Open Access

DNA methylation in endometriosis (Review)

  • Authors:
    • Ourania Koukoura
    • Stavros Sifakis
    • Demetrios A. Spandidos
  • View Affiliations

  • Published online on: February 22, 2016     https://doi.org/10.3892/mmr.2016.4925
  • Pages: 2939-2948
  • Copyright: © Koukoura et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Endometriosis is defined by the presence and growth of functional endometrial tissue, outside the uterine cavity, primarily in the ovaries, pelvic peritoneum and rectovaginal septum. Although it is a benign disease, it presents with malignant characteristics, such as invasion to surrounding tissues, metastasis to distant locations and recurrence following treatment. Accumulating evidence suggests that various epigenetic aberrations may play an essential role in the pathogenesis of endometriosis. Aberrant DNA methylation represents a possible mechanism repsonsible for this disease, linking gene expression alterations observed in endometriosis with hormonal and environmental factors. Several lines of evidence indicate that endometriosis may partially be due to selective epigenetic deregulations influenced by extrinsic factors. Previous studies have shed light into the epigenetic component of endometriosis, reporting variations in the epigenetic patterns of genes known to be involved in the aberrant hormonal, immunologic and inflammatory status of endometriosis. Although recent studies, utilizing advanced molecular techniques, have allowed us to further elucidate the possible association of DNA methylation with altered gene expression, whether these molecular changes represent the cause or merely the consequence of the disease is a question which remains to be answered. This review provides an overview of the current literature on the role of DNA methylation in the pathophysiology and malignant evolution of endometriosis. We also provide insight into the mechanisms through which DNA methylation-modifying agents may be the next step in the research of the pharmaceutical treatment of endometriosis.

1. Introduction

Endometriosis represents a common gynecological disease which affects >10% of women of reproductive age (1). Endometriosis is often characterized as an enigmatic condition, or 'the disease of theories', since it is still a diagnostic and therapeutic challenge despite decades of clinical experience and research (2). Although it is a benign disease, it presents with malignant characteristics, such as invasion to surrounding tissues, metastasis to distant locations and recurrence following treatment. The unique feature of endometriosis of being a benign metastasing disease and the multiple treatment options currently available, clearly indicate how difficult it can be to diagnose and effectively treat endometriosis based on our current understanding of the disease.

Endometriosis is classically defined by the presence and growth of functional endometrial tissue, outside the uterine cavity, primarily in the ovaries, pelvic peritoneum and rectovaginal septum (3). This ectopic endometrial tissue responds to hormones and drugs in a similar manner to the eutopic endometrium. The continued growth of endometriotic tissue, as with that of the endometrium, is dependent upon estrogen. Thus, endometriosis is prevalent in the reproductive years with a peak incidence between 30 and 45 years of age (4). The main pathological processes associated with the disease are peritoneal inflammation and fibrosis, and the formation of adhesions and ovarian cysts (5). Infertility and pelvic pain are the predominant symptoms that greatly affect the quality of life of women with endometriosis (6).

Various theories have been put forth in order to elucidate the possible mechanisms responsible for the development of endometriosis. While the etiology of the disease remains unclear, retrograde menstruation is the most widely accepted mechanism of peritoneal endometriotic implants (7). The other traditionally suggested mechanisms include coelomic metaplasia, induction of the disease through immune system abnormalities, transformation of embryonic rests and vascular and lymphatic spread (3). No single theory however, can explain all the different manifestations of endometriosis. Although the transfer of endometrial cells in the peritoneal cavity is a realistic initiating factor of endometriosis, additional steps are necessary for the survival, implantation and growth of the ectopic endometrium. A defective immune clearance, attachment and invasion to the peritoneal epithelium, the establishment of local neoneurovascularity and an altered hormonal milieu that will stimulate continuous growth, are necessary if endometriosis is to develop from retrograde menstruation (810). Recently, researchers have focused on several well supported hallmarks of the disease, such as genetic predisposition, altered hormonal dependence, inflammation and exposure to environmental toxins (11). The missing link in basic research to the pathophysiology of endometriosis may be a common denominator for a disease that influences its hormonal, immunological and genetic profile. Epigenetics has reformed the understanding of other multifactorial diseases, such as cancer, which shares many common characteristics with endometriosis; thus, epigenetic mechanisms may play a significant role in the origin and progression of endometriosis (12). The epigenome, the collection of DNA methylation and histone modifications can be influenced by environmental factors (13). It is reasonable to speculate that epigenetic alterations represent an attractive candidate, linking intrinsic molecular changes detected in endometriosis, to environmental and lifestyle influences.

2. Genetic and epigenetic basis of endometriosis

The contribution of genetic factors to the susceptibility to endometriosis is supported by a number of different studies (1416). Higher rates of endometriosis are found among the relatives of patients with endometriosis compared with those of control subjects. The number of gene mapping studies for this disease has increased in recent years as the role of genetic factors has become more widely accepted. To date, many deregulated genes have been identified in endometriotic cells with a wide variety of functions, including apoptosis, vascularization, cell cycle regulation, DNA repair, encoding detoxification enzymes, immune system regulation and cell adhesion (1719). Yet despite several publications, few reported data have been replicated by other investigators (20,21). In an earlier review by Falconer et al, the authors concluded that there is a strikingly large amount of conflicting results in the literature and that 'polymorphisms may have a limited value in assessing the possible development of endometriosis' (22).

Genetics and cell science have revealed in detail what cells in our body are composed of, but have not yet succeeded in explaining the diversity in morphology and function that derives from the same genetic material. Explaining the mechanisms that temporally regulate the expression of selected genes in guiding cell differentiation and function, for processes as divergent as placental development and schizophrenia, is the ground work of epigenetics. The word epigenetics is derived from the Greek word 'epi', for over or above, and genetics, or the science of heredity. Epigenetics provide a potential mechanism through which the genome can 'capture' the effects of environmental exposures and perpetuate their influence on cell function. Accumulating evidence on the other hand suggests that various epigenetic aberrations may play an essential role in the pathogenesis of endometriosis (23). Epigenetic alterations refer to stable alterations in gene expression with no underlying modifications in the genetic sequence itself (24). The field of epigenetics has rapidly evolved and has influenced research in different biological phenotypes, such as ageing, memory formation and embryological development (25,26). Epigenetics encompasses several different phenomena, such as DNA methylation, histone modifications, RNA interference and genomic imprinting. Epigenetic processes regulate gene expression and can alter malignancy-associated characteristics, such as growth, migration, invasion, or angiogenesis (27,28). Methylation represents one of the most important epigenetic functions which involves the addition of methyl groups on the cytosine residues of CG (also termed CpG) dinucleotides of DNA (29). Enzymes known as DNA methyltransferases (DNMTs) catalyse the addition of the methyl group to the cytosine ring to form methyl cytosine, using S-adenosylmethionine as a methyl donor. DNMT1 is the predominant mammalian DNA methylating enzyme responsible for the restoration of hemi-methylated sites to full methylation (maintenance methylation) which occurs after DNA replication (30). DNMT3A and DNMT3B are mainly involved in the methylation of new sites, known as de novo methylation (31). In humans and other mammals, DNA modification occurs predominantly on cytosines that precede a guanosine in the DNA sequence (32). These dinucleotides can be clustered in small stretches of DNA, termed CpG islands, which are often associated with promoter regions. In 98% of the genome, CpGs are present approximately once per 80 dinucleotides. By contrast, CpG islands, which comprise 1–2% of the genome, are approximately 200 base pairs to several kb in length and have a frequency of CpGs approximately 5-fold greater than the genome as a whole (33,34). The majority of CpG sites outside of CpG islands are methylated, suggesting a role in the global maintenance of the genome, while the majority of CpG islands in gene promoters are unmethylated, which allows active gene transcription (32,35). Once a CpG becomes methylated in a cell, it will remain methylated in all its descendants (36). Generally, when a given stretch of cytosines in a CpG island located in the promoter region of a gene is methylated, that gene will be silenced by methylation; such a CpG island would be termed 'hypermethylated'. Conversely, when a given stretch of cytosines in a CpG island located in the promoter region of a gene is not methylated, that gene will not be silenced by methylation; the CpG island in this case would be 'hypomethylated' (37). The methylation of promoters inhibits their recognition by transcription factors and RNA polymerase, as methylated cytosines preferentially bind to a protein known as methyl cytosine binding protein (MeCP). When a promoter region normally recognized by an activating transcription factor, is methylated, its transcription will be inhibited (34).

Aberrant DNA methylation represents a possible mechanism, linking gene expression alterations observed in endometriosis with hormonal and environmental factors (38). It is difficult however, to determine whether aberrant methylation is the cause or the consequence of the disease. Given the theoretical multifactorial origin of endometriosis, a linear association that comprises environmental factors, epigenetic alterations and disease, is difficult to established. Despite this fact, several lines of evidence indicate that endometriosis may partially be due to selective epigenetic deregulations influenced by extrinsic factors (39). DNA methylation is a delicate reversible event that is known to be influenced by environmental factors, such as exposure to xenobiotics, social behavior, metabolism and nutritional deficiencies that may exert their effects later in life, during critical periods of development, or may be transmitted transgenerationally to the offspring (40,41). Endometriosis has been frequently associated with exposure to toxins or synthetic compounds and dietary habits. The epigenome in women with endometriosis may be a reflection of the respective woman's age, reproductive history, body mass index and whether or not she was exposed to chemicals during her lifetime (42,43). Recent theories on fetal programming postulate that chronic adult onset diseases with an epigenetic component, originate in utero when the early embryo is exposed to factors that permanently shape its epigenetic mark (44,45).

Since eutopic and ectopic endometriotic stromal cells share the same genetic background, research has focused on mechanisms that may provoke different cellular responses, namely epigenetics. The silencing of progesterone and aromatase genes, which are essential elements in the development of endometriosis, by promoter hypermethylation, may be a strong contributing factor of the disease. It has been shown that a single endometriotic lesion originates form a single progenitor cell, forming a cellular lineage (46). This monoclonality of endometriotic lesions suggests that they may carry neoplastic potentials. This cellular lineage requires that cells transcribe, or it enables the transcription of specific genes, or regions of genes, whereas at the same time it suppesses others. To maintain cellular identity, the gene expression program must be maintained through cell divisions in a heritable manner through epigenetic processes.

3. Aberrant DNMT expression in endometriosis

It has been previously reported that DNMT1, DNMT3A and DNMT3B are overexpressed in the epithelial component of endometriotic implants as compared to normal controls or the in the eutopic endometrium of women with endometriosis. Moreover, the expression levels of DNMT1, DNMT3A and DNMT3B have been shown to positively correlate with each other (47). The upregulated expression of DNMTs in the endometriotic tissue, which leads to hypermethylation, has been confirmed in women with endometriosis only for the DNMT3A transcript and not for DNMT1 and DNMT3B (48). Conversely, significantly lower expression levels of DNMTs have been found by other studies where they compared endometriotic lesions to the eutopic endometrium of women with endometriosis and disease-free controls (49,50). In a previous study, the induction of hypoxia triggered global hypomethylation in ectopic stromal cells through the destabilization of DNMT1 mRNA, thus providing a plausible link of gene-environment interaction by means of DNA methylation (50). Unaltered DNMT expression in response to in vitro decidualization in endometriotic cells may also elucidate the aberrant epigenetic status that alters gene expression and contributes to the progesterone-resistant environment observed in endometriosis (51).

4. Estrogen and progesterone receptor genes

The uterine endometrium is targeted by the ovarian sex steroid hormones, estrogen and progesterone, which regulate the growth of endometrial tissue, basically by stimulating or inhibiting cell proliferation, respectively (52). Each hormone is estimated to regulate the expression of hundreds of genes during various phases of the menstrual cycle. Endometriotic tissue in ectopic locations, such as the peritoneum or ovaries, differs fundamentally from the eutopic endometrium within the uterus in terms of the production of prostaglandins (PGs) and cytokines, estrogen production and metabolism, as well as the clinical response to progestins. The ectopic endometrium is characterized by an imbalance in the function of estrogen and progesterone, namely estrogen dominance and progesterone resistance (53). Abundant quantities of estrogen are available in the endometriotic tissue via several mechanisms, including local aromatase expression. Estrogen and progesterone exert their functions by binding to their intracellular receptors, the estrogen receptor (ER) and progesterone receptor (PR), which are members of the steroid/nuclear receptor (SR) superfamily (54). The involvement of SRs with co-regulators and other recruited proteins in the transcriptional complex is essential for target gene regulation. The SRs interact with DNA-methylating/-demethylating enzymes in the transcriptional complex, and as a final point, histone modification along with DNA methylation both regulate gene expression (55,56). Thus, steroid hormone responsive tissues, such as endometriotic lesions, have epigenetic constituents which may be vulnerable to modifications.

The biologically active estrogen, estradiol (E2), enters cells and binds to the ER in both the eutopic and ectopic endometrial cells. There are two separate ER subtypes, ERα and ERβ, that appear to have overlapping, although different, tissue expression and localization profiles (57). ERα and ERβ are encoded by separate genes, ESR1 and ESR2, respectively, found at different chromosomal locations (58). The E2-receptor complex acts as a transcription factor that becomes associated with the promoters of E2-responsive genes via direct DNA binding or binding to other docking transcription factors, such as activator protein complex 1 (APC-1). Of note, while ERα and ERβ recognize the same estrogen responsive element, the two subtypes display different transactivational properties in a ligand-dependent manner when they are co-expressed (21). In addition, ERβ also has the capacity to regulate ERα (53). Recent studies using ERα, ERβ null mice with surgically induced endometriosis revealed that only ERα inhibits endometrial growth and leads to a decrease in estrogen target gene expression, whereas the deletion of ERβ does not affect the biological responses of the uterus to estrogen (59). Burns et al concluded that estrogen-regulated signaling responses are predominantly mediated by ERα in endometriosis-like lesions (60).

Progesterone has long been used for relieving endometriosis-induced pain, mainly by inducing pseudo-pregnancy, thus suppressing ovarian estrogen production, which in turn suppresses growth and inflammation in endometriosis. The uterine response to progesterone is dependent on PRs. The two predominant isoforms of PR, PRA and PRB, are both encoded by the same PR gene, but use alternative promoters and translation start sites (61).

The majority of altered expression endometriosis-associated genes, are downstream targets of ERs and PRs or overlap with genes known to be regulated by the SRs (62). The ER and PR levels differ markedly in the endometrium compared to endometriosis-derived stromal cells (63). Endometriotic lesions exhibit particularly higher ERβ and significantly lower ERα and PR levels compared to the eutopic endometrium (64,65). The suppressed ERα expression detected in the stromal cells of endometriosis may be a consequence of the strikingly high quantities of estradiol in addition to high ERβ levels produced via local aromatase activity (62). Although ERα seems to be the primary mediator of the estrogenic action, elevated ERβ levels and in particular increased ERβ/ERα ratio in endometriosis compared to that in endometrial tissues is associated with suppressed PR levels, contributing to the loss of progesterone signaling or progesterone resistance noted in endometriosis (53).

Xue et al identified a CpG island occupying the promoter region of the ERβ gene, which exhibited significantly higher methylation levels in endometrial cells versus endometriotic cells (66). Moreover, the activity of the ERβ promoter bearing the CpG island was strongly inactivated by in vitro methylation. The authors of that study concluded that the high ERβ mRNA and protein expression observed in endometriosis is mediated by an epigenetic defect involving the hypomethylation of the gene's promoter. Another piece of evidence linking methylation to SR function is the high methylation levels of the PRβ promoter demonstrated in endometriosis (67). The hypermethylation of the PRβ promoter is in accordance to the already reported downregulation of PRβ in endometriosis, which in turn presents a plausible explanation to progesterone resistance in endometriotic tissues. PRα on the other hand, does not display similar epigenetic patterns. In contrast with the above observations, there are certain studies that report higher levels of PRβ in endometriotic tissues (68,69). For instance, ovarian endometrioma samples have been shown to have significantly higher levels of PRβ mRNA when compared with the eutopic endometrium (68). The diversity of endometriosis is also prominent in studies, whereas different types of endometriosis exhibit diverse DNA methylation patterns of the SR genes. Intestinal endometriosis which is one of the most aggressive forms of the disease has demonstrated no differences in the DNA methylation patterns of the ESR1 and ESR2 genes compared to the eutopic endometrium obtained from the same patient. The methylation of the PGR gene was observed exclusively in a subset of the endometriotic samples which contributed to PGR gene suppression, which in turn was further confirmed by immunostaining of the PGR protein in the same samples (70).

5. Steroidogenic factor 1 (SF-1)

SF-1, also known as Ad4BP or NR5A1, is a member of the nuclear receptor superfamily and is encoded by the NR5A1 gene in humans (71,72). SF-1 is a key transcription factor for steroid biosynthesis and is responsible for inducing the expression of steroidogenic acute regulatory protein (STAR) and cytochrome P450, family 19, subfamily A, polypeptide 1 (CYP19a1), which encode aromatase (73). Aromatase catalyses the final step of estrogen production through thye conversion of C19 steroids to estrogens. SF-1 is also involved in the regulation of other adrenal and testicular steroidogenic genes, such as hydroxysteroid dehydrogenase genes (HSD3B and HSD11B) and melanocortin 2 receptor (adrenocorticotropic hormone) (MC2R). The aberrant expression of steroidogenesis-related genes represents a possible pathogenetic mechanism of endometriosis, wherein estradiol synthesis is locally enhanced within endometriotic cells (74). It has already been reported that SF-1 is highly elevated in endometriotic tissues compared to the normal endometrium. Consistent with the mRNA levels of SF-1, the protein levels of SF-1 are also significantly elevated in the ectopic endometrium and this elevation corresponds to the severity of endometriosis (75).

In an earlier study, Xue et al reported that normal endometrial cells in which SF-1 transcriptional activity was completely suppressed, demonstrated aberrant methylation of the promoter and exon 1 region of the NR5A1 gene (76). On the contrary, endometriotic cells displayed higher SF-1 mRNA and protein levels along with reduced methylation levels at the SF-1 promoter region. Yamagata et al confirmed these results in a genome-wide methylation analysis, in cultured eutopic and ectopic cells. In NR5A1 and STAR, the CpG sites were hypomethylated in cultured cells from endometriotic cysts compared with those from eutopic endometrium (77). Taken together, these two studies have shown that methylation of the proximal promoter of the NR5A1 gene regulates SF-1 expression in endometriotic tissues, as well as in the normal endometrium. A few years later, however, Xue et al reported that hypermethylation of the CpG island that spans from exon 2 to intron 3 of the SF-1 gene activated mRNA expression in endometriotic cells (78). The authors of that study hypothesized that the hypermethylation of this particular region of the gene, distant to the promoter, encloses a silencer which, when hypermethylated, suppresses its silencer function, giving rise to increased SF-1 expression. Furthermore, in a similar study by the same team, the hypermethylation of a novel CpG island located downstream of intron 1 of the SF-1 gene was associated with a high expression of the gene (79). Although this observation is contradicted to the classical association of methylation to gene expression, it is consistent with a large body of literature, indicating that methylation outside of gene promoters leads to increased gene expression (8083).

6. Homeobox A10 (HOXA10)

HOXA10 is a member of a family of homebox genes that serve as transcription factors which are expressed in the endometrium, where they are necessary for endometrial growth, differentiation and implantation (84). Its expression is regulated by estrogen and progesterone and markedly increases during the midsecretory phase which corresponds to the implantation window (85). Therefore, HOXA10 is considered essential in regulating endometrial development during menstrual cycle, thus facilitating conditions necessary for implantation. Increased HOXA10 levels remain elevated when successful implantation occurs, expressed by the developing decidua in early pregnancy. Both estrogen and progesterone individually stimulate the endometrial expression of HOXA10, and progesterone has additional stimulatory effects over estrogen (86). The first indication of the possible role of HOXA10 in endometriosis was postulated when a difference in the expression of the gene was noted in the endometrium of women with endometriosis (87). It has been reported that in patients with endometriosis there is a decrease in HOXA10 expression during the secretory phase, resulting in decreased uterine receptivity and subsequent endometriosis-related infertility (88). Hypermethylation of the HOXA10 gene promoter provides a probable explanation for its reduced gene expression in the endometrium of women with endometriosis (89). The simultaneous occurrence of HOXA10 promoter hypermethylation and reduced HOXA10 expression has been demonstrated in induced endometriosis in baboons and in mice (90,91). It has been confirmed that DNA hypermethylation may be one of the potential molecular mechanisms silencing HOXA10 expression in the mid-luteal endometrium associated with infertility in women with endometriosis (88). In women with ovarian endometriomas, significantly higher HOXA10 promoter methylation levels have been documented during the mid-luteal phase (92). Treatment of endometrial stromal cells from fertile women with endometriosis with 5-azacytidine, a demethylation agent, resulted in increased HOXA10 mRNA and protein levels, thus suggesting the regulatory role of methylation on gene expression in endometriosis (93). In another study, mice, prenatally exposed to diethylstilbestrol, exhibited the overexpression of both Dnmt1 and Dnmt3, along with Hoxa10 hypermethylation (94). Even though hypermethylation of the HOXA10 promoter is a constant finding in different studies on endometriosis, in a recent genome-wide methylation analysis, the epigenetic alteration of the HOXA10 gene was below the arbitrary threshold set by the authors, hence other epigenetically altered genes were considered more relevant to the pathophysiology of the disease (95).

7. Aromatase

Aromatase is the key enzyme in estrogen production which converts androgen to estrogen (96). Estrogen production in women with endometriosis is accomplished by de novo synthesis in the ovaries, by the conversion of circulating androstenedione to estradiol in adipose tissue, skin and skeletal muscle and lastly by a unique de novo system of local estrogen production which takes place in endometriotic lesions (97). Endometriotic stromal cell express the full complement of genes in the steroidogenic cascade, which is sufficient to convert cholesterol to estradiol. Aberrantly expressed aromatase in the endometriotic implants, is thought to be one of the major contributing factors in the development of the hyperestrogenic microenviroment of endometriosis (53). Aromatase expression has been reported to be absent in the eutopic endometrium of healthy women, whereas its mRNA levels are significantly increased in women with endometriosis (98). Moreover, aromatase expression is also increased in the ectopic endometrium of women with endometriosis.

Aromatase is encoded by a single gene, CYP19, which is located on chromosome 15q21. The aromatase gene has the unique characteristic of having multiple exons available for use which are flanked with unique promoters (99). The tissue-specific expression of aromatase is regulated by the alternative use of these exons. It has been shown that endome-triotic cells use the same aromatase promoters, promoters II, I.3 and I.6, as endometrial cells (100). Since both the ectopic and eutopic endometrium share the same promoters, different gene expression relies on an epigenetic regulatory mechanism which switches off and on aromatase gene in healthy and diseased tissues, respectively. Izawa et al confirmed the above speculation by demonstrating a CpG island at approximately 20 kb upstream from the end of exon II which was hypomethylated in endometriotic and hypermethylated in endometrial cells (101).

8. Other genes

Cyclooxygenase-2 (COX-2) is the key enzyme in the conversion of arachidonic acid to PGs and has been mainly associated with the inflammatory response (102). The aberrant expression of COX-2, and thus the over-production of PGE2 has been shown to play critical roles in the development of endometriosis. The peritoneal microenvironment in the setting of endometriosis is notably rich in PGs, and these mediators likely play a central role in disease pathophysiology, as well as in the clinical sequelae of pain and infertility. COXA-2 overexpression has been observed in ectopic endometriotic lesions and has been correlated with the severity of endometriosis-associated pain and also the recurrence of the disease after surgery (103105). The hypo-methylation of the NF-IL6 site within the COX-2 promoter in the eutopic and ectopic endometrial tissues of women with endometriosis has been linked with the already reported increased COX-2 mRNA levels in the same tissues (106,107).

The E-cadherin (CDH1) gene encodes an epithelial cell-cell adhesion glycoprotein that modulates a wide variety of processes, including cell polarization, migration and cancer metastasis. The decreased expression of CDH1 in epithelial cells in peritoneal endometriosis has been reported in the advanced stages of endometriotic lesions (108). In two endometriotic cell lines, the E-cadherin gene has been found to be hypermethylated at the promoter region, and treatment with a histone deacetylase inhibitor, trichostatin A, induced expression (109).

Syncytin-1 plays a critical role in the maintenance of normal pregnancy. The hypomethylation and activation of the syncytin-1 gene has been found in placental trophoblast lineages and malignant cells. While the syncytin-1 gene is absent in the eutopic endometrium from patients with endometriosis, syncytin-1 mRNA and protein levels are detected in endometriotic lesions. The hypomethylation of the gene promoter in the ectopic lesions highlights the epigenetic regulation of the function of this gene in endometriosis (110).

9. Malignant transformation of endometriosis

The malignant transformation ofendometriosis is believed to occur in approximately 1% of all cases (111). The most common site of malignant transformation of endometriosis is the ovaries. Ovarian endometrioid cancer (OEC) and ovarian clear cell cancer (OCCC) account for 76% of all endometriosis-associated ovarian cancer cases (112). The malignant transformation of endometriosis represents a newly launched, attractive research field in epigenetics. Endometriosis as a possible initiating factor in ovarian carcinogenesis, has gained attention since the recently proposed theory on ovarian cancer. Based on this theory, ovarian cancer originates from tissue implantation outside the ovaries. Since ovarian serous adenocarcinoma is thought to originate from tubal fimbria epithelial lesions, OEC and OCCC are likely to originate from endometriotic implants (113). The demethylation of known oncogenes and also epigenetic silencing through the hypermethylation of tumour suppressor genes might induce mechanisms such as uncontrolled cell division, the ability to infiltrate surrounding tissues, avoiding apoptosis or sustaining angiogenesis (114). The hypermethylation of the hMLH1 gene promoter, which represents one of the most important mismatch repair (MMR) genes, and the consequent decrease in gene expression has been associated with the malignant evolution of endometriosis (115). The hypomethylation of long interspersed element-1 (LINE-1) has also been described as an early molecular event involved in malignant progression of endometriosis (116). The runt-related transcription factor 3 gene (RUNX3) has been shown to be a tumour suppressor in a variety of cancers. Guo et al recently reported that the inactivation of the RUNX3 gene by promoter hypermethylation plays a role in the malignant transformation of ovarian endometriosis and is closely related to estrogen metabolism (117). The tumour suppressor gene, Ras-association domain family member 2 (RASSF2), is inactivated by promoter hypermethylation in many types of cancer (118120). A recent study confirmed that the epigenetic inactivation of RASSF2 is associated with the malignant progression of ovarian endome-triosis and that this epigenetic alteration may be an early event in ovarian tumourigenesis (121).

10. Genome-wide methylation in endometriosis

Genome-wide methylation analysis is an emerging technology which may be used to identify novel genes potentially involved in the development and pathogenesis of endometriosis. Analyses of the entire methylome are used to determine the unique epigenetic fingerprint of endometriosis. Recently, Borghese et al reported a whole-genome DNA methylation profiling in >25,000 promoters, using methylated DNA immunoprecipitation with hybridization to promoter microarrays (122). Consistent with the theory of the endometrial origin of endometriosis, the overall methylation profile was highly similar between the endometrium and the endometriotic lesions. Although there was no correlation between promoter methylation and the expression of nearby genes, 35 genes had both methylation and expressional alterations in the lesions. In a more recent study which included 27,578 genes on the methylation array, 120 genes were significantly altered by ≥1.5-fold in the endometrial biopsies of women with endometriosis compared to those from healthy women (95). When comparing methylation profiles of the eutopic endometrium from women with or without endometriosis and ovarian endometrial cysts, only a few genes were differentially methylated in the endometrium, whereas more hypermethylated and hypomethylated CpGs were detected in the endometrial ovarian cysts (77). In a previous study, in 42,248 differentially methylated CpGs that were investigated, 403 genes demonstrated significantly different methylation patterns. A disproportionally large number of transcription factors had different methylation profiles and many of these genes are already known to be involved in the process of decidualization and the pathophysi-ology of endometriosis (123).

11. Therapeutic implications

Unlike DNA mutations or copy number alterations, reversibility is an important characteristic of epigenetic aberrations, since it allows us to employ a number of epigenetic therapies which can potentially reverse the aberrant epigenetic patterns in affected tissues. Enzymes that regulate epigenetic alterations have been targets of research on pharmacological intervention in endometriosis. The target enzymes of epigenetic drugs include DNMTs, histone deacetylases, histone acetyltransferases, histone methyltransferases and histone demethylases (124). In vitro studies have demonstrated promising results with pharmaceutical agents that disrupt the methylation cascade (125). The selective inhibition of the PGE2 receptors, EP2 and EP4, has been shown to decrease the expression of DNMT3a and DNMT3b, but does not modulate the expression of DNMT1 (126). The authors of that study postulated that targeting EP2 and EP4 receptors may emerge as long-term nonsteroidal therapy for the treatment of active endometriotic lesions in women (126). Treatment with a demethylating agent (DMA) has also been shnown to signifi-cantly increase ERβ mRNA levels in endometrial cells and may indicate a possible epigenetic therapeutic target (127). Izawa et al also demonstrated that the treatment of endometrial stromal cells, which normally do not express aromatase, with a DMA, markedly increased the aromatase mRNA expression (101). Thus far, there are two classes of DNA DMAs: nucleoside and non-nucleoside DNMT inhibitors. They both inhibit DNMTs in the S phase of the cell cycle and consequently lower the overall DNA methylation in the target cells (101).

Several practical considerations arise that influence the utilization of DMA in current therapeutic strategies of endometriosis. The administration of these drugs results in a long-lasting demethylating effect, even after a short period of treatment. Since endometriosis is a benign disease, prolonged cytopenia and gastrointestinal system toxicity are considered serious adverse effects in otherwise healthy patients. Moreover, many of the women with endometriosis may need repeated treatments during their reproductive lifespan, therefore safety issues should be the primary concern of future studies.

12. Conclusion

Endometriosis presents a diagnostic and therapeutic enigma for clinicians, but also an emerging field of research on how epigenetics intervene with the pathophysiology and progression of the disease. Previous studies have shed light into the epigenetic component of endometriosis, reporting variations in the epigenetic patterns of genes known to be involved in the aberrant hormonal, immunologic and inflammatory status of endometriosis (Table I and Fig. 1). Although recent studies, utilizing advanced molecular techniques, have allowed us to further elucidate the possible association of DNA methylation with altered gene expression, whether these molecular changes represent the cause or merely the consequence of the disease is a question which remains to be answered. In vitro studies have reported promising results on treatment with epigenetic modifying agents; however, we have a long way to go until we can employ these agents in current medical practice. Till then, endometriosis will be the ideal model on further research, thus being a benign disease with prominent malignant characteristics.

Table I

Genes reported with aberrant methylation in the ectopic or eutopic endometrium.

Table I

Genes reported with aberrant methylation in the ectopic or eutopic endometrium.

GenesFunctionMethylation(Refs.)
ERβEstrogen nuclear receptor. Mediates estrogenic actionHypomethylated in endometriotic cells(50)
ERβEstrogen nuclear receptorHypermethylated in endometrial cells of women with endometriosis(50)
PRβProgesterone nuclear receptor. Mediates progesterone actionHypermethylated in endometriotic cells(51)
ERα, ERβNo difference in methylation levels of intestinal endometriosis compared to eutopic endometrium(54)
SF-1Key transcription factor for steroid biosynthesisHypomethylated in endometriotic cells(58,59)
SF-1Hypermethylated in endometriotic cells(60,61)
HOXA10Transcription factors necessary for endometrial growth, differentiation, and implantationHypermethylated in eutopic endometrium in women with endometriosis(7073)
AromataseKey enzyme in estrogen production which converts androgen to estrogenHypomethylated in endometriotic tissues and hypermethylated in eutopic endometrium of women with endometriosis(80)
COX-2Key enzyme in the conversion of arachidonic acid to prostaglandinsHypomethylated in both endometriotic and endometrial cells(85,86)
E-CadherinEncodes an epithelial cell-cell adhesion glycoprotein that modulates cell polarization, migration and cancer metastasisHypermethylated in cultured endometriotic cells(88)
Syncytin-1Human endogenous retroviral envelope gene (HERVW1) product is expressed in placental trophoblasts and mediates the formation of syncytiotrophoblastsHypomethylated in endometriotic lesions(89)

[i] SF-1, steroidogenic factor 1; HOXA10, homeobox A10; COX-2, cyclooxygenase-2.

References

1 

Viganò P, Parazzini F, Somigliana E and Vercellini P: Endometriosis: Epidemiology and aetiological factors. Best Pract Res Clin Obstet Gynaecol. 18:177–200. 2004. View Article : Google Scholar : PubMed/NCBI

2 

Eskenazi B and Warner ML: Epidemiology of endometriosis. Obstet Gynecol Clin North Am. 24:235–258. 1997. View Article : Google Scholar : PubMed/NCBI

3 

Seli E, Berkkanoglu M and Arici A: Pathogenesis of endometriosis. Obstet Gynecol Clin North Am. 30:41–61. 2003. View Article : Google Scholar : PubMed/NCBI

4 

Houston DE: Evidence for the risk of pelvic endometriosis by age, race and socioeconomic status. Epidemiol Rev. 6:167–191. 1984.PubMed/NCBI

5 

Kobayashi H, Higashiura Y, Shigetomi H and Kajihara H: Pathogenesis of endometriosis: The role of initial infection and subsequent sterile inflammation (Review). Mol Med Rep. 9:9–15. 2014.

6 

Hickey M, Ballard K and Farquhar C: Endometriosis. BMJ. 348:g17522014. View Article : Google Scholar : PubMed/NCBI

7 

Sampson J: Peritoneal endometriosis due to the menstrual dissemination of endometrial tissue into the peritoneal cavity. Obstet Gynecol. 14:422–469. 1927.

8 

Ahn SH, Monsanto SP, Miller C, Singh SS, Thomas R and Tayade C: Pathophysiology and Immune Dysfunction in Endometriosis. BioMed Res Int. 2015:7959762015. View Article : Google Scholar : PubMed/NCBI

9 

Nisolle M, Casanas-Roux F, Anaf V, Mine JM and Donnez J: Morphometric study of the stromal vascularization in peritoneal endometriosis. Fertil Steril. 59:681–684. 1993.PubMed/NCBI

10 

Burney RO and Giudice LC: Pathogenesis and pathophysiology of endometriosis. Fertil Steril. 98:511–519. 2012. View Article : Google Scholar : PubMed/NCBI

11 

Sanchez AM, Viganò P, Somigliana E, Cioffi R, Panina-Bordignon P and Candiani M: The endometriotic tissue lining the internal surface of endometrioma: hormonal, genetic, epigenetic status, and gene expression profile. Reprod Sci. 22:391–401. 2015. View Article : Google Scholar

12 

Lopez J, Percharde M, Coley HM, Webb A and Crook T: The context and potential of epigenetics in oncology. Br J Cancer. 100:571–577. 2009. View Article : Google Scholar : PubMed/NCBI

13 

Jaenisch R and Bird A: Epigenetic regulation of gene expression: How the genome integrates intrinsic and environmental signals. Nat Genet. 33(Suppl): 245–254. 2003. View Article : Google Scholar : PubMed/NCBI

14 

Bellelis P, Barbeiro DF, Rizzo LV, Baracat EC, Abrão MS and Podgaec S: Transcriptional changes in the expression of chemokines related to natural killer and T-regulatory cells in patients with deep infiltrative endometriosis. Fertil Steril. 99:1987–1993. 2013. View Article : Google Scholar : PubMed/NCBI

15 

Hansen KA and Eyster KM: Genetics and genomics of endometriosis. Clin Obstet Gynecol. 53:403–412. 2010. View Article : Google Scholar : PubMed/NCBI

16 

Baranov VS, Ivaschenko TE, Liehr T and Yarmolinskaya MI: Systems genetics view of endometriosis: A common complex disorder. Eur J Obstet Gynecol Reprod Biol. 185:59–65. 2015. View Article : Google Scholar

17 

Campbell IG and Thomas EJ: Endometriosis: Candidate genes. Hum Reprod Update. 7:15–20. 2001. View Article : Google Scholar : PubMed/NCBI

18 

Vigano P, Somigliana E, Vignali M, Busacca M and Blasio AM: Genetics of endometriosis: Current status and prospects. Front Biosci. 12:3247–3255. 2007. View Article : Google Scholar : PubMed/NCBI

19 

Augoulea A, Alexandrou A, Creatsa M, Vrachnis N and Lambrinoudaki I: Pathogenesis of endometriosis: The role of genetics, inflammation and oxidative stress. Arch Gynecol Obstet. 286:99–103. 2012. View Article : Google Scholar : PubMed/NCBI

20 

Altmüller J, Palmer LJ, Fischer G, Scherb H and Wjst M: Genomewide scans of complex human diseases: True linkage is hard to find. Am J Hum Genet. 69:936–950. 2001. View Article : Google Scholar : PubMed/NCBI

21 

Chanock SJ, Manolio T, Boehnke M, Boerwinkle E, Hunter DJ, Thomas G, Hirschhorn JN, Abecasis G, Altshuler D, Bailey-Wilson JE, et al: NCI-NHGRI Working Group on Replication in Association Studies: Replicating genotype-phenotype associations. Nature. 447:655–660. 2007. View Article : Google Scholar : PubMed/NCBI

22 

Falconer H, D'Hooghe T and Fried G: Endometriosis and genetic polymorphisms. Obstet Gynecol Surv. 62:616–628. 2007. View Article : Google Scholar : PubMed/NCBI

23 

Guo SW: Epigenetics of endometriosis. Mol Hum Reprod. 15:587–607. 2009. View Article : Google Scholar : PubMed/NCBI

24 

Inbar-Feigenberg M, Choufani S, Butcher DT, Roifman M and Weksberg R: Basic concepts of epigenetics. Fertil Steril. 99:607–615. 2013. View Article : Google Scholar : PubMed/NCBI

25 

Xu X: DNA methylation and cognitive aging. Oncotarget. 6:13922–13932. 2015. View Article : Google Scholar : PubMed/NCBI

26 

Breiling A and Lyko F: Epigenetic regulatory functions of DNA modifications: 5-methylcytosine and beyond. Epigenetics Chromatin. 8:242015. View Article : Google Scholar : PubMed/NCBI

27 

Burggren WW and Crews D: Epigenetics in comparative biology: why we should pay attention. Integr Comp Biol. 54:7–20. 2014. View Article : Google Scholar : PubMed/NCBI

28 

Koerner MV and Barlow DP: Genomic imprinting-an epigenetic gene-regulatory model. Curr Opin Genet Dev. 20:164–170. 2010. View Article : Google Scholar : PubMed/NCBI

29 

Moore LD, Le T and Fan G: DNA methylation and its basic function. Neuropsychopharmacology. 38:23–38. 2013. View Article : Google Scholar

30 

Xu F, Mao C, Ding Y, Rui C, Wu L, Shi A, Zhang H, Zhang L and Xu Z: Molecular and enzymatic profiles of mammalian DNA methyltransferases: Structures and targets for drugs. Curr Med Chem. 17:4052–4071. 2010. View Article : Google Scholar : PubMed/NCBI

31 

Jeltsch A: Molecular enzymology of mammalian DNA methyltransferases. Curr Top. Microbiol Immunol. 301:203–225. 2006.

32 

Herman JG and Baylin SB: Gene silencing in cancer in association with promoter hypermethylation. N Engl J Med. 349:2042–2054. 2003. View Article : Google Scholar : PubMed/NCBI

33 

Bird AP: CpG-rich islands and the function of DNA methylation. Nature. 321:209–213. 1986. View Article : Google Scholar : PubMed/NCBI

34 

Costello JF and Plass C: Methylation matters. J Med Genet. 38:285–303. 2001. View Article : Google Scholar : PubMed/NCBI

35 

Weber M and Schübeler D: Genomic patterns of DNA methylation: Targets and function of an epigenetic mark. Curr Opin Cell Biol. 19:273–280. 2007. View Article : Google Scholar : PubMed/NCBI

36 

Bird AP and Wolffe AP: Methylation-induced repression - belts, braces, and chromatin. Cell. 99:451–454. 1999. View Article : Google Scholar : PubMed/NCBI

37 

Koukoura O, Sifakis S and Spandidos DA: DNA methylation in the human placenta and fetal growth (Review). Mol Med Rep. 5:883–889. 2012.PubMed/NCBI

38 

Romani M, Pistillo MP and Banelli B: Environmental Epigenetics: Crossroad between Public Health, Lifestyle, and Cancer Prevention. BioMed Res Int. 2015:5879832015. View Article : Google Scholar : PubMed/NCBI

39 

Bruner-Tran KL, Resuehr D, Ding T, Lucas JA and Osteen KG: The role of endocrine disruptors in the epigenetics of reproductive disease and dysfunction: potential relevance to humans. Curr Obstet Gynecol Rep. 1:116–123. 2012. View Article : Google Scholar : PubMed/NCBI

40 

Anway MD, Cupp AS, Uzumcu M and Skinner MK: Epigenetic transgenerational actions of endocrine disruptors and male fertility. Science. 308:1466–1469. 2005. View Article : Google Scholar : PubMed/NCBI

41 

Danchin É, Charmantier A, Champagne FA, Mesoudi A, Pujol B and Blanchet S: Beyond DNA: Integrating inclusive inheritance into an extended theory of evolution. Nat Rev Genet. 12:475–486. 2011. View Article : Google Scholar : PubMed/NCBI

42 

Bulun SE, Zeitoun KM and Kilic G: Expression of dioxin-related transactivating factors and target genes in human eutopic endometrial and endometriotic tissues. Am J Obstet Gynecol. 182:767–775. 2000. View Article : Google Scholar : PubMed/NCBI

43 

Chiaffarino F, Bravi F, Cipriani S, Parazzini F, Ricci E, Viganò P and La Vecchia C: Coffee and caffeine intake and risk of endometriosis: a meta-analysis. Eur J Nutr. 53:1573–1579. 2014. View Article : Google Scholar : PubMed/NCBI

44 

Nugent BM and Bale TL: The omniscient placenta: Metabolic and epigenetic regulation of fetal programming. Front Neuroendocrinol. 39:28–37. 2015. View Article : Google Scholar : PubMed/NCBI

45 

Smith CJ and Ryckman KK: Epigenetic and developmental influences on the risk of obesity, diabetes, and metabolic syndrome. Diabetes Metab Syndr Obes. 8:295–302. 2015.PubMed/NCBI

46 

Chan RW, Ng EH and Yeung WS: Identification of cells with colony-forming activity, self-renewal capacity, and multipotency in ovarian endometriosis. Am J Pathol. 178:2832–2844. 2011. View Article : Google Scholar : PubMed/NCBI

47 

Wu Y, Strawn E, Basir Z, Halverson G and Guo SW: Aberrant expression of deoxyribonucleic acid methyltransferases DNMT1, DNMT3A, and DNMT3B in women with endometriosis. Fertil Steril. 87:24–32. 2007. View Article : Google Scholar

48 

Szczepańska M, Wirstlein P, Skrzypczak J and Jagodziński PP: Expression of HOXA11 in the mid-luteal endometrium from women with endometriosis-associated infertility. Reprod Biol Endocrinol. 10:12012. View Article : Google Scholar

49 

van Kaam KJ, Delvoux B, Romano A, D'Hooghe T, Dunselman GA and Groothuis PG: Deoxyribonucleic acid methyltransferases and methyl-CpG-binding domain proteins in human endometrium and endometriosis. Fertil Steril. 95:1421–1427. 2011. View Article : Google Scholar : PubMed/NCBI

50 

Hsiao KY, Wu MH, Chang N, Yang SH, Wu CW, Sun HS and Tsai SJ: Coordination of AUF1 and miR-148a destabilizes DNA methyltransferase 1 mRNA under hypoxia in endometriosis. Mol Hum Reprod. 21:894–904. 2015. View Article : Google Scholar : PubMed/NCBI

51 

Dyson MT, Kakinuma T, Pavone ME, Monsivais D, Navarro A, Malpani SS, Ono M and Bulun SE: Aberrant expression and localization of deoxyribonucleic acid methyltransferase 3B in endometriotic stromal cells. Fertil Steril. 104:953–963.e2. 2015. View Article : Google Scholar : PubMed/NCBI

52 

Critchley HO and Saunders PT: Hormone receptor dynamics in a receptive human endometrium. Reprod Sci. 16:191–199. 2009. View Article : Google Scholar : PubMed/NCBI

53 

Shao R, Cao S, Wang X, Feng Y and Billig H: The elusive and controversial roles of estrogen and progesterone receptors in human endometriosis. Am J Transl Res. 6:104–113. 2014.PubMed/NCBI

54 

Tsai MJ and O'Malley BW: Molecular mechanisms of action of steroid/thyroid receptor superfamily members. Annu Rev Biochem. 63:451–486. 1994. View Article : Google Scholar : PubMed/NCBI

55 

Fitzpatrick DR and Wilson CB: Methylation and demethylation in the regulation of genes, cells, and responses in the immune system. Clin Immunol. 109:37–45. 2003. View Article : Google Scholar : PubMed/NCBI

56 

Fuks F: DNA methylation and histone modifications: Teaming up to silence genes. Curr Opin Genet Dev. 15:490–495. 2005. View Article : Google Scholar : PubMed/NCBI

57 

Leung YK, Mak P, Hassan S and Ho SM: Estrogen receptor (ER)-beta isoforms: a key to understanding ER-beta signaling. Proc Natl Acad Sci USA. 103:13162–13167. 2006. View Article : Google Scholar : PubMed/NCBI

58 

Herynk MH and Fuqua SA: Estrogen receptor mutations in human disease. Endocr Rev. 25:869–898. 2004. View Article : Google Scholar : PubMed/NCBI

59 

Hewitt SC, Harrell JC and Korach KS: Lessons in estrogen biology from knockout and transgenic animals. Annu Rev Physiol. 67:285–308. 2005. View Article : Google Scholar : PubMed/NCBI

60 

Burns KA, Rodriguez KF, Hewitt SC, Janardhan KS, Young SL and Korach KS: Role of estrogen receptor signaling required for endometriosis-like lesion establishment in a mouse model. Endocrinology. 153:3960–3971. 2012. View Article : Google Scholar : PubMed/NCBI

61 

Graham JD and Clarke CL: Physiological action of progesterone in target tissues. Endocr Rev. 18:502–519. 1997.PubMed/NCBI

62 

Bulun SE, Monsivais D, Kakinuma T, Furukawa Y, Bernardi L, Pavone ME and Dyson M: Molecular biology of endometriosis: from aromatase to genomic abnormalities. Semin Reprod Med. 33:220–224. 2015. View Article : Google Scholar : PubMed/NCBI

63 

Pellegrini C, Gori I, Achtari C, Hornung D, Chardonnens E, Wunder D, Fiche M and Canny GO: The expression of estrogen receptors as well as GREB1, c-MYC, and cyclin D1, estrogen-regulated genes implicated in proliferation, is increased in peritoneal endometriosis. Fertil Steril. 98:1200–1208. 2012. View Article : Google Scholar : PubMed/NCBI

64 

Brandenberger AW, Lebovic DI, Tee MK, Ryan IP, Tseng JF, Jaffe RB and Taylor RN: Oestrogen receptor (ER)-alpha and ER-beta isoforms in normal endometrial and endometriosis-derived stromal cells. Mol Hum Reprod. 5:651–655. 1999. View Article : Google Scholar : PubMed/NCBI

65 

Attia GR, Zeitoun K, Edwards D, Johns A, Carr BR and Bulun SE: Progesterone receptor isoform A but not B is expressed in endometriosis. J Clin Endocrinol Metab. 85:2897–2902. 2000.PubMed/NCBI

66 

Xue Q, Lin Z, Cheng YH, Huang CC, Marsh E, Yin P, Milad MP, Confino E, Reierstad S, Innes J, et al: Promoter methylation regulates estrogen receptor 2 in human endometrium and endometriosis. Biol Reprod. 77:681–687. 2007. View Article : Google Scholar : PubMed/NCBI

67 

Wu Y, Strawn E, Basir Z, Halverson G and Guo SW: Promoter hypermethylation of progesterone receptor isoform B (PR-B) in endometriosis. Epigenetics. 1:106–111. 2006. View Article : Google Scholar

68 

Misao R, Iwagaki S, Fujimoto J, Sun W and Tamaya T: Dominant expression of progesterone receptor form B mRNA in ovarian endometriosis. Horm Res. 52:30–34. 1999. View Article : Google Scholar

69 

Bukulmez O, Hardy DB, Carr BR, Word RA and Mendelson CR: Inflammatory status influences aromatase and steroid receptor expression in endometriosis. Endocrinology. 149:1190–1204. 2008. View Article : Google Scholar

70 

Meyer JL, Zimbardi D, Podgaec S, Amorim RL, Abrão MS and Rainho CA: DNA methylation patterns of steroid receptor genes ESR1, ESR2 and PGR in deep endometriosis compromising the rectum. Int J Mol Med. 33:897–904. 2014.PubMed/NCBI

71 

Rice DA, Mouw AR, Bogerd AM and Parker KL: A shared promoter element regulates the expression of three steroidogenic enzymes. Mol Endocrinol. 5:1552–1561. 1991. View Article : Google Scholar : PubMed/NCBI

72 

Morohashi K, Honda S, Inomata Y, Handa H and Omura T: A common trans-acting factor, Ad4-binding protein, to the promoters of steroidogenic P-450s. J Biol Chem. 267:17913–17919. 1992.PubMed/NCBI

73 

Zeitoun K, Takayama K, Michael MD and Bulun SE: Stimulation of aromatase P450 promoter (II) activity in endometriosis and its inhibition in endometrium are regulated by competitive binding of steroidogenic factor-1 and chicken ovalbumin upstream promoter transcription factor to the same cis-acting element. Mol Endocrinol. 13:239–253. 1999. View Article : Google Scholar : PubMed/NCBI

74 

Kitawaki J, Kado N, Ishihara H, Koshiba H, Kitaoka Y and Honjo H: Endometriosis: the pathophysiology as an estrogen-dependent disease. J Steroid Biochem Mol Biol. 83:149–155. 2002. View Article : Google Scholar

75 

Tian Y, Kong B, Zhu W, Su S and Kan Y: Expression of steroidogenic factor 1 (SF-1) and steroidogenic acute regulatory protein (StAR) in endometriosis is associated with endometriosis severity. J Int Med Res. 37:1389–1395. 2009. View Article : Google Scholar : PubMed/NCBI

76 

Xue Q, Lin Z, Yin P, Milad MP, Cheng YH, Confino E, Reierstad S and Bulun SE: Transcriptional activation of steroidogenic factor-1 by hypomethylation of the 5′ CpG island in endometriosis. J Clin Endocrinol Metab. 92:3261–3267. 2007. View Article : Google Scholar : PubMed/NCBI

77 

Yamagata Y, Nishino K, Takaki E, Sato S, Maekawa R, Nakai A and Sugino N: Genome-wide DNA methylation profiling in cultured eutopic and ectopic endometrial stromal cells. PLoS One. 9:e836122014. View Article : Google Scholar : PubMed/NCBI

78 

Xue Q, Zhou YF, Zhu SN and Bulun SE: Hypermethylation of the CpG island spanning from exon II to intron III is associated with steroidogenic factor 1 expression in stromal cells of endometriosis. Reprod Sci. 18:1080–1084. 2011. View Article : Google Scholar : PubMed/NCBI

79 

Xue Q, Xu Y, Yang H, Zhang L, Shang J, Zeng C, Yin P and Bulun SE: Methylation of a novel CpG island of intron 1 is associated with steroidogenic factor 1 expression in endometriotic stromal cells. Reprod Sci. 21:395–400. 2014. View Article : Google Scholar :

80 

Hu M, Yao J, Cai L, Bachman KE, van den Brûle F, Velculescu V and Polyak K: Distinct epigenetic changes in the stromal cells of breast cancers. Nat Genet. 37:899–905. 2005. View Article : Google Scholar : PubMed/NCBI

81 

Hoivik EA, Bjanesoy TE, Mai O, Okamoto S, Minokoshi Y, Shima Y, Morohashi K, Boehm U and Bakke M: DNA methylation of intronic enhancers directs tissue-specific expression of steroidogenic factor 1/adrenal 4 binding protein (SF-1/Ad4BP). Endocrinology. 152:2100–2112. 2011. View Article : Google Scholar : PubMed/NCBI

82 

Feinberg AP and Tycko B: The history of cancer epigenetics. Nat Rev Cancer. 4:143–153. 2004. View Article : Google Scholar : PubMed/NCBI

83 

Bell AC and Felsenfeld G: Methylation of a CTCF-dependent boundary controls imprinted expression of the Igf2 gene. Nature. 405:482–485. 2000. View Article : Google Scholar : PubMed/NCBI

84 

Zanatta A, Rocha AM, Carvalho FM, Pereira RM, Taylor HS, Motta EL, Baracat EC and Serafini PC: The role of the Hoxa10/HOXA10 gene in the etiology of endometriosis and its related infertility: A review. J Assist Reprod Genet. 27:701–710. 2010. View Article : Google Scholar : PubMed/NCBI

85 

Eun Kwon H and Taylor HS: The role of HOX genes in human implantation. Ann N Y Acad Sci. 1034:1–18. 2004. View Article : Google Scholar

86 

Taylor HS, Arici A, Olive D and Igarashi P: HOXA10 is expressed in response to sex steroids at the time of implantation in the human endometrium. J Clin Invest. 101:1379–1384. 1998. View Article : Google Scholar : PubMed/NCBI

87 

Gui Y, Zhang J, Yuan L and Lessey BA: Regulation of HOXA-10 is and its expression in normal and abnormal endometrium. Mol Hum Reprod. 5:866–873. 1999. View Article : Google Scholar : PubMed/NCBI

88 

Szczepańska M, Wirstlein P, Luczak M, Jagodziński PP and Skrzypczak J: Reduced expression of HOXA10 in the midluteal endometrium from infertile women with minimal endometriosis. Biomed Pharmacother. 64:697–705. 2010. View Article : Google Scholar

89 

Wu Y, Halverson G, Basir Z, Strawn E, Yan P and Guo SW: Aberrant methylation at HOXA10 may be responsible for its aberrant expression in the endometrium of patients with endometriosis. Am J Obstet Gynecol. 193:371–380. 2005. View Article : Google Scholar : PubMed/NCBI

90 

Kim JJ, Taylor HS, Lu Z, Ladhani O, Hastings JM, Jackson KS, Wu Y, Guo SW and Fazleabas AT: Altered expression of HOXA10 in endometriosis: Potential role in decidualization. Mol Hum Reprod. 13:323–332. 2007. View Article : Google Scholar : PubMed/NCBI

91 

Lee B, Du H and Taylor HS: Experimental murine endometriosis induces DNA methylation and altered gene expression in eutopic endometrium. Biol Reprod. 80:79–85. 2009. View Article : Google Scholar

92 

Fambrini M, Sorbi F, Bussani C, Cioni R, Sisti G and Andersson KL: Hypermethylation of HOXA10 gene in mid-luteal endometrium from women with ovarian endometriomas. Acta Obstet Gynecol Scand. 92:1331–1334. 2013. View Article : Google Scholar : PubMed/NCBI

93 

Lu H, Yang X, Zhang Y, Lu R and Wang X: Epigenetic disorder may cause downregulation of HOXA10 in the eutopic endometrium of fertile women with endometriosis. Reprod Sci. 20:78–84. 2013. View Article : Google Scholar

94 

Bromer JG, Wu J, Zhou Y and Taylor HS: Hypermethylation of homeobox A10 by in utero diethylstilbestrol exposure: An epigenetic mechanism for altered developmental programming. Endocrinology. 150:3376–3382. 2009. View Article : Google Scholar : PubMed/NCBI

95 

Naqvi H, Ilagan Y, Krikun G and Taylor HS: Altered genome-wide methylation in endometriosis. Reprod Sci. 21:1237–1243. 2014. View Article : Google Scholar : PubMed/NCBI

96 

Simpson ER, Mahendroo MS, Means GD, Kilgore MW, Hinshelwood MM, Graham-Lorence S, Amarneh B, Ito Y, Fisher CR, Michael MD, et al: Aromatase cytochrome P450, the enzyme responsible for estrogen biosynthesis. Endocr Rev. 15:342–355. 1994.PubMed/NCBI

97 

Abu Hashim H: Potential role of aromatase inhibitors in the treatment of endometriosis. Int J Womens Health. 6:671–680. 2014. View Article : Google Scholar : PubMed/NCBI

98 

Maia H Jr, Haddad C, Coelho G and Casoy J: Role of inflammation and aromatase expression in the eutopic endometrium and its relationship with the development of endometriosis. Womens Health (Lond Engl). 8:647–658. 2012. View Article : Google Scholar

99 

Bulun SE, Takayama K, Suzuki T, Sasano H, Yilmaz B and Sebastian S: Organization of the human aromatase p450 (CYP19) gene. Semin Reprod Med. 22:5–9. 2004. View Article : Google Scholar : PubMed/NCBI

100 

Izawa M, Harada T, Taniguchi F, Ohama Y, Takenaka Y and Terakawa N: An epigenetic disorder may cause aberrant expression of aromatase gene in endometriotic stromal cells. Fertil Steril. 89(Suppl 5): 1390–1396. 2008. View Article : Google Scholar

101 

Izawa M, Taniguchi F, Uegaki T, Takai E, Iwabe T, Terakawa N and Harada T: Demethylation of a nonpromoter cytosine-phosphate-guanine island in the aromatase gene may cause the aberrant up-regulation in endometriotic tissues. Fertil Steril. 95:33–39. 2011. View Article : Google Scholar

102 

Dubois RN, Abramson SB, Crofford L, Gupta RA, Simon LS, Van De Putte LB and Lipsky PE: Cyclooxygenase in biology and disease. FASEB J. 12:1063–1073. 1998.PubMed/NCBI

103 

Ota H, Igarashi S, Sasaki M and Tanaka T: Distribution of cyclooxygenase-2 in eutopic and ectopic endometrium in endometriosis and adenomyosis. Hum Reprod. 16:561–566. 2001. View Article : Google Scholar : PubMed/NCBI

104 

Matsuzaki S, Canis M, Pouly JL, Wattiez A, Okamura K and Mage G: Cyclooxygenase-2 expression in deep endometriosis and matched eutopic endometrium. Fertil Steril. 82:1309–1315. 2004. View Article : Google Scholar : PubMed/NCBI

105 

Buchweitz O, Staebler A, Wülfing P, Hauzman E, Greb R and Kiesel L: COX-2 overexpression in peritoneal lesions is correlated with nonmenstrual chronic pelvic pain. Eur J Obstet Gynecol Reprod Biol. 124:216–221. 2006. View Article : Google Scholar

106 

Zidan HE, Rezk NA, Alnemr AA and Abd El Ghany AM: COX-2 gene promoter DNA methylation status in eutopic and ectopic endometrium of Egyptian women with endometriosis. J Reprod Immunol. 112:63–67. 2015. View Article : Google Scholar : PubMed/NCBI

107 

Wang D, Chen Q, Zhang C, Ren F and Li T: DNA hypo-methylation of the COX-2 gene promoter is associated with up-regulation of its mRNA expression in eutopic endometrium of endometriosis. Eur J Med Res. 17:122012. View Article : Google Scholar

108 

Starzinski-Powitz A, Gaetje R, Zeitvogel A, Kotzian S, Handrow-Metzmacher H, Herrmann G, Fanning E and Baumann R: Tracing cellular and molecular mechanisms involved in endometriosis. Hum Reprod Update. 4:724–729. 1998. View Article : Google Scholar

109 

Wu Y, Starzinski-Powitz A and Guo SW: Trichostatin A, a histone deacetylase inhibitor, attenuates invasiveness and reactivates E-cadherin expression in immortalized endometriotic cells. Reprod Sci. 14:374–382. 2007. View Article : Google Scholar : PubMed/NCBI

110 

Zhou H, Li J, Podratz KC, Tipton T, Marzolf S, Chen HB and Jiang SW: Hypomethylation and activation of syncytin-1 gene in endometriotic tissue. Curr Pharm Des. 20:1786–1795. 2014. View Article : Google Scholar

111 

Stern RC, Dash R, Bentley RC, Snyder MJ, Haney AF and Robboy SJ: Malignancy in endometriosis: Frequency and comparison of ovarian and extraovarian types. Int J Gynecol Pathol. 20:133–139. 2001. View Article : Google Scholar : PubMed/NCBI

112 

Matalliotakis I, Mahutte NG, Koukoura O and Arici A: Endometriosis associated with Stage IA clear cell ovarian carcinoma in a woman with IVF-ET treatments in the Yale Series. Arch Gynecol Obstet. 274:184–186. 2006. View Article : Google Scholar : PubMed/NCBI

113 

Shih IeM and Kurman RJ: Ovarian tumorigenesis: A proposed model based on morphological and molecular genetic analysis. Am J Pathol. 164:1511–1518. 2004. View Article : Google Scholar : PubMed/NCBI

114 

Koukoura O, Spandidos DA, Daponte A and Sifakis S: DNA methylation profiles in ovarian cancer: Implication in diagnosis and therapy (Review). Mol Med Rep. 10:3–9. 2014.PubMed/NCBI

115 

Martini M, Ciccarone M, Garganese G, Maggiore C, Evangelista A, Rahimi S, Zannoni G, Vittori G and Larocca LM: Possible involvement of hMLH1, p16(INK4a) and PTEN in the malignant transformation of endometriosis. Int J Cancer. 102:398–406. 2002. View Article : Google Scholar : PubMed/NCBI

116 

Senthong A, Kitkumthorn N, Rattanatanyong P, Khemapech N, Triratanachart S and Mutirangura A: Differences in LINE-1 methylation between endometriotic ovarian cyst and endometriosis-associated ovarian cancer. Int J Gynecol Cancer. 24:36–42. 2014. View Article : Google Scholar

117 

Guo C, Ren F, Wang D, Li Y, Liu K, Liu S and Chen P: RUNX3 is inactivated by promoter hypermethylation in malignant transformation of ovarian endometriosis. Oncol Rep. 32:2580–2588. 2014.PubMed/NCBI

118 

Perez-Janices N, Blanco-Luquin I, Torrea N, Liechtenstein T, Escors D, Cordoba A, Vicente-Garcia F, Jauregui I, De La Cruz S, Illarramendi JJ, et al: Differential involvement of RASSF2 hypermethylation in breast cancer subtypes and their prognosis. Oncotarget. 6:23944–23958. 2015. View Article : Google Scholar : PubMed/NCBI

119 

Guerrero-Setas D, Pérez-Janices N, Blanco-Fernandez L, Ojer A, Cambra K, Berdasco M, Esteller M, Maria-Ruiz S, Torrea N and Guarch R: RASSF2 hypermethylation is present and related to shorter survival in squamous cervical cancer. Mod Pathol. 26:1111–1122. 2013. View Article : Google Scholar : PubMed/NCBI

120 

Zhao L, Cui Q, Lu Z and Chen J: Aberrant methylation of RASSF2A in human pancreatic ductal adenocarcinoma and its relation to clinicopathologic features. Pancreas. 41:206–211. 2012. View Article : Google Scholar

121 

Ren F, Wang DB, Li T, Chen YH and Li Y: Identification of differentially methylated genes in the malignant transformation of ovarian endometriosis. J Ovarian Res. 7:732014. View Article : Google Scholar : PubMed/NCBI

122 

Borghese B, Barbaux S, Mondon F, Santulli P, Pierre G, Vinci G, Chapron C and Vaiman D: Research resource: Genome-wide profiling of methylated promoters in endometriosis reveals a subtelomeric location of hypermethylation. Mol Endocrinol. 24:1872–1885. 2010. View Article : Google Scholar : PubMed/NCBI

123 

Dyson MT, Roqueiro D, Monsivais D, Ercan CM, Pavone ME, Brooks DC, Kakinuma T, Ono M, Jafari N, Dai Y, et al: Genome-wide DNA methylation analysis predicts an epigenetic switch for GATA factor expression in endometriosis. PLoS Genet. 10:e10041582014. View Article : Google Scholar : PubMed/NCBI

124 

Yoo CB and Jones PA: Epigenetic therapy of cancer: Past, present and future. Nat Rev Drug Discov. 5:37–50. 2006. View Article : Google Scholar : PubMed/NCBI

125 

Nie Jichan, Liu Xishi and Guo SW: Promoter hypermethylation of progesterone receptor isoform B (PR-B) in adenomyosis and its rectification by a histone deacetylase inhibitor and a demethylation agent. Reprod Sci. 17:995–1005. 2010. View Article : Google Scholar

126 

Arosh JA, Lee J, Starzinski-Powitz A and Banu SK: Selective inhibition of prostaglandin E2 receptors EP2 and EP4 modulates DNA methylation and histone modification machinery proteins in human endometriotic cells. Mol Cell Endocrinol. 409:51–58. 2015. View Article : Google Scholar : PubMed/NCBI

127 

Bergman MD, Schachter BS, Karelus K, Combatsiaris EP, Garcia T and Nelson JF: Up-regulation of the uterine estrogen receptor and its messenger ribonucleic acid during the mouse estrous cycle: The role of estradiol. Endocrinology. 130:1923–1930. 1992.PubMed/NCBI

Related Articles

Journal Cover

April-2016
Volume 13 Issue 4

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Koukoura O, Sifakis S and Spandidos DA: DNA methylation in endometriosis (Review). Mol Med Rep 13: 2939-2948, 2016
APA
Koukoura, O., Sifakis, S., & Spandidos, D.A. (2016). DNA methylation in endometriosis (Review). Molecular Medicine Reports, 13, 2939-2948. https://doi.org/10.3892/mmr.2016.4925
MLA
Koukoura, O., Sifakis, S., Spandidos, D. A."DNA methylation in endometriosis (Review)". Molecular Medicine Reports 13.4 (2016): 2939-2948.
Chicago
Koukoura, O., Sifakis, S., Spandidos, D. A."DNA methylation in endometriosis (Review)". Molecular Medicine Reports 13, no. 4 (2016): 2939-2948. https://doi.org/10.3892/mmr.2016.4925