Minireviews Open Access
Copyright ©The Author(s) 2016. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Cardiol. Feb 26, 2016; 8(2): 211-219
Published online Feb 26, 2016. doi: 10.4330/wjc.v8.i2.211
Novel epigenetic-based therapies useful in cardiovascular medicine
Claudio Napoli, Vincenzo Grimaldi, Maria Rosaria De Pascale, Linda Sommese, Division of Immunohematology, Transfusion Medicine and Transplant Immunology, Department of Internal Medicine and Specialistic Unit, Azienda Ospedaliera Universitaria (AOU), Second University of Naples, 80138 Naples, Italy
Claudio Napoli, Teresa Infante, Andrea Soricelli, Institute of Diagnostic and Nuclear Development (SDN), IRCCS, 80143 Naples, Italy
Author contributions: Napoli C, Grimaldi V, De Pascale MR, Infante T, Soricelli A and Sommese L scrutinized the scientific literature; Napoli C, Grimaldi V, De Pascale MR, Soricelli A and Sommese L wrote the paper.
Conflict-of-interest statement: No potential conflicts of interest. No financial support.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Vincenzo Grimaldi, MB, BiolD, Division of Immunohematology, Transfusion Medicine and Transplant Immunology, Department of Internal Medicine and Specialistic Unit, Azienda Ospedaliera Universitaria (AOU), Second University of Naples, Piazza L. Miraglia 2, 80138 Naples, Italy. vincenzo.grimaldi@policliniconapoli.it
Telephone: +39-081-5665067 Fax: +39-081-5665092
Received: July 30, 2015
Peer-review started: July 31, 2015
First decision: September 16, 2015
Revised: December 11, 2015
Accepted: December 16, 2015
Article in press: December 18, 2015
Published online: February 26, 2016

Abstract

Epigenetic modifications include DNA methylation, histone modifications, and microRNA. Gene alterations have been found to be associated with cardiovascular diseases, and epigenetic mechanisms are continuously being studied to find new useful strategies for the clinical management of afflicted patients. Numerous cardiovascular disorders are characterized by the abnormal methylation of CpG islands and so specific drugs that could inhibit DNA methyltransferase directly or by reducing its gene expression (e.g., hydralazine and procainamide) are currently under investigation. The anti-proliferative and anti-inflammatory properties of histone deacetylase inhibitors and their cardio-protective effects have been confirmed in preclinical studies. Furthermore, the regulation of the expression of microRNA targets through pharmacological tools is still under development. Indeed, large controlled trials are required to establish whether current possible candidate antisense microRNAs could offer better therapeutic benefits in clinical practice. Here, we updated therapeutic properties, side effects, and feasibility of emerging epigenetic-based strategies in cardiovascular diseases by highlighting specific problematic issues that still affect the development of large scale novel therapeutic protocols.

Key Words: Epigenetics, Cardiovascular diseases, Heart failure, Inhibitors of histone deacetylases, Antisense microRNAs

Core tip: Recent evidence suggests that specific epigenetic regulatory mechanisms play key roles in cardiac differentiation, homeostasis, injury response, and disease development. Drug therapies that work via epigenetic mechanisms are currently limited to antineoplastic agents; large controlled trials are required to establish whether current possible candidate antisense microRNAs or histone deacetylase inhibitors could offer better therapeutic benefits in cardiovascular disease. We review recent findings on the epigenetic control of several cardiovascular diseases and the new challenges for therapeutic strategies in cardiovascular diseases.



INTRODUCTION

Cardiovascular diseases (CVDs) are the primary cause of death worldwide, with 17.5 million deaths from CVD in 2012 representing 31% of all global deaths that year. CVDs include a number of alterations affecting heart and vascular structures, such as heart valves, heart muscle (e.g., cardiomyopathy), and pericardial and coronary artery diseases. All these conditions may result in cardiomyocyte loss, cardiac-remodeling with consequent heart failure (HF), and an increased risk of arrhythmias and death. Cardiac fibroblasts also have a pivotal role in HF[1]. Indeed, endothelial cell activation and inflammation promotes the transdifferentiation of fibroblasts to myofibroblasts which, after extensive collagen production, results in the release of chemokines and the activation of inflammatory cells, which in turn causes cardiomyocyte stiffness by contributing to HF pathogenesis[1]. Increasing evidence has shown that epigenetic mechanisms control and influence the expression of cell cycle central genes involved in human disease progression[2]. Toward this context, significant epigenetic and epigenomic findings have opened a new area of research by exploring the role of genetic heritability and environmental interaction in CVDs[3]. Some deregulated epigenetic steps are involved in the pathophysiology of CVDs[4]. Specific epigenetic regulatory mechanisms could impact on the endothelium, cardiac muscle, smooth muscle, and fibroblasts[5]. Thus, the pharmacological setting of these pathways might represent a specific target for CVDs. Due to the reversible nature of these modifications, researchers are continuously engaged in the development of novel epigenetic-based drugs (epidrugs) for CVD treatment[6,7]. The primary goal of future studies will be to allow for the identification of selective therapeutic molecules that have been conceived in order to act on specific epigenetic-related pathogenic events.

Here, we summarize the current knowledge concerning epigenetic-based strategies in CVDs by outlining novel therapeutic steps in clinical practice.

EPIGENETICS INVOLVEMENT IN CVDs
DNA methylation as therapeutic target

DNA methylation is the most studied epigenetic modification and mainly involves methylation of CpG islands in the promoter genes. It has good long-term stability and is the most common modification involving the regulation of gene expression in the mammalian genome. All changes in methylation are modulated by specific catalytically-active enzymes, including “maintenance” methyltransferase (DNMT1) and “de novo” methyltransferase (DNMT3a and DNMT3b). DNMTs act by adding methyl groups to CpG residues, thereby modifying the accessibility of DNA to the transcriptional machinery. Altered regulation of cytosine methylation has been linked to CVD development and progression[8], as well as to cancer cell development[9]. In addition, DNA methylation has been shown to regulate biological processes underlying CVDs, such as atherosclerosis, inflammation, hypertension, and diabetes[10,11]. DNA methylation is also involved in essential arterial hypertension[12,13]. To date, DNA methylation remains an attractive target for CVD interventions, owing to its reversible nature. Dietary compounds, including polyphenols and catechins, act on DNA methylation processes[14,15]. In particular, some interesting clinical studies have shown that elevated consumption of polyphenols decreases global DNA methylation of peripheral leukocytes in humans with cardiovascular risk factors (NCT00511420 and NCT00502047)[16]. However, the role of nutrients in the evolvement of CVDs through the epigenetic link remains as yet studied. Conversely, the cardiovascular implication of pharmacological epigenetic compounds appears to be more direct and far-reaching. Indeed, some drugs are known to affect DNA methylation. Hydralazine, a vasodilator used to treat hypertension[17], is an example of compound that has been shown to inhibit DNA methyltransferase directly or by reducing its gene expression[18]. There are several clinical trials focusing on the use of hydralazine to combat hypertensive conditions (Table 1). Many of these completed trials have highlighted the beneficial effect of hydralazine on both hypertension and other cardiovascular conditions compared to other compounds. Several studies have shown that hydralazine might function by modulating the effect of purine-like compounds released from sympathetic nerve endings and/or by inducing an altered Ca2+ balance in vascular smooth muscle cells[19,20]; unfortunately, there are fundamental as-yet unresolved issues concerning this area of research that remain unclarified.

Table 1 Interventional and randomized ongoing clinical trials on the use of hydralazine and procainamide in cardiovascular diseases.
No.StatusConditionNo. of enrolled patientsIntervention
NCT00684489CompletedHypertension52Hydralazine and other drugs
NCT02305095Not open for participant recruitmentHeart failure500 (estimated enrollment)Hydralazine in combination with isosorbide dinitrate
NCT00661895CompletedHypertension99Hydralazine and other drugs
NCT00599235CompletedHypertension30Hydralazine, sildenafil, and placebo
NCT00223717Recruiting participantsHypertension160 (estimated enrollment)Hydralazine and other drugs
NCT01255475CompletedHeart failure, cardiac failure, and congestive heart failure21Hydralazine/amlodipine and placebo
NCT01516346Recruiting participantsHeart failure and congestive heart failure54 (estimated enrollment)Hydralazine, isosorbide dinitrate, and placebo
NCT01822808Recruiting participantsAcute heart failure and left ventricular dysfunction500 (estimated enrollment)Hydralazine, isosorbide dinitrate, and placebo
NCT00000499CompletedCardiovascular diseases, heart diseases, hypertension, and vascular diseasesNot providedHydralazine, reserpine, chlorthalidone, and metoprolol
NCT02050529Recruiting participantsHypertension, Pregnancy induced180 (estimated enrollment)Hydralazine, labetalol
NCT01538875CompletedHypertension, Pregnancy induced261Hydralazine, labetalol
NCT00383799UnknownVentricular tachycardia302 (estimated enrollment)Procainamide, amiodarone
NCT00000464CompletedArrhythmia, Cardiovascular diseases115Procainamide, quinidine, disopyramide, and other drugs
NCT00702117CompletedAtrial fibrillation, tachycardia123Procainamide, ajmaline, flecainide
NCT00589303TerminatedAtrial fibrillation, heart failure27Rhythm control drugs: Procainamide and other drugs
NCT00000556CompletedArrhythmia, atrial fibrillation, cardiovascular diseases4060Procainamide and other drugs
NCT01205529RecruitingAtrial fibrillation750 (estimated enrollment)Procainamide

Procainamide is another drug that inhibits DNA methyltransferase I. It is a sodium channel blocker that belongs to the class of benzamides used against arrhythmias[21]. Clinical trials have evaluated this anti-arrhythmic drug in the acute treatment of monomorphous ventricular tachycardia with positive effects (Table 1). Nevertheless, recent evidence has shown toxic effects of procainamide on the lung after orthotopic cardiac transplantation[22].

Despite the use of the aforementioned drugs in cancer treatment appearing to have promising results[23], the implication of their epigenetic effects in CVDs requires further investigation in future studies.

Histone modifications as therapeutic target

Epigenetic alterations occur in the histone code, and so can modulate histone-DNA interactions and significantly influence chromatin structure by modifying the accessibility of transcriptional regulators to DNA-binding elements[24]. The most common modifications are lysine acetylation and methylation, arginine methylation, and serine phosphorylation. Histone acetylation is catalyzed by histone acetyltransferases (HATs), while histone deacetylation is carried out by histone deacetylases (HDACs)[25].

Inhibitors of histone deacetylases (HDACi) represent a significant group of epidrugs that could be highly relevant to the treatment of CVDs. Indeed, HDACi exert anti-proliferative and anti-inflammatory effects, and their cardio-protective therapeutic use has been recently confirmed in preclinical studies[26,27].

According to their chemistry, HDACi can be divided into four main groups: Hydroxamates, aliphatic acids, benzamides, and cyclic peptides. Hydroxamates like trichostatin A (TSA) and vorinostat (suberoylanilide hydroxamic acid, SAHA) serve as pan-HDACi and are generally most often used for preclinical studies[28-30].

Principal histone modifications and therapeutic targets involved in CVDs are reported in Table 2. Animal studies in vivo showed that TSA treatment improved functional myocardial recovery after myocardial infarction (MI) via a reduction in myocardial and serum tumor necrosis factor-α. Neo-angiogenesis was demonstrated in MI animals after receiving TSA treatment[31]. Taken together, these results indicate that HDACi could preserve cardiac performance and mitigate myocardial remodeling by stimulating endogenous cardiac regeneration[31]. HDAC inhibition was also shown to attenuate ischemic injury in the heart and other tissues. Pre-treatment with TSA resulted in improvements in post-ischemic ventricular function, with a reduction in infarct size in both early and delayed preconditioning models[32]. Despite the high activity of TSA, it was disqualified as a clinical drug due to its many side effects, such as non-transformed cell apoptosis and increased DNA damage[33].

Table 2 Histone modifications and therapeutic targets involved in cardiovascular diseases.
TargetEpigenetic mechanismsConditionOrganism/in vitro, in vivoEffectsRef.
TSAInhibition of HDAC4Ischemic injuryMouse, in vitro and in vivoHDACi would be predicted to have a beneficial effect in the context of active ischemiaGranger et al[28] (2008)
TSA/VPAClass I HDACsCardiac hypertrophyMouse, in vitro and in vivoTherapeutic target for preventing or reversing cardiac hypertrophy and subsequent heart failureKee et al[29] (2006)
TSAInhibition of HDACsAtrial fibrosis and arrhythmiasMouse, in vitro and in vivoReversed myocardial fibrosisLiu et al[30] (2008)
TSAInhibition of HDACsAcute myocardial ischemia and reperfusion injuryMouse, in vitro and in vivoImproved cardiac functional recovery and antagonized myocardial remodeling in chronic myocardial infarctionZhang et al[31] (2012)
TSA/SAHAHDAC inhibitorMyocardial infarctMouse, rabbit, in vivoReduced infarct size in a large animal modelXie et al[35] (2014)
SAHA/sodium valproateInhibition of HDACsIschemic injuryMouse, in vitro and in vivoPotential therapeutic strategy for restoring compromised cardiac proteostasisWang et al[36] (2011)
VPA or tributyrinInhibition of HDACsInfarctRat, in vitroAttenuated ventricular remodeling after infarctionLee et al[37] (2007)
MS-275AInhibition of class I/II HDACsInfarctRat, in vivoSignificant reduction of infarct area observedAune et al[39] (2014)
ApicidinInhibition of class I HDACsCardiac hypertrophy and heart failureRat pups, in vitro Mouse, in vivoPreserved cardiac function in the long-termGallo et al[42] (2008)
Curcuminp300 HAT inhibitorHeart failureRat, in vitroPrevented deterioration of systolic function and heart failureMorimoto et al[45] (2008)

Vorinostat was approved by the Unites States Food and Drug Administration (FDA) for the treatment of advanced cutaneous T cell lymphoma[34]. Suberoylanilide hydroxamic acid (SAHA/vorinostat) reduced myocardial infarct size in a large animal model, even when delivered in the clinically relevant context of reperfusion[35,36].

Aliphatic acids like valproic acid (VPA, 2-propylpentanoic acid) inhibits class I HDACs, causing accumulation of hyperacetylated histone tails (H3 and H4 histones) and other protein targets such as p53. VPA has anti-proliferative and pro-apoptotic activities. Lee et al[37] demonstrated the attenuation of ventricular remodeling following MI in vivo when VPA or tributyrin was administered to rats 24 h after ligation of the left anterior descending artery. However, these short chain fatty acids are known to weakly inhibit HDAC activity with a number of off-target effects[38].

Benzamides are small molecules that are mostly active against class I HDACs. Class I HDACi is entinostat (MS-275A) and prompts protective effects against ischemia reperfusion injury in isolated rat heart. MS-275A is not effective against class IIb HDAC6[39]. Entinostat might be more advantageous than first-generation examples such as TSA, vorinostat, romidepsin, and VPA, as less profound side effects are observed[40]. Some studies have suggested that tranilast also has cardiovascular-protective effects[41]. Depsipeptide is a natural cyclic peptide that inhibits HDAC 1 and 2, and selectively modulates the expression levels of different genes such as c-myc, Hsp90, and p53. The cyclic peptide family includes other HDACi, such as apicidin. The apicidin derivative API-D is capable of reducing hypertrophy and, consequently, the transition to HF in mice subjected to thoracic aortic constriction. Treatment with this substance therefore establishes a relevant therapeutic approach for HF[42].

The cardiovascular protective effects of p300 HAT inhibitor curcumin have been demonstrated[43,44]. In a rat model of HF and primary cultured rat cardiac myocytes and fibroblasts, curcumin prevented ventricular hypertrophy and preserved systolic function[45].

RNA-based mechanisms as novel biomarkers

MicroRNAs are key regulators of gene expression acting at the post-transcriptional level. MiRNAs are implicated in the pathogenesis of several CVDs[46]. The modulation of miRNA expression could represent an innovative therapeutic approach to the treatment of cardiovascular conditions by targeting a single cell type or specific pathways, as demonstrated in an animal model[47,48]. Recently, several study population have investigated the involvement of transcriptionally regulated miRNAs as an attractive target for the treatment of several cardiovascular conditions (Table 3). Preclinical studies using antisense oligonucleotide (antagomir) -mediated knockdown have demonstrated the role of specific miRNAs in HF[47,49,50]. Indeed, it was shown that a single treatment with the infusion of a miR133 antagomir induced cardiac hypertrophy in mice[49]. Recently, Wahlquist et al[47] demonstrated that high levels of miR25 can depress cardiac function, although the inhibition of this miRNA by anti-miR25 effectively restores cardiac function in an HF mouse model. Interestingly, it was demonstrated that miRNAs secreted by cardiac fibroblasts may also act as mediators of cardiomyocyte hypertrophy via a paracrine mechanism[50]. During hypertension or pathological cardiac hypertrophy, reactivation of fetal cardiac genes such as atrial natriuretic peptide, (ANP)/B-type natriuretic peptide (BNP), and beta-myosin heavy chain (β-MHC) can occur. In a hypertensive mouse model, aldosterone-dependent inhibition of miR-208a can occur, resulting in β-MHC inhibition and an increase of cardiac hypertrophy[51]. It was also shown that therapeutic inhibition of miR-208a led to a reduction in cardiac remodeling, which coincided with a significant improvement in survival and cardiac function during heart disease[48]. Additionally, in hypertensive rat models, changes in β-MHC expression were observed after treatment with anti-miR-208a that acted by reverting the levels of several miRNAs, including miR-16, -19b, and -20b[52]. Recently, the regulation of miR-208a and endoglin in AMI were investigated, with the authors demonstrating that the overexpression of antagomir-208a significantly inhibited the increase of myocardial endoglin and β-MHC protein expression induced by infarction[53]. In addition, pre-treatment with atorvastatin and valsartan, members of a drug class known as statins that are primarily used for the prevention of events associated with cardiovascular disease, can decrease myocardial fibrosis induced by AMI by attenuating miR-208a and endoglin expression[53]. Clinical evidence supports the different levels of miR-143, miR-145, miR-21, miR-133, and miR-1 expression in patients with essential hypertension, suggesting that these miRNAs can act in vascular smooth muscle cell phenotypic modulation and could represent potential therapeutic targets in essential hypertension[54]. It was found that the chronic restoration of miR-1 gene expression in an animal model reverted pressure-induced cardiac hypertrophy and prevented the adverse cardiac remodeling induced by pressure overload[55]. Recently, Han et al[56] found higher levels of miR-29a in patients with hypertension and left ventricular (LV) hypertrophy compared to patients with hypertension alone. MiR-29a levels were significantly associated with collagen type I and III and MMP-9 expression. The same authors, employing a mouse model of pressure overload, have shown that antagomir-29a significantly suppressed the hypertrophy of cardiomyocytes and reduced the expression of ANP and β-MHC, suggesting a possible role of miR-29a as a therapeutic target[56]. Several preclinical studies showed the beneficial effects of antagomir-92a administration on small and large animal models before MI[57-59]. Inhibition of miR-92a by repeated intravenous injections of antagomir-92a induced angiogenesis and improved recovery of ventricular function in MI mouse model[57]. In MI large animal models, antagomir-92a treatment revealed cardio-protection against ischemia/reperfusion[58]. Recent evidence has demonstrated favorable post-ischemic myocardial repair after intravenous administration of antagomir-92a in adult large animal models[59]. Indeed, neovasculogenesis and the prevention of adverse ventricular remodeling, the major cause of contractile dysfunction and HF after MI, were observed after intravenous administration of antagomir-92a[59]. These results reveal a promising therapeutic approach for patients affected by MI. Progression of post-infarction LV remodeling in mice was studied by Tolonen et al[60], who observed that the inhibition of Let-7c was associated with decreased apoptosis, reduced fibrosis, and a reduction in the number of discoidin domain receptor 2-positive fibroblasts, while the number of c-kit+ cardiac stem cells and Ki-67+ proliferating cells remained unaltered[60]. Although Let-7c inhibitor injection improved cardiac function after MI, the safety of Let-7c inhibition has yet to be clarified due to its dualistic function that appears to have a causative role in various cancer diseases.

Table 3 Recent evidence investigating the role of circulating miRNAs as biomarkers in several cardiovascular diseases.
miRNAsSourcesConditionsRef.
↑miR-339-5p, miR-483-3p ↓miR-139-5bPlasmaLVISaddic et al[64] (2015)
↓miR-145PlasmaAMIGao et al[65] (2015)
↑miR-122, miR-140-3p, miR-720, miR-2861, miR-3149PlasmaACS, AMILi et al[66] (2015)
↑Let-7e, miR-15a, miR-196b ↓miR-411PlasmaAAA, AtherosclerosisStather et al[67] (2015)
↓ miR-125b, miR-320bPlasmaAMI, CADHuang et al[68] (2014)
↓miR-21SerumCADFan et al[69] (2014)
↓miR-31PlasmaCADWang et al[70] (2014)
↑miR-146a, miR-186, miR-208b, miR-499SerumACS, Stable CAD, CV riskWu et al[71] (2014)
↑miR-210PBMCHFEndo et al[72] (2013)
↑miR-21, miR-25, miR-92a, miR-106b, miR-126, miR-451, miR-590-5pPlasmaAP, UARen et al[73] (2013)
↔ miR-1, miR-208a, miR-423-5pPlasmaAMI, CADNabialek et al[74] (2013)
↑miR-30a, miR-210SerumHFZhao et al[75] (2013)
↑miR-337-5p, miR-433, miR-485-3p, miR-1, miR-122, miR-126, miR-133a/b, miR-199a ↔miR-17-5p, miR-92a, miR-145, miR-155, miR-208a, miR-375, miR-799-5pPlasmaAP, UAD’Alessandra et al[76] (2013)
↓miR-103, miR-142-3p, miR-30b, miR-342-3pPlasmaHFEllis et al[77] (2013)
↑miR-122, miR-200b, miR-520d-5p, miR-622WB and serumHFVogel et al[78] (2013)
↓miR-558
↑miR-21, miR-133a, miR-423-5p, miR-499-5pPlasmaHF, NSTEMIOlivieri et al[79] (2013)
↔miR-1, miR-208a
↑miR-133aPlasmaAMI, APWang et al[80] (2013)
↓miR-214PlasmaAMI, AP, UALu et al[81] (2013)

Circulating miRNA patterns are analyzed as potential disease specific biomarkers in CVDs in two observational prospective studies on aortic aneurism in hereditary aortopathy syndromes (NCT02213484), coronary artery diseases, and myocardial infarction (NCT02076153). Three interventional randomized studies are focusing on the association between miRNA profile modifications and the administration of specific molecules like anti-platelet agents (NCT02071966) (NCT02447809) in coronary syndromes and anti-diabetics drugs in diabetic stable and unstable angina (NCT01331967).

CONCLUSION

To date, several epidrugs (such as vorinostat and panobinostat) have been approved for the treatment of cancer and myelodysplastic syndromes, and are therefore commercially available. However, no epigenetic drugs for CVDs have yet been actually approved by the FDA. Nevertheless, the opportunity to control genetic and epigenetic processes could be considered a promising and attractive tool in cardiovascular medicine. For this reason, the investigation of epigenetic-related mechanisms might help to explain how environmental and lifestyle factors can influence aberrant gene expression patterns over a lifetime that can result in increased cardiovascular risk. Preclinical experiments have identified some HDACi that could have future implications in the treatment of several cardiovascular conditions, including atrial fibrillation, cardiac hypertrophy, and HF. Ongoing human clinical controlled studies are emphasizing the ability of some drugs such as hydralazine and procainamide to act on DNA methylation in CVDs. However, the clinical experience with HDACi in CVDs is limited due to the observed toxic cardiac side effects in oncologic patients.

The study of the human genome will find biomarkers that might affect CVDs. It is likely that only epigenetic profiles obtained from large cohorts of patients with the same genetic mutations will be able to promote the development of surveillance programs and novel effective drugs for the transition of in vitro to in vivo treatments for the early stage of CVDs[61-63].

To date, few clinical trials have investigated the link between drugs and specific miRNA profiles, which might be considered as biomarkers for the classification of CVDs with scarce compliance to standard therapy and affected by the incidence of more aggressive clinical phenotypes. Unfortunately, antagomir in the area of cardiovascular disease has not yet been tested in clinical trials. However, the promising studies covered here reflect the open debate for possible future applications of miRNA therapeutics in CVDs.

Footnotes

P- Reviewer: Gong KZ, O-Uchi J S- Editor: Kong JX L- Editor: A E- Editor: Lu YJ

References
1.  Duygu B, Poels EM, da Costa Martins PA. Genetics and epigenetics of arrhythmia and heart failure. Front Genet. 2013;4:219.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 45]  [Cited by in F6Publishing: 51]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
2.  Brookes E, Shi Y. Diverse epigenetic mechanisms of human disease. Annu Rev Genet. 2014;48:237-268.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 83]  [Cited by in F6Publishing: 91]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
3.  Ptak C, Petronis A. Epigenetics and complex disease: from etiology to new therapeutics. Annu Rev Pharmacol Toxicol. 2008;48:257-276.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 120]  [Cited by in F6Publishing: 103]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
4.  Maciejak A, Kiliszek M, Michalak M, Tulacz D, Opolski G, Matlak K, Dobrzycki S, Segiet A, Gora M, Burzynska B. Gene expression profiling reveals potential prognostic biomarkers associated with the progression of heart failure. Genome Med. 2015;7:26.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 82]  [Article Influence: 9.1]  [Reference Citation Analysis (0)]
5.  Marín-García J, Akhmedov AT. Epigenetics of the failing heart. Heart Fail Rev. 2015;20:435-459.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 13]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
6.  Heerboth S, Lapinska K, Snyder N, Leary M, Rollinson S, Sarkar S. Use of epigenetic drugs in disease: an overview. Genet Epigenet. 2014;6:9-19.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 183]  [Cited by in F6Publishing: 184]  [Article Influence: 18.4]  [Reference Citation Analysis (0)]
7.  Schiano C, Vietri MT, Grimaldi V, Picascia A, Pascale MR, Napoli C. Epigenetic-related therapeutic challenges in cardiovascular disease. Trends Pharmacol Sci. 2015;36:226-235.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 81]  [Cited by in F6Publishing: 79]  [Article Influence: 8.8]  [Reference Citation Analysis (0)]
8.  Baccarelli A, Tarantini L, Wright RO, Bollati V, Litonjua AA, Zanobetti A, Sparrow D, Vokonas PS, Schwartz J. Repetitive element DNA methylation and circulating endothelial and inflammation markers in the VA normative aging study. Epigenetics. 2010;5:222-228.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 92]  [Cited by in F6Publishing: 94]  [Article Influence: 6.7]  [Reference Citation Analysis (0)]
9.  Kikuyama M, Takeshima H, Kinoshita T, Okochi-Takada E, Wakabayashi M, Akashi-Tanaka S, Ogawa T, Seto Y, Ushijima T. Development of a novel approach, the epigenome-based outlier approach, to identify tumor-suppressor genes silenced by aberrant DNA methylation. Cancer Lett. 2012;322:204-212.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 29]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
10.  Grimaldi V, Vietri MT, Schiano C, Picascia A, De Pascale MR, Fiorito C, Casamassimi A, Napoli C. Epigenetic reprogramming in atherosclerosis. Curr Atheroscler Rep. 2015;17:476.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 44]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
11.  Ling C, Groop L. Epigenetics: a molecular link between environmental factors and type 2 diabetes. Diabetes. 2009;58:2718-2725.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 422]  [Cited by in F6Publishing: 437]  [Article Influence: 29.1]  [Reference Citation Analysis (0)]
12.  Cowley AW, Nadeau JH, Baccarelli A, Berecek K, Fornage M, Gibbons GH, Harrison DG, Liang M, Nathanielsz PW, O’Connor DT. Report of the National Heart, Lung, and Blood Institute Working Group on epigenetics and hypertension. Hypertension. 2012;59:899-905.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 78]  [Cited by in F6Publishing: 81]  [Article Influence: 6.8]  [Reference Citation Analysis (0)]
13.  Udali S, Guarini P, Moruzzi S, Choi SW, Friso S. Cardiovascular epigenetics: from DNA methylation to microRNAs. Mol Aspects Med. 2013;34:883-901.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 117]  [Cited by in F6Publishing: 114]  [Article Influence: 9.5]  [Reference Citation Analysis (0)]
14.  Fang M, Chen D, Yang CS. Dietary polyphenols may affect DNA methylation. J Nutr. 2007;137:223S-228S.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Lee WJ, Shim JY, Zhu BT. Mechanisms for the inhibition of DNA methyltransferases by tea catechins and bioflavonoids. Mol Pharmacol. 2005;68:1018-1030.  [PubMed]  [DOI]  [Cited in This Article: ]
16.  Crescenti A, Solà R, Valls RM, Caimari A, Del Bas JM, Anguera A, Anglés N, Arola L. Cocoa Consumption Alters the Global DNA Methylation of Peripheral Leukocytes in Humans with Cardiovascular Disease Risk Factors: A Randomized Controlled Trial. PLoS One. 2013;8:e65744.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 43]  [Cited by in F6Publishing: 44]  [Article Influence: 4.0]  [Reference Citation Analysis (0)]
17.  Klein L, O’Connor CM, Gattis WA, Zampino M, de Luca L, Vitarelli A, Fedele F, Gheorghiade M. Pharmacologic therapy for patients with chronic heart failure and reduced systolic function: review of trials and practical considerations. Am J Cardiol. 2003;91:18F-40F.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 81]  [Cited by in F6Publishing: 88]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
18.  Arce C, Segura-Pacheco B, Perez-Cardenas E, Taja-Chayeb L, Candelaria M, Dueñnas-Gonzalez A. Hydralazine target: from blood vessels to the epigenome. J Transl Med. 2006;4:10.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 69]  [Cited by in F6Publishing: 73]  [Article Influence: 4.1]  [Reference Citation Analysis (0)]
19.  Ellershaw DC, Gurney AM. Mechanisms of hydralazine induced vasodilation in rabbit aorta and pulmonary artery. Br J Pharmacol. 2001;134:621-631.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 42]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
20.  Bang L, Nielsen-Kudsk JE, Gruhn N, Trautner S, Theilgaard SA, Olesen SP, Boesgaard S, Aldershvile J. Hydralazine-induced vasodilation involves opening of high conductance Ca2+-activated K+ channels. Eur J Pharmacol. 1998;361:43-49.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Guidelines 2000 for Cardiopulmonary Resuscitation and Emergency Cardiovascular Care. Part 6: advanced cardiovascular life support: section 5: pharmacology I: agents for arrhythmias. The American Heart Association in collaboration with the International Liaison Committee on Resuscitation. Circulation. 2000;102:I112-I128.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Thangam M, Nathan S, Petrovica M, Kar B, Patel M, Loyalka P, Buja LM, Gregoric ID. Procainamide-induced pulmonary fibrosis after orthotopic heart transplantation: a case report and literature review. Cardiovasc Pathol. 2015;24:250-253.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 4]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
23.  Singh V, Sharma P, Capalash N. DNA methyltransferase-1 inhibitors as epigenetic therapy for cancer. Curr Cancer Drug Targets. 2013;13:379-399.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 105]  [Cited by in F6Publishing: 98]  [Article Influence: 9.8]  [Reference Citation Analysis (0)]
24.  Bernstein BE, Meissner A, Lander ES. The mammalian epigenome. Cell. 2007;128:669-681.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1564]  [Cited by in F6Publishing: 1459]  [Article Influence: 85.8]  [Reference Citation Analysis (0)]
25.  Verdone L, Caserta M, Di Mauro E. Role of histone acetylation in the control of gene expression. Biochem Cell Biol. 2005;83:344-353.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Lehmann LH, Worst BC, Stanmore DA, Backs J. Histone deacetylase signaling in cardioprotection. Cell Mol Life Sci. 2014;71:1673-1690.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 43]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
27.  Zhang LX, DeNicola M, Qin X, Du J, Ma J, Tina Zhao Y, Zhuang S, Liu PY, Wei L, Qin G. Specific inhibition of HDAC4 in cardiac progenitor cells enhances myocardial repairs. Am J Physiol Cell Physiol. 2014;307:C358-C372.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 43]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
28.  Granger A, Abdullah I, Huebner F, Stout A, Wang T, Huebner T, Epstein JA, Gruber PJ. Histone deacetylase inhibition reduces myocardial ischemia-reperfusion injury in mice. FASEB J. 2008;22:3549-3560.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 207]  [Cited by in F6Publishing: 217]  [Article Influence: 13.6]  [Reference Citation Analysis (0)]
29.  Kee HJ, Sohn IS, Nam KI, Park JE, Qian YR, Yin Z, Ahn Y, Jeong MH, Bang YJ, Kim N. Inhibition of histone deacetylation blocks cardiac hypertrophy induced by angiotensin II infusion and aortic banding. Circulation. 2006;113:51-59.  [PubMed]  [DOI]  [Cited in This Article: ]
30.  Liu F, Levin MD, Petrenko NB, Lu MM, Wang T, Yuan LJ, Stout AL, Epstein JA, Patel VV. Histone-deacetylase inhibition reverses atrial arrhythmia inducibility and fibrosis in cardiac hypertrophy independent of angiotensin. J Mol Cell Cardiol. 2008;45:715-723.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 119]  [Cited by in F6Publishing: 117]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
31.  Zhang L, Qin X, Zhao Y, Fast L, Zhuang S, Liu P, Cheng G, Zhao TC. Inhibition of histone deacetylases preserves myocardial performance and prevents cardiac remodeling through stimulation of endogenous angiomyogenesis. J Pharmacol Exp Ther. 2012;341:285-293.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 72]  [Cited by in F6Publishing: 76]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
32.  Zhao TC, Cheng G, Zhang LX, Tseng YT, Padbury JF. Inhibition of histone deacetylases triggers pharmacologic preconditioning effects against myocardial ischemic injury. Cardiovasc Res. 2007;76:473-481.  [PubMed]  [DOI]  [Cited in This Article: ]
33.  Rodríguez-Paredes M, Esteller M. Cancer epigenetics reaches mainstream oncology. Nat Med. 2011;17:330-339.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 868]  [Cited by in F6Publishing: 845]  [Article Influence: 65.0]  [Reference Citation Analysis (0)]
34.  Grozinger CM, Schreiber SL. Deacetylase enzymes: biological functions and the use of small-molecule inhibitors. Chem Biol. 2002;9:3-16.  [PubMed]  [DOI]  [Cited in This Article: ]
35.  Xie M, Kong Y, Tan W, May H, Battiprolu PK, Pedrozo Z, Wang ZV, Morales C, Luo X, Cho G. Histone deacetylase inhibition blunts ischemia/reperfusion injury by inducing cardiomyocyte autophagy. Circulation. 2014;129:1139-1151.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 233]  [Cited by in F6Publishing: 253]  [Article Influence: 25.3]  [Reference Citation Analysis (0)]
36.  Wang D, Zong C, Koag MC, Wang Y, Drews O, Fang C, Scruggs SB, Ping P. Proteome dynamics and proteome function of cardiac 19S proteasomes. Mol Cell Proteomics. 2011;10:M110.006122.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 14]  [Cited by in F6Publishing: 16]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
37.  Lee TM, Lin MS, Chang NC. Inhibition of histone deacetylase on ventricular remodeling in infarcted rats. Am J Physiol Heart Circ Physiol. 2007;293:H968-H977.  [PubMed]  [DOI]  [Cited in This Article: ]
38.  McKinsey TA. The biology and therapeutic implications of HDACs in the heart. Handb Exp Pharmacol. 2011;206:57-78.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 23]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
39.  Aune SE, Herr DJ, Mani SK, Menick DR. Selective inhibition of class I but not class IIb histone deacetylases exerts cardiac protection from ischemia reperfusion. J Mol Cell Cardiol. 2014;72:138-145.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 61]  [Cited by in F6Publishing: 66]  [Article Influence: 6.6]  [Reference Citation Analysis (0)]
40.  Ivanov M, Barragan I, Ingelman-Sundberg M. Epigenetic mechanisms of importance for drug treatment. Trends Pharmacol Sci. 2014;35:384-396.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 105]  [Cited by in F6Publishing: 111]  [Article Influence: 11.1]  [Reference Citation Analysis (0)]
41.  Darakhshan S, Pour AB. Tranilast: a review of its therapeutic applications. Pharmacol Res. 2015;91:15-28.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 152]  [Cited by in F6Publishing: 197]  [Article Influence: 19.7]  [Reference Citation Analysis (0)]
42.  Gallo P, Latronico MV, Gallo P, Grimaldi S, Borgia F, Todaro M, Jones P, Gallinari P, De Francesco R, Ciliberto G. Inhibition of class I histone deacetylase with an apicidin derivative prevents cardiac hypertrophy and failure. Cardiovasc Res. 2008;80:416-424.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 128]  [Cited by in F6Publishing: 131]  [Article Influence: 8.2]  [Reference Citation Analysis (0)]
43.  Pan MH, Lai CS, Wu JC, Ho CT. Epigenetic and disease targets by polyphenols. Curr Pharm Des. 2013;19:6156-6185.  [PubMed]  [DOI]  [Cited in This Article: ]
44.  Wongcharoen W, Phrommintikul A. The protective role of curcumin in cardiovascular diseases. Int J Cardiol. 2009;133:145-151.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 198]  [Cited by in F6Publishing: 190]  [Article Influence: 12.7]  [Reference Citation Analysis (0)]
45.  Morimoto T, Sunagawa Y, Kawamura T, Takaya T, Wada H, Nagasawa A, Komeda M, Fujita M, Shimatsu A, Kita T. The dietary compound curcumin inhibits p300 histone acetyltransferase activity and prevents heart failure in rats. J Clin Invest. 2008;118:868-878.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 69]  [Cited by in F6Publishing: 167]  [Article Influence: 10.4]  [Reference Citation Analysis (0)]
46.  Bronze-da-Rocha E. MicroRNAs expression profiles in cardiovascular diseases. Biomed Res Int. 2014;2014:985408.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 109]  [Cited by in F6Publishing: 125]  [Article Influence: 12.5]  [Reference Citation Analysis (0)]
47.  Wahlquist C, Jeong D, Rojas-Muñoz A, Kho C, Lee A, Mitsuyama S, van Mil A, Park WJ, Sluijter JP, Doevendans PA. Inhibition of miR-25 improves cardiac contractility in the failing heart. Nature. 2014;508:531-535.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 302]  [Cited by in F6Publishing: 315]  [Article Influence: 31.5]  [Reference Citation Analysis (0)]
48.  Montgomery RL, Hullinger TG, Semus HM, Dickinson BA, Seto AG, Lynch JM, Stack C, Latimer PA, Olson EN, van Rooij E. Therapeutic inhibition of miR-208a improves cardiac function and survival during heart failure. Circulation. 2011;124:1537-1547.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 459]  [Cited by in F6Publishing: 443]  [Article Influence: 34.1]  [Reference Citation Analysis (0)]
49.  Carè A, Catalucci D, Felicetti F, Bonci D, Addario A, Gallo P, Bang ML, Segnalini P, Gu Y, Dalton ND. MicroRNA-133 controls cardiac hypertrophy. Nat Med. 2007;13:613-618.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Bang C, Batkai S, Dangwal S, Gupta SK, Foinquinos A, Holzmann A, Just A, Remke J, Zimmer K, Zeug A. Cardiac fibroblast-derived microRNA passenger strand-enriched exosomes mediate cardiomyocyte hypertrophy. J Clin Invest. 2014;124:2136-2146.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 642]  [Cited by in F6Publishing: 736]  [Article Influence: 73.6]  [Reference Citation Analysis (0)]
51.  Azibani F, Devaux Y, Coutance G, Schlossarek S, Polidano E, Fazal L, Merval R, Carrier L, Solal AC, Chatziantoniou C. Aldosterone inhibits the fetal program and increases hypertrophy in the heart of hypertensive mice. PLoS One. 2012;7:e38197.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 21]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
52.  Dickinson BA, Semus HM, Montgomery RL, Stack C, Latimer PA, Lewton SM, Lynch JM, Hullinger TG, Seto AG, van Rooij E. Plasma microRNAs serve as biomarkers of therapeutic efficacy and disease progression in hypertension-induced heart failure. Eur J Heart Fail. 2013;15:650-659.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 114]  [Cited by in F6Publishing: 125]  [Article Influence: 11.4]  [Reference Citation Analysis (0)]
53.  Shyu KG, Wang BW, Cheng WP, Lo HM. MicroRNA-208a Increases Myocardial Endoglin Expression and Myocardial Fibrosis in Acute Myocardial Infarction. Can J Cardiol. 2015;31:679-690.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 53]  [Cited by in F6Publishing: 55]  [Article Influence: 5.5]  [Reference Citation Analysis (0)]
54.  Kontaraki JE, Marketou ME, Zacharis EA, Parthenakis FI, Vardas PE. Differential expression of vascular smooth muscle-modulating microRNAs in human peripheral blood mononuclear cells: novel targets in essential hypertension. J Hum Hypertens. 2014;28:510-516.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 86]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
55.  Karakikes I, Chaanine AH, Kang S, Mukete BN, Jeong D, Zhang S, Hajjar RJ, Lebeche D. Therapeutic cardiac-targeted delivery of miR-1 reverses pressure overload-induced cardiac hypertrophy and attenuates pathological remodeling. J Am Heart Assoc. 2013;2:e000078.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 183]  [Cited by in F6Publishing: 205]  [Article Influence: 18.6]  [Reference Citation Analysis (0)]
56.  Han W, Han Y, Liu X, Shang X. Effect of miR-29a inhibition on ventricular hypertrophy induced by pressure overload. Cell Biochem Biophys. 2015;71:821-826.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 15]  [Cited by in F6Publishing: 15]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
57.  Bonauer A, Carmona G, Iwasaki M, Mione M, Koyanagi M, Fischer A, Burchfield J, Fox H, Doebele C, Ohtani K. MicroRNA-92a controls angiogenesis and functional recovery of ischemic tissues in mice. Science. 2009;324:1710-1713.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 896]  [Cited by in F6Publishing: 923]  [Article Influence: 61.5]  [Reference Citation Analysis (0)]
58.  Hinkel R, Penzkofer D, Zühlke S, Fischer A, Husada W, Xu QF, Baloch E, van Rooij E, Zeiher AM, Kupatt C. Inhibition of microRNA-92a protects against ischemia/reperfusion injury in a large-animal model. Circulation. 2013;128:1066-1075.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 225]  [Cited by in F6Publishing: 235]  [Article Influence: 21.4]  [Reference Citation Analysis (0)]
59.  Bellera N, Barba I, Rodriguez-Sinovas A, Ferret E, Asín MA, Gonzalez-Alujas MT, Pérez-Rodon J, Esteves M, Fonseca C, Toran N. Single intracoronary injection of encapsulated antagomir-92a promotes angiogenesis and prevents adverse infarct remodeling. J Am Heart Assoc. 2014;3:e000946.  [PubMed]  [DOI]  [Cited in This Article: ]
60.  Tolonen AM, Magga J, Szabó Z, Viitala P, Gao E, Moilanen AM, Ohukainen P, Vainio L, Koch WJ, Kerkelä R. Inhibition of Let-7 microRNA attenuates myocardial remodeling and improves cardiac function postinfarction in mice. Pharmacol Res Perspect. 2014;2:e00056.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 46]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
61.  Napoli C, Crudele V, Soricelli A, Al-Omran M, Vitale N, Infante T, Mancini FP. Primary prevention of atherosclerosis: a clinical challenge for the reversal of epigenetic mechanisms? Circulation. 2012;125:2363-2373.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 86]  [Article Influence: 7.2]  [Reference Citation Analysis (0)]
62.  Napoli C, Infante T, Casamassimi A. Maternal-foetal epigenetic interactions in the beginning of cardiovascular damage. Cardiovasc Res. 2011;92:367-374.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 46]  [Cited by in F6Publishing: 47]  [Article Influence: 3.6]  [Reference Citation Analysis (0)]
63.  Napoli C, Lerman LO, de Nigris F, Gossl M, Balestrieri ML, Lerman A. Rethinking primary prevention of atherosclerosis-related diseases. Circulation. 2006;114:2517-2527.  [PubMed]  [DOI]  [Cited in This Article: ]
64.  Saddic LA, Chang TW, Sigurdsson MI, Heydarpour M, Raby BA, Shernan SK, Aranki SF, Body SC, Muehlschlegel JD. Integrated microRNA and mRNA responses to acute human left ventricular ischemia. Physiol Genomics. 2015;47:455-462.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 27]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
65.  Gao H, Guddeti RR, Matsuzawa Y, Liu LP, Su LX, Guo D, Nie SP, Du J, Zhang M. Plasma Levels of microRNA-145 Are Associated with Severity of Coronary Artery Disease. PLoS One. 2015;10:e0123477.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 52]  [Article Influence: 5.8]  [Reference Citation Analysis (0)]
66.  Li X, Yang Y, Wang L, Qiao S, Lu X, Wu Y, Xu B, Li H, Gu D. Plasma miR-122 and miR-3149 Potentially Novel Biomarkers for Acute Coronary Syndrome. PLoS One. 2015;10:e0125430.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 31]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
67.  Stather PW, Sylvius N, Sidloff DA, Dattani N, Verissimo A, Wild JB, Butt HZ, Choke E, Sayers RD, Bown MJ. Identification of microRNAs associated with abdominal aortic aneurysms and peripheral arterial disease. Br J Surg. 2015;102:755-766.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 46]  [Cited by in F6Publishing: 50]  [Article Influence: 5.6]  [Reference Citation Analysis (0)]
68.  Huang S, Chen M, Li L, He M, Hu D, Zhang X, Li J, Tanguay RM, Feng J, Cheng L. Circulating MicroRNAs and the occurrence of acute myocardial infarction in Chinese populations. Circ Cardiovasc Genet. 2014;7:189-198.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 40]  [Cited by in F6Publishing: 44]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
69.  Fan X, Wang E, Wang X, Cong X, Chen X. MicroRNA-21 is a unique signature associated with coronary plaque instability in humans by regulating matrix metalloproteinase-9 via reversion-inducing cysteine-rich protein with Kazal motifs. Exp Mol Pathol. 2014;96:242-249.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 47]  [Cited by in F6Publishing: 43]  [Article Influence: 4.3]  [Reference Citation Analysis (0)]
70.  Wang HW, Huang TS, Lo HH, Huang PH, Lin CC, Chang SJ, Liao KH, Tsai CH, Chan CH, Tsai CF. Deficiency of the microRNA-31-microRNA-720 pathway in the plasma and endothelial progenitor cells from patients with coronary artery disease. Arterioscler Thromb Vasc Biol. 2014;34:857-869.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 65]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
71.  Wu J, Song J, Wang C, Niu D, Li H, Liu Y, Ma L, Yu R, Chen X, Zen K. Identification of serum microRNAs for cardiovascular risk stratification in dyslipidemia subjects. Int J Cardiol. 2014;172:232-234.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 8]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
72.  Endo K, Naito Y, Ji X, Nakanishi M, Noguchi T, Goto Y, Nonogi H, Ma X, Weng H, Hirokawa G. MicroRNA 210 as a biomarker for congestive heart failure. Biol Pharm Bull. 2013;36:48-54.  [PubMed]  [DOI]  [Cited in This Article: ]
73.  Ren J, Zhang J, Xu N, Han G, Geng Q, Song J, Li S, Zhao J, Chen H. Signature of circulating microRNAs as potential biomarkers in vulnerable coronary artery disease. PLoS One. 2013;8:e80738.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 131]  [Cited by in F6Publishing: 156]  [Article Influence: 14.2]  [Reference Citation Analysis (0)]
74.  Nabiałek E, Wańha W, Kula D, Jadczyk T, Krajewska M, Kowalówka A, Dworowy S, Hrycek E, Włudarczyk W, Parma Z. Circulating microRNAs (miR-423-5p, miR-208a and miR-1) in acute myocardial infarction and stable coronary heart disease. Minerva Cardioangiol. 2013;61:627-637.  [PubMed]  [DOI]  [Cited in This Article: ]
75.  Zhao DS, Chen Y, Jiang H, Lu JP, Zhang G, Geng J, Zhang Q, Shen JH, Zhou X, Zhu W. Serum miR-210 and miR-30a expressions tend to revert to fetal levels in Chinese adult patients with chronic heart failure. Cardiovasc Pathol. 2013;22:444-450.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 46]  [Cited by in F6Publishing: 53]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
76.  D’Alessandra Y, Carena MC, Spazzafumo L, Martinelli F, Bassetti B, Devanna P, Rubino M, Marenzi G, Colombo GI, Achilli F. Diagnostic potential of plasmatic MicroRNA signatures in stable and unstable angina. PLoS One. 2013;8:e80345.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 87]  [Cited by in F6Publishing: 103]  [Article Influence: 9.4]  [Reference Citation Analysis (0)]
77.  Ellis KL, Cameron VA, Troughton RW, Frampton CM, Ellmers LJ, Richards AM. Circulating microRNAs as candidate markers to distinguish heart failure in breathless patients. Eur J Heart Fail. 2013;15:1138-1147.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 123]  [Cited by in F6Publishing: 139]  [Article Influence: 12.6]  [Reference Citation Analysis (0)]
78.  Vogel B, Keller A, Frese KS, Leidinger P, Sedaghat-Hamedani F, Kayvanpour E, Kloos W, Backe C, Thanaraj A, Brefort T. Multivariate miRNA signatures as biomarkers for non-ischaemic systolic heart failure. Eur Heart J. 2013;34:2812-2822.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 87]  [Article Influence: 7.9]  [Reference Citation Analysis (0)]
79.  Olivieri F, Antonicelli R, Lorenzi M, D’Alessandra Y, Lazzarini R, Santini G, Spazzafumo L, Lisa R, La Sala L, Galeazzi R. Diagnostic potential of circulating miR-499-5p in elderly patients with acute non ST-elevation myocardial infarction. Int J Cardiol. 2013;167:531-536.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 156]  [Cited by in F6Publishing: 180]  [Article Influence: 15.0]  [Reference Citation Analysis (0)]
80.  Wang F, Long G, Zhao C, Li H, Chaugai S, Wang Y, Chen C, Wang DW. Plasma microRNA-133a is a new marker for both acute myocardial infarction and underlying coronary artery stenosis. J Transl Med. 2013;11:222.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 84]  [Cited by in F6Publishing: 98]  [Article Influence: 8.9]  [Reference Citation Analysis (0)]
81.  Lu HQ, Liang C, He ZQ, Fan M, Wu ZG. Circulating miR-214 is associated with the severity of coronary artery disease. J Geriatr Cardiol. 2013;10:34-38.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in F6Publishing: 24]  [Reference Citation Analysis (0)]