a unique ability to use many different animal models in cardiomyopathy research [32]. Important advantages of the use of animal models are the ability to manipulate gene expression and identify genes and mechanisms regulating heart development, cardiac pathology, and pathophysiology [33, 34]. Advanced systems to identify genes causing human cardiomyopathies such as UAS/GAL4 [35], techniques for accurate phenotyping of cardiac diseases such as optical coherence tomography [36], powerful electrophysiological, mechanical, and histological approaches to characterize heart development, cardiac tissue properties, and structure in the Drosophila heart have emerged as a pioneering model system in basic, genetic, and molecular studies of cardiac development, function, aging, and disease [37]. Numerous Drosophila models have been used to elucidate the pathophysiology of human HCM and DCM and other heart diseases, such as heart failure, cardiac tachycardia, atrial fibrillation, and congenital heart diseases [38, 39, 40].

The zebra fish (Danio rerio) model remains one of the most effective technologies for discovering and functional studying novel cardiomyopathy candidate genes, especially the ability to use morpholino knockdown techniques in fish models [26, 41, 42]. Compared with other vertebrate models, the zebra fish embryos are transparent allowing genetic engineering approaches to apply fluorescent reporter transgenes with genetic fate mapping strategies combined with high-resolution, high-throughput microscopy imaging in vivo of the heart [43, 44]. The transparency of the embryos allows to observe fluorescent proteins that are expressed in various cell types of the cardiovascular system, and these research advances have opened avenues to improve our knowledge of regulatory mechanisms of cardiomyocyte and other cardiac cells’ differentiation [45, 46], regeneration [44], morphogenesis [47], drug effects and toxicity [48], and gene regulation [49]. The advancement in high-speed video imaging and automated image analysis techniques including light sheet planar illumination microscopy not only allows to precisely monitor morphologic and functional characteristics such as heart rate, arrhythmias, and ejection fraction in zebrafish but also progresses our current understanding of the different types of cardiomyopathy.

Rodent models are the most used model species for cardiomyopathy research, including genetics, pharmacology, and long-term survival considering that rodents have a short gestation time, have the ability to be genetically manipulated to generate transgenic or mutant strains, and are easy to handle and house with low maintenance costs [24, 50]. In addition, a fact that mice have short life span allows investigators to generate genetic models in a shorter time period and follow the natural history of genetic diseases at an accelerated pace, enabling to rapidly launch proof-of-principle experiments and potentially translating and exploiting the results into human studies. Significant advantages to rodents as the species of choice can limit the murine data’s applicability to human cardiovascular function; there are significant differences between the mouse models and human disease presentation [25]. Rodents are phylogenetically farthest distant from humans compared to other mammals, and some pathophysiological features of cardiomyopathy phenotypes and their response to environmental stress and treatments may not be reliable for human diseases [23].

The rabbit and pig experimental models of cardiomyopathy offer significant advantages for cardiovascular research [50]. Compared with the mouse, the larger size and slower heart rate of the rabbit and pigs are advantageous for physiological analyses such as echocardiography and cardiac catheterization.

2.2.1 Hypertrophic cardiomyopathy animal models

Animal models of HCM mostly carry human mutations in sarcomeric protein-encoding genes such as a-MHC, a-tropomyosin, troponins, myosin binding protein C (MyBP-C), and other genes shown in Table 1 [51, 52, 53, 54, 55]. Many models carry cardiac-specific (CS) expression or ablation of the proteins of interest. These models have demonstrated that HCM mutations enhance contractile properties with increased force generation, ATP hydrolysis, and actin-myosin sliding velocity, showing that the hypertrophy is not a compensatory response to diminished contractile function [56, 57, 58]. Models of HCM also show abnormal Ca2+ cycling in cardiomyocytes before overt histopathologic changes occurred in the myocardium and delayed myocardial relaxation that occurs before the onset of hypertrophy, suggesting that diastolic dysfunction is a direct consequence of HCM mutations [59, 60]. Hearts from models of HCM progressively accumulate myocardial fibrosis in the same manner as human patients, and fibrosis is considered to be a cellular substrate for cardiac arrhythmias and sudden cardiac death in humans [61, 62, 63].

2.2.2 Dilated cardiomyopathy animal models

Animal models of DCM mostly resemble human mutations in genes encoding cytoskeletal, sarcomeric, and Z-disk proteins and present with ventricular dilation and thinning of the ventricular walls correlated with loss of heart muscle mass. In addition, functional changes in non-myocytes induce fibrotic scars that stiffen the heart tissue and impede normal cardiomyocyte contractility. Novel DCM mechanisms such as impaired Z-disk assembly, sensitivity to apoptosis and abnormalities in myofibrillogenesis under metabolic stress, protein folding, inhibition of protein aggregation, and degradation of misfolded proteins have been explored (Table 2).

Table 2.

Animal models of hypertrophic cardiomyopathy [51, 52, 53, 54, 55, 56, 57, 58, 60, 61, 62, 63, 64, 65, 66, 67, 68, 69, 70, 71, 72, 73, 74, 75, 76, 77, 78, 79, 80, 81].

2.2.3 Restrictive cardiomyopathy animal models

RCM is the least common but most lethal form of cardiomyopathy where impaired ventricular relaxation due to increased stiffness of the myocardium and pressure in the ventricles overcomes the changes in myofibrillar arrangement and cardiomyocyte gross abnormalities [113]. Animal models carrying human RCM-associated mutations have also been generated to mimic human RCM phenotype. These mutations are identified mainly in sarcomeric protein-encoding genes such as troponins, myosin and MYPN (summarized in Table 3).

Table 3.

Animal models of dilated cardiomyopathy [82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 98, 99, 100, 101, 102, 103, 104, 105, 106, 107, 108, 109, 110, 111, 112].

2.2.4 Arrhythmogenic ventricular cardiomyopathy models

Many models of ARVC with mutations in genes encoding desmosomal (DSP, PKP, DSC, DSG, and JUP) and non-desmosomal (RYR2, TMEM43, and ZASP) proteins have been developed [118]. Structural and functional alterations include progressive, diffuse, or segmental loss of cardiomyocytes, probably due to cardiomyocyte apoptosis or necrosis, and replacement with fibrotic and adipose tissue (Table 4). Fibro-fatty tissue primarily is seen in the right ventricle (RV), with common LV involvement in later stages of the disease [119] (Table 5).

Table 4.

Animal models of restrictive cardiomyopathy [67, 114, 115, 116, 117].

Table 5.

Animal models of arrhythmogenic ventricular cardiomyopathy [119, 120, 121, 122, 123, 124, 125, 126, 127, 128, 129, 130, 131, 132, 133, 134, 135, 136, 137, 138, 139].

2.2.5 Left ventricular noncompaction cardiomyopathy models

Animal models of LVNC typically demonstrate a spongiform ventricular myocardium and deep trabeculations, and many reports suggested that LV trabeculation and compaction processes are two distinct but tightly interconnected morphogenetic events resulting in the development of a functionally proficient ventricular chamber wall [140]. Animal models exhibiting LVNC phenotypes and potential pathogenetic mechanisms are summarized in Table 6.

Table 6.

Animal models of left ventricular noncompaction cardiomyopathy [141, 142, 143, 144, 145, 146, 147, 148, 149, 150, 151, 152, 153, 154, 155, 156, 157, 158, 159].

Advertisement

3. Conclusion

Advances in molecular and genetic techniques have vastly improved the understanding of molecular mechanisms responsible for cardiomyopathies and cardiac dysfunction. The wide range of innovative technologies and techniques used in animal models in vivo has led to advances in our knowledge on the etiology, pathophysiology, and therapeutics of inherited cardiomyopathies. It is clear that mutant proteins in cardiomyocytes can perturb cardiac function whether the prime distress occurs in the contractile apparatus or neighboring cellular complexes, yet persistent cellular stress leads to tissue-, organ-, and organism-level pathology and pathophysiology. However, development and investigation of animal models are complex processes and the outcomes of which could be difficult to translate to humans due to differences in human and animal cardiovascular anatomy and physiology as well as differing pathophysiology of human cardiomyopathies and experimentally induced diseases in animals [160]. Therefore, the choice of appropriate animal model(s) for cardiomyopathy research should utterly rely on clinical knowledge of human cardiovascular diseases, proper research questions, sufficient number of study animals, and correct and relevant interpretation of results and outcomes in animals to human population. Although animal models of human cardiomyopathies often represent incomplete or inaccurate pathological and pathophysiological features seen in humans, the use of animal models not only has improved our knowledge on the etiology and mechanisms of cardiac muscle diseases and therapeutic interventions but also has greatly promoted an advancement in cardiac tissue engineering, induced pluripotent stem cells (iPSCs) technology, in silico and in vitro techniques, and preclinical assessment of drug discovery and development [161].

References

  1. 1. Harvey PA, Leinwand LA. The cell biology of disease: Cellular mechanisms of cardiomyopathy. The Journal of Cell Biology. 2011;194(3):355-365
  2. 2. Bowles NE, Bowles KR, Towbin JA. The “final common pathway” hypothesis and inherited cardiovascular disease. The role of cytoskeletal proteins in dilated cardiomyopathy. Herz. 2000;25(3):168-175
  3. 3. Maron BJ, Towbin JA, Thiene G, Antzelevitch C, Corrado D, Arnett D, et al. Contemporary definitions and classification of the cardiomyopathies: An American heart association scientific statement from the council on clinical cardiology, heart failure and transplantation committee; quality of care and outcomes research and functional genomics and translational biology interdisciplinary working groups; and council on epidemiology and prevention. Circulation. 2006;113(14):1807-1816
  4. 4. Hershberger RE, Cowan J, Morales A, Siegfried JD. Progress with genetic cardiomyopathies: Screening, counseling, and testing in dilated, hypertrophic, and arrhythmogenic right ventricular dysplasia/cardiomyopathy. Circulation. Heart Failure. 2009;2(3):253-261
  5. 5. Marcus FI, McKenna WJ, Sherrill D, Basso C, Bauce B, Bluemke DA, et al. Diagnosis of arrhythmogenic right ventricular cardiomyopathy/dysplasia: Proposed modification of the task force criteria. Circulation. 2010;121(13):1533-1541
  6. 6. Towbin JA. Inherited cardiomyopathies. Circulation Journal. 2014;78(10):2347-2356
  7. 7. Towbin JA, Bowles NE. The failing heart. Nature. 2002;415(6868):227-233
  8. 8. Martin MW, Stafford Johnson MJ, Strehlau G, King JN. Canine dilated cardiomyopathy: A retrospective study of prognostic findings in 367 clinical cases. The Journal of Small Animal Practice. 2010;51(8):428-436
  9. 9. Maron BJ, Fox PR. Hypertrophic cardiomyopathy in man and cats. Journal of Veterinary Cardiology. 2015;17(Suppl 1):S6-S9
  10. 10. Kittleson MD, Meurs KM, Harris SP. The genetic basis of hypertrophic cardiomyopathy in cats and humans. Journal of Veterinary Cardiology. 2015;17(Suppl 1):S53-S73
  11. 11. Simpson S, Edwards J, Ferguson-Mignan TF, Cobb M, Mongan NP, Rutland CS. Genetics of human and canine dilated cardiomyopathy. International Journal of Genomics. 2015;2015:204823
  12. 12. Liu SK, Roberts WC, Maron BJ. Comparison of morphologic findings in spontaneously occurring hypertrophic cardiomyopathy in humans, cats and dogs. The American Journal of Cardiology. 1993;72(12):944-951
  13. 13. Freeman LM, Rush JE, Stern JA, Huggins GS, Maron MS. Feline hypertrophic cardiomyopathy: A spontaneous large animal model of human HCM. Cardiology Research. 2017;8(4):139-142
  14. 14. Longeri M, Ferrari P, Knafelz P, Mezzelani A, Marabotti A, Milanesi L, et al. Myosin-binding protein C DNA variants in domestic cats (A31P, A74T, R820W) and their association with hypertrophic cardiomyopathy. Journal of Veterinary Internal Medicine. 2013;27(2):275-285
  15. 15. Nghiem PP, Hoffman EP, Mittal P, Brown KJ, Schatzberg SJ, Ghimbovschi S, et al. Sparing of the dystrophin-deficient cranial sartorius muscle is associated with classical and novel hypertrophy pathways in GRMD dogs. The American Journal of Pathology. 2013;183(5):1411-1424
  16. 16. Stabej P, Imholz S, Versteeg SA, Zijlstra C, Stokhof AA, Domanjko-Petric A, et al. Characterization of the canine desmin (DES) gene and evaluation as a candidate gene for dilated cardiomyopathy in the Dobermann. Gene. 2004;340(2):241-249
  17. 17. O’Sullivan ML, O’Grady MR, Pyle WG, Dawson JF. Evaluation of 10 genes encoding cardiac proteins in Doberman pinschers with dilated cardiomyopathy. American Journal of Veterinary Research. 2011;72(7):932-939
  18. 18. Philipp U, Vollmar A, Distl O. Evaluation of the titin-cap gene (TCAP) as candidate for dilated cardiomyopathy in Irish wolfhounds. Animal Biotechnology. 2008;19(4):231-236
  19. 19. Cattanach BM, Dukes-McEwan J, Wotton PR, Stephenson HM, Hamilton RM. A pedigree-based genetic appraisal of boxer ARVC and the role of the Striatin mutation. The Veterinary Record. 2015;176(19):492
  20. 20. Shyu JJ, Cheng CH, Erlandson RA, Lin JH, Liu SK. Ultrastructure of intramural coronary arteries in pigs with hypertrophic cardiomyopathy. Cardiovascular Pathology. 2002;11(2):104-111
  21. 21. Collins DE, Eaton KA, Hoenerhoff MJ. Spontaneous dilated cardiomyopathy and right-sided heart failure as a differential diagnosis for hepatosis dietetica in a production pig. Comparative Medicine. 2015;65(4):327-332
  22. 22. Lin JH, Huang SY, Lee WC, Liu SK, Chu RM. Echocardiographic features of pigs with spontaneous hypertrophic cardiomyopathy. Comparative Medicine. 2002;52(3):238-242
  23. 23. Camacho P, Fan H, Liu Z, He JQ. Small mammalian animal models of heart disease. American Journal of Cardiovascular Disease. 2016;6(3):70-80
  24. 24. Recchia FA, Lionetti V. Animal models of dilated cardiomyopathy for translational research. Veterinary Research Communications. 2007;31(Suppl 1):35-41
  25. 25. Houser SR, Margulies KB, Murphy AM, Spinale FG, Francis GS, Prabhu SD, et al. Animal models of heart failure: A scientific statement from the American heart association. Circulation Research. 2012;111(1):131-150
  26. 26. Duncker DJ, Bakkers J, Brundel BJ, Robbins J, Tardiff JC, Carrier L. Animal and in silico models for the study of sarcomeric cardiomyopathies. Cardiovascular Research. 2015;105(4):439-448
  27. 27. Doudna JA, Charpentier E. Genome editing. The new frontier of genome engineering with CRISPR-Cas9. Science. 2014;346(6213):1258096
  28. 28. Hsu PD, Lander ES, Zhang F. Development and applications of CRISPR-Cas9 for genome engineering. Cell. 2014;157(6):1262-1278
  29. 29. Suzuki K, Tsunekawa Y, Hernandez-Benitez R, Wu J, Zhu J, Kim EJ, et al. In vivo genome editing via CRISPR/Cas9 mediated homology-independent targeted integration. Nature. 2016;540(7631):144-149
  30. 30. Bier E, Bodmer R. Drosophila, an emerging model for cardiac disease. Gene. 2004;342(1):1-11
  31. 31. Piazza N, Wessells RJ. Drosophila models of cardiac disease. Progress in Molecular Biology and Translational Science. 2011;100:155-210
  32. 32. Tao Y, Schulz RA. Heart development in Drosophila. Seminars in Cell & Developmental Biology. 2007;18(1):3-15
  33. 33. Bryantsev AL, Cripps RM. Cardiac gene regulatory networks in Drosophila. Biochimica et Biophysica Acta. 2009;1789(4):343-353
  34. 34. Bodmer R, Venkatesh TV. Heart development in Drosophila and vertebrates: Conservation of molecular mechanisms. Developmental Genetics. 1998;22(3):181-186
  35. 35. Wolf MJ, Amrein H, Izatt JA, Choma MA, Reedy MC, Rockman HA. Drosophila as a model for the identification of genes causing adult human heart disease. Proceedings of the National Academy of Sciences of the United States of America. 2006;103(5):1394-1399
  36. 36. Choma MA, Izatt SD, Wessells RJ, Bodmer R, Izatt JA. Images in cardiovascular medicine: In vivo imaging of the adult Drosophila melanogaster heart with real-time optical coherence tomography. Circulation. 2006;114(2):e35-e36
  37. 37. Ocorr K, Vogler G, Bodmer R Methods to assess Drosophila heart development, function and aging. Methods. 2014;68(1):265-272
  38. 38. Viswanathan MC, Kaushik G, Engler AJ, Lehman W, Cammarato A. A Drosophila melanogaster model of diastolic dysfunction and cardiomyopathy based on impaired troponin-T function. Circulation Research. 2014;114(2):e6-e17
  39. 39. Vu Manh TP, Mokrane M, Georgenthum E, Flavigny J, Carrier L, Semeriva M, et al. Expression of cardiac myosin-binding protein-C (cMyBP-C) in Drosophila as a model for the study of human cardiomyopathies. Human Molecular Genetics. 2005;14(1):7-17
  40. 40. Zhu JY, Fu Y, Nettleton M, Richman A, Han Z. High throughput in vivo functional validation of candidate congenital heart disease genes in Drosophila. eLife. 2017;6. pii: e22617
  41. 41. Dvornikov AV, de Tombe PP, Xu X. Phenotyping cardiomyopathy in adult zebrafish. Progress in Biophysics and Molecular Biology. 2018;138:116-125
  42. 42. Shi X, Chen R, Zhang Y, Yun J, Brand-Arzamendi K, Liu X, et al. Zebrafish heart failure models: Opportunities and challenges. Amino Acids. 2018;50(7):787-798
  43. 43. Holtzman NG, Schoenebeck JJ, Tsai HJ, Yelon D. Endocardium is necessary for cardiomyocyte movement during heart tube assembly. Development. 2007;134(12):2379-2386
  44. 44. Zhang R, Han P, Yang H, Ouyang K, Lee D, Lin YF, et al. In vivo cardiac reprogramming contributes to zebrafish heart regeneration. Nature. 2013;498(7455):497-501
  45. 45. de Pater E, Clijsters L, Marques SR, Lin YF, Garavito-Aguilar ZV, Yelon D, et al. Distinct phases of cardiomyocyte differentiation regulate growth of the zebrafish heart. Development. 2009;136(10):1633-1641
  46. 46. Staudt DW, Liu J, Thorn KS, Stuurman N, Liebling M, Stainier DY. High-resolution imaging of cardiomyocyte behavior reveals two distinct steps in ventricular trabeculation. Development. 2014;141(3):585-593
  47. 47. Auman HJ, Coleman H, Riley HE, Olale F, Tsai HJ, Yelon D. Functional modulation of cardiac form through regionally confined cell shape changes. PLoS Biology. 2007;5(3):e53
  48. 48. Lin KY, Chang WT, Lai YC, Liau I. Toward functional screening of cardioactive and cardiotoxic drugs with zebrafish in vivo using pseudodynamic three-dimensional imaging. Analytical Chemistry. 2014;86(4):2213-2220
  49. 49. Olson EN. Gene regulatory networks in the evolution and development of the heart. Science. 2006;313(5795):1922-1927
  50. 50. Milani-Nejad N, Janssen PM. Small and large animal models in cardiac contraction research: Advantages and disadvantages. Pharmacology & Therapeutics. 2014;141(3):235-249
  51. 51. Geisterfer-Lowrance AA, Christe M, Conner DA, Ingwall JS, Schoen FJ, Seidman CE, et al. A mouse model of familial hypertrophic cardiomyopathy. Science. 1996;272(5262):731-734
  52. 52. Woodman SE, Park DS, Cohen AW, Cheung MW, Chandra M, Shirani J, et al. Caveolin-3 knock-out mice develop a progressive cardiomyopathy and show hyperactivation of the p42/44 MAPK cascade. The Journal of Biological Chemistry. 2002;277(41):38988-38997
  53. 53. Kuga A, Ohsawa Y, Okada T, Kanda F, Kanagawa M, Toda T, et al. Endoplasmic reticulum stress response in P104L mutant caveolin-3 transgenic mice. Human Molecular Genetics. 2011;20(15):2975-2983
  54. 54. Nixon SJ, Wegner J, Ferguson C, Mery PF, Hancock JF, Currie PD, et al. Zebrafish as a model for caveolin-associated muscle disease; caveolin-3 is required for myofibril organization and muscle cell patterning. Human Molecular Genetics. 2005;14(13):1727-1743
  55. 55. Xu X, Meiler SE, Zhong TP, Mohideen M, Crossley DA, Burggren WW, et al. Cardiomyopathy in zebrafish due to mutation in an alternatively spliced exon of titin. Nature Genetics. 2002;30(2):205-209
  56. 56. Prabhakar R, Boivin GP, Grupp IL, Hoit B, Arteaga G, Solaro RJ, et al. A familial hypertrophic cardiomyopathy alpha-tropomyosin mutation causes severe cardiac hypertrophy and death in mice. Journal of Molecular and Cellular Cardiology. 2001;33(10):1815-1828
  57. 57. Muthuchamy M, Pieples K, Rethinasamy P, Hoit B, Grupp IL, Boivin GP, et al. Mouse model of a familial hypertrophic cardiomyopathy mutation in alpha-tropomyosin manifests cardiac dysfunction. Circulation Research. 1999;85(1):47-56
  58. 58. Tardiff JC, Factor SM, Tompkins BD, Hewett TE, Palmer BM, Moore RL, et al. A truncated cardiac troponin T molecule in transgenic mice suggests multiple cellular mechanisms for familial hypertrophic cardiomyopathy. The Journal of Clinical Investigation. 1998;101(12):2800-2811
  59. 59. Wang L, Seidman JG, Seidman CE. Narrative review: Harnessing molecular genetics for the diagnosis and management of hypertrophic cardiomyopathy. Annals of Internal Medicine. 2010;152(8):513-520
  60. 60. Tardiff JC, Hewett TE, Palmer BM, Olsson C, Factor SM, Moore RL, et al. Cardiac troponin T mutations result in allele-specific phenotypes in a mouse model for hypertrophic cardiomyopathy. The Journal of Clinical Investigation. 1999;104(4):469-481
  61. 61. Manso AM, Li R, Monkley SJ, Cruz NM, Ong S, Lao DH, et al. Talin1 has unique expression versus Talin 2 in the heart and modifies the hypertrophic response to pressure overload. The Journal of Biological Chemistry. 2013;288(6):4252-4264
  62. 62. Song Y, Xu J, Li Y, Jia C, Ma X, Zhang L, et al. Cardiac ankyrin repeat protein attenuates cardiac hypertrophy by inhibition of ERK1/2 and TGF-beta signaling pathways. PLoS ONE. 2012;7(12):e50436
  63. 63. Purevjav E, Arimura T, Augustin S, Huby AC, Takagi K, Nunoda S, et al. Molecular basis for clinical heterogeneity in inherited cardiomyopathies due to myopalladin mutations. Human Molecular Genetics. 2012;21(9):2039-2053
  64. 64. Sehnert AJ, Huq A, Weinstein BM, Walker C, Fishman M, Stainier DY. Cardiac troponin T is essential in sarcomere assembly and cardiac contractility. Nature Genetics. 2002;31(1):106-110
  65. 65. James J, Zhang Y, Osinska H, Sanbe A, Klevitsky R, Hewett TE, et al. Transgenic modeling of a cardiac troponin I mutation linked to familial hypertrophic cardiomyopathy. Circulation Research. 2000;87(9):805-811
  66. 66. Sanbe A, James J, Tuzcu V, Nas S, Martin L, Gulick J, et al. Transgenic rabbit model for human troponin I-based hypertrophic cardiomyopathy. Circulation. 2005;111(18):2330-2338
  67. 67. Huang X, Pi Y, Lee KJ, Henkel AS, Gregg RG, Powers PA, et al. Cardiac troponin I gene knockout: A mouse model of myocardial troponin I deficiency. Circulation Research. 1999;84(1):1-8
  68. 68. Yang Q , Sanbe A, Osinska H, Hewett TE, Klevitsky R, Robbins J. A mouse model of myosin binding protein C human familial hypertrophic cardiomyopathy. The Journal of Clinical Investigation. 1998;102(7):1292-1300
  69. 69. Palmer BM, McConnell BK, Li GH, Seidman CE, Seidman JG, Irving TC, et al. Reduced cross-bridge dependent stiffness of skinned myocardium from mice lacking cardiac myosin binding protein-C. Molecular and Cellular Biochemistry. 2004;263(1-2):73-80
  70. 70. Meurs KM, Sanchez X, David RM, Bowles NE, Towbin JA, Reiser PJ, et al. A cardiac myosin binding protein C mutation in the Maine Coon cat with familial hypertrophic cardiomyopathy. Human Molecular Genetics. 2005;14(23):3587-3593
  71. 71. Bang ML, Gu Y, Dalton ND, Peterson KL, Chien KR, Chen J. The muscle ankyrin repeat proteins CARP, Ankrd2, and DARP are not essential for normal cardiac development and function at basal conditions and in response to pressure overload. PLoS ONE. 2014;9(4):e93638
  72. 72. Ramratnam M, Sharma RK, D’Auria S, Lee SJ, Wang D, Huang XY, et al. Transgenic knockdown of cardiac sodium/glucose cotransporter 1 (SGLT1) attenuates PRKAG2 cardiomyopathy, whereas transgenic overexpression of cardiac SGLT1 causes pathologic hypertrophy and dysfunction in mice. Journal of the American Heart Association. 2014;3(4). pii:e000899
  73. 73. Lu D, Wang J, Li J, Guan F, Zhang X, Dong W, et al. Meox1 accelerates myocardial hypertrophic decompensation through Gata4. Cardiovascular Research. 2018;114(2):300-311
  74. 74. Bailey KE, MacGowan GA, Tual-Chalot S, Phillips L, Mohun TJ, Henderson DJ, et al. Disruption of embryonic ROCK signaling reproduces the sarcomeric phenotype of hypertrophic cardiomyopathy. JCI Insight. 2019;5. pii:125172
  75. 75. Farrell E, Armstrong AE, Grimes AC, Naya FJ, de Lange WJ, Ralphe JC. Transcriptome analysis of cardiac hypertrophic growth in MYBPC3-Null mice suggests early responders in hypertrophic remodeling. Frontiers in Physiology. 2018;9:1442
  76. 76. Lowey S, Bretton V, Joel PB, Trybus KM, Gulick J, Robbins J, et al. Hypertrophic cardiomyopathy R403Q mutation in rabbit ¦Â-myosin reduces contractile function at the molecular and myofibrillar levels. Proceedings of the National Academy of Sciences of the United States of America. 2018;115(44):11238-11243
  77. 77. Ehsan M, Kelly M, Hooper C, Yavari A, Beglov J, Bellahcene M, et al. Mutant muscle LIM protein C58G causes cardiomyopathy through protein depletion. Journal of Molecular and Cellular Cardiology. 2018;121:287-296
  78. 78. Montag J, Petersen B, Flögel AK, Becker E, Lucas-Hahn A, Cost GJ, et al. Successful knock-in of hypertrophic cardiomyopathy-mutation R723G into the MYH7 gene mimics HCM pathology in pigs. Scientific Reports. 2018;8(1):4786
  79. 79. Ferrantini C, Coppini R, Pioner JM, Gentile F, Tosi B, Mazzoni L, et al. Pathogenesis of hypertrophic cardiomyopathy is mutation rather than disease specific: A comparison of the cardiac troponin T E163R and R92Q mouse models. Journal of the American Heart Association. 2017;6(7). pii:e005407
  80. 80. Hueneke R, Adenwala A, Mellor RL, Seidman JG, Seidman CE, Nerbonne JM. Early remodeling of repolarizing K+ currents in the ¦ÁMHC403/+ mouse model of familial hypertrophic cardiomyopathy. Journal of Molecular and Cellular Cardiology. 2017;103:93-101
  81. 81. Sr LL, Bedja D, Sysa-Shah P, Liu H, Maxwell A, Yi X, et al. Echocardiographic characterization of a murine model of hypertrophic obstructive cardiomyopathy induced by cardiac-specific overexpression of epidermal growth ractor receptor 2. Comparative Medicine. 2016;66(4):268-277
  82. 82. Cordier L, Hack AA, Scott MO, Barton-Davis ER, Gao G, Wilson JM, et al. Rescue of skeletal muscles of gamma-sarcoglycan-deficient mice with adeno-associated virus-mediated gene transfer. Molecular Therapy. 2000;1(2):119-129
  83. 83. Rutschow D, Bauer R, Gohringer C, Bekeredjian R, Schinkel S, Straub V, et al. S151A delta-sarcoglycan mutation causes a mild phenotype of cardiomyopathy in mice. European Journal of Human Genetics. 2014;22(1):119-125
  84. 84. Araishi K, Sasaoka T, Imamura M, Noguchi S, Hama H, Wakabayashi E, et al. Loss of the sarcoglycan complex and sarcospan leads to muscular dystrophy in beta-sarcoglycan-deficient mice. Human Molecular Genetics. 1999;8(9):1589-1598
  85. 85. Miyagoe-Suzuki Y, Nakagawa M, Takeda S. Merosin and congenital muscular dystrophy. Microscopy Research and Technique. 2000;48(3-4):181-191
  86. 86. Sicinski P, Geng Y, Ryder-Cook AS, Barnard EA, Darlison MG, Barnard PJ. The molecular basis of muscular dystrophy in the mdx mouse: A point mutation. Science. 1989;244(4912):1578-1580
  87. 87. Guyon JR, Mosley AN, Zhou Y, O’Brien KF, Sheng X, Chiang K, et al. The dystrophin associated protein complex in zebrafish. Human Molecular Genetics. 2003;12(6):601-615
  88. 88. Jones BR, Brennan S, Mooney CT, Callanan JJ, McAllister H, Guo LT, et al. Muscular dystrophy with truncated dystrophin in a family of Japanese spitz dogs. Journal of the Neurological Sciences. 2004;217(2):143-149
  89. 89. Rethinasamy P, Muthuchamy M, Hewett T, Boivin G, Wolska BM, Evans C, et al. Molecular and physiological effects of alpha-tropomyosin ablation in the mouse. Circulation Research. 1998;82(1):116-123
  90. 90. Rajan S, Ahmed RP, Jagatheesan G, Petrashevskaya N, Boivin GP, Urboniene D, et al. Dilated cardiomyopathy mutant tropomyosin mice develop cardiac dysfunction with significantly decreased fractional shortening and myofilament calcium sensitivity. Circulation Research. 2007;101(2):205-214
  91. 91. Wang X, Osinska H, Dorn GW 2nd, Nieman M, Lorenz JN, Gerdes AM, et al. Mouse model of desmin-related cardiomyopathy. Circulation. 2001;103(19):2402-2407
  92. 92. Li M, Andersson-Lendahl M, Sejersen T, Arner A. Knockdown of desmin in zebrafish larvae affects interfilament spacing and mechanical properties of skeletal muscle. The Journal of General Physiology. 2013;141(3):335-345
  93. 93. Milner DJ, Weitzer G, Tran D, Bradley A, Capetanaki Y. Disruption of muscle architecture and myocardial degeneration in mice lacking desmin. The Journal of Cell Biology. 1996;134(5):1255-1270
  94. 94. Arber S, Hunter JJ, Ross J Jr, Hongo M, Sansig G, Borg J, et al. MLP-deficient mice exhibit a disruption of cardiac cytoarchitectural organization, dilated cardiomyopathy, and heart failure. Cell. 1997;88(3):393-403
  95. 95. Purevjav E, Varela J, Morgado M, Kearney DL, Li H, Taylor MD, et al. Nebulette mutations are associated with dilated cardiomyopathy and endocardial fibroelastosis. Journal of the American College of Cardiology. 2010;56(18):1493-1502
  96. 96. Hassel D, Dahme T, Erdmann J, Meder B, Huge A, Stoll M, et al. Nexilin mutations destabilize cardiac Z-disks and lead to dilated cardiomyopathy. Nature Medicine. 2009;15(11):1281-1288
  97. 97. Knoll R, Linke WA, Zou P, Miocic S, Kostin S, Buyandelger B, et al. Telethonin deficiency is associated with maladaptation to biomechanical stress in the mammalian heart. Circulation Research. 2011;109(7):758-769
  98. 98. Zhang R, Yang J, Zhu J, Xu X. Depletion of zebrafish Tcap leads to muscular dystrophy via disrupting sarcomere-membrane interaction, not sarcomere assembly. Human Molecular Genetics. 2009;18(21):4130-4140
  99. 99. Zhou Q , Chu PH, Huang C, Cheng CF, Martone ME, Knoll G, et al. Ablation of Cypher, a PDZ-LIM domain Z-line protein, causes a severe form of congenital myopathy. The Journal of Cell Biology. 2001;155(4):605-612
  100. 100. Fujita M, Mitsuhashi H, Isogai S, Nakata T, Kawakami A, Nonaka I, et al. Filamin C plays an essential role in the maintenance of the structural integrity of cardiac and skeletal muscles, revealed by the medaka mutant zacro. Developmental Biology. 2012;361(1):79-89
  101. 101. Ho CY, Jaalouk DE, Vartiainen MK, Lammerding J, Lamin A. C and emerin regulate MKL1-SRF activity by modulating actin dynamics. Nature. 2013;497(7450):507-511
  102. 102. Sullivan T, Escalante-Alcalde D, Bhatt H, Anver M, Bhat N, Nagashima K, et al. Loss of A-type lamin expression compromises nuclear envelope integrity leading to muscular dystrophy. The Journal of Cell Biology. 1999;147(5):913-920
  103. 103. Rehmani T, Salih M, Tuana BS. Cardiac-specific cre induces age-dependent dilated cardiomyopathy (DCM) in mice. Molecules: A Journal of Synthetic Chemistry and Natural Product Chemistry. 2019;24(6):1189
  104. 104. Dong W, Guan FF, Zhang X, Gao S, Liu N, Chen W, et al. Dhcr24 activates the PI3K/Akt/HKII pathway and protects against dilated cardiomyopathy in mice. Animal Models and Experimental Medicine. 2018;1(1):40-52
  105. 105. Nguyen MN, Ziemann M, Kiriazis H, Su Y, Thomas Z, Lu Q , et al. Galectin-3 deficiency ameliorates fibrosis and remodeling in dilated cardiomyopathy mice with enhanced Mst1 signaling. American Journal of Physiology Heart and Circulatory Physiology. 2019;316(1):H45-H60
  106. 106. Ednie AR, Deng W, Yip KP, Bennett ES. Reduced myocyte complex N-glycosylation causes dilated cardiomyopathy. FASEB Journal: Official Publication of the Federation of American Societies for Experimental Biology. 2019;33(1):1248-1261
  107. 107. Murayama R, Kimura-Asami M, Togo-Ohno M, Yamasaki-Kato Y, Naruse TK, Yamamoto T, et al. Phosphorylation of the RSRSP stretch is critical for splicing regulation by RNA-binding motif protein 20 (RBM20) through nuclear localization. Scientific Reports. 2018;8(1):8970
  108. 108. Li J, Gresham KS, Mamidi R, Doh CY, Wan X, Deschenes I, et al. Sarcomere-based genetic enhancement of systolic cardiac function in a murine model of dilated cardiomyopathy. International Journal of Cardiology. 2018;273:168-176
  109. 109. Chu M, Novak SM, Cover C, Wang AA, Chinyere IR, Juneman EB, et al. Increased cardiac arrhythmogenesis associated with gap junction remodeling with upregulation of RNA-binding protein FXR1. Circulation. 2018;137(6):605-618
  110. 110. Mohamed RM, Morimoto S, Ibrahim IA, Zhan DY, Du CK, Arioka M, et al. GSK-3¦Â heterozygous knockout is cardioprotective in a knockin mouse model of familial dilated cardiomyopathy. American Journal of Physiology Heart and Circulatory Physiology. 2016;310(11):H1808-H1815
  111. 111. Aherrahrou Z, Schlossarek S, Stoelting S, Klinger M, Geertz B, Weinberger F, et al. Knock-out of nexilin in mice leads to dilated cardiomyopathy and endomyocardial fibroelastosis. Basic Research in Cardiology. 2016;111(1):6
  112. 112. Laury-Kleintop LD, Mulgrew JR, Heletz I, Nedelcoviciu RA, Chang MY, Harris DM, et al. Cardiac-specific disruption of Bin1 in mice enables a model of stress- and age-associated dilated cardiomyopathy. Journal of Cellular Biochemistry. 2015;116(11):2541-2551
  113. 113. Huang XP, Du JF. Troponin I, cardiac diastolic dysfunction and restrictive cardiomyopathy. Acta Pharmacologica Sinica. 2004;25(12):1569-1575
  114. 114. Davis J, Wen H, Edwards T, Metzger JM. Thin filament disinhibition by restrictive cardiomyopathy mutant R193H troponin I induces Ca2+-independent mechanical tone and acute myocyte remodeling. Circulation Research. 2007;100(10):1494-1502
  115. 115. Wen Y, Xu Y, Wang Y, Pinto JR, Potter JD, Kerrick WG. Functional effects of a restrictive-cardiomyopathy-linked cardiac troponin I mutation (R145W) in transgenic mice. Journal of Molecular Biology. 2009;392(5):1158-1167
  116. 116. Huby AC, Mendsaikhan U, Takagi K, Martherus R, Wansapura J, Gong N, et al. Disturbance in Z-disk mechanosensitive proteins induced by a persistent mutant myopalladin causes familial restrictive cardiomyopathy. Journal of the American College of Cardiology. 2014;64(25):2765-2776
  117. 117. Yuan CC, Kazmierczak K, Liang J, Kanashiro-Takeuchi R, Irving TC, Gomes AV, et al. Hypercontractile mutant of ventricular myosin essential light chain leads to disruption of sarcomeric structure and function and results in restrictive cardiomyopathy in mice. Cardiovascular Research. 2017;113(10):1124-1136
  118. 118. Padron-Barthe L, Dominguez F, Garcia-Pavia P, Lara-Pezzi E. Animal models of arrhythmogenic right ventricular cardiomyopathy: What have we learned and where do we go? Insight for therapeutics. Basic Research in Cardiology. 2017;112(5):50
  119. 119. Pilichou K, Remme CA, Basso C, Campian ME, Rizzo S, Barnett P, et al. Myocyte necrosis underlies progressive myocardial dystrophy in mouse dsg2-related arrhythmogenic right ventricular cardiomyopathy. The Journal of Experimental Medicine. 2009;206(8):1787-1802
  120. 120. Thomas SA, Schuessler RB, Berul CI, Beardslee MA, Beyer EC, Mendelsohn ME, et al. Disparate effects of deficient expression of connexin43 on atrial and ventricular conduction: Evidence for chamber-specific molecular determinants of conduction. Circulation. 1998;97(7):686-691
  121. 121. Lyon RC, Mezzano V, Wright AT, Pfeiffer E, Chuang J, Banares K, et al. Connexin defects underlie arrhythmogenic right ventricular cardiomyopathy in a novel mouse model. Human Molecular Genetics. 2014;23(5):1134-1150
  122. 122. Ewart JL, Cohen MF, Meyer RA, Huang GY, Wessels A, Gourdie RG, et al. Heart and neural tube defects in transgenic mice overexpressing the Cx43 gap junction gene. Development. 1997;124(7):1281-1292
  123. 123. Garcia-Gras E, Lombardi R, Giocondo MJ, Willerson JT, Schneider MD, Khoury DS, et al. Suppression of canonical Wnt/beta-catenin signaling by nuclear plakoglobin recapitulates phenotype of arrhythmogenic right ventricular cardiomyopathy. The Journal of Clinical Investigation. 2006;116(7):2012-2021
  124. 124. Yang Z, Bowles NE, Scherer SE, Taylor MD, Kearney DL, Ge S, et al. Desmosomal dysfunction due to mutations in desmoplakin causes arrhythmogenic right ventricular dysplasia/cardiomyopathy. Circulation Research. 2006;99(6):646-655
  125. 125. Rietscher K, Wolf A, Hause G, Rother A, Keil R, Magin TM, et al. Growth retardation, loss of desmosomal adhesion, and impaired tight junction function identify a unique role of plakophilin 1 in vivo. The Journal of Investigative Dermatology. 2016;136(7):1471-1478
  126. 126. Dubash AD, Kam CY, Aguado BA, Patel DM, Delmar M, Shea LD, et al. Plakophilin-2 loss promotes TGF-beta1/p38 MAPK-dependent fibrotic gene expression in cardiomyocytes. The Journal of Cell Biology. 2016;212(4):425-438
  127. 127. Moriarty MA, Ryan R, Lalor P, Dockery P, Byrnes L, Grealy M. Loss of plakophilin 2 disrupts heart development in zebrafish. The International Journal of Developmental Biology. 2012;56(9):711-718
  128. 128. Heuser A, Plovie ER, Ellinor PT, Grossmann KS, Shin JT, Wichter T, et al. Mutant desmocollin-2 causes arrhythmogenic right ventricular cardiomyopathy. American Journal of Human Genetics. 2006;79(6):1081-1088
  129. 129. Calore M, Lorenzon A, Vitiello L, Poloni G, Khan MAF, Beffagna G, et al. A novel murine model for arrhythmogenic cardiomyopathy points to a pathogenic role of Wnt signalling and miRNA dysregulation. Cardiovascular Research. 2019;115(4):739-751
  130. 130. Kant S, Holthofer B, Magin TM, Krusche CA, Leube RE. Desmoglein 2-dependent arrhythmogenic cardiomyopathy is caused by a loss of adhesive function. Circulation. Cardiovascular Genetics. 2015;8(4):553-563
  131. 131. Krusche CA, Holthofer B, Hofe V, van de Sandt AM, Eshkind L, Bockamp E, et al. Desmoglein 2 mutant mice develop cardiac fibrosis and dilation. Basic Research in Cardiology. 2011;106(4):617-633
  132. 132. Martin ED, Moriarty MA, Byrnes L, Grealy M. Plakoglobin has both structural and signalling roles in zebrafish development. Developmental Biology. 2009;327(1):83-96
  133. 133. Kirchhof P, Fabritz L, Zwiener M, Witt H, Schafers M, Zellerhoff S, et al. Age- and training-dependent development of arrhythmogenic right ventricular cardiomyopathy in heterozygous plakoglobin-deficient mice. Circulation. 2006;114(17):1799-1806
  134. 134. Montnach J, Chizelle FF, Belbachir N, Castro C, Li L, Loussouarn G, et al. Arrhythmias precede cardiomyopathy and remodeling of Ca2+ handling proteins in a novel model of long QT syndrome. Journal of Molecular and Cellular Cardiology. 2018;123:13-25
  135. 135. Mazurek SR, Calway T, Harmon C, Farrell P, Kim GH. MicroRNA-130a regulation of desmocollin 2 in a novel model of arrhythmogenic cardiomyopathy. MicroRNA. 2017;6(2):143-150
  136. 136. Ozcan C, Battaglia E, Young R, Suzuki G. LKB1 knockout mouse develops spontaneous atrial fibrillation and provides mechanistic insights into human disease process. Journal of the American Heart Association. 2015;4(3):e001733
  137. 137. Asimaki A, Kapoor S, Plovie E, Karin Arndt A, Adams E, Liu Z, et al. Identification of a new modulator of the intercalated disc in a zebrafish model of arrhythmogenic cardiomyopathy. Science Translational Medicine. 2014;6(240):240ra274
  138. 138. Stroud MJ, Fang X, Zhang J, Guimaraes-Camboa N, Veevers J, Dalton ND, et al. Luma is not essential for murine cardiac development and function. Cardiovascular Research. 2018;114(3):378-388
  139. 139. Zheng G, Jiang C, Li Y, Yang D, Ma Y, Zhang B, et al. TMEM43-S358L mutation enhances NF-kappaB-TGFbeta signal cascade in arrhythmogenic right ventricular dysplasia/cardiomyopathy. Protein & Cell. 9 Feb 2018;10(2):104-119
  140. 140. Liu Y, Chen H, Shou W. Potential common pathogenic pathways for the left ventricular noncompaction cardiomyopathy (LVNC). Pediatric Cardiology. 2018;39(6):1099-1106
  141. 141. Ashraf H, Pradhan L, Chang EI, Terada R, Ryan NJ, Briggs LE, et al. A mouse model of human congenital heart disease: High incidence of diverse cardiac anomalies and ventricular noncompaction produced by heterozygous Nkx2-5 homeodomain missense mutation. Circulation. Cardiovascular Genetics. 2014;7(4):423-433
  142. 142. Choquet C, Nguyen THM, Sicard P, Buttigieg E, Tran TT, Kober F, et al. Deletion of Nkx2-5 in trabecular myocardium reveals the developmental origins of pathological heterogeneity associated with ventricular non-compaction cardiomyopathy. PLoS Genetics. 2018;14(7):e1007502
  143. 143. Shou W, Aghdasi B, Armstrong DL, Guo Q , Bao S, Charng MJ, et al. Cardiac defects and altered ryanodine receptor function in mice lacking FKBP12. Nature. 1998;391(6666):489-492
  144. 144. Lee Y, Song AJ, Baker R, Micales B, Conway SJ, Lyons GE. Jumonji, a nuclear protein that is necessary for normal heart development. Circulation Research. 2000;86(9):932-938
  145. 145. King T, Bland Y, Webb S, Barton S, Brown NA. Expression of Peg1 (Mest) in the developing mouse heart: Involvement in trabeculation. Developmental Dynamics. 2002;225(2):212-215
  146. 146. Luxan G, Casanova JC, Martinez-Poveda B, Prados B, D’Amato G, MacGrogan D, et al. Mutations in the NOTCH pathway regulator MIB1 cause left ventricular noncompaction cardiomyopathy. Nature Medicine. 2013;19(2):193-201
  147. 147. Inoue S, Moriya M, Watanabe Y, Miyagawa-Tomita S, Niihori T, Oba D, et al. New BRAF knockin mice provide a pathogenetic mechanism of developmental defects and a therapeutic approach in cardio-facio-cutaneous syndrome. Human Molecular Genetics. 2014;23(24):6553-6566
  148. 148. Liu Z, Li W, Ma X, Ding N, Spallotta F, Southon E, et al. Essential role of the zinc finger transcription factor Casz1 for mammalian cardiac morphogenesis and development. The Journal of Biological Chemistry. 2014;289(43):29801-29816
  149. 149. Kokoszka JE, Waymire KG, Flierl A, Sweeney KM, Angelin A, MacGregor GR, et al. Deficiency in the mouse mitochondrial adenine nucleotide translocator isoform 2 gene is associated with cardiac noncompaction. Biochimica et Biophysica Acta. 2016;1857(8):1203-1212
  150. 150. Li D, Hallett MA, Zhu W, Rubart M, Liu Y, Yang Z, et al. Dishevelled-associated activator of morphogenesis 1 (Daam1) is required for heart morphogenesis. Development. 2011;138(2):303-315
  151. 151. Clay H, Wilsbacher LD, Wilson SJ, Duong DN, McDonald M, Lam I, et al. Sphingosine 1-phosphate receptor-1 in cardiomyocytes is required for normal cardiac development. Developmental Biology. 2016;418(1):157-165
  152. 152. Hirai M, Arita Y, McGlade CJ, Lee KF, Chen J, Evans SM. Adaptor proteins NUMB and NUMBL promote cell cycle withdrawal by targeting ERBB2 for degradation. The Journal of Clinical Investigation. 2017;127(2):569-582
  153. 153. Bourke LM, Del Monte-Nieto G, Outhwaite JE, Bharti V, Pollock PM, Simmons DG, et al. Loss of rearranged L-Myc fusion (RLF) results in defects in heart development in the mouse. Differentiation; Research in Biological Diversity. 2017;94:8-20
  154. 154. Baardman ME, Zwier MV, Wisse LJ, Gittenberger-de Groot AC, Kerstjens-Frederikse WS, Hofstra RM, et al. Common arterial trunk and ventricular non-compaction in Lrp2 knockout mice indicate a crucial role of LRP2 in cardiac development. Disease Models & Mechanisms. 2016;9(4):413-425
  155. 155. Lin W, Li D, Cheng L, Li L, Liu F, Hand NJ, et al. Zinc transporter Slc39a8 is essential for cardiac ventricular compaction. The Journal of Clinical Investigation. 2018;128(2):826-833
  156. 156. Cao Q , Shen Y, Liu X, Yu X, Yuan P, Wan R, et al. Phenotype and functional analyses in a transgenic mouse model of left ventricular noncompaction caused by a DTNA mutation. International Heart Journal. 2017;58(6):939-947
  157. 157. Chen X, Qin L, Liu Z, Liao L, Martin JF, Xu J. Knockout of SRC-1 and SRC-3 in mice decreases cardiomyocyte proliferation and causes a noncompaction cardiomyopathy phenotype. International Journal of Biological Sciences. 2015;11(9):1056-1072
  158. 158. Rhee S, Chung JI, King DA, D’Amato G, Paik DT, Duan A, et al. Endothelial deletion of Ino80 disrupts coronary angiogenesis and causes congenital heart disease. Nature Communications. 2018;9(1):368
  159. 159. Phoon CK, Acehan D, Schlame M, Stokes DL, Edelman-Novemsky I, Yu D, et al. Tafazzin knockdown in mice leads to a developmental cardiomyopathy with early diastolic dysfunction preceding myocardial noncompaction. Journal of the American Heart Association. 2012;1(2). pii:jah3-e000455
  160. 160. Hearse DJ, Sutherland FJ. Experimental models for the study of cardiovascular function and disease. Pharmacological Research. 2000;41(6):597-603
  161. 161. Savoji H, Mohammadi MH, Rafatian N, Toroghi MK, Wang EY, Zhao Y, et al. Cardiovascular disease models: A game changing paradigm in drug discovery and screening. Biomaterials. 2019;198:3-26

Written By

Enkhsaikhan Purevjav

Submitted: 26 March 2019 Reviewed: 06 August 2019 Published: 15 October 2019