Next Article in Journal
Effects of E’Jiao on Skeletal Mineralisation, Osteocyte and WNT Signalling Inhibitors in Ovariectomised Rats
Next Article in Special Issue
Sphingosine-1-phosphate Treatment Improves Cryopreservation Efficiency in Human Mesenchymal Stem Cells
Previous Article in Journal
A Retrospective Single-Arm Cohort Study in a Single Center of Radiofrequency Ablation in Treatment of Chronic Radiation Proctitis
Previous Article in Special Issue
Distribution of Embryonic Stem Cell-Derived Mesenchymal Stem Cells after Intravenous Infusion in Hypoxic–Ischemic Encephalopathy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epigenetic Modification Factors and microRNAs Network Associated with Differentiation of Embryonic Stem Cells and Induced Pluripotent Stem Cells toward Cardiomyocytes: A Review

1
The Persian Gulf Marine Biotechnology Research Center, The Persian Gulf Biomedical Sciences Research Institute, Bushehr University of Medical Sciences, Bushehr 7514633196, Iran
2
Reproductive Development, Murdoch Children’s Research Institute, Melbourne, VIC 3052, Australia
3
Department of Paediatrics, University of Melbourne, Melbourne, VIC 3010, Australia
4
Department of Veterinary and Animal Sciences, University of Copenhagen, 1870 Frederiksberg C, Denmark
5
Department of Basic Sciences, School of Medicine, Bushehr University of Medical Sciences, Bushehr 7514633341, Iran
6
Stem Cell Research Center, Tabriz University of Medical Sciences, Tabriz 5166653431, Iran
7
Department of Reproductive Biology, Faculty of Advanced Medical Sciences, Tabriz University of Medical Sciences, Tabriz 5166653431, Iran
8
Department of Anatomy, School of Medical Sciences, Medicine & Health, UNSW Sydney, Sydney, NSW 2052, Australia
9
PerciaVista R&D Co., Shiraz 7135644144, Iran
*
Authors to whom correspondence should be addressed.
These authors have the same contributions as the first author.
Life 2023, 13(2), 569; https://doi.org/10.3390/life13020569
Submission received: 30 September 2022 / Revised: 16 November 2022 / Accepted: 16 November 2022 / Published: 17 February 2023
(This article belongs to the Special Issue Human Mesenchymal (hMSC) and Induced Pluripotent (hiPSC) Stem Cells)

Abstract

:
More research is being conducted on myocardial cell treatments utilizing stem cell lines that can develop into cardiomyocytes. All of the forms of cardiac illnesses have shown to be quite amenable to treatments using embryonic (ESCs) and induced pluripotent stem cells (iPSCs). In the present study, we reviewed the differentiation of these cell types into cardiomyocytes from an epigenetic standpoint. We also provided a miRNA network that is devoted to the epigenetic commitment of stem cells toward cardiomyocyte cells and related diseases, such as congenital heart defects, comprehensively. Histone acetylation, methylation, DNA alterations, N6-methyladenosine (m6a) RNA methylation, and cardiac mitochondrial mutations are explored as potential tools for precise stem cell differentiation.

1. Introduction

The American Heart Association reported a 363.4-billion-dollar cost for the United States’ cardiovascular diseases, recently [1]. Despite major advancements in the treatment of various cardiac disorders, there being rising mortality rates [2] and a greater burden on healthcare expenditures [3] are anticipated. Some people have described these undesired statistics as the inertia of cardiovascular disease therapy, and therefore, more advanced therapies are being demanded [4]. To comprehend the treatment challenges, we should take a look at heart contraction-producing cells called “cardiomyocytes,” which are inherently different at various stages of maturation. At the adult stage, they reveal characteristics such as longer sarcomeres (2.2 vs. 1.65 μm), lower resting membrane potential, and a dependence on oxidative phosphorylation instead of glycolysis [5,6]. One major characteristic of adult cardiomyocytes which causes problems with the treatment is terminal differentiation. So, after birth, the proliferation is replaced with hypertrophy in the cardiomyocytes. The terminal differentiation phenomenon is executed via either the upregulation of cardiac-specific adult genes or the termination of cardiac cell division [7,8,9]. Additionally, epigenetic mechanisms have been shown to contribute to the occurrence of terminal differentiation [7].
One of the suggested treatments after an infarction is the rejuvenation of the myocardial tissue. To this end, researchers have proposed various sources of stem cells with the capability of differentiating into cardiomyocytes. Three distinct categories may be used to group these cell types. (I) The “First-generation stem cells” typically include the mesenchymal stem cells (MSCs) [10], the hematopoietic stem cells, the endothelial progenitor cells, and the mononuclear cells [11], which are stem cells produced from adult tissues and bone marrow. The “Second-generation stem cells” are committed stem cells that are similar to pluripotent stem cells [12] and induced pluripotent stem cells (iPSCs), patches, exosomes, and they are derivatives of Wharton’s jelly which make up the “third generation of stem cell treatment” [13].
An optimal stem cell candidate for the cell treatment of cardiac disorders in regenerative medicine should meet two important requirements: (I) they should not trigger immunological responses, and (II) they should be being capable of differentiating into the desired cardiovascular cells. One of the pluripotent stem cell resources which are known to be capable of differentiating into cardiomyocytes are human embryonic stem cells (ESCs) [14]. However, their immunogenic potential and ethical constraints are the main objections when one is considering these cell lines [15]. In contrast, researchers have focused on personalized iPSCs since they can be produced by the same person by the nuclear reprogramming of their somatic cells, thus, they are an alternative source for cardiomyocyte production [16]. It has been well documented that the direct transplantation of stem cells into the heart tissue via injection could result in teratoma. Therefore, before their implantation, the effective differentiation of the stem cells into cardiomyocytes is essential [17].
There are now three basic methods for differentiating pluripotent stem cells into cardiomyocytes. The first technique involves co-culturing the pluripotent stem cells with the stromal cells that resemble mouse visceral endoderm (END-2) or in END-2-conditioned media [18]. Additionally, it has been demonstrated that MAPK inhibitors can improve the efficiency of Stem cell differentiation to cardiomyocyte yield. [19]. The second technique for differentiating pluripotent stem cells into cardiomyocytes uses embryoid bodies in suspension. The pluripotent stem cells begin to differentiate into cardiomyocytes when a set of chemical and physical conditions that simulate early embryonic development are present. The pluripotent stem cells gather to form embryoid bodies, which are then co-cultured for a prolonged period, causing 5–70% of the embryoid bodies to differentiate into cardiomyocytes. However, this approach has numerous limitations, including a disparity in the number of beating cardiomyocytes across the embryoid bodies, a poor yield efficiency, and an immature cardiomyocyte phenotype [20].
The last proposed method is two-dimensional monolayer differentiation which is based on adding small molecules and growth factors to the culture medium [21]. This approach results in the successful differentiation of the majority of the cells (85–95%), and subsequently, it yields more mature cardiomyocytes, [22].
Cell nuclear reprogramming via the epigenetic modifications of the cells has a complex mechanism and is conducted via the regulation of the chromatin structure [23]. Furthermore, this mechanism is performed to provide the conditions for the specialization and differentiation of various cell types [24,25]. In the iPSCs, the epigenetics and chromatin modifications are critical but reversible dynamics [26]. Numerous epigenetic factors cooperate with the transcription factors and the signaling pathways to facilitate the cardiomyocyte differentiation process. These epigenetic factors encompass DNA methylation, ATP-dependent chromatin remodeling complexes, and histone modifications [27]. By altering the availability of the DNA sequences for DNA-binding proteins, the epigenetic regulators may affect the expression or repression of the genes. These changes result in looser or tighter chromatin via the altering of the DNA-histone covalent bonds, respectively. As a result, the DNA will either become more or less accessible, respectively. The major focus of this review is the epigenetic regulation of the ESCs and iPSCs into cardiomyocytes.
Another field of interest in promoting regeneration in the cardiac tissue are miRNAs. MiRNAs are basically 19–24 nucleotide RNAs that are encoded endogenously [28], and they downregulate the target mRNA by binding to the 3′-UTR region [29]. MiRNAs promote mRNA breakdown or impede mRNA translation to negatively control the gene expression at the post-transcriptional stage. More than 2600 mature miRNAs have been found in human cells so far [30], and it is known that these miRNAs use intricate methods to control how one-third of the human genes are expressed [31].
MiRNAs are essential controllers of myocardial development and disorders as well [32]. For instance, it has been shown that in vitro myocardial differentiation is promoted by upregulating the cardiac-abundant miRNAs in the ESCs, including mir-1 and mir-499 [33]. Additionally, it was shown that overexpressing miRNA-1 in ESCs improved the cardiac differentiation following a transplantation. Additionally, miRNA-1 overexpression in ESCs may have a positive paracrine impact by preventing cardiomyocytes from apoptosis in vivo [34,35].

2. Embryonic Stem Cells (ESCs)

The ESCs are considered to be an effective cell resource in regenerative medicine due to their inherent plasticity [36,37]. Pluripotent stem cells may be created utilizing a patient’s nuclear genome by cell nuclear transfer during animal cloning using ESCs. These cells may develop into cardiomyocytes, and ultimately, they may be employed to repair heart disease. Although this technology has been applied in animal models [38,39], it comes with some undesirable outcomes, which should be taken into consideration. The low efficiency and low pluripotency in the produced lines, the high demand for superovulation, the abnormalities of cloned cells, and there being an ethical debate for human oocyte manipulations are the main objections.

3. Induced Pluripotent Stem Cells (iPSCs)

Several protocols have been documented for somatic cell reprogramming to differentiate them into iPSCs. These procedures involve employing cell and transgene-free ESC protein extracts, delivering reprogramming factor transgenes by employing adenovirus vectors, lentivirus vectors, Sendai virus vectors, the transduction of Oct4, Sox2, Klf4, and c-Myc by retrovirus [16,40,41,42], the transfection of plasmids without the use of c-Myc [43], and by utilizing the PiggyBac system which is accompanied by a recombinant protein such as tetracycline transactivator (rtTA) [44,45,46,47,48].
The iPSCs are capable of generating all of the three germ lineages and since they contain autologous sources, they could have a syngeneic nature. However, as previously mentioned, the low efficiency of the derived iPSCs, the lengthy treatment procedures, and the propensity for teratoma formation are considered to be the disadvantages of using iPSCs [49]. The epigenetic mechanisms governing the production of iPSCs have not been fully elucidated yet. Therefore, understanding the underlying mechanisms of epigenetic reprogramming and their sustainability following reprogramming is crucial before applying them in a clinical setting.
ESCs and iPSCs are epigenetically highly complicated [50]. A pluripotency status is guaranteed when continuous proliferation and de-differentiation are accompanied by the ability to differentiate into a particular cell lineage. To consider a cell as being pluripotent, three fundamental criteria must be met: (i) it’s epigenetic factors should continuously and actively remain in a flexible balance; (ii) the pluripotency transcripts should remain upregulated, while their differentiating counterparts should be downregulated; (iii) the proper and in-time de-condensation of the chromatin sequences, which are responsible for the initiation of differentiating machinery in the cell [51,52,53,54,55,56,57,58,59]. In other words, for sustaining pluripotency in the ESCs and the iPSCs, a dynamic balance should exist between at least three epigenetic elements, including DNA methylation, chromatin structure, and microRNAs. Furthermore, these elements need to operate in conjunction with the transcription factors to sustain pluripotency [60].

4. Epigenetics of Gene Expression and Silencing

By definition, epigenetic alterations modify the gene expression without changing the genetic sequence. These modifications are carried out by a variety of mechanisms, such as post-transcriptional modifications (siRNA, miRNA, and piRNA), DNA methylation, modifications in the ATP-dependent chromatin remodeling complexes, and modifications to the histone proteins (acetylation, methylation, sumoylation, phosphorylation, deamination, ribosylation, and proline isomerization) (Table 1) [24,61,62]. The process of cardiac differentiation has been shown to involve the activation of cell cycle inhibitors, cardiac-specific genes, and the inhibition of cell cycle progression and non-cardiac transcripts (Figure 1) [63,64].

4.1. Histone Modifications

Firstly, we have to take a closer look to see where histones are located, basically. To this end, we have to become familiar with the nucleosomes. The nucleosomes are the repeating units of chromatin in which two superhelical strands of DNA twist around eight core histone proteins, which consist of an H3-H4 tetramer and two H2A-H2B dimers [65]. The N- and C-terminals of the histones are the sites for acetylation, methylation, ubiquitination, sumoylation, and phosphorylation, which histones are modified by, thus, influencing the gene transcription, epigenetically [66]. All of the relevant published information related to the histone modifications of stem cells for differentiation towards cardiomyocyte cells has been categorized in a more detailed manner.

4.1.1. Histone Acetylation of Stem Cells toward Cardiomyocyte Differentiation

The acetylation of histones was first coined by Phillips [67]. Histone acetyltransferases (HATs) and histone deacetylases have competing roles in the dynamic regulation of the acetylation of histones (HDACs) [68]. The acetylation by HATs weakens the positive charge of a lysine residue and binds the histones and DNA, which leads to the additional exposure of the DNA. Therefore, acetylation is regarded as an activation marker for transcription [69].
P300, which is a HAT that suppresses the activity of the HDACs [70] and a transcriptional coactivator that interacts with: (i) GATA4 (which functions in several ways throughout the development of the heart) [71], (ii) myocyte-specific enhancer factor 2C (Mef2c), which is essential for cardiac morphogenesis, myogenesis, and angiogenesis [72], and (iii) Nkx-2.5 (which are important genes for controlling tissue differentiation and dictating the patterns of temporal and spatial development) [73]. Furthermore, p300 increases the expression of α-actinin and myosin [74]. P300 histone acetyltransferase activity contributes to embryonic myocardium development [75], and it is necessary for heart tissue regeneration as well [76].
The males absent on the first (MOF) protein acts as a HAT, which limits both mouse cardiac hypertrophy and human cardiomyopathy [77]. Another member of the HAT family is the cAMP response element-binding protein (CBP). Nonetheless, its absence does not affect the heart’s formation [78,79]. The other HAT, which is known as Gcn5, is involved in cardiac differentiation by altering the acetylation levels of H3 [80].

4.1.2. Histone Deacetylation of Stem Cells toward Cardiomyocyte Differentiation

It has been revealed that the cardiac-specific simultaneous ablation of the HDAC1 and HDAC2 genes is accompanied by several serious consequences, including dilated cardiomyopathy, cardiac arrhythmias, and neonatal lethality, while the individual deletion of each gene does not result in the development of an abnormal specific phenotype [81]. HDAC2 in combination with Hopx (homeodomain-only protein), restricts the cardiomyocytes’ proliferation via two processes of GATA4 deacetylation and reduces its transcriptional activity [82,83].
HDAC3, which is another class I HDAC, was shown to cause the thickening of the ventricular myocardium while it was overexpressed [84], and the deletion of HDAC3 was correlated with a hypertrophic effect on the cardiac cells in the mice as well [85]. HDAC3 has suppression effects on the Tbx5 activity by binding at a subset of the Tbx5-bound sites [86], which plays an indispensable role in the early life stages of cardiac development [87]. Additionally, an increase in the level of cardiac hyperplasia through the suppression of cyclin-dependent kinase inhibitors such as (p21cip1), (p27Kip1), (p57kip2), (p18inc4c), and (p15inc4b) without hypertrophy is one of the main results for the overexpression of HDAC3 [84,87].
One of the HDACs which prevents cardiac hypertrophy is HDAC4, which suppresses Mef2 and regulates cardiac hypertrophy [88,89]. The functions of HDAC5 and HDAC9 are complementary during the heart’s development via Mef2 suppression. However, impairments in their function are connected to cardiomyocyte abnormalities, thin-walled myocardium, and ventricular septal defects [90,91].
The HDACs are activated by binding the Hey proteins proximally to the transcription starting sites which leads to the suppression of gene expression by the deacetylation of histones and nuclear condensation. In cardiac myocytes, the binding of cardiac inducers activates the histone acetylases, and hence, it inactivates the Hey proteins as a consequence [92].
In humans, there are seven identified sirtuins (SIRTs) that regulate the cellular activity. SIRT1, SIRT6, and SIRT7 are mostly found in the nucleus, SIRT2 is mostly found in the cytoplasm, and it is also translocated in the nucleoplasm, and SIRT3, SIRT4, and SIRT5 are mostly found in the mitochondria [93]. SIRT1, SIRT2, SIRT6, and SIRT7 of the seven mammalian sirtuins have been shown to have key epigenetic functions [94].
Previous research has shown how the Sirt1/PGC1/pAkt axis protects the cells from oxidative stress and senescence, which may be beneficial when one is employing MSCs [95,96]. A study has revealed that Resveratrol may lessen the oxidative stress and boost the survival pathways for adipose-derived MSCs during the high glucose conditioning of the cells by upregulating the Sirt1/PGC1/pAkt axis [97]. Furthermore, pre-exposing MSCs to glucose deprivation conditions may postpone senescence, increase proliferation, and restore the ability of the aged cells to heal the senescent-infarcted myocardium [98].
Doxorubicin is a powerful antitumor anthracycline antibiotic that is routinely used to treat a wide range of tumors. However, Doxorubicin’s medical usage is restricted due to its significant cardiotoxicity, which frequently results in permanent progressive cardiomyopathy and cardiac arrest [99]. Left ventricular dysfunction represents the most prevalent consequence of Doxorubicin cardiotoxicity [100]. In a study, it was revealed that the exosomes alleviated the Doxorubicin-derived cardiomyocyte degeneration, and these positive outcomes were conveyed via the putative exosome/lncRNA-NEAT1/miR-221-3p/Sirt2 pathway [101]. Another study also pointed out that LncRNA KLF3-AS1 in the exosomes released from the MSCs may control Sirt1 to prevent cell pyroptosis and slow the development of myocardial infarction [102].
Another member of the SIRTs family, SIRT3, has a beneficial impact on cellular aging and expectancy, and it can enhance the differentiation of MSCs as well [103,104]. According to research, human MSCs become less resistant to damage as they age. The upregulation of SIRT3 may shield the MSCs from oxidative degradation by moving FOXO3a into the cell nucleus, which activates catalase and manganese superoxide dismutase [105].
Persistent AMPK activation enhances the iPSC-derived myocytes differentiation, as demonstrated by the GLUT4 and CD36 membrane recruitment and the enhanced mitochondrial function. Additionally, AMPK upregulation raises NAD/NADH levels, which might be used by SIRT1, SIRT2, and SIRT6 to reduce the lysine acetylation [106].
The acetylation of H3K56 corresponds with the increased expression of the pluripotency-associated loci in ESCs [107]. SIRT6 is a NAD-dependent histone deacetylase that addresses acetylated H3K56 in mouse ESCs [108]. One of the significant roles of SIRT6 is the expression regulation of the pluripotency-inducing genes such as Oct4, Sox2, and Nanog via the deacetylation of H3K56. This results in ESC differentiation through the Tet-mediated oxidation of 5 mC into 5hmC [109].
In another approach, hMSCs were cultivated on polycaprolactone platforms that were three-dimensionally coordinated. Based on the upregulation of the cardiovascular indicators and intracellular calcium permeability, the collagen-covered filaments were shown to be the most efficient in inducing cardiomyogenesis among the various toppings. Increased SIRT6 levels were linked to the improved differentiation on the collagen-covered filaments at the cellular scale. By triggering the Wnt signaling pathway, the siRNA-mediated SIRT6 deficiency slowed the differentiation [110].

4.1.3. Histone Methylation of Stem Cells toward Cardiomyocyte Differentiation

It was first documented that methionine is the methyl donor for N-methyl lysine in histones [111]. The variety of amino acids and mono-, di-, or trimethylation of the amino acid residues in the histones are two major elements that regulate the transcription activation. Methylation is one of the key post-transcriptional changes that happens at the H3 and H4 histones, as well as on the side chains of lysines and arginines [112,113]. For instance, the H4K20me1, H2BK5me1, and H3K36me3-enriched regions of chromatin are considered as active markers of transcription [114], while the promoters with H3K27me3 or H3K9me3 enrichment are considered to be inactive-associated regions of chromatin [51]. Generally, the transcription factors have more access to the loosened euchromatin structure, while the condensed heterochromatin structure prevents the transcription factors from accessing the genes of interest [115].
Histone methylation plays a crucial role in modifying cardiac evolution [116,117,118,119,120,121,122]. De novo mutations in the genes that alter H3K4 and H3K27 cause congenital cardiac abnormalities [123]. PAX-interacting protein 1 (PTIP), which is known as the H3K4 histone methyltransferase component, is important for altering the expression of the genes that assist in cardiac electrical conduction, notably, Kv channel-interacting protein 2 (Kcnip2) [124]. In the cardiac cells with PTIP having been knocked out, salt and calcium handle the dysregulations that take place [125], but no change happens during cardiac growth [124]. H3K4 histone methyltransferase does not affect the genes for myosin heavy chain beta (MHC-beta) and atrial natriuretic peptide (ANP), which are thought to be the cardiac hypertrophy genes [124].
Smyd1, which is a muscle-specific gene activator, methylates H3K4 and increases the levels of the Iroquois homeobox 4 and the heart and neural crest derivatives-expressed protein 2 (Hand2) (Irx4). Through the methylation of H3K4 and the overexpression of Hand2 and Irx4, Smyd1 may interact with the skeletal nascent polypeptide-associated complex alpha (skNAC). The right ventricle develops as a result of this contact, which also causes cardiomyocyte maturation [117,126,127,128,129]. It is yet unclear how directly suppressing Symd1 affects the activity of histone methyltransferase during the formation of the heart [108,111]. Symd1 is crucial for sarcomere organization by binding to myosin protein (which is known as the basic contractile unit of muscle) [127,130,131]. A neonatal cardiomyocyte histone methyltransferase of H3K4 and H3K36 is known as Smyd2. Interestingly, if Smyd2 is deficient, the other histone methyltransferases, such as Smyd1, can compensate for it [132,133].
Wolf-WHSC1, which is another histone methyltransferase, may mono-, di-, or trimethylate H3K36, which silences WHSC1. This leads to a congenital cardiac defect that is known as Wolf–Hirschhorn Syndrome, consequently [134]. The patients with WHSC1 deficiency have ventricular and atrial septal defects-related signs and symptoms [116]. Transcription is repressed when the Nkx2.5 target genes are occupied by WHSC1 via the trimethylation of H3K36 [120].
Additionally, Polycomb repressive complex 2 (PRC2) is a histone methyltransferase compound that occupies the Oct4, Sox2, and Nanog areas to activate the target genes and preserve pluripotency throughout the differentiation of the ESCs [53]. The PRC2 complex is the only identified methyltransferase of H3K27 in mouse ESCs [135] It comprises four main subsets: embryonic ectoderm development (Eed), the retinoblastoma protein-associated protein 46/48 (RbAp46/48), the catalytic subunit enhancer of Zeste 1 Ezh1/Ezh2, and the Zeste 12 (Suz12) suppressor [136].
The Ezh1 subset mediates the methylation on histone H3K27, complements the Ezh2 function, and plays a critical role in maintaining stem cell pluripotency [137]. On the other hand, Suz12 acts toward the differentiation of the stem cells, and its ablation is correlated with impaired activation of the downstream differentiation genes [138]. Contrary to Ezh2 and Eed, Eed is important for H3K27 di, tri, and monomethylation [139]. Additionally, PRC2 contributes significantly to the heart’s development.
The genes involved in the development of cardiac disease are significantly influenced by Ezh2 in terms of how they are expressed. When cardiac differentiation occurs, the Six1 transcription factor, which is active in the stem cell state, is silenced by Ezh2 [140,141]. The inactivated X chromosomes in female mammalian somatic cells have a high level of methylation H3K27 [142]. An increase in trimethylated H3K27 triggers PRC2’s steady recruitment, which causes cardiac progenitor cells to differentiate into cardiomyocytes [140]. In the yeast, S. cerevisiae, trimethylated H3K4 completely activates the promoter region of the genes (such as r1), while H3K4 demethylation enables transcription [143]. Before the transcriptional activation, the acquisition of trimethylated H3K4, and the recruitment of phosphorylated RNA polymerase II at serine 5, a monomethylated H3K4 pattern is present at the transcription sites of the genes (RNAP). This pattern of pre-activation is critical for the genes that are not regulated by the polycomb complexes [64]. The expression of Ezh1 and Ezh2 in adult and embryonic hearts is predominant, and this condition results in ceasing cardiac differentiation by gene silencing [140,141,144] [141,142,145]. Additionally, Ezh2 attaches to GATA4 and methylates it. This action has three main consequences: (i) deducting in the interaction between GATA4 and p300, (ii) the reduction in the transcriptional activity of GATA4, and (iii) the inhibition of the expression of the myosin heavy chain in fetal heart cells [145].
Moreover, the methyltransferases, G9a and GLP, silence the genes during differentiation by mono- and dimethylating H3K9 in the cardiomyocytes [146]. The G zero phase outset controls the cardiac differentiation process via the trimethylation of H3K9. Suv39h1 is involved in a quiescent state induction. To reduce the transcriptional activity, the suppressive markers such as di- and trimethylated H3K9 and trimethylated H3K27 are used. In the end, this suppression results in non-cardiac gene suppression and cell cycle halt [144].
There have also been reports of other crucial elements for the development of induced pluripotent stem cells into cardiomyocytes. Some of these supposedly essential components are tiny compounds that are used to control the epigenetic regulators. One of these is the histone methyltransferase inhibitor, BIX01294 [147], RG108 and 5-azacytidine, the DNA methyltransferase inhibitors [148,149], and valproic acid, which is a histone deacetylase inhibitor [150].

4.1.4. Histone Demethylation of Stem Cells toward Cardiomyocyte Differentiation

Jumonji proteins are essential for the growth of the cardiomyocytes [151]. A DNA-binding domain, an AT-rich interaction domain (ARID), and two conserved domains are found in the protein jumonji, which is encoded by the Jarid2 gene (JmjN and JmjC) [152]. Histone demethylation requires the JmjC domain [152,153]. Another member of this family, Jmjd6, functions as an H3 and H4 arginine histone demethylase. Studies have shown that Jmjd6 is also important in the formation of cardiac cells [154]. The trimethylated histones, H3K9 and H3K36, are demethylated by JMJD2A [155]. JMJD2A is a class of histone trimethyl demethylase that has been shown to become upregulated, while cardiac hypertrophy and long-term pressure overload happen. More accurately, in response to contractility, it is concentrated in the promoters of the atrial natriuretic peptide (ANP) and the brain natriuretic peptide (BNP) [89,156].
The ubiquitously transcribed tetratricopeptide repeat, X chromosome (UTX), an H3K27 demethylase, is the name of the second JmjC protein. The process of encoding UTX occurs on the X chromosome [157,158,159]. Cardiomyocyte development is reported to be affected following the up-regulation of UTX, which is necessary for cardiomyocyte development [102]. UTX influences the cardiomyocyte development by many methods, which are as follows: (i) by affecting the process of ESC differentiation into cardiomyocytes, (ii) by influencing the expression of cardiac-specific genes (ANP, MLC2, and a-CA) the through demethylation of trimethylated H3K27, (iii) by interacting with the various factors involved in organogenesis, such as TFs, Nkx2.5, Tbx5, GATA4, the serum response factor (SRF), and the Brg1-associated factor Baf60c (which results in the activation of cardiac-specific genes), (iv) by the demethylation of H3K4, which leads to the activation of the cardiac enhancers [122].
The aforementioned epigenetic histone markers and their impact on various cardiac-specific genes, the result of their influence on the cardiac tissue, and the utilized laboratory model are comprehensively gathered in Table 1.

4.2. Methylations

4.2.1. DNA Methylation

Although methylation is not required for the ESCs to retain their stem cell identity, it is required for the differentiation to begin [160]. The enzymes called DNA methyltransferase (dnmt1, dnmt3a, and dnmt3b) attach a methyl group to the 50-CpG-30 dinucleotides. DNA methylation takes place on the fifth carbon position of the cytosine nucleotide. When a CG dinucleotide is formed, it leads to the repression of chromatin and the inhibition of the gene expression [161]. DNA methylation might perform a variety of activities, including X-chromosome inactivation, cell differentiation, chromatin structural modifications, cancer, genomic imprinting, tissue-specific gene expression, and pluripotency in the somatic cells [161,162]. The demethylation of the genes, however, could induce pluripotency in the somatic cells via nuclear reprogramming, thus generating iPSCs. AID, which is a cytidine deaminase molecule, plays an important part in either somatic hypermutation (SHM) or class-switch recombination (CSR) during reprogramming [163]. DNA methylation is catalyzed by factors including TET enzymes (TET1, TET2, and TET3), Fe2+- and α-ketoglutarate-dependent dioxygenases [164,165].
5-formylcytosine (5fC), 5-hydroxymethylcytosine (5 hmC), and 5-carboxycytosine (5caC) are intermediates in the mechanism underlying active DNA demethylation [166,167]. These three intermediary factors are produced via 5-methylcytosine (5 mC) oxidation, which is catalyzed by the TET enzymes, affecting the methylation status of the DNA [166,167]. In order to sustain pluripotency, two important factors should increase: (i) the level of the expression of the genes that are associated with TET1 and TET2, and (ii) the oxidation of 5-methylcytosine (5 mC) to produce 5 hmC [168]. It has been revealed that during the ESC culture, the amount of 5 hmC decreases in the first four days, and then, it increases in the following four days. Simultaneously, Tet1 showed constant downregulation, and Tet3 showed constant upregulation during the period. Interestingly, Tet2 showed downregulation in the first four days of being in the culture and an upregulation during the following four days [169].

4.2.2. N6-Methyladenosine (m6a) RNA Methylation

M6a, which may occur in several forms of RNA, which include the miRNAs, represents the most common internal RNA post-transcriptional alteration in the eukaryotic cells [170]. Adenosine methyltransferases (METTL3, METTL14, and WTAP) act as writers [171,172], m6a -binding proteins act as readers, and m6a demethylating enzymes (FTO and ALKBH5) [173,174] act as erasers, and these are the key regulators of this reversible alteration [175].
It has been shown in other research that a proper hypertrophy response in the cardiomyocytes requires the methyltransferase-like 3 (METTL3)-associated methylation of the mRNA on N6-adenosines as an alteration that is amplified in correspondence to excessive stimulation. The fact that increased m6a RNA methylation produces a compensatory cardiac enlargement and a reduced m6A level of it promotes dynamic myocardium reconfiguration and impairment highlights the importance of this unique stress reaction mechanism in the heart for maintaining proper normal heart function [176].
In another study, METTL3 showed no substantial alterations in the cardiac tissue while inducing the transverse aortic constriction in a mouse model, while the FTO and WTAP expressions were attenuated. Additionally, it was discovered that cardiac hypertrophy started to develop in the mouse model 4 weeks after the transverse aortic constriction, and that the amount of total RNA that had been modified with the m6a codon increased in the hypertrophic heart tissues [177].
In another comprehensive research study, it was suggested that METTL3 was elevated in mice after they gave birth, which is the opposite of the alterations in the proliferation of the cardiomyocytes. The METTL3-pri-miR-143-(miR-143)-Yap/Ctnnd1 axis aids in m6A alteration to promote cardiac healing after an MI. To clarify, METTL3/m6A methylation loss inhibits the development of pri-miR-143 and regulates its production which results in a lower miR-143-3p yield. The transcription of Yap and Ctnnd1 (the receptors of miR-143-3p) rises significantly, promoting the multiplication of the heart cells and an innate recovery after an MI [178].
In another study, it was shown that ALKBH5 was essential for the proliferation and renewal of the cardiac myocytes. The upregulation of ALKBH5 transcription by the adeno-associated virus type 9 (AAV9) enhanced the percentage of replicating cardiac cells, decreased scar diameters, and normalized cardiac function following infarction damage, whereas ALKBH5 knockout in mice significantly restricted the rate of cardiomyocyte propagation and healing [179].
A well-studied cluster of miRNAs in the human ESC is the miR302-367 cluster. It has been revealed that stem cell-associated upstream transcription factors regulate this cluster in human ESCs. Nanog, Oct3/4, Sox2, and Rex1alleviate this cluster promoter expression [180], and consequently, the miR-302a from this cluster inhibits Cyclin D1, which is a significant G1 phase regulator in the human embryonic stem cells [181].

4.3. ATP-Dependent Chromatin Remodeling Complexes

ATP-dependent chromatin remodelers (ACRs) impact the DNA’s accessibility. They work by using ATP to break or modify the histone-DNA interaction. In other words, by sliding, twisting, or looping the nucleosomes, the ATP-dependent chromatin remodelers may modify the DNA’s accessibility [182]. These remodelers come in four families, with the switching defective/sucrose nonfermenting family receiving the most attention in the formation of the cardiomyocytes [183]. Nkx2-5, Gata4, Tbx5, and Tbx20 are a few cardiac transcription factors that interact with the Brahma-related gene 1 (Brg1)/Brahma (Brm)-associated factor complex [184]. Brg1 has a variety of functions in the development of the cardiomyocytes, including the activation of myosin heavy chains in fetuses and the inhibition of myosin heavy chains in adult cardiomyocytes and Bmp10-stimulated cardiomyocyte proliferation [185]. Tbx5, Gata4, and the Brg1/Brm-related factor component, Baf60c, have been shown to enhance the conversion of non-cardiac mesoderm into the heart muscle [186].

4.4. Cardiac Mitochondrial Mutations

In humans, cardiomyocyte development and myocardial infarction are both associated with an altered mitochondrial function [187,188]. In a study, it was discovered that the mitochondria-related genes in the cardiac cells in comparison to the other organs had different levels of methylation [189]. Numerous important processes, including ATP-dependent chromatin remodeling, are engaged in the epigenetic control of heart failure [190]. The SWI/SNF (SWItch/Sucrose Non-Fermentable) complex, which is made up of BRM and BRG1, is an ATP-dependent chromatin remodeling complex in the heart cells [191,192].
According to the research, Brg1/Brm double-mutant mice demonstrated enhanced mitophagy and dysregulated mitochondrial splitting and merging, which resulted in shattered mitochondria and increased mitochondrial biogenesis. These results outlined a function for BRG1 and BRM in mitophagy, mitochondrial dynamics, and abundance regulation as an epigenetic process [193].
The functions of mitochondrial-derived radicals in regulating the epigenetic pattern and gene function in the heart were proposed by another study. The findings demonstrated that mitochondrial-induced oxidative stress in the heart compromises myocardial DNA methylation, alters the heart gene transcription, and promotes the pathological heart changes that are indicative of cardiomyopathy [194]. Another research study found that nuclear DNA methylation was altered by a cardiac mitochondrial polymerase malfunction. The data suggest that mitochondrial impairment may affect DNA methylation as an epigenetic indicator, which is crucial for embryonic development [195].
It is documented that the pyruvate dehydrogenase stimulation that is caused by dichloroacetate may lead to epigenetic modification in the myocardium. Furthermore, dichloroacetate-dependent histone acetylation is linked to the up-regulation (by 2.3%) of the transcription-related genes, and the dichloroacetate-dependent cardiac mitochondrial production of acetyl-CoA is used to produce hydroxybutyrate, which is an innate blocker of the type I HDACs [196].
Table 1. Heart regeneration epigenetic factors, their modifications, target and effects.
Table 1. Heart regeneration epigenetic factors, their modifications, target and effects.
Epigenetic FactorsFunctionAffected GeneEffect(s) on Cardiac TissueSpeciesReferences
EvolutionPhysiologyPathologyAnomalyUnknown
HAT
p300ActivationαMHC+ Mice[74,197]
ActivationαSA+ Mice
ActivationGATA4+ Mice[71,72,73]
ActivationNkx2.5+ Mice
ActivationMef2c+ Mice
KAT2B + Human[198]
Gcn5 + Rat[80]
CBP + Mice[78,79]
MOF + Mice[77]
HDACHDAC1ActivationHopx+ Mice[82,83]
ActivationGATA4+ Mice
HDAC2ActivationHopx+ Mice
ActivationGATA4+ Mice
HDAC3InhibitionMef2 + Mice[88,89]
InhibitionCdkn1a + Mice[84]
InhibitionCdkn1b + Mice
InhibitionCdkn1c + Mice
InhibitionCdkn2b + Mice
InhibitionCdkn2c + Mice
InhibitionTbx5+ + Mice[87]
HDAC4InhibitionMef2 + Mice[88,89]
HDAC5InhibitionMef2 ++ Mice[90,91]
HDAC9InhibitionMef2 ++ Mice
SIRT1Deacetylationp21+ Mice[199]
SIRT1 + Mice
SIRT6ActivationOct4+ Mice[109]
ActivationSox2+ Mice
ActivationNanog+ Mice
PTIPActivationKcnip2 + Mice[124]
Smyd1ActivationskNAC+ Mice[117,128,129]
ActivationHand2+ Mice
ActivationIrx4+ Mice
Smyd2ActivationNCHMTs+ Mice[132]
Smyd5ActivationH4K20me3+ + Mice[200]
Wolf-WHSC1ActivationNkx2.5+ + Mice[134]
PRC2ActivationOct4+ Human[53]
ActivationSox2+ Human
ActivationNanog+ Human
ActivationEZH1+ Mice[137,138,201]
[138,139,202]
ActivationEZH2+ Mice
ActivationEed+ Mice[136]
ActivationRbAp46/48+ Mice
ActivationSuz12+ Mice
G9a + + Mice[146]
GLP + + Mice
Suv39h1 + Mice[144]
DOT1LActivationNkx2.5+ Human[202]
HDMKDM4aActivationANP + Mice[89,156]
ActivationBNP + Mice
KDM4DActivationH3K9me3+ Mice[203]
UTXActivationANP+ Mice[122]
ActivationMLC2+ Mice
Activationa-CA+ Mice
ActivationTFs+ Mice
ActivationNkx2.5+ Mice
ActivationTbx5+ Mice
ActivationGATA4+ Mice
ActivationSRF+ Mice
ActivationBaf60c+ Mice
JmjCNANA+ Mice[152]
NO66 + Mice[204]
Jmjd6 + + Mice[154]
Jmjd2aActivationBNP + Mice
Human
[205]
Jarid2ActivationNotch1+ Mice[205,206]
ActivationFn1+ Mice
Jmjd3Activationβ-MHC+ + Mice
Rat
[207,208]
PHF8Activationpmaip1+ Mice
[209]
αMHC, α-myosin heavy chain; αSA, α-sarcomeric actin; ANP, atrial natriuretic peptide; Baf60c, Brg1-associated factor 60c; BNP, brain natriuretic peptide; CBP, cAMP response element binding protein; Cdkn, cyclin-dependent kinase; CHD, Congenital heart defects; Eed, Embryonic ectoderm development; Ezh, catalytic subunit enhancer of Zeste; Fn1, Fibronectin 1; Hand2, heart- and neural crest derivatives-expressed protein 2; HAT, histone acetyltransferases; HDAC, histone deacetylases; HDM, histone demethylase; HMT, histone methyltransferase; Hopx, homeodomain-only protein; H3K9me3, lysine 9 of histone H3; Irx4, iroquois homeobox 4; Kcnip2, Kv channel-interacting protein 2; Mef2, myocyte-specific enhancer factor 2; MOF, males absent on the first; NCHMT, Neonatal cardiomyocytes histone methyltransferase; PHF19, PHD finger protein 19; PHF8, homeo domain finger protein 8; PRC2, polycomb repressive complex 2; PTIP, PAX-interacting protein 1; RbAp46/48, retinoblastoma protein associated protein 46/48; skNAC, skeletal nascent polypeptide-associated complex alpha; SRF, serum response factor; Suz12, suppressor of Zeste 12; UTX, ubiquitously transcribed tetratricopeptide repeat, X chromosome.

5. MicroRNAs (miRNAs)

5.1. miRNAs and Cardiomyocyte Differentiation

The microRNAs (miRNAs) play a larger role in iPSC differentiation and reprogramming. MiRNAs, which are small non-coding RNAs, are well understood to play a vital part in a number of processes. They are created from both the intragenic and the intergenic areas, and they may affect things such as body growth and development, the pathogenesis of autoimmune illnesses, different malignancies, hypertension, and more factors [210,211,212,213].
Furthermore, their role in respecting the pluripotency state of the stem cells has been assessed. Dicer-null and DGCR8-null ES cells are two instances of cells that are lacking mature miRNAs [214,215,216]. During differentiation, the Dicer-null cells show a very minor drop in the Oct4 expression levels [214], while the DGCR8-null ES cells exhibit an impaired development [216,217]. Thus, they are considered to be unique platforms for studying the significance of the miRNAs in pluripotency.
Just 18 miRNA families account for around 90% of the miRNAs in the heart [218]. The cardiomyocytes possess the highest concentration of miR-1 (which is a microRNA that is particular to the muscles) [219]. It has been demonstrated that a serum response factor controls its production in the rodent heart, where the levels steadily rise with the myocyte development, and they are at a peak in the adult heart as opposed to the embryonic or the neonatal stages [220]. In other words, the neonatal hypertrophic development of the heart is linked to comparatively lower levels of miR-1. It was discovered that miR-1 decreases during the stimulation of hypertrophy, perhaps through a pathway that is reliant upon a serum response factor [221].
Although the miR-133 expression is less than that of miR-1, both miRs are derived from the same bicistronic gene. MiR-1 and miR-133 in the ESCs collaborate to promote mesoblast development in the embryonic stem cells. Later on in their growth, they carry out opposite roles: miR-1 encourages the mesoblast development into cardiomyocytes, whereas miR-133 prevents it [222].
The levels of miRNA-1 and miRNA-133a steadily rise as the heart develops. Additionally, the induced amplification of miRNA-133a inhibits myocardial development, whilst miRNA-1 promotes it. The overexpression of both of the miRNAs increases the mesodermal commitment. The reduced production of miRNA-1 and miRNA-133a shows that both of the miRNAs have the opposite effect on heart growth compared to when they are upregulated [223]. Furthermore, the data show miRNA-1 and miRNA-133 control the transcription of Kdm6A/B and Ezh2 [224]. The adjustment of the Fgf8 activation, which is controlled by Bmp2, is regulated by miRNA-130 and miRNA-133. This regulatory feedback is necessary for attaining initial cardiac differentiation [225].
In the development of cardiomyocytes, miRNA-128 appears to play a direct function. It was demonstrated that the loss of functionality of miRNA-128 did not affect the atrial myocytes. Interestingly, the ventricles of miR-128 KO zebrafish larvae became smaller. The important cardiac transcription factors had their expression levels differently altered, which reflected this misdirected cardiogenesis in vitro (Isl1, Sfrp5, Nkx2.5, Mef2c, Irx4, and Hcn4). By influencing the timing of the cardiac progenitor cell differentiation into different cardiomyocyte subtypes, it was shown that miR-128a performs a yet unidentified function during early heart development [226].
MiR-134 may regulate the proliferation of human cardiomyocyte progenitor cells, as evidenced by the discovery that Meis2 is a target of the miRNA. This suggests that miR-134 may be involved in cardiogenesis [227].
Other investigations suggested that the miR-222 activity performs protection against negative ventricular remodeling and cardiac dysfunction following an ischemia damage. Additionally, these benefits were linked to the suppression of cardiomyocyte death and a significant reduction in fibrosis long after the ischemia damage [228]. Delivering miR-222 in conjunction with three other miRNAs also enhanced the calcium absorption rate and the sarcomere orientation, and consequently, the resting voltage became more negative and the production of the proteins in favor of cardiomyocyte development rose [229].
MicroRNA-294 was shown to have a unique function in the cell cycle of heart cells and myocardial regeneration. After a myocardial damage, the temporary expression of miRNA-294 promotes pro-reparative alterations in the cardiac tissue along with the restart of the myocyte cell cycle, improving the cardiac shape and function. Along with an increased cardiac cell cycle rate, an enhanced vasculature, and a decreased stroke lesion volume in the myocardium following the trauma, miR-294 also increases the cell viability [230].
While the miR-31a-5p transcription in cardiac myocytes was substantially elevated on the tenth postnatal day when it was compared to that of the first postnatal day, it was unexpectedly shown that miR-31a-5p promoted cardiac growth, whereas miRNA-31a-5p suppression inhibited proliferation. Using intact neonatal rat ventricular cardiac cells, this miRNA is the first miRNA from a subsequent myocardial development study to be shown to enhance cell proliferation [231].
In terms of the suppressive effect side of the miRNAs on their targets which comprise a major part of their functions, the results show that miR-200c inhibits the development and maturation of the hESCs. First of all, it blocks the ion channel and the transcription factors translation when it is acting alone. Second, miR-200c regulates many gene programs that are linked to heart growth and function by regulating the cardiac transcription factors [232].
Since MiR-375 is produced at the beginning of the cardiac development process, its upregulation greatly increases the mortality and deformity rates. Additionally, there was a considerable downregulation in the expression levels of the genes linked to heart growth. The investigations have revealed that the excessive transcription of miR-375 may result in deficiencies in the morphology of several organs and abnormalities of the downstream NOTCH signaling pathway (which shows potential as being a CHD therapeutic target and a diagnostic marker) [233].
In a zebrafish model, it was shown that miR-25 targets the cell cycle regulator FBXW7, and that miR-25 overexpression encourages cardiomyocyte multiplication by suppressing FBXW7. As a result, miR-25 may be a cutting-edge agent for heart regeneration [234]. In the research that was conducted by Yu et al., it was shown that miR-23 overexpression dramatically increased the proliferation of the AC16 cells and decreased cell apoptosis. According to these findings, miR-23 decreased the cardiomyocyte death rate, while it promoted proliferation by targeting TGF-1 [235].

5.2. miRNAs and Congenital Heart Defects

According to the strong findings, the miRNAs may be aberrantly expressed during the development of congenital heart defects (CHD), which are one of the most prevalent inherent abnormalities in newborns and children [236]. The regulation of cardiac morphogenesis, conductivities, and the cell-cycle adjustment are all influenced by the disruption in the miRNA synthesis [237].
For instance, DNA methyltransferases and the miR-29b-3p levels were negatively correlated in the CHD patients [238]. For cardiac cell proliferation and cardiac abnormalities, certain miRNAs (hsa-miR-221-3p, hsa-miR-218-5p, and hsa-miR-873-5p) have been described in [239]. The data support the idea that the CHD patients’ illness state is caused by a changed miRNA expression. In another study, it was revealed that hsa-miR-148a and BCL2L11 are linked to cardiomyocyte apoptosis and Hsa-miR-148a might affect BCL2L11 [240]. Recent findings have shown that miR-153-3p targets βII spectrin, which has a negative effect on myocardial development [241].
Researchers have found that improving the myocardial index, histological alterations in the myocardial tissue, and the apoptosis level by up-regulating miR-30c-5p or down-regulating BCL2L11 might have a molecular therapeutic impact on people with congenital heart defects [242]. In another related research, it was discovered that inhibiting miR-375-3p targeted and elevated FOXP1 and Bcl2l2 to suppress an hypoxia-evoked death of the cardiomyocytes. These findings may point to maternal serum, miR375-3p, as a promising biomarker for early fetal CHD identification [243]. All of these discoveries could pave the way for the future use of these miRNAs as diagnostics and targets for the detection and management of cardiac disorders.
Various functions of the related miRNAs involved in the differentiation, proliferation, development, and possible reprogramming of the cardiomyocytes are summarized in Figure 2 and Table 2.

6. Conclusions

In this review, we have discussed numerous strategies for applying in iPSC and ESC-based future investigations that focus on cardiomyocyte regeneration. These approaches consist of various post-transcriptional modifications involving the DNA, RNAs, histone modifications, the ATP-dependent chromatin remodeling complexes, and mutations in the mitochondria of the cardiac cells. These studies together certainly provide a broad range of elements that should be taken into consideration while one is performing heart regeneration studies in the future.
There are other methods outside of this review for pursuing cardiac cell regeneration as well. For instance, an approach for reactivating the ability of the cardiomyocytes toward regeneration is the utilization of extracellular vesicles (EVs), which also deliver miRNAs as part of their cargo. In recent research, it was shown that human and mouse epicardial cells secrete EVs, and that these EVs may stimulate the proliferation of cardiac myocytes in vitro. The effective epicardial EV absorption by heart cells was shown to encourage cell cycle activation [288]. Another study suggested that miRNA-21-loaded EVs persistently prevented cell death, and they have significantly enhanced the cardiovascular output in an animal model of preclinical MI by injecting produced EVs as well [289].
Various types of cells have been discovered to release the EVs that encourage the cardiomyocyte cell cycle activity as well [290,291]. Other studies have supported the hypothesis that turning the attention away from the iPSCs and the ESCs and onto their secreted EVs may be the answer to the ever-increasing problems that are associated with dealing with stem cells [292].
Another intriguing approach was the utilization of exosome-derived miRNA-19a/19b and the MSCs, simultaneously, which significantly improved the restoration of the cardiovascular output and decreased the heart fibrosis in an infarction model [293]. A final solution has not been advised by any of these methods, even though they have all shown us extremely encouraging results. Therefore, further preliminary studies must be conducted to evaluate the efficacy of all of these treatments and their effects on the various cell signaling pathways in the heart muscle in the long run.
The discrepancies between the complicated in vivo environments and the in vitro experiments could be effectively reduced by using microfluidic technology as well [294]. For instance, tissue engineering microfluidic devices which can be utilized while executing electrical stimulation [295], mechanical stimulations (such as a cell stretching device with special substrates or by providing shear stress) [296,297], biochemical stimulation [298], physical factors, and structural stimulations [5] have been suggested to simulate the natural heart environment. In a study, the researchers proposed a system with the capability to deliver every stimulation separately or to integrate three separate impulses to examine the collaboration of multiple stimuli, which would better simulate the complicated in vivo environment [299].
Biochemical reagents for the differentiation of the MSCs are utilized commonly as well. For instance, Azacytidine (5-AZA) has been recognized as a biochemical reagent for reprogramming cells toward contractile striated muscle cells since 1979 [300]. It is still in use today, with new approaches having been taken for the better differentiation of the MSCs into cardiomyocytes. To differentiate the hMSCs, pharmacological agents such as the vascular endothelial growth factor (VEGF) and 5-AZA were utilized in one study by using a microfluidic system, which is capable of the precise regulation of the media flow, its timing, the pattern of the reagent supply, and the incubation period of the cultured cells. The authors have announced promising results using this combined approach as well [301].
Another study supported the idea that TGF-β1 is non-toxic to cells, effective, and may be a better inducer of stem cell to cardiomyocyte differentiation than 5-AZA [302] due to the effect of it on recognized TGF/BMP signaling pathway versus that of 5-AZA, which is a demethylating agent with poorly characterized mechanism [303,304].
Collectively, it should be considered that the cellular micro-environment is composed of an intricate network of stem cells, satellite cells, bioactive substances, and the extracellular matrix. The microenvironment is influenced by protein adhesives (such as laminin, fibronectin, and vitronectin), fibrous structural proteins (such as elastin and collagen), polysaccharides in the form of glycosaminoglycan, and proteoglycan additionally. All of these variables should be precisely taken into account while one is conducting stem cell differentiation experiments [301,305].

Author Contributions

A.T., M.M. and R.S. conceived and designed the format of the manuscript. A.S., A.Z., A.K., S.B., V.N., S.A.-T., A.T. and R.S. collected the data and drafted and edited the manuscript. A.S., A.Z. and A.K. drew the figures and tables. All of the authors reviewed the manuscript and all of them contributed to the critical reading and discussion of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research has received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used to support the findings of this study are included within the article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

MSCs Mesenchymal stem cells
hMSCs Human mesenchymal stem cells
ESCs Embryonic stem cells
iPSCs Induced pluripotent stem cells
miRNAs MicroRNAs
siRNA Small interfering RNA
piRNA Piwi-interacting RNA
HATs Histone acetyltransferases
HDACs Histone deacetylases
MOF Males absent on the first
SIRTs Sirtuins
SIRT1 Sirtuin 1
SIRT2 Sirtuin 2
SIRT3 Sirtuin 3
SIRT4 Sirtuin 4
SIRT5 Sirtuin 5
SIRT6 Sirtuin 6
SIRT7 Sirtuin 7
PTIP PAX-interacting protein 1
PRC2 Polycomb repressive complex 2
UTX Ubiquitously transcribed tetratricopeptide repeat, X chromosome
CHD Congenital heart defect
EVs Extracellular vesicles
VEGF vascular endothelial growth factor
5-AZA 5-azacytidine

References

  1. Virani, S.S.; Alonso, A.; Aparicio, H.J.; Benjamin, E.J.; Bittencourt, M.S.; Callaway, C.W.; Carson, A.P.; Chamberlain, A.M.; Cheng, S.; Delling, F.N.; et al. Heart Disease and Stroke Statistics-2021 Update: A Report From the American Heart Association. Circulation 2021, 143, e254–e743. [Google Scholar] [CrossRef] [PubMed]
  2. Hall, M.J.; Levant, S.; DeFrances, C.J. Hospitalization for Congestive Heart Failure: United States, 2000–2010; NCHS Data Brief: Washington, DC, USA; US Department of Health and Human Services, Centers for Disease Control and Prevention, National Center for Health Statistics: Washington, DC, USA, 2012; pp. 1–8.
  3. Benjamin, E.J.; Virani, S.S.; Callaway, C.W.; Chamberlain, A.M.; Chang, A.R.; Cheng, S.; Chiuve, S.E.; Cushman, M.; Delling, F.N.; Deo, R.; et al. Heart Disease and Stroke Statistics-2018 Update: A Report From the American Heart Association. Circulation 2018, 137, e67–e492. [Google Scholar] [CrossRef] [PubMed]
  4. Verhestraeten, C.; Heggermont, W.A.; Maris, M. Clinical inertia in the treatment of heart failure: A major issue to tackle. Heart. Fail. Rev. 2021, 26, 1359–1370. [Google Scholar] [CrossRef] [PubMed]
  5. Karbassi, E.; Fenix, A.; Marchiano, S.; Muraoka, N.; Nakamura, K.; Yang, X.; Murry, C.E. Cardiomyocyte maturation: Advances in knowledge and implications for regenerative medicine. Nat. Rev. Cardiol. 2020, 17, 341–359. [Google Scholar] [CrossRef]
  6. Yang, X.; Pabon, L.; Murry, C.E. Engineering adolescence: Maturation of human pluripotent stem cell-derived cardiomyocytes. Circ. Res. 2014, 114, 511–523. [Google Scholar] [CrossRef] [Green Version]
  7. Ahuja, P.; Sdek, P.; MacLellan, W.R. Cardiac myocyte cell cycle control in development, disease, and regeneration. Physiol. Rev. 2007, 87, 521–544. [Google Scholar] [CrossRef] [Green Version]
  8. Mollova, M.; Bersell, K.; Walsh, S.; Savla, J.; Das, L.T.; Park, S.Y.; Silberstein, L.E.; Dos Remedios, C.G.; Graham, D.; Colan, S.; et al. Cardiomyocyte proliferation contributes to heart growth in young humans. Proc. Natl. Acad. Sci. USA 2013, 110, 1446–1451. [Google Scholar] [CrossRef] [Green Version]
  9. Naqvi, N.; Li, M.; Calvert, J.W.; Tejada, T.; Lambert, J.P.; Wu, J.; Kesteven, S.H.; Holman, S.R.; Matsuda, T.; Lovelock, J.D.; et al. A proliferative burst during preadolescence establishes the final cardiomyocyte number. Cell 2014, 157, 795–807. [Google Scholar] [CrossRef] [Green Version]
  10. Zuk, P.A.; Zhu, M.; Mizuno, H.; Huang, J.; Futrell, J.W.; Katz, A.J.; Benhaim, P.; Lorenz, H.P.; Hedrick, M.H. Multilineage cells from human adipose tissue: Implications for cell-based therapies. Tissue Eng. 2001, 7, 211–228. [Google Scholar] [CrossRef] [Green Version]
  11. Dolati, S.; Yousefi, M.; Mahdipour, M.; Afrasiabi Rad, A.; Pishgahi, A.; Nouri, M.; Jodati, A.R. Mesenchymal stem cell and bone marrow mononuclear cell therapy for cardiomyopathy: From bench to bedside. J. Cell. Biochem. 2019, 120, 45–55. [Google Scholar] [CrossRef] [Green Version]
  12. Bearzi, C.; Rota, M.; Hosoda, T.; Tillmanns, J.; Nascimbene, A.; De Angelis, A.; Yasuzawa-Amano, S.; Trofimova, I.; Siggins, R.W.; Lecapitaine, N.; et al. Human cardiac stem cells. Proc. Natl. Acad. Sci. USA 2007, 104, 14068–14073. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Cambria, E.; Pasqualini, F.S.; Wolint, P.; Gunter, J.; Steiger, J.; Bopp, A.; Hoerstrup, S.P.; Emmert, M.Y. Translational cardiac stem cell therapy: Advancing from first-generation to next-generation cell types. NPJ Regen. Med. 2017, 2, 17. [Google Scholar] [CrossRef]
  14. Evans, M.J.; Kaufman, M.H. Establishment in culture of pluripotential cells from mouse embryos. Nature 1981, 292, 154–156. [Google Scholar] [CrossRef] [PubMed]
  15. McLaren, A. Ethical and social considerations of stem cell research. Nature 2001, 414, 129–131. [Google Scholar] [CrossRef] [PubMed]
  16. 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] [Green Version]
  17. Bernstein, H.S.; Srivastava, D. Stem cell therapy for cardiac disease. Pediatr. Res. 2012, 71, 491–499. [Google Scholar] [CrossRef] [Green Version]
  18. Mummery, C.; Ward-van Oostwaard, D.; Doevendans, P.; Spijker, R.; van den Brink, S.; Hassink, R.; van der Heyden, M.; Opthof, T.; Pera, M.; de la Riviere, A.B.; et al. Differentiation of human embryonic stem cells to cardiomyocytes: Role of coculture with visceral endoderm-like cells. Circulation 2003, 107, 2733–2740. [Google Scholar] [CrossRef] [Green Version]
  19. Graichen, R.; Xu, X.; Braam, S.R.; Balakrishnan, T.; Norfiza, S.; Sieh, S.; Soo, S.Y.; Tham, S.C.; Mummery, C.; Colman, A.; et al. Enhanced cardiomyogenesis of human embryonic stem cells by a small molecular inhibitor of p38 MAPK. Differentiation 2008, 76, 357–370. [Google Scholar] [CrossRef]
  20. Kawamura, M.; Miyagawa, S.; Miki, K.; Saito, A.; Fukushima, S.; Higuchi, T.; Kawamura, T.; Kuratani, T.; Daimon, T.; Shimizu, T.; et al. Feasibility, safety, and therapeutic efficacy of human induced pluripotent stem cell-derived cardiomyocyte sheets in a porcine ischemic cardiomyopathy model. Circulation 2012, 126, S29–S37. [Google Scholar] [CrossRef] [Green Version]
  21. Mummery, C.L.; Zhang, J.; Ng, E.S.; Elliott, D.A.; Elefanty, A.G.; Kamp, T.J. Differentiation of human embryonic stem cells and induced pluripotent stem cells to cardiomyocytes: A methods overview. Circ. Res. 2012, 111, 344–358. [Google Scholar] [CrossRef]
  22. 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]
  23. Kim, Y.J.; Tamadon, A.; Kim, Y.Y.; Kang, B.C.; Ku, S.Y. Epigenetic Regulation of Cardiomyocyte Differentiation from Embryonic and Induced Pluripotent Stem Cells. Int. J. Mol. Sci. 2021, 22, 8599. [Google Scholar] [CrossRef] [PubMed]
  24. Chen, T.; Dent, S.Y. Chromatin modifiers and remodellers: Regulators of cellular differentiation. Nat. Rev. Genet. 2014, 15, 93–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Li, G.; Reinberg, D. Chromatin higher-order structures and gene regulation. Curr. Opin. Genet. Dev. 2011, 21, 175–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Kooistra, S.M.; Helin, K. Molecular mechanisms and potential functions of histone demethylases. Nat. Rev. Mol. Cell Biol. 2012, 13, 297–311. [Google Scholar] [CrossRef] [PubMed]
  27. Paige, S.L.; Plonowska, K.; Xu, A.; Wu, S.M. Molecular regulation of cardiomyocyte differentiation. Circ. Res. 2015, 116, 341–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Bartel, D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell 2004, 116, 281–297. [Google Scholar] [CrossRef] [Green Version]
  29. Castell-Auvi, A.; Cedo, L.; Movassat, J.; Portha, B.; Sanchez-Cabo, F.; Pallares, V.; Blay, M.; Pinent, M.; Ardevol, A. Procyanidins modulate microRNA expression in pancreatic islets. J. Agric Food Chem. 2013, 61, 355–363. [Google Scholar] [CrossRef]
  30. Plotnikova, O.; Baranova, A.; Skoblov, M. Comprehensive Analysis of Human microRNA-mRNA Interactome. Front. Genet. 2019, 10, 933. [Google Scholar] [CrossRef]
  31. Friedman, R.C.; Farh, K.K.; Burge, C.B.; Bartel, D.P. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2009, 19, 92–105. [Google Scholar] [CrossRef] [Green Version]
  32. Boon, R.A.; Dimmeler, S. MicroRNAs in myocardial infarction. Nat. Rev. Cardiol. 2015, 12, 135–142. [Google Scholar] [CrossRef] [PubMed]
  33. Sluijter, J.P.; van Mil, A.; van Vliet, P.; Metz, C.H.; Liu, J.; Doevendans, P.A.; Goumans, M.J. MicroRNA-1 and -499 regulate differentiation and proliferation in human-derived cardiomyocyte progenitor cells. Arterioscler. Thromb. Vasc. Biol. 2010, 30, 859–868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Ivey, K.N.; Muth, A.; Arnold, J.; King, F.W.; Yeh, R.F.; Fish, J.E.; Hsiao, E.C.; Schwartz, R.J.; Conklin, B.R.; Bernstein, H.S.; et al. MicroRNA regulation of cell lineages in mouse and human embryonic stem cells. Cell Stem Cell 2008, 2, 219–229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Zhu, K.; Liu, D.; Lai, H.; Li, J.; Wang, C. Developing miRNA therapeutics for cardiac repair in ischemic heart disease. J. Thorac. Dis. 2016, 8, E918–E927. [Google Scholar] [CrossRef] [Green Version]
  36. Fard, M.; Akhavan-Tavakoli, M.; Khanjani, S.; Zare, S.; Edalatkhah, H.; Arasteh, S.; Mehrabani, D.; Zarnani, A.-H.; Kazemnejad, S.; Shirazi, R. Bilayer amniotic membrane/nano-fibrous fibroin scaffold promotes differentiation capability of menstrual blood stem cells into keratinocyte-like cells. Mol. Biotechnol. 2018, 60, 100–110. [Google Scholar] [CrossRef]
  37. Kashani, I.R.; Zarnani, A.H.; Soleimani, M.; Abdolvahabi, M.A.; Nayernia, K.; Shirazi, R. Retinoic acid induces mouse bone marrow-derived CD15+, Oct4+ and CXCR4+ stem cells into male germ-like cells in a two-dimensional cell culture system. Cell Biol. Int. 2014, 38, 782–789. [Google Scholar] [CrossRef]
  38. Kofidis, T.; de Bruin, J.L.; Yamane, T.; Tanaka, M.; Lebl, D.R.; Swijnenburg, R.J.; Weissman, I.L.; Robbins, R.C. Stimulation of paracrine pathways with growth factors enhances embryonic stem cell engraftment and host-specific differentiation in the heart after ischemic myocardial injury. Circulation 2005, 111, 2486–2493. [Google Scholar] [CrossRef] [Green Version]
  39. Rajasingh, J.; Bord, E.; Hamada, H.; Lambers, E.; Qin, G.; Losordo, D.W.; Kishore, R. STAT3-dependent mouse embryonic stem cell differentiation into cardiomyocytes: Analysis of molecular signaling and therapeutic efficacy of cardiomyocyte precommitted mES transplantation in a mouse model of myocardial infarction. Circ. Res. 2007, 101, 910–918. [Google Scholar] [CrossRef] [Green Version]
  40. Park, I.H.; Zhao, R.; West, J.A.; Yabuuchi, A.; Huo, H.; Ince, T.A.; Lerou, P.H.; Lensch, M.W.; Daley, G.Q. Reprogramming of human somatic cells to pluripotency with defined factors. Nature 2008, 451, 141–146. [Google Scholar] [CrossRef]
  41. Takahashi, K.; Okita, K.; Nakagawa, M.; Yamanaka, S. Induction of pluripotent stem cells from fibroblast cultures. Nat. Protoc. 2007, 2, 3081–3089. [Google Scholar] [CrossRef]
  42. Takahashi, K.; Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006, 126, 663–676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. 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]
  44. Brambrink, T.; Foreman, R.; Welstead, G.G.; Lengner, C.J.; Wernig, M.; Suh, H.; Jaenisch, R. Sequential expression of pluripotency markers during direct reprogramming of mouse somatic cells. Cell Stem Cell 2008, 2, 151–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Stadtfeld, M.; Nagaya, M.; Utikal, J.; Weir, G.; Hochedlinger, K. Induced pluripotent stem cells generated without viral integration. Science 2008, 322, 945–949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Woltjen, K.; Michael, I.P.; Mohseni, P.; Desai, R.; Mileikovsky, M.; Hamalainen, R.; Cowling, R.; Wang, W.; Liu, P.; Gertsenstein, M.; et al. piggyBac transposition reprograms fibroblasts to induced pluripotent stem cells. Nature 2009, 458, 766–770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Rajasingh, J.; Lambers, E.; Hamada, H.; Bord, E.; Thorne, T.; Goukassian, I.; Krishnamurthy, P.; Rosen, K.M.; Ahluwalia, D.; Zhu, Y.; et al. Cell-free embryonic stem cell extract-mediated derivation of multipotent stem cells from NIH3T3 fibroblasts for functional and anatomical ischemic tissue repair. Circ. Res. 2008, 102, e107–e117. [Google Scholar] [CrossRef]
  48. Ban, H.; Nishishita, N.; Fusaki, N.; Tabata, T.; Saeki, K.; Shikamura, M.; Takada, N.; Inoue, M.; Hasegawa, M.; Kawamata, S.; et al. Efficient generation of transgene-free human induced pluripotent stem cells (iPSCs) by temperature-sensitive Sendai virus vectors. Proc. Natl. Acad. Sci. USA 2011, 108, 14234–14239. [Google Scholar] [CrossRef] [Green Version]
  49. Stadtfeld, M.; Hochedlinger, K. Induced pluripotency: History, mechanisms, and applications. Genes Dev. 2010, 24, 2239–2263. [Google Scholar] [CrossRef] [Green Version]
  50. Yagi, M.; Yamanaka, S.; Yamada, Y. Epigenetic foundations of pluripotent stem cells that recapitulate in vivo pluripotency. Lab. Investig. 2017, 97, 1133–1141. [Google Scholar] [CrossRef]
  51. Bernstein, B.E.; Mikkelsen, T.S.; Xie, X.; Kamal, M.; Huebert, D.J.; Cuff, J.; Fry, B.; Meissner, A.; Wernig, M.; Plath, K.; et al. A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell 2006, 125, 315–326. [Google Scholar] [CrossRef] [Green Version]
  52. Holliday, R. Epigenetics: A historical overview. Epigenetics 2006, 1, 76–80. [Google Scholar] [CrossRef] [Green Version]
  53. Lee, T.I.; Jenner, R.G.; Boyer, L.A.; Guenther, M.G.; Levine, S.S.; Kumar, R.M.; Chevalier, B.; Johnstone, S.E.; Cole, M.F.; Isono, K.; et al. Control of developmental regulators by Polycomb in human embryonic stem cells. Cell 2006, 125, 301–313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Ma, P.; Schultz, R.M. Histone deacetylase 1 (HDAC1) regulates histone acetylation, development, and gene expression in preimplantation mouse embryos. Dev. Biol. 2008, 319, 110–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Mikkelsen, T.S.; Ku, M.; Jaffe, D.B.; Issac, B.; Lieberman, E.; Giannoukos, G.; Alvarez, P.; Brockman, W.; Kim, T.K.; Koche, R.P.; et al. Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature 2007, 448, 553–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Vakoc, C.R.; Sachdeva, M.M.; Wang, H.; Blobel, G.A. Profile of histone lysine methylation across transcribed mammalian chromatin. Mol. Cell. Biol. 2006, 26, 9185–9195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Xi, S.; Geiman, T.M.; Briones, V.; Guang Tao, Y.; Xu, H.; Muegge, K. Lsh participates in DNA methylation and silencing of stem cell genes. Stem Cells 2009, 27, 2691–2702. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Hajkova, P. Epigenetic reprogramming in the germline: Towards the ground state of the epigenome. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2011, 366, 2266–2273. [Google Scholar] [CrossRef] [Green Version]
  59. Simon, J.A.; Kingston, R.E. Mechanisms of polycomb gene silencing: Knowns and unknowns. Nat. Rev. Mol. Cell Biol. 2009, 10, 697–708. [Google Scholar] [CrossRef] [PubMed]
  60. Bibikova, M.; Laurent, L.C.; Ren, B.; Loring, J.F.; Fan, J.B. Unraveling epigenetic regulation in embryonic stem cells. Cell Stem Cell 2008, 2, 123–134. [Google Scholar] [CrossRef] [Green Version]
  61. Mahdipour, M.; van Tol, H.T.; Stout, T.A.; Roelen, B.A. Validating reference microRNAs for normalizing qRT-PCR data in bovine oocytes and preimplantation embryos. BMC Dev. Biol. 2015, 15, 25. [Google Scholar] [CrossRef] [Green Version]
  62. Roovers, E.F.; Rosenkranz, D.; Mahdipour, M.; Han, C.T.; He, N.; Chuva de Sousa Lopes, S.M.; van der Westerlaken, L.A.; Zischler, H.; Butter, F.; Roelen, B.A.; et al. Piwi proteins and piRNAs in mammalian oocytes and early embryos. Cell Rep. 2015, 10, 2069–2082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Paige, S.L.; Thomas, S.; Stoick-Cooper, C.L.; Wang, H.; Maves, L.; Sandstrom, R.; Pabon, L.; Reinecke, H.; Pratt, G.; Keller, G.; et al. A temporal chromatin signature in human embryonic stem cells identifies regulators of cardiac development. Cell 2012, 151, 221–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Wamstad, J.A.; Alexander, J.M.; Truty, R.M.; Shrikumar, A.; Li, F.; Eilertson, K.E.; Ding, H.; Wylie, J.N.; Pico, A.R.; Capra, J.A.; et al. Dynamic and coordinated epigenetic regulation of developmental transitions in the cardiac lineage. Cell 2012, 151, 206–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Luger, K.; Mader, A.W.; Richmond, R.K.; Sargent, D.F.; Richmond, T.J. Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 1997, 389, 251–260. [Google Scholar] [CrossRef]
  66. Strahl, B.D.; Allis, C.D. The language of covalent histone modifications. Nature 2000, 403, 41–45. [Google Scholar] [CrossRef]
  67. Phillips, D.M. The presence of acetyl groups of histones. Biochem. J. 1963, 87, 258–263. [Google Scholar] [CrossRef] [Green Version]
  68. Xhemalce, B.; Dawson, M.A.; Bannister, A.J. Histone Modifications. In Encyclopedia of Molecular Cell Biology and Molecular Medicine; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2011. [Google Scholar] [CrossRef]
  69. Clayton, A.L.; Hebbes, T.R.; Thorne, A.W.; Crane-Robinson, C. Histone acetylation and gene induction in human cells. FEBS Lett. 1993, 336, 23–26. [Google Scholar] [CrossRef] [Green Version]
  70. Sartorelli, V.; Huang, J.; Hamamori, Y.; Kedes, L. Molecular mechanisms of myogenic coactivation by p300: Direct interaction with the activation domain of MyoD and with the MADS box of MEF2C. Mol. Cell. Biol. 1997, 17, 1010–1026. [Google Scholar] [CrossRef] [Green Version]
  71. Kawamura, T.; Ono, K.; Morimoto, T.; Wada, H.; Hirai, M.; Hidaka, K.; Morisaki, T.; Heike, T.; Nakahata, T.; Kita, T.; et al. Acetylation of GATA-4 is involved in the differentiation of embryonic stem cells into cardiac myocytes. J. Biol. Chem. 2005, 280, 19682–19688. [Google Scholar] [CrossRef] [Green Version]
  72. Ma, K.; Chan, J.K.; Zhu, G.; Wu, Z. Myocyte enhancer factor 2 acetylation by p300 enhances its DNA binding activity, transcriptional activity, and myogenic differentiation. Mol. Cell. Biol. 2005, 25, 3575–3582. [Google Scholar] [CrossRef] [Green Version]
  73. Sun, H.; Yang, X.; Zhu, J.; Lv, T.; Chen, Y.; Chen, G.; Zhong, L.; Li, Y.; Huang, X.; Huang, G.; et al. Inhibition of p300-HAT results in a reduced histone acetylation and down-regulation of gene expression in cardiac myocytes. Life Sci. 2010, 87, 707–714. [Google Scholar] [CrossRef] [PubMed]
  74. Yao, T.P.; Oh, S.P.; Fuchs, M.; Zhou, N.D.; Ch’ng, L.E.; Newsome, D.; Bronson, R.T.; Li, E.; Livingston, D.M.; Eckner, R. Gene dosage-dependent embryonic development and proliferation defects in mice lacking the transcriptional integrator p300. Cell 1998, 93, 361–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Shikama, N.; Lutz, W.; Kretzschmar, R.; Sauter, N.; Roth, J.F.; Marino, S.; Wittwer, J.; Scheidweiler, A.; Eckner, R. Essential function of p300 acetyltransferase activity in heart, lung and small intestine formation. EMBO J. 2003, 22, 5175–5185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Schueler, M.; Zhang, Q.; Schlesinger, J.; Tonjes, M.; Sperling, S.R. Dynamics of Srf, p300 and histone modifications during cardiac maturation in mouse. Mol. Biosyst. 2012, 8, 495–503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Qiao, W.; Zhang, W.; Gai, Y.; Zhao, L.; Fan, J. The histone acetyltransferase MOF overexpression blunts cardiac hypertrophy by targeting ROS in mice. Biochem. Biophys. Res. Commun. 2014, 448, 379–384. [Google Scholar] [CrossRef] [PubMed]
  78. Tanaka, Y.; Naruse, I.; Hongo, T.; Xu, M.; Nakahata, T.; Maekawa, T.; Ishii, S. Extensive brain hemorrhage and embryonic lethality in a mouse null mutant of CREB-binding protein. Mech. Dev. 2000, 95, 133–145. [Google Scholar] [CrossRef] [PubMed]
  79. Chen, G.; Zhu, J.; Lv, T.; Wu, G.; Sun, H.; Huang, X.; Tian, J. Spatiotemporal expression of histone acetyltransferases, p300 and CBP, in developing embryonic hearts. J. Biomed. Sci. 2009, 16, 24. [Google Scholar] [CrossRef] [Green Version]
  80. Li, L.; Zhu, J.; Tian, J.; Liu, X.; Feng, C. A role for Gcn5 in cardiomyocyte differentiation of rat mesenchymal stem cells. Mol. Cell. Biochem. 2010, 345, 309–316. [Google Scholar] [CrossRef]
  81. Montgomery, R.L.; Davis, C.A.; Potthoff, M.J.; Haberland, M.; Fielitz, J.; Qi, X.; Hill, J.A.; Richardson, J.A.; Olson, E.N. Histone deacetylases 1 and 2 redundantly regulate cardiac morphogenesis, growth, and contractility. Genes Dev. 2007, 21, 1790–1802. [Google Scholar] [CrossRef] [Green Version]
  82. Trivedi, C.M.; Zhu, W.; Wang, Q.; Jia, C.; Kee, H.J.; Li, L.; Hannenhalli, S.; Epstein, J.A. Hopx and Hdac2 interact to modulate Gata4 acetylation and embryonic cardiac myocyte proliferation. Dev. Cell 2010, 19, 450–459. [Google Scholar] [CrossRef] [Green Version]
  83. Chen, F.; Kook, H.; Milewski, R.; Gitler, A.D.; Lu, M.M.; Li, J.; Nazarian, R.; Schnepp, R.; Jen, K.; Biben, C.; et al. Hop is an unusual homeobox gene that modulates cardiac development. Cell 2002, 110, 713–723. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Trivedi, C.M.; Lu, M.M.; Wang, Q.; Epstein, J.A. Transgenic overexpression of Hdac3 in the heart produces increased postnatal cardiac myocyte proliferation but does not induce hypertrophy. J. Biol. Chem. 2008, 283, 26484–26489. [Google Scholar] [CrossRef] [Green Version]
  85. Montgomery, R.L.; Potthoff, M.J.; Haberland, M.; Qi, X.; Matsuzaki, S.; Humphries, K.M.; Richardson, J.A.; Bassel-Duby, R.; Olson, E.N. Maintenance of cardiac energy metabolism by histone deacetylase 3 in mice. J. Clin. Investig. 2008, 118, 3588–3597. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Greulich, F.; Rudat, C.; Kispert, A. Mechanisms of T-box gene function in the developing heart. Cardiovasc Res. 2011, 91, 212–222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Lewandowski, S.L.; Janardhan, H.P.; Smee, K.M.; Bachman, M.; Sun, Z.; Lazar, M.A.; Trivedi, C.M. Histone deacetylase 3 modulates Tbx5 activity to regulate early cardiogenesis. Hum. Mol. Genet. 2014, 23, 3801–3809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Backs, J.; Worst, B.C.; Lehmann, L.H.; Patrick, D.M.; Jebessa, Z.; Kreusser, M.M.; Sun, Q.; Chen, L.; Heft, C.; Katus, H.A.; et al. Selective repression of MEF2 activity by PKA-dependent proteolysis of HDAC4. J. Cell Biol. 2011, 195, 403–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Hohl, M.; Wagner, M.; Reil, J.C.; Muller, S.A.; Tauchnitz, M.; Zimmer, A.M.; Lehmann, L.H.; Thiel, G.; Bohm, M.; Backs, J.; et al. HDAC4 controls histone methylation in response to elevated cardiac load. J. Clin. Investig. 2013, 123, 1359–1370. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Zhang, C.L.; McKinsey, T.A.; Chang, S.; Antos, C.L.; Hill, J.A.; Olson, E.N. Class II histone deacetylases act as signal-responsive repressors of cardiac hypertrophy. Cell 2002, 110, 479–488. [Google Scholar] [CrossRef] [Green Version]
  91. Chang, S.; McKinsey, T.A.; Zhang, C.L.; Richardson, J.A.; Hill, J.A.; Olson, E.N. Histone deacetylases 5 and 9 govern responsiveness of the heart to a subset of stress signals and play redundant roles in heart development. Mol. Cell. Biol. 2004, 24, 8467–8476. [Google Scholar] [CrossRef] [Green Version]
  92. Weber, D.; Heisig, J.; Kneitz, S.; Wolf, E.; Eilers, M.; Gessler, M. Mechanisms of epigenetic and cell-type specific regulation of Hey target genes in ES cells and cardiomyocytes. J. Mol. Cell. Cardiol. 2015, 79, 79–88. [Google Scholar] [CrossRef]
  93. Li, X.; Kazgan, N. Mammalian sirtuins and energy metabolism. Int. J. Biol. Sci 2011, 7, 575–587. Available online: https://www.ijbs.com/v07p0575.htm (accessed on 28 September 2022). [CrossRef] [PubMed] [Green Version]
  94. Herskovits, A.Z.; Guarente, L. Sirtuin deacetylases in neurodegenerative diseases of aging. Cell. Res. 2013, 23, 746–758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Liu, B.; Wang, L.; Jiang, W.; Xiong, Y.; Pang, L.; Zhong, Y.; Zhang, C.; Ou, W.; Tian, C.; Chen, X.; et al. Myocyte enhancer factor 2A delays vascular endothelial cell senescence by activating the PI3K/p-Akt/SIRT1 pathway. Aging (Albany NY) 2019, 11, 3768–3784. [Google Scholar] [CrossRef]
  96. Tsai, K.L.; Hung, C.H.; Chan, S.H.; Hsieh, P.L.; Ou, H.C.; Cheng, Y.H.; Chu, P.M. Chlorogenic Acid Protects Against oxLDL-Induced Oxidative Damage and Mitochondrial Dysfunction by Modulating SIRT1 in Endothelial Cells. Mol. Nutr. Food. Res. 2018, 62, e1700928. [Google Scholar] [CrossRef] [PubMed]
  97. Chen, T.S.; Chuang, S.Y.; Shen, C.Y.; Ho, T.J.; Chang, R.L.; Yeh, Y.L.; Kuo, C.H.; Mahalakshmi, B.; Kuo, W.W.; Huang, C.Y. Antioxidant Sirt1/Akt axis expression in resveratrol pretreated adipose-derived stem cells increases regenerative capability in a rat model with cardiomyopathy induced by diabetes mellitus. J. Cell. Physiol. 2021, 236, 4290–4302. [Google Scholar] [CrossRef]
  98. Choudhery, M.S.; Khan, M.; Mahmood, R.; Mohsin, S.; Akhtar, S.; Ali, F.; Khan, S.N.; Riazuddin, S. Mesenchymal stem cells conditioned with glucose depletion augments their ability to repair-infarcted myocardium. J. Cell. Mol. Med. 2012, 16, 2518–2529. [Google Scholar] [CrossRef]
  99. Yeh, E.T.; Bickford, C.L. Cardiovascular complications of cancer therapy: Incidence, pathogenesis, diagnosis, and management. J. Am. Coll. Cardiol. 2009, 53, 2231–2247. [Google Scholar] [CrossRef] [Green Version]
  100. Hequet, O.; Le, Q.H.; Moullet, I.; Pauli, E.; Salles, G.; Espinouse, D.; Dumontet, C.; Thieblemont, C.; Arnaud, P.; Antal, D.; et al. Subclinical late cardiomyopathy after doxorubicin therapy for lymphoma in adults. J. Clin. Oncol. 2004, 22, 1864–1871. [Google Scholar] [CrossRef]
  101. Zhuang, L.; Xia, W.; Chen, D.; Ye, Y.; Hu, T.; Li, S.; Hou, M. Exosomal LncRNA-NEAT1 derived from MIF-treated mesenchymal stem cells protected against doxorubicin-induced cardiac senescence through sponging miR-221-3p. J. Nanobiotechnology. 2020, 18, 157. [Google Scholar] [CrossRef]
  102. Mao, Q.; Liang, X.L.; Zhang, C.L.; Pang, Y.H.; Lu, Y.X. LncRNA KLF3-AS1 in human mesenchymal stem cell-derived exosomes ameliorates pyroptosis of cardiomyocytes and myocardial infarction through miR-138-5p/Sirt1 axis. Stem. Cell. Res. Ther. 2019, 10, 393. [Google Scholar] [CrossRef] [Green Version]
  103. Denu, R.A. SIRT3 Enhances Mesenchymal Stem Cell Longevity and Differentiation. Oxid. Med. Cell. Longev. 2017, 2017, 5841716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Bellizzi, D.; Rose, G.; Cavalcante, P.; Covello, G.; Dato, S.; De Rango, F.; Greco, V.; Maggiolini, M.; Feraco, E.; Mari, V.; et al. A novel VNTR enhancer within the SIRT3 gene, a human homologue of SIR2, is associated with survival at oldest ages. Genomics 2005, 85, 258–263. [Google Scholar] [CrossRef] [PubMed]
  105. Zhang, D.Y.; Gao, T.; Xu, R.J.; Sun, L.; Zhang, C.F.; Bai, L.; Chen, W.; Liu, K.Y.; Zhou, Y.; Jiao, X.; et al. SIRT3 Transfection of Aged Human Bone Marrow-Derived Mesenchymal Stem Cells Improves Cell Therapy-Mediated Myocardial Repair. Rejuvenation Res. 2020, 23, 453–464. [Google Scholar] [CrossRef] [PubMed]
  106. Sarikhani, M.; Garbern, J.C.; Ma, S.; Sereda, R.; Conde, J.; Krähenbühl, G.; Escalante, G.O.; Ahmed, A.; Buenrostro, J.D.; Lee, R.T. Sustained Activation of AMPK Enhances Differentiation of Human iPSC-Derived Cardiomyocytes via Sirtuin Activation. Stem. Cell Rep. 2020, 15, 498–514. [Google Scholar] [CrossRef] [PubMed]
  107. Xie, W.; Song, C.; Young, N.L.; Sperling, A.S.; Xu, F.; Sridharan, R.; Conway, A.E.; Garcia, B.A.; Plath, K.; Clark, A.T.; et al. Histone h3 lysine 56 acetylation is linked to the core transcriptional network in human embryonic stem cells. Mol. Cell 2009, 33, 417–427. [Google Scholar] [CrossRef] [Green Version]
  108. Yang, B.; Zwaans, B.M.; Eckersdorff, M.; Lombard, D.B. The sirtuin SIRT6 deacetylates H3 K56Ac in vivo to promote genomic stability. Cell Cycle 2009, 8, 2662–2663. [Google Scholar] [CrossRef] [Green Version]
  109. Etchegaray, J.P.; Chavez, L.; Huang, Y.; Ross, K.N.; Choi, J.; Martinez-Pastor, B.; Walsh, R.M.; Sommer, C.A.; Lienhard, M.; Gladden, A.; et al. The histone deacetylase SIRT6 controls embryonic stem cell fate via TET-mediated production of 5-hydroxymethylcytosine. Nat. Cell Biol. 2015, 17, 545–557. [Google Scholar] [CrossRef] [Green Version]
  110. Ghosh, L.D.; Ravi, V.; Jain, A.; Panicker, A.G.; Sundaresan, N.R.; Chatterjee, K. Sirtuin 6 mediated stem cell cardiomyogenesis on protein coated nanofibrous scaffolds. Nanomedicine 2019, 19, 145–155. [Google Scholar] [CrossRef] [Green Version]
  111. Murray, K. The Occurrence of Epsilon-N-Methyl Lysine in Histones. Biochemistry 1964, 3, 10–15. [Google Scholar] [CrossRef]
  112. Turner, B.M. Reading signals on the nucleosome with a new nomenclature for modified histones. Nat. Struct. Mol. Biol. 2005, 12, 110–112. [Google Scholar] [CrossRef]
  113. Waterborg, J.H. Dynamic methylation of alfalfa histone H3. J. Biol. Chem. 1993, 268, 4918–4921. [Google Scholar] [CrossRef] [PubMed]
  114. Hon, G.; Wang, W.; Ren, B. Discovery and annotation of functional chromatin signatures in the human genome. PLoS Comput. Biol. 2009, 5, e1000566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Johnson, A.; Wu, R.; Peetz, M.; Gygi, S.P.; Moazed, D. Heterochromatic gene silencing by activator interference and a transcription elongation barrier. J. Biol. Chem. 2013, 288, 28771–28782. [Google Scholar] [CrossRef] [Green Version]
  116. Nimura, K.; Ura, K.; Shiratori, H.; Ikawa, M.; Okabe, M.; Schwartz, R.J.; Kaneda, Y. A histone H3 lysine 36 trimethyltransferase links Nkx2-5 to Wolf-Hirschhorn syndrome. Nature 2009, 460, 287–291. [Google Scholar] [CrossRef]
  117. Gottlieb, P.D.; Pierce, S.A.; Sims, R.J.; Yamagishi, H.; Weihe, E.K.; Harriss, J.V.; Maika, S.D.; Kuziel, W.A.; King, H.L.; Olson, E.N.; et al. Bop encodes a muscle-restricted protein containing MYND and SET domains and is essential for cardiac differentiation and morphogenesis. Nat. Genet. 2002, 31, 25–32. [Google Scholar] [CrossRef] [PubMed]
  118. Barski, A.; Cuddapah, S.; Cui, K.; Roh, T.Y.; Schones, D.E.; Wang, Z.; Wei, G.; Chepelev, I.; Zhao, K. High-resolution profiling of histone methylations in the human genome. Cell 2007, 129, 823–837. [Google Scholar] [CrossRef] [Green Version]
  119. Fujii, T.; Tsunesumi, S.; Yamaguchi, K.; Watanabe, S.; Furukawa, Y. Smyd3 is required for the development of cardiac and skeletal muscle in zebrafish. PLoS ONE 2011, 6, e23491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Movassagh, M.; Choy, M.K.; Knowles, D.A.; Cordeddu, L.; Haider, S.; Down, T.; Siggens, L.; Vujic, A.; Simeoni, I.; Penkett, C.; et al. Distinct epigenomic features in end-stage failing human hearts. Circulation 2011, 124, 2411–2422. [Google Scholar] [CrossRef] [Green Version]
  121. Tao, Y.; Neppl, R.L.; Huang, Z.P.; Chen, J.; Tang, R.H.; Cao, R.; Zhang, Y.; Jin, S.W.; Wang, D.Z. The histone methyltransferase Set7/9 promotes myoblast differentiation and myofibril assembly. J. Cell Biol. 2011, 194, 551–565. [Google Scholar] [CrossRef] [Green Version]
  122. Lee, S.; Lee, J.W.; Lee, S.K. UTX, a histone H3-lysine 27 demethylase, acts as a critical switch to activate the cardiac developmental program. Dev. Cell 2012, 22, 25–37. [Google Scholar] [CrossRef] [Green Version]
  123. Zaidi, S.; Choi, M.; Wakimoto, H.; Ma, L.; Jiang, J.; Overton, J.D.; Romano-Adesman, A.; Bjornson, R.D.; Breitbart, R.E.; Brown, K.K.; et al. De novo mutations in histone-modifying genes in congenital heart disease. Nature 2013, 498, 220–223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Stein, A.B.; Jones, T.A.; Herron, T.J.; Patel, S.R.; Day, S.M.; Noujaim, S.F.; Milstein, M.L.; Klos, M.; Furspan, P.B.; Jalife, J.; et al. Loss of H3K4 methylation destabilizes gene expression patterns and physiological functions in adult murine cardiomyocytes. J. Clin. Investig. 2011, 121, 2641–2650. [Google Scholar] [CrossRef] [PubMed]
  125. Stein, A.B.; Goonewardena, S.N.; Jones, T.A.; Prusick, P.J.; Bazzi, A.A.; Belyavskaya, J.M.; McCoskey, M.M.; Dandar, R.A. The PTIP-Associated Histone Methyltransferase Complex Prevents Stress-Induced Maladaptive Cardiac Remodeling. PLoS ONE 2015, 10, e0127839. [Google Scholar] [CrossRef] [PubMed]
  126. Sims, R.J., III; Weihe, E.K.; Zhu, L.; O’Malley, S.; Harriss, J.V.; Gottlieb, P.D. m-Bop, a repressor protein essential for cardiogenesis, interacts with skNAC, a heart- and muscle-specific transcription factor. J. Biol. Chem. 2002, 277, 26524–26529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Tan, X.; Rotllant, J.; Li, H.; De Deyne, P.; Du, S.J. SmyD1, a histone methyltransferase, is required for myofibril organization and muscle contraction in zebrafish embryos. Proc. Natl. Acad. Sci. USA 2006, 103, 2713–2718. [Google Scholar] [CrossRef] [Green Version]
  128. Sirinupong, N.; Brunzelle, J.; Ye, J.; Pirzada, A.; Nico, L.; Yang, Z. Crystal structure of cardiac-specific histone methyltransferase SmyD1 reveals unusual active site architecture. J. Biol. Chem. 2010, 285, 40635–40644. [Google Scholar] [CrossRef] [Green Version]
  129. Park, C.Y.; Pierce, S.A.; von Drehle, M.; Ivey, K.N.; Morgan, J.A.; Blau, H.M.; Srivastava, D. skNAC, a Smyd1-interacting transcription factor, is involved in cardiac development and skeletal muscle growth and regeneration. Proc. Natl. Acad. Sci. USA 2010, 107, 20750–20755. [Google Scholar] [CrossRef] [Green Version]
  130. Just, S.; Meder, B.; Berger, I.M.; Etard, C.; Trano, N.; Patzel, E.; Hassel, D.; Marquart, S.; Dahme, T.; Vogel, B. The myosin-interacting protein SMYD1 is essential for sarcomere organization. J. Cell Sci. 2011, 124, 3127–3136. [Google Scholar] [CrossRef] [Green Version]
  131. Li, H.; Zhong, Y.; Wang, Z.; Gao, J.; Xu, J.; Chu, W.; Zhang, J.; Fang, S.; Du, S.J. Smyd1b is required for skeletal and cardiac muscle function in zebrafish. Mol. Biol. Cell. 2013, 24, 3511–3521. [Google Scholar] [CrossRef] [PubMed]
  132. Costantini, D.L.; Arruda, E.P.; Agarwal, P.; Kim, K.H.; Zhu, Y.; Zhu, W.; Lebel, M.; Cheng, C.W.; Park, C.Y.; Pierce, S.A.; et al. The homeodomain transcription factor Irx5 establishes the mouse cardiac ventricular repolarization gradient. Cell 2005, 123, 347–358. [Google Scholar] [CrossRef] [Green Version]
  133. Diehl, F.; Brown, M.A.; van Amerongen, M.J.; Novoyatleva, T.; Wietelmann, A.; Harriss, J.; Ferrazzi, F.; Bottger, T.; Harvey, R.P.; Tucker, P.W.; et al. Cardiac deletion of Smyd2 is dispensable for mouse heart development. PLoS ONE 2010, 5, e9748. [Google Scholar] [CrossRef] [Green Version]
  134. Bergemann, A.D.; Cole, F.; Hirschhorn, K. The etiology of Wolf-Hirschhorn syndrome. Trends Genet. 2005, 21, 188–195. [Google Scholar] [CrossRef]
  135. Hojfeldt, J.W.; Laugesen, A.; Willumsen, B.M.; Damhofer, H.; Hedehus, L.; Tvardovskiy, A.; Mohammad, F.; Jensen, O.N.; Helin, K. Accurate H3K27 methylation can be established de novo by SUZ12-directed PRC2. Nat. Struct. Mol. Biol. 2018, 25, 225–232. [Google Scholar] [CrossRef] [Green Version]
  136. Margueron, R.; Reinberg, D. The Polycomb complex PRC2 and its mark in life. Nature 2011, 469, 343–349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Shen, X.; Liu, Y.; Hsu, Y.J.; Fujiwara, Y.; Kim, J.; Mao, X.; Yuan, G.C.; Orkin, S.H. EZH1 mediates methylation on histone H3 lysine 27 and complements EZH2 in maintaining stem cell identity and executing pluripotency. Mol. Cell 2008, 32, 491–502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Pasini, D.; Bracken, A.P.; Hansen, J.B.; Capillo, M.; Helin, K. The polycomb group protein Suz12 is required for embryonic stem cell differentiation. Mol. Cell. Biol. 2007, 27, 3769–3779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Montgomery, N.D.; Yee, D.; Chen, A.; Kalantry, S.; Chamberlain, S.J.; Otte, A.P.; Magnuson, T. The murine polycomb group protein Eed is required for global histone H3 lysine-27 methylation. Curr. Biol. 2005, 15, 942–947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Delgado-Olguin, P.; Huang, Y.; Li, X.; Christodoulou, D.; Seidman, C.E.; Seidman, J.G.; Tarakhovsky, A.; Bruneau, B.G. Epigenetic repression of cardiac progenitor gene expression by Ezh2 is required for postnatal cardiac homeostasis. Nat. Genet. 2012, 44, 343–347. [Google Scholar] [CrossRef] [Green Version]
  141. He, A.; Ma, Q.; Cao, J.; von Gise, A.; Zhou, P.; Xie, H.; Zhang, B.; Hsing, M.; Christodoulou, D.C.; Cahan, P.; et al. Polycomb repressive complex 2 regulates normal development of the mouse heart. Circ. Res. 2012, 110, 406–415. [Google Scholar] [CrossRef] [Green Version]
  142. Lachner, M.; O’Sullivan, R.J.; Jenuwein, T. An epigenetic road map for histone lysine methylation. J. Cell Sci. 2003, 116, 2117–2124. [Google Scholar] [CrossRef] [Green Version]
  143. Santos-Rosa, H.; Schneider, R.; Bannister, A.J.; Sherriff, J.; Bernstein, B.E.; Emre, N.C.; Schreiber, S.L.; Mellor, J.; Kouzarides, T. Active genes are tri-methylated at K4 of histone H3. Nature 2002, 419, 407–411. [Google Scholar] [CrossRef] [PubMed]
  144. Sdek, P.; Zhao, P.; Wang, Y.; Huang, C.J.; Ko, C.Y.; Butler, P.C.; Weiss, J.N.; Maclellan, W.R. Rb and p130 control cell cycle gene silencing to maintain the postmitotic phenotype in cardiac myocytes. J. Cell Biol. 2011, 194, 407–423. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. He, A.; Shen, X.; Ma, Q.; Cao, J.; von Gise, A.; Zhou, P.; Wang, G.; Marquez, V.E.; Orkin, S.H.; Pu, W.T. PRC2 directly methylates GATA4 and represses its transcriptional activity. Genes Dev. 2012, 26, 37–42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Inagawa, M.; Nakajima, K.; Makino, T.; Ogawa, S.; Kojima, M.; Ito, S.; Ikenishi, A.; Hayashi, T.; Schwartz, R.J.; Nakamura, K.; et al. Histone H3 lysine 9 methyltransferases, G9a and GLP are essential for cardiac morphogenesis. Mech. Dev. 2013, 130, 519–531. [Google Scholar] [CrossRef]
  147. Mezentseva, N.V.; Yang, J.; Kaur, K.; Iaffaldano, G.; Remond, M.C.; Eisenberg, C.A.; Eisenberg, L.M. The histone methyltransferase inhibitor BIX01294 enhances the cardiac potential of bone marrow cells. Stem Cells Dev. 2013, 22, 654–667. [Google Scholar] [CrossRef] [Green Version]
  148. Savickiene, J.; Treigyte, G.; Jazdauskaite, A.; Borutinskaite, V.V.; Navakauskiene, R. DNA methyltransferase inhibitor RG108 and histone deacetylase inhibitors cooperate to enhance NB4 cell differentiation and E-cadherin re-expression by chromatin remodelling. Cell Biol. Int. 2012, 36, 1067–1078. [Google Scholar] [CrossRef]
  149. Qian, Q.; Qian, H.; Zhang, X.; Zhu, W.; Yan, Y.; Ye, S.; Peng, X.; Li, W.; Xu, Z.; Sun, L.; et al. 5-Azacytidine induces cardiac differentiation of human umbilical cord-derived mesenchymal stem cells by activating extracellular regulated kinase. Stem Cells Dev. 2012, 21, 67–75. [Google Scholar] [CrossRef] [Green Version]
  150. Minami, I.; Yamada, K.; Otsuji, T.G.; Yamamoto, T.; Shen, Y.; Otsuka, S.; Kadota, S.; Morone, N.; Barve, M.; Asai, Y.; et al. A small molecule that promotes cardiac differentiation of human pluripotent stem cells under defined, cytokine- and xeno-free conditions. Cell Rep. 2012, 2, 1448–1460. [Google Scholar] [CrossRef] [Green Version]
  151. Lee, Y.; Song, A.J.; Baker, R.; Micales, B.; Conway, S.J.; Lyons, G.E. Jumonji, a nuclear protein that is necessary for normal heart development. Circ. Res. 2000, 86, 932–938. [Google Scholar] [CrossRef] [Green Version]
  152. Takeuchi, T.; Watanabe, Y.; Takano-Shimizu, T.; Kondo, S. Roles of jumonji and jumonji family genes in chromatin regulation and development. Dev. Dyn. 2006, 235, 2449–2459. [Google Scholar] [CrossRef]
  153. Klose, R.J.; Kallin, E.M.; Zhang, Y. JmjC-domain-containing proteins and histone demethylation. Nat. Rev. Genet. 2006, 7, 715–727. [Google Scholar] [CrossRef] [PubMed]
  154. Chang, B.; Chen, Y.; Zhao, Y.; Bruick, R.K. JMJD6 is a histone arginine demethylase. Science 2007, 318, 444–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Whetstine, J.R.; Nottke, A.; Lan, F.; Huarte, M.; Smolikov, S.; Chen, Z.; Spooner, E.; Li, E.; Zhang, G.; Colaiacovo, M.; et al. Reversal of histone lysine trimethylation by the JMJD2 family of histone demethylases. Cell 2006, 125, 467–481. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Zhang, Q.J.; Chen, H.Z.; Wang, L.; Liu, D.P.; Hill, J.A.; Liu, Z.P. The histone trimethyllysine demethylase JMJD2A promotes cardiac hypertrophy in response to hypertrophic stimuli in mice. J. Clin. Investig. 2011, 121, 2447–2456. [Google Scholar] [CrossRef] [Green Version]
  157. Agger, K.; Cloos, P.A.; Christensen, J.; Pasini, D.; Rose, S.; Rappsilber, J.; Issaeva, I.; Canaani, E.; Salcini, A.E.; Helin, K. UTX and JMJD3 are histone H3K27 demethylases involved in HOX gene regulation and development. Nature 2007, 449, 731–734. [Google Scholar] [CrossRef]
  158. Lan, F.; Bayliss, P.E.; Rinn, J.L.; Whetstine, J.R.; Wang, J.K.; Chen, S.; Iwase, S.; Alpatov, R.; Issaeva, I.; Canaani, E.; et al. A histone H3 lysine 27 demethylase regulates animal posterior development. Nature 2007, 449, 689–694. [Google Scholar] [CrossRef] [PubMed]
  159. Hong, S.; Cho, Y.W.; Yu, L.R.; Yu, H.; Veenstra, T.D.; Ge, K. Identification of JmjC domain-containing UTX and JMJD3 as histone H3 lysine 27 demethylases. Proc. Natl. Acad. Sci. USA 2007, 104, 18439–18444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Jackson, M.; Krassowska, A.; Gilbert, N.; Chevassut, T.; Forrester, L.; Ansell, J.; Ramsahoye, B. Severe global DNA hypomethylation blocks differentiation and induces histone hyperacetylation in embryonic stem cells. Mol. Cell. Biol. 2004, 24, 8862–8871. [Google Scholar] [CrossRef] [Green Version]
  161. Klose, R.J.; Bird, A.P. Genomic DNA methylation: The mark and its mediators. Trends Biochem. Sci. 2006, 31, 89–97. [Google Scholar] [CrossRef]
  162. Trojer, P.; Reinberg, D. Histone lysine demethylases and their impact on epigenetics. Cell 2006, 125, 213–217. [Google Scholar] [CrossRef] [Green Version]
  163. Dudley, D.D.; Chaudhuri, J.; Bassing, C.H.; Alt, F.W. Mechanism and control of V(D)J recombination versus class switch recombination: Similarities and differences. Adv. Immunol. 2005, 86, 43–112. [Google Scholar] [CrossRef] [PubMed]
  164. Iyer, L.M.; Tahiliani, M.; Rao, A.; Aravind, L. Prediction of novel families of enzymes involved in oxidative and other complex modifications of bases in nucleic acids. Cell Cycle 2009, 8, 1698–1710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Tahiliani, M.; Koh, K.P.; Shen, Y.; Pastor, W.A.; Bandukwala, H.; Brudno, Y.; Agarwal, S.; Iyer, L.M.; Liu, D.R.; Aravind, L.; et al. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 2009, 324, 930–935. [Google Scholar] [CrossRef] [Green Version]
  166. He, Y.F.; Li, B.Z.; Li, Z.; Liu, P.; Wang, Y.; Tang, Q.; Ding, J.; Jia, Y.; Chen, Z.; Li, L.; et al. Tet-mediated formation of 5-carboxylcytosine and its excision by TDG in mammalian DNA. Science 2011, 333, 1303–1307. [Google Scholar] [CrossRef] [Green Version]
  167. Ito, S.; Shen, L.; Dai, Q.; Wu, S.C.; Collins, L.B.; Swenberg, J.A.; He, C.; Zhang, Y. Tet proteins can convert 5-methylcytosine to 5-formylcytosine and 5-carboxylcytosine. Science 2011, 333, 1300–1303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Koh, K.P.; Yabuuchi, A.; Rao, S.; Huang, Y.; Cunniff, K.; Nardone, J.; Laiho, A.; Tahiliani, M.; Sommer, C.A.; Mostoslavsky, G.; et al. Tet1 and Tet2 regulate 5-hydroxymethylcytosine production and cell lineage specification in mouse embryonic stem cells. Cell Stem Cell 2011, 8, 200–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Szwagierczak, A.; Bultmann, S.; Schmidt, C.S.; Spada, F.; Leonhardt, H. Sensitive enzymatic quantification of 5-hydroxymethylcytosine in genomic DNA. Nucleic Acids Res. 2010, 38, e181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Du, H.; Zhao, Y.; He, J.; Zhang, Y.; Xi, H.; Liu, M.; Ma, J.; Wu, L. YTHDF2 destabilizes m(6)A-containing RNA through direct recruitment of the CCR4-NOT deadenylase complex. Nat. Commun. 2016, 7, 12626. [Google Scholar] [CrossRef]
  171. Liu, J.; Yue, Y.; Han, D.; Wang, X.; Fu, Y.; Zhang, L.; Jia, G.; Yu, M.; Lu, Z.; Deng, X.; et al. A METTL3-METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation. Nat. Chem. Biol. 2014, 10, 93–95. [Google Scholar] [CrossRef] [Green Version]
  172. Ping, X.L.; Sun, B.F.; Wang, L.; Xiao, W.; Yang, X.; Wang, W.J.; Adhikari, S.; Shi, Y.; Lv, Y.; Chen, Y.S.; et al. Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Res. 2014, 24, 177–189. [Google Scholar] [CrossRef] [Green Version]
  173. Li, Z.; Weng, H.; Su, R.; Weng, X.; Zuo, Z.; Li, C.; Huang, H.; Nachtergaele, S.; Dong, L.; Hu, C.; et al. FTO Plays an Oncogenic Role in Acute Myeloid Leukemia as a N(6)-Methyladenosine RNA Demethylase. Cancer Cell 2017, 31, 127–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Meyer, K.D.; Jaffrey, S.R. Rethinking m(6)A Readers, Writers, and Erasers. Annu. Rev. Cell Dev. Biol. 2017, 33, 319–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Zaccara, S.; Ries, R.J.; Jaffrey, S.R. Reading, writing and erasing mRNA methylation. Nat. Rev. Mol. Cell Biol. 2019, 20, 608–624. [Google Scholar] [CrossRef] [PubMed]
  176. Dorn, L.E.; Lasman, L.; Chen, J.; Xu, X.; Hund, T.J.; Medvedovic, M.; Hanna, J.H.; van Berlo, J.H.; Accornero, F. The N(6)-Methyladenosine mRNA Methylase METTL3 Controls Cardiac Homeostasis and Hypertrophy. Circulation 2019, 139, 533–545. [Google Scholar] [CrossRef]
  177. Li, W.; Xing, C.; Bao, L.; Han, S.; Luo, T.; Wang, Z.; Fan, H. Comprehensive analysis of RNA m6A methylation in pressure overload-induced cardiac hypertrophy. BMC Genomics 2022, 23, 576. [Google Scholar] [CrossRef]
  178. Gong, R.; Wang, X.; Li, H.; Liu, S.; Jiang, Z.; Zhao, Y.; Yu, Y.; Han, Z.; Yu, Y.; Dong, C.; et al. Loss of m(6)A methyltransferase METTL3 promotes heart regeneration and repair after myocardial injury. Pharmacol. Res. 2021, 174, 105845. [Google Scholar] [CrossRef]
  179. Han, Z.; Wang, X.; Xu, Z.; Cao, Y.; Gong, R.; Yu, Y.; Yu, Y.; Guo, X.; Liu, S.; Yu, M.; et al. ALKBH5 regulates cardiomyocyte proliferation and heart regeneration by demethylating the mRNA of YTHDF1. Theranostics 2021, 11, 3000–3016. [Google Scholar] [CrossRef]
  180. Barroso-delJesus, A.; Romero-Lopez, C.; Lucena-Aguilar, G.; Melen, G.J.; Sanchez, L.; Ligero, G.; Berzal-Herranz, A.; Menendez, P. Embryonic stem cell-specific miR302-367 cluster: Human gene structure and functional characterization of its core promoter. Mol. Cell. Biol. 2008, 28, 6609–6619. [Google Scholar] [CrossRef] [Green Version]
  181. Card, D.A.; Hebbar, P.B.; Li, L.; Trotter, K.W.; Komatsu, Y.; Mishina, Y.; Archer, T.K. Oct4/Sox2-regulated miR-302 targets cyclin D1 in human embryonic stem cells. Mol. Cell. Biol. 2008, 28, 6426–6438. [Google Scholar] [CrossRef] [Green Version]
  182. Czaja, W.; Mao, P.; Smerdon, M.J. The emerging roles of ATP-dependent chromatin remodeling enzymes in nucleotide excision repair. Int. J. Mol. Sci. 2012, 13, 11954–11973. [Google Scholar] [CrossRef] [Green Version]
  183. Chang, C.P.; Bruneau, B.G. Epigenetics and cardiovascular development. Annu. Rev. Physiol. 2012, 74, 41–68. [Google Scholar] [CrossRef] [PubMed]
  184. Bevilacqua, A.; Willis, M.S.; Bultman, S.J. SWI/SNF chromatin-remodeling complexes in cardiovascular development and disease. Cardiovasc. Pathol. 2014, 23, 85–91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Hang, C.T.; Yang, J.; Han, P.; Cheng, H.L.; Shang, C.; Ashley, E.; Zhou, B.; Chang, C.P. Chromatin regulation by Brg1 underlies heart muscle development and disease. Nature 2010, 466, 62–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Takeuchi, J.K.; Bruneau, B.G. Directed transdifferentiation of mouse mesoderm to heart tissue by defined factors. Nature 2009, 459, 708–711. [Google Scholar] [CrossRef] [Green Version]
  187. Pisano, A.; Cerbelli, B.; Perli, E.; Pelullo, M.; Bargelli, V.; Preziuso, C.; Mancini, M.; He, L.; Bates, M.G.; Lucena, J.R.; et al. Impaired mitochondrial biogenesis is a common feature to myocardial hypertrophy and end-stage ischemic heart failure. Cardiovasc. Pathol. 2016, 25, 103–112. [Google Scholar] [CrossRef] [Green Version]
  188. Romanello, V.; Sandri, M. Mitochondrial Quality Control and Muscle Mass Maintenance. Front. Physiol. 2015, 6, 422. [Google Scholar] [CrossRef]
  189. Ehrlich, K.C.; Deng, H.W.; Ehrlich, M. Epigenetics of Mitochondria-Associated Genes in Striated Muscle. Epigenomes 2021, 6, 1. [Google Scholar] [CrossRef] [PubMed]
  190. Yang, J.; Xu, W.W.; Hu, S.J. Heart failure: Advanced development in genetics and epigenetics. BioMed Res. Int. 2015, 2015, 352734. [Google Scholar] [CrossRef] [Green Version]
  191. Masliah-Planchon, J.; Bieche, I.; Guinebretiere, J.M.; Bourdeaut, F.; Delattre, O. SWI/SNF chromatin remodeling and human malignancies. Annu. Rev. Pathol. 2015, 10, 145–171. [Google Scholar] [CrossRef]
  192. Shi, J.; Whyte, W.A.; Zepeda-Mendoza, C.J.; Milazzo, J.P.; Shen, C.; Roe, J.S.; Minder, J.L.; Mercan, F.; Wang, E.; Eckersley-Maslin, M.A.; et al. Role of SWI/SNF in acute leukemia maintenance and enhancer-mediated Myc regulation. Genes Dev. 2013, 27, 2648–2662. [Google Scholar] [CrossRef] [Green Version]
  193. Bultman, S.J.; Holley, D.W.; de Ridder, G.G.; Pizzo, S.V.; Sidorova, T.N.; Murray, K.T.; Jensen, B.C.; Wang, Z.; Bevilacqua, A.; Chen, X.; et al. BRG1 and BRM SWI/SNF ATPases redundantly maintain cardiomyocyte homeostasis by regulating cardiomyocyte mitophagy and mitochondrial dynamics in vivo. Cardiovasc. Pathol. 2016, 25, 258–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Koczor, C.A.; Jiao, Z.; Fields, E.; Russ, R.; Ludaway, T.; Lewis, W. AZT-induced mitochondrial toxicity: An epigenetic paradigm for dysregulation of gene expression through mitochondrial oxidative stress. Physiol. Genom. 2015, 47, 447–454. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Koczor, C.A.; Ludlow, I.; Fields, E.; Jiao, Z.; Ludaway, T.; Russ, R.; Lewis, W. Mitochondrial polymerase gamma dysfunction and aging cause cardiac nuclear DNA methylation changes. Physiol. Genom. 2016, 48, 274–280. [Google Scholar] [CrossRef] [Green Version]
  196. Matsuhashi, T.; Hishiki, T.; Zhou, H.; Ono, T.; Kaneda, R.; Iso, T.; Yamaguchi, A.; Endo, J.; Katsumata, Y.; Atsushi, A.; et al. Activation of pyruvate dehydrogenase by dichloroacetate has the potential to induce epigenetic remodeling in the heart. J. Mol. Cell Cardiol. 2015, 82, 116–124. [Google Scholar] [CrossRef] [PubMed]
  197. Partanen, A.; Motoyama, J.; Hui, C.C. Developmentally regulated expression of the transcriptional cofactors/histone acetyltransferases CBP and p300 during mouse embryogenesis. Int. J. Dev. Biol. 1999, 43, 487–494. [Google Scholar] [PubMed]
  198. Hou, Y.S.; Wang, J.Z.; Shi, S.; Han, Y.; Zhang, Y.; Zhi, J.X.; Xu, C.; Li, F.F.; Wang, G.Y.; Liu, S.L. Identification of epigenetic factor KAT2B gene variants for possible roles in congenital heart diseases. Biosci. Rep. 2020, 40, 4. [Google Scholar] [CrossRef] [Green Version]
  199. Li, B.; Li, M.; Li, X.; Li, H.; Lai, Y.; Huang, S.; He, X.; Si, X.; Zheng, H.; Liao, W.; et al. Sirt1-inducible deacetylation of p21 promotes cardiomyocyte proliferation. Aging (Albany NY) 2019, 11, 12546–12567. [Google Scholar] [CrossRef]
  200. Kidder, B.L.; Hu, G.; Cui, K.; Zhao, K. SMYD5 regulates H4K20me3-marked heterochromatin to safeguard ES cell self-renewal and prevent spurious differentiation. Epigenetics Chromatin. 2017, 10, 8. [Google Scholar] [CrossRef] [Green Version]
  201. Boyer, L.A.; Plath, K.; Zeitlinger, J.; Brambrink, T.; Medeiros, L.A.; Lee, T.I.; Levine, S.S.; Wernig, M.; Tajonar, A.; Ray, M.K.; et al. Polycomb complexes repress developmental regulators in murine embryonic stem cells. Nature 2006, 441, 349–353. [Google Scholar] [CrossRef]
  202. Pursani, V.; Bhartiya, D.; Tanavde, V.; Bashir, M.; Sampath, P. Transcriptional activator DOT1L putatively regulates human embryonic stem cell differentiation into the cardiac lineage. Stem Cell Res. Ther. 2018, 9, 97. [Google Scholar] [CrossRef] [Green Version]
  203. El-Nachef, D.; Oyama, K.; Wu, Y.Y.; Freeman, M.; Zhang, Y.; MacLellan, W.R. Repressive histone methylation regulates cardiac myocyte cell cycle exit. J. Mol. Cell Cardiol. 2018, 121, 1–12. [Google Scholar] [CrossRef]
  204. Brien, G.L.; Gambero, G.; O’connell, D.J.; Jerman, E.; Turner, S.A.; Egan, C.M.; Dunne, E.J.; Jurgens, M.C.; Wynne, K.; Piao, L.; et al. Polycomb PHF19 binds H3K36me3 and recruits PRC2 and demethylase NO66 to embryonic stem cell genes during differentiation. Nat. Struct. Mol. Biol. 2012, 19, 1273–1281. [Google Scholar] [CrossRef] [PubMed]
  205. Rosales, W.; Lizcano, F. The Histone Demethylase JMJD2A Modulates the Induction of Hypertrophy Markers in iPSC-Derived Cardiomyocytes. Front. Genet. 2018, 9, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Cho, E.; Mysliwiec, M.R.; Carlson, C.D.; Ansari, A.; Schwartz, R.J.; Lee, Y. Cardiac-specific developmental and epigenetic functions of Jarid2 during embryonic development. J. Biol. Chem. 2018, 293, 11659–11673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Wang, Y.; Li, Y.; Guo, C.; Lu, Q.; Wang, W.; Jia, Z.; Chen, P.; Ma, K.; Reinberg, D.; Zhou, C. ISL1 and JMJD3 synergistically control cardiac differentiation of embryonic stem cells. Nucleic Acids Res. 2016, 44, 6741–6755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Guo, Z.; Lu, J.; Li, J.; Wang, P.; Li, Z.; Zhong, Y.; Guo, K.; Wang, J.; Ye, J.; Liu, P. JMJD3 inhibition protects against isoproterenol-induced cardiac hypertrophy by suppressing beta-MHC expression. Mol. Cell Endocrinol. 2018, 477, 1–14. [Google Scholar] [CrossRef]
  209. Tang, Y.; Hong, Y.Z.; Bai, H.J.; Wu, Q.; Chen, C.D.; Lang, J.Y.; Boheler, K.R.; Yang, H.T. Plant Homeo Domain Finger Protein 8 Regulates Mesodermal and Cardiac Differentiation of Embryonic Stem Cells Through Mediating the Histone Demethylation of pmaip1. Stem Cells. 2016, 34, 1527–1540. [Google Scholar] [CrossRef] [Green Version]
  210. Chen, J.Q.; Papp, G.; Szodoray, P.; Zeher, M. The role of microRNAs in the pathogenesis of autoimmune diseases. Autoimmun. Rev. 2016, 15, 1171–1180. [Google Scholar] [CrossRef] [Green Version]
  211. Deb, B.; Uddin, A.; Chakraborty, S. miRNAs and ovarian cancer: An overview. J. Cell Physiol. 2018, 233, 3846–3854. [Google Scholar] [CrossRef]
  212. Li, X.; Wei, Y.; Wang, Z. microRNA-21 and hypertension. Hypertens. Res. 2018, 41, 649–661. [Google Scholar] [CrossRef]
  213. Cirillo, F.; Catellani, C.; Lazzeroni, P.; Sartori, C.; Street, M.E. The Role of MicroRNAs in Influencing Body Growth and Development. Horm. Res. Paediatr. 2020, 93, 7–15. [Google Scholar] [CrossRef] [PubMed]
  214. Kanellopoulou, C.; Muljo, S.A.; Kung, A.L.; Ganesan, S.; Drapkin, R.; Jenuwein, T.; Livingston, D.M.; Rajewsky, K. Dicer-deficient mouse embryonic stem cells are defective in differentiation and centromeric silencing. Genes Dev. 2005, 19, 489–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Murchison, E.P.; Partridge, J.F.; Tam, O.H.; Cheloufi, S.; Hannon, G.J. Characterization of Dicer-deficient murine embryonic stem cells. Proc. Natl. Acad. Sci. USA 2005, 102, 12135–12140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Wang, Y.; Medvid, R.; Melton, C.; Jaenisch, R.; Blelloch, R. DGCR8 is essential for microRNA biogenesis and silencing of embryonic stem cell self-renewal. Nat. Genet. 2007, 39, 380–385. [Google Scholar] [CrossRef] [PubMed]
  217. Mallanna, S.K.; Rizzino, A. Emerging roles of microRNAs in the control of embryonic stem cells and the generation of induced pluripotent stem cells. Dev. Biol. 2010, 344, 16–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Small, E.M.; Olson, E.N. Pervasive roles of microRNAs in cardiovascular biology. Nature 2011, 469, 336–342. [Google Scholar] [CrossRef] [Green Version]
  219. Lagos-Quintana, M.; Rauhut, R.; Yalcin, A.; Meyer, J.; Lendeckel, W.; Tuschl, T. Identification of tissue-specific microRNAs from mouse. Curr. Biol. 2002, 12, 735–739. [Google Scholar] [CrossRef] [Green Version]
  220. Zhao, Y.; Samal, E.; Srivastava, D. Serum response factor regulates a muscle-specific microRNA that targets Hand2 during cardiogenesis. Nature 2005, 436, 214–220. [Google Scholar] [CrossRef]
  221. Sayed, D.; Hong, C.; Chen, I.Y.; Lypowy, J.; Abdellatif, M. MicroRNAs play an essential role in the development of cardiac hypertrophy. Circ. Res. 2007, 100, 416–424. [Google Scholar] [CrossRef] [Green Version]
  222. Porrello, E.R. microRNAs in cardiac development and regeneration. Clin. Sci. 2013, 125, 151–166. [Google Scholar] [CrossRef] [Green Version]
  223. Izarra, A.; Moscoso, I.; Canon, S.; Carreiro, C.; Fondevila, D.; Martin-Caballero, J.; Blanca, V.; Valiente, I.; Diez-Juan, A.; Bernad, A. miRNA-1 and miRNA-133a are involved in early commitment of pluripotent stem cells and demonstrate antagonistic roles in the regulation of cardiac differentiation. J. Tissue Eng. Regen. Med. 2017, 11, 787–799. [Google Scholar] [CrossRef] [PubMed]
  224. Dal-Pra, S.; Hodgkinson, C.P.; Mirotsou, M.; Kirste, I.; Dzau, V.J. Demethylation of H3K27 is essential for the induction of direct cardiac reprogramming by miR combo. Circ. Res. 2017, 120, 1403–1413. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Lopez-Sanchez, C.; Franco, D.; Bonet, F.; Garcia-Lopez, V.; Aranega, A.; Garcia-Martinez, V. Reciprocal repression between Fgf8 and miR-133 regulates cardiac induction through Bmp2 signaling. Data Brief 2015, 5, 59–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Hoelscher, S.C.; Stich, T.; Diehm, A.; Lahm, H.; Dressen, M.; Zhang, Z.; Neb, I.; Aherrahrou, Z.; Erdmann, J.; Schunkert, H.; et al. miR-128a Acts as a Regulator in Cardiac Development by Modulating Differentiation of Cardiac Progenitor Cell Populations. Int. J. Mol. Sci. 2020, 21, 1158. [Google Scholar] [CrossRef] [Green Version]
  227. Wu, Y.H.; Zhao, H.; Zhou, L.P.; Zhao, C.X.; Wu, Y.F.; Zhen, L.X.; Li, J.; Ge, D.X.; Xu, L.; Lin, L.; et al. miR-134 Modulates the Proliferation of Human Cardiomyocyte Progenitor Cells by Targeting Meis2. Int. J. Mol. Sci. 2015, 16, 25199–25213. [Google Scholar] [CrossRef] [Green Version]
  228. Liu, X.; Xiao, J.; Zhu, H.; Wei, X.; Platt, C.; Damilano, F.; Xiao, C.; Bezzerides, V.; Bostrom, P.; Che, L.; et al. miR-222 is necessary for exercise-induced cardiac growth and protects against pathological cardiac remodeling. Cell Metab. 2015, 21, 584–595. [Google Scholar] [CrossRef] [Green Version]
  229. Lee, D.S.; Chen, J.H.; Lundy, D.J.; Liu, C.H.; Hwang, S.M.; Pabon, L.; Shieh, R.C.; Chen, C.C.; Wu, S.N.; Yan, Y.T.; et al. Defined MicroRNAs Induce Aspects of Maturation in Mouse and Human Embryonic-Stem-Cell-Derived Cardiomyocytes. Cell Rep. 2015, 12, 1960–1967. [Google Scholar] [CrossRef] [Green Version]
  230. Borden, A.; Kurian, J.; Nickoloff, E.; Yang, Y.; Troupes, C.D.; Ibetti, J.; Lucchese, A.M.; Gao, E.; Mohsin, S.; Koch, W.J.; et al. Transient Introduction of miR-294 in the Heart Promotes Cardiomyocyte Cell Cycle Reentry After Injury. Circ. Res. 2019, 125, 14–25. [Google Scholar] [CrossRef]
  231. Xiao, J.; Liu, H.; Cretoiu, D.; Toader, D.O.; Suciu, N.; Shi, J.; Shen, S.; Bei, Y.; Sluijter, J.P.; Das, S.; et al. miR-31a-5p promotes postnatal cardiomyocyte proliferation by targeting RhoBTB1. Exp. Mol. Med. 2017, 49, e386. [Google Scholar] [CrossRef] [Green Version]
  232. Poon, E.N.; Hao, B.; Guan, D.; Jun Li, M.; Lu, J.; Yang, Y.; Wu, B.; Wu, S.C.; Webb, S.E.; Liang, Y.; et al. Integrated transcriptomic and regulatory network analyses identify microRNA-200c as a novel repressor of human pluripotent stem cell-derived cardiomyocyte differentiation and maturation. Cardiovasc. Res. 2018, 114, 894–906. [Google Scholar] [CrossRef]
  233. Zhuang, S.; Fu, Y.; Li, J.; Li, M.; Hu, X.; Zhu, J.; Tong, M. MicroRNA-375 overexpression disrupts cardiac development of Zebrafish (Danio rerio) by targeting notch2. Protoplasma 2020, 257, 1309–1318. [Google Scholar] [CrossRef] [PubMed]
  234. Wang, B.; Xu, M.; Li, M.; Wu, F.; Hu, S.; Chen, X.; Zhao, L.; Huang, Z.; Lan, F.; Liu, D.; et al. miR-25 Promotes Cardiomyocyte Proliferation by Targeting FBXW7. Mol. Ther. -Nucleic Acids 2020, 19, 1299–1308. [Google Scholar] [CrossRef]
  235. Yu, R.B.; Li, K.; Wang, G.; Gao, G.M.; Du, J.X. MiR-23 enhances cardiac fibroblast proliferation and suppresses fibroblast apoptosis via targeting TGF-beta1 in atrial fibrillation. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 4419–4424. [Google Scholar] [CrossRef] [PubMed]
  236. Smith, T.; Rajakaruna, C.; Caputo, M.; Emanueli, C. MicroRNAs in congenital heart disease. Ann. Transl. Med. 2015, 3, 333. [Google Scholar] [CrossRef]
  237. Zhao, Y.; Ransom, J.F.; Li, A.; Vedantham, V.; von Drehle, M.; Muth, A.N.; Tsuchihashi, T.; McManus, M.T.; Schwartz, R.J.; Srivastava, D. Dysregulation of cardiogenesis, cardiac conduction, and cell cycle in mice lacking miRNA-1-2. Cell 2007, 129, 303–317. [Google Scholar] [CrossRef] [Green Version]
  238. Wu, F.; Yang, Q.; Mi, Y.; Wang, F.; Cai, K.; Zhang, Y.; Wang, Y.; Wang, X.; Gui, Y.; Li, Q. miR-29b-3p Inhibitor Alleviates Hypomethylation-Related Aberrations Through a Feedback Loop Between miR-29b-3p and DNA Methylation in Cardiomyocytes. Front. Cell Dev. Biol. 2022, 10, 788799. [Google Scholar] [CrossRef] [PubMed]
  239. Ramachandran, V.; Bhagavatheeswaran, S.; Shanmugam, S.; Vasudevan, M.; Ragunathan, M.; Cherian, K.M.; Munirajan, A.K.; Ravi, S.; Balakrishnan, A. Deep sequencing unveils altered cardiac miRNome in congenital heart disease. Mol. Genet. Genom. 2022, 297, 1123–1139. [Google Scholar] [CrossRef]
  240. Kan, Z.; Yan, W.; Wang, N.; Fang, Y.; Gao, H.; Song, Y. Identification of circRNA-miRNA-mRNA Regulatory Network and Crucial Signaling Pathway Axis Involved in Tetralogy of Fallot. Front. Genet. 2022, 13, 917454. [Google Scholar] [CrossRef]
  241. Yang, P.; Yang, Y.; He, X.; Sun, P.; Zhang, Y.; Song, X.; Tian, Y.; Zong, T.; Ma, J.; Chen, X.; et al. miR-153-3p Targets betaII Spectrin to Regulate Formaldehyde-Induced Cardiomyocyte Apoptosis. Front. Cardiovasc. Med. 2021, 8, 764831. [Google Scholar] [CrossRef]
  242. Meng, S.; Hu, Y.; Zhu, J.; Feng, T.; Quan, X. miR-30c-5p acts as a therapeutic target for ameliorating myocardial ischemia-reperfusion injury. Am. J. Transl. Res. 2021, 13, 2198–2212. [Google Scholar]
  243. Li, Y.; Li, X.; Wang, L.; Han, N.; Yin, G. miR-375-3p contributes to hypoxia-induced apoptosis by targeting forkhead box P1 (FOXP1) and Bcl2 like protein 2 (Bcl2l2) in rat cardiomyocyte h9c2 cells. Biotechnol. Lett. 2021, 43, 353–367. [Google Scholar] [CrossRef] [PubMed]
  244. Ceci, M.; Carlantoni, C.; Missinato, M.A.; Bonvissuto, D.; Di Giacomo, B.; Contu, R.; Romano, N. Micro RNAs are involved in activation of epicardium during zebrafish heart regeneration. Cell Death Discov. 2018, 4, 41. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Gan, J.; Tang, F.M.K.; Su, X.; Lu, G.; Xu, J.; Lee, H.S.S.; Lee, K.K.H. microRNA-1 inhibits cardiomyocyte proliferation in mouse neonatal hearts by repressing CCND1 expression. Ann. Transl. Med. 2019, 7, 455. [Google Scholar] [CrossRef] [PubMed]
  246. Xie, Y.; Wang, Q.; Gao, N.; Wu, F.; Lan, F.; Zhang, F.; Jin, L.; Huang, Z.; Ge, J.; Wang, H.; et al. MircroRNA-10b Promotes Human Embryonic Stem Cell-Derived Cardiomyocyte Proliferation via Novel Target Gene LATS1. Mol. Ther. -Nucleic Acids 2020, 19, 437–445. [Google Scholar] [CrossRef] [PubMed]
  247. Morrison, J.L.; Zhang, S.; Tellam, R.L.; Brooks, D.A.; McMillen, I.C.; Porrello, E.R.; Botting, K.J. Regulation of microRNA during cardiomyocyte maturation in sheep. BMC Genom. 2015, 16, 541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Zhang, Z.; Ursin, R.; Mahapatra, S.; Gallicano, G.I. CRISPR/CAS9 ablation of individual miRNAs from a miRNA family reveals their individual efficacies for regulating cardiac differentiation. Mech. Dev. 2018, 150, 10–20. [Google Scholar] [CrossRef]
  249. Guzzolino, E.; Chiavacci, E.; Ahuja, N.; Mariani, L.; Evangelista, M.; Ippolito, C.; Rizzo, M.; Garrity, D.; Cremisi, F.; Pitto, L. Post-transcriptional Modulation of Sphingosine-1-Phosphate Receptor 1 by miR-19a Affects Cardiovascular Development in Zebrafish. Front. Cell Dev. Biol. 2018, 6, 58. [Google Scholar] [CrossRef] [Green Version]
  250. Gao, F.; Kataoka, M.; Liu, N.; Liang, T.; Huang, Z.P.; Gu, F.; Ding, J.; Liu, J.; Zhang, F.; Ma, Q.; et al. Therapeutic role of miR-19a/19b in cardiac regeneration and protection from myocardial infarction. Nat. Commun. 2019, 10, 1802. [Google Scholar] [CrossRef] [Green Version]
  251. Ai, F.; Zhang, Y.; Peng, B. miR-20a regulates proliferation, differentiation and apoptosis in P19 cell model of cardiac differentiation by targeting Smoothened. Biol. Open 2016, 5, 1260–1265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Qin, X.; Gao, S.; Yang, Y.; Wu, L.; Wang, L. microRNA-25 promotes cardiomyocytes proliferation and migration via targeting Bim. J. Cell. Physiol. 2019, 234, 22103–22115. [Google Scholar] [CrossRef]
  253. Wang, D.; Liu, C.; Wang, Y.; Wang, W.; Wang, K.; Wu, X.; Li, Z.; Zhao, C.; Li, L.; Peng, L. Impact of miR-26b on cardiomyocyte differentiation in P19 cells through regulating canonical/non-canonical Wnt signalling. Cell Prolif. 2017, 50, e12371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Shen, Y.; Lu, H.; Chen, R.; Zhu, L.; Song, G. MicroRNA-29c affects zebrafish cardiac development via targeting Wnt4. Mol. Med. Rep. 2020, 22, 4675–4684. [Google Scholar] [CrossRef] [PubMed]
  255. Yang, Q.; Wu, F.; Mi, Y.; Wang, F.; Cai, K.; Yang, X.; Zhang, R.; Liu, L.; Zhang, Y.; Wang, Y.; et al. Aberrant expression of miR-29b-3p influences heart development and cardiomyocyte proliferation by targeting NOTCH2. Cell Prolif. 2020, 53, e12764. [Google Scholar] [CrossRef] [Green Version]
  256. Liu, X.; Li, M.; Peng, Y.; Hu, X.; Xu, J.; Zhu, S.; Yu, Z.; Han, S. miR-30c regulates proliferation, apoptosis and differentiation via the Shh signaling pathway in P19 cells. Exp. Mol. Med. 2016, 48, e248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  257. Zhang, C.; Zhang, Y.; Zhu, H.; Hu, J.; Xie, Z. MiR-34a/miR-93 target c-Ski to modulate the proliferaton of rat cardiac fibroblasts and extracellular matrix deposition in vivo and in vitro. Cell. Signal. 2018, 46, 145–153. [Google Scholar] [CrossRef]
  258. Wu, K.H.; Xiao, Q.R.; Yang, Y.; Xu, J.L.; Zhang, F.; Liu, C.M.; Zhang, Z.M.; Lu, Y.Q.; Huang, N.P. MicroRNA-34a modulates the Notch signaling pathway in mice with congenital heart disease and its role in heart development. J. Mol. Cell. Cardiol. 2018, 114, 300–308. [Google Scholar] [CrossRef]
  259. Beauchemin, M.; Smith, A.; Yin, V.P. Dynamic microRNA-101a and Fosab expression controls zebrafish heart regeneration. Development 2015, 142, 4026–4037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Yuan, R.; Zhang, X.; Fang, Y.; Nie, Y.; Cai, S.; Chen, Y.; Mo, D. mir-127-3p inhibits the proliferation of myocytes by targeting KMT5a. Biochem. Biophys. Res. Commun. 2018, 503, 970–976. [Google Scholar] [CrossRef]
  261. Huang, W.; Feng, Y.; Liang, J.; Yu, H.; Wang, C.; Wang, B.; Wang, M.; Jiang, L.; Meng, W.; Cai, W.; et al. Loss of microRNA-128 promotes cardiomyocyte proliferation and heart regeneration. Nat. Commun. 2018, 9, 700. [Google Scholar] [CrossRef] [Green Version]
  262. Huang, Y.; Jiang, B.; Xia, Y.; Wang, J.; Ji, C.; Tong, J.; Chen, T.; Jiang, Y. Downregulation of miR-133a contributes to the cardiac developmental toxicity of trichloroethylene in zebrafish. Chemosphere 2020, 251, 126610. [Google Scholar] [CrossRef]
  263. Wojciechowska, A.; Braniewska, A.; Kozar-Kaminska, K. MicroRNA in cardiovascular biology and disease. Adv. Clin. Exp. Med. 2017, 26, 865–874. [Google Scholar] [CrossRef] [Green Version]
  264. Bostjancic, E.; Zidar, N.; Stajer, D.; Glavac, D. MicroRNAs miR-1, miR-133a, miR-133b and miR-208 are dysregulated in human myocardial infarction. Cardiology 2010, 115, 163–169. [Google Scholar] [CrossRef] [PubMed]
  265. Chen, C.; Shen, H.; Huang, Q.; Li, Q. The Circular RNA CDR1as Regulates the Proliferation and Apoptosis of Human Cardiomyocytes Through the miR-135a/HMOX1 and miR-135b/HMOX1 Axes. Genet. Test. Mol. Biomarkers 2020, 24, 537–548. [Google Scholar] [CrossRef] [PubMed]
  266. Cao, M.-L.; Zhu, B.-L.; Sun, Y.-Y.; Qiu, G.-R.; Fu, W.-N.; Jiang, H.-K. MicroRNA-144 regulates cardiomyocyte proliferation and apoptosis by targeting TBX1 through the JAK2/STAT1 pathway. Cytogenet. Genome Res. 2019, 159, 190–200. [Google Scholar] [CrossRef] [PubMed]
  267. Ling, X.; Yao, D.; Kang, L.; Zhou, J.; Zhou, Y.; Dong, H.; Zhang, K.; Zhang, L.; Chen, H. Involment of RAS/ERK1/2 signaling and MEF2C in miR-155-3p inhibition-triggered cardiomyocyte differentiation of embryonic stem cell. Oncotarget 2017, 8, 84403–84416. [Google Scholar] [CrossRef] [Green Version]
  268. Guzzolino, E.; Pellegrino, M.; Ahuja, N.; Garrity, D.; D’Aurizio, R.; Groth, M.; Baumgart, M.; Hatcher, C.J.; Mercatanti, A.; Evangelista, M.; et al. miR-182-5p is an evolutionarily conserved Tbx5 effector that impacts cardiac development and electrical activity in zebrafish. Cell Mol. Life Sci. 2020, 77, 3215–3229. [Google Scholar] [CrossRef]
  269. Liu, X.; Yang, Y.; Wang, X.; Guo, X.; Lu, C.; Kang, J.; Wang, G. MiR-184 directly targets Wnt3 in cardiac mesoderm differentiation of embryonic stem cells. Stem Cells 2020, 38, 1568–1577. [Google Scholar] [CrossRef]
  270. Tao, Y.; Zhang, H.; Huang, S.; Pei, L.; Feng, M.; Zhao, X.; Ouyang, Z.; Yao, S.; Jiang, R.; Wei, K. miR-199a-3p promotes cardiomyocyte proliferation by inhibiting Cd151 expression. Biochem. Biophys. Res. Commun. 2019, 516, 28–36. [Google Scholar] [CrossRef]
  271. Chen, H.P.; Wen, J.; Tan, S.R.; Kang, L.M.; Zhu, G.C. MiR-199a-3p inhibition facilitates cardiomyocyte differentiation of embryonic stem cell through promotion of MEF2C. J. Cell. Physiol. 2019, 234, 23315–23325. [Google Scholar] [CrossRef]
  272. Liang, D.; Li, J.; Wu, Y.; Zhen, L.; Li, C.; Qi, M.; Wang, L.; Deng, F.; Huang, J.; Lv, F.; et al. miRNA-204 drives cardiomyocyte proliferation via targeting Jarid2. Int. J. Cardiol. 2015, 201, 38–48. [Google Scholar] [CrossRef]
  273. Tian, Y.; Liu, Y.; Wang, T.; Zhou, N.; Kong, J.; Chen, L.; Snitow, M.; Morley, M.; Li, D.; Petrenko, N.; et al. A microRNA-Hippo pathway that promotes cardiomyocyte proliferation and cardiac regeneration in mice. Sci. Transl. Med. 2015, 7, 279ra238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Hao, D.; Wang, X.; Wang, X.; Thomsen, B.; Yang, Y.; Lan, X.; Huang, Y.; Chen, H. MicroRNA bta-miR-365-3p inhibits proliferation but promotes differentiation of primary bovine myoblasts by targeting the activin A receptor type I. J. Anim. Sci. Biotechnol. 2021, 12, 16. [Google Scholar] [CrossRef] [PubMed]
  275. Xiao, J.; Zhang, Y.; Tang, Y.; Dai, H.; OuYang, Y.; Li, C.; Yu, M. hsa-miR-4443 inhibits myocardial fibroblast proliferation by targeting THBS1 to regulate TGF-beta1/alpha-SMA/collagen signaling in atrial fibrillation. Braz. J. Med. Biol. Res. 2021, 54, e10692. [Google Scholar] [CrossRef] [PubMed]
  276. Meng, X.; Zhang, P.; Zhang, L. Fetal Hypoxia Impacts on Proliferation and Differentiation of Sca-1+ Cardiac Progenitor Cells and Maturation of Cardiomyocytes: A Role of MicroRNA-210. Genes 2020, 11, 328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Fasanaro, P.; D’Alessandra, Y.; Di Stefano, V.; Melchionna, R.; Romani, S.; Pompilio, G.; Capogrossi, M.C.; Martelli, F. MicroRNA-210 modulates endothelial cell response to hypoxia and inhibits the receptor tyrosine kinase ligand Ephrin-A3. J. Biol. Chem. 2008, 283, 15878–15883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  278. Cai, J.; Yang, J.; Liu, Q.; Gong, Y.; Zhang, Y.; Zhang, Z. Selenium deficiency inhibits myocardial development and differentiation by targeting the mir-215-5p/CTCF axis in chicken. Metallomics 2019, 11, 415–428. [Google Scholar] [CrossRef] [PubMed]
  279. Wang, Y.; Liu, J.; Cui, J.; Sun, M.; Du, W.; Chen, T.; Ming, X.; Zhang, L.; Tian, J.; Li, J.; et al. MiR218 Modulates Wnt Signaling in Mouse Cardiac Stem Cells by Promoting Proliferation and Inhibiting Differentiation through a Positive Feedback Loop. Sci. Rep. 2016, 6, 20968. [Google Scholar] [CrossRef] [Green Version]
  280. Cruz, F.M.; Tome, M.; Bernal, J.A.; Bernad, A. miR-300 mediates Bmi1 function and regulates differentiation in primitive cardiac progenitors. Cell Death Dis. 2015, 6, e1953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  281. Shen, X.; Soibam, B.; Benham, A.; Xu, X.; Chopra, M.; Peng, X.; Yu, W.; Bao, W.; Liang, R.; Azares, A.; et al. miR-322/-503 cluster is expressed in the earliest cardiac progenitor cells and drives cardiomyocyte specification. Proc. Natl. Acad. Sci. USA 2016, 113, 9551–9556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Wang, L.; Song, G.; Liu, M.; Chen, B.; Chen, Y.; Shen, Y.; Zhu, J.; Zhou, X. MicroRNA-375 overexpression influences P19 cell proliferation, apoptosis and differentiation through the Notch signaling pathway. Int. J. Mol. Med. 2016, 37, 47–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Clark, A.L.; Naya, F.J. MicroRNAs in the Myocyte Enhancer Factor 2 (MEF2)-regulated Gtl2-Dio3 Noncoding RNA Locus Promote Cardiomyocyte Proliferation by Targeting the Transcriptional Coactivator Cited2. J. Biol. Chem. 2015, 290, 23162–23172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Yang, F.; Qi, J. miR-430a regulates the development of left-right asymmetry by targeting sqt in the teleost. Gene 2020, 745, 144628. [Google Scholar] [CrossRef]
  285. Pisano, F.; Altomare, C.; Cervio, E.; Barile, L.; Rocchetti, M.; Ciuffreda, M.C.; Malpasso, G.; Copes, F.; Mura, M.; Danieli, P.; et al. Combination of miRNA499 and miRNA133 exerts a synergic effect on cardiac differentiation. Stem Cells 2015, 33, 1187–1199. [Google Scholar] [CrossRef] [PubMed]
  286. Neshati, V.; Mollazadeh, S.; Fazly Bazzaz, B.S.; de Vries, A.A.F.; Mojarrad, M.; Naderi-Meshkin, H.; Neshati, Z.; Mirahmadi, M.; Kerachian, M.A. MicroRNA-499a-5p Promotes Differentiation of Human Bone Marrow-Derived Mesenchymal Stem Cells to Cardiomyocytes. Appl. Biochem. Biotechnol. 2018, 186, 245–255. [Google Scholar] [CrossRef] [PubMed]
  287. Li, B.; Wang, Z.; Yang, F.; Huang, J.; Hu, X.; Deng, S.; Tian, M.; Si, X. miR449a5p suppresses CDK6 expression to inhibit cardiomyocyte proliferation. Mol. Med. Rep. 2021, 23, 1. [Google Scholar] [CrossRef]
  288. Del Campo, C.V.; Liaw, N.Y.; Gunadasa-Rohling, M.; Matthaei, M.; Braga, L.; Kennedy, T.; Salinas, G.; Voigt, N.; Giacca, M.; Zimmermann, W.H.; et al. Regenerative potential of epicardium-derived extracellular vesicles mediated by conserved miRNA transfer. Cardiovasc. Res. 2022, 118, 597–611. [Google Scholar] [CrossRef] [PubMed]
  289. Song, Y.; Zhang, C.; Zhang, J.; Jiao, Z.; Dong, N.; Wang, G.; Wang, Z.; Wang, L. Localized injection of miRNA-21-enriched extracellular vesicles effectively restores cardiac function after myocardial infarction. Theranostics 2019, 9, 2346–2360. [Google Scholar] [CrossRef] [PubMed]
  290. Kang, K.; Ma, R.; Cai, W.; Huang, W.; Paul, C.; Liang, J.; Wang, Y.; Zhao, T.; Kim, H.W.; Xu, M.; et al. Exosomes Secreted from CXCR4 Overexpressing Mesenchymal Stem Cells Promote Cardioprotection via Akt Signaling Pathway following Myocardial Infarction. Stem Cells Int. 2015, 2015, 659890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Lai, R.C.; Arslan, F.; Lee, M.M.; Sze, N.S.; Choo, A.; Chen, T.S.; Salto-Tellez, M.; Timmers, L.; Lee, C.N.; El Oakley, R.M.; et al. Exosome secreted by MSC reduces myocardial ischemia/reperfusion injury. Stem Cell Res. 2010, 4, 214–222. [Google Scholar] [CrossRef] [Green Version]
  292. Shao, L.; Zhang, Y.; Lan, B.; Wang, J.; Zhang, Z.; Zhang, L.; Xiao, P.; Meng, Q.; Geng, Y.J.; Yu, X.Y.; et al. MiRNA-Sequence Indicates That Mesenchymal Stem Cells and Exosomes Have Similar Mechanism to Enhance Cardiac Repair. BioMed Res. Int. 2017, 2017, 4150705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Wang, S.; Li, L.; Liu, T.; Jiang, W.; Hu, X. miR-19a/19b-loaded exosomes in combination with mesenchymal stem cell transplantation in a preclinical model of myocardial infarction. Regen. Med. 2020, 15, 1749–1759. [Google Scholar] [CrossRef] [PubMed]
  294. Huh, D.; Kim, H.J.; Fraser, J.P.; Shea, D.E.; Khan, M.; Bahinski, A.; Hamilton, G.A.; Ingber, D.E. Microfabrication of human organs-on-chips. Nat. Protoc. 2013, 8, 2135–2157. [Google Scholar] [CrossRef]
  295. Tandon, N.; Cannizzaro, C.; Chao, P.H.G.; Maidhof, R.; Marsano, A.; Au, H.T.H.; Radisic, M.; Vunjak-Novakovic, G. Electrical stimulation systems for cardiac tissue engineering. Nat. Protoc. 2009, 4, 155–173. [Google Scholar] [CrossRef] [Green Version]
  296. Huang, Y.; Jia, X.; Bai, K.; Gong, X.; Fan, Y. Effect of fluid shear stress on cardiomyogenic differentiation of rat bone marrow mesenchymal stem cells. Arch. Med. Res. 2010, 41, 497–505. [Google Scholar] [CrossRef]
  297. Toh, Y.C.; Voldman, J. Fluid shear stress primes mouse embryonic stem cells for differentiation in a self-renewing environment via heparan sulfate proteoglycans transduction. FASEB J. 2011, 25, 1208–1217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  298. Cheng, F.; Ping, Z.; Handong, Y.; Zhengtong, Y.; Zhaodong, Z. Induced differentiation of human cord blood mesenchymal stem/progenitor cells into cardiomyocyte-like cells in vitro. J. Huazhong. Univ. Sci. Technolog. Med. Sci. 2003, 23, 154–157. [Google Scholar] [PubMed]
  299. Pavesi, A.; Adriani, G.; Rasponi, M.; Zervantonakis, I.K.; Fiore, G.B.; Kamm, R.D. Controlled electromechanical cell stimulation on-a-chip. Sci. Rep. 2015, 5, 11800. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  300. Taylor, S.M.; Jones, P.A. Multiple new phenotypes induced in 10T1/2 and 3T3 cells treated with 5-azacytidine. Cell 1979, 17, 771–779. [Google Scholar] [CrossRef]
  301. Sokolowska, P.; Zukowski, K.; Lasocka, I.; Szulc-Dabrowska, L.; Jastrzebska, E. Human mesenchymal stem cell (hMSC) differentiation towards cardiac cells using a new microbioanalytical method. Analyst 2020, 145, 3017–3028. [Google Scholar] [CrossRef]
  302. Kakkar, A.; Nandy, S.B.; Gupta, S.; Bharagava, B.; Airan, B.; Mohanty, S. Adipose tissue derived mesenchymal stem cells are better respondents to TGFbeta1 for in vitro generation of cardiomyocyte-like cells. Mol. Cell Biochem. 2019, 460, 53–66. [Google Scholar] [CrossRef]
  303. Kim, Y.S.; Ahn, Y.; Kwon, J.S.; Cho, Y.K.; Jeong, M.H.; Cho, J.G.; Park, J.C.; Kang, J.C. Priming of mesenchymal stem cells with oxytocin enhances the cardiac repair in ischemia/reperfusion injury. Cells Tissues Organs. 2012, 195, 428–442. [Google Scholar] [CrossRef] [PubMed]
  304. Li, T.S.; Komota, T.; Ohshima, M.; Qin, S.L.; Kubo, M.; Ueda, K.; Hamano, K. TGF-beta induces the differentiation of bone marrow stem cells into immature cardiomyocytes. Biochem. Biophys. Res. Commun. 2008, 366, 1074–1080. [Google Scholar] [CrossRef] [PubMed]
  305. Huang, G.; Li, F.; Zhao, X.; Ma, Y.; Li, Y.; Lin, M.; Jin, G.; Lu, T.J.; Genin, G.M.; Xu, F. Functional and Biomimetic Materials for Engineering of the Three-Dimensional Cell Microenvironment. Chem. Rev. 2017, 117, 12764–12850. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Histone modifiers network and their effect on heart development genes and their function. Histone acetyltransferases (HAT), histone deacetylases (HDAC), histone demethylase (HDM), and histone methyltransferase (HMT) (gray octagons) are four families that include (black line) different enzymes. Some of these enzymes (white circles) modify histones and effect cardiac evolution by impacting cardiac genes (light blue circles). Activation (dark blue line), inhibition (red line), deacetylation (purple line), up-regulation (light blue line), down-regulation (yellow line), silencing (dark brown line), execution (orange line), and maintaining (green line) are some of these modifications that are performed by these enzymes on cardiac genes. The results of the influence of these enzymes (white circles) on cardiac genes (light blue circles) are shown in Figure 1 as green triangles. (Figure 1 was generated using Cytoscape application—cytoscape.org (accessed on 28 September 2022)).
Figure 1. Histone modifiers network and their effect on heart development genes and their function. Histone acetyltransferases (HAT), histone deacetylases (HDAC), histone demethylase (HDM), and histone methyltransferase (HMT) (gray octagons) are four families that include (black line) different enzymes. Some of these enzymes (white circles) modify histones and effect cardiac evolution by impacting cardiac genes (light blue circles). Activation (dark blue line), inhibition (red line), deacetylation (purple line), up-regulation (light blue line), down-regulation (yellow line), silencing (dark brown line), execution (orange line), and maintaining (green line) are some of these modifications that are performed by these enzymes on cardiac genes. The results of the influence of these enzymes (white circles) on cardiac genes (light blue circles) are shown in Figure 1 as green triangles. (Figure 1 was generated using Cytoscape application—cytoscape.org (accessed on 28 September 2022)).
Life 13 00569 g001
Figure 2. Epigenetic roles of miRNAs in cardiac-specific gene expression. Various miRNA subtypes (green circle) role in the inhibition (red arrows) and activation (dark blue arrows) of specific transcription factors (light blue triangle), which are responsible for differentiation, proliferation, development, growth, remodeling, reprogramming, and regeneration of cardiomyocytes (green octagon). (Figure 2 was generated using Cytoscape application—cytoscape.org (accessed on 28 September 2022)).
Figure 2. Epigenetic roles of miRNAs in cardiac-specific gene expression. Various miRNA subtypes (green circle) role in the inhibition (red arrows) and activation (dark blue arrows) of specific transcription factors (light blue triangle), which are responsible for differentiation, proliferation, development, growth, remodeling, reprogramming, and regeneration of cardiomyocytes (green octagon). (Figure 2 was generated using Cytoscape application—cytoscape.org (accessed on 28 September 2022)).
Life 13 00569 g002
Table 2. MicroRNAs involved in heart regeneration, their target, and their effect.
Table 2. MicroRNAs involved in heart regeneration, their target, and their effect.
MicroRNAsEffect(s) on GeneAffected GeneGene FunctionReferences
miR-1ActivationDELTADifferentiation[221,223,224,244,245]
InhibitionCDK9Inhibition of differentiation
InhibitionGATA4Proliferation
InhibitionHAND2Proliferation
InhibitionHSP70Proliferation
InhibitionRALDH2Proliferation
InhibitionIGF2BProliferation
InhibitionIGF1Proliferation
InhibitionFRS2Proliferation
InhibitionFOXP1Proliferation
InhibitionHAND2Inhibition of differentiation
InhibitionHDAC4Inhibition of differentiation
ActivationNKX2.5Differentiation
InhibitionCCND1Proliferation
InhibitionNANOGInhibition of differentiation
InhibitionOCT4Inhibition of differentiation
ActivationTROPTDifferentiation
ActivationA-ACTININDifferentiation
ActivationKDM6ADifferentiation
ActivationKDM6BDifferentiation
ActivationTBX5Differentiation
ActivationMEF2CDifferentiation
ActivationGATA4Differentiation
miR-10bInhibitionLATS1Inhibition of proliferation[246]
miR-15InhibitionCHEK1Proliferation[247]
InhibitionCDC2AProliferation
InhibitionBIRC5Proliferation
InhibitionSPAG5Proliferation
miR-17InhibitionFOG2Inhibition of differentiation[248]
miR-19aInhibitionS1PR1Inhibition of development[249,250]
InhibitionBIMInhibition of proliferation
InhibitionPTENInhibition of proliferation
ActivationCCNB1Proliferation
ActivationCCND1Proliferation
ActivationCDK1Proliferation
miR-19bInhibitionWNTDifferentiation[228,250]
InhibitionCTNNBDifferentiation
InhibitionPTENInhibition of proliferation
InhibitionBIMInhibition of proliferation
ActivationCCNB1Proliferation
ActivationCCND1Proliferation
ActivationCDK1Proliferation
miR-20aInhibitionSMOProliferation
Differentiation
[251]
MiR-23ActivationTGF-Β1Proliferation[235]
miR-25InhibitionFBXW7Inhibition of proliferation[234,252]
InhibitionBIMInhibition of proliferation
miR-26bInhibitionWNTDifferentiation[253]
miR-29cInhibitionWNT4Development[254]
miR-29b-3pInhibitionNOTCH2Proliferation[255]
Development
miR-30cInhibitionGLI2Inhibition of proliferation[256]
Differentiation
miR-31a-5pInhibitionRHOBTB1Inhibition of proliferation[231]
miR-34aInhibitionC-SKIInhibition of proliferation[257,258]
InhibitionNOTCH1Development
miR-93InhibitionFOG2Inhibition of differentiation[248,257]
InhibitionC-SKIInhibition of proliferation
miR-101aActivationFOSABProliferation[259]
miR-106aInhibitionFOG2Inhibition of differentiation[248]
miR-125b-5pInhibitionERBB4Inhibition of development[229]
mir-127-3pInhibitionKMT5AProliferation[260]
miR-128InhibitionSUZ12Proliferation[261]
Regeneration
miR-128aInhibitionISL1Development[226]
InhibitionSFRP5Development
InhibitionHCN4Development
InhibitionNKX2.5Development
InhibitionMEF2CDevelopment
InhibitionMYL2Development
ActivationIRX4Inhibition of development
miR-133ActivationBMP2Differentiation[224,225]
ActivationGATA4Differentiation
ActivationNKX2.5Differentiation
InhibitionFGF8Inhibition of differentiation
ActivationKDM6ADifferentiation
ActivationKDM6BDifferentiation
ActivationTBX5Differentiation
ActivationMEF2CDifferentiation
ActivationCX43Differentiation
ActivationCTNTDifferentiation
miR-133aInhibitionNKX2.5Development[223,244,247,262,263,264]
InhibitionPGAM1Development
InhibitionGJA1Development
InhibitionSRFDevelopment
InhibitionCCND2Development
InhibitionIGF1RProliferation
InhibitionHAND2Development
InhibitionA-ACTININDevelopment
InhibitionCTNTDevelopment
InhibitionVEGFProliferation
InhibitionCTGFProliferation
InhibitionFGFR1Proliferation
InhibitionPP2ACProliferation
InhibitionTBX18Proliferation
miR-134InhibitionMEIS2Proliferation[227]
InhibitionCYCLIN AProliferation
InhibitionCYCLIN BProliferation
InhibitionCYCLIN EProliferation
InhibitionCDC2Proliferation
InhibitionCDK4Proliferation
InhibitionPCNAProliferation
miR-135aInhibitionCDR1Proliferation[265]
miR-135bInhibitionCDR1Proliferation[265]
miR-144ActivationTBX1Proliferation[266]
miR-155-3pInhibitionMEF2CDifferentiation[267]
InhibitionCTNTDevelopment
InhibitionGATA4Development
InhibitionNKX2.5Development
miR-182-5pInhibitionCACNB2.2Development[268]
InhibitionCACNB4BDevelopment
InhibitionCACNA2D1Development
miR-184InhibitionWNT3Differentiation[269]
miR-199aActivationCLIC5Inhibition of proliferation[247]
ActivationHOMER1Inhibition of proliferation
miR-199a-3pInhibitionCD151Inhibition of proliferation[270]
InhibitionMEF2CDifferentiation[271]
miR-199a-5pInhibitionERBB4Inhibition of development[229]
miR -200cInhibitionGATA4Differentiation[232]
Development
InhibitionSRFDifferentiation
Development
InhibitionTBX5Differentiation
Development
InhibitionCACNA1CDifferentiation
Development
InhibitionKCNJ2Differentiation
Development
InhibitionSCN5ADifferentiation
Development
miR-204InhibitionJARID2Inhibition of proliferation[272]
miR302-367InhibitionMST1Inhibition of proliferation[273]
InhibitionLATS2Inhibition of proliferation
InhibitionMOB1BInhibition of proliferation
InhibitionCCND1Proliferation
InhibitionGATA4Differentiation
InhibitionNKX2.5Development
InhibitionMYH6Differentiation
InhibitionMYH7Differentiation
bta-miR-365-3pInhibitionACVR1Inhibition of differentiation[274]
Proliferation
InhibitionCCND1Proliferation
InhibitionCDK2Proliferation
InhibitionPCNAProliferation
ActivationMYOD1Differentiation
ActivationMYOGDifferentiation
hsa-miR-4443InhibitionTHBS1Proliferation[275]
InhibitionCOLLAGEN IProliferation
InhibitionCOLLAGEN IIIProliferation
InhibitionSMAD2Proliferation
InhibitionSMAD3Proliferation
InhibitionA-SMAProliferation
miR-208ActivationKDM6ADifferentiation[224]
ActivationKDM6BDifferentiation
ActivationTBX5Differentiation
ActivationMEF2CDifferentiation
ActivationGATA4Differentiation
miR-210InhibitionEPHRIN-A3Differentiation[276,277]
miR-215-5pInhibitionCTCFDevelopment[278]
Differentiation
MiR218InhibitionSFRP2Inhibition of proliferation[279]
Differentiation
InhibitionA-MHCDevelopment
Differentiation
InhibitionΒ-MHCDevelopment
InhibitionACTC1Development
InhibitionTNIDifferentiation
miR-221InhibitionERBB4Inhibition of development[229]
miR-222ActivationANPProliferation[228,229]
Growth
ActivationBNPProliferation
Growth
ActivationACTA1Proliferation
Growth
ActivationMHC-AProliferationGrowth
ActivationMHC- ΒProliferation
Growth
ActivationCX-43Development
ActivationKIR2.1Development
InhibitionERBB4Inhibition of development
InhibitionHIPK1Remodeling
InhibitionHIPK2Remodeling
InhibitionHMBOX1Remodeling
InhibitionP27Remodeling
miR-294InhibitionWEE1Inhibition of proliferation[230]
ActivationCCNB1Proliferation
ActivationCCND1Proliferation
ActivationCCNE1Proliferation
ActivationCCNA2Proliferation
ActivationCDK1Proliferation
ActivationE2F1Proliferation
ActivationE2F3Proliferation
miR-300InhibitionBMI1Differentiation[280]
miR302-367InhibitionMST1Inhibition of proliferation[273]
InhibitionMOB1BInhibition of proliferation
InhibitionLATS2Inhibition of proliferation
miR-322InhibitionCELF1Inhibition of differentiation[281]
ActivationTBX5Differentiation
ActivationMEF2CDifferentiation
ActivationNKX2.5Differentiation
ActivationA-MHCDifferentiation
ActivationA-ACTININDifferentiation
ActivationCTNTDifferentiation
miR-375InhibitionNOTCH2Differentiation[233,282]
Proliferation
Development
InhibitionDLL1Differentiation
Proliferation
InhibitionHES1Differentiation
Proliferation
InhibitionCTNTDifferentiation
InhibitionNKX2.5Differentiation
InhibitionGATA4Differentiation
miR-410InhibitionCITED2Inhibition of proliferation[283]
ActivationPCNAProliferation
InhibitionCITED2Development
InhibitionERFI1Inhibition of proliferation
InhibitionPPP1CBGrowth
InhibitionREREInhibition of proliferation
InhibitionSTAT3Differentiation
miR-430aInhibitionPOSQTInhibition of development[284]
miR-495InhibitionCITED2Inhibition of proliferation[283]
ActivationPCNAProliferation
InhibitionCITED2Development
InhibitionERFI1Inhibition of proliferation
InhibitionPPP1CBGrowth
InhibitionREREInhibition of proliferation
InhibitionSTAT3Differentiation
miR-499InhibitionKDM6AReprogramming[224]
InhibitionKDM6BReprogramming
ActivationTBX5Differentiation
ActivationMEF2CDifferentiation
ActivationGATA4Differentiation
ActivationCX43Differentiation[285]
ActivationCTNTDifferentiation
ActivationNKX2.5Differentiation
ActivationGATA4Differentiation
miR -499a-5pActivationA-ACTININDifferentiation[286,287]
ActivationCTNIDifferentiation
InhibitionCDK6Proliferation
miR-503InhibitionCELF1Inhibition of differentiation[281]
ActivationTBX5Differentiation
ActivationMEF2CDifferentiation
ActivationNKX2.5Differentiation
ActivationA-ACTININDifferentiation
ActivationA-MHCDifferentiation
ActivationCTNTDifferentiation
miR-590ActivationCLIC5Inhibition of proliferation[247]
ActivationHOMER1Inhibition of proliferation
miR-699aInhibitionMYODInhibition of differentiation
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zare, A.; Salehpour, A.; Khoradmehr, A.; Bakhshalizadeh, S.; Najafzadeh, V.; Almasi-Turk, S.; Mahdipour, M.; Shirazi, R.; Tamadon, A. Epigenetic Modification Factors and microRNAs Network Associated with Differentiation of Embryonic Stem Cells and Induced Pluripotent Stem Cells toward Cardiomyocytes: A Review. Life 2023, 13, 569. https://doi.org/10.3390/life13020569

AMA Style

Zare A, Salehpour A, Khoradmehr A, Bakhshalizadeh S, Najafzadeh V, Almasi-Turk S, Mahdipour M, Shirazi R, Tamadon A. Epigenetic Modification Factors and microRNAs Network Associated with Differentiation of Embryonic Stem Cells and Induced Pluripotent Stem Cells toward Cardiomyocytes: A Review. Life. 2023; 13(2):569. https://doi.org/10.3390/life13020569

Chicago/Turabian Style

Zare, Afshin, Aria Salehpour, Arezoo Khoradmehr, Shabnam Bakhshalizadeh, Vahid Najafzadeh, Sahar Almasi-Turk, Mahdi Mahdipour, Reza Shirazi, and Amin Tamadon. 2023. "Epigenetic Modification Factors and microRNAs Network Associated with Differentiation of Embryonic Stem Cells and Induced Pluripotent Stem Cells toward Cardiomyocytes: A Review" Life 13, no. 2: 569. https://doi.org/10.3390/life13020569

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