Next Article in Journal
Protective Effects of Pituitary Adenylate-Cyclase-Activating Polypeptide on Retinal Vasculature and Molecular Responses in a Rat Model of Moderate Glaucoma
Next Article in Special Issue
Plasma miR-486-5p Expression Is Upregulated in Atrial Fibrillation Patients with Broader Low-Voltage Areas
Previous Article in Journal
Identification of Breast Cancer LCK Proto-Oncogene as a Master Regulator of TNBC Neutrophil Enrichment and Polarization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

miRNAs Epigenetic Tuning of Wall Remodeling in the Early Phase after Myocardial Infarction: A Novel Epidrug Approach

by
Francesca Salvatori
1,
Elisabetta D’Aversa
1,
Maria Luisa Serino
2,
Ajay Vikram Singh
3,
Paola Secchiero
1,
Giorgio Zauli
4,
Veronica Tisato
1,5,6,* and
Donato Gemmati
1,2,6
1
Department of Translational Medicine, University of Ferrara, 44121 Ferrara, Italy
2
Centre Haemostasis & Thrombosis, University of Ferrara, 44121 Ferrara, Italy
3
Department of Chemical and Product Safety, German Federal Institute for Risk Assessment (BfR), 10589 Berlin, Germany
4
Department of Environmental Science and Prevention, University of Ferrara, 44121 Ferrara, Italy
5
LTTA Centre, University of Ferrara, 44121 Ferrara, Italy
6
University Centre for Studies on Gender Medicine, University of Ferrara, 44121 Ferrara, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(17), 13268; https://doi.org/10.3390/ijms241713268
Submission received: 3 August 2023 / Revised: 21 August 2023 / Accepted: 22 August 2023 / Published: 26 August 2023
(This article belongs to the Special Issue Current Research for Heart Disease Biology and Therapeutics)

Abstract

:
Myocardial infarction (MI) is one of the leading causes of death in Western countries. An early diagnosis decreases subsequent severe complications such as wall remodeling or heart failure and improves treatments and interventions. Novel therapeutic targets have been recognized and, together with the development of direct and indirect epidrugs, the role of non-coding RNAs (ncRNAs) yields great expectancy. ncRNAs are a group of RNAs not translated into a product and, among them, microRNAs (miRNAs) are the most investigated subgroup since they are involved in several pathological processes related to MI and post-MI phases such as inflammation, apoptosis, angiogenesis, and fibrosis. These processes and pathways are finely tuned by miRNAs via complex mechanisms. We are at the beginning of the investigation and the main paths are still underexplored. In this review, we provide a comprehensive discussion of the recent findings on epigenetic changes involved in the first phases after MI as well as on the role of the several miRNAs. We focused on miRNAs function and on their relationship with key molecules and cells involved in healing processes after an ischemic accident, while also giving insight into the discrepancy between males and females in the prognosis of cardiovascular diseases.

1. Introduction

The heart is a highly specialized organ, whose continuous contraction allows perfusion, oxygenation and, therefore, the survival of cells and tissues. The cardiomyocytes (CMs) responsible for contraction are only one of the numerous cell types present in the heart and involved in its homeostasis. In fact, they do not act alone but work as part of a complex network made up of specialized cells, including those involved in tissue vascularization, the remodeling of the interstitial space, the regulation of contractions and in the innate and adaptive immune response [1] (Figure 1).
Although CMs represent the functional “mass” of the heart, fibroblasts are present in equivalent numbers. Cardiac fibroblasts (CFs) are cells of mesenchymal origin capable of secreting components of the extracellular matrix (ECM), but in the heart, they are also involved in regulating cardiac form and function, mainly through paracrine effects via the secretion of growth factors and cytokines. Fibroblasts communicate with myocytes and endothelial cells (ECs), allowing for the regulation of electrophysiological properties and increased blood vessel formation [1,2]. Another important cellular component of the heart is represented by resident immune cells, which have an important role in cardiac homeostasis, immunological and repair processes [3]. Apart from a small population of T cells and neutrophils, most of the resident immune cells are made up of macrophages and mast cells (MCs) [1,4]. The latter are located mainly in the pericardial adipose tissue and appear to be of the MCTC subtype containing tryptase, chymase, cathepsin G, and carboxypeptidase and are usually found in connective tissues. Cardiac MCs also contain tumor necrosis factor-alpha (TNF-α) and transforming growth factor-beta (TGF-β), involved in myocardial remodeling [5,6,7,8].
The most common resident immune cells in the heart are macrophages. They can be classified into two subgroups based on the expression of C–C motif chemokine receptor 2 (CCR2+ and CCR2−), which binds to monocyte chemoattractant protein 1 (CCL2), which allows monocyte infiltration during inflammation. The presence of this receptor also allows us to discriminate between the macrophages that derive from yolk-sac precursors (CCR2−) and those that develop from circulating monocytes (CCR2−). The CCR2− and CCR2+ macrophages show different functions in the healing heart, which will be discussed later [9,10]. Macrophages of embryonic origin proliferate autonomously in the heart, keeping their number constant, although the proliferation rate decreases over time [11,12]. Conversely, CCR2+ macrophages are regularly replaced by bone marrow or extramedullary tissues derived monocytes [13]. CCR2− cardiac-resident macrophages can engulf dying cells, cellular debris, or pathogens [14], and they play a key role in the proper development of the heart’s vasculature [15], angiogenesis [16], lymphangiogenesis and maturation of lymphatic vessels [17]. Moreover, being widely present in the atrioventricular as well as in the myocardium, they appear to improve cardiac conductivity, acting as a “bridge” between CMs that are not in direct contact via gap junctions (mainly gap junction alpha-1 protein (Cx43) protein) [4,9,18].

Epigenetic Regulation

Various features of cardiac cells are regulated by epigenetic factors. It has been long established that CMs, which are very active in replication during the embryonic-fetal period, drastically reduce their proliferative capacity after birth, leading to the maximum number of cells being reached one month after birth. Since then, the increase in heart size is mainly due to cell hypertrophy, although evidence highlights a minimal replicative capacity of adult CMs, though not sufficient enough to regenerate heart tissue following injury [19].
DNA methylation is the main epigenetic form of gene regulation, and it mainly occurs at the level of the CpG islands in the promoter regions, so gene expression in adult CMs is partially regulated by the methylation of CpG islands. Of note, Gilsbach and colleagues identified 79,655 differently methylated regions (DMRs) by comparing adult CMs with embryonic stem cells (ESCs) [20]. Of these, 90% are hypermethylated while only 10% are hypomethylated, where demethylation directly correlates with active histone marks and increased expression of cardiomyocyte genes [20]. This correlation is not always maintained, since not all demethylated genes in CMs with respect to ESCs are effectively expressed. Many of these methylation changes are associated with trimethylated histone H3 at amino acid K27 (H3K27me3), which is in turn associated with polycomb-mediated gene repression and thus with blocking of the expression [20].
Another gene regulatory mechanism highly involved in CMs biology is histone acetylation. Acetylation of lysines reduces the charge of histones, resulting in the loss of their electrostatic attraction to DNA. This allows for the opening of the chromatin and, therefore, the activation of gene transcription. Chromatin hyperacetylation is present in embryonic cells (H3K9/14, H3K18, and H3K27), concomitant with high expression of p300 acetyltransferase, which decreases rapidly after birth, resulting in hypoacetylation of histones [21,22]. Accordingly, p300 acetyltransferase has been considered a putative epidrug target to contrast cellular damage related to aging and cardiovascular (CV) disease [23]. Epigenetic-based therapy is capturing growing interest due to the benefits of direct (e.g., Apabetalone) or indirect epidrugs such as statins, metformin, SGLT2i, and W-3 via epigenome modifications. On the other hand, also already available drugs have potential indirect epigenetic effects (e.g., metformin, statins, sodium-glucose transporter inhibitors 2, phytochemicals and nutritive vitamins/supplements). This paves the way for the development of novel therapeutics for the treatment of major adverse CV events [24,25].
Histone deacetylases (HDACs) are also involved in the regulation of CMs gene expression and proliferation. In particular, class I HDACs seem to be involved in the suppression of proliferation, such as HDAC2, which interacts with the cardiac negative regulator HOPX (homeodomain-only), reducing acetylation of GATA4 with consequent suppression of CMs cycling [26]. In contrast, class II HDACs, such as HDAC5 and HDAC9, appear to be involved in the stimulation of CMs proliferation. The activities of HDAC5 and HDAC9 have been shown to be redundant, which is why the loss of one of their genes does not drastically alter the effects on the heart [27].
Another important aspect of gene regulation in CMs is related to the effects of non-coding RNAs (ncRNAs), long non-coding RNAs (lncRNAs) and microRNAs (miRNAs). Among these, miRNAs are certainly the best characterized and studied non-coding RNAs. They consist of 18–22 nucleotides and allow for the regulation of gene expression by binding to the 3′ untranslated region (3′UTR) of the transcript, stimulating its degradation or blocking its protein translation [28]. The miRNAs involved in the regulation of CMs can be divided into those that stimulate the proliferation and/or hypertrophy, defined as “pro-proliferative” (Figure 2), and those that instead reduce the proliferative capacity or increase apoptosis, defined as “pro-apoptotic” (Figure 3). In both cases, miRNAs show multiple ways of action directly on cell cycle regulatory proteins, such as the pro-proliferative miR-499 which blocks the expression of SOX6, activating accordingly the cyclin D1 promoter [29]. Another example is the pro-apoptotic miR-195 which inhibits the expression of checkpoint kinase 1 (Check1) [30,31], which regulates G1/S transition, S phase, mitotic entry and mitosis [32]. miRNAs also regulate CMs proliferation through several signaling pathways, some of which are targeted by both pro-proliferative and pro-apoptotic miRNAs. This is the case of the Akt pathway, whose kinase activity can be increased by inhibitors of the expression of the PTEN tumor suppressor, such as miR17-92, miR-19a/b or miR-221-3p [33,34,35] or silenced by miR-208a and miR-489, which block phosphoinositide 3-kinase (PI3K) and spindlin-1 (SPIN1), respectively [36,37]. Conversely, other pathways appear to be regulated only by pro-proliferative or pro-apoptotic miRNAs, such as the NF-kB pathway which can be activated by blockade of the suppressor of cytokine signaling 3 (SOCS3) and the NF-kB inhibitor interacting Ras-like 2 (NKIRAS2) by miR-324 [38] and miR-1180 [39], respectively. However, the Nodal signaling pathway, an important signal transduction pathway active during embryonic heart development [40], is suppressed by miR-33a-5p which targets nodal modulator 1 (NOMO1) [41].
Summarizing, Table 1 and Table 2 show the main pro-proliferative and pro-apoptotic miRNAs studied so far, most of which are mainly found in the embryonic-neonatal stage, while others have also been identified in the adult stage. Certainly, the various aspects of the epigenetic regulation of CMs proliferation are not totally separate from each other. For example, both the pro-proliferative miR-204 [42] and the pro-apoptotic miR-26a [43] target the polycomb repressive complex with histone methyltransferase activity, involved in the silencing of genes in hypomethylated regions of CMs genome [20], while HDAC3 promotes myocardial growth by stimulating fibroblast growth factor-9 (FGF9) and insulin-like growth factor-2 (IGF2) through repressing the pro-apoptotic miR-322 and miR-503 [44].

2. Myocardial Infarction

Myocardial infarction (MI) is the result of acute or chronic ischemia, due to a significant reduction of blood flow responsible for cutting in oxygen and glucose supply, leading to myocardial injury up to tissue death [81]. Although it may have a complex multifactorial origin, atherosclerosis represents the primary etiopathological risk factor. Atherosclerosis in turn can be caused by modifiable risk factors, such as hyperlipidemia, hypertension, diabetes mellitus, obesity, smoking, psychosocial stress, lack of exercise and a sedentary lifestyle, and non-modifiable risk factors such as genetic predisposition together with age progression, male gender and positive family history [82,83,84,85,86] (Figure 4).
MI of atherosclerosis origin arises from intracoronary occlusive thrombosis superimposed on a pre-existent atherosclerotic plaque. The thrombus is usually produced following fissuring or ulceration of the plaque with consequent exposure to the bloodstream of its core or tissue factor and subsequent activation of the thrombogenic response [87,88]. In detail, specific actions of inherited risk/protective factors have been reported by investigating selective haplotypes in coagulation factors genes such as F13A1, F13B, F7 and F5, resulting in modified levels of coagulation factors or different activation rates with significant modifications in the thrombus formation rates [89,90,91,92]. Other risk factors, deriving from rare conditions, can give rise to occlusions or severe reduction in myocardial perfusion. These include vasculitis that can cause coronary occlusions, ventricular hypertrophy, production of emboli following trauma to coronary arteries, coronary artery abnormalities (including aneurysms, aortic dissection), severe anemia, or pulmonary infection. Furthermore, not only a reduced oxygen supply to the myocardium can cause ischemia, but also an increased demand, due to hypothyroidism, fever, and heavy exertion [85,87]. Moreover, restoration of blood flow after a period of ischemia can cause potentially very harmful effects such as necrosis of damaged cells, marked cell swelling and restoration of non-uniform flow in all portions of the tissue. The latter, called the reflux phenomenon, is the result of a vicious cycle of vascular, endothelial, and mitochondrial dysfunctions, with reduced local perfusion, major dysfunctional changes, edema, and much more. All these functional and structural changes are referred to as ischemia/reperfusion (I/R) injury [93].
Although in mammals CMs contain large stores of energy phosphate, in in vivo models, following ischemia, the onset of systolic dysfunction is much more rapid than expected. This is because the breakdown of the creatinine phosphate molecules leads to the production of inorganic phosphate which inhibits the proteins with contractile function. Furthermore, contractility is inhibited by calcium depletion following intracellular acidosis. The blockage of CMs’ aerobic metabolism induces a rapid reduction of ATP and a concomitant increase of anaerobic metabolites, such as lactate. If this block is maintained for less than 10–20 min, cardiac functions can be totally reversible, while if its duration is longer, the damage becomes irreversible and CMs are no longer able to maintain their structural integrity [87].

2.1. Inflammation Phase

During the very early stages following ischemic injury, MCs and soluble complement proteins become important initiators of inflammation [94], leading (in the absence of a rapid reconstitution of coronary blood flow) to loss of CMs, which undergo cell death by apoptosis, necrosis, and necroptosis. The former includes the extrinsic or death receptor pathway and the intrinsic or mitochondrial pathway, although it has been shown that the two are linked and mutually influence each other [95]. The extrinsic and intrinsic apoptotic pathways join into a common trail starting with caspase-3 cleavage up to DNA fragmentation, degradation of cytoskeletal and nuclear proteins, proteins cross-linking, expression of ligands for phagocytic cell and apoptotic bodies formation. The latter are then phagocytosed by macrophages and/or the neighboring cells before they fragment [96].
Among the different types of cell death, necrosis is an uncontrolled form of cell death due to sudden damage, such as a hypoxic event. It is characterized by the swelling of inner cell structures, rupture of plasma membrane up to the complete lysis of the cell and release of intracellular contents leading to tissue damage [97]. Finally, necroptosis is activated by death receptors but determines the production of reactive oxygen species (ROS) and the depletion of ATP, as in necrosis. However, it remains to be clarified how these three cell death mechanisms interact with each other during the ischemic event [98]. On the contrary, what is well known is that necrotic cells release endogenous damage-associated molecular patterns (DAMPs), that can be recognized by pathogen recognition receptors (PRR) and Toll-like receptors (TLRs), a family of transmembrane receptors that activate downstream pro-inflammatory cascade and recruit neutrophils, macrophages, monocytes, MCs and dendritic cells (DCs) [4,87]. After the blood flow has been restored, the release by the cardiac resident MCs of preformed pro-inflammatory mediators, such as TNF-α, stimulates the CMs apoptosis, and the release of histamine and various proteases, which amplify the inflammatory signal involving ECs, resident macrophages, and subsequently infiltrating monocytes [4,94]. Finally, the activated complement system contributes to inflammatory signaling by accentuating CMs necrosis, inducing proinflammatory responses and mediating leukocyte influx in the injured myocardium [87].
Neutrophils accumulate in large numbers in damaged tissue in the first hours after ischemia, localizing in post-capillary venules. In the first phase, they are activated before infiltrating damaged tissue. Following the release of formylated peptides and mitochondrial DNA, detected, respectively, by formyl peptide receptor 1 (FPR1) and TLR9, there is not only the activation of neutrophils but their chemotaxis towards the damaged area, with consequent infiltration [99]. The second phase of neutrophil recruitment involves cardiac ECs, activated by cytokines such as TNF-α and histamine, resulting in increased expression of selectins. The interaction between selectins and their receptors allows neutrophils to roll along the venule wall, where they more easily meet activating factors, such as IL-8 and C5a. These interactions lead to neutrophil integrin activation, which binds to members of the immunoglobulin superfamily expressed in stimulated ECs and allow the firm adhesion of the leukocytes. Trans endothelial migration follows and leads to neutrophil infiltration in the inflamed tissues [100,101], where they begin to engulf the dead or dying CMs, thus reducing the presence of pro-inflammatory DAMPs. At the same time, neutrophils cause the release of oxygen free radicals, inflammatory mediators, and proteolytic enzymes, such as metalloproteinases (MMPs), resulting in degradation of the ECM and cardiac rupture essentially in the absence or severe reduction of expression of F13A1 gene as reported in a mice animal model [102]. The causal relationship between MMP-9 levels and heart rupture in infarcted myocardium has been established by Heymans in KO mice for the MMP9 gene which protects against cardiac rupture [103]. This was further corroborated by the observation that plasma FXIIIA levels significantly drop in the first hours after MI [104], and appropriate circulating levels of FXIIIA molecules or F13A1 gene variants can contrast the unrestrained MMPs action in MI patients [105,106,107], or in fibroblast cell culture adjuvating wound healing in several chronic skin lesions [108,109,110].
MCs and mast cell-committed progenitors (MCPs) recruited in the ischemic cardiac region following inflammatory triggers seem to come mostly from white adipose tissue (WAT), where stem cells with multilineage properties are contained. Once infiltrated in cardiac tissue, WAT-derived MCPs proliferate and differentiate into mature cells in an SCF-dependent manner, they can therefore directly regulate CMs contractility as well as produce cytokines for monocyte recruitment [111].
Among the early mediators of tissue injury, IL-6 produced by CMs and by recruited myeloid cells plays a pivotal role by upregulating intercellular adhesion molecule 1 (ICAM-1) on CMs leading to neutrophil binding and stimulation of their cytotoxic activity [94]. IL-6 and TNF-α contribute to the expression and production of the chemokine CCL2 [112] by mainly ECs and infiltrating leukocytes, with the function of recruiting, activating, and differentiating monocytes and macrophages expressing the CCR2 [113]. Suppression or reduction of this chemokine appears to attenuate adverse remodeling following MI [114]. Some adhesion molecules, such as vascular cell adhesion molecule 1 (VCAM-1), contribute to monocyte recruitment by binding the integrin very late antigen-4 (VLA-4) on their cell surface. Pro-inflammatory Ly-6Chigh monocytes, also defined as M1-type, dominate on days 1 to 4 and promote phagocytosis of apoptotic and necrotic myocytes and neutrophils and other debris, release pro-inflammatory mediators such as TNF-α, IL-1b, IL-6 and macrophage inflammatory protein 1 alpha (MIP-1α), and molecules able to digest preexisting collagen network, like myeloperoxidase, MMPs, cathepsins, and plasminogen activator urokinase [115]. In the early stages following ischemia, monocyte macrophages also begin to produce macrophage-colony stimulating factor (M-CSF), which stimulates phenotypic changes of monocytes proliferation and differentiation into mature macrophages, in an autocrine mechanism, responsible for macrophages growth and activation [116]. It is therefore possible to state that cardiac tissue-resident macrophages drive monocyte recruitment, fate specification, inflammation, and adverse left ventricular (LV) remodeling [10]. Interestingly, FXIIIA is contained in macrophages and platelets, linking in turn coagulation, thrombosis, inflammation, and healing processes [92,117].
Similarly, cardiac resident fibroblasts presenting TLRs can easily respond to DAMPs produced by CMs necrosis, generating granulocyte-macrophage colony-stimulating factor (GM-CSF) involved in the production of neutrophils and monocytes, which follow chemotactic signals, such as CCL2 (for CCR2+ monocytes) and CXCL2 (for neutrophils), up to the infarcted myocardium [118]. Fibroblasts may modify their functions following myocardial insults with changes in the secretion of cytokines and growth factors acting in a paracrine and/or autocrine manner altering protein expression, cell proliferation and cell migration. In all post-ischemia stages, fibroblasts activated by TGF-β and other factors, control changes in the ECM. The phenotypic conversion into myofibroblasts, obtained following activation, leads to increased fibrin production, and persistent myofibroblast proliferation leads to cardiac remodeling [2].
Inflammation results, therefore, as a key pathophysiologic process of MI, in common with other CV diseases, and is potentially involved in disease progression, severity and clinical outcome [7,8,119,120,121]. At this point, it is natural to wonder what mechanisms initiate the resolution of the infarct inflammation since temporal and spatial containment of inflammation is a prerequisite for arrayed wound healing and prevention of adverse ventricular remodeling. In recent years, much attention has been paid to the role of regulatory T cells (Treg) as effectors of the switch between the inflammation phase and the resolution phase of MI [122,123,124,125,126].

2.2. Epigenetic Regulation

The inflammatory response to MI is in part regulated by at least one active histone modification, such as H3K9ac, H3K27ac and H3K4me3. These three marks, mainly present at the promoter level, regulate the expression of chemokines essential in the early stages of MI, such as CCL2, or of proteins involved in the development and differentiation of CMs [127]. CMs survival is also promoted by histone acetylation stimulated by acetyl-CoA synthesis from sodium octanoate (8C) metabolism [128]. HDACs are also involved in the first post-MI phase, such as Sirtuin 2 (SIRT2), a class III HDAC, which is able of deacetylating both histones and numerous non-histone proteins, such as DNA transcription and repair factors [129,130]. It has been reported that SIRT1 is involved in endothelium-dependent vasodilation, via the deacetylation of lysines 496 and 506 of the calmodulin-binding domain of the endothelial nitric oxide synthase (eNOS), leading to enzyme-increased activation and increased endothelial nitric oxide [131]. The latter, unlike its vasodilator effect, has negative results, such as S-nitrosylation and therefore inhibition of SIRT1 [132], which instead seems to possess cardioprotective effects. Sirtuin 1 indeed increases the expression of molecules such as MnSOD, TrX and Bcl-xL while reducing that of pro-apoptotic factors such as Bax, which is positively regulated by p53, which in turn is deacetylated and then inhibited by SIRT1. Sirtuin 1 likewise deacetylates FoxO by stimulating the expression of cell-protective genes [133].
Nonetheless, the most studied epigenetic factors regulating post-MI mechanisms are miRNAs, which mainly act in a synergistic or antagonistic way on the key processes occurring after ischemic damage. Figure 5 and Table S1 show the main miRNAs that are presumably involved in the inflammatory phase stratified according to the mechanism they regulate: (i) cardiomyocyte apoptosis; (ii) inflammation; (iii) fibrosis.
In the first case, the miRNAs are pro- and anti-apoptotic, based on their ability to inhibit genes activating apoptosis, such as miR-133a [134], or genes involved in anti-apoptotic processes, such as miR-124 [135]. Many of these miRNAs show overlapping effects, although they are included in different signaling pathways. For example, miR-34a which blocks the expression of aldehyde dehydrogenase 2 (ALDH2), involved in alcohol metabolism [136], and miR-429 which downregulates the transmembrane receptor Notch1 [137]. In addition, a miRNA can target different proteins with the same result, as for the anti-apoptotic miR-101 [138,139].
In the case of miRNAs that can regulate inflammation and oxidative stress, most of them appear to be involved in the suppression of histone deacetylase SIRT1 (miR-29a [140], miR-132 [141], miR-155 [142]), which is downregulated following I/R damage, with consequent reduction of its cardioprotective effect.
Finally, the activation of CFs and their conversion into myofibroblasts is an important aspect of the inflammatory phase, on which the progression of myocardial damage or its resolution will depend. Most miRNAs seem to have pro-fibrotic effects, which allow for the proliferation, migration, and differentiation of CFs, as well as the production of collagen and ECM, except four with anti-fibrotic function: miR-130a [143], miR-133a [144], miR-148b [145] and miR-590-3p [146]. However, the main pathway regulated by pro- and anti-fibrotic miRNAs is TGF-β signaling and its receptors such as TGFbR and Smads, apart from miR-21 [22], miR-92a [147] and miR-195 [148] which all inhibit Smad7. In turn, TGF-β activates other pathways such as PI3K/AK and JAK- STAT signaling [149], which in turn are targeted by different miRNAs involved in heart healing.

3. Resolution Phase

The suppressor Treg cell activation, directly promoted by the heart and its draining lymphonodes, alleviates local inflammation, protects CMs from apoptosis, enhances M2-like monocyte differentiation, which promotes wound healing, angiogenesis, fibrotic, and scavenger processes, and modules myofibroblasts activation [4,122,124,126]. Treg cells produce TGF-β, IL-10, and IL-13, which stimulate anti-inflammatory macrophage activation, repress anti-inflammatory signals, and stimulate neo-angiogenesis. The activated Treg cells also induce an M2-polarizing milieu locally within the heart, to which IL-10 and TGF-β produced by M2-like macrophages themselves contribute, while TGF-β and IL-13 work together to stimulate collagen deposition by myofibroblasts. Collagen production is also influenced by osteopontin, whose release from M2-type monocyte is stimulated by TGF-β, IL-10 and IL-13 [122]. In addition to the production of collagen by myofibroblasts and fibrin deposition, the presence of FXIII, which promotes cross-linking between fibrin and ECM components, is extremely important for the integrity and elasticity properties of the scar tissue. FXIII is also essential for other important tasks, such as the recruitment of adult stem cells and neo-angiogenesis by inhibition of the potent anti-angiogenic factor thrombospondin-1 (TSP-1) [150,151], so that dynamic of FXIIIA during the first hours after MI can be considered a prognostic indicator of the infarct evolution [104,122].
Treg lymphocytes have numerous influences on the cells present in the injured myocardium, which allow for the resolution of inflammation and the repair of ischemic damage. In fact, they inhibit cytokine production by fibroblasts; promote angiogenesis by acting on ECs; slow down cytokine production and apoptosis as well as promote the proliferation of CMs; inhibit the production of cytokines by Th1/Th17/CD8+ T lymphocytes and of cytokines and antibodies by B cells, as well as T cells migration and B cells proliferation [125].
In addition to the presence and action of Treg lymphocytes, an essential step for the transition to the resolution phase is the removal of debris produced by the degradation of the ECM and apoptotic cells, including CMs and neutrophils that are programmed to undergo apoptosis [87]. They are phagocytosed by macrophages, leading to the release of anti-inflammatory cytokines as well as anti-inflammatory and pro-resolving lipid mediators as lipoxin A4, resolving E1 and protectin D1. The latter contribute to abrogate neutrophil influx in the injured myocardium through the production of lactoferrin, an anti-inflammatory glycoprotein that specifically inhibits chemotaxis of neutrophils but not mononuclear phagocytes [152]. Finally, phagocytic clearance of apoptotic neutrophils reprogrammes monocyte-derived macrophages from a pro-inflammatory (M1-type) to an anti-inflammatory (M2-type) phenotype [153].
M2-type macrophages are widely present in infarcted myocardium and thanks to the secretion of anti-proliferative cytokines and mediators that terminate pro-inflammatory IL-1 signaling, such as the decoy type II IL-1 receptor and IL-1 receptor antagonist (IRA), they can be considered among the undisputed protagonists of the resolution phase [87]. Furthermore, local upregulation of M-CSF creates a microenvironment suitable for macrophage growth, differentiation, and survival. M-CSF can also promote cardiac repair by modulating the endothelial cell phenotype, stimulating the production of the monocyte chemoattractant protein-1 (MCP-1), which, in addition to inducing monocyte chemotaxis, promotes angiogenesis. This may be due either to the induction of production of growth factors such as vascular-endothelial growth factor (VEGF) or by directly acting on proliferation of the ECs [116,154].
Among the anti-inflammatory cytokines produced by macrophages and Treg lymphocytes, TGF-β plays an important role since it attenuates trans-endothelial leukocyte migration, suppresses the production of pro-inflammatory chemokines and cytokines, enhances M-CSF-induced proliferation, inhibits nitrite release, reduces cytotoxic activity, and regulates the phenotype and functions of the T lymphocyte subpopulations [155]. TGF-β also stimulates massive proliferation of myofibroblasts, which become the predominant cell type in healing infarct, as well as the preservation of the ECM, through the production of ECM proteins, such as collagen I, collagen III and fibronectin, the suppression of collagenase expression and the induction of metalloproteinase inhibitor (TIMP-1) secretion [155,156].
During the resolution phase, the ECM is enriched by the presence of matricellular proteins, a family of structurally unrelated extracellular macromolecules that interact with surface receptors, growth factors, proteases, but also with structural matrix proteins, without having a direct structural role. They act as bridges between cells and matrix, integrating signals that modulate cellular behaviour. Matricellular proteins, including TSP-1 and TSP-2, SPARC (secreted protein acidic and rich in cysteine), tenascin-C, osteopontin, and periostin, are usually poorly expressed in the healthy heart, while they are strongly upregulated following cardiac injury, where they perform the function of transducers of key molecular signals in cardiac repair and modulators of cell migration, proliferation, and adhesion. They also modulate cytokine and growth factor signaling, regulate matrix assembly and metabolism, and modulate the fibrogenic potential of inflammatory cells and fibroblasts [157]. The latter contribute to the matrix metabolism, producing cytokines, growth factors, matricellular proteins and MMPs [155].
ECs make up the largest percentage of non-cardiomyocytic cells in the heart [158] and play important roles in the various stages following ischemic damage, restoring oxygen and nutrient supply [159]. Neo-angiogenesis is an integral and essential part of wound healing and, already a few hours after ischemic damage, growth factors such as basic fibroblast growth factor (bFGF) and VEGF stimulate the angiogenic activation of ECs and the formation of new vessels. During the proliferative phase, enlarged vessels are formed, with thin walls poor in pericytes, called “mother vessels”, with transluminal projections capable of dividing the vascular structure into multiple lumens. The absence of pericytes makes these vessels more plastic and therefore capable of adapting to continuous tissue changes. Over time, mother vessels evolve into smaller daughter vessels, such as medium-sized muscular arteries and veins [160]. Many other factors, such as TGF-β, MCP-1, IL-8 and IL-10 are involved in the regulation of angiogenesis and may participate in the formation of new vessels after cardiac infarction. The composition of the ECM is also critical for a dynamic modulation of vascular growth, especially the presence of collagen [87,156]. The key role of cytokines/chemokines/growth factors on endothelial homeostasis and tissue repair suggests that they may be targeted for therapeutic purposes as demonstrated in other complex vascular diseases [161,162].

Epigenetics Regulation

Since myofibroblasts are primarily responsible for normal wound healing and tissue repair after MI, it is very important that they are highly regulated to prevent their excessive activation, which could lead to pathological fibrosis, or reduced ECM deposition which can cause poor repair or detrimental myocardial wall rupture. On the other hand, excessive activity or long-lasting of fibrotic constituents, may result in a permanent damage of the reparative tissue with loss in efficiency of the elastic components as demonstrated in other tissue modelling [163]. As a result, molecule involved in the crosslinking of ECM components, as F13A1 gene product, is predicted to be targeted by at least 133 different miRNAs as reported in the miRDB database (http://www.mirdb.org, accessed on 26 July 2023) suggesting a complex and articulate epigenetic regulation and epi-druggable target.
Moreover, hyperacetylation of histone H4 at the collagen type 1-alpha-2 locus by p300 facilitates TGF-β-induced collagen production by myofibroblasts. Conversely, HDAC4-mediated deacetylation is associated with TGF-β-mediated suppression of collagen production and increased expression of repressors of the TGF-β1 signaling pathway [164]. HDACs are generally upregulated during fibroblast activation, a hypothesis supported by the beneficial effects on remodelling of deacetylase inhibitors [165]. Class I and II HDACs seem to be largely involved in the induction of cardiac hypertrophy, through the repression of antihypertrophic genes, such as Krupple-like factor 2 (Klf2) [166]. HDAC2 seems to induce cardiac hypertrophy by increasing the Akt signaling pathway, which is involved both in cardiac hypertrophy and in the maintenance or improvement of cardiac functions [167,168].
Class III HDACs have opposite protective effects towards CMs as SIRT1, whose expression is upregulated in stressful situations such as inflammation, inhibiting both hypertrophy and apoptosis following oxidative stress [169]. Indeed, it has been demonstrated that the production of ROS is an essential mechanism for the development of cardiac hypertrophy and that SIRT3 is able to control its accumulation, consequently blocking the activation of pathways, such as MAPK/ERK and PI3K/AkT, involved in cardiac hypertrophy [170]. SIRT6 in turn, also acts as a negative regulator of cardiac hypertrophy reducing the differentiation of CFs into myofibroblasts, through suppression of the transcriptional activity of NF-kB, a key regulator of the inflammatory reaction involved in the development of cardiac remodeling [171]. DNA methylation is also involved in the regulation of MI effects, particularly in cardiac fibrosis. In fact, it has been seen that in hypoxic conditions CFs express hypoxia-inducible factor 1-alpha (HIF-1α) which in turn determines the upregulation of the DNMT1 and DNMT3a/3b genes with consequent alteration of DNA methylation and activation of pro-fibrotic genes [165].
HIF-1α is also involved in angiogenesis and is negatively modulated by miR-126 [172] (Figure 6 and Table S2), although miRNAs that appear to regulate new blood vessel formation following MI preferentially target VEGF and its receptor VEGFR, such as miR-129-1, miR-133, miR-139-5p, miR-199b and miR-200a-3p [173,174,175,176]. In addition, some miRNAs act on the proliferation and migration of ECs, regulating the PI3K/Akt signaling pathway (miR-130a and miR-208 [177,178]).
Even in the resolution phase, regulation of cardiac fibrosis is important to prevent adverse consequences such as LV remodeling and heart failure [7,8]. As conceivable, miRNAs likely to act on fibrosis at this stage mainly regulate the TGF-β signaling pathway or directly the expression of collagen genes in myofibroblasts or activated CFs. For example, miR-26a appears to directly target collagen type 1 (COL1) and connective tissue growth factor (CTGF) [179], while miR29b [180] and miR-208 [181] inhibit COL1 expression, by downregulating adapter protein SH2B3 and transcription factor GATA4, respectively. However, miR-29b is also able to directly target the expression of fibrosis-related genes, such as COL1A1 and COL3A1 [182].
A very important aspect of the resolution phase is the polarization of macrophages from the pro-inflammatory to the anti-inflammatory phenotype. This step is essential to limit the progression of inflammation and proceed towards its resolution and although its regulation largely depends on the production of cytokines such as TGF-beta, IL-10, IL-13 and M-CSF [116,122,154], in recent years several studies on non-coding RNAs have been performed.
Although these investigations are still in their infancy, most of the miRNAs analyzed appear to come from exosomes released from mesenchymal cells and CMs and seem to act primarily by regulating the PI3k/Akt signaling pathway (such as miR-21-5p, miR-150 and miR-182), or the NF-kB pathway (namely miR-24-3p) [183,184,185,186]. Unfortunately, there are few miRNAs that can stimulate the re-entry of CMs into the cell cycle, allowing their proliferation and differentiation, with consequent repopulation of the damaged region and these include miR-19a, miR-19b and miR-106a-363 cluster [34,187].
Among the various miRNAs studied, one seems to regulate several aspects of the resolution phase, even if it targets only one pathway: miR-375 regulates the PDK-1/Akt signaling, promoting neovascularization, reducing cardiomyocyte apoptosis and inducing the switch from M1- to M2-macrophage, all actions that limit the expansion of the infarct area and the dysfunction of the left ventricle [188].

4. Maturation Phase

The formation of a complex network of ECM proteins and the deposition/cross-linking of collagen molecules [189] determine the end of the proliferative phase and the beginning of the maturation phase, with consequent production of a stable scar [190]. Different data have been reported concerning the fate of myofibroblasts during this phase. Initially, it was thought that the latter undergo some mechanisms which determined their loss at the end of the physiological tissue repair: cell release from stress, the increase of cross-linking between matrix molecules and an increased formation of specific cell-cell contacts, up to their death by apoptosis [191]. Recently, it has been shown that both activated fibroblasts and myofibroblasts persist at a 3.5-fold higher concentration in the infarcted region than in a healthy heart, over a long period of time without cell turnover. The scar that forms in this last phase must persist in the long term due to the impossibility of the adult heart to regenerate and hosts highly differentiated fibroblasts adapted to a deeply fibrotic environment, matrifibrocytes, which have the task of maintaining the integrity of the mature scar over time [192].
Another important change that occurs during maturation is the coating of new blood vessels with mural cells, such as pericytes and vascular smooth muscle cells (VSMC), which stabilize the microvasculature of the scar and attenuate inflammation, preventing extravasation of blood cells, while uncoated vessels regress [190,193]. The regulation of the two major events associated with scar formation, collagen deposition and vasculature maturation, is largely dependent on platelet-derived growth factor (PDGF) signaling. Both PDGFR-α and PDGFR-β receptors mediate fibrogenic effects, but only the latter is involved in the regulation of microvessel coating [194]. PDGFR-β is expressed by pericytes and vascular smooth muscle cells (VSMCs), while ECs secrete PDGF, which mediates mural cell proliferation and migration [195]. Alteration of this signaling pathway can lead to continued inflammation and haemorrhage and adverse cardiac remodeling.

Epigenetics Regulation

As already highlighted in Figure 3, the Notch pathway can stimulate the proliferation of CMs in the pre-natal period, while it is blocked shortly after birth and cannot be reactivated following MI due to epigenetic alterations at the level of its promoter. This region is in fact modified with the H3K27Me3 mark, associated with chromatin condensation and gene silencing [196]. Also, histone acetylation, as already stated, and CpG island methylation are involved in the silencing of genes that allow the reactivation and proliferation of CMs and could therefore be considered potential therapeutic target to improve the outcomes of MI [197].
The miRNAs presumably involved in the maturation phase mainly regulate cardiomyocyte hypertrophy, vascular smooth muscle cell differentiation, myofibroblast senescence and, more generally, cardiac dysfunction, remodelling and damage (Figure 7 and Table S3).
Is it possible that the mechanisms and processes that took place in the two previous phases have already decided the fate of the infarcted myocardium? In 2010, Nahrendorf had proposed that the very first days immediately following the heart ischemic injury determine the fate of the patient [115]. This view has been confirmed by contemporary CV therapy showing that it is possible to reverse the LV remodelling with a consequent improvement in myocardial function, up to a total recovery in rare cases, but only if timely intervention is performed after MI establishment [198].

5. Sex-Related Differences in Risk Factors and Long-Term Outcomes for MI

Males and females have marked differences in the risk factors, symptoms, and outcomes of MI. The most well-known cardiovascular risk factors appear to be differently associated with a greater risk of developing MI in the two sexes [86].
Smoking, for example, is associated with higher risk in women and increases with the number of cigarettes consumed. Type 1 and type 2 diabetes, as well as atrial fibrillation, also lead to an increased risk of ischemic damage in females, although the incidence of MI is less than half that in males [199,200]. Unfortunately, females have additional sex- and gender-specific risks, including early menopause and pregnancy-related disorders [201,202,203]. In postmenopausal females, specific independent predictor for increased risk of MI have been recognized in MMP mediated collagen degradation products [204]. Females with a history of gestational hypertension and preeclampsia have a 60% greater risk of developing CV diseases, and this persists into their 70s [205]. Increased blood glucose content during pregnancy, even without the development of gestational diabetes, also has a strong impact on the risk of MI [206]. Moreover, regarding the symptoms of MI there are significant differences between the two sexes reporting a greater number of symptomatic phenotypes and some of these not specific for MI [207]. Although females report chest pain more often and earlier after ischemia than males, it is usually a non-specific pain that is unlikely to be related to ischemic damage [208]. In addition to the difficulty in recognizing symptoms (by both females themselves and by health care specialists), other social aspects may contribute to a marked delay in presentation and revascularization in females with acute MI. Among these, the deep-rooted idea that female sex is rarely affected by ischemia and the women social role which relegates women to providing assistance rather than receiving it [201].
Post-MI outcomes and complications are also influenced by sex differences. First, females experience longer hospitalizations, more bleeding complications, and higher in-hospital mortality following coronary revascularization than the sex counterpart. Furthermore, females have an increased risk of early death, complications such as heart failure and cardiogenic shock, and recurrent MIs. The latter could be related to several factors. Compared to men, women are more likely to have non-obstructive atherosclerotic coronary artery disease or spontaneous coronary artery dissection, a different coronary physiology with increased plaque vulnerability, a smaller coronary luminal area regardless of body size, with larger susceptibility to thrombotic occlusion, and worse outcomes after revascularization [209].
Is it possible that these anatomical differences alone underlie the differences in risk factors, symptoms, and outcomes between females and males? Might there be an extensive different epigenetic regulation of the various stages of MI involved in the two sexes? Epigenetic modifications of DNA and histones, as well as non-coding RNAs, have been studied mainly in cells and animal models, so is it possible that there are substantial differences in sex-related epigenetic regulations? And how could this affect the development of targeted efficient therapeutic strategies?
Overall, the number and the complexity of the clinical unmet needs, ask for appropriate and dedicated answers, and this is worthily approached by the international scientific community in several complex diseases by considering both genetics and epigenetics aspects [86,210,211,212].

6. Conclusions

In the present review, the main post-MI processes and the epigenetic factors that govern them have been reported, to present a global idea of the heart healing process and its regulation. Epigenetic control mechanisms can have synergistic, additive, or antagonistic effects, and the regulation of a single physiological or pathological phenomenon is therefore very intricate to decipher. Even in the case of the epigenetic regulation of pre- and post-MI cardiac processes, the combined action of many different factors makes it extremely difficult to intervene in order to have a significant effect on the course of symptoms. Just think of the number of all the miRNAs involved in the proliferation of CMs or in the regulation of cardiac fibrosis.
Many research groups are currently concentrating on identifying molecules that can either block or mimic the effects of miRNAs, which are being considered as potential therapeutic markers. However, there have been no studies that have progressed beyond animal testing. Moreover, in some cases, the effects of these non-coding RNAs have been known for a decade now, though dedicated clinical trials on humans have not yet been started. Interestingly, the three anti-miRs that were in the pipeline by miRagen Therapeutics in 2017 (MGN-4220 against miR-29 for the treatment of cardiac fibrosis, MGN-1374 able to inhibit miR-15 and miR-195 for the treatment of post-MI cardiac remodeling, and MGN-9103 which targets miR-208, for the treatment of chronic heart failure) [213], are interesting candidates to test. Certainly, one of the main obstacles is the difficulty in the delivery of RNAs, easily degraded by RNases present in body fluids, but on which significant goals are being achieved [214]. Another difficult problem to deal with is the complex regulation of the various myocardium processes. It is not only important to understand the function of epigenetic factors and verify their up or down regulation following MI, but also to verify how they vary in subjects who experienced MI with different outcomes. Several important questions remain to understand what exactly allows one individual to completely recover after MI compared to the majority of patients. What changes between subjects who have undergone fibrosis or LV remodeling? The comparison of the miRNA expression profiles, or other epigenetic factors will allow us to identify key element in heart healing. Precision medicine fits right into this context, being able to verify in the period immediately following the ischemic damage how miRNAs vary, allowing for effective intervention.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241713268/s1. References [215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253] are cited in the supplementary materials.

Author Contributions

Conceptualization, D.G., F.S. and V.T.; Original draft preparation and writing, F.S., D.G. and V.T.; Figure preparation, F.S., E.D. and M.L.S.; Review and editing, D.G., V.T., F.S., E.D., M.L.S., A.V.S., P.S. and G.Z. All authors have read and agreed to the published version of the manuscript.

Funding

FAR and FIRD grants from University of Ferrara (D.G. and V.T.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the National Recovery and Resilience Plan (NRRP), Mission 04 Component 2 Investment 1.5—NextGenerationEU, Call for tender n. 3277 dated 30/12/2021; Award Number: 0001052 dated 23 June 2022.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

ALDH2Aldehyde Dehydrogenase 2
bFGFBasic Fibroblast Growth Factor
CCL2Monocyte Chemoattractant Protein-1 (CCL2/MCP1)
CCR2CC chemokine Receptor 2
CFsCardiac Fibroblasts
Check1Checkpoint Kinase 1
CMsCardiomyocytes
COL1Collagen Type 1
CTGFConnective Tissue Growth Factor
CVCardiovascular
DAMPsDamage-Associated Molecular Patterns
DCsDendritic Cells
DMRsDifferently Methylated Regions
ECMExtracellular Matrix
ECsEndothelial Cells
eNOSEndothelial Nitric Oxide Synthase
ESCsEmbryonic Stem Cells
FGF9Fibroblast Growth Factor-9
FPR1Formyl Peptide Receptor 1
GM-CSFGranulocyte-Macrophage Colony-Stimulating Factor
HDACsHistone deacetylases
HIF-1αHypoxia-Inducible Factor 1-alpha
I/RIschemia/Reperfusion
ICAM-1Intercellular Adhesion Molecule 1
IGF2Insulin-like Growth Factor-2
IRAIL-1 Receptor Antagonist
Klf2Krupple-like factor 2
lncRNAsLong Non-coding RNAs
LVLeft Ventricular
MCP-1Monocyte Chemoattractant Protein-1
MCPsMast Cell-Committed Progenitors
MCsMast Cells
M-CSFMacrophage-Colony Stimulating Factor
MCTCTrypase positive, chymase positive mast cells
MIMyocardial Infarction
MIP-1αMacrophage Inflammatory Protein 1 Alpha
miRNAsMicroRNAs
MMPsMetalloproteinases
ncRNAsNon-coding RNAs
NKIRAS2NF-kB Inhibitor Interacting Ras like 2
NOMO1Nodal Modulator 1
PDGFPlatelet-Derived Growth Factor
PI3KPhosphoinositide 3-kinase
PRRPathogen Recognition Receptors
ROSReactive Oxygen Species
SIRTSirtuin
SOCS3Suppressor Of Cytokine Signaling 3
SPARCSecreted Protein Acidic and Rich in Cysteine
SPIN1Spindlin-1
TGF-βTransforming Growth Factor-beta
TIMP-1Tissue metalloproteinase inhibitor-1
TLRsToll-Like Receptors
TNF-αTumor Necrosis Factor-alpha
TregRegulatory T cells
TSP-1Thrombospondin-1
TSP-2Thrombospondin-2
VCAM-1Vascular Cell Adhesion Molecule 1
VEGFVascular-Endothelial Growth Factor
VLA-4Very Late Antigen-4
VSMCsVascular Smooth Muscle Cells
WATWhite Adipose Tissue

References

  1. Tucker, N.R.; Chaffin, M.; Fleming, S.J.; Hall, A.W.; Parsons, V.A.; Bedi, K.C., Jr.; Akkad, A.D.; Herndon, C.N.; Arduini, A.; Papangeli, I.; et al. Transcriptional and Cellular Diversity of the Human Heart. Circulation 2020, 142, 466–482. [Google Scholar] [CrossRef] [PubMed]
  2. Dostal, D.; Glaser, S.; Baudino, T.A. Cardiac fibroblast physiology and pathology. Compr. Physiol. 2015, 5, 887–909. [Google Scholar] [CrossRef]
  3. Sun, K.; Li, Y.Y.; Jin, J. A double-edged sword of immuno-microenvironment in cardiac homeostasis and injury repair. Signal Transduct. Target. Ther. 2021, 6, 79. [Google Scholar] [CrossRef] [PubMed]
  4. Koc, A.; Cagavi, E. Cardiac Immunology: A New Era for Immune Cells in the Heart. Adv. Exp. Med. Biol. 2021, 1312, 75–95. [Google Scholar] [CrossRef]
  5. Janicki, J.S.; Brower, G.L.; Levick, S.P. The emerging prominence of the cardiac mast cell as a potent mediator of adverse myocardial remodeling. Methods Mol. Biol. 2015, 1220, 121–139. [Google Scholar] [CrossRef] [PubMed]
  6. Swirski, F.K.; Nahrendorf, M. Cardioimmunology: The immune system in cardiac homeostasis and disease. Nat. Rev. Immunol. 2018, 18, 733–744. [Google Scholar] [CrossRef]
  7. Kumar, V.; Rosenzweig, R.; Asalla, S.; Nehra, S.; Prabhu, S.D.; Bansal, S.S. TNFR1 Contributes to Activation-Induced Cell Death of Pathological CD4(+) T Lymphocytes During Ischemic Heart Failure. JACC Basic. Transl. Sci. 2022, 7, 1038–1049. [Google Scholar] [CrossRef]
  8. Kumar, V.; Prabhu, S.D.; Bansal, S.S. CD4(+) T-lymphocytes exhibit biphasic kinetics post-myocardial infarction. Front. Cardiovasc. Med. 2022, 9, 992653. [Google Scholar] [CrossRef]
  9. Alvarez-Argote, S.; O’Meara, C.C. The Evolving Roles of Cardiac Macrophages in Homeostasis, Regeneration, and Repair. Int. J. Mol. Sci. 2021, 22, 7923. [Google Scholar] [CrossRef]
  10. Bajpai, G.; Schneider, C.; Wong, N.; Bredemeyer, A.; Hulsmans, M.; Nahrendorf, M.; Epelman, S.; Kreisel, D.; Liu, Y.; Itoh, A.; et al. The human heart contains distinct macrophage subsets with divergent origins and functions. Nat. Med. 2018, 24, 1234–1245. [Google Scholar] [CrossRef]
  11. Molawi, K.; Wolf, Y.; Kandalla, P.K.; Favret, J.; Hagemeyer, N.; Frenzel, K.; Pinto, A.R.; Klapproth, K.; Henri, S.; Malissen, B.; et al. Progressive replacement of embryo-derived cardiac macrophages with age. J. Exp. Med. 2014, 211, 2151–2158. [Google Scholar] [CrossRef]
  12. Epelman, S.; Lavine, K.J.; Beaudin, A.E.; Sojka, D.K.; Carrero, J.A.; Calderon, B.; Brija, T.; Gautier, E.L.; Ivanov, S.; Satpathy, A.T.; et al. Embryonic and adult-derived resident cardiac macrophages are maintained through distinct mechanisms at steady state and during inflammation. Immunity 2014, 40, 91–104. [Google Scholar] [CrossRef]
  13. Dick, S.A.; Macklin, J.A.; Nejat, S.; Momen, A.; Clemente-Casares, X.; Althagafi, M.G.; Chen, J.; Kantores, C.; Hosseinzadeh, S.; Aronoff, L.; et al. Self-renewing resident cardiac macrophages limit adverse remodeling following myocardial infarction. Nat. Immunol. 2019, 20, 29–39. [Google Scholar] [CrossRef]
  14. Ma, Y.; Mouton, A.J.; Lindsey, M.L. Cardiac macrophage biology in the steady-state heart, the aging heart, and following myocardial infarction. Transl. Res. 2018, 191, 15–28. [Google Scholar] [CrossRef]
  15. Leid, J.; Carrelha, J.; Boukarabila, H.; Epelman, S.; Jacobsen, S.E.; Lavine, K.J. Primitive Embryonic Macrophages are Required for Coronary Development and Maturation. Circ. Res. 2016, 118, 1498–1511. [Google Scholar] [CrossRef] [PubMed]
  16. Fantin, A.; Vieira, J.M.; Gestri, G.; Denti, L.; Schwarz, Q.; Prykhozhij, S.; Peri, F.; Wilson, S.W.; Ruhrberg, C. Tissue macrophages act as cellular chaperones for vascular anastomosis downstream of VEGF-mediated endothelial tip cell induction. Blood 2010, 116, 829–840. [Google Scholar] [CrossRef]
  17. Cahill, T.J.; Sun, X.; Ravaud, C.; Villa Del Campo, C.; Klaourakis, K.; Lupu, I.E.; Lord, A.M.; Browne, C.; Jacobsen, S.E.W.; Greaves, D.R.; et al. Tissue-resident macrophages regulate lymphatic vessel growth and patterning in the developing heart. Development 2021, 148, dev194563. [Google Scholar] [CrossRef] [PubMed]
  18. Hulsmans, M.; Clauss, S.; Xiao, L.; Aguirre, A.D.; King, K.R.; Hanley, A.; Hucker, W.J.; Wulfers, E.M.; Seemann, G.; Courties, G.; et al. Macrophages Facilitate Electrical Conduction in the Heart. Cell 2017, 169, 510–522. [Google Scholar] [CrossRef]
  19. Bergmann, O.; Bhardwaj, R.D.; Bernard, S.; Zdunek, S.; Barnabe-Heider, F.; Walsh, S.; Zupicich, J.; Alkass, K.; Buchholz, B.A.; Druid, H.; et al. Evidence for cardiomyocyte renewal in humans. Science 2009, 324, 98–102. [Google Scholar] [CrossRef]
  20. Gilsbach, R.; Preissl, S.; Gruning, B.A.; Schnick, T.; Burger, L.; Benes, V.; Wurch, A.; Bonisch, U.; Gunther, S.; Backofen, R.; et al. Dynamic DNA methylation orchestrates cardiomyocyte development, maturation and disease. Nat. Commun. 2014, 5, 5288. [Google Scholar] [CrossRef] [PubMed]
  21. May, D.; Blow, M.J.; Kaplan, T.; McCulley, D.J.; Jensen, B.C.; Akiyama, J.A.; Holt, A.; Plajzer-Frick, I.; Shoukry, M.; Wright, C.; et al. Large-scale discovery of enhancers from human heart tissue. Nat. Genet. 2011, 44, 89–93. [Google Scholar] [CrossRef]
  22. Yuan, X.; Braun, T. Multimodal Regulation of Cardiac Myocyte Proliferation. Circ. Res. 2017, 121, 293–309. [Google Scholar] [CrossRef]
  23. Ghosh, A.K. Acetyltransferase p300 Is a Putative Epidrug Target for Amelioration of Cellular Aging-Related Cardiovascular Disease. Cells 2021, 10, 2839. [Google Scholar] [CrossRef]
  24. Chew, N.W.S.; Loong, S.S.E.; Foo, R. Epigenetics in cardiovascular health and disease. Prog. Mol. Biol. Transl. Sci. 2023, 197, 105–134. [Google Scholar] [CrossRef]
  25. Gorica, E.; Mohammed, S.A.; Ambrosini, S.; Calderone, V.; Costantino, S.; Paneni, F. Epi-Drugs in Heart Failure. Front. Cardiovasc. Med. 2022, 9, 923014. [Google Scholar] [CrossRef] [PubMed]
  26. 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]
  27. 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] [PubMed]
  28. Kansakar, U.; Varzideh, F.; Mone, P.; Jankauskas, S.S.; Santulli, G. Functional Role of microRNAs in Regulating Cardiomyocyte Death. Cells 2022, 11, 983. [Google Scholar] [CrossRef]
  29. Li, X.; Wang, J.; Jia, Z.; Cui, Q.; Zhang, C.; Wang, W.; Chen, P.; Ma, K.; Zhou, C. MiR-499 regulates cell proliferation and apoptosis during late-stage cardiac differentiation via Sox6 and cyclin D1. PLoS ONE 2013, 8, e74504. [Google Scholar] [CrossRef] [PubMed]
  30. Porrello, E.R.; Johnson, B.A.; Aurora, A.B.; Simpson, E.; Nam, Y.J.; Matkovich, S.J.; Dorn, G.W., 2nd; van Rooij, E.; Olson, E.N. MiR-15 family regulates postnatal mitotic arrest of cardiomyocytes. Circ. Res. 2011, 109, 670–679. [Google Scholar] [CrossRef]
  31. Porrello, E.R.; Mahmoud, A.I.; Simpson, E.; Johnson, B.A.; Grinsfelder, D.; Canseco, D.; Mammen, P.P.; Rothermel, B.A.; Olson, E.N.; Sadek, H.A. Regulation of neonatal and adult mammalian heart regeneration by the miR-15 family. Proc. Natl. Acad. Sci. USA 2013, 110, 187–192. [Google Scholar] [CrossRef]
  32. Patil, M.; Pabla, N.; Dong, Z. Checkpoint kinase 1 in DNA damage response and cell cycle regulation. Cell Mol. Life Sci. 2013, 70, 4009–4021. [Google Scholar] [CrossRef]
  33. Chen, J.; Huang, Z.P.; Seok, H.Y.; Ding, J.; Kataoka, M.; Zhang, Z.; Hu, X.; Wang, G.; Lin, Z.; Wang, S.; et al. mir-17-92 cluster is required for and sufficient to induce cardiomyocyte proliferation in postnatal and adult hearts. Circ. Res. 2013, 112, 1557–1566. [Google Scholar] [CrossRef]
  34. 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]
  35. Sun, L.; Zhu, W.; Zhao, P.; Zhang, J.; Lu, Y.; Zhu, Y.; Zhao, W.; Liu, Y.; Chen, Q.; Zhang, F. Down-Regulated Exosomal MicroRNA-221-3p Derived From Senescent Mesenchymal Stem Cells Impairs Heart Repair. Front. Cell Dev. Biol. 2020, 8, 263. [Google Scholar] [CrossRef] [PubMed]
  36. Shi, H.H.; Wang, Z.Q.; Zhang, S. MiR-208a participates with sevoflurane post-conditioning in protecting neonatal rat cardiomyocytes with simulated ischemia-reperfusion injury via PI3K/AKT signaling pathway. Eur. Rev. Med. Pharmacol. Sci. 2020, 24, 943–955. [Google Scholar] [CrossRef]
  37. Chen, Y.; Wu, S.; Zhang, X.S.; Wang, D.M.; Qian, C.Y. MicroRNA-489 promotes cardiomyocyte apoptosis induced by myocardial ischemia-reperfusion injury through inhibiting SPIN1. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 6683–6690. [Google Scholar] [CrossRef] [PubMed]
  38. Han, X.; Chen, X.; Han, J.; Zhong, Y.; Li, Q.; An, Y. MiR-324/SOCS3 Axis Protects Against Hypoxia/Reoxygenation-Induced Cardiomyocyte Injury and Regulates Myocardial Ischemia via TNF/NF-kappaB Signaling Pathway. Int. Heart J. 2020, 61, 1258–1269. [Google Scholar] [CrossRef]
  39. Ding, Y.; Bi, L.; Wang, J. MiR-1180 promotes cardiomyocyte cell cycle re-entry after injury through the NKIRAS2-NFkappaB pathway. Biochem. Cell Biol. 2020, 98, 449–457. [Google Scholar] [CrossRef] [PubMed]
  40. Shen, M.M. Nodal signaling: Developmental roles and regulation. Development 2007, 134, 1023–1034. [Google Scholar] [CrossRef]
  41. Xing, W.; Li, T.; Wang, Y.; Qiang, Y.; Ai, C.; Tang, H. MiR-33a-5p targets NOMO1 to modulate human cardiomyocyte progenitor cells proliferation and differentiation and apoptosis. J. Recept. Signal Transduct. Res. 2021, 41, 476–487. [Google Scholar] [CrossRef] [PubMed]
  42. 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]
  43. Crippa, S.; Nemir, M.; Ounzain, S.; Ibberson, M.; Berthonneche, C.; Sarre, A.; Boisset, G.; Maison, D.; Harshman, K.; Xenarios, I.; et al. Comparative transcriptome profiling of the injured zebrafish and mouse hearts identifies miRNA-dependent repair pathways. Cardiovasc. Res. 2016, 110, 73–84. [Google Scholar] [CrossRef] [PubMed]
  44. Jang, J.; Song, G.; Pettit, S.M.; Li, Q.; Song, X.; Cai, C.L.; Kaushal, S.; Li, D. Epicardial HDAC3 Promotes Myocardial Growth Through a Novel MicroRNA Pathway. Circ. Res. 2022, 131, 151–164. [Google Scholar] [CrossRef]
  45. 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]
  46. Qiao, L.; Hu, S.; Liu, S.; Zhang, H.; Ma, H.; Huang, K.; Li, Z.; Su, T.; Vandergriff, A.; Tang, J.; et al. microRNA-21-5p dysregulation in exosomes derived from heart failure patients impairs regenerative potential. J. Clin. Investig. 2019, 129, 2237–2250. [Google Scholar] [CrossRef]
  47. Gao, J.; Zhu, M.; Liu, R.F.; Zhang, J.S.; Xu, M. Cardiac Hypertrophy is Positively Regulated by MicroRNA-24 in Rats. Chin. Med. J. 2018, 131, 1333–1341. [Google Scholar] [CrossRef]
  48. 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]
  49. Wu, Y.; Fan, Z.; Chen, Z.; Hu, J.; Cui, J.; Liu, Y.; Wang, Y.; Guo, B.; Shen, J.; Xie, L. Astragaloside IV protects human cardiomyocytes from hypoxia/reoxygenation injury by regulating miR-101a. Mol. Cell Biochem. 2020, 470, 41–51. [Google Scholar] [CrossRef]
  50. Liang, C.; Wang, S.; Zhao, L.; Han, Y.; Zhang, M. Effects of miR-145-5p on cardiomyocyte proliferation and apoptosis, GIGYF1 expression and oxidative stress response in rats with myocardial ischemia-reperfusion. Cell. Mol. Biol. 2022, 68, 147–159. [Google Scholar] [CrossRef]
  51. Wang, X.; Ha, T.; Liu, L.; Hu, Y.; Kao, R.; Kalbfleisch, J.; Williams, D.; Li, C. TLR3 Mediates Repair and Regeneration of Damaged Neonatal Heart through Glycolysis Dependent YAP1 Regulated miR-152 Expression. Cell Death Differ. 2018, 25, 966–982. [Google Scholar] [CrossRef]
  52. 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]
  53. Torrini, C.; Cubero, R.J.; Dirkx, E.; Braga, L.; Ali, H.; Prosdocimo, G.; Gutierrez, M.I.; Collesi, C.; Licastro, D.; Zentilin, L.; et al. Common Regulatory Pathways Mediate Activity of MicroRNAs Inducing Cardiomyocyte Proliferation. Cell Rep. 2019, 27, 2759–2771.e2755. [Google Scholar] [CrossRef]
  54. Sun, R.; Zhang, L. Long non-coding RNA MALAT1 regulates cardiomyocytes apoptosis after hypoxia/reperfusion injury via modulating miR-200a-3p/PDCD4 axis. Biomed. Pharmacother. 2019, 111, 1036–1045. [Google Scholar] [CrossRef] [PubMed]
  55. Arif, M.; Pandey, R.; Alam, P.; Jiang, S.; Sadayappan, S.; Paul, A.; Ahmed, R.P.H. MicroRNA-210-mediated proliferation, survival, and angiogenesis promote cardiac repair post myocardial infarction in rodents. J. Mol. Med. 2017, 95, 1369–1385. [Google Scholar] [CrossRef]
  56. 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]
  57. Zhen, L.; Zhao, Q.; Lu, J.; Deng, S.; Xu, Z.; Zhang, L.; Zhang, Y.; Fan, H.; Chen, X.; Liu, Z.; et al. miR-301a-PTEN-AKT Signaling Induces Cardiomyocyte Proliferation and Promotes Cardiac Repair Post-MI. Mol. Ther. Nucleic Acids 2020, 22, 251–262. [Google Scholar] [CrossRef]
  58. Xu, F.; Yang, J.; Shang, J.; Lan, F.; Li, M.; Shi, L.; Shen, L.; Wang, Y.; Ge, J. MicroRNA-302d promotes the proliferation of human pluripotent stem cell-derived cardiomyocytes by inhibiting LATS2 in the Hippo pathway. Clin. Sci. 2019, 133, 1387–1399. [Google Scholar] [CrossRef] [PubMed]
  59. Shi, J.; Qin, S.; Wang, R.; Zhang, Z.; Li, H.; Hao, C. MicroRNA-362-3p Implicated in Cardioprotection Against Hypoxia/Reoxygenation-Induced Cardiomyocyte Apoptosis by Repressing TP53INP2 and Regulating SDF-1/CXCR4 Pathway. Altern. Ther. Health Med. 2023, 29, 254–261. [Google Scholar]
  60. Zhao, Z.; Zhao, Y.; Ying-Chun, L.; Zhao, L.; Zhang, W.; Yang, J.G. Protective role of microRNA-374 against myocardial ischemia-reperfusion injury in mice following thoracic epidural anesthesia by downregulating dystrobrevin alpha-mediated Notch1 axis. J. Cell Physiol. 2019, 234, 10726–10740. [Google Scholar] [CrossRef] [PubMed]
  61. 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]
  62. Teng, Y.L.; Ren, F.; Xu, H.; Song, H.J. Overexpression of miRNA-410-3p protects hypoxia-induced cardiomyocyte injury via targeting TRAF5. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 9050–9057. [Google Scholar] [CrossRef]
  63. Jin, Y.; Ni, S. miR-496 remedies hypoxia reoxygenation-induced H9c2 cardiomyocyte apoptosis via Hook3-targeted PI3k/Akt/mTOR signaling pathway activation. J. Cell Biochem. 2020, 121, 698–712. [Google Scholar] [CrossRef]
  64. Deng, S.; Zhao, Q.; Zhen, L.; Zhang, C.; Liu, C.; Wang, G.; Zhang, L.; Bao, L.; Lu, Y.; Meng, L.; et al. Neonatal Heart-Enriched miR-708 Promotes Proliferation and Stress Resistance of Cardiomyocytes in Rodents. Theranostics 2017, 7, 1953–1965. [Google Scholar] [CrossRef]
  65. Gao, X.; Li, H.; Wang, X.; Ren, Z.; Tian, Y.; Zhao, J.; Qi, W.; Wang, H.; Yu, Y.; Gong, R.; et al. Light Emitting Diodes Irradiation Regulates miRNA-877-3p to Promote Cardiomyocyte Proliferation. Int. J. Med. Sci. 2022, 19, 1254–1264. [Google Scholar] [CrossRef]
  66. Hu, Y.; Jin, G.; Li, B.; Chen, Y.; Zhong, L.; Chen, G.; Chen, X.; Zhong, J.; Liao, W.; Liao, Y.; et al. Suppression of miRNA let-7i-5p promotes cardiomyocyte proliferation and repairs heart function post injury by targetting CCND2 and E2F2. Clin. Sci. 2019, 133, 425–441. [Google Scholar] [CrossRef]
  67. 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]
  68. Xu, J.; Cao, D.; Zhang, D.; Zhang, Y.; Yue, Y. MicroRNA-1 facilitates hypoxia-induced injury by targeting NOTCH3. J. Cell Biochem. 2020, 121, 4458–4469. [Google Scholar] [CrossRef]
  69. Zheng, J.; Peng, B.; Zhang, Y.; Ai, F.; Hu, X. miR-9 knockdown inhibits hypoxia-induced cardiomyocyte apoptosis by targeting Yap1. Life Sci. 2019, 219, 129–135. [Google Scholar] [CrossRef]
  70. Cao, X.; Wang, J.; Wang, Z.; Du, J.; Yuan, X.; Huang, W.; Meng, J.; Gu, H.; Nie, Y.; Ji, B.; et al. MicroRNA profiling during rat ventricular maturation: A role for miR-29a in regulating cardiomyocyte cell cycle re-entry. FEBS Lett. 2013, 587, 1548–1555. [Google Scholar] [CrossRef]
  71. Zhang, Y.; Matsushita, N.; Eigler, T.; Marban, E. Targeted MicroRNA Interference Promotes Postnatal Cardiac Cell Cycle Re-Entry. J. Regen. Med. 2013, 2, 2. [Google Scholar] [CrossRef]
  72. 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]
  73. Yang, Y.; Cheng, H.W.; Qiu, Y.; Dupee, D.; Noonan, M.; Lin, Y.D.; Fisch, S.; Unno, K.; Sereti, K.I.; Liao, R. MicroRNA-34a Plays a Key Role in Cardiac Repair and Regeneration Following Myocardial Infarction. Circ. Res. 2015, 117, 450–459. [Google Scholar] [CrossRef]
  74. 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] [PubMed]
  75. Liu, N.; Bezprozvannaya, S.; Williams, A.H.; Qi, X.; Richardson, J.A.; Bassel-Duby, R.; Olson, E.N. microRNA-133a regulates cardiomyocyte proliferation and suppresses smooth muscle gene expression in the heart. Genes. Dev. 2008, 22, 3242–3254. [Google Scholar] [CrossRef]
  76. Ma, W.Y.; Song, R.J.; Xu, B.B.; Xu, Y.; Wang, X.X.; Sun, H.Y.; Li, S.N.; Liu, S.Z.; Yu, M.X.; Yang, F.; et al. Melatonin promotes cardiomyocyte proliferation and heart repair in mice with myocardial infarction via miR-143-3p/Yap/Ctnnd1 signaling pathway. Acta Pharmacol. Sin. 2021, 42, 921–931. [Google Scholar] [CrossRef] [PubMed]
  77. 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]
  78. Zhu, X.; Lu, X. MiR-423-5p inhibition alleviates cardiomyocyte apoptosis and mitochondrial dysfunction caused by hypoxia/reoxygenation through activation of the wnt/beta-catenin signaling pathway via targeting MYBL2. J. Cell Physiol. 2019, 234, 22034–22043. [Google Scholar] [CrossRef] [PubMed]
  79. Xu, M.; Li, X.Y.; Song, L.; Tao, C.; Fang, J.; Tao, L. miR-484 targeting of Yap1-induced LPS-inhibited proliferation, and promoted apoptosis and inflammation in cardiomyocyte. Biosci. Biotechnol. Biochem. 2021, 85, 378–385. [Google Scholar] [CrossRef]
  80. Zhang, J.S.; Zhao, Y.; Lv, Y.; Liu, P.Y.; Ruan, J.X.; Sun, Y.L.; Gong, T.X.; Wan, N.; Qiu, G.R. miR-873 suppresses H9C2 cardiomyocyte proliferation by targeting GLI1. Gene 2017, 626, 426–432. [Google Scholar] [CrossRef]
  81. Chapman, A.R.; Adamson, P.D.; Mills, N.L. Assessment and classification of patients with myocardial injury and infarction in clinical practice. Heart 2017, 103, 10–18. [Google Scholar] [CrossRef]
  82. Smyth, A.; O’Donnell, M.; Lamelas, P.; Teo, K.; Rangarajan, S.; Yusuf, S.; Investigators, I. Physical Activity and Anger or Emotional Upset as Triggers of Acute Myocardial Infarction: The INTERHEART Study. Circulation 2016, 134, 1059–1067. [Google Scholar] [CrossRef] [PubMed]
  83. Ghiasmand, M.; Moghadamnia, M.T.; Pourshaikhian, M.; Kazemnejad Lili, E. Acute triggers of myocardial infarction: A case-crossover study. Egypt. Heart J. 2017, 69, 223–228. [Google Scholar] [CrossRef] [PubMed]
  84. Claeys, M.J.; Rajagopalan, S.; Nawrot, T.S.; Brook, R.D. Climate and environmental triggers of acute myocardial infarction. Eur. Heart J. 2017, 38, 955–960. [Google Scholar] [CrossRef]
  85. Fathima, S.N. An Update on Myocardial Infarction. Curr. Res. Trends Med. Sci. Technol. 2021, 1, 216. [Google Scholar]
  86. Gemmati, D.; Varani, K.; Bramanti, B.; Piva, R.; Bonaccorsi, G.; Trentini, A.; Manfrinato, M.C.; Tisato, V.; Care, A.; Bellini, T. “Bridging the Gap” Everything that Could Have Been Avoided If We Had Applied Gender Medicine, Pharmacogenetics and Personalized Medicine in the Gender-Omics and Sex-Omics Era. Int. J. Mol. Sci. 2019, 21, 296. [Google Scholar] [CrossRef]
  87. Frangogiannis, N.G. Pathophysiology of Myocardial Infarction. Compr. Physiol. 2015, 5, 1841–1875. [Google Scholar] [CrossRef]
  88. Thygesen, K.; Alpert, J.S.; Jaffe, A.S.; Chaitman, B.R.; Bax, J.J.; Morrow, D.A.; White, H.D.; Group, E.S.C.S.D. Fourth universal definition of myocardial infarction (2018). Eur. Heart J. 2019, 40, 237–269. [Google Scholar] [CrossRef]
  89. Girelli, D.; Russo, C.; Ferraresi, P.; Olivieri, O.; Pinotti, M.; Friso, S.; Manzato, F.; Mazzucco, A.; Bernardi, F.; Corrocher, R. Polymorphisms in the factor VII gene and the risk of myocardial infarction in patients with coronary artery disease. N. Engl. J. Med. 2000, 343, 774–780. [Google Scholar] [CrossRef]
  90. Hoekema, L.; Castoldi, E.; Tans, G.; Girelli, D.; Gemmati, D.; Bernardi, F.; Rosing, J. Functional properties of factor V and factor Va encoded by the R2-gene. Thromb. Haemost. 2001, 85, 75–81. [Google Scholar] [CrossRef]
  91. Gemmati, D.; Federici, F.; Campo, G.; Tognazzo, S.; Serino, M.L.; De Mattei, M.; Valgimigli, M.; Malagutti, P.; Guardigli, G.; Ferraresi, P.; et al. Factor XIIIA-V34L and factor XIIIB-H95R gene variants: Effects on survival in myocardial infarction patients. Mol. Med. 2007, 13, 112–120. [Google Scholar] [CrossRef]
  92. Gemmati, D.; Vigliano, M.; Burini, F.; Mari, R.; El Mohsein, H.H.; Parmeggiani, F.; Serino, M.L. Coagulation Factor XIIIA (F13A1): Novel Perspectives in Treatment and Pharmacogenetics. Curr. Pharm. Des. 2016, 22, 1449–1459. [Google Scholar] [CrossRef] [PubMed]
  93. Soares, R.O.S.; Losada, D.M.; Jordani, M.C.; Evora, P.; Castro, E.S.O. Ischemia/Reperfusion Injury Revisited: An Overview of the Latest Pharmacological Strategies. Int. J. Mol. Sci. 2019, 20, 5034. [Google Scholar] [CrossRef] [PubMed]
  94. Epelman, S.; Liu, P.P.; Mann, D.L. Role of innate and adaptive immune mechanisms in cardiac injury and repair. Nat. Rev. Immunol. 2015, 15, 117–129. [Google Scholar] [CrossRef]
  95. Igney, F.H.; Krammer, P.H. Death and anti-death: Tumour resistance to apoptosis. Nat. Rev. Cancer 2002, 2, 277–288. [Google Scholar] [CrossRef] [PubMed]
  96. Elmore, S. Apoptosis: A review of programmed cell death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef] [PubMed]
  97. Khalid, N.; Azimpouran, M. Necrosis. In StatPearls; StatPearls Publishing LLC: Treasure Island, FL, USA, 2023. [Google Scholar]
  98. Lodrini, A.M.; Goumans, M.J. Cardiomyocytes Cellular Phenotypes After Myocardial Infarction. Front. Cardiovasc. Med. 2021, 8, 750510. [Google Scholar] [CrossRef]
  99. Oka, T.; Hikoso, S.; Yamaguchi, O.; Taneike, M.; Takeda, T.; Tamai, T.; Oyabu, J.; Murakawa, T.; Nakayama, H.; Nishida, K.; et al. Mitochondrial DNA that escapes from autophagy causes inflammation and heart failure. Nature 2012, 485, 251–255. [Google Scholar] [CrossRef]
  100. Frangogiannis, N.G.; Lindsey, M.L.; Michael, L.H.; Youker, K.A.; Bressler, R.B.; Mendoza, L.H.; Spengler, R.N.; Smith, C.W.; Entman, M.L. Resident cardiac mast cells degranulate and release preformed TNF-alpha, initiating the cytokine cascade in experimental canine myocardial ischemia/reperfusion. Circulation 1998, 98, 699–710. [Google Scholar] [CrossRef]
  101. Frangogiannis, N.G.; Smith, C.W.; Entman, M.L. The inflammatory response in myocardial infarction. Cardiovasc. Res. 2002, 53, 31–47. [Google Scholar] [CrossRef]
  102. Nahrendorf, M.; Hu, K.; Frantz, S.; Jaffer, F.A.; Tung, C.H.; Hiller, K.H.; Voll, S.; Nordbeck, P.; Sosnovik, D.; Gattenlohner, S.; et al. Factor XIII deficiency causes cardiac rupture, impairs wound healing, and aggravates cardiac remodeling in mice with myocardial infarction. Circulation 2006, 113, 1196–1202. [Google Scholar] [CrossRef] [PubMed]
  103. Heymans, S.; Luttun, A.; Nuyens, D.; Theilmeier, G.; Creemers, E.; Moons, L.; Dyspersin, G.D.; Cleutjens, J.P.; Shipley, M.; Angellilo, A.; et al. Inhibition of plasminogen activators or matrix metalloproteinases prevents cardiac rupture but impairs therapeutic angiogenesis and causes cardiac failure. Nat. Med. 1999, 5, 1135–1142. [Google Scholar] [CrossRef] [PubMed]
  104. Gemmati, D.; Zeri, G.; Orioli, E.; Mari, R.; Moratelli, S.; Vigliano, M.; Marchesini, J.; Grossi, M.E.; Pecoraro, A.; Cuneo, A.; et al. Factor XIII-A dynamics in acute myocardial infarction: A novel prognostic biomarker? Thromb. Haemost. 2015, 114, 123–132. [Google Scholar] [CrossRef] [PubMed]
  105. Vanhoutte, D.; Heymans, S. Factor XIII: The cement of the heart after myocardial infarction? Eur. Heart J. 2008, 29, 427–428. [Google Scholar] [CrossRef] [PubMed]
  106. Nahrendorf, M.; Weissleder, R.; Ertl, G. Does FXIII deficiency impair wound healing after myocardial infarction? PLoS ONE 2006, 1, e48. [Google Scholar] [CrossRef]
  107. Ansani, L.; Marchesini, J.; Pestelli, G.; Luisi, G.A.; Scillitani, G.; Longo, G.; Milani, D.; Serino, M.L.; Tisato, V.; Gemmati, D. F13A1 Gene Variant (V34L) and Residual Circulating FXIIIA Levels Predict Short- and Long-Term Mortality in Acute Myocardial Infarction after Coronary Angioplasty. Int. J. Mol. Sci. 2018, 19, 2766. [Google Scholar] [CrossRef]
  108. Zamboni, P.; Gemmati, D. Clinical implications of gene polymorphisms in venous leg ulcer: A model in tissue injury and reparative process. Thromb. Haemost. 2007, 98, 131–137. [Google Scholar]
  109. Zamboni, P.; De Mattei, M.; Ongaro, A.; Fogato, L.; Carandina, S.; De Palma, M.; Tognazzo, S.; Scapoli, G.L.; Serino, M.L.; Caruso, A.; et al. Factor XIII contrasts the effects of metalloproteinases in human dermal fibroblast cultured cells. Vasc. Endovasc. Surg. 2004, 38, 431–438. [Google Scholar] [CrossRef]
  110. Gemmati, D.; Tognazzo, S.; Serino, M.L.; Fogato, L.; Carandina, S.; De Palma, M.; Izzo, M.; De Mattei, M.; Ongaro, A.; Scapoli, G.L.; et al. Factor XIII V34L polymorphism modulates the risk of chronic venous leg ulcer progression and extension. Wound Repair. Regen. 2004, 12, 512–517. [Google Scholar] [CrossRef]
  111. Ngkelo, A.; Richart, A.; Kirk, J.A.; Bonnin, P.; Vilar, J.; Lemitre, M.; Marck, P.; Branchereau, M.; Le Gall, S.; Renault, N.; et al. Mast cells regulate myofilament calcium sensitization and heart function after myocardial infarction. J. Exp. Med. 2016, 213, 1353–1374. [Google Scholar] [CrossRef]
  112. Zhang, H.; Yang, K.; Chen, F.; Liu, Q.; Ni, J.; Cao, W.; Hua, Y.; He, F.; Liu, Z.; Li, L.; et al. Role of the CCL2-CCR2 axis in cardiovascular disease: Pathogenesis and clinical implications. Front. Immunol. 2022, 13, 975367. [Google Scholar] [CrossRef]
  113. Chen, B.; Frangogiannis, N.G. Chemokines in Myocardial Infarction. J. Cardiovasc. Transl. Res. 2021, 14, 35–52. [Google Scholar] [CrossRef]
  114. van der Laan, A.M.; Nahrendorf, M.; Piek, J.J. Healing and adverse remodelling after acute myocardial infarction: Role of the cellular immune response. Heart 2012, 98, 1384–1390. [Google Scholar] [CrossRef]
  115. Nahrendorf, M.; Pittet, M.J.; Swirski, F.K. Monocytes: Protagonists of infarct inflammation and repair after myocardial infarction. Circulation 2010, 121, 2437–2445. [Google Scholar] [CrossRef]
  116. Frangogiannis, N.G.; Mendoza, L.H.; Ren, G.; Akrivakis, S.; Jackson, P.L.; Michael, L.H.; Smith, C.W.; Entman, M.L. MCSF expression is induced in healing myocardial infarcts and may regulate monocyte and endothelial cell phenotype. Am. J. Physiol. Heart Circ. Physiol. 2003, 285, H483–H492. [Google Scholar] [CrossRef]
  117. Bazzan, E.; Casara, A.; Radu, C.M.; Tine, M.; Biondini, D.; Faccioli, E.; Pezzuto, F.; Bernardinello, N.; Conti, M.; Balestro, E.; et al. Macrophages-derived Factor XIII links coagulation to inflammation in COPD. Front. Immunol. 2023, 14, 1131292. [Google Scholar] [CrossRef]
  118. Anzai, A.; Choi, J.L.; He, S.; Fenn, A.M.; Nairz, M.; Rattik, S.; McAlpine, C.S.; Mindur, J.E.; Chan, C.T.; Iwamoto, Y.; et al. The infarcted myocardium solicits GM-CSF for the detrimental oversupply of inflammatory leukocytes. J. Exp. Med. 2017, 214, 3293–3310. [Google Scholar] [CrossRef] [PubMed]
  119. Matthia, E.L.; Setteducato, M.L.; Elzeneini, M.; Vernace, N.; Salerno, M.; Kramer, C.M.; Keeley, E.C. Circulating Biomarkers in Hypertrophic Cardiomyopathy. J. Am. Heart Assoc. 2022, 11, e027618. [Google Scholar] [CrossRef]
  120. Lillo, R.; Graziani, F.; Franceschi, F.; Iannaccone, G.; Massetti, M.; Olivotto, I.; Crea, F.; Liuzzo, G. Inflammation across the spectrum of hypertrophic cardiac phenotypes. Heart Fail. Rev. 2023, 28, 1065–1075. [Google Scholar] [CrossRef] [PubMed]
  121. Becker, R.C.; Owens, A.P., 3rd; Sadayappan, S. Tissue-level inflammation and ventricular remodeling in hypertrophic cardiomyopathy. J. Thromb. Thrombolysis 2020, 49, 177–183. [Google Scholar] [CrossRef] [PubMed]
  122. Weirather, J.; Hofmann, U.D.; Beyersdorf, N.; Ramos, G.C.; Vogel, B.; Frey, A.; Ertl, G.; Kerkau, T.; Frantz, S. Foxp3+ CD4+ T cells improve healing after myocardial infarction by modulating monocyte/macrophage differentiation. Circ. Res. 2014, 115, 55–67. [Google Scholar] [CrossRef] [PubMed]
  123. Nahrendorf, M.; Swirski, F.K. Regulating repair: Regulatory T cells in myocardial infarction. Circ. Res. 2014, 115, 7–9. [Google Scholar] [CrossRef]
  124. Xia, N.; Lu, Y.; Gu, M.; Li, N.; Liu, M.; Jiao, J.; Zhu, Z.; Li, J.; Li, D.; Tang, T.; et al. A Unique Population of Regulatory T Cells in Heart Potentiates Cardiac Protection From Myocardial Infarction. Circulation 2020, 142, 1956–1973. [Google Scholar] [CrossRef]
  125. Lu, Y.; Xia, N.; Cheng, X. Regulatory T Cells in Chronic Heart Failure. Front. Immunol. 2021, 12, 732794. [Google Scholar] [CrossRef]
  126. Delgobo, M.; Weiss, E.; Ashour, D.; Richter, L.; Popiolkowski, L.; Arampatzi, P.; Stangl, V.; Arias-Loza, P.; Mariotti-Ferrandiz, E.; Rainer, P.P.; et al. Myocardial Milieu Favors Local Differentiation of Regulatory T Cells. Circ. Res. 2023, 132, 565–582. [Google Scholar] [CrossRef]
  127. Wang, J.; Lin, B.; Zhang, Y.; Ni, L.; Hu, L.; Yang, J.; Xu, L.; Shi, D.; Chen, Y.H. The Regulatory Role of Histone Modification on Gene Expression in the Early Stage of Myocardial Infarction. Front. Cardiovasc. Med. 2020, 7, 594325. [Google Scholar] [CrossRef]
  128. Lei, I.; Tian, S.; Gao, W.; Liu, L.; Guo, Y.; Tang, P.; Chen, E.; Wang, Z. Acetyl-CoA production by specific metabolites promotes cardiac repair after myocardial infarction via histone acetylation. Elife 2021, 10, e60311. [Google Scholar] [CrossRef]
  129. Cohen, H.Y.; Miller, C.; Bitterman, K.J.; Wall, N.R.; Hekking, B.; Kessler, B.; Howitz, K.T.; Gorospe, M.; de Cabo, R.; Sinclair, D.A. Calorie restriction promotes mammalian cell survival by inducing the SIRT1 deacetylase. Science 2004, 305, 390–392. [Google Scholar] [CrossRef]
  130. Motta, M.C.; Divecha, N.; Lemieux, M.; Kamel, C.; Chen, D.; Gu, W.; Bultsma, Y.; McBurney, M.; Guarente, L. Mammalian SIRT1 represses forkhead transcription factors. Cell 2004, 116, 551–563. [Google Scholar] [CrossRef]
  131. Mattagajasingh, I.; Kim, C.S.; Naqvi, A.; Yamamori, T.; Hoffman, T.A.; Jung, S.B.; DeRicco, J.; Kasuno, K.; Irani, K. SIRT1 promotes endothelium-dependent vascular relaxation by activating endothelial nitric oxide synthase. Proc. Natl. Acad. Sci. USA 2007, 104, 14855–14860. [Google Scholar] [CrossRef]
  132. Shinozaki, S.; Chang, K.; Sakai, M.; Shimizu, N.; Yamada, M.; Tanaka, T.; Nakazawa, H.; Ichinose, F.; Yamada, Y.; Ishigami, A.; et al. Inflammatory stimuli induce inhibitory S-nitrosylation of the deacetylase SIRT1 to increase acetylation and activation of p53 and p65. Sci. Signal 2014, 7, ra106. [Google Scholar] [CrossRef]
  133. Hsu, C.P.; Zhai, P.; Yamamoto, T.; Maejima, Y.; Matsushima, S.; Hariharan, N.; Shao, D.; Takagi, H.; Oka, S.; Sadoshima, J. Silent information regulator 1 protects the heart from ischemia/reperfusion. Circulation 2010, 122, 2170–2182. [Google Scholar] [CrossRef]
  134. Izarra, A.; Moscoso, I.; Levent, E.; Canon, S.; Cerrada, I.; Diez-Juan, A.; Blanca, V.; Nunez-Gil, I.J.; Valiente, I.; Ruiz-Sauri, A.; et al. miR-133a enhances the protective capacity of cardiac progenitors cells after myocardial infarction. Stem Cell Rep. 2014, 3, 1029–1042. [Google Scholar] [CrossRef]
  135. Han, F.; Chen, Q.; Su, J.; Zheng, A.; Chen, K.; Sun, S.; Wu, H.; Jiang, L.; Xu, X.; Yang, M.; et al. MicroRNA-124 regulates cardiomyocyte apoptosis and myocardial infarction through targeting Dhcr24. J. Mol. Cell. Cardiol. 2019, 132, 178–188. [Google Scholar] [CrossRef]
  136. Fan, F.; Sun, A.; Zhao, H.; Liu, X.; Zhang, W.; Jin, X.; Wang, C.; Ma, X.; Shen, C.; Zou, Y.; et al. MicroRNA-34a promotes cardiomyocyte apoptosis post myocardial infarction through down-regulating aldehyde dehydrogenase 2. Curr. Pharm. Des. 2013, 19, 4865–4873. [Google Scholar] [CrossRef]
  137. Xu, H.; Jin, L.; Chen, Y.; Li, J. Downregulation of microRNA-429 protects cardiomyocytes against hypoxia-induced apoptosis by increasing Notch1 expression. Int. J. Mol. Med. 2016, 37, 1677–1685. [Google Scholar] [CrossRef]
  138. Li, X.; Zhang, S.; Wa, M.; Liu, Z.; Hu, S. MicroRNA-101 Protects Against Cardiac Remodeling Following Myocardial Infarction via Downregulation of Runt-Related Transcription Factor 1. J. Am. Heart Assoc. 2019, 8, e013112. [Google Scholar] [CrossRef]
  139. Li, Q.; Gao, Y.; Zhu, J.; Jia, Q. MiR-101 Attenuates Myocardial Infarction-induced Injury by Targeting DDIT4 to Regulate Autophagy. Curr. Neurovascular Res. 2020, 17, 123–130. [Google Scholar] [CrossRef]
  140. Ding, S.; Liu, D.; Wang, L.; Wang, G.; Zhu, Y. Inhibiting MicroRNA-29a Protects Myocardial Ischemia-Reperfusion Injury by Targeting SIRT1 and Suppressing Oxidative Stress and NLRP3-Mediated Pyroptosis Pathway. J. Pharmacol. Exp. Ther. 2020, 372, 128–135. [Google Scholar] [CrossRef]
  141. Zhou, Y.; Li, K.S.; Liu, L.; Li, S.L. MicroRNA-132 promotes oxidative stress-induced pyroptosis by targeting sirtuin 1 in myocardial ischaemia-reperfusion injury. Int. J. Mol. Med. 2020, 45, 1942–1950. [Google Scholar] [CrossRef]
  142. Huang, G.; Hao, F.; Hu, X. Downregulation of microRNA-155 stimulates sevoflurane-mediated cardioprotection against myocardial ischemia/reperfusion injury by binding to SIRT1 in mice. J. Cell Biochem. 2019, 120, 15494–15505. [Google Scholar] [CrossRef]
  143. Feng, Y.; Bao, Y.; Ding, J.; Li, H.; Liu, W.; Wang, X.; Guan, H.; Chen, Z. MicroRNA-130a attenuates cardiac fibrosis after myocardial infarction through TGF-beta/Smad signaling by directly targeting TGF-beta receptor 1. Bioengineered 2022, 13, 5779–5791. [Google Scholar] [CrossRef]
  144. Yu, B.T.; Yu, N.; Wang, Y.; Zhang, H.; Wan, K.; Sun, X.; Zhang, C.S. Role of miR-133a in regulating TGF-beta1 signaling pathway in myocardial fibrosis after acute myocardial infarction in rats. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 8588–8597. [Google Scholar] [CrossRef]
  145. Zhang, S.; Gao, S.; Wang, Y.; Jin, P.; Lu, F. lncRNA SRA1 Promotes the Activation of Cardiac Myofibroblasts Through Negative Regulation of miR-148b. DNA Cell Biol. 2019, 38, 385–394. [Google Scholar] [CrossRef]
  146. Yuan, X.; Pan, J.; Wen, L.; Gong, B.; Li, J.; Gao, H.; Tan, W.; Liang, S.; Zhang, H.; Wang, X. MiR-590-3p regulates proliferation, migration and collagen synthesis of cardiac fibroblast by targeting ZEB1. J. Cell Mol. Med. 2020, 24, 227–237. [Google Scholar] [CrossRef]
  147. Wang, X.; Morelli, M.B.; Matarese, A.; Sardu, C.; Santulli, G. Cardiomyocyte-derived exosomal microRNA-92a mediates post-ischemic myofibroblast activation both in vitro and ex vivo. ESC Heart Fail. 2020, 7, 284–288. [Google Scholar] [CrossRef]
  148. Morelli, M.B.; Shu, J.; Sardu, C.; Matarese, A.; Santulli, G. Cardiosomal microRNAs Are Essential in Post-Infarction Myofibroblast Phenoconversion. Int. J. Mol. Sci. 2019, 21, 201. [Google Scholar] [CrossRef]
  149. Tzavlaki, K.; Moustakas, A. TGF-β Signaling. Biomolecules 2020, 10, 487. [Google Scholar] [CrossRef]
  150. Dardik, R.; Loscalzo, J.; Eskaraev, R.; Inbal, A. Molecular mechanisms underlying the proangiogenic effect of factor XIII. Arterioscler. Thromb. Vasc. Biol. 2005, 25, 526–532. [Google Scholar] [CrossRef]
  151. Dardik, R.; Solomon, A.; Loscalzo, J.; Eskaraev, R.; Bialik, A.; Goldberg, I.; Schiby, G.; Inbal, A. Novel proangiogenic effect of factor XIII associated with suppression of thrombospondin 1 expression. Arterioscler. Thromb. Vasc. Biol. 2003, 23, 1472–1477. [Google Scholar] [CrossRef]
  152. Bournazou, I.; Pound, J.D.; Duffin, R.; Bournazos, S.; Melville, L.A.; Brown, S.B.; Rossi, A.G.; Gregory, C.D. Apoptotic human cells inhibit migration of granulocytes via release of lactoferrin. J. Clin. Investig. 2009, 119, 20–32. [Google Scholar] [CrossRef]
  153. Soehnlein, O.; Lindbom, L. Phagocyte partnership during the onset and resolution of inflammation. Nat. Rev. Immunol. 2010, 10, 427–439. [Google Scholar] [CrossRef]
  154. Morimoto, H.; Takahashi, M. Role of monocyte chemoattractant protein-1 in myocardial infarction. Int. J. Biomed. Sci. 2007, 3, 159–167. [Google Scholar]
  155. Hanna, A.; Frangogiannis, N.G. The Role of the TGF-beta Superfamily in Myocardial Infarction. Front. Cardiovasc. Med. 2019, 6, 140. [Google Scholar] [CrossRef] [PubMed]
  156. Frangogiannis, N.G. The mechanistic basis of infarct healing. Antioxid. Redox Signal. 2006, 8, 1907–1939. [Google Scholar] [CrossRef] [PubMed]
  157. Frangogiannis, N.G. Matricellular proteins in cardiac adaptation and disease. Physiol. Rev. 2012, 92, 635–688. [Google Scholar] [CrossRef]
  158. Pinto, A.R.; Ilinykh, A.; Ivey, M.J.; Kuwabara, J.T.; D’Antoni, M.L.; Debuque, R.; Chandran, A.; Wang, L.; Arora, K.; Rosenthal, N.A.; et al. Revisiting Cardiac Cellular Composition. Circ. Res. 2016, 118, 400–409. [Google Scholar] [CrossRef] [PubMed]
  159. Tombor, L.S.; Dimmeler, S. Why is endothelial resilience key to maintain cardiac health? Basic. Res. Cardiol. 2022, 117, 35. [Google Scholar] [CrossRef]
  160. Ren, G.; Michael, L.H.; Entman, M.L.; Frangogiannis, N.G. Morphological characteristics of the microvasculature in healing myocardial infarcts. J. Histochem. Cytochem. 2002, 50, 71–79. [Google Scholar] [CrossRef]
  161. Tisato, V.; Zauli, G.; Rimondi, E.; Gianesini, S.; Brunelli, L.; Menegatti, E.; Zamboni, P.; Secchiero, P. Inhibitory effect of natural anti-inflammatory compounds on cytokines released by chronic venous disease patient-derived endothelial cells. Mediat. Inflamm. 2013, 2013, 423407. [Google Scholar] [CrossRef]
  162. Cervellati, C.; Valacchi, G.; Tisato, V.; Zuliani, G.; Marsillach, J. Evaluating the link between Paraoxonase-1 levels and Alzheimer’s disease development. Minerva Med. 2019, 110, 238–250. [Google Scholar] [CrossRef]
  163. Gemmati, D.; Occhionorelli, S.; Tisato, V.; Vigliano, M.; Longo, G.; Gonelli, A.; Sibilla, M.G.; Serino, M.L.; Zamboni, P. Inherited genetic predispositions in F13A1 and F13B genes predict abdominal adhesion formation: Identification of gender prognostic indicators. Sci. Rep. 2018, 8, 16916. [Google Scholar] [CrossRef]
  164. Duong, T.E.; Hagood, J.S. Epigenetic Regulation of Myofibroblast Phenotypes in Fibrosis. Curr. Pathobiol. Rep. 2018, 6, 79–96. [Google Scholar] [CrossRef] [PubMed]
  165. Grimaldi, V.; De Pascale, M.R.; Zullo, A.; Soricelli, A.; Infante, T.; Mancini, F.P.; Napoli, C. Evidence of epigenetic tags in cardiac fibrosis. J. Cardiol. 2017, 69, 401–408. [Google Scholar] [CrossRef] [PubMed]
  166. Kang, S.H.; Seok, Y.M.; Song, M.J.; Lee, H.A.; Kurz, T.; Kim, I. Histone deacetylase inhibition attenuates cardiac hypertrophy and fibrosis through acetylation of mineralocorticoid receptor in spontaneously hypertensive rats. Mol. Pharmacol. 2015, 87, 782–791. [Google Scholar] [CrossRef] [PubMed]
  167. Shiojima, I.; Yefremashvili, M.; Luo, Z.; Kureishi, Y.; Takahashi, A.; Tao, J.; Rosenzweig, A.; Kahn, C.R.; Abel, E.D.; Walsh, K. Akt signaling mediates postnatal heart growth in response to insulin and nutritional status. J. Biol. Chem. 2002, 277, 37670–37677. [Google Scholar] [CrossRef] [PubMed]
  168. Catalucci, D.; Condorelli, G. Effects of Akt on cardiac myocytes: Location counts. Circ. Res. 2006, 99, 339–341. [Google Scholar] [CrossRef]
  169. Shen, T.; Ding, L.; Ruan, Y.; Qin, W.; Lin, Y.; Xi, C.; Lu, Y.; Dou, L.; Zhu, Y.; Cao, Y.; et al. SIRT1 functions as an important regulator of estrogen-mediated cardiomyocyte protection in angiotensin II-induced heart hypertrophy. Oxid. Med. Cell Longev. 2014, 2014, 713894. [Google Scholar] [CrossRef]
  170. Sundaresan, N.R.; Gupta, M.; Kim, G.; Rajamohan, S.B.; Isbatan, A.; Gupta, M.P. Sirt3 blocks the cardiac hypertrophic response by augmenting Foxo3a-dependent antioxidant defense mechanisms in mice. J. Clin. Investig. 2009, 119, 2758–2771. [Google Scholar] [CrossRef]
  171. Tian, K.; Liu, Z.; Wang, J.; Xu, S.; You, T.; Liu, P. Sirtuin-6 inhibits cardiac fibroblasts differentiation into myofibroblasts via inactivation of nuclear factor kappaB signaling. Transl. Res. 2015, 165, 374–386. [Google Scholar] [CrossRef]
  172. Alique, M.; Bodega, G.; Giannarelli, C.; Carracedo, J.; Ramirez, R. MicroRNA-126 regulates Hypoxia-Inducible Factor-1alpha which inhibited migration, proliferation, and angiogenesis in replicative endothelial senescence. Sci. Rep. 2019, 9, 7381. [Google Scholar] [CrossRef]
  173. Soufi-Zomorrod, M.; Hajifathali, A.; Kouhkan, F.; Mehdizadeh, M.; Rad, S.M.; Soleimani, M. MicroRNAs modulating angiogenesis: miR-129-1 and miR-133 act as angio-miR in HUVECs. Tumour Biol. 2016, 37, 9527–9534. [Google Scholar] [CrossRef] [PubMed]
  174. Wang, C.; Yang, D.; Xu, C.; Duan, H. MicroRNA-139-5p inhibits vascular endothelial cell viability and serves as a diagnostic biomarker in acute myocardial infarction patients. Exp. Gerontol. 2021, 152, 111453. [Google Scholar] [CrossRef] [PubMed]
  175. Chen, T.; Margariti, A.; Kelaini, S.; Cochrane, A.; Guha, S.T.; Hu, Y.; Stitt, A.W.; Zhang, L.; Xu, Q. MicroRNA-199b Modulates Vascular Cell Fate During iPS Cell Differentiation by Targeting the Notch Ligand Jagged1 and Enhancing VEGF Signaling. Stem Cells 2015, 33, 1405–1418. [Google Scholar] [CrossRef]
  176. Ranjan, P.; Kumari, R.; Goswami, S.K.; Li, J.; Pal, H.; Suleiman, Z.; Cheng, Z.; Krishnamurthy, P.; Kishore, R.; Verma, S.K. Myofibroblast-Derived Exosome Induce Cardiac Endothelial Cell Dysfunction. Front. Cardiovasc. Med. 2021, 8, 676267. [Google Scholar] [CrossRef] [PubMed]
  177. Lu, C.; Wang, X.; Ha, T.; Hu, Y.; Liu, L.; Zhang, X.; Yu, H.; Miao, J.; Kao, R.; Kalbfleisch, J.; et al. Attenuation of cardiac dysfunction and remodeling of myocardial infarction by microRNA-130a are mediated by suppression of PTEN and activation of PI3K dependent signaling. J. Mol. Cell. Cardiol. 2015, 89, 87–97. [Google Scholar] [CrossRef] [PubMed]
  178. Zhang, X.-T.; Xu, M.-G. Potential link between microRNA-208 and cardiovascular diseases. J. Xiangya Med. 2021, 6, 10–21037. [Google Scholar] [CrossRef]
  179. Wei, C.; Kim, I.K.; Kumar, S.; Jayasinghe, S.; Hong, N.; Castoldi, G.; Catalucci, D.; Jones, W.K.; Gupta, S. NF-kappaB mediated miR-26a regulation in cardiac fibrosis. J. Cell. Physiol. 2013, 228, 1433–1442. [Google Scholar] [CrossRef] [PubMed]
  180. Wang, Y.; Jin, B.J.; Chen, Q.; Yan, B.J.; Liu, Z.L. MicroRNA-29b upregulation improves myocardial fibrosis and cardiac function in myocardial infarction rats through targeting SH2B3. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 10115–10122. [Google Scholar] [CrossRef]
  181. Zhang, S.; Zhang, R.; Wu, F.; Li, X. MicroRNA-208a Regulates H9c2 Cells Simulated Ischemia-Reperfusion Myocardial Injury via Targeting CHD9 through Notch/NF-kappa B Signal Pathways. Int. Heart J. 2018, 59, 580–588. [Google Scholar] [CrossRef]
  182. Zhu, J.N.; Chen, R.; Fu, Y.H.; Lin, Q.X.; Huang, S.; Guo, L.L.; Zhang, M.Z.; Deng, C.Y.; Zou, X.; Zhong, S.L.; et al. Smad3 inactivation and MiR-29b upregulation mediate the effect of carvedilol on attenuating the acute myocardium infarction-induced myocardial fibrosis in rat. PLoS ONE 2013, 8, e75557. [Google Scholar] [CrossRef]
  183. Shen, D.; He, Z. Mesenchymal stem cell-derived exosomes regulate the polarization and inflammatory response of macrophages via miR-21-5p to promote repair after myocardial reperfusion injury. Ann. Transl. Med. 2021, 9, 1323. [Google Scholar] [CrossRef] [PubMed]
  184. Qiu, M.; Ma, J.; Zhang, J.; Guo, X.; Liu, Q.; Yang, Z. MicroRNA-150 deficiency accelerates intimal hyperplasia by acting as a novel regulator of macrophage polarization. Life Sci. 2020, 240, 116985. [Google Scholar] [CrossRef] [PubMed]
  185. Zhao, J.; Li, X.; Hu, J.; Chen, F.; Qiao, S.; Sun, X.; Gao, L.; Xie, J.; Xu, B. Mesenchymal stromal cell-derived exosomes attenuate myocardial ischaemia-reperfusion injury through miR-182-regulated macrophage polarization. Cardiovasc. Res. 2019, 115, 1205–1216. [Google Scholar] [CrossRef]
  186. Zhu, F.; Chen, Y.; Li, J.; Yang, Z.; Lin, Y.; Jiang, B.; Shao, L.; Hu, S.; Shen, Z. Human Umbilical Cord Mesenchymal Stem Cell-Derived Exosomes Attenuate Myocardial Infarction Injury via miR-24-3p-Promoted M2 Macrophage Polarization. Adv. Biol. 2022, 6, e2200074. [Google Scholar] [CrossRef] [PubMed]
  187. Jung, J.H.; Ikeda, G.; Tada, Y.; von Bornstadt, D.; Santoso, M.R.; Wahlquist, C.; Rhee, S.; Jeon, Y.J.; Yu, A.C.; O’Brien, C.G.; et al. miR-106a-363 cluster in extracellular vesicles promotes endogenous myocardial repair via Notch3 pathway in ischemic heart injury. Basic. Res. Cardiol. 2021, 116, 19. [Google Scholar] [CrossRef]
  188. Garikipati, V.N.S.; Verma, S.K.; Jolardarashi, D.; Cheng, Z.; Ibetti, J.; Cimini, M.; Tang, Y.; Khan, M.; Yue, Y.; Benedict, C.; et al. Therapeutic inhibition of miR-375 attenuates post-myocardial infarction inflammatory response and left ventricular dysfunction via PDK-1-AKT signalling axis. Cardiovasc. Res. 2017, 113, 938–949. [Google Scholar] [CrossRef]
  189. Zimmerman, S.D.; Thomas, D.P.; Velleman, S.G.; Li, X.; Hansen, T.R.; McCormick, R.J. Time course of collagen and decorin changes in rat cardiac and skeletal muscle post-MI. Am. J. Physiol. Heart Circ. Physiol. 2001, 281, H1816–H1822. [Google Scholar] [CrossRef]
  190. Alex, L.; Frangogiannis, N.G. Pericytes in the infarcted heart. Vasc. Biol. 2019, 1, H23–H31. [Google Scholar] [CrossRef]
  191. Hinz, B. Formation and function of the myofibroblast during tissue repair. J. Investig. Dermatol. 2007, 127, 526–537. [Google Scholar] [CrossRef]
  192. Fu, X.; Khalil, H.; Kanisicak, O.; Boyer, J.G.; Vagnozzi, R.J.; Maliken, B.D.; Sargent, M.A.; Prasad, V.; Valiente-Alandi, I.; Blaxall, B.C.; et al. Specialized fibroblast differentiated states underlie scar formation in the infarcted mouse heart. J. Clin. Investig. 2018, 128, 2127–2143. [Google Scholar] [CrossRef]
  193. Frangogiannis, N.G. The extracellular matrix in myocardial injury, repair, and remodeling. J. Clin. Investig. 2017, 127, 1600–1612. [Google Scholar] [CrossRef]
  194. Zymek, P.; Bujak, M.; Chatila, K.; Cieslak, A.; Thakker, G.; Entman, M.L.; Frangogiannis, N.G. The role of platelet-derived growth factor signaling in healing myocardial infarcts. J. Am. Coll. Cardiol. 2006, 48, 2315–2323. [Google Scholar] [CrossRef]
  195. Hellstrom, M.; Kalen, M.; Lindahl, P.; Abramsson, A.; Betsholtz, C. Role of PDGF-B and PDGFR-beta in recruitment of vascular smooth muscle cells and pericytes during embryonic blood vessel formation in the mouse. Development 1999, 126, 3047–3055. [Google Scholar] [CrossRef]
  196. Felician, G.; Collesi, C.; Lusic, M.; Martinelli, V.; Ferro, M.D.; Zentilin, L.; Zacchigna, S.; Giacca, M. Epigenetic modification at Notch responsive promoters blunts efficacy of inducing notch pathway reactivation after myocardial infarction. Circ. Res. 2014, 115, 636–649. [Google Scholar] [CrossRef]
  197. Kraus, L. Targeting Epigenetic Regulation of Cardiomyocytes through Development for Therapeutic Cardiac Regeneration after Heart Failure. Int. J. Mol. Sci. 2022, 23, 11878. [Google Scholar] [CrossRef] [PubMed]
  198. Boulet, J.; Mehra, M.R. Left Ventricular Reverse Remodeling in Heart Failure: Remission to Recovery. Struct. Heart 2021, 5, 466–481. [Google Scholar] [CrossRef]
  199. Millett, E.R.C.; Peters, S.A.E.; Woodward, M. Sex differences in risk factors for myocardial infarction: Cohort study of UK Biobank participants. BMJ 2018, 363, k4247. [Google Scholar] [CrossRef]
  200. Vaccarezza, M.; Papa, V.; Milani, D.; Gonelli, A.; Secchiero, P.; Zauli, G.; Gemmati, D.; Tisato, V. Sex/Gender-Specific Imbalance in CVD: Could Physical Activity Help to Improve Clinical Outcome Targeting CVD Molecular Mechanisms in Women? Int. J. Mol. Sci. 2020, 21, 1477. [Google Scholar] [CrossRef]
  201. Stehli, J.; Duffy, S.J.; Burgess, S.; Kuhn, L.; Gulati, M.; Chow, C.; Zaman, S. Sex Disparities in Myocardial Infarction: Biology or Bias? Heart Lung Circ. 2021, 30, 18–26. [Google Scholar] [CrossRef] [PubMed]
  202. Cervellati, C.; Bonaccorsi, G.; Bergamini, C.M.; Fila, E.; Greco, P.; Valacchi, G.; Massari, L.; Gonelli, A.; Tisato, V. Association between circulatory levels of adipokines and bone mineral density in postmenopausal women. Menopause 2016, 23, 984–992. [Google Scholar] [CrossRef] [PubMed]
  203. Agostinis, C.; Bulla, R.; Tisato, V.; De Seta, F.; Alberico, S.; Secchiero, P.; Zauli, G. Soluble TRAIL is elevated in recurrent miscarriage and inhibits the in vitro adhesion and migration of HTR8 trophoblastic cells. Hum. Reprod. 2012, 27, 2941–2947. [Google Scholar] [CrossRef]
  204. Bertelsen, D.M.; Neergaard, J.S.; Bager, C.L.; Nielsen, S.H.; Secher, N.H.; Svendsen, J.H.; Bihlet, A.R.; Andersen, J.R.; Karsdal, M.A.; Christiansen, C.; et al. Matrix Metalloproteinase Mediated Type I Collagen Degradation is an Independent Predictor of Increased Risk of Acute Myocardial Infarction in Postmenopausal Women. Sci. Rep. 2018, 8, 5371. [Google Scholar] [CrossRef]
  205. Haug, E.B.; Horn, J.; Markovitz, A.R.; Fraser, A.; Klykken, B.; Dalen, H.; Vatten, L.J.; Romundstad, P.R.; Rich-Edwards, J.W.; Asvold, B.O. Association of Conventional Cardiovascular Risk Factors With Cardiovascular Disease After Hypertensive Disorders of Pregnancy: Analysis of the Nord-Trondelag Health Study. JAMA Cardiol. 2019, 4, 628–635. [Google Scholar] [CrossRef]
  206. Retnakaran, R.; Shah, B.R. Glucose screening in pregnancy and future risk of cardiovascular disease in women: A retrospective, population-based cohort study. Lancet Diabetes Endocrinol. 2019, 7, 378–384. [Google Scholar] [CrossRef]
  207. Brush, J.E., Jr.; Krumholz, H.M.; Greene, E.J.; Dreyer, R.P. Sex Differences in Symptom Phenotypes Among Patients With Acute Myocardial Infarction. Circ. Cardiovasc. Qual. Outcomes 2020, 13, e005948. [Google Scholar] [CrossRef]
  208. Crea, F.; Battipaglia, I.; Andreotti, F. Sex differences in mechanisms, presentation and management of ischaemic heart disease. Atherosclerosis 2015, 241, 157–168. [Google Scholar] [CrossRef] [PubMed]
  209. Asleh, R.; Manemann, S.M.; Weston, S.A.; Bielinski, S.J.; Chamberlain, A.M.; Jiang, R.; Gerber, Y.; Roger, V.L. Sex Differences in Outcomes After Myocardial Infarction in the Community. Am. J. Med. 2021, 134, 114–121. [Google Scholar] [CrossRef] [PubMed]
  210. Gemmati, D.; Tisato, V. Chapter 24—Genomic and epigenomic signature at the branch-point among genome, phenome, and sexome in health and disease: A multiomics approach. In Principles of Gender-Specific Medicine, 4th ed.; Legato, M.J., Ed.; Academic Press: Cambridge, MA, USA, 2023; pp. 393–408. [Google Scholar] [CrossRef]
  211. Gemmati, D.; Tisato, V. Chapter 15—Genetics and epigenetics of the one-carbon metabolism pathway in autism spectrum disorder: Role of a sex-specific brain epigenome. In Sex, Gender, and Epigenetics; Legato, M.J., Feldberg, D., Glezerman, M., Eds.; Academic Press: Cambridge, MA, USA, 2023; pp. 181–191. [Google Scholar] [CrossRef]
  212. Tisato, V.; Silva, J.A.; Longo, G.; Gallo, I.; Singh, A.V.; Milani, D.; Gemmati, D. Genetics and Epigenetics of One-Carbon Metabolism Pathway in Autism Spectrum Disorder: A Sex-Specific Brain Epigenome? Genes 2021, 12, 782. [Google Scholar] [CrossRef]
  213. Chakraborty, C.; Sharma, A.R.; Sharma, G.; Doss, C.G.P.; Lee, S.S. Therapeutic miRNA and siRNA: Moving from Bench to Clinic as Next Generation Medicine. Mol. Ther. Nucleic Acids 2017, 8, 132–143. [Google Scholar] [CrossRef]
  214. Dasgupta, I.; Chatterjee, A. Recent Advances in miRNA Delivery Systems. Methods Protoc. 2021, 4, 10. [Google Scholar] [CrossRef]
  215. Hullinger, T.G.; Montgomery, R.L.; Seto, A.G.; Dickinson, B.A.; Semus, H.M.; Lynch, J.M.; Dalby, C.M.; Robinson, K.; Stack, C.; Latimer, P.A.; et al. Inhibition of miR-15 protects against cardiac ischemic injury. Circ. Res. 2012, 110, 71–81. [Google Scholar] [CrossRef] [PubMed]
  216. Wang, J.; Huang, W.; Xu, R.; Nie, Y.; Cao, X.; Meng, J.; Xu, X.; Hu, S.; Zheng, Z. MicroRNA-24 regulates cardiac fibrosis after myocardial infarction. J. Cell Mol. Med. 2012, 16, 2150–2160. [Google Scholar] [CrossRef] [PubMed]
  217. Han, M.; Yang, Z.; Sayed, D.; He, M.; Gao, S.; Lin, L.; Yoon, S.; Abdellatif, M. GATA4 expression is primarily regulated via a miR-26b-dependent post-transcriptional mechanism during cardiac hypertrophy. Cardiovasc. Res. 2012, 93, 645–654. [Google Scholar] [CrossRef]
  218. Zhang, B.; Zhou, M.; Li, C.; Zhou, J.; Li, H.; Zhu, D.; Wang, Z.; Chen, A.; Zhao, Q. MicroRNA-92a inhibition attenuates hypoxia/reoxygenation-induced myocardiocyte apoptosis by targeting Smad7. PLoS ONE 2014, 9, e100298. [Google Scholar] [CrossRef]
  219. Liang, W.; Guo, J.; Li, J.; Bai, C.; Dong, Y. Downregulation of miR-122 attenuates hypoxia/reoxygenation (H/R)-induced myocardial cell apoptosis by upregulating GATA-4. Biochem. Biophys. Res. Commun. 2016, 478, 1416–1422. [Google Scholar] [CrossRef] [PubMed]
  220. Zhou, T.; Qin, G.; Yang, L.; Xiang, D.; Li, S. LncRNA XIST regulates myocardial infarction by targeting miR-130a-3p. J. Cell Physiol. 2019, 234, 8659–8667. [Google Scholar] [CrossRef]
  221. Zhao, Z.; Du, S.; Shen, S.; Wang, L. microRNA-132 inhibits cardiomyocyte apoptosis and myocardial remodeling in myocardial infarction by targeting IL-1beta. J. Cell Physiol. 2020, 235, 2710–2721. [Google Scholar] [CrossRef]
  222. Gong, X.; Zhu, Y.; Chang, H.; Li, Y.; Ma, F. Long noncoding RNA MALAT1 promotes cardiomyocyte apoptosis after myocardial infarction via targeting miR-144-3p. Biosci. Rep. 2019, 39, BSR20191103. [Google Scholar] [CrossRef]
  223. Li, R.; Yan, G.; Li, Q.; Sun, H.; Hu, Y.; Sun, J.; Xu, B. MicroRNA-145 protects cardiomyocytes against hydrogen peroxide (H(2)O(2))-induced apoptosis through targeting the mitochondria apoptotic pathway. PLoS ONE 2012, 7, e44907. [Google Scholar] [CrossRef]
  224. Yang, L.; Wang, B.; Zhou, Q.; Wang, Y.; Liu, X.; Liu, Z.; Zhan, Z. MicroRNA-21 prevents excessive inflammation and cardiac dysfunction after myocardial infarction through targeting KBTBD7. Cell Death Dis. 2018, 9, 769. [Google Scholar] [CrossRef] [PubMed]
  225. Gu, W.; Zhan, H.; Zhou, X.Y.; Yao, L.; Yan, M.; Chen, A.; Liu, J.; Ren, X.; Zhang, X.; Liu, J.X.; et al. MicroRNA-22 regulates inflammation and angiogenesis via targeting VE-cadherin. FEBS Lett. 2017, 591, 513–526. [Google Scholar] [CrossRef]
  226. Yuan, H.; Du, S.; Deng, Y.; Xu, X.; Zhang, Q.; Wang, M.; Wang, P.; Su, Y.; Liang, X.; Sun, Y.; et al. Effects of microRNA-208a on inflammation and oxidative stress in ketamine-induced cardiotoxicity through Notch/NF-kappaB signal pathways by CHD9. Biosci. Rep. 2019, 39. [Google Scholar] [CrossRef]
  227. Valkov, N.; King, M.E.; Moeller, J.; Liu, H.; Li, X.; Zhang, P. MicroRNA-1-Mediated Inhibition of Cardiac Fibroblast Proliferation Through Targeting Cyclin D2 and CDK6. Front. Cardiovasc. Med. 2019, 6, 65. [Google Scholar] [CrossRef]
  228. Tijsen, A.J.; van der Made, I.; van den Hoogenhof, M.M.; Wijnen, W.J.; van Deel, E.D.; de Groot, N.E.; Alekseev, S.; Fluiter, K.; Schroen, B.; Goumans, M.J.; et al. The microRNA-15 family inhibits the TGFbeta-pathway in the heart. Cardiovasc. Res. 2014, 104, 61–71. [Google Scholar] [CrossRef] [PubMed]
  229. Zhong, C.; Wang, K.; Liu, Y.; Lv, D.; Zheng, B.; Zhou, Q.; Sun, Q.; Chen, P.; Ding, S.; Xu, Y.; et al. miR-19b controls cardiac fibroblast proliferation and migration. J. Cell Mol. Med. 2016, 20, 1191–1197. [Google Scholar] [CrossRef] [PubMed]
  230. Yuan, J.; Chen, H.; Ge, D.; Xu, Y.; Xu, H.; Yang, Y.; Gu, M.; Zhou, Y.; Zhu, J.; Ge, T.; et al. miR-21 Promotes Cardiac Fibrosis After Myocardial Infarction Via Targeting Smad7. Cell Physiol. Biochem. 2017, 42, 2207–2219. [Google Scholar] [CrossRef] [PubMed]
  231. Cao, W.; Shi, P.; Ge, J.J. miR-21 enhances cardiac fibrotic remodeling and fibroblast proliferation via CADM1/STAT3 pathway. BMC Cardiovasc. Disord. 2017, 17, 88. [Google Scholar] [CrossRef]
  232. Zhou, X.L.; Xu, H.; Liu, Z.B.; Wu, Q.C.; Zhu, R.R.; Liu, J.C. miR-21 promotes cardiac fibroblast-to-myofibroblast transformation and myocardial fibrosis by targeting Jagged1. J. Cell Mol. Med. 2018, 22, 3816–3824. [Google Scholar] [CrossRef] [PubMed]
  233. Nagpal, V.; Rai, R.; Place, A.T.; Murphy, S.B.; Verma, S.K.; Ghosh, A.K.; Vaughan, D.E. MiR-125b Is Critical for Fibroblast-to-Myofibroblast Transition and Cardiac Fibrosis. Circulation 2016, 133, 291–301. [Google Scholar] [CrossRef]
  234. Yuan, X.; Pan, J.; Wen, L.; Gong, B.; Li, J.; Gao, H.; Tan, W.; Liang, S.; Zhang, H.; Wang, X. MiR-144-3p Enhances Cardiac Fibrosis After Myocardial Infarction by Targeting PTEN. Front. Cell Dev. Biol. 2019, 7, 249. [Google Scholar] [CrossRef] [PubMed]
  235. Liao, Y.; Li, H.; Cao, H.; Dong, Y.; Gao, L.; Liu, Z.; Ge, J.; Zhu, H. Therapeutic silencing miR-146b-5p improves cardiac remodeling in a porcine model of myocardial infarction by modulating the wound reparative phenotype. Protein Cell 2021, 12, 194–212. [Google Scholar] [CrossRef] [PubMed]
  236. Wei, Y.; Yan, X.; Yan, L.; Hu, F.; Ma, W.; Wang, Y.; Lu, S.; Zeng, Q.; Wang, Z. Inhibition of microRNA-155 ameliorates cardiac fibrosis in the process of angiotensin II-induced cardiac remodeling. Mol. Med. Rep. 2017, 16, 7287–7296. [Google Scholar] [CrossRef] [PubMed]
  237. Sun, M.; Yu, H.; Zhang, Y.; Li, Z.; Gao, W. MicroRNA-214 Mediates Isoproterenol-induced Proliferation and Collagen Synthesis in Cardiac Fibroblasts. Sci. Rep. 2015, 5, 18351. [Google Scholar] [CrossRef]
  238. Liu, X.; Xu, Y.; Deng, Y.; Li, H. MicroRNA-223 Regulates Cardiac Fibrosis After Myocardial Infarction by Targeting RASA1. Cell Physiol. Biochem. 2018, 46, 1439–1454. [Google Scholar] [CrossRef]
  239. Iaconetti, C.; Polimeni, A.; Sorrentino, S.; Sabatino, J.; Pironti, G.; Esposito, G.; Curcio, A.; Indolfi, C. Inhibition of miR-92a increases endothelial proliferation and migration in vitro as well as reduces neointimal proliferation in vivo after vascular injury. Basic. Res. Cardiol. 2012, 107, 296. [Google Scholar] [CrossRef] [PubMed]
  240. Li, P.; Yin, Y.L.; Guo, T.; Sun, X.Y.; Ma, H.; Zhu, M.L.; Zhao, F.R.; Xu, P.; Chen, Y.; Wan, G.R.; et al. Inhibition of Aberrant MicroRNA-133a Expression in Endothelial Cells by Statin Prevents Endothelial Dysfunction by Targeting GTP Cyclohydrolase 1 in Vivo. Circulation 2016, 134, 1752–1765. [Google Scholar] [CrossRef]
  241. Bernardo, B.C.; Gao, X.M.; Winbanks, C.E.; Boey, E.J.; Tham, Y.K.; Kiriazis, H.; Gregorevic, P.; Obad, S.; Kauppinen, S.; Du, X.J.; et al. Therapeutic inhibition of the miR-34 family attenuates pathological cardiac remodeling and improves heart function. Proc. Natl. Acad. Sci. USA 2012, 109, 17615–17620. [Google Scholar] [CrossRef]
  242. Huang, Y.; Qi, Y.; Du, J.Q.; Zhang, D.F. MicroRNA-34a regulates cardiac fibrosis after myocardial infarction by targeting Smad4. Expert Opin. Ther. Targets 2014, 18, 1355–1365. [Google Scholar] [CrossRef] [PubMed]
  243. Yuan, J.; Liu, H.; Gao, W.; Zhang, L.; Ye, Y.; Yuan, L.; Ding, Z.; Wu, J.; Kang, L.; Zhang, X.; et al. MicroRNA-378 suppresses myocardial fibrosis through a paracrine mechanism at the early stage of cardiac hypertrophy following mechanical stress. Theranostics 2018, 8, 2565–2582. [Google Scholar] [CrossRef]
  244. Goto, S.; Ichihara, G.; Katsumata, Y.; Ko, S.; Anzai, A.; Shirakawa, K.; Endo, J.; Kataoka, M.; Moriyama, H.; Hiraide, T.; et al. Time-Series Transcriptome Analysis Reveals the miR-27a-5p-Ppm1l Axis as a New Pathway Regulating Macrophage Alternative Polarization After Myocardial Infarction. Circ. J. 2021, 85, 929–938. [Google Scholar] [CrossRef]
  245. Chen, C.; Cai, S.; Wu, M.; Wang, R.; Liu, M.; Cao, G.; Dong, M.; Yiu, K.H. Role of Cardiomyocyte-Derived Exosomal MicroRNA-146a-5p in Macrophage Polarization and Activation. Dis. Markers 2022, 2022, 2948578. [Google Scholar] [CrossRef] [PubMed]
  246. Yang, F.; Chen, Q.; He, S.; Yang, M.; Maguire, E.M.; An, W.; Afzal, T.A.; Luong, L.A.; Zhang, L.; Xiao, Q. miR-22 Is a Novel Mediator of Vascular Smooth Muscle Cell Phenotypic Modulation and Neointima Formation. Circulation 2018, 137, 1824–1841. [Google Scholar] [CrossRef] [PubMed]
  247. Song, G.; Zhu, L.; Ruan, Z.; Wang, R.; Shen, Y. MicroRNA-122 promotes cardiomyocyte hypertrophy via targeting FoxO3. Biochem. Biophys. Res. Commun. 2019, 519, 682–688. [Google Scholar] [CrossRef]
  248. Wang, G.; Wang, R.; Ruan, Z.; Liu, L.; Li, Y.; Zhu, L. MicroRNA-132 attenuated cardiac fibrosis in myocardial infarction-induced heart failure rats. Biosci. Rep. 2020, 40, BSR20201696. [Google Scholar] [CrossRef]
  249. Cheng, Y.; Liu, X.; Yang, J.; Lin, Y.; Xu, D.Z.; Lu, Q.; Deitch, E.A.; Huo, Y.; Delphin, E.S.; Zhang, C. MicroRNA-145, a novel smooth muscle cell phenotypic marker and modulator, controls vascular neointimal lesion formation. Circ. Res. 2009, 105, 158–166. [Google Scholar] [CrossRef]
  250. Seok, H.Y.; Chen, J.; Kataoka, M.; Huang, Z.P.; Ding, J.; Yan, J.; Hu, X.; Wang, D.Z. Loss of MicroRNA-155 protects the heart from pathological cardiac hypertrophy. Circ. Res. 2014, 114, 1585–1595. [Google Scholar] [CrossRef]
  251. Das, S.; Kohr, M.; Dunkerly-Eyring, B.; Lee, D.I.; Bedja, D.; Kent, O.A.; Leung, A.K.; Henao-Mejia, J.; Flavell, R.A.; Steenbergen, C. Divergent Effects of miR-181 Family Members on Myocardial Function Through Protective Cytosolic and Detrimental Mitochondrial microRNA Targets. J. Am. Heart Assoc. 2017, 6, e004694. [Google Scholar] [CrossRef]
  252. Ucar, A.; Gupta, S.K.; Fiedler, J.; Erikci, E.; Kardasinski, M.; Batkai, S.; Dangwal, S.; Kumarswamy, R.; Bang, C.; Holzmann, A.; et al. The miRNA-212/132 family regulates both cardiac hypertrophy and cardiomyocyte autophagy. Nat. Commun. 2012, 3, 1078. [Google Scholar] [CrossRef] [PubMed]
  253. Chen, H.; Lv, L.; Liang, R.; Guo, W.; Liao, Z.; Chen, Y.; Zhu, K.; Huang, R.; Zhao, H.; Pu, Q.; et al. miR-486 improves fibrotic activity in myocardial infarction by targeting SRSF3/p21-Mediated cardiac myofibroblast senescence. J. Cell Mol. Med. 2022, 26, 5135–5149. [Google Scholar] [CrossRef]
Figure 1. Cardiac cell types and microenvironment. Schematic representation of myocardial tissue architecture, which is mainly supported by ECM proteins. The most relevant cell types present in the adult steady-state heart are shown. Created with BioRender.com.
Figure 1. Cardiac cell types and microenvironment. Schematic representation of myocardial tissue architecture, which is mainly supported by ECM proteins. The most relevant cell types present in the adult steady-state heart are shown. Created with BioRender.com.
Ijms 24 13268 g001
Figure 2. Pro-proliferative miRNAs regulating heart development. Schematic representation of key signaling pathways involved in CMs regulation. Relevant miRNAs involved in cell proliferation are shown. Black: embryonic/neonatal; Blue: neonatal adult; Red: adult. Created with BioRender.com.
Figure 2. Pro-proliferative miRNAs regulating heart development. Schematic representation of key signaling pathways involved in CMs regulation. Relevant miRNAs involved in cell proliferation are shown. Black: embryonic/neonatal; Blue: neonatal adult; Red: adult. Created with BioRender.com.
Ijms 24 13268 g002
Figure 3. Anti-proliferative miRNAs regulating heart development. Schematic representation of the key signaling pathways involved in the regulation of CMs. Relevant miRNAs involved in the anti-proliferative, pro-apoptotic effect are shown. Black: embryonic/neonatal; Blue: neonatal adult; Red: adult. Created with BioRender.com.
Figure 3. Anti-proliferative miRNAs regulating heart development. Schematic representation of the key signaling pathways involved in the regulation of CMs. Relevant miRNAs involved in the anti-proliferative, pro-apoptotic effect are shown. Black: embryonic/neonatal; Blue: neonatal adult; Red: adult. Created with BioRender.com.
Ijms 24 13268 g003
Figure 4. MI etiopathogenesis. Schematic representation of the main risk factors for MI divided into atherosclerotic dependent (left panel) and non-atherosclerotic causes (right panel). Created with BioRender.com.
Figure 4. MI etiopathogenesis. Schematic representation of the main risk factors for MI divided into atherosclerotic dependent (left panel) and non-atherosclerotic causes (right panel). Created with BioRender.com.
Ijms 24 13268 g004
Figure 5. The inflammation phase. Schematic representation of the key cell subsets involved in the post-MI inflammatory phase, including neutrophils and M1-macrophages, as well as pro-inflammatory cytokines and chemokines. miRNAs regulate various aspects of this phase: cardiomyocyte apoptosis, inflammation, and fibrosis, both by activating CFs and by stimulating the fibroblast-to-myofibroblasts switch. The miRNAs involved in the regulation of the different processes are listed in the boxes. Created with BioRender.com.
Figure 5. The inflammation phase. Schematic representation of the key cell subsets involved in the post-MI inflammatory phase, including neutrophils and M1-macrophages, as well as pro-inflammatory cytokines and chemokines. miRNAs regulate various aspects of this phase: cardiomyocyte apoptosis, inflammation, and fibrosis, both by activating CFs and by stimulating the fibroblast-to-myofibroblasts switch. The miRNAs involved in the regulation of the different processes are listed in the boxes. Created with BioRender.com.
Ijms 24 13268 g005
Figure 6. The resolution phase. Schematic representation of the main events of the resolution phase regarding Treg recruitment, the anti-inflammatory polarization of macrophages, the release of anti-inflammatory cytokines, the deposition of ECM and the initiation of neo-angiogenesis. The miRNAs involved in the regulation of the different processes are listed in the boxes. Created with BioRender.com.
Figure 6. The resolution phase. Schematic representation of the main events of the resolution phase regarding Treg recruitment, the anti-inflammatory polarization of macrophages, the release of anti-inflammatory cytokines, the deposition of ECM and the initiation of neo-angiogenesis. The miRNAs involved in the regulation of the different processes are listed in the boxes. Created with BioRender.com.
Ijms 24 13268 g006
Figure 7. The maturation phase. Schematic representation of the maturation phase in which the ECM is stabilized with the release of collagen and matrix proteins, the neo vessels are coated with mural cells, to strengthen them, and part of the fibroblasts are transformed into matrifibroblasts involved in the scar stability. Adverse outcomes such as cardiac remodeling are regulated at this stage by miRNAs. Created with BioRender.com.
Figure 7. The maturation phase. Schematic representation of the maturation phase in which the ECM is stabilized with the release of collagen and matrix proteins, the neo vessels are coated with mural cells, to strengthen them, and part of the fibroblasts are transformed into matrifibroblasts involved in the scar stability. Adverse outcomes such as cardiac remodeling are regulated at this stage by miRNAs. Created with BioRender.com.
Ijms 24 13268 g007
Table 1. Cardiomyocytes regulation: mechanistic insights on pro-proliferative miRNAs.
Table 1. Cardiomyocytes regulation: mechanistic insights on pro-proliferative miRNAs.
miRNAModelTargetsPathwayEffectsMechanism(s)Ref.
miR-10bHuLats1Hippo↓apoptosismiR-10b blocks the expression of Lats1 and thereby inhibits the Hippo pathway[45]
miR-17-92M; RPTENAkt↑proliferationmiR-17-92 represses PTEN-inducing cardiomyocyte proliferation mediated by the AKt signaling pathway[33]
miR19a/19bMPTENAkt↑proliferationmiR-19a/19b blocks the expression of PTEN inducing cardiomyocyte proliferation mediated by the AKt signaling [34]
miR-21-5pHuPDCD4/
PTEN
Akt↑proliferationmiR-21-5p directly represses PCDN4 e PTEN-inducing cardiomyocyte proliferation mediated by the AKt signaling [46]
miR-24Rp27Cell cycle↑proliferationmiR-24 promotes cells in G0/G1 phase into S phase by repressing p27 expression[47]
miR-25Hu; ZbFBXW7Cell cycle↑proliferationmiR-25 inhibits FBXW7, a cell-cycle regulatory factor that mediates the proteolysis of positive cell-cycle regulators[48]
miR-101aHuTGFBR1MAPK↑proliferationmiR-101a blocks the expression of TGFBR1 and thereby activates the MAPK signaling pathway[49]
miR-145-5pRGIGYF1/↓apoptosismiR-145-5p inhibits GIGYF1, a regulator of mRNA turnover, blocking translation and activating transcript decay[50]
miR-152Mp27Hippo/The increase of miR-152 by the activation of YAP1 represses the expression of p27[51]
miR-199a-3pRCD151MAPK↑proliferationmiR-199a-3p inhibits the tetraspanin CD151, involved in several cellular process[52]
miR-199a-3pRTAOK1Hippo↓apoptosismiR-199a-3p represses Serine/threonine-protein kinase TAOK1 involved in the Hippo pathway[53]
miR-199a-3p;
-302d; -373;
-590-3p; -1825
RSTK38LHippo↓apoptosis
↑proliferation
They downregulate STK38L, a direct targeting on the 3’UTR of the related mRNA has not been identified[53]
miR-200a-3pHu; MPDCD4Akt↑proliferationmiR-200a-3p represses PCDN4 activating AKt signaling[54]
miR-204RJarid2/↑proliferationmiR-204 downregulates Jarid2, a non-catalytic member of the polycomb repressive complex 2[42]
miR-210MAPC/β-cateninWNT↑proliferationmiR-210 represses the cell cycle inhibitor APC, involved in the WNT pathway[55]
miR-221-3pRPTENAkt↑proliferationIt enhances Akt kinase activity by inhibiting PTEN[35]
miR-294MWee1Cell cycle/miR-294 inhibits Wee1, increasing the activity of the cyclin B1/CDK1[56]
miR-301aM; RPTENAkt↑proliferationPTEN is a direct target of miR-301a in regulating cardiomyocyte proliferation[57]
miR-302d;
-373; -590-3p
HuLATS2Hippo↓apoptosis
↑proliferation
miR-302d, miR-373 and miR590-3p target LATS2, inhibiting the Hippo pathway[58]
[53]
miR-324HuSOCS3NFkB↑proliferation
↓apoptosis
miR-32a represses SOCS3, activating the NFkB pathway[38]
miR-362-3pRTP53INP2SDF1/
CXCR4
↑proliferation
↓apoptosis
miR-362-3p downregulates TP53INP2, activating at the same time the SDF-1/CXCR4 pathway[59]
miR-374MDTNANotch↑viability
↓apoptosis
miR-374 inhibits DTNA and the Notch1 axis[60]
miR-410; -495M; RCited2Cell cycle↑proliferationmiR-410 and miR-495 downregulate the transcriptional coactivator Cited2 [61]
miR-410-3p HuTRAF5NFkB↓apoptosismiR-410-3p protects CMs from apoptosis by repressing TRAF5 expression[62]
miR-496RHOOK3Akt↓apoptosis
↑proliferation
miR-496 binds to Hook3 to inhibit the activation of PI3K/Akt/mTOR signalling pathway[63]
miR-499M; RSOX6Cell cycle↑proliferation
↓apoptosis
miR-499 promotes cell proliferation and inhibits apoptosis in the late stage of cardiac differentiation targeting SOX6 and leading to activation of cyclin D1[29]
miR-708M; RMAPK14MAPK↑proliferationmiR-708 inhibits the expression of MAPK14, which arrests the cell cycle in CMs[64]
miR-877-3pMGADD45gCell cycle↑proliferationLED-Red irradiation increases miR-877-3p expression promoting proliferation of neonatal CMs via GADD45g[65]
miR-1180RNKIRAS2NFkB↑proliferation
↑viability
↓apoptosis
miR-1180 represses the NF-κB inhibitor interacting with Ras-like 2 (NKIRA2), activating the NFkB pathway[39]
Abbreviations: Hu: human; M: murine; R: rat; Zb: zebrafish.; ↑: upregulation/increase; ↓: downregulation/decrease.
Table 2. Cardiomyocytes regulation: mechanistic insights on anti-proliferative/pro-apoptotic miRNAs.
Table 2. Cardiomyocytes regulation: mechanistic insights on anti-proliferative/pro-apoptotic miRNAs.
miRNAModelTargetsPathwayEffectsMechanism(s)Ref.
Let-71-5pME2F2–CCND2Cell cycle↓proliferationLet-7i-5p downregulates the expression of E2F2 and CCND2 and further represses CM proliferation[66]
miR-1MCCND1Cell cycle↓proliferationIt directly suppresses the cell-cycle regulator CCND1[67]
miR-1RNOTCH3Notch↑apoptosis
↓proliferation
↓autophagy
miR-1 suppresses NOTCH3 thereby negatively regulating the Notch pathway[68]
miR-9RYAP1Hippo↑apoptosismiR-9 promotes hypoxia-induced cardiomyocyte apoptosis by targeting Yap1[69]
miR-26aR, ZbEzh2/↓proliferationmiR-26a targets activators of the cell cycle and Ezh2, a component of PRC2, with repressive functions on negative regulators of the cell cycle[43]
miR29aRCCND2Cell cycle↓proliferationmiE29a inhibits proliferation by acting on cyclin D2[70]
miR-29a;
-30a; -141
RCCNA2–CDK6Cell cycle↓proliferationmiR-29a, miR-30a and miR-141 reduce expression of Cyclin A2, reducing cell proliferation[71]
miR29b-3pR; ZbNOTCH2Notch↓proliferationmiR29b-3p inhibits CMs proliferation by targeting NOTCH2[72]
miR-33a-5pHuNOMO1Nodal↓proliferation
↑apoptosis
↓differentiation
miR-33a-5p targets NOMO1, a component of a protein complex involved in Nodal signaling leading to apoptosis[41]
miR-34aM; RBcl2, Sirt1, CCND1/↓proliferation
↑apoptosis
miR-34a directly regulated cell cycle and death via modulation of its targets such as Bcl2, Cyclin D1, and Sirt1[73]
miR-128RSUZ12Cell cycle↓proliferationmiR-128 downregulates SUZ12, which cannot suppress p27 expression and activate the positive cell cycle regulators Cyclin E and CDK2[74]
miR-133aMCCND2
SRF
Cell cycle
Akt
↓proliferationmiR-133 functions as an inhibitor of cardiomyocyte proliferation, targeting cyclin D2 and SRF.[75]
miR-143-3pMYAP–Ctnnd1Hippo↑apoptosismiR-143-3p inhibited the mitosis of CMs, targeting YAP and Ctnnd1[76]
miR-144RTBX1Jak2/Stat1↑apoptosismiR-144 leads to proliferation via TBX1/JAK2/STAT1 axis[77]
miR-195MChek1Cell cycle↓proliferationmiR-195 regulates the expression of many cell cycle genes, including Chek1[30]
miR-208aRPI3KAkt↑apoptosis
↑autophagy
miR-208a inhibits the PI3K/AKT pathway[36]
miR-322;
-503
MIGF2
FGF9
Akt↓proliferationThey suppress FGF9 and IGF2, disrupting epicardial signaling and impairing myocardial growth[44]
miR-423RMYBL2WNT↑apoptosismiR-423 silences MYBL2 with suppression of wnt/β-[78]
miR-484RYAP1Hippo↑apoptosis
↑inflammation
Catenin signaling pathway[79]
miR-489RSPIN1Akt↑apoptosisIt directly targets Yap1mRNA to inhibit cell viability[37]
miR-873RGLI1Hedgehog↓proliferationIt inhibits SPIN1 thereby inactivating the Akt pathway[80]
Abbreviations: Hu: human; M: murine; R: rat; Zb: zebrafish.; ↑: upregulation/increase; ↓: downregulation/decrease.
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

Salvatori, F.; D’Aversa, E.; Serino, M.L.; Singh, A.V.; Secchiero, P.; Zauli, G.; Tisato, V.; Gemmati, D. miRNAs Epigenetic Tuning of Wall Remodeling in the Early Phase after Myocardial Infarction: A Novel Epidrug Approach. Int. J. Mol. Sci. 2023, 24, 13268. https://doi.org/10.3390/ijms241713268

AMA Style

Salvatori F, D’Aversa E, Serino ML, Singh AV, Secchiero P, Zauli G, Tisato V, Gemmati D. miRNAs Epigenetic Tuning of Wall Remodeling in the Early Phase after Myocardial Infarction: A Novel Epidrug Approach. International Journal of Molecular Sciences. 2023; 24(17):13268. https://doi.org/10.3390/ijms241713268

Chicago/Turabian Style

Salvatori, Francesca, Elisabetta D’Aversa, Maria Luisa Serino, Ajay Vikram Singh, Paola Secchiero, Giorgio Zauli, Veronica Tisato, and Donato Gemmati. 2023. "miRNAs Epigenetic Tuning of Wall Remodeling in the Early Phase after Myocardial Infarction: A Novel Epidrug Approach" International Journal of Molecular Sciences 24, no. 17: 13268. https://doi.org/10.3390/ijms241713268

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