Next Article in Journal
Effectiveness of Extracellular Vesicle Application in Skin Aging Treatment and Regeneration: Do We Have Enough Evidence from Clinical Trials?
Next Article in Special Issue
Sodium–Glucose Cotransporter 2 Inhibitors in Aortic Stenosis: Toward a Comprehensive Cardiometabolic Approach
Previous Article in Journal
Molecular Anatomy of Synaptic and Extrasynaptic Neurotransmission Between Nociceptive Primary Afferents and Spinal Dorsal Horn Neurons
Previous Article in Special Issue
Genome-Wide Expression Profiling and Phenotypic Analysis of Downstream Targets Identify the Fox Transcription Factor Jumeau as a Master Regulator of Cardiac Progenitor Cell Division
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Blood-Based DNA Methylation Biomarkers to Identify Risk and Progression of Cardiovascular Disease

1
Biomedical Research and Innovation Platform, South African Medical Research Council, Tygerberg 7505, South Africa
2
Centre for Cardio-metabolic Research in Africa, Division of Medical Physiology, Faculty of Medicine and Health Sciences, Stellenbosch University, Tygerberg 7505, South Africa
3
Division of Cell Biology, Department of Human Biology, Faculty of Health Sciences, University of Cape Town, Cape Town 7925, South Africa
4
Department of Biochemistry and Microbiology, University of Zululand, Kwa-Dlangezwa 3886, South Africa
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(5), 2355; https://doi.org/10.3390/ijms26052355
Submission received: 17 January 2025 / Revised: 28 February 2025 / Accepted: 2 March 2025 / Published: 6 March 2025

Abstract

:
Non-communicable diseases (NCDs) are the leading cause of death worldwide, with cardiovascular disease (CVD) accounting for half of all NCD-related deaths. The biological onset of CVD may occur long before the development of clinical symptoms, hence the urgent need to understand the molecular alterations underpinning CVD, which would facilitate intervention strategies to prevent or delay the onset of the disease. There is evidence to suggest that CVD develops through a complex interplay between genetic, lifestyle, and environmental factors. Epigenetic modifications, including DNA methylation, serve as proxies linking genetics and the environment to phenotypes and diseases. In the past decade, a growing list of studies has implicated DNA methylation in the early events of CVD pathogenesis. In this regard, screening for these epigenetic marks in asymptomatic individuals may assist in the early detection of CVD and serve to predict the response to therapeutic interventions. This review discusses the current literature on the relationship between blood-based DNA methylation alterations and CVD in humans. We highlight a set of differentially methylated genes that show promise as candidates for diagnostic and prognostic CVD biomarkers, which should be prioritized and replicated in future studies across additional populations. Finally, we discuss key limitations in DNA methylation studies, including genetic diversity, interpatient variability, cellular heterogeneity, study confounders, different methodological approaches used to isolate and measure DNA methylation, sample sizes, and cross-sectional study design.

1. Background

Cardiovascular diseases (CVDs), including coronary heart disease (CHD), hypertension, and heart failure, represent the highest prevalence of non-communicable diseases worldwide and are increasing at an alarming rate, particularly in low- and middle-income countries [1,2]. In 2019, CVDs were responsible for approximately 17.9 million deaths, representing 32% of all deaths globally [3]. CVDs are progressive disorders which involve an intricate interplay between genetic, environmental, and lifestyle factors, including tobacco use, physical inactivity, alcohol consumption, and obesity [4,5]. Additional risk factors include social adversity, high blood pressure, pollution, and urbanization in low- and middle-income countries [5,6].
Investments in prevention are considered to have a major impact on reducing the CVD burden. By preventing and/or delaying the disease, it may be possible to reduce CVD-associated mortality and the economic burden of treating this disease. Furthermore, the use of biomarkers may aid in the identification of high-risk individuals, profile key modifiable risk factors, and represent novel therapeutic targets. To this end, epigenetic alterations that are detectable in the blood, a minimally-invasive biological source, may present as attractive biomarkers and are thus drawing increasing attention for the early detection of many diseases [7].
Epigenetics, defined as heritable changes that influence gene expression without alteration(s) within the underlying DNA sequence, may shed light on the interplay between the genome and its environment in complex diseases, such as in CVDs [8]. DNA methylation is an epigenetic modification defined as the enzymatic addition of a methyl group to the fifth carbon of cytosine dinucleotides [9]. DNA methylation within the promoter regions of genes has been extensively linked with gene silencing, and accumulating studies have uncovered important roles for DNA methylation in the etiopathogenesis of CVDs [10,11,12,13,14,15]. There is also evidence to suggest that DNA methylation may be an early event in the onset of CVDs and, importantly, since DNA methylation is reversible, screening for disease-associated DNA methylation marks in asymptomatic individuals may assist in the early detection of disease and serve to predict response to therapeutic interventions in affected individuals.
For studies focused on CVDs, the accessibility and procurement of cardiac tissue, which is central to disease pathogenesis, is not feasible. This has led to a search for less invasive clinical biomarkers associated with CVDs that can be detected in easily accessible tissues, such as in peripheral blood. While some studies have reported that whole blood or its constituents are good proxies for DNA methylation in disease-associated tissues, there is still a paucity of information on whether this is the case for CVDs [16,17].
This review aims to explore the current evidence on blood-based DNA methylation signatures that are associated with CVDs. Three bibliographic databases (Cochrane, PubMed, and Scopus) were searched to identify articles reporting on CVD-associated DNA methylation signatures in blood using the following keywords: “DNA methylation”, AND “blood”, OR “peripheral blood”, OR “peripheral blood mononuclear cells”, OR “peripheral blood leukocytes”, OR “peripheral blood lymphocytes”, OR “white blood cells” AND “cardiovascular disease” AND “human”. As CVD is an umbrella term encompassing a wide range of phenotypes, this review included studies that examined both general CVD outcomes as well as related conditions affecting the heart and blood vessels. To this end, we included studies investigating myocardial infarction, coronary heart disease, stroke, heart failure, and atherosclerosis, recognizing the shared pathophysiological mechanisms and overlapping risk factors of these conditions. To maximize the sample size of the included articles, no restrictions on dates were applied to the literature search. Articles were considered eligible if they were original, full-length studies investigating the DNA methylation modifications associated with CVD or related conditions in humans, using randomized controlled trials, cohort, case-control, cross-sectional, or longitudinal study designs. Intervention studies, case reports, letters, conference abstracts, and articles not written in English were excluded. The reference lists of all eligible original research and review articles were also manually scanned to identify studies that were missed using the original search strategy.

2. Studies Profiling DNA Methylation in CVD

A total of 4747 articles were identified in the initial literature search. After removing duplicates, article titles and abstracts were screened for eligible full-text articles, of which 37 met the inclusion criteria. An additional 11 studies, identified from reference lists of eligible articles, were also included in this review (Figure 1). The detailed characteristics of the eligible studies investigating the DNA methylation alterations in CVDs are summarized in Table 1, Table 2 and Table 3. A total of 12 studies analyzed global DNA methylation, 14 investigated DNA methylation of candidate genes, and 22 investigated genome-wide DNA methylation signatures. The studies were conducted in numerous populations, and included European (n = 17), Chinese (n = 16), American (n = 12), Indian (n = 4), African (n = 4), Japanese (n = 1), Iranian (n = 1), Turkish (n = 1), and Brazilian (n = 1) participants. The included articles studied either cross-sectional, case-control datasets (n = 30) or longitudinal cohorts (n = 18), with the participant sample sizes varying substantially between studies (ranging from 2 to 11,273 individuals). Thirty-five of the studies included a sample larger than 100 individuals and three studies analyzed samples from 20 or fewer individuals. The profiling of DNA methylation was conducted in either whole blood or blood constituents, including white blood cells, peripheral blood mononuclear cells, lymphocytes, and leukocytes.

2.1. Global DNA Methylation Studies

Global DNA methylation is defined as the total percentage of cytosines within the genome that are methylated [63]. A change in global DNA methylation has been frequently associated with aging and cancer, and more recently, with metabolic diseases, including CVDs [64]. We identified several articles investigating the status of global DNA methylation in major CVD processes, including atherosclerosis, hypertension, and CHD (Table 1). These studies used various molecular techniques to quantify global DNA methylation, including the measurement of repetitive elements as a surrogate for DNA methylation, enzyme-linked immunosorbent assays (ELISAs), methylation-specific restriction enzymes, as well as liquid chromatography mass spectrometry (LCMS).
Castro and colleagues (2003) investigated global DNA methylation levels in the peripheral blood leukocytes from 17 male atherosclerosis patients compared to healthy controls, using a methylation sensitive restriction enzyme-based assay [19]. The authors reported that global DNA methylation was significantly reduced in the atherosclerosis patients and inversely correlated with plasma homocysteine levels [19]. Since homocysteine plays a prominent role in one-carbon metabolism and is an established, independent risk factor for vascular disease [65], the authors proposed that elevated plasma homocysteine concentrations observed in the atherosclerosis group may consequently result in DNA methyltransferase inhibition and reduced systemic methylation capacity. Similarly, using an ELISA method, Ramos (2016) reported reduced global DNA methylation in the peripheral blood leukocytes from 90 postmenopausal women, which revealed an association between global DNA hypomethylation and cardiovascular risk that remained significant even after adjustment for time since menopause [25].
It is estimated that more than half of the human genome is comprised of repetitive elements [66]. These include retrotransposons, such as LINE-1 and Alu elements, which have gained increased attention in the last decade for their use as surrogate markers of global DNA methylation. In line with the findings of Castro (2003) and Ramos (2016), six additional studies analyzing the status of LINE-1 and Alu in blood identified an association between global DNA hypomethylation and CVD risk predictors, such as altered levels of low density- and high-density lipoproteins, hypertension, and CVD events, such as stroke, CHD, and myocardial infarction [21,22,23,24,67,68]. Interestingly, there is also evidence to suggest that methylation of LINE-1 and Alu elements may be sensitive to environmental exposures. Indeed, numerous CVD risk factors, including exposure to air pollution [69,70], cigarette smoking [71], and alcohol consumption [72], have also been associated with DNA hypomethylation, suggesting this epigenetic signature may be an early event in the development of CVDs, as well as other lifestyle-related diseases.
Other data from global DNA methylation studies have been conflicting. Kim et al. (2010) investigated methylation of Alu elements in 286 participants from the Singapore Chinese Health Study and identified a positive correlation between DNA hypermethylation in peripheral blood leukocytes and the prevalence of CVDs [10]. Using a methylation-sensitive restriction enzyme assay, Sharma and colleagues (2008) also observed an increase in the global DNA methylation levels in the peripheral lymphocytes from 137 patients with CHD compared to 150 healthy controls, which also correlated with plasma homocysteine levels [20]. Similarly, using a pyrosequencing approach, Alexeeff et al. (2013) assessed the DNA methylation of LINE-1 and Alu elements in relation to blood pressure in a longitudinal study comprising 789 elderly participants from the Normative Aging Study, and identified a positive association between Alu hypermethylation and diastolic blood pressure, whilst no changes in LINE-1 methylation were observed [27].
Moreover, Jiang and colleagues (2019) investigated the levels of 5-methylcytosine (5mC) and 5-methylhydroxycytosine (5hmC) in a cross-sectional cohort consisting of 91 Chinese participants diagnosed with atherosclerosis and 22 healthy controls. They identified a positive correlation between 5mC/5hmC and Gensini scoring, a clinical indicator of atherosclerosis severity. Furthermore, 5mC/5hmC significantly correlated with atherosclerosis risk factors, including fasting glucose, HbA1c levels, and Crouse score [28]. In an additional study by the same group, 5mC/5hmC levels were analyzed in the peripheral blood mononuclear cells from 44 elderly subjects with CVDs (23 with acute myocardial infarction and 21 with angina) and 42 control subjects matched for age and gender. Using two independent quantification methods, i.e., ELISA and dot plot, the authors equivocally showed that 5mC levels were significantly increased in the CVD patients compared to the controls [29]. Lastly, in contrast to the aforementioned findings, Istas and coworkers (2017) investigated global DNA methylation in whole blood samples from eight atherosclerosis patients compared to eight healthy controls from the Impact of Le Ventricular Assist Devices Implantation on Micro- and Macrovascular Function cohort [LVAD, (NCT02174133)] and found no significant difference in global DNA methylation between the two groups [26].

2.2. Candidate Gene Studies

Despite the popularity of global DNA methylation studies, a shift towards investigating gene candidates that correlate with specific clinical features or disease traits has gained increased interest in recent years [73]. The methodological approaches used for this application include methylated DNA immunoprecipitation (MeDIP), quantitative methylation-specific PCR, mass spectrometry, and bisulfite pyrosequencing. Using these approaches, several studies have identified associations between the methylation status of specific gene loci and cardiovascular risks. These loci include fatty acid binding protein 3 (FABP3), interferon gamma (IFN-γ), interleukin-6 (IL-6), inducible nitric oxide synthase (iNOS), thioredoxin-interacting protein (TXNIP), bone morphogenic protein receptor 2 (BMPR2), forkhead/winged helix transcription factor 3 (FOXP3), miR-510, miR-223, sterol O-acyltransferase 1 (SOAT1), human telomerase reverse transcriptase (hTERT), toll-like receptor 2 (TLR2), glucocorticoid receptor (NR3C1) and atrial natriuretic peptide (NPPA), and are elaborated on in the subsequent sections and summarized in Table 2.

2.2.1. Genes Involved in Inflammation and Immunity

Chronic inflammation is largely mediated by the expression of interleukin genes, which encode a family of pro-inflammatory cytokines [74]. Of these, IL-6, which regulates various inflammatory responses, has been the most extensively studied. Zuo and colleagues (2016) investigated the methylation status of two CpG sites within the IL-6 promoter in the peripheral blood leukocytes obtained from 212 CHD patients [36]. Both CpG sites were hypomethylated in the CHD patients compared to the controls, with the combined methylation scores correlating significantly with CHD risk [36]. Whilst the gene and protein expression in this study were not determined, other studies have reported a positive relationship between circulating IL-6 expression and CVD characteristics, including blood pressure and arterial stiffness [75,76]. Importantly, IL-6 expression is also influenced by exposure to CVD risk factors, including air pollution and salt consumption, highlighting that aberrant DNA methylation and the expression of IL-6 may be an early and preventable event in the development of CVDs [75,77]. In an independent study, DNA methylation of the IL-6 promoter was investigated in a sample of 35 Iranian patients with atherosclerosis versus 30 healthy controls. The study identified hypomethylation of six CpG sites in the distal IL-6 promoter in the peripheral blood from atherosclerosis patients compared to healthy subjects, which inversely correlated with mRNA expression levels [39]. Interestingly, the CpG sites investigated in this study were highly conserved amongst several species, indicating that these sites may functionally be important in the regulation of IL-6 gene expression.
In a 10-year longitudinal study involving 789 elderly participants, Alexeeff and colleagues (2013) investigated the methylation status of a panel of inflammatory genes in whole blood using a pyrosequencing approach [27]. After adjusting for confounding factors, including age, body mass index, diabetes mellitus, alcohol consumption, ethnicity, and neutrophil count, they reported that DNA methylation of the TLR2 and iNOS promoters positively correlated with blood pressure, whilst a negative correlation was observed between IFN-γ promoter methylation and blood pressure. Although the expression of these genes was not quantified in the study, TLR2, IFN-γ, and iNOS have previously been implicated in atherosclerosis and myocardial inflammation in both humans and animal models [78,79,80,81]. Indeed, overexpression of TLR2 in mice leads to early atherosclerosis onset, whilst TLR2-deficient mice display reduced atherosclerosis-associated inflammation [82]. IFN-γ is a T-cell-derived, macrophage-activated pro-inflammatory cytokine that is overexpressed in human atherosclerotic lesions [83]. In mice, the exogenous administration of IFN-γ exacerbates the formation of atherosclerotic lesions, whilst the knockout of IFN-γ reduces lesion size [83]. Furthermore, rabbits exposed to a high-cholesterol diet, whilst treated simultaneously with a selective iNOS inhibitor, were ameliorated from the development of atherosclerosis compared to control atherosclerotic rabbits [84].
Increased expression of circulating TXNIP, a member of the alpha-arrestin protein family, has been associated with numerous metabolic disorders [85]. TXNIP is upregulated in response to high-glucose-induced stress and plays a role in redox homeostasis by binding to and activating the NOD-like receptor family pyrin domain containing three (NLRP3) inflammasome [85]. The methylation status of TXNIP was investigated in the peripheral blood leukocytes of 54 Chinese patients with CHD compared to 54 age- and gender-matched healthy controls. The results showed that a single CpG site, located at cg19693031 within the 3′ UTR of TXNIP, was hypomethylated in the CHD patients compared to the controls. The DNA methylation at this site also inversely correlated with TXNIP mRNA levels, as well as fasting plasma glucose and total cholesterol [42]. In line with this study, the DNA methylation of this same CpG site was found to be associated with type 2 diabetes in four independent GWAS studies involving Mexican-American, German, Indian-Asian, and Spanish populations [86,87,88,89].
Valvular heart disease is a common cause of pulmonary hypertension and is associated with high mortality rates, partly due to a lack of reliable and effective early diagnosis markers [90]. A study by Li et al. [30] investigated the methylation status of BMPR2, a member of the TGF-β family, in the peripheral blood from Chinese patients with valvular heart disease complicated by pulmonary hypertension (VHD-PAH). The study cohort consisted of 26 male and 29 female VHD-PAH patients and 28 controls. Using pyrosequencing, the authors showed that the BMPR2 promoter was hypermethylated in cases versus controls, which inversely correlated with BMPR2 mRNA levels. Importantly, receiver operating characteristic curve (ROC) analysis revealed that DNA methylation combined with BMPR2 mRNA levels could distinguish VHD-PAH patients from controls with high specificity. These findings agree with studies showing a link between BMPR2 mutations and the development of pulmonary arterial hypertension [91]. These findings also agree with other studies highlighting monoallelic mutations, resulting in haploinsufficiency of BMPR2, as the main genetic risk factor for heritable pulmonary arterial hypertension [92]. Moreover, homozygous and heterozygous mutations in BMPR2 in rats lead to spontaneous pulmonary hypertension, as evidenced by pulmonary artery pressure, pulmonary vascular resistance, right ventricular hypertrophy, and decreased cardiac output [93].
Regulatory T (Treg) cells have been demonstrated to play a protective role against atherosclerosis in numerous experimental models. These cells are regulated by FOXP3, and recently, genetic variants of the FOXP3 gene have been identified as risk factors for the development of atherosclerosis [94,95]. Zhu and colleagues analyzed the DNA methylation levels at a conserved regulatory region (known as the Regulatory T-Cell-Specific Demethylated Region) within the FOXP3 locus in the peripheral blood from a longitudinal cohort consisting of 171 patients with acute CHD, and showed that elevated methylation levels of FOXP3 associated with poor clinical outcomes and severity of atherosclerosis [31]. This agrees with an independent study showing that DNA methylation within the same regulatory region of FOXP3 was associated with CHD, even after adjusting for age, smoking, BMI, mean blood pressure, blood glucose, and lipid profiles [32]. The authors of this study additionally showed that exposure of cultured peripheral blood mononuclear cells to oxidized low-density lipoprotein resulted in a dose-dependent increase in FOXP3 methylation. The authors thus proposed a mechanism whereby dysregulation of Treg cells may occur via epigenetic suppression of FOXP3, leading to an increased risk of developing acute coronary syndrome [32].

2.2.2. MicroRNAs

MicroRNAs (miRNAs) are short, non-coding regulatory molecules that control gene expression post-transcriptionally [96]. The altered expression of miRNAs has been associated with the development of a growing list of non-communicable diseases, including cancer, diabetes, and cardiovascular disease [96]. Krishnan et al. (2017) investigated the promoter DNA methylation levels of miR-510 in the peripheral blood from 20 hypertensive patients versus 20 healthy controls using a methylation-specific PCR-based method. Their study revealed that the miR-510 promoter was hypomethylated in hypertensive patients, which was accompanied by increased miR-510 gene expression levels [33]. MiR-510 has an established role in regulating cell proliferation, migration, and apoptosis, and not surprisingly, it has been implicated as an onco-miRNA in several cancers [97,98,99]. While the detailed role of miR-510 in hypertension remains to be elucidated, the authors proposed that circulating miR-510 may be a promising candidate diagnostic biomarker for hypertension.
MiR-223 plays a crucial role in cholesterol homeostasis by targeting genes involved in the synthesis, degradation, and transport of lipoproteins. Expectedly, it has been implicated in diseases involving the vascular system and has been proposed as a circulating biomarker for a range of cardiovascular diseases, including atherosclerosis, myocardial infarction, arrhythmia, and hypertension [100]. A study by Li and colleagues (2017) investigated the methylation status of the miRNA-233 promoter in the peripheral blood leukocytes from 23 patients with atherosclerotic cerebral infarction compared to 32 healthy controls [37]. Using bisulfite sequencing, they identified a total of seven hypomethylated CpG sites within the miRNA-233 promoter in cases versus controls, which negatively correlated with miRNA-233 mRNA levels. Using Spearman’s correlation analysis, the authors showed that methylation at these CpG sites inversely correlated with total plasma cholesterol levels. The mean methylation levels of these sites were also lower in participants with carotid atherosclerosis compared to those without carotid atherosclerosis. The authors thus proposed that miRNA-233 should be considered as a potential biomarker and target for the prevention and treatment of atherosclerotic cerebral infarction.

2.2.3. Genes Involved in Lipid Metabolism

Dysregulated lipid metabolism is a common feature of obesity, insulin resistance, and CVDs [101]. Fatty acid binding proteins (FABPs) are abundantly expressed in the heart and serve as chaperones for intracellular lipids, particularly for the transportation of free long-chain fatty acids for cellular metabolism, resulting in protection against lipid toxicity [102]. A previous preclinical study in mice showed that the depletion of FABP3 results in increased glycolysis, lipid accumulation, reduction in fatty acid uptake, and cardiac dysfunction [103]. Based on this preliminary evidence, Zhang et al. (2013) sought to investigate the DNA methylation levels of FABP3 in PMBCs, and its relationship with cardiometabolic risk factors in a study cohort consisting of 517 individuals of Northern European descent [34]. Using an Epityper mass array, they investigated the DNA methylation of 17 CpG sites located within the proximal promoter and first exon of FABP3. The average methylation of the 17 CpG sites correlated positively with cholesterol levels, whilst methylation at specific individual sites linked with diastolic blood pressure, insulin resistance, and low-density lipoprotein. Importantly, the gene expression of FABP3 was found to be inversely correlated with FABP3 methylation and positively correlated with protective metabolic factors, HDL-cholesterol, and adiponectin, and negatively with total cholesterol, triglycerides, and waist-to-hip ratio [34]. The inverse relationship between DNA methylation and FABP3 gene expression is consistent with a previous study showing that a 50-base pair region directly upstream of the FABP3 transcription start site was sufficient to drive expression of the FABP3 gene in a mouse model [104].
The Sterol-O-Acyl transferase (SOAT) enzyme family consists of membrane-bound proteins localized to the endoplasmic reticulum, where they play a key role in converting endoplasmic reticulum cholesterol to cholesterol esters to store lipid droplets [105]. Of these enzymes, SOAT1 has been linked to inflammation and the formation of atherosclerotic plaques. Moreover, depletion of SOAT1 was previously shown to ameliorate dysfunctional insulin signaling and dyslipidemia in mice exposed to a high-fat diet [106]. Based on these studies, Abuzhalihan and colleagues (2019) investigated SOAT1 methylation in a Chinese population of 99 individuals with CHD and 89 healthy controls. Of 26 CpG sites analyzed within the SOAT1 promoter, 7 CpG sites had lower methylation levels in the CHD patients compared to the control group. These methylation profiles correlated with early disease onset after adjusting for total cholesterol [41].

2.2.4. Genes Associated with Stress and Aging

It is well established that telomere length is associated with age-related diseases, including a broad spectrum of cardiovascular diseases [107]. Amongst various levels of regulation, telomere activity and length are maintained by hTERT. The gene encoding hTERT has been shown to be regulated by methylation in a wide range of tissue types [108]. Zhang and colleagues (2013) investigated promoter methylation of hTERT in 197 atherosclerosis patients and 165 age- and sex-matched healthy controls. They found that hTERT was demethylated in atherosclerosis patients compared to healthy individuals, which corresponded with a decrease in telomere length and hTERT mRNA levels [35]. Although the mRNA changes observed in this study did not match the longstanding relationship between the promoter DNA methylation and gene silencing, the authors found that hypomethylation of the hTERT promoter occurred within binding sites specific to CTCF, a transcriptional repressor. They therefore speculated that demethylation of the hTERT promoter, in this scenario, may promote the recruitment of this repressor and consequently block hTERT transcription.
The glucocorticoid receptor is a critical regulator of the HPA axis, a neuroendocrine pathway associated with the stress response [109]. Genetic and epigenetic alterations within the NR3C1 gene, encoding the glucocorticoid receptor, have been implicated in neuropsychiatric diseases and more recently, a range of cardiometabolic disorders [110]. A study by Zou et al. (2015) investigated the relationship between promoter methylation of NR3C1 and brachial artery flow-mediated dilation as a measure of subclinical atherosclerosis in the peripheral blood leukocytes from 84 monozygotic male twin pairs from the Vietnam Era Twin Registry. After adjusting for covariates (smoking, physical activity, body mass index, lipids, blood pressure, and fasting glucose) and multiple testing, they identified 12 CpG sites in the NR3C1 exon 1F promoter that displayed intra-pair differences in DNA methylation, which positively correlated with brachial artery flow-mediated dilation. The authors speculated that altered glucocorticoid signaling, through dysregulated NR3C1 methylation, may result in endothelial dysfunction through the disruption of the HPA axis, which influences metabolic signaling [38].

2.2.5. Genes Involved in Endothelial Function

NPPA encodes a peptide hormone that is produced within the atria of the heart. Polymorphisms within the NPPA gene have been associated with CHD, atherosclerosis, and ischemic stroke, and functional studies have indicated that NPPA promotes diuresis, natriuresis, and vasodilatation, and may thus play fundamental roles in the pathogenesis of CVDs through these mechanisms [111,112]. In a prospective, longitudinal study conducted in China, 2498 middle-aged and elderly participants were followed for a total of 10 years, after which 263 participants developed CVD-related events. Using bisulfite sequencing, the authors investigated the DNA methylation status of nine CpG sites within the proximal promoter of the NPPA gene in peripheral blood. The results revealed that hypermethylation of a CpG site situated 459 base pairs upstream of the transcriptional start site was significantly associated with reduced CVD risk, which was independent of age, sex, education level, tobacco and alcohol use, body mass index, fasting glucose levels, and circulating lipids [111,112]. In this regard, the methylation status of NPPA, as proposed by the authors, may be useful in predicting the risk of lifestyle diseases, including CVDs.

2.3. Genome-Wide Studies

The advent of next-generation sequencing has led to a major paradigm shift in epigenomic research, facilitating a hypothesis-free assessment of DNA methylation signatures at a genome-wide scale [113]. Almost a decade ago, a study published by Lister et al. (2009) described, for the first time, the use of Whole-Genome Bisulfite Sequencing (WGBS) to generate comprehensive DNA methylation maps of the entire human genome at single-base pair sites [114]. Now, whole-genome sequencing is routinely utilized to identify genes and genomic regions that are differentially methylated in relation to disease onset and progression. These sequencing platforms include Infinium Beadchip Arrays (HumanMethylation27, HumanMethylation450, and HumanMethylationEPIC arrays), sequencing of immunoprecipitated methylated DNA, and whole-genome bisulfite sequencing [115]. This section will provide an overview of genome-wide, differentially methylated CpG regions associated with cardiovascular outcomes, which is summarized in Table 3. Of the 22 studies included 1 study utilized a methylation specific restriction-enzyme based microarray, 2 studies conducted methyl-binding domain-capture sequencing (using the Illumina HISEq 2500), 2 studies carried out reduced representation bisulfite sequencing (RRBS), while 4 studies used the HumanMethylationEPIC BeadChip array, and 13 studies used the older HumanMethylation450 BeadChip array.
Using a CpG island microarray approach, Sharma et al. (2014) identified hypermethylation at 19 CpG regions in the peripheral blood mononuclear cells from male patients with coronary artery disease (CAD) compared to controls [43]. The putative biological function(s) of the differentially methylated genes were further investigated by gene ontology analysis, which revealed an enrichment of genes involved in signal transduction, transcriptional regulation, organelle development, transport, and cell regulation. Of the 19 differentially methylated regions identified, 12 selected regions were subjected to bisulfite sequencing using samples from an independent, validation cohort comprising 48 CAD cases and 48 control individuals. The results showed hypermethylation of six CpG sites mapping to intronic regions within the STE20-related kinase adaptor alpha (STRADA) gene and Heat Shock Protein 90 Beta Family Member 3 Pseudogene (HSP90B3P), as well as exon 1 of the Complement C1q Like 4 (C1QL4) gene. While there is a paucity of information on the roles of STRADA and HSP90B3P in CVD development, C1QL4 has previously been implicated in cardiovascular system development through its ability to promote the angiogenesis of endothelial cells [116].
Using the HumanMethylation450 platform and two different population cohorts, Guarrera and colleagues (2015) investigated DNA methylation profiles in white blood cells and their association with myocardial infarction occurrence [24]. They identified a differentially methylated region within the gene body of Zinc Finger and BTB Domain Containing 12 (ZBTB12), which was hypermethylated in myocardial infarction cases compared to controls. Furthermore, the association of this differentially methylated mark was more pronounced in subjects with a shorter time to disease, revealing its potential as a prognostic marker. A study by Rask-Andersen and colleagues (2016) performed genome-wide DNA methylation analysis on blood samples from a cross-sectional study population cohort from Northern Sweden. This cohort included individuals with hypertension, myocardial infarction, stroke, thrombosis, and cardiac arrhythmia. They found that 211 CpG sites, representing 196 genes, were differentially methylated in individuals with a history of myocardial infarction. Gene ontology analysis revealed an enrichment of genes involved in neuronal systems, transmembrane ion transport, and synaptic signaling. Furthermore, a large proportion of differentially methylated loci (n = 42) mapped to genes previously associated with cardiac function and CVDs in the literature, such as Dysferlin (DYSF) and Neuregulin 1 (NRG1), supporting the biological relevance and candidacy of these methylation marks as potential biomarkers for myocardial infarction [44].
Armstrong (2018) studied the DNA methylation patterns of 180 individuals (aged 35 years and older) in a multicenter, double-blinded, randomized controlled clinical study involving the Vitamin Intervention for Stroke Prevention cohort. The study population comprised Europeans (n = 104) and African Americans (n = 76), all of whom had suffered recurrent episodes of ischemic stroke within 120 days of enrolment and displayed high plasma homocysteine levels. Data from this study revealed two highly significant hypermethylated CpG sites, mapping to the ASB10 (cg22812874) and TTC37 (cg00340919) genes, which were associated with an increased risk of stroke. Of particular interest, this association was found exclusively in the European subgroup but was absent in the African American subgroup [45]. Moreover, since the European patients suffered recurrent stroke episodes, it is still unknown whether the DNA methylation changes observed in ASB10 and TTC37 were causal or a consequence of ischemic stroke.
Age is an established and independent cardiovascular risk factor that autonomously alters epigenetic signatures. In a study conducted by Fernández-Sanlés (2018), the author investigated the influence of age on DNA methylation in individuals with atherosclerotic traits using whole blood from two independent longitudinal cohorts, the Registre Gironí del Cor (REGICOR) cohort and the Framingham Offspring Study (FOS) cohort [46]. Using the Illumina HumanMethylation (450K) BeadChip array, they identified 48 differentially methylated regions associated with cardiovascular risk in the REGICOR study. These signatures were further validated in the FOS cohort, and findings revealed the hypomethylation of six CpG loci within four genes, namely aryl hydrocarbon receptor repressor (AHRR, cg05575921), carnitine palmitoyltransferase 1A (CPTIA, cg00574958), peptidylprolyl isomerase F (PPIF, cg12547807, cg19939077, cg18608055), and strawberry notch homolog 2 (SBNO2, cg18608055), which correlated with increased cardiovascular risk after adjusting for age. Importantly, post-analysis showed that the DNA methylation of these CpG loci was able to explain 8.53% and 15.6% of the CVD risk in the REGICOR and FOS cohorts, respectively. This study also identified an association between hypomethylation of these CpG loci and several CVD risk factors, including smoking (AHRR and PPIF), BMI (PPIF), and the combination of BMI, diabetes, and triglyceride levels (CPTIA). BMI additionally correlated inversely with the methylation of a single CpG site mapping to the Growth/Differentiation Factor-15 (GDF-15) gene, an established CVD biomarker with well-reported roles in inflammation. Indeed, GDF-15 has been shown to be overexpressed in the myocardium and blood of patients with acute myocardial infarction [117,118] and its levels are associated with increased risk of death in individuals with CHD and chronic heart failure [119].
Bain and colleagues (2020) performed a pilot investigation of the DNA methylation differences in the peripheral blood from 20 male CAD patients with extreme heart failure phenotypes compared to 20 age-, sex-, and ethnicity-matched CAD patients without these extreme phenotypes. Using a methyl-binding domain-capture sequencing approach, the authors identified 68 differentially methylated regions associated with extreme heart failure. Gene enrichment analysis identified 103 significantly enriched gene sets, of which hypomethylated CpGs annotated to histone deacetylase 9 (HDAC9), jumonji and AT-rich interaction domain containing 2 (JARID2), and gremlin 1 (GREM1), whilst hypermethylated CpGs annotated to decaprenyl diphosphate synthase subunit 2 (PDSS2) [47]. An additional study by Gao (2021), followed by Ammous (2022), made use of the Infinium HumanMethylationEPIC Beadchip 850K array to identify novel methylated biomarkers that can predict the risk of developing CVDs in Chinese and African American populations, respectively. In the study by Gao (2021), eight CpGs were differentially methylated, with one hypermethylated CpG [annotating to cytochrome P450 family 8 subfamily B member 1 (CYP8B1, cg07655795)], being associated with increased CVD risk, whilst hypomethylation of the remaining seven CpGs [annotating to spondin 1 (SPON1, g06901278, cg11651314), phosphofurin acidic cluster sorting protein 1 (PACS1, cg03914662), uveal autoantigen with coiled-coil domains and ankyrin repeats (UACA, cg09306458, cg05359217), and coiled-coil domain containing 50 (CCDC50, cg05946546, cg02518222)], associated with an increased risk of developing CVDs [48]. Furthermore, Ammous et al. (2022) identified a significant correlation between hypomethylation at four CpGs, annotating to aryl hydrocarbon receptor repressor (AHRR, cg05575921 and cg21161138), TIMP metallopeptidase inhibitor 3 (GF11, cg09936388), and leucine rich repeat containing 52 (LRRC52, cg181168448), and atherosclerosis risk in an African American population. Interestingly, two of these identified CpGs (cg05575921 and cg21161138, annotating to AHRR,) were also similarly associated with CVDs in two independent studies by Fernández-Sanlés (2018, 2021) in American and European populations [46,50].
A growing list of epigenome-wide studies have been used to identify DNA methylation changes associated with serum lipids, which remain the most established risk factors for CVDs. Hedman (2017) proposed that CHD-associated DNA methylation changes may not necessarily be causal to CHD, but a consequence of previous insults on the myocardium, such as lipotoxicity. To test this, the authors conducted an epigenome-wide association study using the Illumina HumanMethylation450K BeadChip array to identify the DNA methylation alterations that are associated with lipid concentrations, which may serve as potential predictors of CHD events [51]. Findings from the discovery study identified 193 CpGs in the two independent population cohorts, of which 33 could be validated in an additional 2025 study participants. Of these, 25 CpG sites showed novel associations with lipids, including those annotated to the ATP binding cassette subfamily G member 1 locus (ABCG1, cg27243685), a gene associated with reverse cholesterol transport, which subsequently correlated with a 38% higher risk of incident CHD per every standard deviation increase in methylation [51]. Interestingly, this gene has also previously been identified as aberrantly methylated in the peripheral blood from type 2 diabetic individuals compared to healthy individuals [64].
To determine the association between DNA methylation and body mass index (BMI), Mendelson et al. (2017) utilized whole blood-derived DNA samples collected from patients enrolled in the Framingham Heart Study (FHS) and Lothian Birth Cohorts (LBCs), and validated their findings using samples collected from patients in the Atherosclerosis Risk in Communities (ARICs), Genetics of Lipid Lowering Drugs and Diet Network (GOLDN), and Prospective Investigation of the Vasculature in Uppsala Seniors (PIVUSs) studies [52]. The study reported a three-way association between 11 CpG sites, the expression of seven genes, and BMI. Of the identified CpGs, two annotated to CPT1A (cg00574958 and cg17058475), one annotated to ABCG1 (cg01881899), while one annotated to sterol regulatory element binding transcription factor 1 (SREBF1, cg11024682), with the latter shown to associate with increased adiposity and CAD. The methylation of the cg11024682 site in the SREBF1 gene has previously been associated with lipid measures in other studies, suggesting that the role of SREBF1 in lipid metabolism, in the context of cardiometabolic risk, may be alleviated by modulation of its methylation profile [120].
A study by Allum (2019) investigated the association between cardiometabolic risk-linked epigenetic loci and circulating plasma lipid levels in matched adipose tissue and whole blood samples from participants with obesity from the Quebec Heart and Lung Institute (IUCPQ) study. Using methylC-capture-sequencing, they analyzed 1.3 million CpG sites mapping to adipose tissue-specific regulatory regions. After adjusting for sequencing depth, age, and BMI, the authors identified 1230 CpG sites associated with at least one plasma lipid trait, i.e., triglycerides, HDL-cholesterol, LDL-cholesterol, and total cholesterol [53]. These lipid-associated CpG sites were annotated to 264 putative adipose tissue-specific enhancers and 303 promoters, of which 341 were shared elements between other tissues and adipose tissues, and 226 were specific to adipose tissue. These differentially methylated regulatory regions annotated to several obesity-related loci, including tyrosine-protein kinase (CSK), solute carrier organic anion transporter family member 3A1 (SLCO3A1), G protein subunit gamma 7 (GNG7), G protein subunit alpha i1 (GNAI5), lipocalin 2 (LCN2), enoyl-CoA hydratase short chain 1 (ECHS1), isocitrate dehydrogenase 2 (IDH2), and CD7 molecule (CD7). Next, they aimed to determine whether these signatures could be replicated in matched whole blood from the same sample population. Out of the 565 regulatory regions identified in the adipose tissue, 340 were reflected in the whole blood, of which 68 (mapping to 13 putative enhancers and 55 promoters) shared the same associations with plasma lipid traits, including those annotating to serine/threonine kinase AKT1, histone deacetylase HDAC4, bone morphogenetic protein 4 (BMP4), growth differentiation factor 7 (GDF7), ceramide kinase (CERK), vestigial like family member 3 (VGLL3), and ATP-binding cassette transporter 5 (ABCC5). Of the 68 CpG sites identified in the IUCPQ cohort, 22 could be further validated in the whole blood of an independent cohort. Taken together, these findings offer distinct characteristics of both tissue-specific and tissue-independent regulatory regions associated with lipid profiles and cardiometabolic risk [53].
C-reactive protein (CRP), an established marker of heightened inflammation, has been associated with CVDs and stroke. In a recent study by Chilunga (2021), the impact of DNA methylation loci, specifically those linked to CRP, on the development of CVDs was explored in a population of 589 African individuals with type 2 diabetes. The study highlighted 14 previously unidentified differentially methylated loci associated with CRP levels reaching up to 40 mg/L, with three of the differentially methylated CpGs [annotating to pyruvate carboxylase (PC, cg14653250), BTG anti-proliferation factor 4 (BTG4, cg13767940), and peptidyl arginine deiminase 1 (PAD1, cg21010178)] demonstrating a link with CVD risk [54]. While BTG4 and PADI1 have not yet been linked to CVD-related events, the PC gene has been shown to play an important role in gluconeogenesis, lipogenesis, and insulin secretion and has recently been associated with T2D [121].
In an attempt to assess the blood-based DNA methylation profiles of atherosclerosis patients enrolled in the Left Ventricular Assist Device Study (LVAD), Istas and colleagues (2017) conducted genome-wide DNA methylation sequencing using the Illumina Infinium HumanMethylation450 array, in the whole blood of eight atherosclerosis patients and eight healthy volunteers. After adjusting for age, they identified 287 hypo- and 229 hypermethylated CpG sites in cases versus controls, which mapped to genes involved in the cell cycle, cell adhesion, and cell death [26]. They validated the DNA methylation of nine CpG sites in BRCA1 DNA repair associated 1 (BRCA1) and seven CpG sites in cysteine-rich secretory protein 2 (CRISP2), using an Epityper MassARRAY approach. In addition, they identified associations between three CpG sites in CRISP2 (cg12440062, cg25390787, and cg01076129) and one CpG site in BRCA1 (cg16630982), as well as subclinical atherosclerosis measures in an independent cohort comprising 24 participants from the Aragon Workers Health Study (AWHS). While BRCA1 is widely recognized as a tumor suppressor, recent evidence suggests that it may act as an anti-inflammatory protein to retard endothelial dysfunction and atherosclerosis [122]. Data on CRISP2 methylation and its role in atherosclerosis is a novel concept.
Using the Illumina Infinium HumanMethylation850K BeadChip array, Antoun (2022) investigated DNA methylation signatures in the whole blood of 293 Gambian and 698 Indian children. Several differentially methylated regions were identified for each cohort and were associated with systolic blood pressure [annotating to mediator complex subunit 22 (MED22, g13455829) and Egl-9 family hypoxia inducible factor 2 (EGLN2, cg22671726)], fasting insulin [annotating to the gene bodies of family with sequence similarity 46 member A (FAM46A, cg22388948) and LysM domain containing 2 (LYSMD2, cg13934266)], circulating triglycerides (cg15237100 located in an intergenic region), and LDL-cholesterol (cg01469688 and cg06952751, mapping to the promoter regions of suppressor of cancer cell invasion (SCAI) and C18orf8, respectively) in the Gambian cohort, and with insulin sensitivity [mapping to the gene body of transmembrane protein 57 (TMEM57, cg10304969)] and HDL-cholesterol [cg04988216 annotating to the gene body of Receptor Tyrosine Kinase Like Orphan Receptor 1 (ROR1)] in the Indian cohort. Interestingly, the differentially methylated CpGs identified in the study were unique to each cohort and no overlap was observed for any CpGs between groups. Since DNA methylation at specific CpGs can be influenced by genotype, the authors also identified single nucleotide polymorphisms (SNPs) within an approximately 2 MB region from the CpGs of interest that were associated with the differential methylation. Significant quantitative trait loci (methQTLs) were found for three differentially methylated CpGs associated with LDL-cholesterol in Gambians; however, methylation did not mediate genotype effects on CVD outcomes in the Indian population [55]. Therefore, the authors argued that DNA methylation patterns could be population-specific, and that environmental and ethnic differences could influence these changes.
Agha (2019) conducted a longitudinal multi-cohort epigenome-wide search for DNA methylation alterations associated with CHD events across nine different cohorts, comprising 11273 study participants from Europe and the United States. A total of 442192 CpGs were analyzed for all nine cohorts, of which DNA methylation at 52 CpG sites was associated with CHD and myocardial infarction incidence. These CpG sites mapped to key genes associated with calcium flux [including those mapping to ATPase Plasma Membrane Ca2+ Transporting 2 (ATP2B2), calcium sensing receptor (CASR), guanylate cyclase activator 1B (GUCAIB), and hippocalcin like 1 (HPCAL1)], arterial plaque calcification [such as protein tyrosine phosphatase receptor type N2 (PTPRN2)], and kidney function [including CpG sites mapping to cadherin related 23 (CDH23) and hippocalcin-like 1 (HPCAL1)]. Interestingly, and in contrast to the findings of Antoun (2022), the associations between the 52 CpG sites and CAD incidence were consistent across all cohorts in the study. Moreover, these associations were not influenced by ethnicity, as they remained similar when comparing European groups to those of African American ancestry [56].
Using reduced representation bisulfite sequencing (RRBS), Li et al. (2017) assessed the DNA methylation profile in the blood leukocytes of 27 heart failure patients compared to 20 healthy controls. They identified three genes, solute carrier family 2 member 1 (SLC22A1), MPV17 mitochondrial inner membrane protein like (MPV17L), and plectin (PLEC), as differentially methylated in the heart failure patients compared with the healthy controls, with the DNA methylation of these genes being inversely associated with their mRNA expression levels [57]. While the role of SLC22A1 in heart failure is not well established, it has been shown to promote tumor cell proliferation and glycolysis in cancer [123]. An increase in glycolysis is believed to reduce cardiac hypertrophy and dysfunction induced by pressure overload, suggesting that SLC22A1 may be beneficial in cardiac function, but this requires further investigation [124]. On the other hand, MPV17L has previously been shown to prevent pancreatic β-cell hyperplasia, suggesting that the protein may play a critical role in preventing diabetes-related cardiac dysfunction [125]. The PLEC gene encodes a cytoskeletal linker protein that has been previously implicated in cardiomyopathy [126].
Using the Illumina HumanMethylation450K BeadChip system, Huan (2019) studied the methylome of 4170 participants of European ancestry enrolled in the FHS, after which findings were replicated in two independent populations from the Atherosclerosis Risk in Communities (ARICs) and Grady Trauma Project (GTP) study cohorts. Their collective findings highlighted DNA methylation at 92 CpGs as a putative causal factor for cardiovascular disease and myocardial infarction, with some of the CpGs annotating to genes linked to heart failure [58]. Indeed, lysosomal acid lipase (LIPA) has been linked to hypercholesterolemia [127], whilst ABO and serologically defined colon cancer antigen 8 (SDCCAG8) have been identified as genes underlying the development of CHD and diastolic blood pressure, respectively [128,129].
Using monozygotic and dizygotic twin pairs from a Swedish population, Qin et al. (2021) identified 20 top-ranked CpG sites associated with non-stroke CVDs, overall stroke, and ischemic stroke [59]. The methylation of these CpGs was shown to determine the levels of cardiometabolic traits, such as BMI and blood pressure, with the latter mediating CVD risk. Similarly, using RRBS, Koseler and colleagues assessed the DNA methylation differences of thousands of CpG sites between one pair of monozygotic twins discordant for CVDs [60]. Here, DNA methylations, annotating to the genes involved in fatty acid and cholesterol metabolism [namely, lipid droplet associated hydrolase (LDAH), apolipoprotein B (APOB), acyl-CoA synthetase members 2A, 5, and 3 (ACSM2A, ACSM5, and ACSM3), carboxylesterase 1 (CES1), carboxylesterase 1 pseudogene 1 (CES1P1), AFG3 like matrix AAA peptidase subunit 2 (AFG3L2), iron-sulfur cluster assembly enzyme (ISCU), SEC14 like lipid binding 2 (SEC14L2), and microsomal triglyceride transfer protein (MTTP)], were hypomethylated in the healthy twin as compared with the twin presenting with premature CAD. Some of these genes, including APOB, have been previously reported to be strong drivers of myocardial infarction risk [130]. The authors noted that the DNA methylation differences observed between the twins could be attributed to the influence of lifestyle or statin treatment [60].
To identify the association between DNA methylation and myocardial infarction, Nakatochi and colleagues (2017) used samples collected from the whole blood of elderly male Japanese subjects with a history of myocardial infarction, as compared with a control group. The control group was chosen from an ongoing Kita-Nagoya Genomic Epidemiology (KING) study and was selected based on their nondiabetic status [61]. Data revealed an association between two CpG sites [mapping to zinc finger homeobox 3 (ZFHX3, cg07786668) and SWI/SNF-related matrix-associated actin-dependent regulator of chromatin subfamily a member 4 (SMARCA4, cg17218495)] and myocardial infarction, which was independent of CVD risk factors, such as BMI, blood lipid levels, and type 2 diabetes. While ZFHX3 has been implicated in the pathophysiology of obesity, SMARCA4 has been shown to regulate cholesterol levels, suggesting that the methylation of both genes could be essential biomarkers of myocardial infarction risk [131,132].
Rask-Andersen et al. (2016) conducted an epigenome-wide association study to identify differentially methylated genes in the peripheral blood from participants from the Northern Sweden Population Healthy Study cohort, comprising individuals with a history of myocardial infarction. They identified altered methylation at 211 CpG sites, which correlated with myocardial infarction, a correlation that remained independent of BMI [44]. Some of these CpG sites annotated to genes previously implicated in cardiovascular diseases, including axonally derived neuregulin 1 (NRG1), known to play a role in myocardial repair following myocardial infarction [44]. Moreover, they reported that the methylation and expression of dysferlin (DYSF), encoding a protein involved in membrane integrity of skeletal muscle cells, provided a strong predictive value for atherosclerotic cardiovascular disease in their cohort [133,134]. Finally, Ward-Caviness and colleagues (2018) attempted to develop an epigenetic fingerprint predictive of myocardial infarction incidence in the Cooperative Health Research in the Augsburg Region (KORA) cohort [62]. Using the Illumina Infinium HumanMethylation450 BeadChip, they identified the methylation of 11 CpGs that related to incidents of myocardial infarction, although following postadjustment for medication, only 9 CpGs retained statistical significance. Importantly, these modifications were shown to be generalized across various ethnicities and geographical regions. Of these, two CpGs were annotated to low-density lipoprotein receptor-related protein 8 (LRP8) and potassium calcium-activated channel subfamily N member 1 (KCNN1), genes shown by others to be associated with CVDs [135,136]. Indeed, a previous study showed that phosphorylation-induced activation of LRP8, upon binding to LDL, leads to the activation of p38 mitogen-activated protein kinase, a critical protein involved in myocardial injury during ischemia and reperfusion [135]. The KCCN genes encode for small-conductance Ca2+-activated K+ channels that are known to play a key role in cardiac excitability [137].

3. Future Perspectives and Conclusions

The pathogenesis of CVDs involves a complex interaction between genetic, environmental, and lifestyle factors, and there has been some success in identifying DNA methylation alterations that are associated with these diseases. Screening individuals using non-invasive methods may represent opportunities for disease prevention and intervention. The current review aimed to provide a comprehensive analysis of the latest reported DNA methylation alterations as blood-based biomarkers for CVDs. The review included forty-eight studies covering various CVD phenotypes, conducted in five continents (Europe, Asia, North and South America, and Africa), featuring diverse populations, age groups, blood cell types, and study designs. While no strong consensus exists on the association between global DNA methylation and CVDs, candidate gene and genome-wide studies collectively identified a repertoire of over 100 genes with methylation statuses associated with CVDs, their risk factors, or related conditions (Table 4). These genes included FOXP3, IL-6, CPT1A, ABCG1, and AHRR, which were differentially methylated in individuals with CVDs compared to healthy individuals in at least two or more studies. Moreover, these five genes were consistently differentially methylated across various populations, methodological approaches, and blood cell types, supporting their candidacy as biomarkers of CVDs.
Most studies covered in this review performed their analyses in whole blood, utilized repetitive elements as proxies for quantification of global DNA methylation, applied bisulfite sequencing or pyrosequencing for candidate gene studies, and employed the Illumina Infinium HumanMethylation450K for measuring DNA methylation in genome-wide studies. Due to practical and ethical reasons, the procurement of tissues central to CVD development, such as biopsies of the myocardium or blood vessel walls, remains difficult due to the invasive procedures required for their collection. In this regard, peripheral blood can be collected using minimally invasive, inexpensive methods and has been shown to be a reliable source of DNA methylation information, possibly due to the cardiovascular system being significantly regulated via the blood and its constituents. Indeed, a previous study reported a high correlation between the DNA methylation in the peripheral blood leukocytes and atrial biopsies collected from patients undergoing coronary bypass surgery [138], thus supporting the utility of blood-based DNA methylation as a proxy for cardiac cells. However, it is important to note that whole blood contains various cell types, which may lead to variability in DNA methylation signatures, based on inter-individual differences in the fraction of DNA-containing blood cells collected in studies. There are several methods that can be used to account for these confounding effects, including the flow-cytometry-based sorting of blood cells and subsequent quantification of the DNA methylation in the resulting individual cell types, or the post hoc adjustment of cell count based on published regression models, such as that reported by Houseman et al. (2012) [139].
The global DNA methylation studies covered in this review presented contradictory findings, with DNA hypomethylation associated with CVD events in six studies, hypermethylation associated with the same CVD traits in five studies, and one study reporting no significant differences in the global DNA methylation levels in individuals with atherosclerosis compared to healthy controls. These reports are consistent with other reviews on the relationship between DNA methylation and metabolic diseases, including dyslipidaemia, type 2 diabetes, and CVDs [64,140,141,142]. These inconsistencies might be explained, in part, by heterogeneous methodologies applied by the different studies, including differences in sample size, methods used for DNA isolation and methylation quantification, and biological sources (i.e., whole blood versus PBMCs), as well as adjustments for confounding factors. Moreover, it is widely accepted that the influence of DNA methylation on gene expression is dependent on its genomic location, and thus, methylation at specific disease-causing genes/loci may be a more informative measure of health-related outcomes compared to the global, pooled analysis of methylated cytosines within the genome.
Most of the studies included in this review utilized either a candidate gene or a genome-wide approach. For candidate gene studies, the DNA methylation of targeted regions was selected based on hypotheses generated from previous genetic or biological studies. Only two genes, IL-6 and FOXP3, were examined and replicated in more than one study; therefore, most of the remaining candidate gene findings require additional replication. IL-6 encodes a pro-inflammatory cytokine and a growing list of studies have implicated this gene in CVD events [143]. It is speculated that, under atherosclerotic conditions, hypomethylation of the IL-6 promoter would cause upregulation of its expression, leading to a higher production of the cytokine by macrophages and consequent vascular inflammation. FOXP3 encodes a transcription factor that is responsible for lineage specification, cell maturation, and the function of regulatory T-cells (Tregs). Studies in numerous experimental models have shown that Tregs play a powerful protective role in atherosclerosis. For example, Kligenberg and colleagues (2013) showed that the depletion of FOXP3+ Treg cells resulted in a 2.1-fold increase in plaque size and hypercholesterolemia in mice [144]. We, therefore, postulate that in humans, hypermethylation of FOXP3, as observed by Zhu (2019) and Jia (2013), and reported in this review, may result in suppression of the gene and concomitant depletion of Treg cells, resulting in a decreased protective effect of these cells against the development of atherosclerosis [31,32].
This review also included 22 genome-wide DNA methylation studies. This approach has become increasingly popular and preferred over candidate gene methods as it is more effective in detecting novel disease-associated DNA methylation marks. However, the drawbacks include increased cost, the need for large sample sizes, as well as the requirement for validation in independent replication cohorts [145]. Most genome-wide studies covered here validated their results in one or more replication cohorts, however, some used relatively small sample sizes. Amongst the differentially methylated loci identified, CpGs mapping to three genes (cg00574958 in intron 1 of CPT1A, cg27243685 in the 5′UTR of ABCG1, and cgo5575921 in the gene body of AHRR) were differentially methylated in more than one genome-wide study. Moreover, while the differentially methylated CpG sites within CPT1A and ABCG1 do not reside in promoter regions, we note that these patterns were reported to be accompanied by the altered expression of these genes, suggesting that gene expression in this context, may be regulated by methylation at distal sites.
CPT1A, encoding a constituent of the carnitine palmitoyltransferase (CPT) system, is an important rate-limiting enzyme involved in the conversion of cytoplasmic long-chain acetyl-CoA to acylcarnitine, which in turn, enters the mitochondria for fatty acid β-oxidation (FAO) [146]. FAO is crucial for endothelial cell metabolism as it aids in resisting oxidative stress in the development of blood vessels. It is therefore not surprising that CPT1A has been proposed as a candidate target for the treatment of vascular-related diseases. In mice, the inhibition of CPT1A activity, and the consequent reduction of FAO, resulted in improved whole-body glucose tolerance and insulin sensitivity [147]. In addition, a CPT1A inhibitor, etomoxir, showed efficacy in improving outcomes for heart failure patients in a clinical trial; however, the trial was terminated early due to unacceptably high liver transaminase levels in enrolled patients [148]. ABCG1 encodes a protein involved in the efflux of cholesterol to high-density lipoprotein (HDL) in macrophages. As observed by Hedman (2017) and Mendelson (2017), hypermethylation of one CpG site located in the 5′UTR of ABCG1, as well as the concomitant reduced expression of ABCG1 mRNA, was associated with cardiovascular risk in the FHS, PIVUSs, GOLDN, LBCs, and ARICs cohorts [51]. We speculate that the downregulation of ABCG1 may contribute to CVD risk by lowering the capacity for reverse cholesterol transport, potentially leading to the accumulation of excessive cholesterol in the body and the development of atherosclerotic plaques [149]. The AHRR gene encodes for a repressor of the aryl hydrocarbon receptor, which amongst other roles, inhibits the metabolism of polycyclic aromatic hydrocarbons, a carcinogen found in cigarette smoke [150]. Interestingly, the DNA methylation of cg05575921 within AHRR is a well-documented marker of cigarette smoking, as well as overall mortality risk [151]. Two studies covered in this review implicated the DNA methylation of cg05575921 with CVD risk, both of which adjusted their data for smoking status. We speculate that the connection between cg05575921 and CVD risk in these studies may be a true association, or it may be due to residual confounding with smoking, failures in self-reporting of smoking, and/or inter-individual sensitivities to smoking, as well as its lasting biological effects.
Some limitations of this review should be acknowledged. First, whilst effort was made to conduct a comprehensive literature search, we cannot disregard the possibility of publication bias due to underreporting of negative findings. Our literature search was also conducted in only three databases and excluded studies that were not published in English. Most reviewed studies utilized a cross-sectional design, and due to the unstable nature of the epigenome, we cannot ascertain whether the identified signatures are a cause or consequence of CVDs. Moreover, only DNA methylation modifications in CVDs were reviewed and we therefore cannot rule out the effect of other epigenetic marks, such as posttranslational histone modifications and non-coding RNAs, as biomarker candidates of CVDs. A key limitation of many of the reviewed studies was the use of small sample sizes and a lack of adjustment for potential confounders. It is well established that factors such as sex, age, smoking, alcohol consumption, comorbidities, and medication use significantly influence DNA methylation profiles across numerous genes [152,153,154]. However, among the 46 studies analyzed, only 33 accounted for age. Ten studies were conducted exclusively in males, three exclusively in females, and three did not specify the participants’ sex. Of the 32 studies that included both sexes, only 23 adjusted for sex. The studies also encompassed a range of CVD phenotypes and methodological approaches, posing challenges for comparative analysis. Nevertheless, the DNA methylation of FOXP3, IL-6, CPT1A, ABCG1, and AHRR was consistently associated with CVDs across multiple studies. Importantly, each of these studies accounted for age, sex, and additional covariates, such as BMI, smoking, comorbidities, and blood cell count, reinforcing the reliability of these findings and supporting their potential as CVD biomarkers.
Our review also has several strengths. Due to the difficulty in accessing tissues central to the pathogenesis of CVDs, our review focused on associations between blood-based DNA methylation signatures and CVDs, and managed to consolidate findings from 48 research articles, identifying FOXP3, IL-6, CPT1A, ABCG1, and AHRR as promising biomarker candidates for CVDs (Figure 2). These findings may have important implications for the development of non-invasive epigenetic risk assessment tests for screening individuals for CVDs. We also highlighted differentially methylated genes that were associated with major risk factors for CVDs, including HDL/LDL-cholesterol levels, triglyceride concentrations, and BMI, as well as HbA1c and glucose levels, further supporting the notion that blood-based screening for DNA methylation could be utilized to monitor individuals who are at high risk of developing CVDs and who require medical and lifestyle interventions. In this context, the DNA methylation candidates identified in this review should be further evaluated in large-scale longitudinal studies using high-quality DNA methylation quantification platforms, with careful consideration of confounders, such as age and sex. To date, DNA methylation biomarkers for CVDs have yet to be assessed in a clinical trial setting, however, we anticipate that the continuous advancement of next generation sequencing methods will expedite the identification of novel predictive CVD biomarkers, and when implemented, will have a sizable public health impact.

Author Contributions

T.W. designed the study and extracted the data. T.W., R.J. and L.M. wrote the manuscript. C.J.F.M. and J.S. corrected the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the South African Medical Research Council (SAMRC) and the National Research Foundation CSUR (Grant number: 23030881391 awarded to TW). The content hereof is the sole responsibility of the authors and does not necessarily represent the official views of the SAMRC or the funders.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

5mC5-methylcytosine
5hmC5-hydroxymethylcytosine
AluArthrobacter luteus
ARICsAtherosclerosis Risk in Communities Study
ATAdipose tissue
AWHSAragon Workers Health Study
BMIBody mass index
BMPR2Bone morphogenic protein receptor 2
CaGCARTaGENE
CHDCoronary heart disease
CHSCardiovascular Health Study
CpGsCytosine-guanine dinucleotides
CRPC-reactive protein
CVDCardiovascular disease
DILGOMNational FINRISK Study
ELISAEnzyme-linked immunosorbent assay
EPICORsLong-Term Follow-Up of Antithrombotic Management Patterns in Acute Coronary Syndrome Patients
FABP3Fatty Acid Binding Protein 3
FHSFramingham Heart Study
FOXP3Forkhead/winged helix transcription factor 3
FOSFramingham Offspring Study
FPGFasting plasma glucose
GOLDNGenetics of Lipid Lowering Drugs and Diet Network
GTPGrady Trauma Project
GWASsGenome-wide association studies
HDLHigh-density lipoprotein
HFHeart failure
HOMA-IRHomeostatic model assessment for insulin resistance
hTERTHuman telomerase reverse transcriptase
IFN-γInterferon Gamma
IL-6Interleukin-6
iNOSInducible nitric oxide synthase
InCHIANTIThe Invecchiare in Chianti study
IUCPQQuebec Heart and Lung Institute
KORAKooperative Gesundheitsforschung in der Region Augsburg study
LBCsLothian Birth Cohorts
LDLLow-density lipoprotein
LINE-1Long interspersed nuclear element-1
LVADImpact of Left Ventricular Assist Devices Implantation on Micro- and Macrovascular Function
MCCMethyl-capture sequencing
MetSMetabolic syndrome
MIMyocardial infarction
MeDIPMethylated DNA immunoprecipitation
MiRNAMicroRNA
MZMonozygotic
NASNormative Aging Study
NFBCNorthern Finland Birth Cohort
NPPAAtrial natriuretic peptide
NR3C1Glucocorticoid receptor
NSPHSNorthern Sweden Population Health Study
PBPeripheral blood
PBLsPeripheral blood leukocytes
PBMCsPeripheral blood mononuclear cells
PIVUSsProspective Investigation of the Vasculature in Uppsala Seniors
REGICORGirona Heart Registry/REgistre GIroní del COR
SOAT1Sterol-O-Acyltransferase
SNPSingle nucleotide polymorph
T2DType 2 Diabetes
TCTotal cholesterol
TGsTriglycerides
TLR2Toll-like receptor 2
TSSTranscriptional start site
TXNIPThioredoxin-interacting protein
VATVisceral adipose tissue
VSIPVitamin Intervention for Stroke Prevention
VHDVascular heart disease
WBWhole blood
WBCsWhite blood cells
WHI-BAA23Women’s Health Initiative “Integrative genomics and risk of CHD and related phenotypes in the Women’s Health Initiative”
WHI-EMPCWomen’s Health Initiative “Epigenetic Mechanisms of Particulate Matter-Mediated CVD”
BPBlood plasma
HDL-CHigh-density lipoprotein cholesterol

References

  1. Vineis, P.; Stringhini, S.; Porta, M. The environmental roots of non-communicable diseases (NCDs) and the epigenetic impacts of globalization. Environ. Res. 2014, 133, 424–430. [Google Scholar] [CrossRef] [PubMed]
  2. Boutayeb, A.; Boutayeb, S. The burden of non communicable diseases in developing countries. Int. J. Equity Health 2005, 4, 2. [Google Scholar] [CrossRef] [PubMed]
  3. World Health Organization. Cardiovascular Diseases. 11 June 2021. Available online: https://www.who.int/news-room/fact-sheets/detail/cardiovascular-diseases-(cvds) (accessed on 30 May 2024).
  4. World Health Organization. Global Status Report on Non Communicable Diseases 2010; World Health Organization: Geneva, Switzerland, 2011; Chapter 1. [Google Scholar]
  5. Kontis, V.; Mathers, C.D.; Rehm, J.; Stevens, G.A.; Shield, K.D.; Bonita, R.; Riley, L.M.; Poznyak, V.; Beaglehole, R.; Ezzati, M. Contribution of six risk factors to achieving the 25×25 non-communicable disease mortality reduction target: A modelling study. Lancet 2014, 384, 427–437. [Google Scholar] [CrossRef] [PubMed]
  6. Marmot, M.; Friel, S.; Bell, R.; Houweling, T.A.; Taylor, S.; on behalf of the Commission on Social Determinants of Health. Closing the gap in a generation-health equity through action on the social determinants of health. Lancet 2008, 372, 661–669. [Google Scholar] [CrossRef]
  7. Sharp, G.C.; Relton, C.L. Epigenetics and noncommunicable diseases. Epigenomics 2017, 9, 789–791. [Google Scholar] [CrossRef]
  8. Eccleston, A.; DeWitt, N.; Gunter, C.; Marte, B.; Nath, D. Introduction Epigenetics. Nature 2007, 447, 395–396. [Google Scholar] [CrossRef]
  9. Jones, P.A. Functions of DNA methylation: Islands, start sites, gene bodies and beyond. Nat. Rev. Genet. 2012, 13, 484–492. [Google Scholar] [CrossRef]
  10. Kim, M.; Long, T.I.; Arakawa, K.; Wang, R.; Yu, M.C.; Laird, P.W. DNA methylation as a biomarker for cardiovascular disease risk. PLoS ONE 2010, 5, e9692. [Google Scholar] [CrossRef]
  11. Al Muftah, W.A.; Al-Shafai, M.; Zaghlool, S.B.; Visconti, A.; Tsai, P.-C.; Kumar, P.; Spector, T.; Bell, J.; Falchi, M.; Suhre, K. Epigenetic associations of type 2 diabetes and BMI in an Arab population. Clin. Epigenet. 2016, 8, 13. [Google Scholar] [CrossRef]
  12. Kasinska, M.A.; Drzewoski, J.; Sliwinska, A. Epigenetic modifications in adipose tissue—Relation to obesity and diabetes. Arch. Med. Sci. 2016, 12, 1293–1301. [Google Scholar] [CrossRef]
  13. Duthie, S.J. Epigenetic modifications and human pathologies: Cancer and CVD. Proc. Nutr. Soc. 2011, 70, 47–56. [Google Scholar] [CrossRef] [PubMed]
  14. Turunen, M.P.; Aavik, E.; Ylä-Herttuala, S. Epigenetics and atherosclerosis. Biochim. Biophys. Acta 2009, 1790, 886–891. [Google Scholar] [CrossRef] [PubMed]
  15. Cao, Y.; Lu, L.; Liu, M.; Li, X.-C.; Sun, R.-R.; Zheng, Y.; Zhang, P.-Y. Impact of epigenetics in the management of cardiovascular disease—A review. Eur. Rev. Med. Pharmacol. Sci. 2014, 18, 3097–3104. [Google Scholar]
  16. Heyn, H.; Esteller, M. DNA methylation profiling in the clinic: Applications and challenges. Nat. Rev. Genet. 2012, 13, 679–692. [Google Scholar] [CrossRef]
  17. Shenker, N.S.; Polidoro, S.; van Veldhoven, K.; Sacerdote, C.; Ricceri, F.; Birrell, M.A.; Belvisi, M.G.; Brown, R.; Vineis, P.; Flanagan, J.M. Epigenome-wide association study in the European Prospective Investigation into Cancer and Nutrition (EPIC-Turin) identifies novel genetic loci associated with smoking. Hum. Mol. Genet. 2013, 22, 843–851. [Google Scholar] [CrossRef] [PubMed]
  18. Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 statement: An updated guideline for reporting systematic reviews. Int. J. Surg. 2021, 88, 105906. [Google Scholar] [CrossRef]
  19. Castro, R.; Rivera, I.; Struys, E.A.; Jansen, E.E.; Ravasco, P.; Camilo, M.E.; Blom, H.J.; Jakobs, C.; de Almeida, I.T. Increased homocysteine and S-adenosylhomocysteine concentrations and DNA hypomethylation in vascular disease. Clin. Chem. 2003, 49, 1292–1296. [Google Scholar] [CrossRef]
  20. Sharma, P.; Kumar, J.; Garg, G.; Kumar, A.; Patowary, A.; Karthikeyan, G.; Ramakrishnan, L.; Brahmachari, V.; Sengupta, S. Detection of Altered Global DNA Methylation in Coronary Artery Disease Patients. DNA Cell Biol. 2008, 27, 357–365. [Google Scholar] [CrossRef]
  21. Baccarelli, A.; Wright, R.; Bollati, V.; Litonjua, A.; Zanobetti, A.; Tarantini, L.; Sparrow, D.; Vokonas, P.; Schwartz, J. Ischemic Heart Disease and Stroke in Relation to Blood DNA Methylation. Epidemiology 2010, 21, 819–828. [Google Scholar] [CrossRef]
  22. Lin, R.-T.; Hsi, E.; Lin, H.-F.; Liao, Y.-C.; Wang, Y.-S.; Juo, S.-H.H. LINE-1 methylation is associated with an increased risk of ischemic stroke in men. Curr. Neurovascular Res. 2014, 11, 4–9. [Google Scholar] [CrossRef]
  23. Wei, L.; Liu, S.; Su, Z.; Cheng, R.; Bai, X.; Li, X. LINE-1 hypomethylation is associated with the risk of coronary heart disease in Chinese population. Arq. Bras. Cardiol. 2014, 102, 481–488. [Google Scholar] [CrossRef]
  24. Guarrera, S.; Fiorito, G.; Onland-Moret, N.C.; Russo, A.; Agnoli, C.; Allione, A.; Di Gaetano, C.; Mattiello, A.; Ricceri, F.; Chiodini, P.; et al. Gene-specific DNA methylation profiles and LINE-1 hypomethylation are associated with myocardial infarction risk. Clin. Epigenet. 2015, 7, 133. [Google Scholar] [CrossRef]
  25. Ramos, R.B.; Fabris, V.; Lecke, S.B.; Maturana, M.A.; Spritzer, P.M. Association between global leukocyte DNA methylation and cardiovascular risk in postmenopausal women. BMC Med. Genet. 2016, 17, 71. [Google Scholar] [CrossRef]
  26. Istas, G.; Declerck, K.; Pudenz, M.; Szic, K.S.V.; Lendinez-Tortajada, V.; Leon-Latre, M.; Heyninck, K.; Haegeman, G.; Casasnovas, J.A.; Tellez-Plaza, M.; et al. Identification of differentially methylated BRCA1 and CRISP2 DNA regions as blood surrogate markers for cardiovascular disease. Sci. Rep. 2017, 7, 5120. [Google Scholar] [CrossRef]
  27. Alexeeff, S.E.; Baccarelli, A.A.; Halonen, J.; Coull, B.A.; Wright, R.O.; Tarantini, L.; Bollati, V.; Sparrow, D.; Vokonas, P.; Schwartz, J. Association between blood pressure and DNA methylation of retrotransposons and pro-inflammatory genes. Int. J. Epidemiol. 2013, 42, 270–280. [Google Scholar] [CrossRef]
  28. Jiang, D.; Wang, Y.; Chang, G.; Duan, Q.; You, L.; Sun, M.; Hu, C.; Gao, L.; Wu, S.; Tao, H.; et al. DNA hydroxymethylation combined with carotid plaques as a novel biomarker for coronary atherosclerosis. Aging 2019, 11, 3170–3181. [Google Scholar] [CrossRef]
  29. Jiang, D.; Sun, M.; You, L.; Lu, K.; Gao, L.; Hu, C.; Wu, S.; Chang, G.; Tao, H.; Zhang, D. DNA methylation and hydroxymethylation are associated with the degree of coronary atherosclerosis in elderly patients with coronary heart disease. Life Sci. 2019, 224, 241–248. [Google Scholar] [CrossRef]
  30. Li, N.; Zhu, L.; Zhu, C.; Zhou, H.; Zheng, D.; Xu, G.; Shi, H.; Gao, J.; Li, A.J.; Wang, Z.; et al. BMPR2 promoter methylation and its expression in valvular heart disease complicated with pulmonary artery hypertension. Aging 2021, 13, 24580–24604. [Google Scholar] [CrossRef]
  31. Zhu, L.; Jia, L.; Liu, Z.; Zhang, Y.; Wang, J.; Yuan, Z.; Hui, R. Elevated Methylation of FOXP3 (Forkhead Box P3)-TSDR (Regulatory T-Cell-Specific Demethylated Region) Is Associated with Increased Risk for Adverse Outcomes in Patients with Acute Coronary Syndrome. Hypertension 2019, 74, 581–589. [Google Scholar] [CrossRef]
  32. Jia, L.; Zhu, L.; Wang, J.Z.; Wang, X.J.; Chen, J.Z.; Song, L.; Wu, Y.J.; Sun, K.; Yuan, Z.Y.; Hui, R. Methylation of FOXP3 in regulatory T cells is related to the severity of coronary artery disease. Atherosclerosis 2013, 228, 346–352. [Google Scholar] [CrossRef]
  33. Krishnan, R.; Mani, P.; Sivakumar, P.; Gopinath, V.; Sekar, D. Expression and methylation of circulating microRNA-510 in essential hypertension. Hypertens. Res. 2017, 40, 361–363. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, Y.; Kent, J.W.; Lee, A.; Cerjak, D.; Ali, O.; Diasio, R.; Olivier, M.; Blangero, J.; A Carless, M.; Kissebah, A.H. Fatty acid binding protein 3 (fabp3) is associated with insulin, lipids and cardiovascular phenotypes of the metabolic syndrome through epigenetic modifications in a northern european family population. BMC Med. Genom. 2013, 6, 9. [Google Scholar] [CrossRef]
  35. Zhang, D.; Wen, X.; Wu, W.; Xu, E.; Zhang, Y.; Cui, W. Homocysteine-related hTERT DNA demethylation contributes to shortened leukocyte telomere length in atherosclerosis. Atherosclerosis 2013, 231, 173–179. [Google Scholar] [CrossRef] [PubMed]
  36. Zuo, H.-P.; Guo, Y.-Y.; Che, L.; Wu, X.-Z. Hypomethylation of Interleukin-6 Promoter is Associated with the Risk of Coronary Heart Disease. Arq. Bras. Cardiol. 2016, 107, 131–136. [Google Scholar] [CrossRef]
  37. Li, Z.; Yu, F.; Zhou, X.; Zeng, S.; Zhan, Q.; Yuan, M.; Yang, Q.; Liu, Y.; Xia, J. Promoter hypomethylation of microRNA223 gene is associated with atherosclerotic cerebral infarction. Atherosclerosis 2017, 263, 237–243. [Google Scholar] [CrossRef]
  38. Zhao, J.; An, Q.; Goldberg, J.; Quyyumi, A.A.; Vaccarino, V. Promoter methylation of glucocorticoid receptor gene is associated with subclinical atherosclerosis: A monozygotic twin study. Atherosclerosis 2015, 242, 71–76. [Google Scholar] [CrossRef] [PubMed]
  39. Mohammadpanah, M.; Heidari, M.M.; Khatami, M.; Hadadzadeh, M. Relationship of hypomethylation CpG islands in interleukin-6 gene promoter with IL-6 mRNA levels in patients with coronary atherosclerosis. J. Cardiovasc. Thorac. Res. 2020, 12, 221–228. [Google Scholar] [CrossRef]
  40. Li, J.; Zhu, J.; Zhang, Q.; Chen, L.; Ma, S.; Lu, Y.; Shen, B.; Zhang, R.; Zhang, M.; He, Y.; et al. NPPA Promoter Hypomethylation Predicts Central Obesity Development: A Prospective Longitudinal Study in Chinese Adults. Obes. Facts 2022, 15, 257–270. [Google Scholar] [CrossRef]
  41. Abuzhalihan, J.; Wang, Y.-T.; Ma, Y.-T.; Fu, Z.-Y.; Yang, Y.-N.; Ma, X.; Li, X.-M.; Liu, F.; Chen, B.-D. SOAT1 methylation is associated with coronary heart disease. Lipids Health Dis. 2019, 18, 192. [Google Scholar] [CrossRef]
  42. Rong, J.; Xu, X.; Xiang, Y.; Yang, G.; Ming, X.; He, S.; Liang, B.; Zhang, X.; Zheng, F. Thioredoxin-interacting protein promotes activation and inflammation of monocytes with DNA demethylation in coronary artery disease. J. Cell. Mol. Med. 2020, 24, 3560–3571. [Google Scholar] [CrossRef]
  43. Sharma, P.; Garg, G.; Kumar, A.; Mohammad, F.; Kumar, S.R.; Tanwar, V.S.; Sati, S.; Sharma, A.; Karthikeyan, G.; Brahmachari, V.; et al. Genome wide DNA methylation profiling for epigenetic alteration in coronary artery disease patients. Gene 2014, 541, 31–40. [Google Scholar] [CrossRef]
  44. Rask-Andersen, M.; Martinsson, D.; Ahsan, M.; Enroth, S.; Ek, W.E.; Gyllensten, U.; Johansson, Å. Epigenome-wide association study reveals differential DNA methylation in individuals with a history of myocardial infarction. Hum. Mol. Genet. 2016, 25, 4739–4748. [Google Scholar] [CrossRef] [PubMed]
  45. Armstrong, N.M.D.; Chen, W.-M.; Brewer, M.S.; Williams, S.R.; Sale, M.M.; Worrall, B.B.; Keene, K.L. Epigenome-Wide Analyses Identify Two Novel Associations with Recurrent Stroke in the Vitamin Intervention for Stroke Prevention Clinical Trial. Front. Genet. 2018, 9, 358. [Google Scholar] [CrossRef]
  46. Fernández-Sanlés, A.; Sayols-Baixeras, S.; Curcio, S.; Subirana, I.; Marrugat, J.; Elosua, R. DNA methylation and age-Independent cardiovascular risk, an epigenome-Wide approach the REGICOR study (REgistre GIroní del COR). Arterioscler. Thromb. Vasc. Biol. 2018, 38, 645–652. [Google Scholar] [CrossRef]
  47. Bain, C.R.; Ziemann, M.; Kaspi, A.; Khan, A.W.; Taylor, R.; Trahair, H.; Khurana, I.; Kaipananickal, H.; Wallace, S.; El-Osta, A.; et al. DNA methylation patterns from peripheral blood separate coronary artery disease patients with and without heart failure. ESC Heart Fail. 2020, 7, 2468–2478. [Google Scholar] [CrossRef] [PubMed]
  48. Gao, Y.; Pang, H.; Chen, B.; Wu, C.; Wang, Y.; Hou, L.; Wang, S.; Sun, D.; Zheng, X. Genome-wide analysis of DNA methylation and risk of cardiovascular disease in a Chinese population. BMC Cardiovasc. Disord. 2021, 21, 240. [Google Scholar] [CrossRef]
  49. Ammous, F.; Zhao, W.; Lin, L.; Ratliff, S.M.; Mosley, T.H.; Bielak, L.F.; Zhou, X.; Peyser, P.A.; Kardia, S.L.R.; Smith, J.A. Epigenetics of single-site and multi-site atherosclerosis in African Americans from the Genetic Epidemiology Network of Arteriopathy (GENOA). Clin. Epigenetics 2022, 14, 10. [Google Scholar] [CrossRef] [PubMed]
  50. Fernández-Sanlés, A.; Sayols-Baixeras, S.; Subirana, I.; Sentí, M.; Pérez-Fernández, S.; Moura, M.d.C.; Esteller, M.; Marrugat, J.; Elosua, R. DNA methylation biomarkers of myocardial infarction and cardiovascular disease. Clin. Epigenet. 2021, 13, 86. [Google Scholar] [CrossRef]
  51. Hedman, Å.K.; Mendelson, M.M.; Marioni, R.E.; Gustafsson, S.; Joehanes, R.; Irvin, M.R.; Zhi, D.; Sandling, J.K.; Yao, C.; Liu, C.; et al. Epigenetic Patterns in Blood Associated with Lipid Traits Predict Incident Coronary Heart Disease Events and Are Enriched for Results from Genome-Wide Association Studies. Circ. Cardiovasc. Genet. 2017, 10, e001487. [Google Scholar] [CrossRef]
  52. Mendelson, M.M.; Marioni, R.E.; Joehanes, R.; Liu, C.; Hedman, Å.K.; Aslibekyan, S.; Demerath, E.W.; Guan, W.; Zhi, D.; Yao, C.; et al. Association of Body Mass Index with DNA Methylation and Gene Expression in Blood Cells and Relations to Cardiometabolic Disease: A Mendelian Randomization Approach. PLoS Med. 2017, 14, e1002215. [Google Scholar] [CrossRef]
  53. Allum, F.; Hedman, Å.K.; Shao, X.; Cheung, W.A.; Vijay, J.; Guénard, F.; Kwan, T.; Simon, M.-M.; Ge, B.; Moura, C.; et al. Dissecting features of epigenetic variants underlying cardiometabolic risk using full-resolution epigenome profiling in regulatory elements. Nat. Commun. 2019, 10, 1209. [Google Scholar] [CrossRef] [PubMed]
  54. Chilunga, F.P.; Henneman, P.; Venema, A.; Meeks, K.A.C.; Requena-Méndez, A.; Beune, E.; Mockenhaupt, F.P.; Smeeth, L.; Bahendeka, S.; Danquah, I.; et al. Genome-wide DNA methylation analysis on C-reactive protein among Ghanaians suggests molecular links to the emerging risk of cardiovascular diseases. npj Genom. Med. 2021, 6, 46. [Google Scholar] [CrossRef] [PubMed]
  55. Antoun, E.; Issarapu, P.; di Gravio, C.; Shrestha, S.; Betts, M.; Saffari, A.; Sahariah, S.A.; Sankareswaran, A.; Arumalla, M.; Prentice, A.M.; et al. DNA methylation signatures associated with cardiometabolic risk factors in children from India and The Gambia: Results from the EMPHASIS study. Clin. Epigenet. 2022, 14, 6. [Google Scholar] [CrossRef]
  56. Agha, G.; Mendelson, M.M.; Ward-Caviness, C.K.; Joehanes, R.; Huan, T.; Gondalia, R.; Salfati, E.; Brody, J.A.; Fiorito, G.; Bressler, J.; et al. Blood Leukocyte DNA Methylation Predicts Risk of Future Myocardial Infarction and Coronary Heart Disease. Circulation 2019, 140, 645–657. [Google Scholar] [CrossRef] [PubMed]
  57. Li, B.; Feng, Z.-H.; Sun, H.; Zhao, Z.-H.; Yang, S.-B.; Yang, P. The blood genome-wide DNA methylation analysis reveals novel epigenetic changes in human heart failure. Eur. Rev. Med. Pharmacol. Sci. 2017, 21, 1828–1836. [Google Scholar] [CrossRef]
  58. Huan, T.; Joehanes, R.; Song, C.; Peng, F.; Guo, Y.; Mendelson, M.; Yao, C.; Liu, C.; Ma, J.; Richard, M.; et al. Genome-wide identification of DNA methylation QTLs in whole blood highlights pathways for cardiovascular disease. Nat. Commun. 2019, 10, 4267. [Google Scholar] [CrossRef]
  59. Qin, X.; Karlsson, I.K.; Wang, Y.; Li, X.; Pedersen, N.; Reynolds, C.A.; Hägg, S. The epigenetic etiology of cardiovascular disease in a longitudinal Swedish twin study. Clin. Epigenet. 2021, 13, 129. [Google Scholar] [CrossRef]
  60. Koseler, A.; Ma, F.; Kilic, I.D.; Morselli, M.; Kilic, O.; Pellegrini, M. Genome-wide DNA methylation profiling of blood from monozygotic twins discordant for myocardial infarction. In Vivo 2020, 34, 361–367. [Google Scholar] [CrossRef]
  61. Nakatochi, M.; Ichihara, S.; Yamamoto, K.; Naruse, K.; Yokota, S.; Asano, H.; Matsubara, T.; Yokota, M. Epigenome-wide association of myocardial infarction with DNA methylation sites at loci related to cardiovascular disease. Clin. Epigenet. 2017, 9, 54. [Google Scholar] [CrossRef]
  62. Ward-Caviness, C.K.; Agha, G.; Chen, B.H.; Pfeiffer, L.; Wilson, R.; Wolf, P.; Gieger, C.; Schwartz, J.; Vokonas, P.S.; Hou, L.; et al. Analysis of repeated leukocyte DNA methylation assessments reveals persistent epigenetic alterations after an incident myocardial infarction. Clin. Epigenet. 2018, 10, 161. [Google Scholar] [CrossRef]
  63. Suzuki, M.M.; Bird, A. DNA methylation landscapes: Provocative insights from epigenomics. Nat. Rev. Genet. 2008, 9, 465–476. [Google Scholar] [CrossRef]
  64. Willmer, T.; Johnson, R.; Louw, J.; Pheiffer, C. Blood-Based DNA Methylation Biomarkers for Type 2 Diabetes: Potential for Clinical Applications. Front. Endocrinol. 2018, 9, 744. [Google Scholar] [CrossRef] [PubMed]
  65. Ganguly, P.; Alam, S.F. Role of homocysteine in the development of cardiovascular disease. Nutr. J. 2015, 14, 6. [Google Scholar] [CrossRef]
  66. Cordaux, R.; Batzer, M.A. The impact of retrotransposons on human genome evolution. Nat. Rev. Genet. 2009, 10, 691–703. [Google Scholar] [CrossRef] [PubMed]
  67. Cash, H.L.; McGarvey, S.T.; Houseman, E.A.; Marsit, C.J.; Hawley, N.L.; Lambert-Messerlian, G.M.; Viali, S.; Tuitele, J.; Kelsey, K.T. Cardiovascular disease risk factors and DNA methylation at the LINE-1 repeat region in peripheral blood from Samoan Islanders. Epigenetics 2011, 6, 1257–1264. [Google Scholar] [CrossRef] [PubMed]
  68. Thongsroy, J.; Mutirangura, A. The association between Alu hypomethylation and the severity of hypertension. PLoS ONE 2022, 17, e0270004. [Google Scholar] [CrossRef]
  69. Chen, R.; Meng, X.; Zhao, A.; Wang, C.; Yang, C.; Li, H.; Cai, J.; Zhao, Z.; Kan, H. DNA hypomethylation and its mediation in the effects of fine particulate air pollution on cardiovascular biomarkers: A randomized crossover trial. Environ. Int. 2016, 94, 614–619. [Google Scholar] [CrossRef]
  70. Bellavia, A.; Urch, B.; Speck, M.; Brook, R.D.; Scott, J.A.; Albetti, B.; Behbod, B.; North, M.; Valeri, L.; Bertazzi, P.A.; et al. DNA hypomethylation, ambient particulate matter, and increased blood pressure: Findings from controlled human exposure experiments. J. Am. Heart Assoc. 2013, 2, e000212. [Google Scholar] [CrossRef]
  71. Tsaprouni, L.G.; Yang, T.-P.; Bell, J.; Dick, K.J.; Kanoni, S.; Nisbet, J.; Viñuela, A.; Grundberg, E.; Nelson, C.P.; Meduri, E.; et al. Cigarette smoking reduces DNA methylation levels at multiple genomic loci but the effect is partially reversible upon cessation. Epigenetics 2014, 9, 1382–1396. [Google Scholar] [CrossRef]
  72. Zakhari, S. Alcohol Metabolism and Epigenetics Changes. Alcohol Res. Curr. Rev. 2013, 35, 6–16. [Google Scholar]
  73. Bock, C.; Halbritter, F.; Carmona, F.J.; Tierling, S.; Datlinger, P.; Assenov, Y.; Berdasco, M.; Bergmann, A.K.; Booher, K.; Busato, F.; et al. Quantitative comparison of DNA methylation assays for biomarker development and clinical applications. Nat. Biotechnol. 2016, 34, 726–737. [Google Scholar]
  74. Hartman, J.; Frishman, W.H. Inflammation and atherosclerosis: A review of the role of interleukin-6 in the development of atherosclerosis and the potential for targeted drug therapy. Cardiol. Rev. 2014, 22, 147–151. [Google Scholar] [CrossRef] [PubMed]
  75. Hashmat, S.; Rudemiller, N.; Lund, H.; Abais-Battad, J.M.; Van Why, S.; Mattson, D.L. Interleukin-6 inhibition attenuates hypertension and associated renal damage in Dahl salt-sensitive rats. Am. J. Physiol. Renal. Physiol. 2016, 311, F555–F561. [Google Scholar] [CrossRef] [PubMed]
  76. Feng, Y.; Ye, D.; Wang, Z.; Pan, H.; Lu, X.; Wang, M.; Xu, Y.; Yu, J.; Zhang, J.; Zhao, M.; et al. The Role of Interleukin-6 Family Members in Cardiovascular Diseases. Front. Cardiovasc. Med. 2022, 9, 818890. [Google Scholar] [CrossRef] [PubMed]
  77. Thompson, A.M.; Zanobetti, A.; Silverman, F.; Schwartz, J.; Coull, B.; Urch, B.; Speck, M.; Brook, J.R.; Manno, M.; Gold, D.R. Baseline repeated measures from controlled human exposure studies: Associations between ambient air pollution exposure and the systemic inflammatory biomarkers IL-6 and fibrinogen. Environ. Health Perspect. 2010, 118, 120–124. [Google Scholar] [CrossRef]
  78. Frantz, S.; Ertl, G.; Bauersachs, J. Mechanisms of disease: Toll-like receptors in cardiovascular disease. Nat. Clin. Pract. Cardiovasc. Med. 2007, 4, 444–454. [Google Scholar] [CrossRef]
  79. Liu, V.W.; Huang, P.L. Cardiovascular roles of nitric oxide: A review of insights from nitric oxide synthase gene disrupted mice. Cardiovasc. Res. 2008, 77, 19–29. [Google Scholar] [CrossRef]
  80. McLaren, J.E.; Ramji, D.P. Interferon gamma: A master regulator of atherosclerosis. Cytokine Growth Factor Rev. 2009, 20, 125–135. [Google Scholar] [CrossRef]
  81. Kleemann, R.; Zadelaar, S.; Kooistra, T. Cytokines and atherosclerosis: A comprehensive review of studies in mice. Cardiovasc. Res. 2008, 79, 360–376. [Google Scholar] [CrossRef]
  82. Roshan, M.H.K.; Tambo, A.; Pace, N.P. The Role of TLR2, TLR4, and TLR9 in the Pathogenesis of Atherosclerosis. Int. J. Inflamm. 2016, 2016, 1532832. [Google Scholar] [CrossRef]
  83. Voloshyna, I.; Littlefield, M.J.; Reiss, A.B. Atherosclerosis and interferon-γ: New insights and therapeutic targets. Trends Cardiovasc. Med. 2014, 24, 45–51. [Google Scholar] [CrossRef] [PubMed]
  84. Hayashi, T.; Matsui-Hirai, H.; Fukatsu, A.; Sumi, D.; Kano-Hayashi, H.; P, J.A.R.; Iguchi, A. Selective iNOS inhibitor, ONO1714 successfully retards the development of high-cholesterol diet induced atherosclerosis by novel mechanism. Atherosclerosis 2006, 187, 316–324. [Google Scholar] [CrossRef]
  85. Chong, C.-R.; Chan, W.P.A.; Nguyen, T.H.; Liu, S.; Procter, N.E.K.; Ngo, D.T.; Sverdlov, A.L.; Chirkov, Y.Y.; Horowitz, J.D. Thioredoxin-interacting protein: Pathophysiology and emerging pharmacotherapeutics in cardiovascular disease and diabetes. Cardiovasc. Drugs Ther. 2014, 28, 347–360. [Google Scholar] [CrossRef] [PubMed]
  86. Soriano-Tárraga, C.; Jiménez-Conde, J.; Giralt-Steinhauer, E.; Mola-Caminal, M.; Vivanco-Hidalgo, R.M.; Ois, A.; Rodríguez-Campello, A.; Cuadrado-Godia, E.; Sayols-Baixeras, S.; Elosua, R.; et al. Epigenome-wide association study identifies TXNIP gene associated with type 2 diabetes mellitus and sustained hyperglycemia. Hum. Mol. Genet. 2015, 25, 609–619. [Google Scholar] [CrossRef]
  87. Kulkarni, H.; Kos, M.Z.; Neary, J.; Dyer, T.D.; Kent, J.W.; Göring, H.H.; Cole, S.A.; Comuzzie, A.G.; Almasy, L.; Mahaney, M.C.; et al. Novel epigenetic determinants of type 2 diabetes in Mexican-American families. Hum. Mol. Genet. 2015, 24, 5330–5344. [Google Scholar] [CrossRef] [PubMed]
  88. Chambers, J.C.; Loh, M.; Lehne, B.; Drong, A.; Kriebel, J.; Motta, V.; Wahl, S.; Elliott, H.R.; Rota, F.; Scott, W.R.; et al. Epigenome-wide association of DNA methylation markers in peripheral blood from Indian Asians and Europeans with incident type 2 diabetes: A nested case-control study. Lancet Diabetes Endocrinol. 2015, 3, 526–534. [Google Scholar] [CrossRef]
  89. Florath, I.; Butterbach, K.; Heiss, J.; Bewerunge-Hudler, M.; Zhang, Y.; Schöttker, B.; Brenner, H. Type 2 diabetes and leucocyte DNA methylation: An epigenome-wide association study in over 1500 older adults. Diabetologia 2016, 59, 130–138. [Google Scholar] [CrossRef]
  90. Iung, B.; Vahanian, A. Epidemiology of valvular heart disease in the adult. Nat. Rev. Cardiol. 2011, 8, 162–172. [Google Scholar] [CrossRef] [PubMed]
  91. Evans, J.D.W.; Girerd, B.; Montani, D.; Wang, X.-J.; Galiè, N.; Austin, E.D.; Elliott, G.; Asano, K.; Grünig, E.; Yan, Y.; et al. BMPR2 mutations and survival in pulmonary arterial hypertension: An individual participant data meta-analysis. Lancet Respir. Med. 2016, 4, 129–137. [Google Scholar] [CrossRef]
  92. Guignabert, C.; Bailly, S.; Humbert, M. Restoring BMPRII functions in pulmonary arterial hypertension: Opportunities, challenges and limitations. Expert Opin. Ther. Targets 2017, 21, 181–190. [Google Scholar] [CrossRef]
  93. Hautefort, A.; Mendes-Ferreira, P.; Sabourin, J.; Manaud, G.; Bertero, T.; Rucker-Martin, C.; Riou, M.; Adão, R.; Manoury, B.; Lambert, M.; et al. Bmpr2 mutant rats develop pulmonary and cardiac characteristics of pulmonary arterial hypertension. Circulation 2019, 139, 932–948. [Google Scholar] [CrossRef]
  94. Ali, A.J.; Makings, J.; Ley, K. Regulatory T cell stability and plasticity in atherosclerosis. Cells 2020, 9, 2665. [Google Scholar] [CrossRef] [PubMed]
  95. Joly, A.-L.; Seitz, C.; Liu, S.; Kuznetsov, N.V.; Gertow, K.; Westerberg, L.S.; Paulsson-Berne, G.; Hansson, G.K.; Andersson, J. Alternative splicing of FOXP3 controls regulatory T cell effector functions and is associated with human atherosclerotic plaque stability. Circ. Res. 2018, 122, 1385–1394. [Google Scholar] [CrossRef] [PubMed]
  96. Mahmoud, M.M.; Sanad, E.F.; Hamdy, N.M. MicroRNAs Role in Non-Communicable Diseases and Link to Multidrug Resistance, Regulation or Alteration. Preprints 2021. [Google Scholar] [CrossRef]
  97. Chen, D.; Li, Y.; Yu, Z.; Li, Y.; Su, Z.; NI, L.; Yang, S.; Gui, Y.; Lai, Y. Downregulated microRNA-510-5p acts as a tumor suppressor in renal cell carcinoma. Mol. Med. Rep. 2015, 12, 3061–3066. [Google Scholar] [CrossRef]
  98. Gaj, P.; Zagozdzon, R. In silicoanalysis of microRNA-510 as a potential oncomir in human breast cancer. Breast Cancer Res. 2014, 16, 403. [Google Scholar] [CrossRef] [PubMed]
  99. Guo, Q.J.; Mills, J.N.; Bandurraga, S.G.; Nogueira, L.M.; Mason, N.J.; Camp, E.R.; Larue, A.C.; Turner, D.P.; Findlay, V.J. MicroRNA-510 promotes cell and tumor growth by targeting peroxiredoxin1 in breast cancer. Breast Cancer Res. 2013, 15, R70. [Google Scholar] [CrossRef]
  100. Zhang, M.-W.; Shen, Y.-J.; Shi, J.; Yu, J.-G. MiR-223-3p in cardiovascular diseases: A biomarker and potential therapeutic target. Front. Cardiovasc. Med. 2021, 7, 610561. [Google Scholar] [CrossRef]
  101. Sookoian, S.; Pirola, C.J. Metabolic syndrome: From the genetics to the pathophysiology. Curr. Hypertens. Rep. 2011, 13, 149–157. [Google Scholar] [CrossRef]
  102. Blüher, S.; Käpplinger, J.; Herget, S.; Reichardt, S.; Böttcher, Y.; Grimm, A.; Kratzsch, J.; Petroff, D. Cardiometabolic risk markers, adipocyte fatty acid binding protein (aFABP) and the impact of high-intensity interval training (HIIT) in obese adolescents. Metab. Clin. Exp. 2017, 68, 77–87. [Google Scholar] [CrossRef]
  103. Zhuang, L.; Mao, Y.; Liu, Z.; Li, C.; Jin, Q.; Lu, L.; Tao, R.; Yan, X.; Chen, K. FABP3 deficiency exacerbates metabolic derangement in cardiac hypertrophy and heart failure via PPARα pathway. Front. Cardiovasc. Med. 2021, 8, 722908. [Google Scholar] [CrossRef]
  104. Qian, Q.; Kuo, L.; Yu, Y.-T.; Rottman, J.N. A concise promoter region of the heart fatty acid–binding protein gene dictates tissue-appropriate expression. Circ. Res. 1999, 84, 276–289. [Google Scholar] [CrossRef]
  105. Hai, Q.; Smith, J.D. Acyl-Coenzyme A: Cholesterol Acyltransferase (ACAT) in Cholesterol Metabolism: From Its Discovery to Clinical Trials and the Genomics Era. Metabolites 2021, 11, 543. [Google Scholar] [CrossRef]
  106. Xu, N.; Meng, H.; Liu, T.-Y.; Feng, Y.-L.; Qi, Y.; Zhang, D.-H.; Wang, H.-L. Sterol O-acyltransferase 1 deficiency improves defective insulin signaling in the brains of mice fed a high-fat diet. Biochem. Biophys. Res. Commun. 2018, 499, 105–111. [Google Scholar] [CrossRef] [PubMed]
  107. Yeh, J.-K.; Wang, C.-Y. Telomeres and telomerase in cardiovascular diseases. Genes 2016, 7, 58. [Google Scholar] [CrossRef] [PubMed]
  108. Kyo, S.; Takakura, M.; Fujiwara, T.; Inoue, M. Understanding and exploiting hTERT promoter regulation for diagnosis and treatment of human cancers. Cancer Sci. 2008, 99, 1528–1538. [Google Scholar] [CrossRef] [PubMed]
  109. Li-Tempel, T.; Larra, M.F.; Sandt, E.; Mériaux, S.B.; Schote, A.B.; Schächinger, H.; Muller, C.P.; Turner, J.D. The cardiovascular and hypothalamus-pituitary-adrenal axis response to stress is controlled by glucocorticoid receptor sequence variants and promoter methylation. Clin. Epigenetics 2016, 8, 12. [Google Scholar] [CrossRef]
  110. Willmer, T.; Goedecke, J.H.; Dias, S.; Louw, J.; Pheiffer, C. DNA methylation of FKBP5 in South African women: Associations with obesity and insulin resistance. Clin. Epigenet. 2020, 12, 141. [Google Scholar] [CrossRef]
  111. Volpe, M.; Rubattu, S.; Burnett, J.J. Natriuretic peptides in cardiovascular diseases: Current use and perspectives. Eur. Heart J. 2014, 35, 419–425. [Google Scholar] [CrossRef]
  112. Zois, N.E.; Bartels, E.D.; Hunter, I.; Kousholt, B.S.; Olsen, L.H.; Goetze, J.P. Natriuretic peptides in cardiometabolic regulation and disease. Nat. Rev. Cardiol. 2014, 11, 403–412. [Google Scholar] [CrossRef]
  113. Gouil, Q.; Keniry, A. Latest techniques to study DNA methylation. Essays Biochem. 2019, 63, 639–648. [Google Scholar] [PubMed]
  114. Lister, R.; Pelizzola, M.; Dowen, R.H.; Hawkins, R.D.; Hon, G.; Tonti-Filippini, J.; Nery, J.R.; Lee, L.; Ye, Z.; Ngo, Q.-M.; et al. Human DNA methylomes at base resolution show widespread epigenomic differences. Nature 2009, 462, 315–322. [Google Scholar] [CrossRef]
  115. Jin, Z.; Liu, Y. DNA methylation in human diseases. Genes Dis. 2018, 5, 1–8. [Google Scholar] [CrossRef]
  116. Brott, B.K.; Sokol, S.Y. A Vertebrate Homolog of the Cell Cycle Regulator Dbf4 Is an Inhibitor of Wnt Signaling Required for Heart Development. Dev. Cell 2005, 8, 703–715. [Google Scholar] [CrossRef] [PubMed]
  117. Kempf, T.; Eden, M.; Strelau, J.; Naguib, M.; Willenbockel, C.; Tongers, J.; Heineke, J.; Kotlarz, D.; Xu, J.; Molkentin, J.D.; et al. The transforming growth factor-β superfamily member growth-differentiation factor-15 protects the heart from ischemia/reperfusion injury. Circ. Res. 2006, 98, 351–360. [Google Scholar] [CrossRef] [PubMed]
  118. Khan, S.Q.; Ng, K.; Dhillon, O.; Kelly, D.; Quinn, P.; Squire, I.B.; Davies, J.E.; Ng, L.L. Growth differentiation factor-15 as a prognostic marker in patients with acute myocardial infarction. Eur. Heart J. 2009, 30, 1057–1065. [Google Scholar] [CrossRef]
  119. Wollert, K.C.; Kempf, T.; Peter, T.; Olofsson, S.; James, S.; Johnston, N.; Lindahl, B.; Horn-Wichmann, R.; Brabant, G.; Simoons, M.L.; et al. Prognostic value of growth-differentiation factor-15 in patients with non–ST-elevation acute coronary syndrome. Circulation 2007, 115, 962–971. [Google Scholar] [CrossRef]
  120. Wu, Z.; Chen, L.; Hong, X.; Si, J.; Cao, W.; Yu, C.; Huang, T.; Sun, D.; Liao, C.; Pang, Y.; et al. Temporal associations between leukocytes DNA methylation and blood lipids: A longitudinal study. Clin. Epigenet. 2022, 14, 132. [Google Scholar] [CrossRef]
  121. Jitrapakdee, S.; Wallace, J.C. Structure, function and regulation of pyruvate carboxylase. Biochem. J. 1999, 340 Pt 1, 1–16. [Google Scholar] [CrossRef]
  122. Singh, K.K.; Shukla, P.C.; Quan, A.; Al-Omran, M.; Lovren, F.; Pan, Y.; Brezden-Masley, C.; Ingram, A.J.; Stanford, W.L.; Teoh, H.; et al. BRCA1 is a novel target to improve endothelial dysfunction and retard atherosclerosis. J. Thorac. Cardiovasc. Surg. 2013, 146, 949–960.e4. [Google Scholar] [CrossRef]
  123. Liu, X.S.; Yang, J.W.; Zeng, J.; Chen, X.Q.; Gao, Y.; Kui, X.Y.; Liu, X.Y.; Zhang, Y.; Zhang, Y.H.; Pei, Z.J. SLC2A1 is a Diagnostic Biomarker Involved in Immune Infiltration of Colorectal Cancer and Associated With m6A Modification and ceRNA. Front. Cell Dev. Biol. 2022, 10, 853596. [Google Scholar] [CrossRef]
  124. Wang, J.; Xu, J.; Wang, Q.; Brainard, R.E.; Watson, L.J.; Jones, S.P.; Epstein, P.N. Reduced cardiac fructose 2,6 bisphosphate increases hypertrophy and decreases glycolysis following aortic constriction. PLoS ONE 2013, 8, e53951. [Google Scholar] [CrossRef]
  125. Iida, R.; Ueki, M.; Yasuda, T. Deficiency of M-LP/Mpv17L leads to development of β-cell hyperplasia and improved glucose tolerance via activation of the Wnt and TGF-β pathways. Biochim. Biophys. Acta Mol. Basis Dis. 2022, 1868, 166318. [Google Scholar] [CrossRef]
  126. Szabadosova, V.; Boronova, I.; Ferenc, P.; Tothova, I.; Bernasovska, J.; Zigova, M.; Kmec, J.; Bernasovsky, I. Analysis of selected genes associated with cardiomyopathy by next-generation sequencing. J. Clin. Lab. Anal. 2018, 32, e22254. [Google Scholar] [CrossRef]
  127. Cao, W.; Zhang, T.; Feng, R.; Xia, T.; Huang, H.; Liu, C.; Sun, C. Hoxa5 alleviates obesity-induced chronic inflammation by reducing ER stress and promoting M2 macrophage polarization in mouse adipose tissue. J. Cell. Mol. Med. 2019, 23, 7029–7042. [Google Scholar] [CrossRef] [PubMed]
  128. Allen, A.J.H.; Ghafoor, A.A.; Liu, Y.; Sandford, A.; Jen, R.; Daniele, P.; Taylor, C.; Peres, B.U.; Ayas, N.T. The rs579459 ABO gene polymorphism and risk of incident cardiovascular events in obstructive sleep apnea: A pilot study. Sleep Breath 2023, 27, 721–725. [Google Scholar] [CrossRef] [PubMed]
  129. Nandakumar, P.; Lee, D.; Hoffmann, T.J.; Ehret, G.B.; Arking, D.; Ranatunga, D.; Li, M.; Grove, M.L.; Boerwinkle, E.; Schaefer, C.; et al. Analysis of putative cis-regulatory elements regulating blood pressure variation. Hum. Mol. Genet. 2020, 29, 1922–1932. [Google Scholar] [CrossRef]
  130. Marston, N.A.; Giugliano, R.P.; Melloni, G.E.; Park, J.G.; Morrill, V.; Blazing, M.A.; Ference, B.; Stein, E.; Stroes, E.S.; Braunwald, E.; et al. Association of Apolipoprotein B-Containing Lipoproteins and Risk of Myocardial Infarction in Individuals with and Without Atherosclerosis: Distinguishing Between Particle Concentration, Type, and Content. JAMA Cardiol. 2022, 7, 250–256. [Google Scholar] [CrossRef] [PubMed]
  131. Talmud, P.J.; Drenos, F.; Shah, S.; Shah, T.; Palmen, J.; Verzilli, C.; Gaunt, T.R.; Pallas, J.; Lovering, R.; Li, K.; et al. Gene-centric association signals for lipids and apolipoproteins identified via the HumanCVD BeadChip. Am. J. Hum. Genet. 2009, 85, 628–642. [Google Scholar] [CrossRef]
  132. Turcot, V.; Lu, Y.; Highland, H.M.; Schurmann, C.; Justice, A.E.; Fine, R.S.; Bradfield, J.P.; Esko, T.; Giri, A.; Graff, M.; et al. Protein-altering variants associated with body mass index implicate pathways that control energy intake and expenditure in obesity. Nat. Genet. 2018, 50, 26–41. [Google Scholar] [CrossRef]
  133. Liang, X.; Ding, Y.; Zhang, Y.; Chai, Y.-H.; He, J.; Chiu, S.-M.; Gao, F.; Tse, H.-F.; Lian, Q. Activation of NRG1-ERBB4 signaling potentiates mesenchymal stem cell-mediated myocardial repairs following myocardial infarction. Cell Death Dis. 2015, 6, e1765. [Google Scholar] [CrossRef]
  134. Zhang, X.; He, D.; Xiang, Y.; Wang, C.; Liang, B.; Li, B.; Qi, D.; Deng, Q.; Yu, H.; Lu, Z.; et al. DYSF promotes monocyte activation in atherosclerotic cardiovascular disease as a DNA methylation-driven gene. Transl. Res. 2022, 247, 19–38. [Google Scholar] [CrossRef] [PubMed]
  135. Shen, G.-Q.; Li, L.; Girelli, D.; Seidelmann, S.B.; Rao, S.; Fan, C.; Park, J.E.; Xi, Q.; Li, J.; Hu, Y.; et al. An LRP8 variant is associated with familial and premature coronary artery disease and myocardial infarction. Am. J. Hum. Genet. 2007, 81, 780–791. [Google Scholar] [CrossRef]
  136. Rahm, A.-K.; Gramlich, D.; Wieder, T.; Müller, M.E.; Schoeffel, A.; A El Tahry, F.; Most, P.; Heimberger, T.; Sandke, S.; Weis, T.; et al. Trigger-Specific Remodeling of KCa2 Potassium Channels in Models of Atrial Fibrillation. Pharmacogenom. Pers. Med. 2021, 14, 579–590. [Google Scholar] [CrossRef]
  137. Zhong, K.; Kant, S.; Sellke, F.; Feng, J. SK channels and heart disease. In Ion Transporters—From Basic Properties to Medical Treatment; IntechOpen: London, UK, 2022. [Google Scholar]
  138. Ma, B.; Wilker, E.H.; Willis-Owen, S.A.G.; Byun, H.-M.; Wong, K.C.C.; Motta, V.; Baccarelli, A.A.; Schwartz, J.; Cookson, W.O.C.M.; Khabbaz, K.; et al. Predicting DNA methylation level across human tissues. Nucleic Acids Res. 2014, 42, 3515–3528. [Google Scholar] [CrossRef] [PubMed]
  139. Houseman, E.A.; Accomando, W.P.; Koestler, D.C.; Christensen, B.C.; Marsit, C.J.; Nelson, H.H.; Wiencke, J.K.; Kelsey, K.T. DNA methylation arrays as surrogate measures of cell mixture distribution. BMC Bioinform. 2012, 13, 86. [Google Scholar] [CrossRef] [PubMed]
  140. Fernández-Sanlés, A.; Sayols-Baixeras, S.; Subirana, I.; Degano, I.R.; Elosua, R. Association between DNA methylation and coronary heart disease or other atherosclerotic events: A systematic review. Atherosclerosis 2017, 263, 325–333. [Google Scholar] [CrossRef]
  141. Muka, T.; Koromani, F.; Portilla, E.; O’Connor, A.; Bramer, W.M.; Troup, J.; Chowdhury, R.; Dehghan, A.; Franco, O.H. The role of epigenetic modifications in cardiovascular disease: A systematic review. Int. J. Cardiol. 2016, 212, 174–183. [Google Scholar] [CrossRef]
  142. Braun, K.V.; Voortman, T.; Dhana, K.; Troup, J.; Bramer, W.M.; Troup, J.; Chowdhury, R.; Dehghan, A.; Muka, T.; Franco, O.H. The role of DNA methylation in dyslipidaemia: A systematic review. Prog. Lipid Res. 2016, 64, 178–191. [Google Scholar] [CrossRef]
  143. Anderson, D.R.; Poterucha, J.T.; Mikuls, T.R.; Duryee, M.J.; Garvin, R.P.; Klassen, L.W.; Shurmur, S.W.; Thiele, G.M. IL-6 and its receptors in coronary artery disease and acute myocardial infarction. Cytokine 2013, 62, 395–400. [Google Scholar] [CrossRef]
  144. Klingenberg, R.; Gerdes, N.; Badeau, R.M.; Gisterå, A.; Strodthoff, D.; Ketelhuth, D.F.; Lundberg, A.M.; Rudling, M.; Nilsson, S.K.; Olivecrona, G.; et al. Depletion of FOXP3+ regulatory T cells promotes hypercholesterolemia and atherosclerosis. J. Clin. Investig. 2013, 123, 1323–1334. [Google Scholar] [CrossRef] [PubMed]
  145. Michels, K.B.; Binder, A.M.; Dedeurwaerder, S.; Epstein, C.B.; Greally, J.M.; Gut, I.; Houseman, E.A.; Izzi, B.; Kelsey, K.T.; Meissner, A.; et al. Recommendations for the design and analysis of epigenome-wide association studies. Nat. Methods 2013, 10, 949–955. [Google Scholar] [CrossRef]
  146. McGarry, J.D.; Brown, N.F. The mitochondrial carnitine palmitoyltransferase system. From concept to molecular analysis. Eur. J. Biochem. 1997, 244, 1–14. [Google Scholar] [CrossRef] [PubMed]
  147. Keung, W.; Ussher, J.R.; Jaswal, J.S.; Raubenheimer, M.; Lam, V.H.; Wagg, C.S.; Lopaschuk, G.D. Inhibition of carnitine palmitoyltransferase-1 activity alleviates insulin resistance in diet-induced obese mice. Diabetes 2013, 62, 711–720. [Google Scholar] [CrossRef]
  148. Rohlenova, K.; Veys, K.; Miranda-Santos, I.; De Bock, K.; Carmeliet, P. Endothelial cell metabolism in health and disease. Trends Cell Biol. 2018, 28, 224–236. [Google Scholar] [CrossRef]
  149. Xia, Y.; Brewer, A.; Bell, J.T. DNA methylation signatures of incident coronary heart disease: Findings from epigenome-wide association studies. Clin. Epigenet. 2021, 13, 186. [Google Scholar] [CrossRef]
  150. Grieshober, L.; Graw, S.; Barnett, M.J.; Thornquist, M.D.; Goodman, G.E.; Chen, C.; Koestler, D.C.; Marsit, C.J.; Doherty, J.A. AHRR methylation in heavy smokers: Associations with smoking, lung cancer risk, and lung cancer mortality. BMC Cancer 2020, 20, 905. [Google Scholar] [CrossRef] [PubMed]
  151. Philibert, R.A.; Dogan, M.V.; Mills, J.A.; Long, J.D. AHRR Methylation is a Significant Predictor of Mortality Risk in Framingham Heart Study. J. Insur. Med. 2019, 48, 79–89. [Google Scholar] [CrossRef]
  152. Bacalini, M.G.; Boattini, A.; Gentilini, D.; Giampieri, E.; Pirazzini, C.; Giuliani, C.; Fontanesi, E.; Remondini, D.; Capri, M.; Del Rio, A.; et al. A meta-analysis on age-associated changes in blood DNA methylation: Results from an original analysis pipeline for Infinium 450k data. Aging 2015, 7, 97–109. [Google Scholar] [CrossRef]
  153. Chamberlain, J.D.; Nusslé, S.; Chapatte, L.; Kinnaer, C.; Petrovic, D.; Pradervand, S.; Bochud, M.; Harris, S.E.; Corley, J.; Cox, S.R.; et al. Blood DNA methylation signatures of lifestyle exposures: Tobacco and alcohol consumption. Clin. Epigenet. 2022, 14, 155. [Google Scholar] [CrossRef]
  154. Wang, Y.; Hannon, E.; Grant, O.A.; Gorrie-Stone, T.J.; Kumari, M.; Mill, J.; Zhai, X.; McDonald-Maier, K.D.; Schalkwyk, L.C. DNA methylation-based sex classifier to predict sex and identify sex chromosome aneuploidy. BMC Genom. 2021, 22, 484. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Flow diagram depicting the selection of research articles for inclusion in the review. Adapted from [18].
Figure 1. Flow diagram depicting the selection of research articles for inclusion in the review. Adapted from [18].
Ijms 26 02355 g001
Figure 2. DNA methylation alterations associated with cardiovascular disease.
Figure 2. DNA methylation alterations associated with cardiovascular disease.
Ijms 26 02355 g002
Table 1. Main findings from CVD studies using global DNA methylation analysis in peripheral blood.
Table 1. Main findings from CVD studies using global DNA methylation analysis in peripheral blood.
AuthorPopulationTissue TypeMethodStudy Outcome
[19]European; 7 male atherosclerotic vascular patients, 7 healthy male controlsWBCs5mCAtherosclerosis patients had significantly higher plasma homocysteine concentrations and significantly lower genomic DNA methylation.
[20] Indian; 137 coronary artery disease patients, 150 controlsPBLsHpaII/MspI ratioGlobal hypermethylation was associated with coronary artery disease.
[21] American; 712 elderly menWBLINE-1Blood LINE-1 hypomethylation was associated with ischemic heart disease.
[10] Singapore Chinese; 129 men, 157 womenPBLsALU and Satellite 2 repetitive element methylationPositive correlation in global DNA methylation in men with a history of CVD or its predisposing conditions but not in women.
[22] Chinese; 280 ischemic stroke patients and 280 healthy controls; 40 and 80 yearsWBLINE-1Global hypomethylation was associated with higher risk for ischemic stroke in men but not in women.
[23] Chinese; 334 CHD patients and 788 healthy controlsPBLsLINE-1LINE-1 hypomethylation was significantly associated with age, elevated total cholesterol, diagnosis of diabetes, and risk of CHD.
[24] Italian European; 292 myocardial infarction patients and 292 matched controlsWBCsLINE-1Global hypomethylation was associated with myocardial infarction risk in men.
[25] Brazilian; 90 postmenopausal women between 45 and 65 yearsPBLs5mCHypomethylation was associated with higher CVD risk.
[26] German; 8 atherosclerosis patients and 8 healthy controls; 35–60 yearsWB5mC/5hmCNo statistically significant changes in global DNA methylation in atherosclerosis patients compared to healthy individuals.
[27] American; longitudinal study of 789 elderly subjects followed over 10 yearsWBLINE-1 and AluPositive association identified between Alu hypermethylation and diastolic blood pressure.
[28] Chinese; 91 atherosclerosis patients and 22 controlsPBMCs5mC/5hmC5mC and 5hmC levels were significantly correlated with Gensini (severity of atherosclerosis).
[29] Chinese; 44 patients with cardiovascular disease and 42 controlsPBMCs5mC/5hmC5mC and 5hmC levels were higher in CVD patients than in control subjects.
Table 2. Main findings from CVD studies investigating candidate gene methylation in peripheral blood.
Table 2. Main findings from CVD studies investigating candidate gene methylation in peripheral blood.
Author (Year)Genes InvestigatedCountrySample Size SexTissue TypeMethodStudy Outcome
[27] TLR2, IFN-γ, and iNOSUnited StatesLongitudinal study of 789 subjects M WBBisulphite pyrosequencingMethylation of TLR2 and iNOS positively correlated with blood pressure, whilst IFN-γ promoter methylation negatively associated with blood pressure.
[30] BMPR2ChinaVHD = 82
Controls = 57
M and FPBBisulphite pyrosequencingThe BMPR2 promoter was hypermethylated in the VHD patients compared to healthy controls and correlated with VHD severity.
[31] FOXP3China171 patients with CHDMPBMethylation specific PCRIncreased DNA methylation of FOXP3 associated with severity of atherosclerosis.
[32] FOXP3ChinaCHD = 188
Controls = 68
MWBBisulfite sequencingIncreased DNA methylation of FOXP3 associated with severity of CHD.
[33] microRNA-510IndiaHypertension = 20
Controls = 20
M and FWBMethylation specific PCRThe miR-510 promoter was hypermethylated in the hypertensive patients versus healthy controls.
[34] FABP3 (17 CpGs)Europe517 individuals with varying parameters of metabolic syndromeM and FPBMCsEpiTYPER MassArrayAverage methylation of 17 CpG sites within the FABP3 positively associated with total cholesterol, whilst individual CpG methylation associated with diastolic blood pressure, insulin resistance, and LDL-cholesterol.
[35] hTERT (5 CpGs)ChinaAtherosclerosis = 197
Controls = 165
M and FPBLsMethylation-specific PCRhTERT was hypomethylated in atherosclerosis patients compared to healthy individuals and associated with a concomitant decrease in hTERT mRNA levels.
[36] IL-6 (2 CpGs)ChinaCHD = 212
Controls = 218
M and FPBLsBisulfite pyrosequencingHypomethylation at two promoter CpGs in CHD patients compared to controls. Combined methylation scores of both CpGs correlated with CHD risk.
[37] miRNA-233 (9 CpGs)ChinaAtherosclerosis = 23
Controls = 32
M and FPBLsBisulfite sequencingHypomethylation of seven CpG sites were detected in the miRNA-233 promoter in participants with atherosclerotic cerebral infarction compared to healthy controls. DNA methylation of these CpG sites correlated with total plasma cholesterol levels.
[38] NR3C1 (22 CpGs)United States84 MZ twin pairsMPBLsBisulfite pyrosequencingAltered DNA methylation in exon 1F of the NR3C1 promoter positively associated with markers of early subclinical atherosclerosis.
[39]IL-6 (6 CpGs)IranAtherosclerosis = 35
Controls = 30
M and FPBBisulfite sequencingSix evolutionary conserved CpG sites in the distal IL-6 promoter were hypomethylated in the atherosclerosis patients compared to the controls. DNA methylation of these CpG sites inversely correlated with IL-6 mRNA levels.
[40]NPPA (9 CpGs)ChinaLongitudinal study of 2498 subjectsM and FPBBisulfite sequencingHypermethylation of a single CpG site situated −459 bp from the NPPA TSS associated with reduced risk of developing CVDs.
[41] SOAT1 (26 CpGs)ChinaCHD = 99
Controls = 89
M and FWBBisulfite sequencingHypomethylation of seven CpG sites associated with incidence of CHD.
[42] TXNIP (1 CpG)ChinaCAD = 54
Controls = 54
M and FPBLsPyrosequencingDNA methylation of TXNIP at a single CpG site, located at cg19693031, was significantly decreased in cases compared to controls. DNA methylation inversely correlated with TXNIP mRNA expression, fasting plasma glucose, and total cholesterol.
Table 3. Main findings from CVD studies investigating whole-genome methylation in blood.
Table 3. Main findings from CVD studies investigating whole-genome methylation in blood.
Author (Year)CountrySample Size SexTissue TypeMethodStudy Outcome
[43] IndiaDiscovery cohort
CHD = 18
Controls = 18
Validation cohort
CHD = 48
Controls = 48
M PBMCsBisulphite sequencing via 454 platformHypermethylation was identified within 19 regions in CHD patients versus controls. Twelve differentially methylated regions were assessed in the validation cohort, which identified the hypermethylation of 6 CpG sites in the CHD group compared to controls. Three of these sites mapped to the intronic region of STE20 related adaptor alpha (STRADA), one was situated in the intronic region of heat shock protein 90 beta family member 3 pseudogene (HSP90B3P), and two were situated within exon 1 of Complement C1q like 4 (C1QL4).
[24] Italian and DutchDiscovery cohort
MI = 292 Controls = 292
Validation cohort
MI = 317 Controls = 262
M and FWBCsDiscovery: Illumina Infinium HumanMethylation450 BeadChip
Validation: MALDI-TOF mass spectrometry
Results from the discovery cohort revealed the hypomethylation of 15 CpG sites within exon 1 of the zinc finger and BTB domain-containing protein 12 (ZBTB12) gene, which correlated with MI, and was replicated in the validation cohort.
[44] SwedishCVD = 238 Controls = 491MPBIllumina Infinium HumanMethylation450 BeadchipAltered DNA methylation was observed at 211 CpG sites, mapping to 196 genes, which collectively associated with a history of MI. Forty-two of the differentially methylated regions mapped to genes previously associated with MI in the literature. Some of these genes identified included Dysferlin (DYSF) and Neuregulin 1 (NRG1). The link between MI and methylation was independent of BMI.
[45] European and AfricanStroke patients = 180 M and FWBIllumina Infinium HumanMethylation450 BeadChipHypermethylation at two CpG sites located in Ankyrin Repeat And SOCS Box Containing 10 (cg22812874, ASB10) and Tetratricopeptide repeat domain 37 (cg00340919, TTC37) associated with plasma homocysteine levels in individuals with recurrent stroke. This association was found in Europeans but not in Africans.
[46] American and EuropeanDiscovery REGICOR cohort
n = 645
Validation
FOS cohort
n = 2542
M and FWBIllumina Infinium HumanMethylation450 BeadChipEight CpG sites annotating to four genes [Aryl hydrocarbon receptor repressor (AHRR), Carnitine Palmitoyltransferase 1A (CPT1A), Peptidylprolyl Isomerase F (PPIF), and Strawberry Notch Homolog 2 (SBNO2)], and three intergenic regions were differentially methylated and associated with cardiovascular risk, independent of age.
[47] EuropeanT2D + HF = 10
Control = 10
MWBDiscovery: Methyl-binding domain-capture sequencing
Validation: Bisulfite sequencing
Differential methylation was identified in both gene body and enhancer elements, with reduced methylation signatures observed in coding regions of [Histone deacetylase 9 (HDAC9), MicroRNA 3675 (MIR3675), Jumonji- and AT-rich interaction domain (ARID)-domain-containing protein (JARID2), and Gremlin 1 (GREM1)], and with increased methylation in Decaprenyl Diphosphate Synthase Subunit 2 (PDSS2), linked to the extreme phenotypes of heart failure.
[48] Chinese CVD = 83 Controls = 83M and FWBIllumina Infinium HumanMethylation850 EPIC BeadChip Eight CVD-related CpGs sites were identified, with the combined use of all eight showing an excellent predictive power for CVD development. The 8 CpGs included: 1 hypermethylated CpG of cytochrome P450, family 8, subfamily B, and polypeptide 1 (cg07655795, CYP8B1), being associated with increased CVD risk, whilst the remaining seven hypomethylated CpGs of Spondin 1 (cg06901278, cg11651314, and SPON1), Phosphofurin acidic cluster sorting protein 1 (cg03914662, PACS1), Uveal Autoantigen With Coiled-Coil Domains And Ankyrin Repeats (cg09306458, UACA), Coiled-coil domain containing 50 (cg05359217, cg05946546, CCDC50), and Haemagglutinin stalk domain 17811 (cg02518222, HASD17811) were associated with an increased CVD risk.
[49]African American and EuropeanAtherosclerosis = 93 M and FWBIllumina Infinium HumanMethylation850 EPIC BeadChipFour CpGs were associated with atherosclerosis after adjusting for CVD risk. The CpGs included Aryl hydrocarbon receptor repressor (cg05575921, cg21161138, AHRR), Growth differentiation factor 11 (cg09936388, GF11), and Leucine Rich Repeat Containing 52 (cg181168448, LRRC52).
[50] American and EuropeanDiscovery cohort (REGICOR-1)
MI = 191
Controls = 195
Validation cohort (REGICOR-2)
MI = 101 Controls = 103
Follow-up association studies
WHI = 1863
FOS = 2540
M and FWBIllumina Infinium HumanMethylation850 EPIC BeadChipA total of 34 of the identified CpG sites associated with MI in both REGICOR-1 and REGICOR-2 samples. Of these, 25 were located within gene-coding regions, whereas 9 mapped to intergenic regions. Importantly, 3 of the identified CpG sites associated with CVD in follow up association studies using the WHI and FOS cohorts. The 3 identified CpGs were Aryl hydrocarbon receptor repressor (cg05575921, AHRR), F2R like thrombin or trypsin receptor 3 (cg21566642, F2RL3), and myeloperoxidase (cg04988978, MPO). More importantly, DNA methylation changes in cg21566642 showed a genetic influence in childhood (rs139595493) and adolescence (rs72617176).
[51] American and EuropeanDiscovery cohorts
FHS =1494 PIVUSs = 812
Validation cohorts
LBCs1921 ≤ 380
LBCs1936 ≤ 654 GOLDN = 991
M and FWBIllumina Infinium HumanMethylation450 BeadChipDifferential methylation was identified at 33 CpG sites in 5 different study cohorts. Of these, 25 CpG sites showed novel associations with circulating lipid traits (HDL, LDL, and TGs), including a CpG site annotated to the ATP Binding Cassette Subfamily G Member 1 locus (cg27243685, ABCG1). Methylation of this site correlated with a 38% increased risk of CHD.
[52] USA, UK,
Africa, Europe, Sweden
Discovery cohorts
FHS = 2377
LBCs = 1366
Validation cohorts
ARICs = 2096
GOLDN = 992
PIVUSs = 967
M and FWB and CD4+ cellsIllumina Infinium HumanMethylation450 BeadChipDifferential methylation in 135 CpGs were identified in the discovery cohorts, of which 83 could be replicated in the three validation cohorts. The study reported a three-way association between 11 differentially methylated CpGs, expression of 7 genes, and BMI. These associations included the DNA methylation and expression status of ATP Binding Cassette Subfamily G Member 1 (cg01881899, ABCG1), Carnitine palmitoyltransferase 1A (cg00574958, cg17058475, CPT1A), and Sterol regulatory element-binding protein 1 (cg11024682, SREBF1).
[53] CanadianDiscovery cohort
IUCPQ = 206
Validation cohort
CaG = 137
M and FVAT
WB
Methyl-binding domain-capture sequencingThis study identified 565 differentially methylated regulatory regions in adipose tissue, which were associated with plasma lipid levels. Of these, 340 were reflected in the whole blood, of which 68 (mapping to 13 putative enhancers and 55 promoters) shared similar associations with plasma lipid traits, including serine/threonine kinase 1 (AKT1), histone deacetylase 4 (HDAC4), Bone morphogenetic protein 4 (BMP4), Growth differentiation factor 7 (GDF7), Ceramide Kinase (CERK), Vestigial-like family member 3 (VGLL3), and ATP binding cassette subfamily C member 5 (ABCC5). Of the 68 CpG sites identified in the IUCPQ cohort, 22 could be validated in the CARTaGENE cohort.
[54] Ghanaian T2D = 589 M and FWBIllumina Infinium HumanMethylation450 BeadChipThe study identified 14 novel differentially methylated loci associated with CRP levels up to 40 mg/L. Three of the differentially methylated CpGs, annotating to Pyruvate Carboxylase (cg14653250, PC), BTG Anti-Proliferation Factor 4 (cg13767940, BTG4), and Peptidyl Arginine Deiminase 1 (cg21010178, PADI1), showed an association with CVD risk.
[26] GermanDiscovery cohort:
Atherosclerosis = 8
Controls = 8
Validation cohort:
AWHS = 24
M and FWBDiscovery:
Illumina Infinium HumanMethylation450 BeadChip
Validation:
Epityper MassARRAY
Differential methylation was identified in CpGs (287 hypo- and 229 hypermethylated) in cases versus controls which related to genes involved in cell cycle, adhesion, and death, among others. DNA methylation of breast cancer 1 (BRCA1) and Cysteine Rich Secretory Protein 2 (CRISP2) were validated using an Epityper MassARRAY and showed, in the additional AWHS cohort, that three CpG sites in CRISP2 (cg12440062, cg25390787, cg01076129) and one CpG site in BRCA1 (cg16630982) associated with subclinical atherosclerosis measures.
[55] Gambian
and Indian
n = 293 Gambian n = 698 IndianM and FWBIllumina Infinium HumanMethylation850 BeadChipSeveral differentially methylated regions were associated with diastolic blood pressure, insulin sensitivity, triglycerides, and LDL-cholesterol in the Gambian population, and insulin sensitivity and HDL-cholesterol in the Indian population. Importantly, there was no overlap of differentially methylated CpGs identified between the cohorts. The study identified significant quantitative trait loci (cis-methQTLs) at three LDL-cholesterol-associated differentially methylated CpGs in Gambians; however, methylation did not mediate genotype effects on the CVD outcomes in the Indian population.
[56] American and EuropeanARICs = 2567
CHS = 197
EPICORs = 584
FHS = 2375
InCHIANTI = 457
KORA = 1377
NAS = 484
WHI-EMPC = 1610
WHI-BAA23 = 1622
M and FPBLsIllumina Infinium HumanMethylation450 BeadChipAn association was found between methylation at 52 CpG sites and the risk of developing CHD or myocardial infarction in nine population-based cohorts. These CpG sites mapped to key regulatory regions of genes associated with calcium flux [ATPase Plasma Membrane Ca2+ Transporting 2 (ATP2B2), calcium sensing receptor (CASR), guanylate cyclase activator 1B (GUCAIB), and hippocalcin like 1 (HPCAL1)], arterial plaque calcification [Protein Tyrosine Phosphatase Receptor Type N2 (PTPRN2)], and kidney function [Cadherin Related 23 (CDH23) and Hippocalcin-like 1 (HPCAL1)]. Identified loci remained significant across all nine cohorts and significance was not influenced by race.
[57] ChinaHF = 27
Controls = 20
M and FPBLsReduced Representation Bisulfite Sequencing Assessment of DNA methylation in African American males younger or equal to 30 years of age revealed an increase in methylation of two CpGs in the Sulfatase 1 (SULF1) gene in HF patients, as compared to age-matched normotensive controls.
[58] European and African AmericanDiscovery cohorts
FHS offspring cohort = 2648
FHS third generation cohort = 1522
Validation cohorts
ARICs = 963
GTP = 384
M and FWB (buffy coat)Illumina Infinium HumanMethylation450 BeadChipDifferential methylation was observed at 92 CpGs and replicated in the validation cohorts. These CpG sites were associated with cardiovascular disease traits. Some of these CpGs were annotated to genes, such as Lipase A (LIPA), the ABO blood group gene (ABO), and Serologically defined colon cancer antigen 8 (SDCCAG8) that showed putative causality for CHD and MI
[59] SwedenMonozygotic twin pairs: 83
Dizygotic twin pairs 155
Single twins: 59
M and FWBIllumina Infinium HumanMethylation450 BeadChipDifferential methylation at 20 top-ranked CpGs associated with non-stroke CVD, overall stroke, and ischemic stroke. The methylation of these CpGs was shown to determine the levels of cardiometabolic trait (eg., BMI and blood pressure), with the latter mediating CVD risk.
[60] TurkeyCAD = 1
Control = 1
MWBReduced Representation Bisulfite SequencingEleven genes [Lipid Droplet Associated Hydrolase (LDAH), Apolipoprotein B (APOB), Acyl-CoA synthetase members 2A, 5 and 3 (ACSM2A, ACSM5, ACSM3), Carboxylesterase 1 (CES1), Carboxylesterase 1 Pseudogene 1 (CES1P1), AFG3 Like Matrix AAA Peptidase Subunit 2 (AFG3L2), Iron-Sulfur Cluster Assembly Enzyme (ISCU), SEC14 Like Lipid Binding 2 (SEC14L2), and Microsomal triglyceride transfer protein (MTTP)] involved in fatty acid and cholesterol metabolism were hypomethylated in the healthy twin as compared with twin that presenting with MI.
[61] JapanMI = 192
Control = 192
MWBIllumina Infinium HumanMethylation450 BeadChipAn association was identified between DNA methylation at 2 CpG sites [cg07786668 in zinc finer homeobox 3 (ZFHX3) and cg17218495 in SWI/SNF-related, matrix-associated, actin-dependent regulator of chromatin, subfamily a, member 4 (SMARCA4)] and MI, which were independent of CVD risk factors, such as BMI, blood lipid levels, and T2D.
[62] GermanyDiscovery Cohorts:
KORA = 1103
NAS = 344
Validation Cohort:
InCHIANTI = 443
M and FPBLsIllumina Infinium HumanMethylation450 BeadChipMethylation at 11 CpGs was related to incidents of MI, although after adjusting for medication usage, 9 CpGs remained and composed an epigenetic fingerprint for MI. This included genes such as LDL Receptor Related Protein 8 (LRP8) and KCNN1 potassium calcium-activated channel subfamily N member 1 (KCCN1).
Table 4. Classification of main study findings based on disease outcome.
Table 4. Classification of main study findings based on disease outcome.
Disease PhenotypeStudiesMain Genes IdentifiedSummary of Conclusions
MI and CHD[24,30,32,36,39,41,42,44,50,56,61,62]ZBTB12, BMPR2, IL-6, DYSF, NRG1, AHRR, F2RL3, MPO, ZFHX3, FOXP3, SOAT1, SMARCA4 and TXNIP, CASR, GUCAIB, HPCAL1, PTPRN2, and CDH23Identified genes involved in pathways related to inflammation, vascular remodeling, and oxidative stress.
Identified gene associations independent from traditional risk factors, suggesting an intrinsic epigenetic contribution to disease risk.
Stroke[45,59] ASB10 and TTC37Identified methylation changes may contribute to stroke pathophysiology through mechanisms involving homocysteine metabolism, blood pressure, and BMI regulation.
Some stroke-related epigenetic associations were population-specific (e.g., observed only in Europeans).
HF, general CVD, and atherosclerosis[26,31,37,38,40,43,46,47,48,49,54]HF severity: HDAC9, JARID2, GREM1, and SULF1
CVD risk: CYP8B1, SPON1, PACS1, UACA, NPPA, AHRR, and CPT1A
Atherosclerosis: IL-6, AHRR, hTERT, GF11, LRRC52, NR3C1, miRNA-233, CRISP2, and BRCA1
Hypo- and hypermethylation of distinct CpGs linked to cardiac stress responses, atherosclerosis, and inflammation.
Associations with CRP levels and kidney function genes suggest inflammation and metabolic dysfunction as key contributors to CVD risk.
Multiple CVD traits[27,34,51,52,53,55,58,60]ABCG1, CPT1A, FABP3, SREBF1, LIPA, SDCCAG8, ABO, TLR2, IFN-γ, and iNOSIdentified differentially methylated genes associated with CVD risk factors: lipid metabolism, insulin sensitivity, diastolic blood pressure, and BMI.
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

Willmer, T.; Mabasa, L.; Sharma, J.; Muller, C.J.F.; Johnson, R. Blood-Based DNA Methylation Biomarkers to Identify Risk and Progression of Cardiovascular Disease. Int. J. Mol. Sci. 2025, 26, 2355. https://doi.org/10.3390/ijms26052355

AMA Style

Willmer T, Mabasa L, Sharma J, Muller CJF, Johnson R. Blood-Based DNA Methylation Biomarkers to Identify Risk and Progression of Cardiovascular Disease. International Journal of Molecular Sciences. 2025; 26(5):2355. https://doi.org/10.3390/ijms26052355

Chicago/Turabian Style

Willmer, Tarryn, Lawrence Mabasa, Jyoti Sharma, Christo J. F. Muller, and Rabia Johnson. 2025. "Blood-Based DNA Methylation Biomarkers to Identify Risk and Progression of Cardiovascular Disease" International Journal of Molecular Sciences 26, no. 5: 2355. https://doi.org/10.3390/ijms26052355

APA Style

Willmer, T., Mabasa, L., Sharma, J., Muller, C. J. F., & Johnson, R. (2025). Blood-Based DNA Methylation Biomarkers to Identify Risk and Progression of Cardiovascular Disease. International Journal of Molecular Sciences, 26(5), 2355. https://doi.org/10.3390/ijms26052355

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