Next Article in Journal
miR-129a-3p Inhibits PEDV Replication by Targeting the EDA-Mediated NF-κB Pathway in IPEC-J2 Cells
Next Article in Special Issue
The Effect of Sera from Children with Obstructive Sleep Apnea Syndrome (OSAS) on Human Cardiomyocytes Differentiated from Human Embryonic Stem Cells
Previous Article in Journal
Pancreatic Adenocarcinoma Therapeutics Targeting RTK and TGF Beta Receptor
Previous Article in Special Issue
Investigating LMNA-Related Dilated Cardiomyopathy Using Human Induced Pluripotent Stem Cell-Derived Cardiomyocytes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Patient-Specific iPSCs for Modelling and Treatment of X-Linked Cardiomyopathies

1
Cardiology Division, Department of Medicine, Li Ka Shing Faculty of Medicine, The University of Hong Kong, Hong Kong, China
2
Centre of Translational Stem Cell Biology, Hong Kong Science and Technology Park, Hong Kong, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally.
Int. J. Mol. Sci. 2021, 22(15), 8132; https://doi.org/10.3390/ijms22158132
Submission received: 29 June 2021 / Revised: 22 July 2021 / Accepted: 24 July 2021 / Published: 29 July 2021

Abstract

:
Inherited cardiomyopathies are among the major causes of heart failure and associated with significant mortality and morbidity. Currently, over 70 genes have been linked to the etiology of various forms of cardiomyopathy, some of which are X-linked. Due to the lack of appropriate cell and animal models, it has been difficult to model these X-linked cardiomyopathies. With the advancement of induced pluripotent stem cell (iPSC) technology, the ability to generate iPSC lines from patients with X-linked cardiomyopathy has facilitated in vitro modelling and drug testing for the condition. Nonetheless, due to the mosaicism of the X-chromosome inactivation, disease phenotypes of X-linked cardiomyopathy in heterozygous females are also usually more heterogeneous, with a broad spectrum of presentation. Recent advancements in iPSC procedures have enabled the isolation of cells with different lyonisation to generate isogenic disease and control cell lines. In this review, we will summarise the current strategies and examples of using an iPSC-based model to study different types of X-linked cardiomyopathy. The potential application of isogenic iPSC lines derived from a female patient with heterozygous Danon disease and drug screening will be demonstrated by our preliminary data. The limitations of an iPSC-derived cardiomyocyte-based platform will also be addressed.

1. Introduction

Cardiomyopathy is a highly heterogeneous myocardial disease that may contribute to the development of heart failure. In general, cardiomyopathies can be inherited or acquired. To date, more than 70 genes have been identified that are associated with various forms of cardiomyopathies, some of which (Table 1) are X-linked. Due to the substantial differences in the cardiac physiology of humans and rodents, the establishment of a disease model for X-linked cardiomyopathy has been difficult. For example, rodents have a rapid heart rate (300–800 bpm, compared with 60–100 bpm in humans) and their hearts have adapted to this with rapid systolic and diastolic filling [1]. Consequently, the action potentials of mice cardiomyocytes lack the plateau phase containing a sustained calcium release [2]. In addition, a recent study also showed that rodents lack the delayed rectifier potassium currents and show different responses to ion channel blockers compared with large animals and humans [3]. These cellular and electrophysiological differences may limit the usefulness of rodent models in studying X-linked cardiomyopathy [4]. With technical advancements in cellular reprogramming, adult somatic cells can now be reprogrammed into pluripotent stem cells [5,6,7], and such induced pluripotent stem cells (iPSCs) can be further differentiated into cardiomyocytes for disease modelling and drug testing [8]. To date, various studies have evaluated the possibility of using patient-specific iPSCs to model different forms of X-linked cardiomyopathy. In addition, there are some gender-specific differences in an iPSC model of X-linked cardiomyopathies. Apart from the different transmission pattern, the phenotypic expression in heterozygous females is usually more heterogeneous due to random X-chromosome inactivation. Due to the presence of a population of wildtype-like cells in the myocardium, heterozygous females with an X-linked dominant disorder (e.g., Danon disease) usually show a later disease onset and less severe symptoms than their male counterparts [9]. Disease phenotypes may also be observed in the heterozygous female carrying an X-linked recessive gene. Strategies, such as X-chromosome re-activation, have been demonstrated on iPSC platforms to solve these issues. In this review, the current strategies and examples of using patient-specific iPSC technology in modelling X-linked cardiomyopathies will be discussed. Our preliminary data using isogenic iPSC lines for drug screening will also be presented. Finally, the limitations of an iPSC modelling platform will be addressed.

2. The Advantages of an iPSC-Based Model

Most stem cells are of embryonic origin or derived from adult tissues with the ability to proliferate (e.g., bone marrow). These stem cells differentiate into a myriad of cell types but do not undergo a de-differentiation process [30]. Induced pluripotent stem cells can be reprogrammed directly from terminally differentiated somatic cells by over-expression of the four transcription factors (Oct4, Sox2, Klf4, and cMyc). The technology was first introduced in 2006 by Yamanaka et al. [7], and since then, iPSCs derived from patients or healthy individuals have been widely used for modelling various inherited disorders. Research in regenerative medicine is possible because of the ability of iPSCs to proliferate indefinitely and differentiate into various cell types, such as neurons, hepatocytes, and cardiomyocytes.
Since iPSCs are human in origin and pluripotent, they have enabled the development of patient-specific cell models to recapitulate the pathophysiology of human diseases [8]. This facilitates the study of inherited diseases, since the derived iPSCs and their differentiated cells will contain the same genomic profile as their patient somatic cell origin [31]. Additionally, novel drugs specific to the patient’s disease can be screened and tested on these disease-expressing cell types. For example, iPSC-derived cardiomyocytes can be used to reproduce the phenotype of autophagic dysfunction in cardiomyopathy caused by LAMP2 mutation [32]; this also allows the identification of novel drugs that can reactivate the silenced X-chromosome.
An iPSC-based model has multiple advantages compared to the traditional cell lines. The traditional primary cell lines were difficult to establish from patients, especially for cell types that were difficult to access clinically. For instance, there are limited scenarios where human adult cardiomyocytes can be isolated, so most cultured cardiomyocytes have been isolated from neonatal cardiomyocytes [33]. Nonetheless neonatal cardiomyocytes express an electrophysiological and proteomics profile that differs from that of their adult counterparts. The iPSC-based model enables the derivation of a broad range of cell types from a readily accessible somatic cell source, such as blood cells and skin fibroblasts [34].
In the early years, a disease iPSC line would be paired with a control cell line derived from an individual without the disease. Nonetheless, with the advancement of gene-editing techniques, iPSC models can now be genetically manipulated to study the disease phenotype [35]. The development of the CRISPR-Cas9 system has enabled precise and specific gene-editing and epigenetic modulations [36,37]. Along with the expandable self-renewal features of iPSCs, the CRISPR-Cas9 system allows the induction of genetic changes in the targeted cell line, such as gene knockout/knock-in, in order to control gene expression [38]. In disease modelling, knockout of the disease-causing gene facilitates the establishment of an isogenic control cell line differing only in the disease-relevant gene [39]. Gene-editing can also introduce disease-causing mutations to a normal cell line to study the disease phenotype. Consequently, diverse genetically defective diseases can be modelled and help in the understanding of multiple genetic conditions [40].
The unique features of iPSCs provide many potential opportunities in disease modelling. An iPSC-based system allows for the development of a 3D platform to form self-organised organoids with intensive cell–cell interactions. These iPSC-derived organoids form primary tissue-like clumps that contain multiple cell types similar to their natural human counterpart [41]. This has advantages over the 2D culture platform, since the 3D organoids are expected to be more mature [42]. Apart from forming organoids, iPSCs also enable the formation of human–animal chimaera models via xenotransplantation. Engrafted cells can interact with the host microenvironment, replicating the diseased organ’s typical environment. Examples include the development of solid organ and neural chimaeras in order to study different diseases and model conditions, such as familial hypercholesterolemia and autistic spectrum disorders (ASD) [43,44].

3. The Procedures for Producing iPSCs

The fundamental principle of iPSC production is to express the reprogramming Yamanaka factors (Oct3/4, Sox2, c-Myc, and Klf4) in cells in order to drive the transcriptome into the pluripotent stage. Theoretically, iPSCs can be reprogrammed from any somatic cells with a complete genome. Although fibroblasts derived from skin biopsy have been widely used to generate iPSCs, collection is invasive and painful; thus, various alternatives are now available, including peripheral blood mononuclear cells (PBMC) from a blood sample and renal epithelial cells from urine [45,46,47,48].
The methods of delivering the factors into the somatic cells dictate the outcomes of the iPSCs. The first method involves the use of a retrovirus to transduce fibroblasts, but viral integration has been a huge issue [7]. Integration-free systems include either using the non-integrating virus, excisable transduction system, or the delivery of recombinant reprogramming proteins. Using lentiviral vectors instead of the retrovirus results in a higher efficiency of reprogramming with fewer viral integrations [49]. Such a simpler approach eliminates the risk of the possible integration of the virus genes into the host genome, so the resultant iPSC lines are more clinically relevant. Although integration-free systems can successfully reprogram somatic cells into iPSCs with equal quality, they are less efficient than the conventional integrating vectors, owing to the less sustainable reprogramming expression.
The iPSC was traditionally cultured on serum-dependent feeder layers comprised of either mouse or human fibroblasts [45]. Nonetheless the use of animal serum raised concerns about possible xeno-contamination [50]. Over time, iPSCs have been cultured in feeder-free conditions using components such as Matrigel, Geltrex, and other extracellular matrix proteins. Most iPSCs are now cultured in Essential 8™ or MTeSR™ medium; both systems maintain iPSC pluripotency after many passages and retain the ability to differentiate without causing any chromosomal aberrations [51].
Refering to X-linked disorders, for male patients with only one X-chromosome in the genome, the iPSC line can usually be routinely generated and the cell lines produced will contain the mutation. Nevertheless, if the patient is a heterozygous female, due to the mosaic patten of X-chromosome inactivation, the generation of iPSCs with a specific active allele is not guaranteed. Based on our observation that the X-chromosome inactivation status remains unchanged during reprogramming, one should consider separating cells (skin fibroblasts or PBMCs) with a different X-chromosome inactivation status prior to transduction of the reprogramming factors. This can be achieved by diluting the cells to single cells and then checking the expression of the wildtype or mutant allele of interest using immunostaining or mRNA-based methods. (Figure 1). Using this approach, we created an isogenic pair of iPSCs that were genetically identical, differing only in their expression of the specific LAMP2 allele, and used it to model Danon disease [7].

4. Differentiation of Established iPSC Lines to Cardiomyocytes

Differentiation protocols of established iPSC lines to cardiomyocytes have progressed from early differentiation of only 8–22% efficiency to current methods that achieve a high yield of cardiac troponin-T positive cells [52,53]. Earlier approaches included the use of embryoid bodies with serum-containing media, but their efficiency was very low [52]. Later efficient protocols included cardiac induction using cytokines with one of three established methods: suspended embryoid bodies, forced aggregation, or cell monolayer [54]. The first stage of cardiac differentiation mimics embryonic signals to induce mesodermal development [55,56,57,58,59], and the second step is cardiac specification [55,60,61,62]. For the maintenance of differentiated cardiomyocytes, most efficient protocols consist of basal medium RPMI 1640 supplemented with B27 serum or fetal bovine serum (FBS) at various concentrations [63,64,65]. Isolation of cardiomyocytes can be easily achieved using a glucose-depleted culture medium containing abundant lactate, since other cell types are not capable of using lactate for energy production. Another method of cardiomyocyte isolation includes cell sorting using antibodies against cardiomyocyte cell surface markers, such as SIRPa and VCAM1 [66,67,68,69].
The current differentiation protocol generates a heterogeneous group of cardiomyocytes. The addition of specific molecules has been shown to promote or inhibit certain cardiomyocyte subtypes in order to generate more specific disease models and drug testing platforms [70]. Studies show that the induction of appropriate mesoderm with different concentrations of small molecules to activate or inhibit certain signalling pathways specifically yields atrial, ventricular, or pacemaker-type cardiomyocytes [71,72,73,74,75,76,77,78]. Cardiomyocyte differentiation kits are now commercially available. Although their components may vary, they are convenient alternatives for laboratories that are routinely performing cardiac differentiation. For example, we are using the GibcoTM PSC Cardiomyocyte Differentiation Kit from Thermo Fisher Scientific with a differentiation efficiency of 30–90%, depending on the genetic background of the iPSCs. For wildtype iPSCs, we consistently yield over 80% of cTNT positive cells (Figure 2).

5. Functional Characterisation of Cardiomyocytes Derived from iPSCs

Depending on the gene involved, the effect of a specific mutation on the cardiac phenotype may require individual evaluation. In general, the characterisation of patient-specific iPSC-derived cardiomyocytes should involve the structural organisation and electrophysiological and calcium handling properties, as well as contractile function.

5.1. Characterisation of Structural Properties

Cardiomyopathy patients usually exhibit characteristic changes in the heart tissue. For example, many X-linked mutations lead to hypertrophic phenotypes [13,17]. Nonetheless, it should be noted that iPSC-derived cardiomyocytes are usually irregular in shape. Although not many X-linked cardiomyopathy-associated genes are directly involved in cytoskeleton protein production, many mutations can precipitate disruption of the cytoskeleton organisation as a secondary effect [79]. As such, the examination of sarcomeric protein organisation may help to evaluate the pathophysiological effects of different mutations. Immunofluorescent microscopy can be used to identify pathological changes, such as disorganised myofilaments, by staining with antibodies specific to the sarcomeric proteins (e.g., sarcomeric α-actinin and cardiac-troponin-T) [32]. Electron microscopy with a resolution power of 50–200 pm can also identify phenotypic changes at an ultra-structural level [80]. Abnormalities in membranous structures, as well as glycogen granule accumulation, can be revealed, and are extremely useful in the evaluation of cardiac defects resulting from protein aggregation, autophagic dysfunction (Danon disease), and dysregulated intracellular storage (Danon disease and Fabry disease).

5.2. Characterisation of Electrophysiological Properties

Although no ion channel-encoding genes are located on the X-chromosome, mutations in X-linked genes, such as MeCP2 [81], may indirectly affect cardiac electrophysiology; as such, it is essential to evaluate changes in electrophysiological properties in iPSC-derived cardiomyocytes when modelling X-linked cardiomyopathy. Patch clamp analysis has been used routinely to record the physiological properties of neurons and cardiomyocytes [82]. This method allows for the real-time measurement of changes in the membrane potential or electrical current of a single cell, a technique that requires great skill and is labour intensive and time consuming. Furthermore, since patch-clamp analysis works only on a single cell, it may not reveal the phenotypes related to conduction abnormality [83].
Multielectrode arrays (MEA) analysis is another means of investigating the electrophysiology of cardiomyocytes. It enables the investigation of the mutations that induce conduction system defects [84]. By culturing a monolayer of iPSC-derived cardiomyocytes on the MEA plate, voltage changes due to depolarisation and repolarisation of cardiomyocytes across the surface can be simultaneously recorded to reveal the action potential conduction velocity [85]. As well as MEA, optical mapping (OM) can be used to study disease phenotypes associated with propagation abnormalities [86]. Although this method was developed to visualise myocardial infarction-induced heart block in isolated animal hearts, it can be modified to record the conduction velocity across the monolayers of iPSC-derived cardiomyocytes (Figure 3A). Unlike the MEA system, the OM system does not directly record the voltage changes with electrodes; instead, a voltage-sensitive fluorochrome (e.g., Di-8-ANEPPS) is used as the voltage indicator to provide a highly sensitive recording of voltage changes across the cardiomyocytes (Figure 3B).

5.3. Characterisation of the Calcium Handling Properties

During excitation–contraction coupling, the release and reuptake of sarcomeric calcium relays the electric signalling and the consequent contraction of cardiomyocytes. Failure in the regulation or handling of intracellular calcium may significantly compromise cardiac function or lead to cardiomyocyte death [87]. Again, mutations in X-linked genes, such as LAMP2, can indirectly alter intracellular calcium handling [32]; thus, the study of intracellular calcium handling is important to understand the pathophysiology of X-linked cardiomyopathies. Currently, the most widely used method to study intracellular calcium handling involves the use of calcium-sensing fluorochromes (e.g., Fura-2-AM) to trace the intracellular calcium transient. Upon electrical stimulation, the release of intracellular calcium is indicated by the fluorescent emission that is recorded with an ultra-fast and sensitive camera [81].

5.4. Characterisations of Contraction Properties

Measurement of the contractile property of an iPSC-derived cardiomyocyte is a direct indication of the effects of a mutation on cardiac function [88]. This is especially important when evaluating the effects of mutations in the genes encoding cytoskeleton networks (e.g., FLNA). Contractile function can be evaluated by different methods. The IonOptix cardiomyocyte contraction system allows for a high-speed recording (1000 hz) of the shortening and re-lengthening of a single cardiomyocyte (Figure 4). Nevertheless, since iPSC-derived cardiomyocytes are usually of irregular shape and contract multi-directionally, some laboratories may prefer to measure the contractile force generated during the contraction of the iPSC-derived cardiomyocyte with anatomic force microscopy [89].

6. Strategies for Using iPSC-Based Models for Disease Modelling and Drug Testing

The iPSC-derived cardiomyocyte-based model is a simple, reproducible, and economically efficient platform for disease modelling and drug testing [90]. The iPSC-derived cardiomyocytes demonstrate characteristic electrophysiological functions and pharmacological responsiveness comparable to primary cardiomyocytes, and thus have been used extensively for disease modelling and drug testing (Table 2 and Table 3) [61]. Patient-specific iPSCs have been used to reproduce a myriad of cardiomyopathy phenotypes, including long QT syndrome, Brugada syndrome, and familial dilated cardiomyopathy [91,92,93]. The electrophysiological indices of different ion channel blockers have been assessed for their antiarrhythmic properties and proarrhythmic risks using an aniPSC-derived cardiomyocyte-based platform using a multi-electrode array system [94,95,96]. Precision medicine, such as gender and population-specific differences in drug responses, can also be tested with iPSC-derived cardiomyocytes [97].
Sub-population-specific cardiomyocytes can be used for disease modelling and drug testing in pathology involving specific subtypes only [98]. Atrial-selective pharmacology is a potential means to treat atrial fibrillation with fewer side effects; drugs specific to atrial ion channels have been shown to reduce early repolarisation of atrial but not ventricular cardiomyocytes [99]. Atrial-specific cardiomyocytes have been used to assess atrial selectivity and efficacy of antiarrhythmic drugs [99,100]. Drug-induced arrhythmia has also been tested with ventricular-like cardiomyocytes [101].
Traditional two-dimensional cell cultures have limited translational capacity to humans. Consequently, the development of human cardiac organoid technologies may mimic the tissue microenvironment and increase confidence in translatability in drug screening and disease modelling [102]. Some 3D cultures show a more mature electrophysiological function, comparable to the normal adult myocardium [103,104,105]. A 3D model from a patient-derived iPSC for cardio-facio-cutaneous syndrome demonstrated hypertrophic cardiomyopathy-related disease phenotypes in cardiac tissues that could not be shown with 2D cultures [106]. Other conditions, such as Barth syndrome, inherited dilated cardiomyopathy, and arrhythmia, have also been modelled using 3D organoids [107,108,109]. An iPSC-derived 3D human engineered cardiac tissue model has also been used to study the potential of gene therapy in restoring the dystrophin expression and mechanical contraction force in DMD [110].
Furthermore, iPSC-derived cardiomyocyte-based cardiac repair is a promising treatment for the injured heart. Allogenic transplantation of iPSC-derived cardiomyocytes has been shown to improve cardiac contractile function and attenuate ventricular dilation in monkeys, mice, and rats with myocardial infarction, without inducing immune rejection [111,112,113]. Infarct size and apoptosis have both been reported to be reduced with the transplantation of patches of iPSC-derived cardiomyocytes that have also been shown to have additional cytoprotective and angiogenesis effects in order to improve the survival of damaged myocardium [114,115]. The improved cardiac function is likely due to the formation of mature grafts by injected iPSC-CMs that continuously express cardiac markers and sarcomere structures [116].

6.1. Duchenne Muscular Dystrophy

Duchenne muscular dystrophy (DMD) is an X-linked condition affecting 19.8 males in every 100,000 and resulting in progressive muscle wasting [132]. It is caused by a DMD gene mutation [133]. The DMD gene is located on the short arm of the X-chromosome and codes for the dystrophin protein. Frame shifting or non-sense mutations of the gene code for a truncated non-functional dystrophin protein without its cysteine domain [134]. The deficient dystrophin cannot act as a cohesive protein and results in the disaggregation of dystrophin-associated protein complex (DAPC), ultimately resulting in an impaired structure and reduced durability of the myofibres [133,135].
Most patients with DMD will ultimately develop dilated cardiomyopathy (DCM), one of the main causes of DMD-associated mortality [136]. Indeed, DMD-associated DCM is responsible for 2% of inherited DCM, often resulting in heart failure and ventricular arrhythmia [137]. It is characterised by inferolateral left ventricle regional contractile impairment and replacement fibrosis [138]. Its gradual progression with increased muscle stiffness and ventricular dilation ultimately results in more severe heart failure. The prognosis of the DMD patient with DCM is poor, with limited therapeutic options [139].
Several iPSC-based models have been established to recapitulate the pathophysiology of DMD-associated DCM, including calcium handling changes, impairment of myofibrils, and membrane fragility. After repeated stretch–contraction cycles, it was observed that the cell morphology was altered due to rupture of the sarcolemma. Compared with patient-derived tissues, iPSC models permit investigation of the early-stage pathogenesis of DMD and the different gene mutations associated with DMD-associated DCM, including Δ Exon 50, Δ Exons 45–52, c.5899C>T, and more [81,120,140].
The in vivo model can recapitulate the disease phenotypes and provide a more synoptic picture of the pathophysiology. The in vitro cellular models show the molecular changes observed in the tissues of the DMD patient, including the DMD signature dystrophin deficiency and the downstream changes, such as increased oxidative stress, impaired autophagy, and immunoproteasome dysregulation [118,140,141,142]. Studies show that the overall ATP production and basal respiration rate are decreased in the disease cell lines [117]. Furthermore, the yes-associated protein (YAP) is also modified due to altered actin stress fibers, leading to reduced proliferation of the iPSCs-derived cardiomyocytes. Cytosolic calcium has been found to be overloaded with increased intracellular Ca2+ concentration, especially after mechanical stretching [118]. Utilising high-throughput methods, muscle genes and mitochondrial metabolism genes have been identified to be dysregulated from the somite stage [119].
The molecular pathogenic profile can also be linked to the cardiac cardiomyopathic and electrophysiologic phenotype observed in affected patients. Patients with DMD-associated cardiomyopathy frequently suffer supraventricular and ventricular arrhythmias [143]. The diseased cell line has been shown to exhibit slower firing rates and increased events of delayed after depolarisations (DADs) and oscillatory prepotentials (OPPs). The length of the repolarisation also increased. This is especially the case in affected males, due to the increased ICa,L density. This explains the arrhythmic pattern observed in DMD patients [120].
After identifying the pathophysiological changes, novel drug targets can be identified to alleviate the disease phenotype. In a study with Δ exons 49–50, a proteasome inhibitor was applied to rescue the immunoproteasome dysregulation observed in iPSC-derived cardiomyocytes [7]. In another study of Δ exons 45–52, a membrane sealant Polaxamer 188 was able to restore the normal phenotype upon understanding how the damaged membrane could lead to dysregulated mechanosensitive channels [81,144]. Furthermore, the advancement of gene therapy may also enable the correction of the DMD phenotype. Human artificial chromosomes (HACs) have been shown to restore the dystrophin expression level and the normal phenotype [89].

6.2. Fabry Disease

Fabry disease is an X-linked multi-system lysosomal storage disorder with an incidence of around 0.04 to 0.25 per 10,000 males, and a higher incidence for the later-onset variant [145,146,147]. The disease mainly affects males, and heterozygous females have a spectrum of symptoms [148]. The causal mutation in Fabry disease is located at the GLA site and results in a defect or deficiency in the production of the alpha-galactosidase A enzyme (a-GAL A). a-GAL A is a lysosomal enzyme responsible for the removal of terminal alpha-galactosyl groups in glycolipids, particularly globotriaosylceramide (Gb3). Deficiency leads to the consequent accumulation of glycolipids in various tissues [149]. Progressive storage of these glycolipids leads to cellular dysfunction and irreversible damage, resulting in symptoms such as neuropathic pain, left ventricular hypertrophy, renal impairment, and angiokeratoma [150]. Some patients have the atypical late-onset cardiac variant of Fabry disease that presents as unexplained left ventricular hypertrophy without previous symptoms in middle-aged males [151]. Other cardiac manifestations include coronary insufficiency, valvular involvement, arrhythmia, and atrioventricular conduction problems [152]. There is no cure for Fabry disease, and the current standard of care is enzyme replacement therapy with intravenous administration of synthetic a-GAL A every other week [153].
Patient iPSC-derived cardiomyocytes can be used to model the disease and screen potential drugs [154]. Early attempts at modelling with iPSCs from patient fibroblasts generated a characteristic accumulation of Gb3 and pathological ultrastructural features [155]. Kuramoto Y. et al. [122] further isolated iPSCs from female patients and discovered individual clones that demonstrated either deficient or normal enzyme activity, that were then respectively used as the disease model and the isogenic control. Gb3 accumulation can be identified by mass spectrometry, and the expression of ANP in diseased iPSCs is increased, indicative of hypertrophy. Others have also generated iPSC-derived cardiomyocytes with decreased contractility, cellular hypertrophy, disturbed ion channel electrical currents, and peripheral displacement of cardiac contractile myofibrils, the typical phenotypic changes in diseased patients [121,128,156]. Currently, we are working with iPSCs established from a patient with Fabry disease. In brief, this patient contains a GLA-IVS4–919-G>A mutation with a significantly reduced a-GAL A protein level in the resultant cardiomyocytes (Figure 5).
Drug testing with patient iPSC-derived cardiomyocytes is made possible with the development of high-throughput compound screening [157]. No established data or drug testing research have been performed, but many studies have demonstrated the potential of the patient iPSC model for drug testing in Fabry disease. After identification of pathological phenotypes in mutant iPSCs, candidate drugs can be tested for normalisation or minimisation of these phenotypic changes and the drug toxicity can be assessed [158,159]. Several molecules, such as interleukin-18 and arachidonate 12/15-lipoxygenase, have been demonstrated in Fabry iPSC-derived cardiomyocytes to be associated with disease progression, and respective inhibitors have been proposed to have an adjunctive effect with enzyme replacement therapy [123,160]. Future investigations could make use of an iPSC-based model to test the effectiveness of these inhibitors in the alleviation of Fabry phenotypes.

6.3. Rett Syndrome

Rett syndrome is a monogenic X-linked dominant neurological disorder affecting 0.44 to 0.72 per 10,000 females aged 2 to 18 years, with similar rates across different races and ethnicities [126,161,162]. The disease mainly affects females, as mutations in hemizygous males often lead to embryonic or neonatal death in the first year of life [125]. Females with typical Rett syndrome initially develop normally after an uneventful full-term delivery but after several months experience loss of speech, purposeful hand movements, gait abnormalities, postnatal microcephaly, stereotypic hand movements, seizures, autistic features, and respiratory, cardiac, and nervous system abnormalities. Neurological development arrests by the age of one [156]. Some patients have atypical Rett syndrome that does not include all of the classic phenotypes seen in the typical form. Approximately 90% of the typical Rett syndrome and the milder preserved speech variant are due to a loss-of-function mutation in the MECP2 gene, located in the long arm of the X-chromosome (Xq28) [163]. The MECP2 gene encodes for methyl-CpG binding protein 2, which is abundantly expressed in the brain as an epigenetic modulator to control gene expression through DNA methylation [164]. A minority of patients with atypical Rett syndrome have mutations in the cyclin-dependent kinase-like 5 (CDKL5), Forkhead box protein G1 (FOXG1), myocyte-specific enhancer factor 2C (MEF2C), and transcription factor 4 (TCF4) genes [165].
Patients with Rett syndrome are at an increased risk of life-threatening cardiac arrhythmia with prolongation of QT corrected interval and alternations of ventricular repolarisation with the possibility of sudden unexpected cardiac death [166]. Other cardiac phenotypes include reduced cardiac vagal tone, sympatho-vagal imbalance, and heart rate variability [161].
Patient fibroblast-derived iPSCs have been used for disease modelling and drug testing in Rett syndrome. Rett-iPSCs from female patients retained the ability to undergo X-inactivation and differentiation into neurons. Compared with their wildtype controls, these diseased iPSC-generated neurons were smaller in size, and had fewer glutamatergic synapses, reduced spine density, smaller soma size, altered calcium signalling, and electrophysiological defects [128,162]. Diseased iPSC-derived neural cultures also resulted in changes to exosome protein cargo and signalling bioactivity [129]. The iPSCs generated from atypical Rett syndrome patients with a CDKL5 mutation exhibited a decrease in axon outgrowth, dendritic morphogenesis, and synapse formation in diseased neurons [167].
A patient-derived iPSC model can also be used to screen the effects of candidate drugs by observation of the reversal of Rett disease phenotypes. IGF-1 and gentamicin administration have been found to be effective in the partial reversal of Rett-like symptoms in a cell culture, similar to results in a previously described mouse model [128]. Administration of control exosomes also rescues neurodevelopmental deficits seen in Rett syndrome iPSC models [129]. Selective inhibitors of histone deacetylase improved GABAergic circuit disruption and acetylated a-tubulin defects present in iPSC-derived neurons [130]. Traditional Indian Medhya Rasayana herbs have also been shown to reverse neurological disorders when tested in iPSC models [131].

6.4. Danon Disease

Danon disease is an X-linked dominant lethal multi-systemic condition that affects 1–4% of patients with hypertrophic cardiomyopathy [168]. It is caused by the mutation of the X-linked LAMP2 gene that encodes lysosomal-associated membrane protein 2 (LAMP2). Deficiency in LAMP2 protein production leads to the failed fusion of autophagosomes and lysosomes, with the consequent accumulation of glycogen in muscle cells [169]. Male Danon disease patients manifest with myopathy, retinopathy, cognitive impairment, and hypertrophic cardiomyopathy (HCM) that ultimately results in sudden death in the absence of heart transplantation [170,171]. In heterozygous females, the age of onset of the disease is 15 years later than in males, and they mainly present with cardiac-restricted manifestations, including cardiomyopathy and arrhythmia [172].
To this end, several Danon disease patient-specific iPSC-based models have been established. Using this model, we have previously demonstrated that the accumulation of autophagy materials inside the affected cardiomyocytes is mainly due to the failure to clear the autophagic cargo, rather than over-activation of the autophagy signalling pathway [32].
In addition, increased mitochondrial oxidative stress, impaired mitophagy flux, and increased apoptosis have also been observed in iPSC-derived cardiomyocytes carrying the LAMP2 mutant [173]. This results in the reduced OCR of basal respiration, ATP production, and maximal respiratory capacity (MRC) in the Danon disease iPSC-derived cardiomyocytes [174]. Treatment of the diseased iPSC-derived cardiomyocytes with antioxidant N-acetylcysteine has been shown to reduce the level of oxidative stress and alleviate the mitochondrial fragmentation pattern observed in the diseased cell line. Given the low rate of ATP generation, Danon disease iPSC-derived cardiomyocytes switched to utilising the glycolysis pathway to sustain their metabolism [175]. The Danon disease iPSC-derived cardiomyocytes exhibited downregulation of protein synthesis, potentially originating from reduced ATP synthesis and increased unfolded protein responses [175,176].
Danon disease iPSC-derived cardiomyocytes also exhibit diseased cardiac phenotypes, demonstrating an impaired contractile function similar to hypertrophic cardiomyopathy. It was shown that the contractility was decreased [32], and the contractile force per post was lower than that of control cell lines. The contractile force has been shown to return to normal after correction of the LAMP2 mutation (c.247C>T) using the CRISPR-cas9system [174]. The LAMP2-deficient iPSC-derived cardiomyocytes exhibit a reduced amplitude of intracellular calcium transient and reduced maximum velocity of calcium release and reuptake. This is due to the reduced cellular level of calcium handling proteins in the mutant line, such as sodium–calcium exchanger (NCX1), sacro/endoplasmic reticulum Ca2+-ATPase 2, and ryanodine receptor 2 (RYR2) [32]. The Danon disease iPSC-CM also demonstrates traits of heart failure, such as significantly increased expression of ANP and BNP in the diseased cell line [127].
As the disease is due to phenotypic demonstration of the diseased allele, activating the previously silenced normal X-chromosome with the wildtype LAMP2 gene will rescue the diseased phenotype in the heterozygous female Danon disease patient. To prove this concept, we have created isogenic iPSCs from female Danon disease patients. By reactivating the wildtype LAMP2 allele by demethylation, we were able to restore LAMP2 protein production and improve the autophagic functions in the affected cells [32].

7. Isogenic iPSCs as a New Platform to Study X-Chromosome Inactivation

As mentioned before, isogenic iPSC-pairs with distinctive phenotypes can now be generated from a heterozygous female carrying one X-linked wildtype allele and one X-linked mutant allele. Making use of such a system, we have established an isogenic-iPSC-based platform to screen chemicals that can activate the silenced X-chromosome. In brief, as demonstrated previously [32], when the LAMP2-deficient iPSC carrying a wildtype LAMP2 allele on the silenced X-chromosome was treated with a drug that could reactivate the silenced X-chromosome, cells started to produce a functional LAMP2 protein that could be detected by a simple immunoassay (Figure 6). Using this system, we have screened a library of approximately 300 chemicals and identified several inhibitors of the JAK2 signalling pathway (e.g., LY2784544 and BMS-911543) that can reactivate the genes on the silenced X-chromosome. This result not only facilitates the development of therapeutic targets for X-linked disorders, but also reveals the signalling pathways that may govern the randomised X-chromosome inactivation.

8. Limitations

Cardiomyocytes derived from patient-specific iPSCs offer an exclusive and convenient means to model disease phenotypes that are associated with X-linked mutations. Nonetheless, the functional immaturity of the iPSC-derived cardiomyocytes remains an important issue to be addressed. As demonstrated earlier by our group, compared with human ESC-derived cardiomyocytes, human iPSC-derived cardiomyocytes usually exhibit a reduced calcium handling ability and sarcoplasmic reticulum function [177]. This is due to the immature nature of the human iPSC-derived cardiomyocytes [178]. Human iPSC-derived cardiomyocytes display a lack of T-tubules and poor colocalisation of calcium channels and ryanodine receptors [179]. Therefore, extra attention should be paid when one considers modelling cardiomyopathies that resulted from the gene mutation that may affect calcium transients, such as DMD [180]. The immature iPSC-derived cardiomyocytes also demonstrate some different electrophysiology, cell morphology, and metabolism. As such, there is a need to improve the maturity of the iPSC-derived cardiomyocytes; this can be achieved with a prolonged culturing time [181], physical stimulation [182,183], biochemical stimulation [184], and metabolic alterations [185]. Moreover, since the iPSC-derived cardiomyocytes area contains a mixed population of atrial, ventricular, and nodal-like cells, these cells may differ substantially in their electrophysiological properties despite being genetically identical. Thus, for modelling cardiomyopathies associated with ion channel abnormalities, one may need to consider the contribution of the different cardiomyocyte subtypes to the results obtained or derive sub-population specific cardiomyocytes in the differentiation process. Similarly, when testing the effects of a putative antiarrhythmic drug, the immature phenotype and mixed cardiomyocyte subtypes should also be taken into account. Finally, although the use of isogenic iPSCs may help to evaluate the effects of a particular mutation under the same genetic background, the results may still not be sufficient to help predict prognosis and disease severity in heterozygous female patients.

Author Contributions

J.Z., O.H.-I.C., Y.-L.T. and K.-M.N., writing—original draft and figure preparation; H.-F.T., writing—review and editing; K.-M.N. and H.-F.T., funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Food and Health bureau, Hong Kong (the Health and medical research fund: EH 472/1–1); Research grant council, Hong Kong (general research funds: 17111320 and 17106917) to K.N. and H.T. and the Innovation Technology Commission Funding to H.T.

Institutional Review Board Statement

The study was conducted according to the guidelines of the Declaration of Helsinki and approved by the Institutional Review Board of HKWC-HKU Cluster Research Ethics Committee/Institutional Review Board (“REC/IRB”) (protocol code 05161936, approved on 11 January 2018).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Milani-Nejad, N.; Janssen, P.M. Small and large animal models in cardiac contraction research: Advantages and disadvantages. Pharm. 2014, 141, 235–249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Nerbonne, J.M. Studying cardiac arrhythmias in the mouse--a reasonable model for probing mechanisms? Trends Cardiovasc. Med. 2004, 14, 83–93. [Google Scholar] [CrossRef] [PubMed]
  3. Sutanto, H.; Laudy, L.; Clerx, M.; Dobrev, D.; Crijns, H.; Heijman, J. Maastricht antiarrhythmic drug evaluator (manta): A computational tool for better understanding of antiarrhythmic drugs. Pharm. Res. 2019, 148, 104444. [Google Scholar] [CrossRef] [PubMed]
  4. Alpert, N.R.; Brosseau, C.; Federico, A.; Krenz, M.; Robbins, J.; Warshaw, D.M. Molecular mechanics of mouse cardiac myosin isoforms. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H1446–H1454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Takahashi, K.; Tanabe, K.; Ohnuki, M.; Narita, M.; Ichisaka, T.; Tomoda, K.; Yamanaka, S. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 2007, 131, 861–872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Nakagawa, M.; Koyanagi, M.; Tanabe, K.; Takahashi, K.; Ichisaka, T.; Aoi, T.; Okita, K.; Mochiduki, Y.; Takizawa, N.; Yamanaka, S. Generation of induced pluripotent stem cells without myc from mouse and human fibroblasts. Nat. Biotechnol. 2008, 26, 101–106. [Google Scholar] [CrossRef]
  7. Farini, A.; Gowran, A.; Bella, P.; Sitzia, C.; Scopece, A.; Castiglioni, E.; Rovina, D.; Nigro, P.; Villa, C.; Fortunato, F.; et al. Fibrosis rescue improves cardiac function in dystrophin-deficient mice and duchenne patient-specific cardiomyocytes by immunoproteasome modulation. Am. J. Pathol. 2019, 189, 339–353. [Google Scholar] [CrossRef] [Green Version]
  8. Peper, J.; Kownatzki-Danger, D.; Weninger, G.; Seibertz, F.; Pronto, J.R.D.; Sutanto, H.; Pacheu-Grau, D.; Hindmarsh, R.; Brandenburg, S.; Kohl, T.; et al. Caveolin3 stabilizes mct1-mediated lactate/proton transport in cardiomyocytes. Circ. Res. 2021, 128, e102–e120. [Google Scholar] [CrossRef]
  9. Brambatti, M.; Caspi, O.; Maolo, A.; Koshi, E.; Greenberg, B.; Taylor, M.R.G.; Adler, E.D. Danon disease: Gender differences in presentation and outcomes. Int. J. Cardiol. 2019, 286, 92–98. [Google Scholar] [CrossRef]
  10. Jefferies, J.L.; Eidem, B.W.; Belmont, J.W.; Craigen, W.J.; Ware, S.M.; Fernbach, S.D.; Neish, S.R.; Smith, E.O.; Towbin, J.A. Genetic predictors and remodeling of dilated cardiomyopathy in muscular dystrophy. Circulation 2005, 112, 2799–2804. [Google Scholar] [CrossRef] [Green Version]
  11. Rivera, S.R.; Jhamb, S.K.; Abdel-Hamid, H.Z.; Acsadi, G.; Brandsema, J.; Ciafaloni, E.; Darras, B.T.; Iannaccone, S.T.; Konersman, C.G.; Kuntz, N.L.; et al. Medical management of muscle weakness in duchenne muscular dystrophy. PLoS ONE 2020, 15, e0240687. [Google Scholar] [CrossRef]
  12. Eiholzer, U.; Boltshauser, E.; Frey, D.; Molinari, L.; Zachmann, M. Short stature: A common feature in duchenne muscular dystrophy. Eur. J. Pediatr. 1988, 147, 602–605. [Google Scholar] [CrossRef] [PubMed]
  13. Ommen, S.R.; Nishimura, R.A.; Edwards, W.D. Fabry disease: A mimic for obstructive hypertrophic cardiomyopathy? Heart 2003, 89, 929–930. [Google Scholar] [CrossRef] [Green Version]
  14. Bokhari, S.R.A.; Zulfiqar, H.; Hariz, A. Fabry disease; Statpearls: Treasure Island, FL, USA, 2021. [Google Scholar]
  15. Kyndt, F.; Gueffet, J.P.; Probst, V.; Jaafar, P.; Legendre, A.; Le Bouffant, F.; Toquet, C.; Roy, E.; McGregor, L.; Lynch, S.A.; et al. Mutations in the gene encoding filamin a as a cause for familial cardiac valvular dystrophy. Circulation 2007, 115, 40–49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Van Kogelenberg, M.; Clark, A.R.; Jenkins, Z.; Morgan, T.; Anandan, A.; Sawyer, G.M.; Edwards, M.; Dudding, T.; Homfray, T.; Castle, B.; et al. Diverse phenotypic consequences of mutations affecting the c-terminus of flna. J. Mol. Med. 2015, 93, 773–782. [Google Scholar] [CrossRef]
  17. Monteleone, P.L.; Fagan, L.F. Possible x-linked congenital heart disease. Circulation 1969, 39, 611–614. [Google Scholar] [CrossRef] [Green Version]
  18. Ritelli, M.; Morlino, S.; Giacopuzzi, E.; Carini, G.; Cinquina, V.; Chiarelli, N.; Majore, S.; Colombi, M.; Castori, M. Ehlers-danlos syndrome with lethal cardiac valvular dystrophy in males carrying a novel splice mutation in flna. Am. J. Med. Genet. A 2017, 173, 169–176. [Google Scholar] [CrossRef]
  19. Arad, M.; Maron, B.J.; Gorham, J.M.; Johnson, W.H., Jr.; Saul, J.P.; Perez-Atayde, A.R.; Spirito, P.; Wright, G.B.; Kanter, R.J.; Seidman, C.E.; et al. Glycogen storage diseases presenting as hypertrophic cardiomyopathy. N. Engl. J. Med. 2005, 352, 362–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Prall, F.R.; Drack, A.; Taylor, M.; Ku, L.; Olson, J.L.; Gregory, D.; Mestroni, L.; Mandava, N. Ophthalmic manifestations of danon disease. Ophthalmology 2006, 113, 1010–1013. [Google Scholar] [CrossRef] [PubMed]
  21. Yardeni, M.; Weisman, O.; Mandel, H.; Weinberger, R.; Quarta, G.; Salazar-Mendiguchia, J.; Garcia-Pavia, P.; Lobato-Rodriguez, M.J.; Simon, L.F.; Dov, F.; et al. Psychiatric and cognitive characteristics of individuals with danon disease (lamp2 gene mutation). Am. J. Med. Genet. A 2017, 173, 2461–2466. [Google Scholar] [CrossRef]
  22. Gourzi, P.; Pantou, M.P.; Gkouziouta, A.; Kaklamanis, L.; Tsiapras, D.; Zygouri, C.; Constantoulakis, P.; Adamopoulos, S.; Degiannis, D. A new phenotype of severe dilated cardiomyopathy associated with a mutation in the lamp2 gene previously known to cause hypertrophic cardiomyopathy in the context of danon disease. Eur. J. Med. Genet. 2019, 62, 77–80. [Google Scholar] [CrossRef]
  23. Codron, P.; Pautot, V.; Tassin, A.; Sternberg, D.; Letournel, F.; Richard, P.; Nadaj-Pakleza, A. Abundant electrical myotonia and left ventricular noncompaction: Unusual features of danon disease due to a novel mutation in lamp2 gene. Rev. Neurol. 2019, 175, 201–203. [Google Scholar] [CrossRef] [PubMed]
  24. Amir, R.E.; Van den Veyver, I.B.; Wan, M.; Tran, C.Q.; Francke, U.; Zoghbi, H.Y. Rett syndrome is caused by mutations in x-linked mecp2, encoding methyl-cpg-binding protein 2. Nat. Genet. 1999, 23, 185–188. [Google Scholar] [CrossRef]
  25. Hagberg, B.; Aicardi, J.; Dias, K.; Ramos, O. A progressive syndrome of autism, dementia, ataxia, and loss of purposeful hand use in girls: Rett’s syndrome: Report of 35 cases. Ann. Neurol. 1983, 14, 471–479. [Google Scholar] [CrossRef]
  26. Dolce, A.; Ben-Zeev, B.; Naidu, S.; Kossoff, E.H. Rett syndrome and epilepsy: An update for child neurologists. Pediatr. Neurol. 2013, 48, 337–345. [Google Scholar] [CrossRef] [PubMed]
  27. Leonard, H.; Thomson, M.R.; Glasson, E.J.; Fyfe, S.; Leonard, S.; Bower, C.; Christodoulou, J.; Ellaway, C. A population-based approach to the investigation of osteopenia in rett syndrome. Dev. Med. Child. Neurol. 1999, 41, 323–328. [Google Scholar] [CrossRef] [PubMed]
  28. Agrawal, P.B.; Pierson, C.R.; Joshi, M.; Liu, X.; Ravenscroft, G.; Moghadaszadeh, B.; Talabere, T.; Viola, M.; Swanson, L.C.; Haliloglu, G.; et al. Speg interacts with myotubularin, and its deficiency causes centronuclear myopathy with dilated cardiomyopathy. Am. J. Hum. Genet. 2014, 95, 218–226. [Google Scholar] [CrossRef] [Green Version]
  29. Dowling, J.J.; Lawlor, M.W.; Das, S. X-linked myotubular myopathy. In GeneReviews®; Adam, M.P., Ardinger, H.H., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Mirzaa, G., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  30. Zakrzewski, W.; Dobrzynski, M.; Szymonowicz, M.; Rybak, Z. Stem cells: Past, present, and future. Stem. Cell Res. Ther. 2019, 10, 68. [Google Scholar] [CrossRef]
  31. Hu, S.; Zhao, M.T.; Jahanbani, F.; Shao, N.Y.; Lee, W.H.; Chen, H.; Snyder, M.P.; Wu, J.C. Effects of cellular origin on differentiation of human induced pluripotent stem cell-derived endothelial cells. JCI Insight 2016, 1. [Google Scholar] [CrossRef]
  32. Ng, K.M.; Mok, P.Y.; Butler, A.W.; Ho, J.C.; Choi, S.W.; Lee, Y.K.; Lai, W.H.; Au, K.W.; Lau, Y.M.; Wong, L.Y.; et al. Amelioration of x-linked related autophagy failure in danon disease with DNA methylation inhibitor. Circulation 2016, 134, 1373–1389. [Google Scholar] [CrossRef]
  33. Ghanbari, H.A.; McCarl, R.L. A comparative study of myosin and its subunits in adult and neonatal-rat hearts and in rat heart cells from young and old cultures. Biochem. J. 1980, 191, 627–635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Peter, A.K.; Bjerke, M.A.; Leinwand, L.A. Biology of the cardiac myocyte in heart disease. Mol. Biol. Cell 2016, 27, 2149–2160. [Google Scholar] [CrossRef] [PubMed]
  35. Xue, H.; Wu, J.; Li, S.; Rao, M.S.; Liu, Y. Genetic modification in human pluripotent stem cells by homologous recombination and crispr/cas9 system. Methods Mol. Biol. 2016, 1307, 173–190. [Google Scholar] [PubMed]
  36. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome engineering using the crispr-cas9 system. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. De Masi, C.; Spitalieri, P.; Murdocca, M.; Novelli, G.; Sangiuolo, F. Application of crispr/cas9 to human-induced pluripotent stem cells: From gene editing to drug discovery. Hum. Genom. 2020, 14, 25. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, Z.; Zhang, Y.; Gao, F.; Han, S.; Cheah, K.S.; Tse, H.F.; Lian, Q. Crispr/cas9 genome-editing system in human stem cells: Current status and future prospects. Mol. Ther. Nucleic. Acids 2017, 9, 230–241. [Google Scholar] [CrossRef] [Green Version]
  39. Sterneckert, J.L.; Reinhardt, P.; Scholer, H.R. Investigating human disease using stem cell models. Nat. Rev. Genet. 2014, 15, 625–639. [Google Scholar] [CrossRef] [PubMed]
  40. Hockemeyer, D.; Jaenisch, R. Induced pluripotent stem cells meet genome editing. Cell Stem Cell 2016, 18, 573–586. [Google Scholar] [CrossRef] [Green Version]
  41. McCauley, H.A.; Wells, J.M. Pluripotent stem cell-derived organoids: Using principles of developmental biology to grow human tissues in a dish. Development 2017, 144, 958–962. [Google Scholar] [CrossRef] [Green Version]
  42. Machiraju, P.; Greenway, S.C. Current methods for the maturation of induced pluripotent stem cell-derived cardiomyocytes. World J. Stem Cells 2019, 11, 33–43. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, J.; Wang, Y.; Zhou, T.; Wong, L.Y.; Tian, X.Y.; Hong, X.; Lai, W.H.; Au, K.W.; Wei, R.; Liu, Y.; et al. Generation of human liver chimeric mice with hepatocytes from familial hypercholesterolemia induced pluripotent stem cells. Stem Cell Rep. 2017, 8, 605–618. [Google Scholar] [CrossRef] [Green Version]
  44. Vitrac, A.; Pons, S.; Balkota, M.; Lemiere, N.; Rais, C.; Bourgeois, J.P.; Maskos, U.; Bourgeron, T.; Cloez-Tayarani, I. A chimeric mouse model to study human ipsc-derived neurons: The case of a truncating shank3 mutation. Sci. Rep. 2020, 10, 13315. [Google Scholar] [CrossRef]
  45. Yu, J.; Hu, K.; Smuga-Otto, K.; Tian, S.; Stewart, R.; Slukvin, I.I.; Thomson, J.A. Human induced pluripotent stem cells free of vector and transgene sequences. Science 2009, 324, 797–801. [Google Scholar] [CrossRef] [Green Version]
  46. El Hokayem, J.; Cukier, H.N.; Dykxhoorn, D.M. Blood derived induced pluripotent stem cells (ipscs): Benefits, challenges and the road ahead. J. Alzheimers Dis. Parkinsonism 2016, 6. [Google Scholar] [CrossRef] [Green Version]
  47. Spitzhorn, L.-S.; Megges, M.; Wruck, W.; Rahman, M.S.; Otte, J.; Degistirici, Ö.; Meisel, R.; Sorg, R.V.; Oreffo, R.O.C.; Adjaye, J. Human ipsc-derived mscs (imscs) from aged individuals acquire a rejuvenation signature. Stem Cell Res. Ther. 2019, 10, 100. [Google Scholar] [CrossRef] [Green Version]
  48. Zhou, T.; Benda, C.; Duzinger, S.; Huang, Y.; Li, X.; Li, Y.; Guo, X.; Cao, G.; Chen, S.; Hao, L.; et al. Generation of induced pluripotent stem cells from urine. J. Am. Soc. Nephrol. 2011, 22, 1221–1228. [Google Scholar] [CrossRef] [PubMed]
  49. Sommer, C.A.; Stadtfeld, M.; Murphy, G.J.; Hochedlinger, K.; Kotton, D.N.; Mostoslavsky, G. Induced pluripotent stem cell generation using a single lentiviral stem cell cassette. Stem Cells 2009, 27, 543–549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Yamasaki, S.; Taguchi, Y.; Shimamoto, A.; Mukasa, H.; Tahara, H.; Okamoto, T. Generation of human induced pluripotent stem (ips) cells in serum- and feeder-free defined culture and tgf-beta1 regulation of pluripotency. PLoS ONE 2014, 9, e87151. [Google Scholar] [CrossRef]
  51. Ghaedi, M.; Niklason, L.E. Human pluripotent stem cells (ipsc) generation, culture, and differentiation to lung progenitor cells. Methods Mol. Biol. 2019, 1576, 55–92. [Google Scholar] [PubMed]
  52. Kehat, I.; Kenyagin-Karsenti, D.; Snir, M.; Segev, H.; Amit, M.; Gepstein, A.; Livne, E.; Binah, O.; Itskovitz-Eldor, J.; Gepstein, L. Human embryonic stem cells can differentiate into myocytes with structural and functional properties of cardiomyocytes. J. Clin. Invest. 2001, 108, 407–414. [Google Scholar] [CrossRef] [PubMed]
  53. Segev, H.; Kenyagin-Karsenti, D.; Fishman, B.; Gerecht-Nir, S.; Ziskind, A.; Amit, M.; Coleman, R.; Itskovitz-Eldor, J. Molecular analysis of cardiomyocytes derived from human embryonic stem cells. Dev. Growth Differ. 2005, 47, 295–306. [Google Scholar] [CrossRef]
  54. Blazeski, A.; Zhu, R.; Hunter, D.W.; Weinberg, S.H.; Boheler, K.R.; Zambidis, E.T.; Tung, L. Electrophysiological and contractile function of cardiomyocytes derived from human embryonic stem cells. Prog. Biophys. Mol. Biol. 2012, 110, 178–195. [Google Scholar] [CrossRef] [Green Version]
  55. Kattman, S.J.; Witty, A.D.; Gagliardi, M.; Dubois, N.C.; Niapour, M.; Hotta, A.; Ellis, J.; Keller, G. Stage-specific optimization of activin/nodal and bmp signaling promotes cardiac differentiation of mouse and human pluripotent stem cell lines. Cell Stem Cell 2011, 8, 228–240. [Google Scholar] [CrossRef] [Green Version]
  56. Burridge, P.W.; Thompson, S.; Millrod, M.A.; Weinberg, S.; Yuan, X.; Peters, A.; Mahairaki, V.; Koliatsos, V.E.; Tung, L.; Zambidis, E.T. A universal system for highly efficient cardiac differentiation of human induced pluripotent stem cells that eliminates interline variability. PLoS ONE 2011, 6, e18293. [Google Scholar] [CrossRef]
  57. Zhu, W.Z.; Van Biber, B.; Laflamme, M.A. Methods for the derivation and use of cardiomyocytes from human pluripotent stem cells. Methods Mol. Biol. 2011, 767, 419–431. [Google Scholar]
  58. Jha, R.; Xu, R.H.; Xu, C. Efficient differentiation of cardiomyocytes from human pluripotent stem cells with growth factors. Methods Mol. Biol. 2015, 1299, 115–131. [Google Scholar]
  59. Yang, L.; Soonpaa, M.H.; Adler, E.D.; Roepke, T.K.; Kattman, S.J.; Kennedy, M.; Henckaerts, E.; Bonham, K.; Abbott, G.W.; Linden, R.M.; et al. Human cardiovascular progenitor cells develop from a kdr+ embryonic-stem-cell-derived population. Nature 2008, 453, 524–528. [Google Scholar] [CrossRef]
  60. Willems, E.; Cabral-Teixeira, J.; Schade, D.; Cai, W.; Reeves, P.; Bushway, P.J.; Lanier, M.; Walsh, C.; Kirchhausen, T.; Izpisua Belmonte, J.C.; et al. Small molecule-mediated tgf-beta type ii receptor degradation promotes cardiomyogenesis in embryonic stem cells. Cell Stem Cell 2012, 11, 242–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Ren, Y.; Lee, M.Y.; Schliffke, S.; Paavola, J.; Amos, P.J.; Ge, X.; Ye, M.; Zhu, S.; Senyei, G.; Lum, L.; et al. Small molecule wnt inhibitors enhance the efficiency of bmp-4-directed cardiac differentiation of human pluripotent stem cells. J. Mol. Cell Cardiol. 2011, 51, 280–287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Hamad, S.; Derichsweiler, D.; Papadopoulos, S.; Nguemo, F.; Saric, T.; Sachinidis, A.; Brockmeier, K.; Hescheler, J.; Boukens, B.J.; Pfannkuche, K. Generation of human induced pluripotent stem cell-derived cardiomyocytes in 2d monolayer and scalable 3d suspension bioreactor cultures with reduced batch-to-batch variations. Theranostics 2019, 9, 7222–7238. [Google Scholar] [CrossRef] [PubMed]
  63. 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. Chemically defined generation of human cardiomyocytes. Nat. Methods 2014, 11, 855–860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Van der Valk, J.; Brunner, D.; De Smet, K.; Fex Svenningsen, A.; Honegger, P.; Knudsen, L.E.; Lindl, T.; Noraberg, J.; Price, A.; Scarino, M.L.; et al. Optimization of chemically defined cell culture media--replacing fetal bovine serum in mammalian in vitro methods. Toxicol Vitr. 2010, 24, 1053–1063. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Dambrot, C.; Braam, S.R.; Tertoolen, L.G.; Birket, M.; Atsma, D.E.; Mummery, C.L. Serum supplemented culture medium masks hypertrophic phenotypes in human pluripotent stem cell derived cardiomyocytes. J. Cell Mol. Med. 2014, 18, 1509–1518. [Google Scholar] [CrossRef] [PubMed]
  66. Tohyama, S.; Hattori, F.; Sano, M.; Hishiki, T.; Nagahata, Y.; Matsuura, T.; Hashimoto, H.; Suzuki, T.; Yamashita, H.; Satoh, Y.; et al. Distinct metabolic flow enables large-scale purification of mouse and human pluripotent stem cell-derived cardiomyocytes. Cell Stem Cell 2013, 12, 127–137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Elliott, D.A.; Braam, S.R.; Koutsis, K.; Ng, E.S.; Jenny, R.; Lagerqvist, E.L.; Biben, C.; Hatzistavrou, T.; Hirst, C.E.; Yu, Q.C.; et al. Nkx2-5(egfp/w) hescs for isolation of human cardiac progenitors and cardiomyocytes. Nat. Methods 2011, 8, 1037–1040. [Google Scholar] [CrossRef]
  68. Uosaki, H.; Fukushima, H.; Takeuchi, A.; Matsuoka, S.; Nakatsuji, N.; Yamanaka, S.; Yamashita, J.K. Efficient and scalable purification of cardiomyocytes from human embryonic and induced pluripotent stem cells by vcam1 surface expression. PLoS ONE 2011, 6, e23657. [Google Scholar] [CrossRef] [PubMed]
  69. Dubois, N.C.; Craft, A.M.; Sharma, P.; Elliott, D.A.; Stanley, E.G.; Elefanty, A.G.; Gramolini, A.; Keller, G. Sirpa is a specific cell-surface marker for isolating cardiomyocytes derived from human pluripotent stem cells. Nat. Biotechnol. 2011, 29, 1011–1018. [Google Scholar] [CrossRef] [Green Version]
  70. Zhao, M.T.; Shao, N.Y.; Garg, V. Subtype-specific cardiomyocytes for precision medicine: Where are we now? Stem Cells 2020, 38, 822–833. [Google Scholar] [CrossRef] [Green Version]
  71. Zhang, Q.; Jiang, J.; Han, P.; Yuan, Q.; Zhang, J.; Zhang, X.; Xu, Y.; Cao, H.; Meng, Q.; Chen, L.; et al. Direct differentiation of atrial and ventricular myocytes from human embryonic stem cells by alternating retinoid signals. Cell Res. 2011, 21, 579–587. [Google Scholar] [CrossRef] [Green Version]
  72. Lee, J.H.; Protze, S.I.; Laksman, Z.; Backx, P.H.; Keller, G.M. Human pluripotent stem cell-derived atrial and ventricular cardiomyocytes develop from distinct mesoderm populations. Cell Stem Cell 2017, 21, 179–194 e174. [Google Scholar] [CrossRef]
  73. Kleinsorge, M.; Cyganek, L. Subtype-directed differentiation of human ipscs into atrial and ventricular cardiomyocytes. STAR Protoc. 2020, 1, 100026. [Google Scholar] [CrossRef]
  74. Fu, J.D.; Rushing, S.N.; Lieu, D.K.; Chan, C.W.; Kong, C.W.; Geng, L.; Wilson, K.D.; Chiamvimonvat, N.; Boheler, K.R.; Wu, J.C.; et al. Distinct roles of microrna-1 and -499 in ventricular specification and functional maturation of human embryonic stem cell-derived cardiomyocytes. PLoS ONE 2011, 6, e27417. [Google Scholar] [CrossRef] [Green Version]
  75. Zhu, W.Z.; Xie, Y.; Moyes, K.W.; Gold, J.D.; Askari, B.; Laflamme, M.A. Neuregulin/erbb signaling regulates cardiac subtype specification in differentiating human embryonic stem cells. Circ. Res. 2010, 107, 776–786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Yechikov, S.; Kao, H.K.J.; Chang, C.W.; Pretto, D.; Zhang, X.D.; Sun, Y.H.; Smithers, R.; Sirish, P.; Nolta, J.A.; Chan, J.W.; et al. Nodal inhibition promotes differentiation of pacemaker-like cardiomyocytes from human induced pluripotent stem cells. Stem Cell Res. 2020, 49, 102043. [Google Scholar] [CrossRef]
  77. Fukushima, H.; Yoshioka, M.; Kawatou, M.; Lopez-Davila, V.; Takeda, M.; Kanda, Y.; Sekino, Y.; Yoshida, Y.; Yamashita, J.K. Specific induction and long-term maintenance of high purity ventricular cardiomyocytes from human induced pluripotent stem cells. PLoS ONE 2020, 15, e0241287. [Google Scholar] [CrossRef]
  78. Schweizer, P.A.; Darche, F.F.; Ullrich, N.D.; Geschwill, P.; Greber, B.; Rivinius, R.; Seyler, C.; Muller-Decker, K.; Draguhn, A.; Utikal, J.; et al. Subtype-specific differentiation of cardiac pacemaker cell clusters from human induced pluripotent stem cells. Stem Cell Res. Ther. 2017, 8, 229. [Google Scholar] [CrossRef] [Green Version]
  79. Towbin, J.A. The role of cytoskeletal proteins in cardiomyopathies. Curr. Opin. Cell Biol. 1998, 10, 131–139. [Google Scholar] [CrossRef]
  80. Mudhar, H.S.; Wagner, B.E.; Suvarna, S.K. Electron microscopy of myocardial tissue. A nine year review. J. Clin. Pathol. 2001, 54, 321–325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Lin, B.; Li, Y.; Han, L.; Kaplan, A.D.; Ao, Y.; Kalra, S.; Bett, G.C.; Rasmusson, R.L.; Denning, C.; Yang, L. Modeling and study of the mechanism of dilated cardiomyopathy using induced pluripotent stem cells derived from individuals with duchenne muscular dystrophy. Dis. Models Mech. 2015, 8, 457–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Verkerk, A.O.; Veerman, C.C.; Zegers, J.G.; Mengarelli, I.; Bezzina, C.R.; Wilders, R. Patch-clamp recording from human induced pluripotent stem cell-derived cardiomyocytes: Improving action potential characteristics through dynamic clamp. Int. J. Mol. Sci. 2017, 18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Kornreich, B.G. The patch clamp technique: Principles and technical considerations. J. Vet. Cardiol. 2007, 9, 25–37. [Google Scholar] [CrossRef]
  84. Emmenegger, V.; Obien, M.E.J.; Franke, F.; Hierlemann, A. Technologies to study action potential propagation with a focus on hd-meas. Front. Cell Neurosci. 2019, 13, 159. [Google Scholar] [CrossRef]
  85. Spira, M.E.; Hai, A. Multi-electrode array technologies for neuroscience and cardiology. Nat. Nanotechnol. 2013, 8, 83–94. [Google Scholar] [CrossRef] [PubMed]
  86. Attin, M.; Clusin, W.T. Basic concepts of optical mapping techniques in cardiac electrophysiology. Biol. Res. Nurs. 2009, 11, 195–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Lullmann, H.; Ziegler, A. Calcium, cell membrane, and excitation-contraction coupling. J. Cardiovasc. Pharmacol. 1987, 10 (Suppl. 1), S2-8. [Google Scholar] [CrossRef]
  88. Sedwick, C. How a mutation undermines cardiac function. J. Gen. Physiol. 2019, 151, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Zatti, S.; Martewicz, S.; Serena, E.; Uno, N.; Giobbe, G.; Kazuki, Y.; Oshimura, M.; Elvassore, N. Complete restoration of multiple dystrophin isoforms in genetically corrected duchenne muscular dystrophy patient-derived cardiomyocytes. Mol. Therapy. Methods Clin. Dev. 2014, 1, 1. [Google Scholar] [CrossRef]
  90. Csobonyeiova, M.; Polak, S.; Danisovic, L. Toxicity testing and drug screening using ipsc-derived hepatocytes, cardiomyocytes, and neural cells. Can J. Physiol. Pharmacol. 2016, 94, 687–694. [Google Scholar] [CrossRef] [Green Version]
  91. Wallace, E.; Howard, L.; Liu, M.; O’Brien, T.; Ward, D.; Shen, S.; Prendiville, T. Long qt syndrome: Genetics and future perspective. Pediatr. Cardiol. 2019, 40, 1419–1430. [Google Scholar] [CrossRef] [Green Version]
  92. Sendfeld, F.; Selga, E.; Scornik, F.S.; Perez, G.J.; Mills, N.L.; Brugada, R. Experimental models of brugada syndrome. Int. J. Mol. Sci. 2019, 20. [Google Scholar] [CrossRef] [Green Version]
  93. Sun, N.; Yazawa, M.; Liu, J.; Han, L.; Sanchez-Freire, V.; Abilez, O.J.; Navarrete, E.G.; Hu, S.; Wang, L.; Lee, A.; et al. Patient-specific induced pluripotent stem cells as a model for familial dilated cardiomyopathy. Sci. Transl. Med. 2012, 4, 130ra147. [Google Scholar] [CrossRef] [Green Version]
  94. Pfeiffer-Kaushik, E.R.; Smith, G.L.; Cai, B.; Dempsey, G.T.; Hortigon-Vinagre, M.P.; Zamora, V.; Feng, S.; Ingermanson, R.; Zhu, R.; Hariharan, V.; et al. Electrophysiological characterization of drug response in hsc-derived cardiomyocytes using voltage-sensitive optical platforms. J. Pharmacol. Toxicol. Methods 2019, 99, 106612. [Google Scholar] [CrossRef] [PubMed]
  95. Izumi-Nakaseko, H.; Hagiwara-Nagasawa, M.; Naito, A.T.; Goto, A.; Chiba, K.; Sekino, Y.; Kanda, Y.; Sugiyama, A. Application of human induced pluripotent stem cell-derived cardiomyocytes sheets with microelectrode array system to estimate antiarrhythmic properties of multi-ion channel blockers. J. Pharmacol. Sci. 2018, 137, 372–378. [Google Scholar] [CrossRef] [PubMed]
  96. Kussauer, S.; David, R.; Lemcke, H. Hipscs derived cardiac cells for drug and toxicity screening and disease modeling: What micro- electrode-array analyses can tell us. Cells 2019, 8. [Google Scholar] [CrossRef] [Green Version]
  97. Huo, J.; Wei, F.; Cai, C.; Lyn-Cook, B.; Pang, L. Sex-related differences in drug-induced qt prolongation and torsades de pointes: A new model system with human ipsc-cms. Toxicol Sci. 2019, 167, 360–374. [Google Scholar] [CrossRef] [PubMed]
  98. Zhang, J.Z.; Termglinchan, V.; Shao, N.Y.; Itzhaki, I.; Liu, C.; Ma, N.; Tian, L.; Wang, V.Y.; Chang, A.C.Y.; Guo, H.; et al. A human ipsc double-reporter system enables purification of cardiac lineage subpopulations with distinct function and drug response profiles. Cell Stem Cell 2019, 24, 802–811 e805. [Google Scholar] [CrossRef]
  99. Devalla, H.D.; Schwach, V.; Ford, J.W.; Milnes, J.T.; El-Haou, S.; Jackson, C.; Gkatzis, K.; Elliott, D.A.; Chuva de Sousa Lopes, S.M.; Mummery, C.L.; et al. Atrial-like cardiomyocytes from human pluripotent stem cells are a robust preclinical model for assessing atrial-selective pharmacology. EMBO Mol. Med. 2015, 7, 394–410. [Google Scholar] [CrossRef]
  100. Lemme, M.; Ulmer, B.M.; Lemoine, M.D.; Zech, A.T.L.; Flenner, F.; Ravens, U.; Reichenspurner, H.; Rol-Garcia, M.; Smith, G.; Hansen, A.; et al. Atrial-like engineered heart tissue: An in vitro model of the human atrium. Stem Cell Rep. 2018, 11, 1378–1390. [Google Scholar] [CrossRef] [Green Version]
  101. Geng, L.; Kong, C.W.; Wong, A.O.T.; Shum, A.M.; Chow, M.Z.Y.; Che, H.; Zhang, C.; Yau, K.L.; Chan, C.W.; Keung, W.; et al. Probing flecainide block of ina using human pluripotent stem cell-derived ventricular cardiomyocytes adapted to automated patch-clamping and 2d monolayers. Toxicol. Lett. 2018, 294, 61–72. [Google Scholar] [CrossRef]
  102. Sacchetto, C.; Vitiello, L.; de Windt, L.J.; Rampazzo, A.; Calore, M. Modeling cardiovascular diseases with hipsc-derived cardiomyocytes in 2d and 3d cultures. Int. J. Mol. Sci. 2020, 21. [Google Scholar] [CrossRef]
  103. Shadrin, I.Y.; Allen, B.W.; Qian, Y.; Jackman, C.P.; Carlson, A.L.; Juhas, M.E.; Bursac, N. Cardiopatch platform enables maturation and scale-up of human pluripotent stem cell-derived engineered heart tissues. Nat. Commun. 2017, 8, 1825. [Google Scholar] [CrossRef] [Green Version]
  104. Goldfracht, I.; Efraim, Y.; Shinnawi, R.; Kovalev, E.; Huber, I.; Gepstein, A.; Arbel, G.; Shaheen, N.; Tiburcy, M.; Zimmermann, W.H.; et al. Engineered heart tissue models from hipsc-derived cardiomyocytes and cardiac ecm for disease modeling and drug testing applications. Acta Biomater. 2019, 92, 145–159. [Google Scholar] [CrossRef] [PubMed]
  105. Tiburcy, M.; Hudson, J.E.; Balfanz, P.; Schlick, S.; Meyer, T.; Chang Liao, M.L.; Levent, E.; Raad, F.; Zeidler, S.; Wingender, E.; et al. Defined engineered human myocardium with advanced maturation for applications in heart failure modeling and repair. Circulation 2017, 135, 1832–1847. [Google Scholar] [CrossRef] [PubMed]
  106. Cashman, T.J.; Josowitz, R.; Johnson, B.V.; Gelb, B.D.; Costa, K.D. Human engineered cardiac tissues created using induced pluripotent stem cells reveal functional characteristics of braf-mediated hypertrophic cardiomyopathy. PLoS ONE 2016, 11, e0146697. [Google Scholar] [CrossRef] [PubMed]
  107. Wang, G.; McCain, M.L.; Yang, L.; He, A.; Pasqualini, F.S.; Agarwal, A.; Yuan, H.; Jiang, D.; Zhang, D.; Zangi, L.; et al. Modeling the mitochondrial cardiomyopathy of barth syndrome with induced pluripotent stem cell and heart-on-chip technologies. Nat. Med. 2014, 20, 616–623. [Google Scholar] [CrossRef]
  108. Hinson, J.T.; Chopra, A.; Nafissi, N.; Polacheck, W.J.; Benson, C.C.; Swist, S.; Gorham, J.; Yang, L.; Schafer, S.; Sheng, C.C.; et al. Heart disease. Titin mutations in ips cells define sarcomere insufficiency as a cause of dilated cardiomyopathy. Science 2015, 349, 982–986. [Google Scholar] [CrossRef] [Green Version]
  109. Lemme, M.; Braren, I.; Prondzynski, M.; Aksehirlioglu, B.; Ulmer, B.M.; Schulze, M.L.; Ismaili, D.; Meyer, C.; Hansen, A.; Christ, T.; et al. Chronic intermittent tachypacing by an optogenetic approach induces arrhythmia vulnerability in human engineered heart tissue. Cardiovasc. Res. 2020, 116, 1487–1499. [Google Scholar] [CrossRef] [Green Version]
  110. Long, C.; Li, H.; Tiburcy, M.; Rodriguez-Caycedo, C.; Kyrychenko, V.; Zhou, H.; Zhang, Y.; Min, Y.L.; Shelton, J.M.; Mammen, P.P.A.; et al. Correction of diverse muscular dystrophy mutations in human engineered heart muscle by single-site genome editing. Sci. Adv. 2018, 4, eaap9004. [Google Scholar] [CrossRef] [Green Version]
  111. Shiba, Y.; Gomibuchi, T.; Seto, T.; Wada, Y.; Ichimura, H.; Tanaka, Y.; Ogasawara, T.; Okada, K.; Shiba, N.; Sakamoto, K.; et al. Allogeneic transplantation of ips cell-derived cardiomyocytes regenerates primate hearts. Nature 2016, 538, 388–391. [Google Scholar] [CrossRef]
  112. Laflamme, M.A.; Chen, K.Y.; Naumova, A.V.; Muskheli, V.; Fugate, J.A.; Dupras, S.K.; Reinecke, H.; Xu, C.; Hassanipour, M.; Police, S.; et al. Cardiomyocytes derived from human embryonic stem cells in pro-survival factors enhance function of infarcted rat hearts. Nat. Biotechnol. 2007, 25, 1015–1024. [Google Scholar] [CrossRef]
  113. Caspi, O.; Huber, I.; Kehat, I.; Habib, M.; Arbel, G.; Gepstein, A.; Yankelson, L.; Aronson, D.; Beyar, R.; Gepstein, L. Transplantation of human embryonic stem cell-derived cardiomyocytes improves myocardial performance in infarcted rat hearts. J. Am. Coll. Cardiol. 2007, 50, 1884–1893. [Google Scholar] [CrossRef] [PubMed]
  114. Gao, L.; Gregorich, Z.R.; Zhu, W.; Mattapally, S.; Oduk, Y.; Lou, X.; Kannappan, R.; Borovjagin, A.V.; Walcott, G.P.; Pollard, A.E.; et al. Large cardiac muscle patches engineered from human induced-pluripotent stem cell-derived cardiac cells improve recovery from myocardial infarction in swine. Circulation 2018, 137, 1712–1730. [Google Scholar] [CrossRef]
  115. Zhao, X.; Chen, H.; Xiao, D.; Yang, H.; Itzhaki, I.; Qin, X.; Chour, T.; Aguirre, A.; Lehmann, K.; Kim, Y.; et al. Comparison of non-human primate versus human induced pluripotent stem cell-derived cardiomyocytes for treatment of myocardial infarction. Stem Cell Rep. 2018, 10, 422–435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Rojas, S.V.; Kensah, G.; Rotaermel, A.; Baraki, H.; Kutschka, I.; Zweigerdt, R.; Martin, U.; Haverich, A.; Gruh, I.; Martens, A. Transplantation of purified ipsc-derived cardiomyocytes in myocardial infarction. PLoS ONE 2017, 12, e0173222. [Google Scholar] [CrossRef]
  117. Willi, L.; Agranovich, B.; Abramovich, I.; Freimark, D.; Arad, M.; Gottlieb, E.; Binah, O. Bioenergetic and metabolic impairments in duchenne muscular dystrophy (dmd) patients’ ipsc-derived cardiomyocytes. Eur. Heart J. 2020, 41. [Google Scholar] [CrossRef]
  118. Tsurumi, F.; Baba, S.; Yoshinaga, D.; Umeda, K.; Hirata, T.; Takita, J.; Heike, T. The intracellular ca2+ concentration is elevated in cardiomyocytes differentiated from hipscs derived from a duchenne muscular dystrophy patient. PLoS ONE 2019, 14, e0213768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Mournetas, V.; Massourides, E.; Dupont, J.B.; Kornobis, E.; Polveche, H.; Jarrige, M.; Dorval, A.R.L.; Gosselin, M.R.F.; Manousopoulou, A.; Garbis, S.D.; et al. Myogenesis modelled by human pluripotent stem cells: A multi-omic study of duchenne myopathy early onset. J. Cachexia Sarcopenia Muscle 2021, 12, 209–232. [Google Scholar] [CrossRef]
  120. Eisen, B.; Ben Jehuda, R.; Cuttitta, A.J.; Mekies, L.N.; Shemer, Y.; Baskin, P.; Reiter, I.; Willi, L.; Freimark, D.; Gherghiceanu, M.; et al. Electrophysiological abnormalities in induced pluripotent stem cell-derived cardiomyocytes generated from duchenne muscular dystrophy patients. J. Cell Mol. Med. 2019, 23, 2125–2135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Birket, M.J.; Raibaud, S.; Lettieri, M.; Adamson, A.D.; Letang, V.; Cervello, P.; Redon, N.; Ret, G.; Viale, S.; Wang, B.; et al. A human stem cell model of fabry disease implicates limp-2 accumulation in cardiomyocyte pathology. Stem Cell Rep. 2019, 13, 380–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Kuramoto, Y.; Naito, A.T.; Tojo, H.; Sakai, T.; Ito, M.; Shibamoto, M.; Nakagawa, A.; Higo, T.; Okada, K.; Yamaguchi, T.; et al. Generation of fabry cardiomyopathy model for drug screening using induced pluripotent stem cell-derived cardiomyocytes from a female fabry patient. J. Mol. Cell Cardiol. 2018, 121, 256–265. [Google Scholar] [CrossRef]
  123. Chien, Y.; Chien, C.S.; Chiang, H.C.; Huang, W.L.; Chou, S.J.; Chang, W.C.; Chang, Y.L.; Leu, H.B.; Chen, K.H.; Wang, K.L.; et al. Interleukin-18 deteriorates fabry cardiomyopathy and contributes to the development of left ventricular hypertrophy in fabry patients with gla ivs4+919 g>a mutation. Oncotarget 2016, 7, 87161–87179. [Google Scholar] [CrossRef] [PubMed]
  124. Bienvenu, T.; Philippe, C.; De Roux, N.; Raynaud, M.; Bonnefond, J.P.; Pasquier, L.; Lesca, G.; Mancini, J.; Jonveaux, P.; Moncla, A.; et al. The incidence of rett syndrome in france. Pediatr. Neurol. 2006, 34, 372–375. [Google Scholar] [CrossRef]
  125. Chahrour, M.; Zoghbi, H.Y. The story of rett syndrome: From clinic to neurobiology. Neuron 2007, 56, 422–437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Itier, J.M.; Ret, G.; Viale, S.; Sweet, L.; Bangari, D.; Caron, A.; Le-Gall, F.; Benichou, B.; Leonard, J.; Deleuze, J.F.; et al. Effective clearance of gl-3 in a human ipsc-derived cardiomyocyte model of fabry disease. J. Inherit. Metab. Dis. 2014, 37, 1013–1022. [Google Scholar] [CrossRef] [PubMed]
  127. Hashem, S.I.; Perry, C.N.; Bauer, M.; Han, S.; Clegg, S.D.; Ouyang, K.; Deacon, D.C.; Spinharney, M.; Panopoulos, A.D.; Izpisua Belmonte, J.C.; et al. Brief report: Oxidative stress mediates cardiomyocyte apoptosis in a human model of danon disease and heart failure. Stem Cells 2015, 33, 2343–2350. [Google Scholar] [CrossRef] [Green Version]
  128. Marchetto, M.C.; Carromeu, C.; Acab, A.; Yu, D.; Yeo, G.W.; Mu, Y.; Chen, G.; Gage, F.H.; Muotri, A.R. A model for neural development and treatment of rett syndrome using human induced pluripotent stem cells. Cell 2010, 143, 527–539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Sharma, P.; Mesci, P.; Carromeu, C.; McClatchy, D.R.; Schiapparelli, L.; Yates, J.R., 3rd; Muotri, A.R.; Cline, H.T. Exosomes regulate neurogenesis and circuit assembly. Proc. Natl. Acad. Sci. USA 2019, 116, 16086–16094. [Google Scholar] [CrossRef] [Green Version]
  130. Landucci, E.; Brindisi, M.; Bianciardi, L.; Catania, L.M.; Daga, S.; Croci, S.; Frullanti, E.; Fallerini, C.; Butini, S.; Brogi, S.; et al. Ipsc-derived neurons profiling reveals gabaergic circuit disruption and acetylated alpha-tubulin defect which improves after ihdac6 treatment in rett syndrome. Exp. Cell Res. 2018, 368, 225–235. [Google Scholar] [CrossRef] [PubMed]
  131. Gomathi, M.; Balachandar, V. Novel therapeutic approaches: Rett syndrome and human induced pluripotent stem cell technology. Stem Cell Investig. 2017, 4, 20. [Google Scholar] [CrossRef] [Green Version]
  132. Crisafulli, S.; Sultana, J.; Fontana, A.; Salvo, F.; Messina, S.; Trifiro, G. Global epidemiology of duchenne muscular dystrophy: An updated systematic review and meta-analysis. Orphanet J. Rare. Dis. 2020, 15, 141. [Google Scholar] [CrossRef]
  133. Duan, D.; Goemans, N.; Takeda, S.; Mercuri, E.; Aartsma-Rus, A. Duchenne muscular dystrophy. Nat. Rev. Dis. Primers 2021, 7, 13. [Google Scholar] [CrossRef] [PubMed]
  134. Monaco, A.P.; Bertelson, C.J.; Liechti-Gallati, S.; Moser, H.; Kunkel, L.M. An explanation for the phenotypic differences between patients bearing partial deletions of the dmd locus. Genomics 1988, 2, 90–95. [Google Scholar] [CrossRef]
  135. Wolinsky, H.; Goldfischer, S.; Schiller, B.; Kasak, L.E. Lysosomes in aortic smooth muscle cells. Effects of hypertension. Am. J. Pathol. 1973, 73, 727–734. [Google Scholar] [PubMed]
  136. Kaspar, R.W.; Allen, H.D.; Montanaro, F. Current understanding and management of dilated cardiomyopathy in duchenne and becker muscular dystrophy. J. Am. Acad. Nurse Pract. 2009, 21, 241–249. [Google Scholar] [CrossRef] [Green Version]
  137. Mendell, J.R.; Shilling, C.; Leslie, N.D.; Flanigan, K.M.; al-Dahhak, R.; Gastier-Foster, J.; Kneile, K.; Dunn, D.M.; Duval, B.; Aoyagi, A.; et al. Evidence-based path to newborn screening for duchenne muscular dystrophy. Ann. Neurol. 2012, 71, 304–313. [Google Scholar] [CrossRef]
  138. Frankel, K.A.; Rosser, R.J. The pathology of the heart in progressive muscular dystrophy: Epimyocardial fibrosis. Hum. Pathol. 1976, 7, 375–386. [Google Scholar] [CrossRef]
  139. Japp, A.G.; Gulati, A.; Cook, S.A.; Cowie, M.R.; Prasad, S.K. The diagnosis and evaluation of dilated cardiomyopathy. J. Am. Coll. Cardiol. 2016, 67, 2996–3010. [Google Scholar] [CrossRef]
  140. Zhu, S.; Law, A.H.Y.; Deng, R.; Poon, E.N.Y.; Lo, C.W.; Kwong, A.K.Y.; Liang, R.; Chan, K.Y.K.; Wong, W.L.; Tan-Un, K.C.; et al. Generation of genomic-integration-free human induced pluripotent stem cells and the derived cardiomyocytes of x-linked dilated cardiomyopathy from dmd gene mutation. Stem Cell Res. 2020, 49, 102040. [Google Scholar] [CrossRef] [PubMed]
  141. De Palma, C.; Morisi, F.; Cheli, S.; Pambianco, S.; Cappello, V.; Vezzoli, M.; Rovere-Querini, P.; Moggio, M.; Ripolone, M.; Francolini, M.; et al. Autophagy as a new therapeutic target in duchenne muscular dystrophy. Cell Death Dis. 2012, 3, e418. [Google Scholar] [CrossRef] [Green Version]
  142. Guan, X.; Mack, D.L.; Moreno, C.M.; Strande, J.L.; Mathieu, J.; Shi, Y.; Markert, C.D.; Wang, Z.; Liu, G.; Lawlor, M.W.; et al. Dystrophin-deficient cardiomyocytes derived from human urine: New biologic reagents for drug discovery. Stem Cell Res. 2014, 12, 467–480. [Google Scholar] [CrossRef] [Green Version]
  143. Takami, Y.; Takeshima, Y.; Awano, H.; Okizuka, Y.; Yagi, M.; Matsuo, M. High incidence of electrocardiogram abnormalities in young patients with duchenne muscular dystrophy. Pediatr. Neurol. 2008, 39, 399–403. [Google Scholar] [CrossRef] [PubMed]
  144. Maroto, R.; Raso, A.; Wood, T.G.; Kurosky, A.; Martinac, B.; Hamill, O.P. Trpc1 forms the stretch-activated cation channel in vertebrate cells. Nat. Cell Biol. 2005, 7, 179–185. [Google Scholar] [CrossRef]
  145. Meikle, P.J.; Hopwood, J.J.; Clague, A.E.; Carey, W.F. Prevalence of lysosomal storage disorders. JAMA 1999, 281, 249–254. [Google Scholar] [CrossRef] [PubMed]
  146. Poorthuis, B.J.; Wevers, R.A.; Kleijer, W.J.; Groener, J.E.; de Jong, J.G.; van Weely, S.; Niezen-Koning, K.E.; van Diggelen, O.P. The frequency of lysosomal storage diseases in the netherlands. Hum. Genet. 1999, 105, 151–156. [Google Scholar] [CrossRef]
  147. Spada, M.; Pagliardini, S.; Yasuda, M.; Tukel, T.; Thiagarajan, G.; Sakuraba, H.; Ponzone, A.; Desnick, R.J. High incidence of later-onset fabry disease revealed by newborn screening. Am. J. Hum. Genet. 2006, 79, 31–40. [Google Scholar] [CrossRef] [Green Version]
  148. Wang, R.Y.; Lelis, A.; Mirocha, J.; Wilcox, W.R. Heterozygous fabry women are not just carriers, but have a significant burden of disease and impaired quality of life. Genet. Med. 2007, 9, 34–45. [Google Scholar] [CrossRef] [Green Version]
  149. Zarate, Y.A.; Hopkin, R.J. Fabry’s disease. Lancet 2008, 372, 1427–1435. [Google Scholar] [CrossRef]
  150. Gieselmann, V. Cellular pathophysiology of lysosomal storage diseases. In Fabry Disease: Perspectives from 5 Years of Fos; Mehta, A., Beck, M., Sunder-Plassmann, G., Eds.; Oxford PharmaGenesis: Oxford, UK, 2006. [Google Scholar]
  151. Morrissey, R.P.; Philip, K.J.; Schwarz, E.R. Cardiac abnormalities in anderson-fabry disease and fabry’s cardiomyopathy. Cardiovasc. J. Afr. 2011, 22, 38–44. [Google Scholar] [PubMed]
  152. Linhart, A. The heat in fabry disease. In Fabry Disease: Perspectives from 5 Years of Fos; Mehta, A.B., Beck, M., Sunder-Plassmann, G., Eds.; Oxford PharmaGenesis: Oxford, UK, 2006. [Google Scholar]
  153. Schiffmann, R.; Kopp, J.B.; Austin, H.A., 3rd; Sabnis, S.; Moore, D.F.; Weibel, T.; Balow, J.E.; Brady, R.O. Enzyme replacement therapy in fabry disease: A randomized controlled trial. JAMA 2001, 285, 2743–2749. [Google Scholar] [CrossRef] [PubMed]
  154. Borger, D.K.; McMahon, B.; Roshan Lal, T.; Serra-Vinardell, J.; Aflaki, E.; Sidransky, E. Induced pluripotent stem cell models of lysosomal storage disorders. Dis. Model. Mech. 2017, 10, 691–704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Kawagoe, S.; Higuchi, T.; Otaka, M.; Shimada, Y.; Kobayashi, H.; Ida, H.; Ohashi, T.; Okano, H.J.; Nakanishi, M.; Eto, Y. Morphological features of ips cells generated from fabry disease skin fibroblasts using sendai virus vector (sevdp). Mol. Genet. Metab. 2013, 109, 386–389. [Google Scholar] [CrossRef] [PubMed]
  156. Renieri, A.; Meloni, I.; Longo, I.; Ariani, F.; Mari, F.; Pescucci, C.; Cambi, F. Rett syndrome: The complex nature of a monogenic disease. J. Mol. Med. 2003, 81, 346–354. [Google Scholar] [CrossRef]
  157. CENTOGENE. Application of iPSC Technology in Orphan Drug Development. Available online: https://www.centogene.com/science/whitepapers/application-of-ipsc-technology-in-orphan-drug-development.html (accessed on 20 July 2021).
  158. Kubota, T. Advantage in human induced pluripotent stem cells research of x-linked genetic diseases for drug screening. Int. J. Stem Cell Res. Ther. 2015, 2, 12. [Google Scholar] [CrossRef]
  159. Huang, H.P.; Chuang, C.Y.; Kuo, H.C. Induced pluripotent stem cell technology for disease modeling and drug screening with emphasis on lysosomal storage diseases. Stem Cell Res. Ther. 2012, 3, 34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Chien, Y.; Chou, S.J.; Chang, Y.L.; Leu, H.B.; Yang, Y.P.; Tsai, P.H.; Lai, Y.H.; Chen, K.H.; Chang, W.C.; Sung, S.H.; et al. Inhibition of arachidonate 12/15-lipoxygenase improves alpha-galactosidase efficacy in ipsc-derived cardiomyocytes from fabry patients. Int. J. Mol. Sci. 2018, 19. [Google Scholar] [CrossRef] [Green Version]
  161. Acampa, M.; Guideri, F. Cardiac disease and rett syndrome. Arch. Dis. Child. 2006, 91, 440–443. [Google Scholar] [CrossRef] [Green Version]
  162. Muffat, J.; Li, Y.; Yuan, B.; Mitalipova, M.; Omer, A.; Corcoran, S.; Bakiasi, G.; Tsai, L.H.; Aubourg, P.; Ransohoff, R.M.; et al. Efficient derivation of microglia-like cells from human pluripotent stem cells. Nat. Med. 2016, 22, 1358–1367. [Google Scholar] [CrossRef] [Green Version]
  163. Bienvenu, T.; Carrie, A.; de Roux, N.; Vinet, M.C.; Jonveaux, P.; Couvert, P.; Villard, L.; Arzimanoglou, A.; Beldjord, C.; Fontes, M.; et al. Mecp2 mutations account for most cases of typical forms of rett syndrome. Hum. Mol. Genet. 2000, 9, 1377–1384. [Google Scholar] [CrossRef]
  164. Zachariah, R.M.; Rastegar, M. Linking epigenetics to human disease and rett syndrome: The emerging novel and challenging concepts in mecp2 research. Neural Plast. 2012, 2012, 415825. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Gold, W.A.; Krishnarajy, R.; Ellaway, C.; Christodoulou, J. Rett syndrome: A genetic update and clinical review focusing on comorbidities. ACS Chem. Neurosci. 2018, 9, 167–176. [Google Scholar] [CrossRef]
  166. Sekul, E.A.; Moak, J.P.; Schultz, R.J.; Glaze, D.G.; Dunn, J.K.; Percy, A.K. Electrocardiographic findings in rett syndrome: An explanation for sudden death? J. Pediatr. 1994, 125, 80–82. [Google Scholar] [CrossRef]
  167. Zhu, Y.C.; Xiong, Z.Q. Molecular and synaptic bases of cdkl5 disorder. Dev. Neurobiol. 2019, 79, 8–19. [Google Scholar] [CrossRef] [Green Version]
  168. Yang, Z.; McMahon, C.J.; Smith, L.R.; Bersola, J.; Adesina, A.M.; Breinholt, J.P.; Kearney, D.L.; Dreyer, W.J.; Denfield, S.W.; Price, J.F.; et al. Danon disease as an underrecognized cause of hypertrophic cardiomyopathy in children. Circulation 2005, 112, 1612–1617. [Google Scholar] [CrossRef] [Green Version]
  169. Myerowitz, R.; Puertollano, R.; Raben, N. Impaired autophagy: The collateral damage of lysosomal storage disorders. EBioMedicine 2021, 63, 103166. [Google Scholar] [CrossRef]
  170. Harris, K.M.; Spirito, P.; Maron, M.S.; Zenovich, A.G.; Formisano, F.; Lesser, J.R.; Mackey-Bojack, S.; Manning, W.J.; Udelson, J.E.; Maron, B.J. Prevalence, clinical profile, and significance of left ventricular remodeling in the end-stage phase of hypertrophic cardiomyopathy. Circulation 2006, 114, 216–225. [Google Scholar] [CrossRef]
  171. Sugie, K.; Yamamoto, A.; Murayama, K.; Oh, S.J.; Takahashi, M.; Mora, M.; Riggs, J.E.; Colomer, J.; Iturriaga, C.; Meloni, A.; et al. Clinicopathological features of genetically confirmed danon disease. Neurology 2002, 58, 1773–1778. [Google Scholar] [CrossRef] [Green Version]
  172. D’Souza, R.S.; Levandowski, C.; Slavov, D.; Graw, S.L.; Allen, L.A.; Adler, E.; Mestroni, L.; Taylor, M.R. Danon disease: Clinical features, evaluation, and management. Circ. Heart Fail. 2014, 7, 843–849. [Google Scholar] [CrossRef] [Green Version]
  173. Hashem, S.I.; Murphy, A.N.; Divakaruni, A.S.; Klos, M.L.; Nelson, B.C.; Gault, E.C.; Rowland, T.J.; Perry, C.N.; Gu, Y.; Dalton, N.D.; et al. Impaired mitophagy facilitates mitochondrial damage in danon disease. J. Mol. Cell Cardiol. 2017, 108, 86–94. [Google Scholar] [CrossRef]
  174. Chi, C.; Leonard, A.; Knight, W.E.; Beussman, K.M.; Zhao, Y.; Cao, Y.; Londono, P.; Aune, E.; Trembley, M.A.; Small, E.M.; et al. Lamp-2b regulates human cardiomyocyte function by mediating autophagosome–lysosome fusion. Proc. Natl. Acad. Sci. USA 2019, 116, 556–565. [Google Scholar] [CrossRef] [Green Version]
  175. Del Favero, G.; Bonifacio, A.; Rowland, T.J.; Gao, S.; Song, K.; Sergo, V.; Adler, E.D.; Mestroni, L.; Sbaizero, O.; Taylor, M.R.G. Danon disease-associated lamp-2 deficiency drives metabolic signature indicative of mitochondrial aging and fibrosis in cardiac tissue and hipsc-derived cardiomyocytes. J. Clin. Med. 2020, 9. [Google Scholar] [CrossRef]
  176. Goldberg, A.L. Protein degradation and protection against misfolded or damaged proteins. Nature 2003, 426, 895–899. [Google Scholar] [CrossRef]
  177. Lee, Y.K.; Ng, K.M.; Lai, W.H.; Chan, Y.C.; Lau, Y.M.; Lian, Q.; Tse, H.F.; Siu, C.W. Calcium homeostasis in human induced pluripotent stem cell-derived cardiomyocytes. Stem Cell Rev. 2011, 7, 976–986. [Google Scholar] [CrossRef] [Green Version]
  178. Koivumaki, J.T.; Naumenko, N.; Tuomainen, T.; Takalo, J.; Oksanen, M.; Puttonen, K.A.; Lehtonen, S.; Kuusisto, J.; Laakso, M.; Koistinaho, J.; et al. Structural immaturity of human ipsc-derived cardiomyocytes: In silico investigation of effects on function and disease modeling. Front. Physiol. 2018, 9, 80. [Google Scholar] [CrossRef]
  179. Itzhaki, I.; Rapoport, S.; Huber, I.; Mizrahi, I.; Zwi-Dantsis, L.; Arbel, G.; Schiller, J.; Gepstein, L. Calcium handling in human induced pluripotent stem cell derived cardiomyocytes. PLoS ONE 2011, 6, e18037. [Google Scholar] [CrossRef]
  180. Law, M.L.; Cohen, H.; Martin, A.A.; Angulski, A.B.B.; Metzger, J.M. Dysregulation of calcium handling in duchenne muscular dystrophy-associated dilated cardiomyopathy: Mechanisms and experimental therapeutic strategies. J. Clin. Med. 2020, 9. [Google Scholar] [CrossRef] [Green Version]
  181. Lewandowski, J.; Rozwadowska, N.; Kolanowski, T.J.; Malcher, A.; Zimna, A.; Rugowska, A.; Fiedorowicz, K.; Labedz, W.; Kubaszewski, L.; Chojnacka, K.; et al. The impact of in vitro cell culture duration on the maturation of human cardiomyocytes derived from induced pluripotent stem cells of myogenic origin. Cell Transpl. 2018, 27, 1047–1067. [Google Scholar] [CrossRef] [Green Version]
  182. Abilez, O.J.; Tzatzalos, E.; Yang, H.; Zhao, M.T.; Jung, G.; Zollner, A.M.; Tiburcy, M.; Riegler, J.; Matsa, E.; Shukla, P.; et al. Passive stretch induces structural and functional maturation of engineered heart muscle as predicted by computational modeling. Stem Cells 2018, 36, 265–277. [Google Scholar] [CrossRef] [Green Version]
  183. Ronaldson-Bouchard, K.; Ma, S.P.; Yeager, K.; Chen, T.; Song, L.; Sirabella, D.; Morikawa, K.; Teles, D.; Yazawa, M.; Vunjak-Novakovic, G. Advanced maturation of human cardiac tissue grown from pluripotent stem cells. Nature 2018, 556, 239–243. [Google Scholar] [CrossRef]
  184. Yang, X.; Rodriguez, M.; Pabon, L.; Fischer, K.A.; Reinecke, H.; Regnier, M.; Sniadecki, N.J.; Ruohola-Baker, H.; Murry, C.E. Tri-iodo-l-thyronine promotes the maturation of human cardiomyocytes-derived from induced pluripotent stem cells. J. Mol. Cell Cardiol. 2014, 72, 296–304. [Google Scholar] [CrossRef] [Green Version]
  185. Kuppusamy, K.T.; Jones, D.C.; Sperber, H.; Madan, A.; Fischer, K.A.; Rodriguez, M.L.; Pabon, L.; Zhu, W.Z.; Tulloch, N.L.; Yang, X.; et al. Let-7 family of microrna is required for maturation and adult-like metabolism in stem cell-derived cardiomyocytes. Proc. Natl. Acad. Sci. USA 2015, 112, E2785–E2794. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Generation of isogenic induced pluripotent stem cells (iPSCs) with distinctive X-chromosome inactivation status from the heterozygous female with X-linked mutation.
Figure 1. Generation of isogenic induced pluripotent stem cells (iPSCs) with distinctive X-chromosome inactivation status from the heterozygous female with X-linked mutation.
Ijms 22 08132 g001
Figure 2. Differentiation of iPSCs into cardiomyocytes using commercially available kits from Thermo Fisher scientific: (A) the iPSCs were cultured as a monolayer in differentiation medium A (day 0-day 2), medium B (day 2-day 4), and cardiomyocyte maintenance medium (from day 4). Beating clusters usually can be observed by day 8, scale bar: 200 µm; (B) flow cytometry analysis indicates that the cardiac troponin-T (cTnT)-SERCA2-double positive cells constitute approximately 84%; (C) immunostaining of the cardiomyocytes dissociated from the beating clusters, scale bar: 50 µm.
Figure 2. Differentiation of iPSCs into cardiomyocytes using commercially available kits from Thermo Fisher scientific: (A) the iPSCs were cultured as a monolayer in differentiation medium A (day 0-day 2), medium B (day 2-day 4), and cardiomyocyte maintenance medium (from day 4). Beating clusters usually can be observed by day 8, scale bar: 200 µm; (B) flow cytometry analysis indicates that the cardiac troponin-T (cTnT)-SERCA2-double positive cells constitute approximately 84%; (C) immunostaining of the cardiomyocytes dissociated from the beating clusters, scale bar: 50 µm.
Ijms 22 08132 g002
Figure 3. MEA and optical mapping for studying the electrophysiology of the iPSC-derived cardiomyocytes: (A) microelectrode array (MEA) analysis for evaluating the field potential changes across the monolayer surface of iPSC-derived cardiomyocytes; (B) optical mapping-based methods for evaluating the changes in membrane potentials of a single iPSC-derived cardiomyocyte. EAD: early after depolarisation.
Figure 3. MEA and optical mapping for studying the electrophysiology of the iPSC-derived cardiomyocytes: (A) microelectrode array (MEA) analysis for evaluating the field potential changes across the monolayer surface of iPSC-derived cardiomyocytes; (B) optical mapping-based methods for evaluating the changes in membrane potentials of a single iPSC-derived cardiomyocyte. EAD: early after depolarisation.
Ijms 22 08132 g003
Figure 4. IonOptix recording of the change in cell length during contraction.
Figure 4. IonOptix recording of the change in cell length during contraction.
Ijms 22 08132 g004
Figure 5. Generation of iPSC-derived cardiomyocytes from Fabry disease patient: (A) pedigree analysis of a family carrying the GLA-IVS4–919 G > A allele; (B) the iPSC line established from non-disease individual and FD-patient 1 were differentiated into cardiomyocytes and subjected to Western blot analysis; (C) a-GAL enzyme activity of the corresponding iPSC-derived cardiomyocytes. *: p < 0.05 between the indicated groups, N = 4.
Figure 5. Generation of iPSC-derived cardiomyocytes from Fabry disease patient: (A) pedigree analysis of a family carrying the GLA-IVS4–919 G > A allele; (B) the iPSC line established from non-disease individual and FD-patient 1 were differentiated into cardiomyocytes and subjected to Western blot analysis; (C) a-GAL enzyme activity of the corresponding iPSC-derived cardiomyocytes. *: p < 0.05 between the indicated groups, N = 4.
Ijms 22 08132 g005
Figure 6. Application of the iPSCs with a silenced wildtype LAMP2 allele for drug screening: (A) the iPSCs with silenced LAMP2 allele were cultured on 96-well plates and treated with DiscoveryProbe™ Epigenetics Compound Library (APExBio). The reactivation of X-chromosomes will be detected using Alexa488-conjugated antibodies specific to human LAMP2; (B) primary screening results.
Figure 6. Application of the iPSCs with a silenced wildtype LAMP2 allele for drug screening: (A) the iPSCs with silenced LAMP2 allele were cultured on 96-well plates and treated with DiscoveryProbe™ Epigenetics Compound Library (APExBio). The reactivation of X-chromosomes will be detected using Alexa488-conjugated antibodies specific to human LAMP2; (B) primary screening results.
Ijms 22 08132 g006
Table 1. X-linked genes that cause cardiomyopathy.
Table 1. X-linked genes that cause cardiomyopathy.
DiseaseGeneType of CardiomyopathyExtra-Cardiac Manifestations
Duchenne muscular dystrophyDMDDilated [10]Muscle weakness [11]
Growth delay [12]
Fabry diseaseGLAHypertrophic [13]Neuropathic pain
Renal impairment
Angiokeratoma [14]
Familial cardiac filaminopathyFLNADilated [15]Periventricular heterotopia [16]
Hypertrophic [17]
Both dilated and hypertrophic [18]
Danon diseaseLAMP2Hypertrophic [19]Skeletal myopathy
Retinopathy [20]
Cognitive impairment [21]
Dilated [22]
Left-ventricular non-compaction [23]
Rett syndromeMECP2Arrhythmogenic [24]Extrapyramidal motor dysfunction [25]
Epilepsy [26]
Bone fracture [27]
X-linked myotubular myopathyMTM1Dilated [28]Respiratory failure [29]
Muscle weakness
Table 2. Application of iPSCs in the investigation of X-linked cardiomyopathies.
Table 2. Application of iPSCs in the investigation of X-linked cardiomyopathies.
DiseasePathophysiological ChangesReferences
Duchenne muscular dystrophyExpress truncated non-functional dystrophin protein, disrupted myofibrils, calcium overloads, disrupted membrane fragility, increased DAD and OPPs[7,117,118,119,120]
Fabry diseaseGb3 accumulation, deficient enzyme, high ANP expression
Decreased contractility, cellular hypertrophy, disturbed ion channel electrical currents, peripheral displacement of myofibrils
[121,122,123]
Danon diseaseDeficiency of LAMP2, accumulation of autophagy materials, increased mitochondrial oxidative stress, increased apoptosis, altered metabolism, impaired contractile function, reduced calcium transients[32,124,125]
Rett syndromeDecreased soma size, fewer glutamatergic synapses, reduced spine density, altered calcium signalling and electrophysiological defects, decreased axon outgrowth, dendritic morphogenesis
Changes in exosome protein cargo and signalling bioactivity
[124,125]
Table 3. Applications of iPSC-based models for drug testing.
Table 3. Applications of iPSC-based models for drug testing.
DiseaseDrug/Treatment TestedReferences
Duchenne muscluar dystrophyProteasome inhibitors, polaxamer188, human artificial chromosomes[7,81,89]
Fabry diseaseglucosylceramide synthase inhibitor[126]
Danon diseaseN-acetylcystein, rotenone, DNA-demethylating drugs[32,127]
Rett syndromeIGF-1, gentamycin, exosomes hstone deacetylase inhibitors, medhya rasayana[128,129,130,131]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, J.; Chou, O.H.-I.; Tse, Y.-L.; Ng, K.-M.; Tse, H.-F. Application of Patient-Specific iPSCs for Modelling and Treatment of X-Linked Cardiomyopathies. Int. J. Mol. Sci. 2021, 22, 8132. https://doi.org/10.3390/ijms22158132

AMA Style

Zhang J, Chou OH-I, Tse Y-L, Ng K-M, Tse H-F. Application of Patient-Specific iPSCs for Modelling and Treatment of X-Linked Cardiomyopathies. International Journal of Molecular Sciences. 2021; 22(15):8132. https://doi.org/10.3390/ijms22158132

Chicago/Turabian Style

Zhang, Jennifer, Oscar Hou-In Chou, Yiu-Lam Tse, Kwong-Man Ng, and Hung-Fat Tse. 2021. "Application of Patient-Specific iPSCs for Modelling and Treatment of X-Linked Cardiomyopathies" International Journal of Molecular Sciences 22, no. 15: 8132. https://doi.org/10.3390/ijms22158132

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop