Skip to main content

The emerging role of epigenetics in human autoimmune disorders

Abstract

Epigenetic pathways play a pivotal role in the development and function of the immune system. Over the last decade, a growing body of studies has been published out seeking to explain a correlation between epigenetic modifications and the development of autoimmune disorders. Epigenetic changes, such as DNA methylation, histone modifications, and noncoding RNAs, are involved in the pathogenesis of autoimmune diseases mainly by regulating gene expression. This paper reviews the importance of epigenetic alterations during the development of the most prevalent human autoimmune diseases, such as systemic lupus erythematosus (SLE), rheumatoid arthritis (RA), systemic sclerosis (SSc), Sjogren’s syndrome (SS), autoimmune thyroid diseases (AITD), and type 1 diabetes (T1D), aiming to provide new insights in the pathogenesis of autoimmune diseases and the possibility to develop novel therapeutic approaches targeting the epigenome.

Background

Epigenetic mechanisms, known for their ability to regulate gene transcription and genomic stability, are key players for maintaining normal cell growth, development, and differentiation [1]. The term “epigenetics” can be outlined as the meiotically/mitotically heritable alterations in gene expression, related to environmental factors, without changes to the sequence of bases in the DNA [1]. Since genome-wide profiling in some cases does not give a sufficient answer to explain the complex biological processes in autoimmune disorders, epigenetic modifications are retained additional regulators in immune responses (Fig. 1). Epigenetic dysregulation directly influences the development of autoimmunity by regulating immune cell functions [2]. The recognition of the complexity of the interaction between epigenetic events and the alteration of the immune system in autoimmune disorders is a prominent challenge for the discovery of novel potential therapeutic strategies. Epigenetic mechanisms, such as DNA methylation, chromatin remodeling, and noncoding RNAs, have been identified as crucial regulators in cellular immunity, owing to their mechanisms in modulating gene expression and transcription in targeted cells and tissues [3]. Extensive evidences indicate that autoimmune diseases are mainly an interplay of genetic and non-genetic factors, although the role of the latter ones often remains unclear. Over the last decade, the influence of epigenetic modifications on innate and adaptive immunity has been intensively investigated, especially in autoimmune disorders.

Fig. 1
figure 1

Schematic representation of key epigenetic mechanisms (DNA methylation, histone post-translational modifications, and RNA-based mechanism) involved in the pathogenesis of autoimmune diseases, including SLE, RA, SSc, SS, AITD, and T1D

Epigenetic mechanisms

DNA methylation

DNA methylation is one of the most studied epigenetic marks and typically occurs at the pyrimidine C5 position of cytosine residues by transferring a methyl group from S-adenosyl-l-methionine (SAM) through the catalytic action of DNA methyltransferases (DNMTs) [4,5,6]. DNA methylation influences a large variety of biological processes such as transcriptional repression, reversible promoter silencing, and chromosomal instability [7, 8]. The various functions and diverse subtypes of DNMTs have been summarized well in various reviews [9,10,11].

One of the crucial functions of DNA methylation is the maintenance of the T cell regulation. Recent findings underlined the critical role of DNA methylation in numerous autoimmune diseases by altering gene expression profiles [12,13,14,15]. Several factors such as environmental influences, genetic variants, drugs, and miRNAs influencing the DNA methylation status have been linked to autoimmune disorders.

Histone modifications

Histone-modifying enzymes have an essential role in modulating chromatin compaction state, nucleosomal processes, and DNA repair [16,17,18]. Some HDAC inhibitors (HDACi) have been widely studied as a means to increase the transcription of various genes, including those involved in autoimmunity [19,20,21]. Moreover, HDACi have been described to influence immune and inflammatory processes, mainly involving T cells [22, 23]. Acetyl and methyl marks inserted on histones could also interact with the corresponding “reader” proteins to add complexity to this scenario. Some evidence suggested that posttranslational modifications are implicated in the development and regulation of different cell lines and in the modulation of immune tolerance and autoimmune disorders [24, 25].

Noncoding RNAs

Long noncoding RNAs (lncRNAs) are a huge and diverse class of transcribed RNA molecules with a length of more than 200 nucleotides. Usually, they are capped, polyadenylated, and spliced without having any coding function for proteins [26]. LncRNAs have been described as key players of gene regulation in various human pathologies due to their implication in numerous cellular and biological events regulating heterochromatin formation, histone modifications, DNA methylation targeting, and gene silencing [27, 28]. Furthermore, lncRNA can activate regulatory complexes regulating the development and differentiation of various immune cell types via an active protein expression control. MiRNAs are responsible for the regulation of about 60% of all mRNAs and are implicated in numerous disorders, such as cancers, inflammatory processes, and metabolic diseases [29, 30]. Due to their essential role in the hematopoietic development, as well as cell activation and differentiation, abnormal expression of miRNAs might lead to the onset of autoimmune diseases [31, 32].

Systemic and organ-specific autoimmune diseases

Autoimmune diseases are characterized by an immune response to antigenic components of the host itself (autoantigens). Two main types of autoimmune diseases can be distinguished: on the one hand, the systemic ones and on the other, the organ-specific ones. In systemic diseases, the immune system attacks in a generalized manner its own antigens in several organs, while in organ-specific diseases the immune response is directed towards a single organ. Examples of systemic and organ-specific autoimmune diseases with known autoantigen targets are illustrated in Table 1.

Table 1 Most relevant autoimmune diseases with known autoantigen targets

Systemic autoimmune rheumatic diseases (SARDs)

Systemic autoimmune rheumatic diseases (SARDs) belong to a group of rare chronic inflammatory conditions including rheumatoid arthritis (RA), systemic lupus erythematosus (SLE), systemic sclerosis (SSc), and Sjogren’s syndrome (SS). Autoimmunity is defined by the breakdown of self-tolerance that produces a state of abnormal humoral and cell-mediated responses against self-components. Collectively, SARDs affect up to 5% of the world population and, in particular, people of work-force age [33, 34]. Until now, no effective treatments have been identified for SARDs, even though the use of glucocorticoids has been considered as a first-line therapy. Nowadays, however, antimalarial and immunosuppressive drugs are most commonly used due to their limited long-term side effects. Such autoimmune disorders are often associated with an autoimmune dysregulation which determines morbidity and, in most cases, premature mortality [35, 36]. In particular, most of these conditions happen when the immune system produces autoantibodies (ANA) directed against intracellular antigens. So, ANA are considered as convincing serological hallmarks of SARDs, which are routinely detected via an indirect immunofluorescence (IF) assay [37]. Although many progresses have been performed in the last 15 years, there is an unmet need to find an innovative and successful therapy to fight SARDs, especially RA, the most prevalent autoimmune disease. Understanding the molecular mechanisms of SARDs appears to be extremely important to achieve beneficial outcomes in these chronic conditions. However, similar biologic pathways that underlie SARDs in RA, SLE, and SSc may suggest novel mechanistic similarities. Indeed, these three chronic diseases have in common demographic distribution (the most affected individuals are women), some signs and symptoms (arthritis, lung and vascular disease), serological elements (ANA and anti-Ro52/TRIM21 antibodies), immunological components (type I interferon signature and complex abnormalities in CD4C T lymphocyte function, in particular Th17 and Treg cell subsets), and genetic similarities (e.g., MHC class II alleles, IRF5, STAT4, PTPN22 loci) [38,39,40,41,42,43,44]. More and more studies demonstrated that epigenetically altered immune components, such as CD4+ T cells and costimulatory molecules, are key drivers of SARDs, thereby compromising cellular immune system function and regulation [45,46,47,48]. Characterization of epigenetic modifications that occur across these autoimmune diseases may yield valuable insights into their pathogenesis and treatment. Thus, in an attempt to determine the most important epigenetic changes in SARDs, researchers investigated the role of epigenetic processes in regulating autoimmunity. For many years, epigenetic implications for the most important related forms of autoimmunity including RA, SLE, SSc, and SS have been studied in order to find a tight association between epigenetics and systemic autoimmune, providing novel regulatory mechanisms for SARDs (Table 2).

Table 2 Epigenetic changes in systemic lupus erythematosus (SLE), rheumatoid arthritis (RA), systemic sclerosis (SSc) and Sjogren’s syndrome (SS)

Systemic lupus erythematosus (SLE)

SLE is the most studied autoimmune disease correlated with epigenetic modification. It especially occurs in women and is mainly caused by dysregulation of T lymphocytes, making the disorder complicated and hard to handle. SLE is a chronic autoimmune dysfunction characterized by the development of autoantibodies against nuclear antigens affecting any organ system and tissue, such as kidney and blood vessels. In a recent study, neutrophils and granulocytes from patients with SLE have been described as totally hypomethylated, especially at the gene locus of the interferons MX1 and IFI44L [49]. Discrepancies in monozygotic twins propone environmental factors as crucial drivers for the development of SLE. Epigenetic alterations such as DNA methylation and histone modifications have been found to be able to regulate gene expression in mature T cells. Numerous genes such as CD11a (ITGAL), perforin (PRF1), CD70 (TNFSF7), and CD40LG (TNFSF5) in T lymphocytes of SLE patients were found to be hypomethylated [50]. Early studies in SLE CD4+ T cells demonstrated a conversion to autoreactivity and an induced lupus-like syndrome after treatment with DNMT inhibitors [51,52,53]. The auto-reactivity has been shown only in cloned and polyclonal human and murine CD4+ cells, but not in CD8+ cells, the reason is still undetermined [53, 54]. However, there is strong evidence that auto-reactivity development is concomitant with an increased expression of the adhesion molecules LFA-1 (CD11a/CD18), due to elevated levels of CD11a (ITGAL) transcripts. ITGAL is an integrin responsible for costimulation and cellular adhesion. The upstream promoter of ITGAL can be found to be demethylated in SLE patients’ CD4+ cells and depending on the disease activity, and progression CD11a can be found more or less overexpressed [55]. The methylation status of other specific genes has been linked with SLE pathogenesis and development. CD40L, a type II transmembrane protein encoded on the X chromosome by CD40LG and functioning as a costimulatory molecule, has been found to be overexpressed in human SLE patients. It still remains elusive why SLE predominantly affects women. CD40LG methylation patterns have been associated with female susceptibility to this disease [56]. Following treatment with DNMT inhibitors (azacytidine, procainamide) or ERK pathway inhibitors (hydralazine, PD98059), demethylated CD40LG led to induced T cell autoreactivity in vitro [56]. Overexpressed E4BP4 (NFIL3), an important human transcription factor, led to the downregulation of the autoimmune responses in SLE patients through inhibiting CD40L expression [57]. Another interesting costimulatory molecule is CD70, encoded by TNFSF7, a ligand of CD27. Disruption of CD70-CD27 interaction via blocking antibody allows attenuating lectin-stimulated B cells’ IgG production in vitro [58]. Overexpression in a mouse model can be caused both by traditional demethylating agents and medications associated with drug-induced lupus, including hydralazine and procainamide [59]. Overexpression of CD11a and CD70 in CD4+ cells after histone methyltransferase SUV39H1 recruitment resulted in decreased regulatory factor X1 (RFX1) levels are important for the determination of CD4+ T cell auto-reactivity [60, 61]. Sunahori et al. could demonstrate that inhibiting the catalytic subunit of protein phosphatase 2A (PP2Ac) increased MEK/ERK phosphorylation levels and elevated DNA methylation levels as well as diminished CD70 gene expression [62]. Perforin (PRF1), a key regulator gene in cytotoxic CD8+ and NK cells, produces a cytolytic protein allowing the disruption and lysis of target cellular membranes [63]. PRF1 overexpression in both CD4+ and CD8+ cells, mediated by the gene promoting methylation state, has been linked to SLE. Overexpression of IL10 in T cells from SLE patients is modulated by abnormal STAT3 activation through the histone acetyltransferase p300 leading to an increase of specific autoantibody production and tissue damage [64]. Another important interleukin, IL13, has been considered to play a pivotal role in the development of both human and murine lupus via Th2 cell differentiation [65]. Histone modification pattern in lupus has been less investigated than DNA methylation. A heavily methylated HDAC6 promoter resulted in lower HDAC6 mRNA expression in SLE patients if compared to healthy controls [66]. MRLlpr lupus-prone mouse splenocytes showed increased methylation as well as decreased acetylation of histones H3 and H4 compared to control mice; treatment with HDAC inhibitors (HDACi) normalizes aberrant gene expression thus reducing disease activity. However, when lupus T cells are treated by HDACi, including trichostatin A (TSA) and suberoylanilide hydroxamic acid (SAHA), alterations of acetylation levels of acid nuclear transport proteins, transcription factors, and cytoskeleton proteins have been reported [67]. Thus, no valid evidence has been found yet to connect the histone modifications with SLE activity. Several recent studies investigated the role of lncRNAs in lupus pathogenesis. MiR-21, miR-148a, and miR126 are three microRNAs regulated by methylation that are matched with a decreased expression of DNMTs in CD4+ T cells of SLE [68]. MiR-148a elicited the expression of CD70 and CD11a, similar to lupus patients [69]. Further, overexpressed miR-155 has been found in Treg cell of MRLlpr mice. Its T cell distribution regulating activity has been proven in miR-155 deficient mice which display reduced serum levels of IL4 and IL17A, two specific cytokines secreted by Th2 and Th17 cells, respectively [70]. Another study describes increased serum levels of RANTES in SLE patients harboring, at the same time, lowered MiR-125a expression [71].

Rheumatoid arthritis (RA)

RA is a chronic and debilitating inflammatory sickness causing destructive arthritis with diffuse damages to joints. Epigenetic mechanisms involved in RA include altered methylation stati in T and B cells as well as in synovial fibroblasts [72,73,74]. Early methylation studies revealed that patients’ T cells showed a remarkable phenotype, similar to SLE, characterized by global hypomethylation [75, 76]. In particular, methotrexate (MTX), used to treat RA, led to the accumulation of protective Treg cells by inducing FOXP3 expression through the promoter demethylation. However, other secondary effects of MTX, which is namely a folate antagonist, on the RA evolution cannot yet be ruled out. Further studies have shown that MTX can reverse the hypomethylated status in peripheral blood mononuclear cells (PBMC) [77, 78]. In 2012, Nakano and colleagues reported the first epigenome-wide study comparing RA and osteoarthritis (OA) FLS cell lines: out of 1859 differently methylated loci, around 60% were identified as hypermethylated [79]. Thanks to a second recent genome-wide study by de La Rica et al., novel target genes have been discovered as being differently methylated, including IL6R, CAPN8, and DPP4. Comparing rheumatoid arthritis synovial fibroblast (RASF) DNA methylation with miRNA expression and RASF transcriptome data from the Gene Expression Omnibus (GEO), more than 200 of the 714 genes identified had an inverse expression. Moreover, several CpG sites have been detected to be hypermethylated with concomitantly reduced miRNA expression [80]. Histone modifications associated with RA have been less studied; although, an involvement of these epigenetic changes has been documented in RA pathogenesis. Dysregulated HDAC activity in RA PBMCs was not affected by conventional anti-TNF therapies whereas treatment with MI192, a HDAC2/3-selective inhibitor, reduced IL6 production in a dose-dependent manner, thus providing a novel therapeutic approach for RA [81]. Wendling et al. evaluated the relationship between Sirt1 activity/expression and IL-3 levels in PBMCs. High levels of serum IL-3 associated with decreased Sirt1 expression and increased apoptosis in patients with RA have been observed [82]. Kawabata et al. reported an increased HDAC1 expression by TNF-α supplementation in RA synovial fibroblasts [83,84,85]. HDACi showed anti-inflammatory properties in FLS attenuating disease in animal models of RA. An interesting small open-label trial of givinostat (a HDACi) in a similar disorder, juvenile idiopathic arthritis, outlined improvement in arthritis and a wide safety profile [86]. In RA synovial tissues, the equilibrium between HATs and HDACs activity is heavily disturbed owing to the soaring of HAT activity resulting in hyperacetylation [87]. An increase in IL6 expression by hyperacetylation of histone H3 has been found in synovial fibroblasts [88]. Ahmed et al. demonstrated that largazole, a marine-derived class I-selective HDACi, provokes the suppression of the TNFα-induced expression of the intracellular adhesion molecule-1 (ICAM-1) and the vascular adhesion molecule-1 (VCAM-1) in RASF, as well as in the reduction of the TNFα-induced MMP2 activity. Additionally, largazole was shown to modulate expression levels of HDAC1, HDAC5, and HDAC6. Of particular interest is the role of HDAC6 in largazole-induced changes of ICAM-1 and VCAM-1 expression levels [89]. Studies trying to explain the effects of microRNAs on RA pathogenesis are emerging more and more in the recent literature. The upregulation of miR-146a with TNF-α and at the same time the downregulation of miR-363 and miR-498 has been found in CD4+ cells of RA patients [90]. Despite these pieces of evidence, various studies have shown that miR-146a and miR-155 were decreased in Treg cells after T cell stimulation in RA patients [91]. Additionally, the expression of miR-126a in RA turned out to be elevated, leading to hypomethylated promotors of CD11a and CD70 which in turn led to their overexpression [92]. Further, the increased expression of miR-21 resulted in Treg cell accumulation in synovial fibroblasts of patients suffering from RA [93].

Systemic sclerosis (SSc)

SSc is a rare and poorly understood autoimmune disease of the connective tissue leading to excessive collagen deposition in the skin and other organs often with a lethal outcome. Aberrant activation of fibroblasts and collagen secretion in SSc conduces to fibrosis. Like SLE, early studies have shown a link to T cell dysfunction and, in particular, autoreactive T cell transfer signals to surrounding fibroblasts inducing the deposit of collagen and initiation of fibrosis [94]. The hypomethylation of CD4+ cells, caused at least in part by the downregulation of DNMTs, determined the overexpression of several genes involved in disease progression [95]. Notably, the downregulation of functional demethylating enzymes such as DNMT1, MBD3, and MBD4 resulted in lower levels of methylation at the promoter sites of sensitive genes implicated in SS leading to the overexpression of CD40L, CD11a, and CD70 as a relevant feature of SSc [96,97,98]. The overexpression of the adaptive immune costimulatory molecule CD40L, which is crucial for the integrative role in fibrosis of SSc, is a clear sign of the disease in plasma and skin fibroblasts [99, 100]. Furthermore, abrogation of CD40/CD40L interaction reduces fibrosis in a SSc mouse model [101]. Recently, other pathways related to SSc have been discussed with particular emphasis on their epigenetic dysregulation mechanisms. One such pathway, Wnt, increases skin fibrosis in SSc patients [102]. Importantly, Dickkopf-1 and SFRP-1 are both hypermethylated in dermal fibroblasts and PBMCs in a mouse model of SSc. Thus, treatment with 5-azacytidine reduces Wnt signaling and deletes the fibrosis phenotype [103]. Likewise, inactivation of the collagen suppressor gene FLI1 by epigenetic hypermethylation reduces the expression of type I collagen in fibroblasts from SSc patients [104]. As aforementioned, fibroblasts play a pivotal role in SSc pathogenesis due to collagen and extracellular matrix component overproduction. As a consequence, various cytokines may be produced to dysregulated immune cells further [105]. Reduced H3K27me3 levels were observed as an important posttranslational modification of CD4+ cells in SSc leading to the accumulation of the jumonij demethylase JMJD3 in the aforementioned cells [106]. Similarly, to SLE, several histone modifications have been noted in SSc. Both H3 and H4 acetylation is reduced in SSc fibroblasts [104]. Indeed, the reduced expression of collagen in cultured SSc fibroblasts, as well as fibrosis in animal models, have been reported following HDACi treatment [107, 108]. MicroRNAs have been implicated in SSc concerning to dysregulated fibrosis. In details, miR-29a is decreased in SSc fibroblasts leading to a reduction in Co1 1 and 3 expressions [109]. MiR-29a can avoid pulmonary fibrosis and its knock-down increases pro-fibrotic TGF-β and PDGF-B level. [110]. MiR-21 is regulated by TGF- β and targets smad-7, a key profibrotic gene [111, 112]. At last, miR-155 mediates fibrosis by recruiting keratinocyte growth factor, and overexpressed miR-196 reduces collagen levels by regulating collagen 1 [113, 114].

Sjogren’s syndrome (SS)

SS is an autoimmune disease affecting salivary and tear glands of unknown origin resulting in eye and mouth dryness. SS is a quite prevalent autoimmune disorder affecting about four million people in the USA. This autoimmune illness is associated with autoantibody production in the blood that is directed against various tissues of the body leading to an inflammatory response. Similarly to other autoimmune diseases, reduced DNA methylation of immune cells is one of the most studied epigenetic marks in SS. Yin et al. found that the T cell costimulatory gene CD70 overexpression on CD4+ T cells is due to the hypomethylation of CD70 promoterin SS patients [115]. By contrast, lower expression levels of FOXP3 in SS CD4+ T cells have been correlated with DNA hypermethylation [116]. A genome-wide DNA methylation study in naïve CD4+CD45RA+ cells resulted in the identification of 553 methylated CpG sites. LTA, encoding lymphotoxin α, has been found overexpressed in both salivary gland tissue and sera of SS patients [117]. Therefore, its downregulation in the salivary gland of SS mouse model prevents the development of SS. Next, a broader comprehensive study of genome-wide DNA methylation patterns was carried out in whole blood, peripheral CD19+ B cells, and minor salivary glands [118]. Differentially hypomethylated sites were identified in type I IFN-induced genes such as MX1, IFI44L, PARP9, and IFITM1. The hypomethylation of their promoter in SS B cells was linked to elevated mRNA expression levels. Several other genes have been noted to be differentially methylated, including hypomethylated STAT1, IFI44L, IFITM1, and USP8. Conversely, the RUNX1 gene, an important transcription factor involved in T cell development, was hypermethylated among SS patients. Due to reduced expression of DNMT1 and an increase in GADD45α, epithelial cells from the salivary gland (SGEC) showed global hypomethylation in SS patients. This finding might be associated with infiltrating B lymphocytes. Increased DNA methylation levels in SGEC from SS patients have been reported after the administration of the anti-CD20 monoclonal antibody rituximab [119]. Konsta et al. showed that reduced DNA methylation levels in minor salivary glands might be correlated with overexpression of the epithelial protein cytokeratin-19 (gene KRT19) in glandular acini, while high methylation levels resulted in low levels of this protein proven by immunohistochemical analyses [120]. A follow-up study of the same authors proved their concept as the treatment of a human salivary gland cell line with the DNMT inhibitor 5-azacytidine led to KRT19 mRNA and cytokeratin-19 protein overexpression [121]. To sum up, altered DNA methylation patterns in SGEC may be an important factor in SS pathophysiology probably at least partially via the controlling of the KRT19/cytokeratin-19 expression. Actually, no large-scale studies of histone modifications as well as of lncRNAs have been carried out in SS. Increased miR-146a expression has been observed in PBMCs during the onset of SS [122].

Organ-specific autoimmune diseases

Organ-specific autoimmune diseases, as suggested by the name, are immune response attacks of the body towards healthy cells of a specific organ system occurring as a result of either genetic predisposition or environmental influences or a mixture of both. Many efforts have been performed to understand the inherited autoimmune responses [123]. A large number of polymorphic genes are key players in the genesis of these chronic autoimmune disorders. To date, most details in these processes still need to be understood better even though a large number these genes are apparently involved in setting a threshold for an immune response. Organ-specific autoimmune diseases are usually related to specific human leukocyte class II antigens [124]. In many cases, specific antibodies may be found in patients with chronic and apparently “idiopathic” organ-specific autoimmune diseases. Basically, these antibodies bind to self-antigens in the organ cells or directly on cells, thus leading to their destruction [125]. Examples are the autoimmune thyroid diseases (AITD) and type 1 diabetes (T1D), in which ANA attack the thyroid gland and the immune system which compromises the pancreas, respectively [126, 127]. Normally, the regulation of the autoimmune process involves antigen-specific regulatory cells.

Anti-inflammatory cytokines such as IL-10 and TGF-β are also involved. Up to now, blockers of the immune response produced a greater success in the clinical use than treatments exploiting natural immune regulation. In fact, blocking the immune response is crucial in autoimmunity, even though immunosuppression leads to various side effects, including the reactivation of latent infections and the reduction of immunosurveillance. Thus, antigen-specific immune therapeutical options, instead of rather unspecific therapies targeting the immune system, are an important goal to reach towards the treatment of autoimmune disorders. Over the last decade, scientists figured out that epigenetic changes are clearly correlated with organ-specific autoimmune diseases (Table 3). With the aim to better elucidate mechanisms behind these organ-specific diseases, several studies have been carried out using approved epi-drugs [128, 129] (Table 4). Importantly, the discovery of an epigenetic therapy to treat such autoimmune disorders may unearth potential biomarkers for disease diagnosis and prediction.

Table 3 Main epigenetic modifications in autoimmune thyroid diseases (AITDs) and type 1 diabetes (T1D)
Table 4 Effects of Epi-drugs on autoimmune disorders discussed in this review

Autoimmune thyroid diseases (AITDs)

AITDs are a form of autoimmune diseases mediated by B and T cells. The two main clinical manifestations of AITDs are Flajani-Basedow-Graves’ disease or Graves’ disease (GD) and Hashimoto’s thyroiditis (HT) [130, 131]. HT led to hypothyroidism by cell-mediated autoimmune destruction, while in GD specific autoantibodies against the thyroid stimulation hormone receptor result in hyperthyroidism [132]. Recent literature shed light on the epigenetic mechanisms involved in the pathogenesis of AITD. DNA methylation, histone modifications, and lncRNA have been deeply analyzed in AITDs. However, the clinical utility of epigenetic modulation still remains elusive. Recent findings confirmed that DNA methylation is also in AITDs setting a very crucial epigenetic mechanism. Global DNA hypomethylation was observed in AITD patients, which may result in overexpressed genes important for a correct immune function, or for the activation of immune cells, ultimately leading into an autoimmune attack towards thyroid tissues [133, 134]. In a genome-wide study, Cai et al. have detected in GD patients more than 200 hyper- and hypo-methylated genetic regions, such as ICAM1, which partly controls cell antigen processing and presentation; DNMT1; and MECP2 genes [133]. Via epigenetic profiling in CD4+ and CD8+ cells from GD patients, hypermethylated gene loci of ICAM1, CD247, and CTLA4 associated with T cell receptor signaling were observed. [133, 134]. Furthermore, hypermethylation of the first intron area in the TSHR gene confirmed that methylation is involved in the development of AITDs. However, DNA hypomethylation levels and susceptibility to AITDs have been correlated to the presence of genetic polymorphisms of DNA methylation-regulatory genes, such as DNMT1 or methionine synthase reductase (MTRR) [135]. Histone modifications play a key role in AITDs, but the exact mechanism(s) in modulating immune tolerance in AITDs is still not fully elucidated. Yan et al. described higher levels of HDAC1 and HDAC2 mRNAs in GD patients thus histone H4 acetylation levels in peripheral blood mononuclear cells of GD patients were lower if compared to those observed in healthy control patient. These results underline the potential importance of aberrant histone modifications in GD patients [24]. In patients with GD, CD4+ and CD8+ cells were presented higher levels of H3K4me3 and H3K27Ac histone marks [134]. IFN-α, an important cytokine secreted during viral infections, has been identified to be able to lead to higher mono and trimethylation levels of H3K4m in thyroid cells [136]. As previously seen for DNA methylation, genetic polymorphism of histone-modifying genes may result in various malfunctions, and/or other further aberrant histone modifications. Sarumaru et al. demonstrated that rs3758391 and rs4746720 in the SIRT1 gene were linked to higher titers of autoantibodies in AITD affected patients [137]. As previously illustrated for other autoimmune disorders, microRNAs actively modulate in various circumstances the differentiation or activation of immune cells and immune response. Two of the most studied miRNA are miR-155-5p and miR-146a-5p, whose overexpression is believed to break immune tolerance thus fostering the development of autoimmune diseases. For example, GD and HT patients exhibit markedly lower levels of miR-155-5p and miR-146a-5p in the thyroid tissues [138]. Although many research’s efforts showed the potential role of other microRNAs, such as miR-125a-3p, miR-197a-3p, miR-22-3p, and miR-183-5p, only a small number of studies investigated their clinical relevance as diagnostic biomarkers [139,140,141]. Another epigenetic process called X chromosome inactivation (XCI) has also been discussed in correlation with AITDs [142,143,144]. X chromosomes are randomly inactivated in females resulting in transcriptional silencing of one of the X chromosomes [145, 146]. Indeed, AITDs are more often observed in females, confirming the compelling role of XCI in these autoimmune disorders. Brix et al. demonstrated that skewed X chromosome inactivation in female twins with GD and HT was significantly higher than in the control populations, suggesting a probable role of XCI in the etiology of AITD [147]. However, Ishido et al. reported no apparent difference between AITD cases and controls. By contrast, a notable relationship between skewed XCI inactivation and the prognosis of GD and HT was observed [148].

Type 1 diabetes (T1D)

T1D is a chronic autoimmune disease that involves β cell destruction together with a strong inflammatory response. The incidence of diabetes is on the rise worldwide becoming one of the major causes of death. In more detail, the body’s own immune system is destroying the insulin-producing pancreatic β cells first lowering and ultimately ceasing the ability of the pancreas to produce insulin. To date, the disease cannot be cured, although exogenous insulin therapy still remains a life-saving therapy. The destructive process causing T1D is thought to have a pertinent adaptive autoimmune component. In fact, persuasive shreds of evidence suggested that aberrant epigenetic modifications are involved in T1D pathogenesis. The epigenetic involvement in T1D has been nicely reviewed recently [149]. However, environmental factors, in different models of diabetes, influence epigenetic changes and consequently contribute significantly to altered gene expression relevant in T1D development and progression. Over the last decade, some studies have reported that DNA methylation, histone modifications, and noncoding RNAs are considered to have a crucial role in T1D. Again, a genome-wide DNA methylation identified SNP-CpG methylation patterns as potentially relevant for the genetic association of insulin expression and/or secretion in human pancreatic islets [150]. In more detail, candidate genes, such as GPX7, GSTT1, and SNX19, are known to possess a direct influence on various biological processes regarding proliferation and apoptosis in pancreatic β cells. Another study in mouse models also confirmed the implication of epigenetic changes in insulin secretion and diabetes risk [151]. Patients with T1D, compared to the healthy controls, exhibited cell-type-specific gene regulatory circuits crucial for immune cell metabolism and the cell cycle. The latter includes mTOR signaling pathway implicated in the development of diabetes-associated damage [152]. It is believed that epigenetic alteration in diabetes might be the cause for an increased risk for the development and progression of vascular complications. Very likely, these effects are mediated through histone methyltransferases which boost the pro-inflammatory networks implicated in vascular injury [153,154,155]. Hypermethylated regions of FOXP3 in CD4+ cells in T1D resulted in decreased FOXP3 expression and reduced production of regulatory T cells [156]. Additionally, Wang et al. supported the idea of the FOXP3 involvement as its gene hypermethylation was induced by Toll-like receptor 9 (TLR9) in association with a reduced binding activity of interferon regulator factor 7 (IRF-7) [157]. About histone modifications, increased H3K9me2 levels in the promoter of CLTA4, a T1D susceptibility gene, has been correlated with T cell activation [158]. Moreover, Patel et al. suggested that TSA, a well-known HDACi, excites IFN-γ production and increases the transcription activity of Tbx21 in T lymphocytes, alleviating the inflammatory damage of islets [159]. Since T1D shows a broad range of miRNA expression profiles, some miRNAs have been associated with T1D. Upregulated miR-510, as well as the downregulation of miR-342 and miR-191, have been described in T1D. Furthermore, once Treg cells are compared with other types of effector T cells, data showed a substantial change in the expression patterns of miR-146a in conjugation with a decreased expression of eight targeting miRNAs (20b, 31, 99a, 100, 125b, 151, 335, and 365) thus demonstrating the involvement of miRNAs in T1D patients’ Treg cells [160].

Benefits and risk of epigenetic therapy

As we have seen in the previous sections, mainly old well-studied and often already approved epigenetic modulators have been used in the context of autoimmune diseases [56, 59, 67, 81, 84, 86, 89, 103, 121, 159]. These first-generation molecules have been originally developed as anti-tumor therapy. They often show heavy side effects by mechanism, which are tolerable in cancer, but not in other diseases [9, 19]. This point is one major drawback and risk of the studies currently present in literature. The main challenge for researchers is not only to understand better the role of epigenetics in autoimmune disorders also with the help of epigenetic modulators, but also to develop novel specific targeted therapies. Interestingly, novel second or third generation epigenetic modulators, which often have a better selectivity and safety profile, have been used in other medical conditions but have not yet been applied in immunotherapy approaches [18, 161]. In our view, benefits outweigh the risks as novel innovative treatments are needed, but researchers should proceed with care considering the selectivity and safety profile of their potential treatments already at an early stage of the therapy development.

Conclusions

This work has been focused on epigenetic mechanisms regulating several autoimmune disorders with the aim to provide new therapeutic ideas in this area of interest (Fig. 1). The link between epigenetics and autoimmunity has been widely documented in the literature; although, more and more extensive studies are required to better understand the influence of these modifications in the different autoimmune disorders. A deeper exploration of the complex epigenetic interactions may be useful for the development of promising treatment strategies targeting the epigenome. The fundamental work of Farh et al. in 2015 described a fine-mapping algorithm to identify non-coding genetic variants that could underlie autoimmune diseases from genotyping data [162]. However, to date, there is still a lot of research necessary in order to provide effective healthcare solutions for such diseases, such as an adequate choice of treatment and a precise prediction of treatment outcomes. Epigenetics will very likely aid to provide further progress in the field of autoimmunity. Until now, only a few studies have been carried out to examine the clinical applicability of epigenetics and its modulators in such autoimmune diseases. Keeping this concept in mind, more comprehensive and more accessible technologies are recommended to promote advances in the therapeutics leading to prevention by early interventions and reducing both costs and patient morbidity. Taken together, the future of epigenetics in autoimmunity is rapidly increasing. However, it still needs an in-depth investigation to allow novel potential therapeutic opportunities, better than those currently used.

Abbreviations

AITD:

Autoimmune thyroid diseases

DNMTs:

DNA methyltransferases

GD:

Graves’ disease

GEO:

Gene Expression Omnibus

HATs:

Histone acetyltransferases

HDACi:

Histone deacetylases inhibitors

HDACs:

Histone deacetylases

HT:

Hashimoto thyroiditis

ICAM-1:

Intracellular adhesion molecule-1

IF:

Immunofluorescence

IRF-7:

Interferon regulator factor 7

LincRNAs:

Long intergenic noncoding RNAs

LncRNAs:

Long noncoding RNAs

LRA:

Largazole

miRNAs:

microRNAs

MTRR:

Methionine synthase reductase

OA:

Osteoarthritis

PBMC:

Peripheral blood mononuclear cells

PP2Ac:

Protein phosphatase 2A

PRF1:

Perforin

RA:

Rheumatoid arthritis

RASF:

Rheumatoid arthritis synovial fibroblasts

RFX1:

Regulatory factor X1

SAM:

S-adenosyl-methionine

SGEC:

Salivary gland

SLE:

Systemic lupus erythematosus

SS:

Sjogren’s syndrome

SSc:

Systemic sclerosis

T1D:

Type 1 diabetes

TLR9:

Toll-like receptor 9

TSA:

Trichostatin A

VCAM-1:

Vascular adhesion molecule-1

XCI:

X chromosome inactivation

References

  1. Bird A. Perceptions of epigenetics. Nature. 2007;447(7143):396–8. https://doi.org/10.1038/nature05913.

    Article  CAS  PubMed  Google Scholar 

  2. Moosavi A, Motevalizadeh AA. Role of epigenetics in biology and human diseases. Iran Biomed J. 2016;20(5):246–58.

    PubMed  PubMed Central  Google Scholar 

  3. Mazzone R, Zwergel C, Mai A, Valente S. Epi-drugs in combination with immunotherapy: a new avenue to improve anticancer efficacy. Clin Epigenetics. 2017;9:59. https://doi.org/10.1186/s13148-017-0358-y.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Du J, Johnson LM, Jacobsen SE, Patel DJ. DNA methylation pathways and their crosstalk with histone methylation. Nat Rev Mol Cell Biol. 2015;16(9):519–32. https://doi.org/10.1038/nrm4043.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Torres IO, Fujimori DG. Functional coupling between writers, erasers and readers of histone and DNA methylation. Curr Opin Struct Biol. 2015;35:68–75. https://doi.org/10.1016/j.sbi.2015.09.007.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Schubeler D. Function and information content of DNA methylation. Nature. 2015;517(7534):321–6. https://doi.org/10.1038/nature14192.

    Article  CAS  PubMed  Google Scholar 

  7. Bjornsson HT, Fallin MD, Feinberg AP. An integrated epigenetic and genetic approach to common human disease. Trends Genet. 2004;20(8):350–8. https://doi.org/10.1016/j.tig.2004.06.009.

    Article  CAS  PubMed  Google Scholar 

  8. Vaissiere T, Sawan C, Herceg Z. Epigenetic interplay between histone modifications and DNA methylation in gene silencing. Mutat Res. 2008;659(1–2):40–8. https://doi.org/10.1016/j.mrrev.2008.02.004.

    Article  CAS  PubMed  Google Scholar 

  9. Zwergel C, Valente S, Mai A. DNA methyltransferases inhibitors from natural sources. Curr Top Med Chem. 2016;16(7):680–96. https://doi.org/10.2174/1568026615666150825141505.

    Article  CAS  PubMed  Google Scholar 

  10. Jurkowski TP, Jeltsch A. On the evolutionary origin of eukaryotic DNA methyltransferases and Dnmt2. PLoS One. 2011;6(11):e28104. https://doi.org/10.1371/journal.pone.0028104.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Jurkowska RZ, Jeltsch A. Mechanisms and biological roles of DNA methyltransferases and DNA methylation: from past achievements to future challenges. Adv Exp Med Biol. 2016;945:1–17. https://doi.org/10.1007/978-3-319-43624-1_1.

    Article  CAS  PubMed  Google Scholar 

  12. Wadhwa E, Nicolaides T. Bromodomain inhibitor review: Bromodomain and extra-terminal family protein inhibitors as a potential new therapy in central nervous system tumors. Cureus. 2016;8(5):e620. https://doi.org/10.7759/cureus.620.

    Article  PubMed  PubMed Central  Google Scholar 

  13. Chen Z, Miao F, Paterson AD, Lachin JM, Zhang L, Schones DE, et al. Epigenomic profiling reveals an association between persistence of DNA methylation and metabolic memory in the DCCT/EDIC type 1 diabetes cohort. Proc Natl Acad Sci U S A. 2016;113(21):E3002–11. https://doi.org/10.1073/pnas.1603712113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Paul DS, Teschendorff AE, Dang MA, Lowe R, Hawa MI, Ecker S, et al. Increased DNA methylation variability in type 1 diabetes across three immune effector cell types. Nat Commun. 2016;7:13555. https://doi.org/10.1038/ncomms13555.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Tahara T, Hirata I, Nakano N, Nagasaka M, Nakagawa Y, Shibata T, et al. Comprehensive DNA methylation profiling of inflammatory mucosa in ulcerative colitis. Inflamm Bowel Dis. 2017;23(1):165–73. https://doi.org/10.1097/MIB.0000000000000990.

    Article  PubMed  Google Scholar 

  16. Lawrence M, Daujat S, Schneider R. Lateral thinking: how histone modifications regulate gene expression. Trends Genet. 2016;32(1):42–56. https://doi.org/10.1016/j.tig.2015.10.007.

    Article  CAS  PubMed  Google Scholar 

  17. Venkatesh S, Workman JL. Histone exchange, chromatin structure and the regulation of transcription. Nat Rev Mol Cell Biol. 2015;16(3):178–89. https://doi.org/10.1038/nrm3941.

    Article  CAS  PubMed  Google Scholar 

  18. Dieker J, Muller S. Epigenetic histone code and autoimmunity. Clin Rev Allergy Immunol. 2010;39(1):78–84. https://doi.org/10.1007/s12016-009-8173-7.

    Article  CAS  PubMed  Google Scholar 

  19. Zwergel C, Valente S, Jacob C, Mai A. Emerging approaches for histone deacetylase inhibitor drug discovery. Expert Opin Drug Discovery. 2015;10(6):599–613. https://doi.org/10.1517/17460441.2015.1038236.

    Article  CAS  Google Scholar 

  20. Chistiakov DA, Orekhov AN, Bobryshev YV. Treatment of cardiovascular pathology with epigenetically active agents: focus on natural and synthetic inhibitors of DNA methylation and histone deacetylation. Int J Cardiol. 2017;227:66–82. https://doi.org/10.1016/j.ijcard.2016.11.204.

    Article  PubMed  Google Scholar 

  21. Falkenberg KJ, Johnstone RW. Histone deacetylases and their inhibitors in cancer, neurological diseases and immune disorders. Nat Rev Drug Discov. 2014;13(9):673–91. https://doi.org/10.1038/nrd4360.

    Article  CAS  PubMed  Google Scholar 

  22. Akimova T, Beier UH, Liu Y, Wang L, Hancock WW. Histone/protein deacetylases and T-cell immune responses. Blood. 2012;119(11):2443–51. https://doi.org/10.1182/blood-2011-10-292003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Panella S, Marcocci ME, Celestino I, Valente S, Zwergel C, Li Puma DD, et al. MC1568 inhibits HDAC6/8 activity and influenza a virus replication in lung epithelial cells: role of Hsp90 acetylation. Future Med Chem. 2016;8(17):2017–31. https://doi.org/10.4155/fmc-2016-0073.

    Article  CAS  PubMed  Google Scholar 

  24. Yan N, Zhou JZ, Zhang JA, Cai T, Zhang W, Wang Y, et al. Histone hypoacetylation and increased histone deacetylases in peripheral blood mononuclear cells from patients with graves’ disease. Mol Cell Endocrinol. 2015;414:143–7. https://doi.org/10.1016/j.mce.2015.05.037.

    Article  CAS  PubMed  Google Scholar 

  25. Cheng F, Lienlaf M, Wang HW, Perez-Villarroel P, Lee C, Woan K, et al. A novel role for histone deacetylase 6 in the regulation of the tolerogenic STAT3/IL-10 pathway in APCs. J Immunol. 2014;193(6):2850–62. https://doi.org/10.4049/jimmunol.1302778.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Rinn JL, Chang HY. Genome regulation by long noncoding RNAs. Annu Rev Biochem. 2012;81:145–66. https://doi.org/10.1146/annurev-biochem-051410-092902.

    Article  CAS  PubMed  Google Scholar 

  27. Khalil AM, Guttman M, Huarte M, Garber M, Raj A, Rivea Morales D, et al. Many human large intergenic noncoding RNAs associate with chromatin-modifying complexes and affect gene expression. Proc Natl Acad Sci U S A. 2009;106(28):11667–72. https://doi.org/10.1073/pnas.0904715106.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Guttman M, Amit I, Garber M, French C, Lin MF, Feldser D, et al. Chromatin signature reveals over a thousand highly conserved large non-coding RNAs in mammals. Nature. 2009;458(7235):223–7. https://doi.org/10.1038/nature07672.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Jonas S, Izaurralde E. Towards a molecular understanding of microRNA-mediated gene silencing. Nat Rev Genet. 2015;16(7):421–33. https://doi.org/10.1038/nrg3965.

    Article  CAS  PubMed  Google Scholar 

  30. Vicente R, Noel D, Pers YM, Apparailly F, Jorgensen C. Deregulation and therapeutic potential of microRNAs in arthritic diseases. Nat Rev Rheumatol. 2016;12(8):496. https://doi.org/10.1038/nrrheum.2016.119.

    Article  CAS  PubMed  Google Scholar 

  31. Chen JQ, Papp G, Szodoray P, Zeher M. The role of microRNAs in the pathogenesis of autoimmune diseases. Autoimmun Rev. 2016;15(12):1171–80. https://doi.org/10.1016/j.autrev.2016.09.003.

    Article  CAS  PubMed  Google Scholar 

  32. Fedeli M, Riba M, Garcia Manteiga JM, Tian L, Vigano V, Rossetti G, et al. miR-17 approximately 92 family clusters control iNKT cell ontogenesis via modulation of TGF-beta signaling. Proc Natl Acad Sci U S A. 2016;113(51):E8286–E95. https://doi.org/10.1073/pnas.1612024114.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Jacobson DL, Gange SJ, Rose NR, Graham NM. Epidemiology and estimated population burden of selected autoimmune diseases in the United States. Clin Immunol Immunopathol. 1997;84(3):223–43. https://doi.org/10.1006/clin.1997.4412.

    Article  CAS  PubMed  Google Scholar 

  34. Helmick CG, Felson DT, Lawrence RC, Gabriel S, Hirsch R, Kwoh CK, et al. Estimates of the prevalence of arthritis and other rheumatic conditions in the United States. Part I Arthritis Rheum. 2008;58(1):15–25. https://doi.org/10.1002/art.23177.

    Article  PubMed  Google Scholar 

  35. Walsh SJ, Rau LM. Autoimmune diseases: a leading cause of death among young and middle-aged women in the United States. Am J Public Health. 2000;90(9):1463–6. https://doi.org/10.2105/AJPH.90.9.1463.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Marie I, Hachulla E, Hatron PY, Hellot MF, Levesque H, Devulder B, et al. Polymyositis and dermatomyositis: short term and longterm outcome, and predictive factors of prognosis. J Rheumatol. 2001;28(10):2230–7.

    CAS  PubMed  Google Scholar 

  37. Meroni PL, Schur PH. ANA screening: an old test with new recommendations. Ann Rheum Dis. 2010;69(8):1420–2. https://doi.org/10.1136/ard.2009.127100.

    Article  CAS  PubMed  Google Scholar 

  38. Kavanaugh A, Tomar R, Reveille J, Solomon DH, Homburger HA. Guidelines for clinical use of the antinuclear antibody test and tests for specific autoantibodies to nuclear antigens. American College of Pathologists. Arch Pathol Lab Med. 2000;124(1):71–81. https://doi.org/10.1043/0003-9985(2000)124<0071:GFCUOT>2.0.CO;2.

    Article  CAS  PubMed  Google Scholar 

  39. Hudson M, Pope J, Mahler M, Tatibouet S, Steele R, Baron M, et al. Clinical significance of antibodies to Ro52/TRIM21 in systemic sclerosis. Arthritis Res Ther. 2012;14(2):R50. https://doi.org/10.1186/ar3763.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Higgs BW, Zhu W, Richman L, Fiorentino DF, Greenberg SA, Jallal B, et al. Identification of activated cytokine pathways in the blood of systemic lupus erythematosus, myositis, rheumatoid arthritis, and scleroderma patients. Int J Rheum Dis. 2012;15(1):25–35. https://doi.org/10.1111/j.1756-185X.2011.01654.x.

    Article  CAS  PubMed  Google Scholar 

  41. Cho JH, Gregersen PK. Genomics and the multifactorial nature of human autoimmune disease. N Engl J Med. 2011;365(17):1612–23. https://doi.org/10.1056/NEJMra1100030.

    Article  CAS  PubMed  Google Scholar 

  42. Davidson A, Diamond B. Autoimmune diseases. N Engl J Med. 2001;345(5):340–50. https://doi.org/10.1056/NEJM200108023450506.

    Article  CAS  PubMed  Google Scholar 

  43. Cotsapas C, Voight BF, Rossin E, Lage K, Neale BM, Wallace C, et al. Pervasive sharing of genetic effects in autoimmune disease. PLoS Genet. 2011;7(8):e1002254. https://doi.org/10.1371/journal.pgen.1002254.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Hudson M, Rojas-Villarraga A, Coral-Alvarado P, Lopez-Guzman S, Mantilla RD, Chalem P, et al. Polyautoimmunity and familial autoimmunity in systemic sclerosis. J Autoimmun. 2008;31(2):156–9. https://doi.org/10.1016/j.jaut.2008.05.002.

    Article  CAS  PubMed  Google Scholar 

  45. Jeffries MA, Sawalha AH. Autoimmune disease in the epigenetic era: how has epigenetics changed our understanding of disease and how can we expect the field to evolve? Expert Rev Clin Immunol. 2015;11(1):45–58. https://doi.org/10.1586/1744666X.2015.994507.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Strickland FM, Richardson BC. Epigenetics in human autoimmunity. Epigenetics in autoimmunity - DNA methylation in systemic lupus erythematosus and beyond. Autoimmunity. 2008;41(4):278–86. https://doi.org/10.1080/08916930802024616.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Brooks WH, Le Dantec C, Pers JO, Youinou P, Renaudineau Y. Epigenetics and autoimmunity. J Autoimmun. 2010;34(3):J207–19. https://doi.org/10.1016/j.jaut.2009.12.006.

    Article  CAS  PubMed  Google Scholar 

  48. Meda F, Folci M, Baccarelli A, Selmi C. The epigenetics of autoimmunity. Cell Mol Immunol. 2011;8(3):226–36. https://doi.org/10.1038/cmi.2010.78.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Coit P, Yalavarthi S, Ognenovski M, Zhao W, Hasni S, Wren JD, et al. Epigenome profiling reveals significant DNA demethylation of interferon signature genes in lupus neutrophils. J Autoimmun. 2015;58:59–66. https://doi.org/10.1016/j.jaut.2015.01.004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Richardson B. Primer: epigenetics of autoimmunity. Nat Clin Pract Rheumatol. 2007;3(9):521–7. https://doi.org/10.1038/ncprheum0573.

    Article  CAS  PubMed  Google Scholar 

  51. Yung RL, Quddus J, Chrisp CE, Johnson KJ, Richardson BC. Mechanism of drug-induced lupus. I. Cloned Th2 cells modified with DNA methylation inhibitors in vitro cause autoimmunity in vivo. J Immunol. 1995;154(6):3025–35.

    CAS  PubMed  Google Scholar 

  52. Quddus J, Johnson KJ, Gavalchin J, Amento EP, Chrisp CE, Yung RL, et al. Treating activated CD4+ T cells with either of two distinct DNA methyltransferase inhibitors, 5-azacytidine or procainamide, is sufficient to cause a lupus-like disease in syngeneic mice. J Clin Invest. 1993;92(1):38–53. https://doi.org/10.1172/JCI116576.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Richardson B. Effect of an inhibitor of DNA methylation on T cells. II. 5-Azacytidine induces self-reactivity in antigen-specific T4+ cells. Hum Immunol. 1986;17(4):456–70. https://doi.org/10.1016/0198-8859(86)90304-6.

    Article  CAS  PubMed  Google Scholar 

  54. Richardson BC, Liebling MR, Hudson JL. CD4+ cells treated with DNA methylation inhibitors induce autologous B cell differentiation. Clin Immunol Immunopathol. 1990;55(3):368–81. https://doi.org/10.1016/0090-1229(90)90125-A.

    Article  CAS  PubMed  Google Scholar 

  55. Lu Q, Kaplan M, Ray D, Ray D, Zacharek S, Gutsch D, et al. Demethylation of ITGAL (CD11a) regulatory sequences in systemic lupus erythematosus. Arthritis Rheum. 2002;46(5):1282–91. https://doi.org/10.1002/art.10234.

    Article  CAS  PubMed  Google Scholar 

  56. Lu Q, Wu A, Tesmer L, Ray D, Yousif N, Richardson B. Demethylation of CD40LG on the inactive X in T cells from women with lupus. J Immunol. 2007;179(9):6352–8. https://doi.org/10.4049/jimmunol.179.9.6352.

    Article  CAS  PubMed  Google Scholar 

  57. Zhao M, Liu Q, Liang G, Wang L, Luo S, Tang Q, et al. E4BP4 overexpression: a protective mechanism in CD4+ T cells from SLE patients. J Autoimmun. 2013;41:152–60. https://doi.org/10.1016/j.jaut.2013.01.004.

    Article  CAS  PubMed  Google Scholar 

  58. Sugita K, Torimoto Y, Nojima Y, Daley JF, Schlossman SF, Morimoto C. The 1A4 molecule (CD27) is involved in T cell activation. J Immunol. 1991;147(5):1477–83.

    CAS  PubMed  Google Scholar 

  59. Oelke K, Lu Q, Richardson D, Wu A, Deng C, Hanash S, et al. Overexpression of CD70 and overstimulation of IgG synthesis by lupus T cells and T cells treated with DNA methylation inhibitors. Arthritis Rheum. 2004;50(6):1850–60. https://doi.org/10.1002/art.20255.

    Article  CAS  PubMed  Google Scholar 

  60. Zhao M, Sun Y, Gao F, Wu X, Tang J, Yin H, et al. Epigenetics and SLE: RFX1 downregulation causes CD11a and CD70 overexpression by altering epigenetic modifications in lupus CD4+ T cells. J Autoimmun. 2010;35(1):58–69. https://doi.org/10.1016/j.jaut.2010.02.002.

    Article  CAS  PubMed  Google Scholar 

  61. Zhao M, Wu X, Zhang Q, Luo S, Liang G, Su Y, et al. RFX1 regulates CD70 and CD11a expression in lupus T cells by recruiting the histone methyltransferase SUV39H1. Arthritis Res Ther. 2010;12(6):R227. https://doi.org/10.1186/ar3214.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Sunahori K, Nagpal K, Hedrich CM, Mizui M, Fitzgerald LM, Tsokos GC. The catalytic subunit of protein phosphatase 2A (PP2Ac) promotes DNA hypomethylation by suppressing the phosphorylated mitogen-activated protein kinase/extracellular signal-regulated kinase (ERK) kinase (MEK)/phosphorylated ERK/DNMT1 protein pathway in T-cells from controls and systemic lupus erythematosus patients. J Biol Chem. 2013;288(30):21936–44. https://doi.org/10.1074/jbc.M113.467266.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Podack ER, Young JD, Cohn ZA. Isolation and biochemical and functional characterization of perforin 1 from cytolytic T-cell granules. Proc Natl Acad Sci U S A. 1985;82(24):8629–33. https://doi.org/10.1073/pnas.82.24.8629.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Hedrich CM, Rauen T, Apostolidis SA, Grammatikos AP, Rodriguez Rodriguez N, Ioannidis C, et al. Stat3 promotes IL-10 expression in lupus T cells through trans-activation and chromatin remodeling. Proc Natl Acad Sci U S A. 2014;111(37):13457–62. https://doi.org/10.1073/pnas.1408023111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Zhao M, Tang J, Gao F, Wu X, Liang Y, Yin H, et al. Hypomethylation of IL10 and IL13 promoters in CD4+ T cells of patients with systemic lupus erythematosus. J Biomed Biotechnol. 2010;2010:931018. https://doi.org/10.1155/2010/931018.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Fang TJ, Lin YZ, Liu CC, Lin CH, Li RN, Wu CC, et al. Methylation and gene expression of histone deacetylases 6 in systemic lupus erythematosus. Int J Rheum Dis. 2016;19(10):968–73. https://doi.org/10.1111/1756-185X.12783.

    Article  CAS  PubMed  Google Scholar 

  67. Marks PA, Dokmanovic M. Histone deacetylase inhibitors: discovery and development as anticancer agents. Expert Opin Investig Drugs. 2005;14(12):1497–511. https://doi.org/10.1517/13543784.14.12.1497.

    Article  CAS  PubMed  Google Scholar 

  68. Zhao S, Wang Y, Liang Y, Zhao M, Long H, Ding S, et al. MicroRNA-126 regulates DNA methylation in CD4+ T cells and contributes to systemic lupus erythematosus by targeting DNA methyltransferase 1. Arthritis Rheum. 2011;63(5):1376–86. https://doi.org/10.1002/art.30196.

    Article  CAS  PubMed  Google Scholar 

  69. Pan W, Zhu S, Yuan M, Cui H, Wang L, Luo X, et al. MicroRNA-21 and microRNA-148a contribute to DNA hypomethylation in lupus CD4+ T cells by directly and indirectly targeting DNA methyltransferase 1. J Immunol. 2010;184(12):6773–81. https://doi.org/10.4049/jimmunol.0904060.

    Article  CAS  PubMed  Google Scholar 

  70. Xin Q, Li J, Dang J, Bian X, Shan S, Yuan J, et al. miR-155 deficiency ameliorates autoimmune inflammation of systemic lupus erythematosus by targeting S1pr1 in Faslpr/lpr mice. J Immunol. 2015;194(11):5437–45. https://doi.org/10.4049/jimmunol.1403028.

    Article  CAS  PubMed  Google Scholar 

  71. Lu MM, Wang J, Pan HF, Chen GM, Li J, Cen H, et al. Increased serum RANTES in patients with systemic lupus erythematosus. Rheumatol Int. 2012;32(5):1231–3. https://doi.org/10.1007/s00296-010-1761-2.

    Article  CAS  PubMed  Google Scholar 

  72. Smolen JS, Aletaha D, McInnes IB. Rheumatoid arthritis. Lancet. 2016;388(10055):2023–38. https://doi.org/10.1016/S0140-6736(16)30173-8.

    Article  CAS  PubMed  Google Scholar 

  73. Joo YB, Park Y, Kim K, Bang SY, Bae SC, Lee HS. Association of CD8(+) T-cells with bone erosion in patients with rheumatoid arthritis. Int J Rheum Dis. 2018;21(2):440–6. https://doi.org/10.1111/1756-185X.13090.

    Article  CAS  PubMed  Google Scholar 

  74. Glossop JR, Emes RD, Nixon NB, Packham JC, Fryer AA, Mattey DL, et al. Genome-wide profiling in treatment-naive early rheumatoid arthritis reveals DNA methylome changes in T and B lymphocytes. Epigenomics. 2016;8(2):209–24. https://doi.org/10.2217/epi.15.103.

    Article  CAS  PubMed  Google Scholar 

  75. Richardson B, Scheinbart L, Strahler J, Gross L, Hanash S, Johnson M. Evidence for impaired T cell DNA methylation in systemic lupus erythematosus and rheumatoid arthritis. Arthritis Rheum. 1990;33(11):1665–73. https://doi.org/10.1002/art.1780331109.

    Article  CAS  Google Scholar 

  76. Corvetta A, Della Bitta R, Luchetti MM, Pomponio G. 5-Methylcytosine content of DNA in blood, synovial mononuclear cells and synovial tissue from patients affected by autoimmune rheumatic diseases. J Chromatogr. 1991;566(2):481–91. https://doi.org/10.1016/0378-4347(91)80265-E.

    Article  CAS  Google Scholar 

  77. Cribbs AP, Kennedy A, Penn H, Amjadi P, Green P, Read JE, et al. Methotrexate restores regulatory T cell function through demethylation of the FoxP3 upstream enhancer in patients with rheumatoid arthritis. Arthritis Rheumatol. 2015;67(5):1182–92. https://doi.org/10.1002/art.39031.

    Article  CAS  PubMed  Google Scholar 

  78. de Andres MC, Perez-Pampin E, Calaza M, Santaclara FJ, Ortea I, Gomez-Reino JJ, et al. Assessment of global DNA methylation in peripheral blood cell subpopulations of early rheumatoid arthritis before and after methotrexate. Arthritis Res Ther. 2015;17:233. https://doi.org/10.1186/s13075-015-0748-5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Nakano K, Whitaker JW, Boyle DL, Wang W, Firestein GS. DNA methylome signature in rheumatoid arthritis. Ann Rheum Dis. 2013;72(1):110–7. https://doi.org/10.1136/annrheumdis-2012-201526.

    Article  CAS  PubMed  Google Scholar 

  80. de la Rica L, Urquiza JM, Gomez-Cabrero D, Islam AB, Lopez-Bigas N, Tegner J, et al. Identification of novel markers in rheumatoid arthritis through integrated analysis of DNA methylation and microRNA expression. J Autoimmun. 2013;41:6–16. https://doi.org/10.1016/j.jaut.2012.12.005.

    Article  CAS  PubMed  Google Scholar 

  81. Gillespie J, Savic S, Wong C, Hempshall A, Inman M, Emery P, et al. Histone deacetylases are dysregulated in rheumatoid arthritis and a novel histone deacetylase 3-selective inhibitor reduces interleukin-6 production by peripheral blood mononuclear cells from rheumatoid arthritis patients. Arthritis Rheum. 2012;64(2):418–22. https://doi.org/10.1002/art.33382.

    Article  CAS  PubMed  Google Scholar 

  82. Wendling D, Abbas W, Godfrin-Valnet M, Kumar A, Guillot X, Khan KA, et al. Dysregulated serum IL-23 and SIRT1 activity in peripheral blood mononuclear cells of patients with rheumatoid arthritis. PLoS One. 2015;10(3):e0119981. https://doi.org/10.1371/journal.pone.0119981.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Kawabata T, Nishida K, Takasugi K, Ogawa H, Sada K, Kadota Y, et al. Increased activity and expression of histone deacetylase 1 in relation to tumor necrosis factor-alpha in synovial tissue of rheumatoid arthritis. Arthritis Res Ther. 2010;12(4):R133. https://doi.org/10.1186/ar3071.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Nishida K, Komiyama T, Miyazawa S, Shen ZN, Furumatsu T, Doi H, et al. Histone deacetylase inhibitor suppression of autoantibody-mediated arthritis in mice via regulation of p16INK4a and p21(WAF1/Cip1) expression. Arthritis Rheum. 2004;50(10):3365–76. https://doi.org/10.1002/art.20709.

    Article  CAS  PubMed  Google Scholar 

  85. Horiuchi M, Morinobu A, Chin T, Sakai Y, Kurosaka M, Kumagai S. Expression and function of histone deacetylases in rheumatoid arthritis synovial fibroblasts. J Rheumatol. 2009;36(8):1580–9. https://doi.org/10.3899/jrheum.081115.

    Article  CAS  PubMed  Google Scholar 

  86. Vojinovic J, Damjanov N, D'Urzo C, Furlan A, Susic G, Pasic S, et al. Safety and efficacy of an oral histone deacetylase inhibitor in systemic-onset juvenile idiopathic arthritis. Arthritis Rheum. 2011;63(5):1452–8. https://doi.org/10.1002/art.30238.

    Article  CAS  PubMed  Google Scholar 

  87. Huber LC, Brock M, Hemmatazad H, Giger OT, Moritz F, Trenkmann M, et al. Histone deacetylase/acetylase activity in total synovial tissue derived from rheumatoid arthritis and osteoarthritis patients. Arthritis Rheum. 2007;56(4):1087–93. https://doi.org/10.1002/art.22512.

    Article  CAS  PubMed  Google Scholar 

  88. Wada TT, Araki Y, Sato K, Aizaki Y, Yokota K, Kim YT, et al. Aberrant histone acetylation contributes to elevated interleukin-6 production in rheumatoid arthritis synovial fibroblasts. Biochem Biophys Res Commun. 2014;444(4):682–6. https://doi.org/10.1016/j.bbrc.2014.01.195.

    Article  CAS  PubMed  Google Scholar 

  89. Ahmed S, Riegsecker S, Beamer M, Rahman A, Bellini JV, Bhansali P, et al. Largazole, a class I histone deacetylase inhibitor, enhances TNF-alpha-induced ICAM-1 and VCAM-1 expression in rheumatoid arthritis synovial fibroblasts. Toxicol Appl Pharmacol. 2013;270(2):87–96. https://doi.org/10.1016/j.taap.2013.04.014.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Li J, Wan Y, Guo Q, Zou L, Zhang J, Fang Y, et al. Altered microRNA expression profile with miR-146a upregulation in CD4+ T cells from patients with rheumatoid arthritis. Arthritis Res Ther. 2010;12(3):R81. https://doi.org/10.1186/ar3006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Zhou Q, Haupt S, Kreuzer JT, Hammitzsch A, Proft F, Neumann C, et al. Decreased expression of miR-146a and miR-155 contributes to an abnormal Treg phenotype in patients with rheumatoid arthritis. Ann Rheum Dis. 2015;74(6):1265–74. https://doi.org/10.1136/annrheumdis-2013-204377.

    Article  CAS  PubMed  Google Scholar 

  92. Yang G, Wu D, Zeng G, Jiang O, Yuan P, Huang S, et al. Correlation between miR-126 expression and DNA hypomethylation of CD4+ T cells in rheumatoid arthritis patients. Int J Clin Exp Pathol. 2015;8(8):8929–36.

    CAS  PubMed  PubMed Central  Google Scholar 

  93. van der Geest KS, Smigielska-Czepiel K, Park JA, Abdulahad WH, Kim HW, Kroesen BJ, et al. SF Treg cells transcribing high levels of Bcl-2 and microRNA-21 demonstrate limited apoptosis in RA. Rheumatology (Oxford). 2015;54(5):950–8. https://doi.org/10.1093/rheumatology/keu407.

    Article  CAS  Google Scholar 

  94. Hugle T, O'Reilly S, Simpson R, Kraaij MD, Bigley V, Collin M, et al. Tumor necrosis factor-costimulated T lymphocytes from patients with systemic sclerosis trigger collagen production in fibroblasts. Arthritis Rheum. 2013;65(2):481–91. https://doi.org/10.1002/art.37738.

    Article  CAS  PubMed  Google Scholar 

  95. Lei W, Luo Y, Lei W, Luo Y, Yan K, Zhao S, et al. Abnormal DNA methylation in CD4+ T cells from patients with systemic lupus erythematosus, systemic sclerosis, and dermatomyositis. Scand J Rheumatol. 2009;38(5):369–74. https://doi.org/10.1080/03009740902758875.

    Article  CAS  Google Scholar 

  96. Lian X, Xiao R, Hu X, Kanekura T, Jiang H, Li Y, et al. DNA demethylation of CD40l in CD4+ T cells from women with systemic sclerosis: a possible explanation for female susceptibility. Arthritis Rheum. 2012;64(7):2338–45. https://doi.org/10.1002/art.34376.

    Article  CAS  PubMed  Google Scholar 

  97. Wang Y, Shu Y, Xiao Y, Wang Q, Kanekura T, Li Y, et al. Hypomethylation and overexpression of ITGAL (CD11a) in CD4(+) T cells in systemic sclerosis. Clin Epigenetics. 2014;6(1):25. https://doi.org/10.1186/1868-7083-6-25.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Jiang H, Xiao R, Lian X, Kanekura T, Luo Y, Yin Y, et al. Demethylation of TNFSF7 contributes to CD70 overexpression in CD4+ T cells from patients with systemic sclerosis. Clin Immunol. 2012;143(1):39–44. https://doi.org/10.1016/j.clim.2012.01.005.

    Article  CAS  PubMed  Google Scholar 

  99. Fukasawa C, Kawaguchi Y, Harigai M, Sugiura T, Takagi K, Kawamoto M, et al. Increased CD40 expression in skin fibroblasts from patients with systemic sclerosis (SSc): role of CD40-CD154 in the phenotype of SSc fibroblasts. Eur J Immunol. 2003;33(10):2792–800. https://doi.org/10.1002/eji.200324088.

    Article  CAS  PubMed  Google Scholar 

  100. Allanore Y, Borderie D, Meune C, Lemarechal H, Weber S, Ekindjian OG, et al. Increased plasma soluble CD40 ligand concentrations in systemic sclerosis and association with pulmonary arterial hypertension and digital ulcers. Ann Rheum Dis. 2005;64(3):481–3. https://doi.org/10.1136/ard.2003.020040.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Komura K, Fujimoto M, Yanaba K, Matsushita T, Matsushita Y, Horikawa M, et al. Blockade of CD40/CD40 ligand interactions attenuates skin fibrosis and autoimmunity in the tight-skin mouse. Ann Rheum Dis. 2008;67(6):867–72. https://doi.org/10.1136/ard.2007.073387.

    Article  CAS  PubMed  Google Scholar 

  102. Akhmetshina A, Palumbo K, Dees C, Bergmann C, Venalis P, Zerr P, et al. Activation of canonical Wnt signalling is required for TGF-beta-mediated fibrosis. Nat Commun. 2012;3:735. https://doi.org/10.1038/ncomms1734.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Dees C, Schlottmann I, Funke R, Distler A, Palumbo-Zerr K, Zerr P, et al. The Wnt antagonists DKK1 and SFRP1 are downregulated by promoter hypermethylation in systemic sclerosis. Ann Rheum Dis. 2014;73(6):1232–9. https://doi.org/10.1136/annrheumdis-2012-203194.

    Article  CAS  PubMed  Google Scholar 

  104. Wang Y, Fan PS, Kahaleh B. Association between enhanced type I collagen expression and epigenetic repression of the FLI1 gene in scleroderma fibroblasts. Arthritis Rheum. 2006;54(7):2271–9. https://doi.org/10.1002/art.21948.

    Article  CAS  PubMed  Google Scholar 

  105. Xiao C, Rajewsky K. MicroRNA control in the immune system: basic principles. Cell. 2009;136(1):26–36. https://doi.org/10.1016/j.cell.2008.12.027.

    Article  CAS  PubMed  Google Scholar 

  106. Wang Q, Xiao Y, Shi Y, Luo Y, Li Y, Zhao M, et al. Overexpression of JMJD3 may contribute to demethylation of H3K27me3 in CD4+ T cells from patients with systemic sclerosis. Clin Immunol. 2015;161(2):396–9. https://doi.org/10.1016/j.clim.2015.03.006.

    Article  CAS  PubMed  Google Scholar 

  107. Huber LC, Distler JH, Moritz F, Hemmatazad H, Hauser T, Michel BA, et al. Trichostatin a prevents the accumulation of extracellular matrix in a mouse model of bleomycin-induced skin fibrosis. Arthritis Rheum. 2007;56(8):2755–64. https://doi.org/10.1002/art.22759.

    Article  CAS  PubMed  Google Scholar 

  108. Hemmatazad H, Rodrigues HM, Maurer B, Brentano F, Pileckyte M, Distler JH, et al. Histone deacetylase 7, a potential target for the antifibrotic treatment of systemic sclerosis. Arthritis Rheum. 2009;60(5):1519–29. https://doi.org/10.1002/art.24494.

    Article  PubMed  Google Scholar 

  109. Maurer B, Stanczyk J, Jungel A, Akhmetshina A, Trenkmann M, Brock M, et al. MicroRNA-29, a key regulator of collagen expression in systemic sclerosis. Arthritis Rheum. 2010;62(6):1733–43. https://doi.org/10.1002/art.27443.

    Article  CAS  PubMed  Google Scholar 

  110. Xiao J, Meng XM, Huang XR, Chung AC, Feng YL, Hui DS, et al. miR-29 inhibits bleomycin-induced pulmonary fibrosis in mice. Mol Ther. 2012;20(6):1251–60. https://doi.org/10.1038/mt.2012.36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Sing T, Jinnin M, Yamane K, Honda N, Makino K, Kajihara I, et al. microRNA-92a expression in the sera and dermal fibroblasts increases in patients with scleroderma. Rheumatology (Oxford). 2012;51(9):1550–6. https://doi.org/10.1093/rheumatology/kes120.

    Article  CAS  Google Scholar 

  112. Zhu H, Luo H, Li Y, Zhou Y, Jiang Y, Chai J, et al. MicroRNA-21 in scleroderma fibrosis and its function in TGF-beta-regulated fibrosis-related genes expression. J Clin Immunol. 2013;33(6):1100–9. https://doi.org/10.1007/s10875-013-9896-z.

    Article  CAS  PubMed  Google Scholar 

  113. Pottier N, Maurin T, Chevalier B, Puissegur MP, Lebrigand K, Robbe-Sermesant K, et al. Identification of keratinocyte growth factor as a target of microRNA-155 in lung fibroblasts: implication in epithelial-mesenchymal interactions. PLoS One. 2009;4(8):e6718. https://doi.org/10.1371/journal.pone.0006718.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Honda N, Jinnin M, Kajihara I, Makino T, Makino K, Masuguchi S, et al. TGF-beta-mediated downregulation of microRNA-196a contributes to the constitutive upregulated type I collagen expression in scleroderma dermal fibroblasts. J Immunol. 2012;188(7):3323–31. https://doi.org/10.4049/jimmunol.1100876.

    Article  CAS  PubMed  Google Scholar 

  115. Yin H, Zhao M, Wu X, Gao F, Luo Y, Ma L, et al. Hypomethylation and overexpression of CD70 (TNFSF7) in CD4+ T cells of patients with primary Sjogren's syndrome. J Dermatol Sci. 2010;59(3):198–203. https://doi.org/10.1016/j.jdermsci.2010.06.011.

    Article  CAS  PubMed  Google Scholar 

  116. Yu X, Liang G, Yin H, Ngalamika O, Li F, Zhao M, et al. DNA hypermethylation leads to lower FOXP3 expression in CD4+ T cells of patients with primary Sjogren's syndrome. Clin Immunol. 2013;148(2):254–7. https://doi.org/10.1016/j.clim.2013.05.005.

    Article  CAS  PubMed  Google Scholar 

  117. Shen L, Suresh L, Wu J, Xuan J, Li H, Zhang C, et al. A role for lymphotoxin in primary Sjogren’s disease. J Immunol. 2010;185(10):6355–63. https://doi.org/10.4049/jimmunol.1001520.

    Article  CAS  PubMed  Google Scholar 

  118. Imgenberg-Kreuz J, Sandling JK, Almlof JC, Nordlund J, Signer L, Norheim KB, et al. Genome-wide DNA methylation analysis in multiple tissues in primary Sjogren's syndrome reveals regulatory effects at interferon-induced genes. Ann Rheum Dis. 2016;75(11):2029–36. https://doi.org/10.1136/annrheumdis-2015-208659.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Thabet Y, Le Dantec C, Ghedira I, Devauchelle V, Cornec D, Pers JO, et al. Epigenetic dysregulation in salivary glands from patients with primary Sjogren's syndrome may be ascribed to infiltrating B cells. J Autoimmun. 2013;41:175–81. https://doi.org/10.1016/j.jaut.2013.02.002.

    Article  CAS  PubMed  Google Scholar 

  120. Konsta OD, Le Dantec C, Charras A, Cornec D, Kapsogeorgou EK, Tzioufas AG, et al. Defective DNA methylation in salivary gland epithelial acini from patients with Sjogren’s syndrome is associated with SSB gene expression, anti-SSB/LA detection, and lymphocyte infiltration. J Autoimmun. 2016;68:30–8. https://doi.org/10.1016/j.jaut.2015.12.002.

    Article  CAS  PubMed  Google Scholar 

  121. Konsta OD, Charras A, Le Dantec C, Kapsogeorgeou E, Bordron A, Brooks WH, et al. Epigenetic modifications in salivary glands from patients with Sjogren’s syndrome affect cytokeratin 19 expression. Bull Group Int Rech Sci Stomatol Odontol. 2016;53(1):e01.

    CAS  PubMed  Google Scholar 

  122. Pauley KM, Stewart CM, Gauna AE, Dupre LC, Kuklani R, Chan AL, et al. Altered miR-146a expression in Sjogren's syndrome and its functional role in innate immunity. Eur J Immunol. 2011;41(7):2029–39. https://doi.org/10.1002/eji.201040757.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Huang W, Connor E, Rosa TD, Muir A, Schatz D, Silverstein J, et al. Although DR3-DQB1*0201 may be associated with multiple component diseases of the autoimmune polyglandular syndromes, the human leukocyte antigen DR4-DQB1*0302 haplotype is implicated only in beta-cell autoimmunity. J Clin Endocrinol Metab. 1996;81(7):2559–63. https://doi.org/10.1210/jcem.81.7.8675578.

    Article  CAS  PubMed  Google Scholar 

  124. Mastrandrea LD. An overview of organ-specific autoimmune diseases including immunotherapy. Immunol Investig. 2015;44(8):803–16. https://doi.org/10.3109/08820139.2015.1099409.

    Article  CAS  Google Scholar 

  125. Lesage S, Goodnow CC. Organ-specific autoimmune disease: a deficiency of tolerogenic stimulation. J Exp Med. 2001;194(5):F31–6. https://doi.org/10.1084/jem.194.5.F31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Wang YI, McDuffie M, Lafferty KJ. Antigen presentation and effector mechanisms in ‘auto-immune’ diabetes. J Gastroenterol Hepatol. 1991;6(6):574–9. https://doi.org/10.1111/j.1440-1746.1991.tb00913.x.

    Article  CAS  PubMed  Google Scholar 

  127. Balazs C. Hashimoto's thyroiditis, the model of organ-specific autoimmune disorders. Orv Hetil. 2007;148(Suppl 1):31–3. https://doi.org/10.1556/OH.2007.28032.

    Article  PubMed  Google Scholar 

  128. Greer JM, McCombe PA. The role of epigenetic mechanisms and processes in autoimmune disorders. Biologics. 2012;6:307–27. https://doi.org/10.2147/BTT.S24067.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Ciechomska M, O'Reilly S. Epigenetic modulation as a therapeutic Prospect for treatment of autoimmune rheumatic diseases. Mediat Inflamm. 2016;2016:9607946. https://doi.org/10.1155/2016/9607946.

    Article  CAS  Google Scholar 

  130. Lee HJ, Li CW, Hammerstad SS, Stefan M, Tomer Y. Immunogenetics of autoimmune thyroid diseases: a comprehensive review. J Autoimmun. 2015;64:82–90. https://doi.org/10.1016/j.jaut.2015.07.009.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. McLachlan SM, Rapoport B. Breaking tolerance to thyroid antigens: changing concepts in thyroid autoimmunity. Endocr Rev. 2014;35(1):59–105. https://doi.org/10.1210/er.2013-1055.

    Article  CAS  PubMed  Google Scholar 

  132. Wang B, Shao X, Song R, Xu D, Zhang JA. The emerging role of epigenetics in autoimmune thyroid diseases. Front Immunol. 2017;8:396. https://doi.org/10.3389/fimmu.2017.00396.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Cai TT, Muhali FS, Song RH, Qin Q, Wang X, Shi LF, et al. Genome-wide DNA methylation analysis in graves’ disease. Genomics. 2015;105(4):204–10. https://doi.org/10.1016/j.ygeno.2015.01.001.

    Article  CAS  PubMed  Google Scholar 

  134. Limbach M, Saare M, Tserel L, Kisand K, Eglit T, Sauer S, et al. Epigenetic profiling in CD4+ and CD8+ T cells from graves’ disease patients reveals changes in genes associated with T cell receptor signaling. J Autoimmun. 2016;67:46–56. https://doi.org/10.1016/j.jaut.2015.09.006.

    Article  CAS  PubMed  Google Scholar 

  135. Arakawa Y, Watanabe M, Inoue N, Sarumaru M, Hidaka Y, Iwatani Y. Association of polymorphisms in DNMT1, DNMT3A, DNMT3B, MTHFR and MTRR genes with global DNA methylation levels and prognosis of autoimmune thyroid disease. Clin Exp Immunol. 2012;170(2):194–201. https://doi.org/10.1111/j.1365-2249.2012.04646.x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Kolarz B, Majdan M. Epigenetic aspects of rheumatoid arthritis: contribution of non-coding RNAs. Semin Arthritis Rheum. 2017;46(6):724–31. https://doi.org/10.1016/j.semarthrit.2017.01.003.

    Article  CAS  PubMed  Google Scholar 

  137. Sarumaru M, Watanabe M, Inoue N, Hisamoto Y, Morita E, Arakawa Y, et al. Association between functional SIRT1 polymorphisms and the clinical characteristics of patients with autoimmune thyroid disease. Autoimmunity. 2016;49(5):329–37. https://doi.org/10.3109/08916934.2015.1134506.

    Article  CAS  PubMed  Google Scholar 

  138. Bernecker C, Lenz L, Ostapczuk MS, Schinner S, Willenberg H, Ehlers M, et al. MicroRNAs miR-146a1, miR-155_2, and miR-200a1 are regulated in autoimmune thyroid diseases. Thyroid. 2012;22(12):1294–5. https://doi.org/10.1089/thy.2012.0277.

    Article  CAS  PubMed  Google Scholar 

  139. Peng H, Liu Y, Tian J, Ma J, Tang X, Yang J, et al. Decreased expression of microRNA-125a-3p upregulates interleukin-23 receptor in patients with Hashimoto's thyroiditis. Immunol Res. 2015;62(2):129–36. https://doi.org/10.1007/s12026-015-8643-3.

    Article  CAS  PubMed  Google Scholar 

  140. Hashemi M, Bahari G, Naderi M, Sadeghi-Bojd S, Taheri M. Pri-miR-34b/c rs4938723 polymorphism is associated with the risk of childhood acute lymphoblastic leukemia. Cancer Genet. 2016;209(11):493–6. https://doi.org/10.1016/j.cancergen.2016.09.009.

    Article  CAS  PubMed  Google Scholar 

  141. Cai T, Li J, An X, Yan N, Li D, Jiang Y, et al. Polymorphisms in MIR499A and MIR125A gene are associated with autoimmune thyroid diseases. Mol Cell Endocrinol. 2017;440:106–15. https://doi.org/10.1016/j.mce.2016.11.017.

    Article  CAS  PubMed  Google Scholar 

  142. Wang Y, Bi Y, Chen X, Li C, Li Y, Zhang Z, et al. Histone deacetylase SIRT1 negatively regulates the differentiation of Interleukin-9-producing CD4(+) T cells. Immunity. 2016;44(6):1337–49. https://doi.org/10.1016/j.immuni.2016.05.009.

    Article  CAS  PubMed  Google Scholar 

  143. Zhang J, Lee SM, Shannon S, Gao B, Chen W, Chen A, et al. The type III histone deacetylase Sirt1 is essential for maintenance of T cell tolerance in mice. J Clin Invest. 2009;119(10):3048–58. https://doi.org/10.1172/JCI38902.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Yang H, Lee SM, Gao B, Zhang J, Fang D. Histone deacetylase sirtuin 1 deacetylates IRF1 protein and programs dendritic cells to control Th17 protein differentiation during autoimmune inflammation. J Biol Chem. 2013;288(52):37256–66. https://doi.org/10.1074/jbc.M113.527531.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Brooks WH, Renaudineau Y. Epigenetics and autoimmune diseases: the X chromosome-nucleolus nexus. Front Genet. 2015;6:22. https://doi.org/10.3389/fgene.2015.00022.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Invernizzi P, Pasini S, Selmi C, Miozzo M, Podda M. Skewing of X chromosome inactivation in autoimmunity. Autoimmunity. 2008;41(4):272–7. https://doi.org/10.1080/08916930802024574.

    Article  CAS  PubMed  Google Scholar 

  147. Brix TH, Knudsen GP, Kristiansen M, Kyvik KO, Orstavik KH, Hegedus L. High frequency of skewed X-chromosome inactivation in females with autoimmune thyroid disease: a possible explanation for the female predisposition to thyroid autoimmunity. J Clin Endocrinol Metab. 2005;90(11):5949–53. https://doi.org/10.1210/jc.2005-1366.

    Article  CAS  PubMed  Google Scholar 

  148. Ishido N, Inoue N, Watanabe M, Hidaka Y, Iwatani Y. The relationship between skewed X chromosome inactivation and the prognosis of Graves’ and Hashimoto's diseases. Thyroid. 2015;25(2):256–61. https://doi.org/10.1089/thy.2014.0318.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Fodor A, Cozma A, Karnieli E. TBC update: personalized epigenetic management of diabetes. Per Med. 2017;14(6):531–49. https://doi.org/10.2217/pme-2017-0043.

    Article  CAS  PubMed  Google Scholar 

  150. Olsson AH, Volkov P, Bacos K, Dayeh T, Hall E, Nilsson EA, et al. Genome-wide associations between genetic and epigenetic variation influence mRNA expression and insulin secretion in human pancreatic islets. PLoS Genet. 2014;10(11):e1004735. https://doi.org/10.1371/journal.pgen.1004735.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Rui J, Deng S, Lebastchi J, Clark PL, Usmani-Brown S, Herold KC. Methylation of insulin DNA in response to proinflammatory cytokines during the progression of autoimmune diabetes in NOD mice. Diabetologia. 2016;59(5):1021–9. https://doi.org/10.1007/s00125-016-3897-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Alejandro EU, Bozadjieva N, Blandino-Rosano M, Wasan MA, Elghazi L, Vadrevu S, et al. Overexpression of kinase-dead mTOR impairs glucose homeostasis by regulating insulin secretion and not beta-cell mass. Diabetes. 2017;66(8):2150–62. https://doi.org/10.2337/db16-1349.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Bacos K, Gillberg L, Volkov P, Olsson AH, Hansen T, Pedersen O, et al. Blood-based biomarkers of age-associated epigenetic changes in human islets associate with insulin secretion and diabetes. Nat Commun. 2016;7:11089. https://doi.org/10.1038/ncomms11089.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Brasacchio D, Okabe J, Tikellis C, Balcerczyk A, George P, Baker EK, et al. Hyperglycemia induces a dynamic cooperativity of histone methylase and demethylase enzymes associated with gene-activating epigenetic marks that coexist on the lysine tail. Diabetes. 2009;58(5):1229–36. https://doi.org/10.2337/db08-1666.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Syreeni A, El-Osta A, Forsblom C, Sandholm N, Parkkonen M, Tarnow L, et al. Genetic examination of SETD7 and SUV39H1/H2 methyltransferases and the risk of diabetes complications in patients with type 1 diabetes. Diabetes. 2011;60(11):3073–80. https://doi.org/10.2337/db11-0073.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Li Y, Zhao M, Hou C, Liang G, Yang L, Tan Y, et al. Abnormal DNA methylation in CD4+ T cells from people with latent autoimmune diabetes in adults. Diabetes Res Clin Pract. 2011;94(2):242–8. https://doi.org/10.1016/j.diabres.2011.07.027.

    Article  CAS  PubMed  Google Scholar 

  157. Wang Z, Zheng Y, Hou C, Yang L, Li X, Lin J, et al. DNA methylation impairs TLR9 induced Foxp3 expression by attenuating IRF-7 binding activity in fulminant type 1 diabetes. J Autoimmun. 2013;41:50–9. https://doi.org/10.1016/j.jaut.2013.01.009.

    Article  CAS  PubMed  Google Scholar 

  158. Miao F, Smith DD, Zhang L, Min A, Feng W, Natarajan R. Lymphocytes from patients with type 1 diabetes display a distinct profile of chromatin histone H3 lysine 9 dimethylation: an epigenetic study in diabetes. Diabetes. 2008;57(12):3189–98. https://doi.org/10.2337/db08-0645.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Patel T, Patel V, Singh R, Jayaraman S. Chromatin remodeling resets the immune system to protect against autoimmune diabetes in mice. Immunol Cell Biol. 2011;89(5):640–9. https://doi.org/10.1038/icb.2010.144.

    Article  CAS  PubMed  Google Scholar 

  160. Hezova R, Slaby O, Faltejskova P, Mikulkova Z, Buresova I, Raja KR, et al. microRNA-342, microRNA-191 and microRNA-510 are differentially expressed in T regulatory cells of type 1 diabetic patients. Cell Immunol. 2010;260(2):70–4. https://doi.org/10.1016/j.cellimm.2009.10.012.

    Article  CAS  PubMed  Google Scholar 

  161. Sippl W, Jung M, Mannhold R, Buschmann H, Holenz J. Epigenetic drug discovery: Wiley; 2019. https://www.wiley-vch.de/en/areas-interest/natural-sciences/epigenetic-drug-discovery-978-3-527-34314-0.

  162. Farh KK, Marson A, Zhu J, Kleinewietfeld M, Housley WJ, Beik S, et al. Genetic and epigenetic fine mapping of causal autoimmune disease variants. Nature. 2015;518(7539):337–43. https://doi.org/10.1038/nature13835.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgments

Not applicable.

Funding

This work was supported by funds from IIT-Sapienza Project (R.M.), the COST action CM1406 Epigenetic Chemical Biology (A.M.), PRIN 2015 (prot. 20152TE5PK) (A.M.), AIRC 2016 (n. 19162) (A.M.), PE-2013-02355271 (A.M.) and NIH (n. R01GM114306) (A.M.). The research for this paper was financially supported by the Italian Ministry of Health and Fondazione Roma.

Availability of data and materials

Not applicable.

Author information

Authors and Affiliations

Authors

Contributions

RM, CZ, and AM researched the literature and drafted the manuscript. ST, MR, MA, AG, and AM critically reviewed and edited the work. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Antonello Mai.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mazzone, R., Zwergel, C., Artico, M. et al. The emerging role of epigenetics in human autoimmune disorders. Clin Epigenet 11, 34 (2019). https://doi.org/10.1186/s13148-019-0632-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13148-019-0632-2

Keywords