Next Article in Journal
Unraveling the Role of Antimicrobial Peptides in Insects
Previous Article in Journal
Non-Coding RNAs as Biomarkers for Embryo Quality and Pregnancy Outcomes: A Systematic Review and Meta-Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A View on Uterine Leiomyoma Genesis through the Prism of Genetic, Epigenetic and Cellular Heterogeneity †

D.O. Ott Research Institute of Obstetrics, Gynecology and Reproductology, 199034 St. Petersburg, Russia
*
Author to whom correspondence should be addressed.
Dedicated to Professor Vladislav S. Baranov.
Int. J. Mol. Sci. 2023, 24(6), 5752; https://doi.org/10.3390/ijms24065752
Submission received: 4 December 2022 / Revised: 7 March 2023 / Accepted: 15 March 2023 / Published: 17 March 2023
(This article belongs to the Section Molecular Genetics and Genomics)

Abstract

:
Uterine leiomyomas (ULs), frequent benign tumours of the female reproductive tract, are associated with a range of symptoms and significant morbidity. Despite extensive research, there is no consensus on essential points of UL initiation and development. The main reason for this is a pronounced inter- and intratumoral heterogeneity resulting from diverse and complicated mechanisms underlying UL pathobiology. In this review, we comprehensively analyse risk and protective factors for UL development, UL cellular composition, hormonal and paracrine signalling, epigenetic regulation and genetic abnormalities. We conclude the need to carefully update the concept of UL genesis in light of the current data. Staying within the framework of the existing hypotheses, we introduce a possible timeline for UL development and the associated key events—from potential prerequisites to the beginning of UL formation and the onset of driver and passenger changes.

1. Introduction

Uterine leiomyoma (UL) is the most common tumour of the female reproductive system, with an incidence rate reaching 70–80% [1]. Despite its benign nature and low risk of malignisation, UL can be characterised by rapid growth, large size and multiple nodes [2]. In approximately 50% of cases, ULs are associated with symptoms that are detrimental to the life quality, including pelvic pain, excessive uterine bleeding and urinary incontinence, and may cause reproductive disorders, in particular, infertility and pregnancy loss [3,4].
UL aetiopathogenesis is being actively investigated. Parity, oral contraceptive intake and smoking are among possible protective factors for UL development [5,6,7,8,9]. The risk factors include advanced reproductive age, African and Latin American ethnicity, UL cases in family medical history, obesity, early menarche, arterial hypertension, chronic inflammation, exposure to xenoestrogens in early ontogenesis, sexually transmitted infections and alcohol [8,9,10,11]. Researchers place the highest significance on hereditary predisposition, endocrine and paracrine factors, somatic gene and chromosomal mutations and epigenetic disorders triggered by exposure to endogenous and exogenous factors [9,10,12,13]. Identifying precise mechanisms of UL tumorigenesis and developing universal treatment approaches is complicated by UL being a multifactorial disease with pronounced heterogeneity in clinical, morphological and molecular characteristics.
The present review analyses the cellular, endocrine, paracrine, genetic and epigenetic aspects of UL tumorigenesis, discusses potential prerequisites and suggests possible key events on the timeline of UL development.

2. UL Cellular Origin and Composition

ULs represent dense nodules (myomas) of rounded shape, consisting of transformed cells that produce an excess of extracellular matrix (ECM) [13]. The ECM content varies in different ULs and is presumably subject to change as the myoma progresses and undergoes involution [14]. The main cellular components of a UL are smooth muscle cells, fibroblasts and vascular endothelial cells; other detected cellular components include blood cells, neurons and telocytes [15,16,17,18,19,20,21,22].
Myomas are commonly considered to be monoclonal in nature, with their diverse cellular composition resulting from the expansion and differentiation of a single transformed myometrium cell [9,10,23]. Tissue-specific stem cells are believed to be key to the formation of monoclonal ULs, as they are characterised primarily by their capacity for asymmetric cell division and further differentiation. A cell population with stem/progenitor cell properties (the so-called “side population”) was detected for the first time in the myometrium of non-pregnant women in 2007 [24]. Transcription and cell surface antigen profiles specific to this population were also described [25,26]. Subsequently, a side population expressing several stem markers (OCT4 (octamer-binding transcription factor 4), NANOG (Nanog homeobox), DNMT3B (DNA methyltransferase 3B), GDF3 (growth differentiation factor-3)) was isolated from a UL as well [27,28].
UL stem cells have several differences from those of the myometrium. In culture conditions, ULs form fewer colonies of non-differentiated cells, and the cells, in turn, have a shorter cell cycle [27,29,30]. Stem cells isolated from ULs have also been found to carry MED12 (mediator complex subunit 12) mutations [29], increased DNA damage and specific expression patterns of genes involved in the repair of single- and double-strand DNA breaks [31]. At the same time, UL stem cells as well as myometrium-derived stem cells are deficient in oestrogen or progesterone receptors [29,32].
To grow and proliferate, myometrial and UL stem cells need the presence of differentiated cells with higher levels of hormone receptors. The interaction between stem and mature cells is believed to occur through paracrine signals (WNT (wingless-type) proteins, growth factors, etc.) [28,32,33] or intercellular contacts with telocytes—a type of interstitial cells [17,19,34,35]. Myometrial telocytes express oestrogen and progesterone receptors and can form contacts with various cell types, including smooth muscle, nervous and stem cells, with their thin long extensions called telopodes [36].
The cellular diversity is a considerable, but not the only, contributor to the intratumoral heterogeneity in ULs. Signalling pathways, gene variants, chromosomal abnormalities and epigenetic regulation are among the other aspects of tumorigenesis that create the unique profile of each UL cell (Figure 1).

3. Diversity of Endo-, Auto- and Paracrine Mechanisms Regulating UL Development

Sex steroid hormones are the key regulators of UL growth. Not only can UL cells bind oestrogen and progesterone from the blood flow, but they can also convert circulating androgens into oestrogens, which explains higher oestrogen levels in UL tissue as compared to the adjacent myometrium [37].
Testosterone and androstenedione are converted into oestradiol with aromatase and 17β-hydroxysteroid dehydrogenase enzymes. The activity of this enzyme system enables UL cells to produce enough oestrogen to support their growth [38]. Treatment of cultured UL cells with selective aromatase inhibitors abolishes the promoting effect of testosterone and androstenedione on UL cell proliferation [38]. The enzymatic conversion of androgens to oestrogens is also typical of fat tissue [39], making excessive weight a significant risk factor for ULs (Figure 1) [40,41,42,43].
Oestrogens, particularly their most active natural form, 17β-oestradiol, are essential components of many pathways associated with UL development. Oestrogens act through their specific receptors (α and β), which bind to 17β-oestradiol and its natural analogues (estrone and estriol) as well as a variety of xenoestrogens, many of which are considered to be endocrine-disrupting chemicals (EDCs) [44]. Oestrogen receptors bind with oestrogen response elements (EREs) in the genome or with other transcription factors, thus activating the expression of genes coding growth factors, ECM proteins and finally, oestrogen and progesterone receptors [12]. However, UL cell proliferation requires the presence of both oestrogen and progesterone [45,46].
During the luteal phase of the menstrual cycle, characterised by an increased progesterone blood level, a UL features a higher expression of proliferation markers and a higher mitotic activity [45,47,48]. Increased proliferative activity has also been observed in ULs of postmenopausal women following combined oestrogen and progesterone therapy in comparison to those with oestrogen therapy or without hormone replacement [45,49]. Progesterone-mediated UL growth regulation does not only occur through increased proliferation but also through decreased apoptosis and higher ECM accumulation [46,50,51]. Directly or via interacting with other transcription factors, progesterone receptors can regulate the expression of many genes, in particular, proliferating cell nuclear antigen (PCNA), epidermal growth factor (EGF), transforming growth factor beta (TGF-β), anti-apoptotic Bcl-2 protein and the miR-29b microRNA, which regulates the accumulation of ECM [37,52].
Oestrogen and progesterone also have non-genomic mechanisms of action in UL cells [53,54], realised through the binding of activated hormone receptors with cytoplasmic and membrane regulatory proteins involved in various kinase cascades—MAPK (mitogen-activated protein kinase), PI3K/AKT (phosphatidylinositol 3-kinase / protein kinase B), PLC/PKC (phospholipase C / protein kinase C) and cAMP/PKA (cyclic AMP / protein kinase A) [54,55]. Paracrine signalling pathways also facilitate intercellular communication in myometrial and UL tissues [24,34]. UL-specific MED12 mutations are linked, among other factors, to changes in the expression and functioning of various WNT/β-catenin signalling components [56,57], whereas the inhibitors of the WNT/β-catenin pathway (vitamin D, Simvastatin, etc.) can suppress UL cell proliferation and ECM accumulation by lowering the expression of oestrogen and progesterone receptors [58,59,60,61].
Non-hormonal UL development regulators are growth factors (fibroblast growth factor (FGF), vascular endothelial growth factor (VEGF), epidermal growth factor (EGF), insulin-like growth factor (IGF), platelet-derived growth factor (PDGF), transforming growth factor β (TGF-β)), proinflammatory cytokines (tumour necrosis factor TNF-α) and ECM components [62]. The UL ECM includes collagens, fibronectin, laminins and proteoglycans, which, in turn, facilitate mechanotransduction and the launch of signalling pathways (WNT/β-catenin, ERK (extracellular signal-regulated kinase)/MAPK and Hippo/YAP (yes-associated protein) pathways) [63,64]. Therefore, the ECM is not just a side product of the UL cell anomalous functioning but an active participant in the neoplastic process thanks to its reciprocal interaction with UL cells.
Thus, oestrogen and progesterone synergistically regulate the mechanisms of cell proliferation, differentiation and fibrosis in the UL through their receptors. The UL’s overlapping network of intra- and intercellular signalling goes beyond sex hormones to include a variety of growth factors, receptors, cellular adhesion molecules, cytokines, ECM proteins and kinase cascades associated with them. Not only can the reprogramming of fundamental hormonal, paracrine and immune response mechanisms result from genetic and epigenetic changes, but it can serve as an independent factor triggering genetic instability and aberrant gene expression in UL cells.

4. The Role of Woman’s Genotype in UL Aetiology

Researchers have theorised about the hereditary nature of ULs for a rather long time, considering the ample evidence of a higher UL frequency among close blood relatives. In the 1950s, The Lancet published medical data on four generations of one family, including 15 women over 15 years old [65]. According to the family history, nine of the women were reliably found to have ULs, which attested to the high significance of hereditary factors in the development of ULs in the female members of this family. A later study, which included 215 women from 97 families, revealed that the risk of UL in women whose mothers had ULs almost doubled [66]. The role of heredity in UL aetiology was confirmed in the study of mono- and dizygotic twins [67] and several ethnic groups [68,69].
Women of ethnic Latin American and African ancestry have a considerably higher risk of developing ULs than women of European ancestry [1]. Furthermore, African ancestry is additionally associated with earlier development of ULs and heavier symptoms [70]. Along with hereditary factors, vitamin D deficiency, which often occurs in black women, dietary specifics (a high share of soy products, in particular), obesity, high-stress levels and a high content of stem cells in the myometrium are also considered to be of significance for the observed differences [9].
Modern genomic research (genome-wide association search (GWAS) and genome-wide single-nucleotide polymorphism analysis) has identified variants determining predisposition for UL in over 50 genes (WNT4 (Wnt family member 4), GREB1 (growth regulating oestrogen receptor binding 1), TERC (telomerase RNA component), TERT (telomerase reverse transcriptase), HMGA1 (high mobility group AT-hook 1), FOXO1 (forkhead box O1), TP53 (tumour protein p53), etc.) [71,72,73,74,75,76,77,78,79,80,81]. These genes are involved in DNA damage repair, telomere length maintenance, hormonal and paracrine regulation, apoptosis, urogenital system development and early menarche. Furthermore, so-called “protective” single-nucleotide variants, which reduce UL risk, have been discovered in the introns of the following genes: KCNMB2 (potassium calcium-activated channel subfamily M regulatory beta subunit 2), FBN2 (fibrillin 2), ESR1 (oestrogen receptor 1) and CELF4 (CUGBP Elav-like family member 4) [82]. Researchers have also described COL6A3 (collagen type VI alpha 3 chain), COL13A (collagen type XIII alpha 1 chain), ARHGAP26 (Rho GTPase activating protein 26), MAN1C1 (mannosidase alpha class 1C member 1), BET1L (Bet1 golgi vesicular membrane trafficking protein like), TNRC6B (trinucleotide repeat containing adaptor 6B) and COMT (catechol-O-methyltransferase) gene variants associated with the accumulation of ECM in the UL tissue, the size, localisation and the multiple form of UL [83,84,85,86,87].
Particular hereditary syndromes (Reed’s, Alport, Proteus and Cowden syndromes), which feature a high risk of UL [10], warrant special attention. Frequently referred to as the HLRCC syndrome (Hereditary Leiomyomatosis and Renal Cell Carcinoma), Reed’s syndrome is caused by germline mutations in FH, the gene encoding fumarate hydratase, a tricarboxylic acid cycle enzyme [88,89]. Patients with this syndrome face a high risk of developing renal cell carcinoma and benign leiomyomatosis, with women displaying a high frequency of multiple ULs. With Reed’s syndrome, the neoplastic process is triggered through biallelic FH gene inactivation, fumarate accumulation in the cells, changes in cell metabolism, the stabilisation of hypoxia-inducible factors (HIFs), the activation of carcinogenic transcription factor NRF2 (nuclear factor erythroid 2 related factor 2) and changes to the expression profile [9,10]. In most cases, however, the development of FH-deficient ULs is associated with somatic but not germline FH mutations (whole-gene deletions, frameshift or missense mutations) [90,91,92,93]. ULs with FH mutations are often histologically distinctive, demonstrate hypercellularity, nuclear atypia, inclusion-like nucleoli, stromal oedema, and occur with a high frequency among atypical leiomyomas (37.3% compared to 1.6% in unselected ULs) [91,92,94].
Alport syndrome is caused by basement membrane damage resulting from germinal mutations in the genes of type IV collagen (COL4A3, COL4A4, COL4A5 (collagen type IV alpha 3, 4, 5 chain genes)). Patients with this syndrome experience primarily progressive loss of kidney function and hearing loss [95]. ULs are considerably less frequent in patients with Alport syndrome than with Reed’s syndrome, and their development is associated with mutations at the COL4A5-COL4A6 locus. Overall, the contribution of FH and COL4A5-COL4A6 mutations in the total UL frequency is low, not exceeding 5% [96,97].
Therefore, the significance of heredity in UL aetiology has been confirmed with extensive data. At present, however, genotype alone is not sufficient for an accurate prognosis of UL development in a woman except in cases of rare hereditary syndromes. ULs are associated with allelic variations in genes involved in a wide range of processes, which presumably determines its high occurrence rate and further emphasises its multifactorial nature (Figure 1).

5. Spectrum of Somatic Genetic Aberrations in UL Cells

ULs are characterised by varied chromosomal abnormalities and high heterogeneity in acquired somatic mutations, with the most frequent being harboured in MED12, HMGA2 (high-mobility group AT-hook 2), FH and COL4A5-COL4A6 genes. Less frequent mutations in MED8 (mediator complex subunit 8), MED18 (mediator complex subunit 18), CDK8 (cyclin-dependent kinase 8), CASC15 (cancer susceptibility 15), COL4A6, DCN (decorin), AHR (aryl hydrocarbon receptor), NRG1 (neuregulin 1), ADAM18 (ADAM metallopeptidase domain 18), HUWE1 (HECT, UBA and WWE domain containing E3 ubiquitin protein ligase 1), FBXW4 (F-box and WD repeat domain containing 4), FBXL13 (F-box and leucine-rich repeat protein 13) and CAPRIN1 (cell cycle-associated protein 1) genes [98] and mitochondrial DNA mutations are also associated with ULs [99].

5.1. MED12 Mutations

In different samples, the frequency of ULs with MED12 mutations (MED12-ULs) varies from 31% to 85% [100,101,102,103,104,105]. The presence of MED12-ULs in women is associated with a history of inflammatory disorders of the pelvic organs, nulliparity and African ancestry [105,106]. MED12 mutations are detected more frequently in multiple ULs and small, mitotically active myomas of normal morphology and subserous (as opposed to intramural) localisation with a high ECM content [102,103,106,107,108,109,110,111]. Furthermore, a UL susceptibility locus was identified 250 kb upstream of the MED12 gene [75]. However, the environment plays an important role in the emergence of MED12 mutations because women with multiple ULs can have MED12 mutations in some of the myomas but not all of them.
The MED12 gene is located in the Xq13.1 chromosome region and encodes the large mediator complex subunit (the so-called “Mediator”), which bridges gene-specific regulatory proteins in the promoter region to the RNA polymerase II initiation complex. Pathological variants in MED12 are represented in ULs primarily by heterozygous missense mutations in codon 44 of exon 2; ULs have also been found to feature missense mutations in codons 36 and 43, in-frame deletion plus insertion mutations and in rare cases, exon 1 mutations [100,112]. MED12 mutations have been detected in UL stem cells [29] but not in the adjacent myometrium or the UL pseudocapsule [113] and are therefore treated as the potential drivers of UL tumorigenesis.
In MED12-ULs, MED12 is expressed primarily from the mutant allele [100], indicating the involvement of the anomalous variant of MED12 protein in the neoplastic process. A UL mouse model demonstrates that the expression of MED12 c.131G>A mutant allele in uterine tissue is sufficient for the development of UL-like tumours with chromosomal aberrations [114]. However, the precise molecular mechanism of how MED12 mutations trigger UL growth remains to be established. MED12 mutations undermine the transduction of several cytoplasmic and transcription signals due to the disrupted ability of the Mediator to activate cyclin C and CDK8/CDK19 cyclin-dependent kinases [115,116,117]. Furthermore, MED12 mutations are associated with an increased expression of WNT/β-catenin signalling and changes in the regulation of downstream pathways (mTOR, TGF-β) linked to cell proliferation, autophagy and ECM accumulation [56,57,118]. Other signalling pathways—MAPK, AKT and IGF (insulin-like growth factor) —mediate the action of oestrogen and progesterone in MED12-ULs. Anomalous MED12 protein interacts with the progesterone receptor more efficiently, driving the progesterone-dependent expression of potential carcinogens in MED12-ULs (in particular, this has been shown for the receptor activator of nuclear factor kappa-Β ligand (RANKL) and tryptophan 2,3-dioxygenase (TDO2) genes) [119,120]. Gene expression in MED12-ULs is additionally affected by epigenetic abnormalities including post-translation modifications of histones, enhancer architecture and DNA methylation [111,121]. Nevertheless, changes in the global level of UL DNA hydroxymethylation are not associated with MED12 mutations [122]. Smooth muscle cells from MED12-ULs also exhibit signs of replication stress, such as higher levels of R-loops, altered replication fork dynamics and delayed S phase of the cell cycle [123].
Importantly, not all cells of a MED12-UL carry MED12 mutations. Within the first several passages in culture, MED12-ULs manifest an elimination of cells with MED12 mutations with the expansion of wild-type cells [124]. Wu et al. have established that MED12 mutations are present in smooth muscle cells but not tumour-associated fibroblasts [109]. However, Goad et al. have recently used single-cell RNA sequencing to show that intercellular heterogeneity of MED12 mutations is typical for all three predominant UL populations: fibroblasts, smooth muscle cells and endothelial cells [22]. Variations in the mechanisms of in vitro maintenance of heterogeneous MED12-UL cell populations could include the differential effect of sex steroids because UL smooth muscle cell proliferation is stimulated mostly by progesterone, while that of fibroblasts is oestrogen-regulated [109,125]. In light of this, it is practical to prioritise the optimisation of culture conditions, mutant UL cell immortalisation and use of tumour xenografts when creating cellular and organotypic MED12-UL models [57,126,127,128,129].
MED12 mutations could either be the only genetic aberration detected in a myoma [118,130,131] or coexist with other “driver” mutations (HMGA2, FH, COL4A5-COL4A6) [132,133]. Furthermore, karyotypically normal and abnormal ULs feature the same frequency of MED12 mutations [101,134], which do not affect the growth potential of cells with chromosomal aberrations in vitro [124,134,135].

5.2. Chromosomal Abnormalities

At least 40% of ULs carry chromosomal abnormalities. Initially, they were detected using karyotyping of short-term and long-term tumour cell cultures. Over the last few decades, cytogeneticists have described over 200 UL-specific abnormalities, of which the most frequent are structural rearrangements involving chromosome regions 1p, 6p21, 7q and 12q15 [136]. In a significant share of ULs, chromosomal abnormalities were presented in the mosaic state alongside karyotypically normal cells, which invited hypotheses of their secondary origin within the tumour or in culture conditions [137,138,139]. The introduction of FISH analysis on interphase nuclei ushered in the discovery that mosaic chromosomal rearrangements may be present in uncultured myomas and presumably emerge in vivo [140]. The further advancement of molecular technologies opened opportunities for the analysis of chromosomal aberrations with the use of genomic DNA from the native tumour. Researchers identified the genes involved in the most frequent rearrangements, chromosomal abnormalities in cell clones incapable of in vitro growth and extremely complicated variants of chromosomal aberrations, such as chromothripsis [134,141,142,143]. In our opinion, conventional and molecular techniques are complementary rather than competitive and create a more objective picture for a better understanding of the cytogenetic aspects of UL pathogenesis.

5.2.1. 12q14-15 Rearrangements

Region 12q14-15 was among the first to be referred to as a “hot spot” of chromosomal rearrangements in ULs [144]. This region is characterised primarily by reciprocal changes involving region 14q24, with fewer frequent translocations involving two or more other chromosomes and intrachromosomal aberrations [136]. The development of ULs carrying 12q14-15 rearrangements has been associated with the HMGA2/HMGI-C locus [141,145,146]. The HMGA2 gene encodes a non-histone protein that regulates chromatin conformation and gene expression via binding with AT-rich DNA sequences of enhancers and promoters. Normally, this gene is active in embryonic tissues, stem cells and many malignant tumours [147,148,149].
In ULs, most 12q14-15 breaks are located upstream of the promoter sequence of the HMGA2 gene, upregulating its expression [146,150,151]. Along with transcriptional alterations, the higher levels of HMGA2 protein in ULs could be associated with the loss of let-7 regulatory microRNA binding sites in truncated or chimeric transcripts of the damaged HMGA2 gene [152,153]. Furthermore, the overexpression of HMGA2 is typical for a considerable subset of ULs without 12q14-15 rearrangements [154] and could be caused by either HMGA2 hypomethylation or decreased levels of repressor microRNAs: let-7a, miR-26a, miR-26b, miR-93 and miR-106b [133,155]. Therefore, regardless of the presence or absence of 12q15 rearrangements, all ULs featuring HMGA2 overexpression are consolidated into a single pathogenetic subgroup, HMGA2-ULs.
HMGA2-ULs contain 90% of HMGA2-positive smooth muscle cells and are characterised by a larger size and higher growth rate, mostly through mitotic activity rather than excessive ECM accumulation [108,109,145]. HMGA2-ULs respond to gonadotropin-releasing hormone agonist therapy less than karyotypically normal ULs, which suggests a lower dependence of HMGA2-ULs on circulating sex steroids [156]. In standard culture conditions without hormone supplementation, the cells of a UL with rearrangements at the HMGA2 locus are stably maintained throughout multiple passages, but their growth is coupled with a decrease in proliferation level and HMGA2 expression and a parallel increase in cellular senescence markers [124,157,158,159].
It has been demonstrated in an animal UL model that the expression of a shortened HMGA2 variant in myometrial stem cells is sufficient for their transformation into progenitor UL cells [160]. HMGA2-ULs are further characterised by the activation of AKT and IGF signalling pathways, increased angiogenesis and the overexpression of the oestrogen receptor α (ESR1), the PLAG1 (pleomorphic adenoma gene 1) proto-oncogene, CCND1, CCND2, CCND3 and CDK6 cell cycle regulators, epidermal (EGF) and transforming (TGFA) growth factors, the fibroblast growth factor (FGF2) and the vascular endothelial growth factor (VEGFA) [20,118,161,162,163,164,165]. Furthermore, the functioning of genes located on the other chromosomes involved in the translocations with the 12q14-15 region may be also affected. Described cases include the genes of a DNA repair enzyme RAD51B/RAD51L1 (14q24), G2/M cell cycle checkpoint regulator CCNB1IP1/HEI10 (14q11), cytochrome C oxidase subunit COX6C (8q22), cytokine response regulator TRAF3IP2 (6q21) and Golgi complex proteins NAA11 (4q21) and COG5 (7q22-31) [166,167,168,169,170,171,172]. At the same time, the rapid progression of a tumour, and, therefore, a possible manifestation of genetic instability is hampered in HMGA2-ULs by the activity of pathways associated with autophagy and cellular senescence [158,165].

5.2.2. 6p21 Rearrangements

As another distinctive trait of ULs, rearrangements in chromosome region 6p21 are associated with the damage and overexpression of the HMGA1/HMGI-C gene, which emphasises the contribution of the HMGA protein family to tumorigenesis [173,174]. Similar to HMGA2, increased HMGA1 expression is mediated by the hypomethylation of CpG islands at the promoter gene region and may present in a UL without 6p21 rearrangements [155,175,176]. HMGA1-ULs are characterised by the activation of the PLAG1 proto-oncogene and PAPPA2 (pappalysine), the insulin-like growth factor bioavailability regulator, and inhibited expression of SMOC2, a gene involved in angiogenesis [118,155].

5.2.3. 10q22 Rearrangements

Apart from locus 14q24 (the RAD51B gene), chromosome region 10q22 is another frequent translocation partner for regions 12q14-15 and 6p21 [136,177]. The rearrangements of this region in ULs affect the histone acetyltransferase gene, KAT6B/MORF/MYST4 (lysine acetyltransferase 6B), which reflects the involvement of anomalous chromatin regulation in UL development [178]. Whereas the range of KAT6B abnormalities and their impact on clinical and morphological UL characteristics have barely been studied, there are data suggesting that chimeric transcripts of KAT6B-KANSL1 (17q21) may be discovered in fast-growing cellular ULs, although they are most typical for malignant uterine tumours [179,180].

5.2.4. Deletions in the Long Arm of Chromosome 7

Detected in 17–50% of karyotypically abnormal ULs, deletions in the long arm of chromosome 7 (7q) are also among the most frequent chromosomal rearrangements in ULs [136,138,181]. In culture conditions, UL cells with a deletion in 7q undergo negative selection revealing their coexistence with the 46,XX cells within the first few passages, enabling researchers to suggest the secondary nature of this abnormality early on [139,181]. The significance of deletions in 7q for the UL karyotype evolution is confirmed by their recurrent emergence as independent clonal events or secondary emergence in cells with pre-existing chromosome abnormalities [134,136,182,183]. There is also a documented case of a UL with deletions in the long arms of both chromosome 7 homologues [184].
While ULs have been found to feature deletions along the entire long arm of chromosome 7, they most frequently affect region 7q22 [185,186,187,188,189]. This region is characterised by an extremely high gene density, considerably complicating the search for mechanisms associated with UL pathogenesis. The genes located in 7q22 are involved in a wide range of regulatory processes, including collagen production, various signalling pathways, and DNA repair [183]. The molecular studies of ULs harbouring deletions, translocations and inversions involving region 7q22 have demonstrated losses and damage in gene CUX1/CUTL1 (cut-like homeobox 1), which could serve as a haploinsufficient tumour suppressor [118,190,191]. Deletions in the origin recognition complex protein gene ORC5/ORC5L and the transmembrane protein gene LHFPL3 are also presumed to be of significance [192,193,194]. Furthermore, deletions in 7q are associated with changes in the expression of several genes located in the region: LMTK2 (lemur tyrosine kinase 2), COPS6 (COP9 signalosome subunit 6), CUX1, MLL5/KMT2E (lysine methyltransferase 2E), LHFPL3 (LHFPL tetraspan subfamily member 3), LAMB1 (laminin subunit beta 1) and more [118,194,195,196].
An important characteristic of ULs with deletions in 7q (7q-ULs) is their similarity to karyotypically normal ULs in terms of size and response to gonadotropin-releasing hormone agonist therapy (as opposed to ULs with rearrangements in 12q14-15 region) [197]. The mosaic deletions in 7q and the lack of unique clinical and morphological traits in this group of ULs attest to the high possibility of these rearrangements occurring as “passenger” genetic aberrations, without any impact on the UL growth potential. The possible exception is cases of secondary deletions in 7q in cells with pre-existing chromosome abnormalities. Thus, the emergence of a deletion in 7q in cells with a t(12;14) translocation offers them a proliferative advantage both in vivo and in vitro, not only on karyotypically normal cells but also on the initial population of cells carrying the translocation [134].

5.2.5. Chromosome 1 Rearrangements

Various chromosome 1 rearrangements, which may occur in both karyotypically normal and karyotypically abnormal cells, also significantly contribute to the clonal UL evolution. ULs are characterised by the formation of ring chromosomes 1 and translocations involving both homologues of chromosome 1 or other chromosomes [134,198,199,200,201]. Chromosome 1 rearrangements in ULs are often accompanied by deletions [202]. Partial monosomy 1p occurs more frequently in large cellular ULs—a rare histological UL subtype presumably characterised by a higher risk of malignisation [203,204].

5.2.6. Chromothripsis

Chromothripsis is a special case of complex chromosomal rearrangements characterised by multiple breaks and deletions in one or few chromosomes [205]. Initially, chromothripsis was discovered using mate-pair sequencing in malignant tumours and was associated with high-grade cancer and a poor prognosis [206,207]. However, in 2013, Mehine et al. documented the first case of chromothripsis in UL cells [142]. Based on various estimates, this variant of complex chromosomal rearrangements occurs in 20–41.7% of ULs [97,142,143].
Compared to malignant tumours, chromothripsis in UL cells is characterised by a lower number of breaks (around twenty or more versus tens to hundreds) and a larger number of chromosomes involved (up to four) [208]. Chromothripsis has been documented in myomas that show no signs of malignancy and have no MED12 or FH mutations. Furthermore, the use of molecular techniques allows for finding chromothripsis in ULs with a previously detected apparently normal karyotype or other chromosomal abnormalities indicating the decrease in the proliferative potential of cells with chromothripsis in vitro [134,143]. Meanwhile, there is a documented case of unbalanced chromothripsis, detected both in cultured and uncultured UL cells [209]. Most likely, the proliferative capacity of UL cells with chromothripsis in culture conditions is determined to a lesser extent by the complexity of chromosomal aberrations and to a greater extent by the genome regions involved in the rearrangement.

5.2.7. Other Chromosomal Abnormalities

Regions involved in chromosomal rearrangements in ULs are located almost throughout the entire genome. Apart from the chromosomes and chromosomal regions described in this section, research papers on UL cytogenetics often mention rearrangements in regions 2p, 2q, 3p, 3q, 5q, 13q, 19q, 22q, Xp, trisomy 12 and monosomy 10 and 22 [134,136,210].

5.3. The Clinical Significance of Somatic Genetic Abnormalities in UL Cells

The high diversity of somatic changes in ULs is meaningful not only for fundamental studies of UL pathogenesis but also in the development of treatment approaches. Unlike malignant neoplasms, the correlation between the UL response to pharmacological therapy and its genotype/karyotype has been barely examined. For the moment, ULs with a translocation t(12;14)(q14-15;q24) have been established to show less shrinkage in response to gonadotropin-releasing hormone agonist therapy than ULs with deletions in 7q or with a normal karyotype [156,197]. The same is true for ULs with MED12 mutations as compared to the “wild-type” ULs [211]. However, MED12-ULs have 4.4 times higher odds of shrinking in response to treatment with the progesterone receptor modulator ulipristal acetate as compared to HMGA2-ULs [212].
Data on genetic aberrations also serve as an auxiliary tool in the clonal origin analysis of ULs and other tumours. Whereas the majority of multiple myomas in the uterus emerge independently from one another, in some cases, different myomas could be clonally linked and could share the same chromosomal abnormalities [139,213,214,215,216]. Molecular studies have shown that such ULs can additionally acquire individual genetic mutations conformed with the branched tumour evolution model [217]. Furthermore, in extremely rare cases, the development of benign tumours clonally linked to the initial myoma is observed not only in the corpus of the uterus but also in remote tissues, primarily in the lungs [218]. Some studies also confirmed shared histological, genetic and cytogenetic profiles between secondary tumours and the primary myoma [219,220,221,222,223,224,225]. Deletions in the long arms of chromosomes 3, 11, 19 and 22 [222,226], FH mutations [227] and the activation of the alternative telomere lengthening mechanism could be associated with the risk of benign UL metastasising [228]. Pathogenesis of benign metastasising (or parasitic) ULs is barely studied. There are assumptions that UL metastases might occur as a complication after laparoscopic surgery due to the morcellation of myomas and even that UL metastasising points towards the underestimated malignant potential of the initial tumour [229,230,231,232]. Regardless of UL metastasising mechanism, this phenomenon raises the need to update UL classification criteria as well as criteria to distinguish benign and malignant tumours.
Areas with UL-specific histology may be present in uterine leiomyosarcomas, attesting to the possibility of UL malignisation [233,234]. Uterine leiomyomas and leiomyosarcomas often feature a similar spectrum of genetic alterations, although leiomyosarcoma undoubtedly demonstrates a more pronounced cytogenetic instability [202,235,236,237]. The fact that warrants special attention is that, whereas deletions in chromosomes 1 and 22 are associated with UL malignisation risk [231,238], chromothripsis is not [142,143,209]. Overall, UL malignisation is a considerably rare event and occurs more typically in histopathological variants [239].
Karyotype abnormalities in UL cells could also lead to false positive results in non-invasive prenatal testing (NIPT), a modern technique for detecting foetal chromosomal abnormalities based on the analysis of circulating cell-free foetal DNA in the mother’s blood [240,241,242]. Pregnant women with ULs run a higher risk of having unbalanced subchromosomal aberrations detected through NIPT—but not trisomy 13, 18 or 21 or sex chromosome aneuploidies. In this category of patients, the most frequently detected abnormality is deletions in 7q, which is determined by their frequent occurrence in ULs [243].
Therefore, the genetic alterations should be taken into account not only in the investigation of UL tumorigenesis mechanisms but also when assessing the risk of its benign metastasising and malignisation, developing efficient treatment approaches and diagnosing other diseases and conditions.

5.4. Speculations on the Origin of Somatic Genetic Abnormalities in UL Cells

The occurrence mechanisms of genetic abnormalities specific to ULs have been investigated in parallel with the discussion about the clonal origin of myomas. As early as in the 1960s, researchers analysed the expressed alleles of glucose-6-phosphate dehydrogenase and established the non-random X chromosome inactivation in myomas, which attested to possible UL monoclonality [244]. The discovery of mosaic chromosomal abnormalities in some ULs brought new relevance to the question of their clonal origin, with the answer provided before long by the inquiry into the androgen receptor gene AR methylation status, which confirmed the non-random inactivation of chromosome X in this group of ULs as well [139]. Since then, the monoclonal origin of ULs has been acknowledged by the vast majority of authors. The accumulation of data confirming the unstable repeatability and discordance of AR methylation analysis results obtained using varied techniques cast a shadow of doubt on the matter [245]. Subsequently, Holdsworth-Carson et al. used the analysis of AR expression (RNA-HUMARA) instead of AR methylation to confirm the monoclonal origin in 23/25 (92%) examined ULs, including for individual tumour cell populations: fibroblasts, smooth muscle cells and endothelial cells [23]. However, the molecular techniques referred to above allow for the analysis of multicellular samples but not individual cells. This may result in misinterpretation of a single cell clone overgrowth as monoclonality and thus the underestimation of minor clones in ULs. Such methodological limitations present an immense challenge for investigating a chain of events that have been driving the UL formation and growth up to the time of analysis.
Since the discovery of MED12 mutations in UL stem cells [29], many researchers began to refer to this genetic aberration as a “driver” of the condition, while actively discussing the potential mechanisms of MED12-associated myometrial cell transformation [13,106,118,246]. However, the analysis of individual MED12-UL populations demonstrated their genetic heterogeneity, providing a new incentive for further investigation of the clonal UL origin [22,109]. MED12-ULs are presumed to be non-monoclonal and capable of recruiting surrounding wild-type cells to support tumour growth [22]. While acknowledging the validity of this hypothesis, we would nevertheless like to highlight the possibility of genetic heterogeneity within the MED12-UL stem cell population, and, therefore, the emergence of MED12 mutation not as a transforming event but in the early stages of tumorigenesis (Figure 2).
As for chromosomal abnormalities, the consensus about their secondary occurrence during tumour growth has held until the present day. The exact timing and mechanisms, however, remain to be established. Not all ULs are cytogenetically stable, with a share of myomas characterised by karyotype evolution [134,136,200]. Minor populations of heteroploid cells have also recently been detected in ULs with an apparently normal karyotype [135], demonstrating that the risk of chromosomal abnormalities persists for ULs in the long-term perspective.
Rearrangements involving region 12q14-15 (primarily balanced translocations and inversions) are presumed to occur in the early stages of UL formation, while deletions in chromosomes 1 and 7 occur in later stages [181,199,200,201]. The start of UL formation is most likely accompanied by global epigenetic changes in stem cells. Actively transcribed genes located on different chromosomes may become topologically closer to each other in the transcription factories, which, in turn, may result in faulty repair of multiple DNA breaks and the emergence of structural rearrangements [247,248]. Data on HMGA1 (6p21) and HMGA2 (12q14-15) overexpression in some ULs without the rearrangements of corresponding chromosome regions [154,175,176], a similar level of HMGA2 expression across multiple myomas in one patient, the localised pattern of its expression in the myometrium of UL patients [132] and the differential gene expression profile in the myometrium of patients with HMGA2- and 7q-ULs speak in favour of this suggestion [162].
Since UL cell proliferation is not coupled with telomerase or ALT (alternative telomere lengthening mechanism) activity, ULs are characterised by shorter telomeres than the myometrium [134,228,249,250,251,252]. Telomere attrition is directly linked to the complex rearrangements (including chromothripsis) and may trigger chromosomal instability in the later stages of UL growth [134,253,254]. Alternatively, the impact of mechanical forces, similar to those occurring during pregnancy, may also serve as a factor triggering cytogenetic aberrations in the cells of developing ULs [255].
Therefore, we are suggesting that, in the period from the beginning of tumour mass formation to its surgical removal, a UL experiences multiple changes in dominating processes that may cause somatic mutations. In the early stages, the more significant factors are a hereditary predisposition, chromatin-level changes, overexpression of individual genes and abnormal hormonal regulation, whereas in the late stages, they give way to telomere shortening, deterioration in DNA repair systems, general metabolic and functional abnormalities in UL cells, the mechanical influence of ECM and adjacent tissues and finally, hormonal and physiological changes in the female body associated with pregnancy, childbirth, lactation and the late reproductive period. Nevertheless, it remains unclear, considering the intercellular genetic heterogeneity of MED12-ULs, whether cytogenetic changes occur in wild-type cells or cells carrying MED12 mutations and how acquired chromosomal abnormalities can change the “developmental programme” of MED12-ULs.

6. Epigenetic Changes Associated with UL Development

An epigenome is a record of various chemical markers to the DNA and histone proteins and molecules involved in gene expression regulation on the pre- and post-transcription levels. The functional role of epigenetic factors goes beyond genome regulation to include changes in chromosome compaction throughout the cell cycle, centromere positioning, ensuring DNA availability for its recognition by repair complexes, intercellular communication and more [208,256]. Errors in epigenetic processes have been described for multiple conditions, including the UL. The aberrant expression of multiple genes in UL cells results from epigenetic changes on all levels, from DNA methylation and histone modifications to non-coding RNAs [257].

6.1. Histone Post-Translational Modifications in ULs

Histones (H1, H2A, H2B, H3, H4) are highly conserved DNA-binding proteins whose primary functions are packaging DNA into nucleosomes and chromatin compaction. Not only do the post-translational modifications (acetylation, methylation, phosphorylation, SUMOylation, ubiquitination, etc.) of amino-acid residues in N- and C-terminal regions of histone proteins ensure the dynamic structure of chromatin, but they are also involved in gene expression regulation, the interaction of DNA with other proteins and the processes of chromosome replication, repair, recombination and disjunction [258]. Acetylation of lysine residues in histones H3 and H4 is a marker of transcriptionally active (open) chromatin. Mono-, di- and trimethylation of lysine residues in histone H3 and H4 can indicate either active (H3K4, H3K36, H3K79) or repressed chromatin (H3K9, H3K27, H4K20). Histone H3 phosphorylation at serine and threonine residues also contributes to chromatin structure regulation and plays an important part in the processes of kinetochore function, mitotic chromosome condensation/segregation and centromere recognition by specific proteins. Furthermore, cross-interaction among different types of histone modifications may be observed [259]. Exploration of the spectrum and functional significance of histone post-translational modifications is a rapidly expanding area of epigenetic research that has gained particular relevance in the context of developmental biology and tumorigenesis.
ULs have been found to feature various changes in histone modification patterns and associated enzymatic systems. ULs are characterised by increased histone deacetylase expression levels as compared to the myometrium [260,261]. In UL cells, histone deacetylase 6 is a positive regulator of oestrogen receptor α gene expression [260]. If a UL is exposed to oestrogen or a combination of oestrogen, progesterone and mifepristone, its histone deacetylase activity reaches higher levels than in the myometrium [262]. Histone deacetylase inhibition in UL cells results in a dose-dependent decrease in the expression of β-catenin, cyclin D1, c-Myc, inhibited growth and proliferation, apoptosis induction, cell cycle arrest and decreased TGF-β3 signalling and ECM production [261,263].
Furthermore, around 70% of ULs feature higher expression of histone methyltransferase EZH2 (the enhancer of zeste homologue 2) and content of H3K27me3 histone than the myometrium [264]. This enzyme is a negative regulator of DNA repair gene activity in ULs. Exposure of UL cells to azacitidine and other EZH2 inhibitors drives the expression of RAD51B, BRCA1 (BRCA1 DNA repair associated) and MSH2 (mutS homolog 2) genes [264,265]. In Eker rats (the animal UL model), neonatal exposure to phytoestrogen genistein decreases EZH2 activity and H3K27me3 content in the uterine tissue, which results in an increased UL incidence and tumour multiplicity later in life [266]. Administration of genistein and cadmium, a “metalloestrogen”, to UL cells is also associated with increased content of H3S10ph histone and higher proliferation level [267,268].
A recent study has documented constitutional variations in the genes YEATS4 and ZNHIT1 encoding chromatin remodelling complex proteins (SRCAP), which predispose to UL development [269]. Somatic mutations of these and other genes encoding SRCAP proteins in ULs result in the loss of histone H2A.Z, an active chromatin marker, and changes in the tumour’s expression profile [269].

6.2. DNA Methylation and Demethylation in ULs

Another gene expression regulation mechanism is the methylation of cytosine at the fifth position (DNA methylation). For example, 5-methylcytosine blocks the binding of transcriptional factors with DNA directly or indirectly through specific protein complexes [270]. The transfer of the CH3 group from S-adenosylmethionine to cytosine is catalysed by DNA methyltransferases. DNA methyltransferase 1 (DNMT1) maintains the methylated state of the DNA in dividing cells by attaching the methyl group to cytosine residues in the newly replicated strand using the parental strand as a template. DNA methyltransferases 3A and 3B (DNMT3A, DNMT3B) catalyse methylation de novo and can methylate cytosine both in hemimethylated and unmethylated DNA [271]. The 5-methylcytosine content in a cell is also determined by the TET (Ten-Eleven Translocation) oxygenase enzymes, the AID/APOBEC deaminating complex (activation-induced deaminase/apolipoprotein B mRNA editing enzyme), TDG (thymine DNA-glycosylase) glycosylases and SMUG1 (selective monofunctional uracil DNA glycosylase), which are involved in the reverse process: DNA demethylation [272].
TET enzymes catalyse the sequential oxidation of 5-methylcytosine to 5-hydroxymethylcytosine, 5-formylcytosine and 5-carboxylcytosine [273,274,275]. The activity of TET oxygenases requires ascorbic acid, Fe(II) and α-ketoglutarate [273]. The inhibition of TET-dependent DNA demethylation may result from either a deficiency of cofactors or the accumulation of 2-hydroxyglutarate, succinate and fumarate in the cell due to a metabolic disorder or mutations (for example, in the fumarate hydratase gene FH) [276,277]. Decreased 5-hydroxymethylcytosine content may also occur as a result of adverse environmental exposure [278]. The AID/APOBEC complex can deaminate 5-methylcytosine to thymine and 5-hydroxymethylcytosine to 5-hydroxymethyluracil [279]. The excision repair of oxygenated and deaminated products of 5-methylcytosine is performed by TDG and SMUG1 DNA glycosylases. SMUG1 converts 5-hydroxymethyluracil to cytosine, whereas TDG catalyses the production of cytosine from thymine, 5-hydroxymethylcytosine, 5-formylcytosine and 5-carboxylcytosine [280,281]. Therefore, the unique DNA methylation profile of a cell is determined by two competing mechanisms: DNA methylation and demethylation. DNA methylation always results from the activity of DNA methyltransferases, whereas DNA demethylation can either be enzymatic or occur via passive loss of the methyl mark during replication in the absence of DNA methyltransferase activity or their low capacity for binding with oxidised forms of 5-methylcytosine [282].
Unlike the myometrium, ULs feature global genome hypomethylation due to decreased expression of de novo DNA methyltransferases DNMT3A and DNMT3B [283]. By contrast, the expression level of maintenance DNA methyltransferase DMNT1 in ULs is either higher or equal to that of the myometrium [283,284,285]. A considerable share of hypomethylated genes in ULs is located on chromosome X (FAM9A, CPXCR1, CXORF45, TAF1, NXF5, VBP1, GABRE, DDX53, FHL1, BRCC3, DMD, GJB1, AP1S2 and PCDH11X) [286]. Hypomethylation/expression activation is also typical of collagen genes (COL4A1, COL4A2 and COL6A3), prolactin, oestrogen receptor α, repair enzyme gene RAD51B and several oncogenes (ATP8B4, CEMIP, ZPMS2-AS1, RIMS2 and TFAP2C) [287,288,289]. Furthermore, the UL has been found to show hypermethylation/inhibited expression of tumour suppressor genes (EFEMP1, FBLN2, ARHGAP10, HTATIP2, DLEC1 and KRT19), RANKL gene encoding tumour-associated cytokine, glycoprotein Neuregulin 1 NRG1 gene, 14-3-3γ protein YWHAG gene, and transcription factor genes SATB2, KLF4, KLF11 and ATF3 [119,289,290,291,292,293]. The distribution of aberrantly methylated CpGs in the chromosomes as well as the expression of genes associated with ECM organization, cell adhesion, angiogenesis, endo- and paracrine signalling pathways differ between ULs with and without MED12 mutations [111] and could also serve as a biomarker for the differential diagnosis of uterine leiomyomas and leiomyosarcomas [292,294].
The DNA methylation patterns described above are more typical of mature tumour cells as the population of stem/progenitor cells in ULs is relatively limited [27,295]. Recent studies have revealed around 10,000 differentially methylated regions, most of which were hypermethylated in UL stem cells in comparison with differentiated ones [295]. In all appearances, the “stem phenotype” of UL cells is determined primarily by the progesterone receptor gene hypermethylation and its inhibited expression. Knocking down this gene in mature UL cells results in a shift of their transcription profile toward non-differentiated cells, decreased ECM production and the activation of genes associated with the cell cycle and proliferation [296]. Hypomethylating agent azacitidine (5-aza-2′-deoxycytidine) activates progesterone receptor expression in UL stem cells and stimulates their differentiation, thus decreasing tumour growth potential [296]. Azacitidine treatment of a primary UL culture decreases tumour cell viability (probably via hypomethylation/activation of tumour-suppressor genes), ECM production and the expression of genes encoding WISP1, c-MYC and MMP7 proteins, which are the targets of the WNT/β-catenin signalling pathway [285].
Global hypomethylation of the UL genome can be also significantly influenced by TET-mediated active DNA demethylation. ULs are characterised by a higher expression of TET1 and TET3 than the myometrium [297]. Meanwhile, the expression of TET1 and TET3 in UL stem cells is lower than in differentiated ones, in full alignment with the differences in their DNA methylation levels [165]. The H19-let-7-TET3 axis is an important expression regulator for the genes associated with the early stages of UL development, signalling pathways and collagen production. In particular, the TET3 protein has the capacity for binding with the promoters of MED12, TGFBR2 (transforming growth factor beta receptor 2) and TSP1 (thrombospondin 1), thus facilitating their demethylation and increased chromatin availability [298]. TET1 and TET3 knockdown or treatment with specific TET inhibitors result in a decrease in both DNA hydroxymethylation levels and UL cell proliferation [297]. TET3 expression in the UL also depends on sex steroid hormones and increases under combined exposure to oestrogen and progesterone [298]. Furthermore, reduced TET1 expression in UL cells is associated with the risk of tumour recurrence [299].
Indeed, 5-Hydroxymethylcytosine is the first product of TET-mediated 5-methylcytosine oxidation and is most widely represented in the genome [300]. Located primarily in the transcriptionally active genome regions, 5-Hydroxymethylcytosine levels vary greatly across tissues and ontogenesis stages [301,302,303,304,305,306]. Low 5-hydroxymethylcytosine content is typical of many malignant tumours and is associated with late stages of the disease and poor prognosis [307,308,309]. By contrast, ULs display higher levels of DNA hydroxymethylation than in the adjacent myometrium or leiomyosarcoma [297]. Finally, 5-Hydroxymethylcytosine content does not differ in ULs with and without MED12 mutations but depends on the menstrual cycle phase: it is higher in the follicular than in the luteal phase [122].

6.3. Non-Coding RNAs

Non-coding RNAs (ncRNAs) are involved in transcription, translation, splicing, lengthening of telomeres, genome imprinting, X chromosome inactivation, intercellular communication and more. ULs have been found to feature dysregulation in several RNA classes, namely microRNAs (miRNAs), circular RNAs (circRNAs) and long ncRNAs (lncRNAs).
MiRNAs are short single-stranded RNA molecules containing 18-25 nucleotides. Changes in miRNA profiles in a UL are associated with the expression of components of DNA demethylation, immune response and cell cycle proteins, with the tumour cell differentiation, EDC exposure, tumour size and the patient’s ethnicity [310,311,312,313,314,315]. The most studied miRNAs in UL belong to the let-7 and miR-29 families. The let-7 miRNA is a negative regulator of HMGA2 mRNA translation, and its expression is reduced in HMGA2-ULs [133]. The let-7 miRNA may also serve as a marker of cellular senescence and inflammation in ULs. Downregulation of the miR-29 miRNA is shown to be responsible for TGF-β3- and oestrogen/progesterone-dependent ECM accumulation in UL cells [316]. High miR-29 content inhibits both the expression of collagen genes and the mechanisms of cell proliferation, invasion and metastasising [317]. Overall, ULs demonstrate expression changes for over 100 miRNAs, often overlapping in their functionality [318].
CircRNAs are stable regulatory RNA molecules covalently closed at the 3′ and 5′ ends. ULs have been found to feature increased expression of 579 and reduced expression of 625 circRNAs as compared to the myometrium [319]. The most significant differences are observed in circRNAs hsa_circ_0083920, hsa_circ_0056686, hsa_circ_0062558, hsa_circ_0020376 and hsa_circ_0043597. In particular, hsa_circ_0056686 circRNA expression is higher in tumour-associated fibroblasts than in fibroblasts isolated from adjacent myometrial tissues and is associated with UL size, the regulation of proliferation, migration and apoptosis and endoplasmic reticulum stress [319,320].
LncRNAs are RNA molecules containing 200 or more nucleotides. Some variants of the HOTAIR long ncRNA gene are associated with the risk of UL development [321]. ULs also display changes in the expression of H19, APTR, XIST, CAR10, CASC15, TCONS_l2_00000923, UC.10, HOTTIP, LINC00890, MEG3, LNCRNA-ATB, MAMDC2-AS1, TSIX, HULC, BX640708, UCA1 and AK023096 lncRNAs as compared to the myometrium [322]. The H19 lncRNA is the key regulator of HMGA2 and MED12 (“UL drivers”) expression [298]. The APTR (Alu-mediated p21 transcriptional regulator) and XIST (X-inactive specific transcript) lncRNAs are involved in the hormone-dependent proliferation of UL cells, the activation of the WNT signalling pathway, collagens and fibronectin gene expression [323,324]. Increased expression of the lnc-AL445665.1-4 lncRNA is typical of multiple but not solitary ULs [325]. The MED12-UL demonstrate changes in the expression of the SRA1 (steroid receptor RNA activator 1) and MIAT (myocardial infarction-associated transcript) lncRNAs, which regulate steroidogenesis and ECM accumulation, respectively [324,326].

6.4. Speculations on the Role of Epigenetic Changes in UL Aetiology

Data presented in 6.1-6.3 demonstrate that epigenetic changes in ULs are highly diverse and could be specific either to the condition at large or to patients, molecular subgroups of UL or even cell populations. On the one hand, the epigenetic features of a cell are fairly stable and specific. On the other hand, they are characterised by plasticity and are subject to changes under the influence of endogenous and exogenous stimuli.
Over the last few years, a considerable quantity of experimental and clinical studies have been issued on the role of endocrine-disrupting chemicals (EDCs), in particular, on xenoestrogens (bisphenol A, diethylstilbestrol, phthalates, etc.) in UL development [267,314,327,328,329,330]. Xenoestrogens can mimic the activity of natural hormones and modulate a wide variety of cellular processes (paracrine signalling, proliferation, apoptosis, DNA repair, DNA methylation, histone protein modifications, and gene expression) [331,332,333]. Thus, prenatal and prepubertal exposure to xenoestrogens increases the risk of developing tumours in adulthood [44,334,335,336,337,338,339]. Whereas the specific mechanism of the delayed EDC effect on UL development is yet to be investigated, epigenetic abnormalities in myometrial stem cells are presumed to play a key part in this process [9]. Consequently, the focus of UL aetiology research has somewhat shifted from genetic to epigenetic aberrations.
The search for a universal event that triggers UL development is considerably complicated by the heterogeneity of molecular features across ULs. For a long time, dependence on sex steroid hormones was considered to be the only common trait of all ULs. However, around three decades ago, Bulun et al. demonstrated that over 90% of myomas are also characterised by an increased expression of aromatase, which converts androgens to oestrogens—up to 25 times as high as in the adjacent myometrium. Aromatase expression was also discovered in the myometrium of 75% of women with UL and 0% of women without ULs [340]. A suggestion was made that localised oestrogen biosynthesis may be of pathological significance in the promotion of UL growth [341]. Subsequently, aromatase overexpression was detected in the endometrium and uterine arteries of women diagnosed with UL, which attests to the condition causing systemic lesions [342,343].
Apart from ULs, aromatase overexpression is also observed in other benign uterine neoplasms: endometriosis and adenomyosis [344,345]. Similar to ULs, aromatase expression is found in the eutopic endometrium of endometriosis patients but not in healthy women [346]. In both myomas and endometriomas, aromatase expression is controlled primarily by the proximal promoter II; a few groups of patients have also shown activity of promoters I.3 and I.4 [340,347,348,349,350]. In a benign neoplastic process in the uterus, the potential stimulant of aromatase expression from promoter II is presumed to be prostaglandin E2 [341,347]. Prostaglandin E2 has a wide range of functions, in particular, acting as a smooth muscle tone regulator, inflammation mediator and immune response component. The start of an inflammatory process, prostaglandin E2 production and accumulation may result from harmful environmental exposure [351] and are associated with benign and malignant tumours and dysmenorrhea [352,353,354,355].
Benign neoplasms of the female reproductive tract share a multitude of molecular features that were addressed in detail by Baranov et al. [356]. According to the authors’ hypothesis, dysregulation of mesenchymal stem cells is the source of both endometrial and myometrial neoplasms. However, the fate of a transformed stem cell is determined by its location in the corpus of the uterus and a combination of factors, including genetic and epigenetic background [356]. We suggest that the induction of aromatase expression in uterine stem cells as a result of adverse environmental exposure or a hereditary predisposition may trigger the neoplastic process and be the basis of UL and endometriosis syntropy.
EDC exposure in the early stages of ontogenesis and abnormal aromatase activation may result in a possibility of prepubertal emergence of areas with a high concentration of oestrogen in uterine tissue. Normally, oestrogens trigger the activity of multiple genes and signalling cascades; however, hyperproduction of oestrogens and their oxidised metabolites (catechols and quinones) may also serve as the transforming factor [357,358,359]. The absence of documented pre-menarche UL cases is explained by the fact that, in addition to oestrogen, a progenitor UL cell also needs progesterone, which is produced in the ovaries, to develop proliferative potential [12,360]. Oestrogen activates progesterone receptor expression, and, therefore, regulates the susceptibility of myometrial and UL cells to progesterone [46,361]. As Omar et al. have recently demonstrated, the myometrium of UL patients contains more progesterone receptors and is characterised by higher progesterone-associated gene activation levels [362], which also speaks in favour of systemic but not localised myometrium lesions in UL patients. Exposure to endogenous and exogenous factors may have a cumulative effect, which determines the highest UL risk in premenopausal women. Meanwhile, the observed diversity in morphologic and molecular features of myomas most likely develops at the stage of tumour promotion as a result of sequential epigenetic and genetic abnormalities in the transformed cells (Figure 2).

7. Conclusions

UL is a neoplastic disease characterised by a high incidence, diverse clinical presentation, pronounced autoregulation and considerable changeability in the tumour’s genome and epigenome, which surprisingly coincides with a low risk of malignisation. It is generally accepted that a UL develops from a single myometrial cell—presumably a stem cell—as a result of its transformation and clonal expansion. In our opinion, however, the clonal origin of UL once again requires serious investigation in the context of modern data and methodological approaches. The search for an event that triggers UL formation is strongly impeded by the broad range and non-unique nature of predisposing factors, their uneven contribution, on the one hand, and the pronounced variability in clinicopathological, morphological, genetic and epigenetic characteristics of ULs, on the other hand. By the time of their detection and surgical removal, myomas have existed in the uterus for months, years or even decades, undergoing sequential developmental stages and exposure to a variety of endogenous and exogenous factors. As a result, every myoma acquires a unique portrait with a pronounced intercellular heterogeneity, which makes identifying the characteristics of the initially transformed cell highly problematic (Figure 1). Unfortunately, the available techniques do not enable us to establish links across all of the aspects that contribute to the individual profile of each UL cell because focusing on one or several characteristics, we inevitably lose sight of the rest (Figure 1).
The key role in the neoplastic transformation belongs to genetic and epigenetic alterations, and in this context, the UL is most likely no exception. The wide variety of such aberrations in ULs can be divided into driver events (MED12 and FH mutations, HMGA2 rearrangements and overexpression, COL4A5-COL4A6 rearrangements) and passenger events (7q deletions, chromosome 1 rearrangements, heteroploidy, etc.), with the latter occurring in later stages and making the greatest contribution to the intratumoral heterogeneity (Figure 2). Driver mutations in MED12, HMGA2 and other genes appear to be emerging in undifferentiated cells at the early stages of UL tumorigenesis and largely define the fate of each myoma. However, MED12-ULs and HMGA12-ULs have also been demonstrated to feature cells without corresponding driver mutations, which attests to the fact that, instead of emerging in the initial transformed myometrial cell, they appeared later in one of its daughter cells. In light of this, we suggest that the prerequisite of a myometrial cell transformation into a UL cell may be an epigenetic alteration. Furthermore, the prerequisite event and the beginning of UL formation could be separated by a considerable stretch of time. Moreover, we cannot rule out the possibility that the potential prerequisite may occur during gametogenesis (Figure 2). In this case, the transgenerational inheritance of specific epigenetic patterns could be of significance. After puberty, the coinciding of a prerequisite condition with the moment of an imbalance between protective and negative factors results in UL formation. Further on, up until the occurrence of driver mutations, the myoma most likely develops through paracrine interactions with surrounding differentiated myometrial cells. It cannot be ruled out that local changes in hormone concentrations in the myometrium, for instance, caused by aromatase activation, or the emergence of cells with hormonal hypersensitivity could play a protumorigenic role, creating pre-conditioned areas. Therefore, the prerequisite for UL development presumably occurs much earlier than the beginning of tumour formation, while the latter is followed by multiple genetic and epigenetic changes, reflecting the UL evolution.

Funding

This research was funded by the Russian Science Foundation, grant number 19-15-00108. A.S.K. is a grantee of the RF President Scholarship.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Kseniia O. Khudadian for her helpful advice during the preparation of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Day Baird, D.; Dunson, D.B.; Hill, M.C.; Cousins, D.; Schectman, J.M. High Cumulative Incidence of Uterine Leiomyoma in Black and White Women: Ultrasound Evidence. Am. J. Obstet. Gynecol. 2003, 188, 100–107. [Google Scholar] [CrossRef]
  2. Bulun, S.E. Uterine Fibroids. N. Engl. J. Med. 2013, 369, 1344–1355. [Google Scholar] [CrossRef] [Green Version]
  3. Al-Hendy, A.; Myers, E.R.; Stewart, E. Uterine Fibroids: Burden and Unmet Medical Need. Semin. Reprod. Med. 2017, 35, 473–480. [Google Scholar] [CrossRef] [Green Version]
  4. Coutinho, L.M.; Assis, W.A.; Spagnuolo-Souza, A.; Reis, F.M. Uterine Fibroids and Pregnancy: How Do They Affect Each Other? Reprod. Sci. Thousand Oaks Calif. 2022, 29, 2145–2151. [Google Scholar] [CrossRef]
  5. Parazzini, F.; Tozzi, L.; Bianchi, S. Pregnancy Outcome and Uterine Fibroids. Best Pract. Res. Clin. Obstet. Gynaecol. 2016, 34, 74–84. [Google Scholar] [CrossRef]
  6. Sato, F.; Mori, M.; Nishi, M.; Kudo, R.; Miyake, H. Familial Aggregation of Uterine Myomas in Japanese Women. J. Epidemiol. 2002, 12, 249–253. [Google Scholar] [CrossRef] [Green Version]
  7. Day Baird, D.; Dunson, D.B. Why Is Parity Protective for Uterine Fibroids? Epidemiology 2003, 14, 247–250. [Google Scholar] [CrossRef]
  8. Stewart, E.; Cookson, C.; Gandolfo, R.; Schulze-Rath, R. Epidemiology of Uterine Fibroids: A Systematic Review. BJOG Int. J. Obstet. Gynaecol. 2017, 124, 1501–1512. [Google Scholar] [CrossRef] [Green Version]
  9. Yang, Q.; Ciebiera, M.; Bariani, M.V.; Ali, M.; Elkafas, H.; Boyer, T.G.; Al-Hendy, A. Comprehensive Review of Uterine Fibroids: Developmental Origin, Pathogenesis, and Treatment. Endocr. Rev. 2022, 43, 678–719. [Google Scholar] [CrossRef]
  10. Commandeur, A.E.; Styer, A.K.; Teixeira, J.M. Epidemiological and Genetic Clues for Molecular Mechanisms Involved in Uterine Leiomyoma Development and Growth. Hum. Reprod. Update 2015, 21, 593–615. [Google Scholar] [CrossRef] [Green Version]
  11. Pavone, D.; Clemenza, S.; Sorbi, F.; Fambrini, M.; Petraglia, F. Epidemiology and Risk Factors of Uterine Fibroids. Best Pract. Res. Clin. Obstet. Gynaecol. 2018, 46, 3–11. [Google Scholar] [CrossRef]
  12. Moravek, M.B.; Bulun, S.E. Endocrinology of Uterine Fibroids: Steroid Hormones, Stem Cells, and Genetic Contribution. Curr. Opin. Obstet. Gynecol. 2015, 27, 276–283. [Google Scholar] [CrossRef]
  13. Baranov, V.S.; Osinovskaya, N.S.; Yarmolinskaya, M.I. Pathogenomics of Uterine Fibroids Development. Int. J. Mol. Sci. 2019, 20, 6151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Flake, G.P.; Moore, A.B.; Sutton, D.; Flagler, N.; Clayton, N.; Kissling, G.E.; Hall, B.W.; Horton, J.; Walmer, D.; Robboy, S.J.; et al. The Life Cycle of the Uterine Fibroid Myocyte. Curr. Obstet. Gynecol. Rep. 2018, 7, 97–105. [Google Scholar] [CrossRef] [PubMed]
  15. Miura, S.; Khan, K.N.; Kitajima, M.; Hiraki, K.; Moriyama, S.; Masuzaki, H.; Samejima, T.; Fujishita, A.; Ishimaru, T. Differential Infiltration of Macrophages and Prostaglandin Production by Different Uterine Leiomyomas. Hum. Reprod. Oxf. Engl. 2006, 21, 2545–2554. [Google Scholar] [CrossRef] [Green Version]
  16. Holdsworth-Carson, S.J.; Zhao, D.; Cann, L.; Bittinger, S.; Nowell, C.J.; Rogers, P.A.W. Differences in the Cellular Composition of Small versus Large Uterine Fibroids. Reprod. Camb. Engl. 2016, 152, 467–480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Othman, E.R.; Elgamal, D.A.; Refaiy, A.M.; Abdelaal, I.I.; Abdel-Mola, A.F.; Al-Hendy, A. Identification and Potential Role of Telocytes in Human Uterine Leiomyoma. Contracept. Reprod. Med. 2016, 1, 12. [Google Scholar] [CrossRef] [Green Version]
  18. Giray, B.; Esim-Buyukbayrak, E.; Hallac-Keser, S.; Karageyim-Karsidag, A.Y.; Turkgeldi, A. Comparison of Nerve Fiber Density between Patients with Uterine Leiomyoma with and without Pain: A Prospective Clinical Study. Geburtshilfe Frauenheilkd. 2018, 78, 407–411. [Google Scholar] [CrossRef] [Green Version]
  19. Aleksandrovych, V.; Kurnik-Łucka, M.; Bereza, T.; Białas, M.; Pasternak, A.; Cretoiu, D.; Walocha, J.A.; Gil, K. The Autonomic Innervation and Uterine Telocyte Interplay in Leiomyoma Formation. Cell Transplant. 2019, 28, 619–629. [Google Scholar] [CrossRef]
  20. Li, Y.; Qiang, W.; Griffin, B.B.; Gao, T.; Chakravarti, D.; Bulun, S.; Kim, J.J.; Wei, J.-J. HMGA2-Mediated Tumorigenesis through Angiogenesis in Leiomyoma. Fertil. Steril. 2020, 114, 1085–1096. [Google Scholar] [CrossRef]
  21. Zannotti, A.; Greco, S.; Pellegrino, P.; Giantomassi, F.; Delli Carpini, G.; Goteri, G.; Ciavattini, A.; Ciarmela, P. Macrophages and Immune Responses in Uterine Fibroids. Cells 2021, 10, 982. [Google Scholar] [CrossRef]
  22. Goad, J.; Rudolph, J.; Zandigohar, M.; Tae, M.; Dai, Y.; Wei, J.-J.; Bulun, S.E.; Chakravarti, D.; Rajkovic, A. Single-Cell Sequencing Reveals Novel Cellular Heterogeneity in Uterine Leiomyomas. Hum. Reprod. Oxf. Engl. 2022, 37, 2334–2349. [Google Scholar] [CrossRef] [PubMed]
  23. Holdsworth-Carson, S.J.; Zaitseva, M.; Vollenhoven, B.J.; Rogers, P.A.W. Clonality of Smooth Muscle and Fibroblast Cell Populations Isolated from Human Fibroid and Myometrial Tissues. Mol. Hum. Reprod. 2014, 20, 250–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Ono, M.; Maruyama, T.; Masuda, H.; Kajitani, T.; Nagashima, T.; Arase, T.; Ito, M.; Ohta, K.; Uchida, H.; Asada, H.; et al. Side Population in Human Uterine Myometrium Displays Phenotypic and Functional Characteristics of Myometrial Stem Cells. Proc. Natl. Acad. Sci. USA 2007, 104, 18700–18705. [Google Scholar] [CrossRef] [Green Version]
  25. Gálvez, B.G.; Martín, N.S.; Salama-Cohen, P.; Lazcano, J.J.; Coronado, M.J.; Lamelas, M.L.; Alvarez-Barrientes, A.; Eiró, N.; Vizoso, F.; Rodríguez, C. An Adult Myometrial Pluripotential Precursor That Promotes Healing of Damaged Muscular Tissues. Vivo Athens Greece 2010, 24, 431–441. [Google Scholar]
  26. Mas, A.; Nair, S.; Laknaur, A.; Simón, C.; Diamond, M.P.; Al-Hendy, A. Stro-1/CD44 as Putative Human Myometrial and Fibroid Stem Cell Markers. Fertil. Steril. 2015, 104, 225–234. [Google Scholar] [CrossRef] [Green Version]
  27. Chang, H.L.; Senaratne, T.N.; Zhang, L.; Szotek, P.P.; Stewart, E.; Dombkowski, D.; Preffer, F.; Donahoe, P.K.; Teixeira, J. Uterine Leiomyomas Exhibit Fewer Stem/Progenitor Cell Characteristics When Compared with Corresponding Normal Myometrium. Reprod. Sci. Thousand Oaks Calif. 2010, 17, 158–167. [Google Scholar] [CrossRef] [Green Version]
  28. Mas, A.; Cervelló, I.; Gil-Sanchis, C.; Faus, A.; Ferro, J.; Pellicer, A.; Simón, C. Identification and Characterization of the Human Leiomyoma Side Population as Putative Tumor-Initiating Cells. Fertil. Steril. 2012, 98, 741–751. [Google Scholar] [CrossRef]
  29. Ono, M.; Qiang, W.; Serna, V.A.; Yin, P.; Coon, J.S.; Navarro, A.; Monsivais, D.; Kakinuma, T.; Dyson, M.; Druschitz, S.; et al. Role of Stem Cells in Human Uterine Leiomyoma Growth. PLoS ONE 2012, 7, e36935. [Google Scholar] [CrossRef]
  30. Orciani, M.; Caffarini, M.; Biagini, A.; Lucarini, G.; Delli Carpini, G.; Berretta, A.; Di Primio, R.; Ciavattini, A. Chronic Inflammation May Enhance Leiomyoma Development by the Involvement of Progenitor Cells. Stem Cells Int. 2018, 2018, 1716246. [Google Scholar] [CrossRef]
  31. Prusinski Fernung, L.E.; Al-Hendy, A.; Yang, Q. A Preliminary Study: Human Fibroid Stro-1+/CD44+ Stem Cells Isolated From Uterine Fibroids Demonstrate Decreased DNA Repair and Genomic Integrity Compared to Adjacent Myometrial Stro-1+/CD44+ Cells. Reprod. Sci. Thousand Oaks Calif. 2019, 26, 619–638. [Google Scholar] [CrossRef]
  32. Ono, M.; Yin, P.; Navarro, A.; Moravek, M.B.; Coon, J.S.; Druschitz, S.A.; Serna, V.A.; Qiang, W.; Brooks, D.C.; Malpani, S.S.; et al. Paracrine Activation of WNT/β-Catenin Pathway in Uterine Leiomyoma Stem Cells Promotes Tumor Growth. Proc. Natl. Acad. Sci. USA 2013, 110, 17053–17058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Ono, M.; Kajitani, T.; Uchida, H.; Arase, T.; Oda, H.; Uchida, S.; Ota, K.; Nagashima, T.; Masuda, H.; Miyazaki, K.; et al. CD34 and CD49f Double-Positive and Lineage Marker-Negative Cells Isolated from Human Myometrium Exhibit Stem Cell-Like Properties Involved in Pregnancy-Induced Uterine Remodeling. Biol. Reprod. 2015, 93, 37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Popescu, L.M.; Faussone-Pellegrini, M.-S. TELOCYTES—A Case of Serendipity: The Winding Way from Interstitial Cells of Cajal (ICC), via Interstitial Cajal-Like Cells (ICLC) to TELOCYTES. J. Cell. Mol. Med. 2010, 14, 729–740. [Google Scholar] [CrossRef] [Green Version]
  35. Varga, I.; Klein, M.; Urban, L.; Danihel, L.; Polak, S.; Danihel, L. Recently Discovered Interstitial Cells “Telocytes” as Players in the Pathogenesis of Uterine Leiomyomas. Med. Hypotheses 2018, 110, 64–67. [Google Scholar] [CrossRef]
  36. Aleksandrovych, V.; Gil, A.; Wrona, A. Sex Steroid Hormone Receptors of Telocytes—Potential Key Role in Leiomyoma Development. Folia Med. Cracov. 2020, 60, 81–95. [Google Scholar] [CrossRef] [PubMed]
  37. Moravek, M.B.; Yin, P.; Ono, M.; Coon, J.S.; Dyson, M.T.; Navarro, A.; Marsh, E.E.; Chakravarti, D.; Kim, J.J.; Wei, J.-J.; et al. Ovarian Steroids, Stem Cells and Uterine Leiomyoma: Therapeutic Implications. Hum. Reprod. Update 2015, 21, 1–12. [Google Scholar] [CrossRef] [Green Version]
  38. Sumitani, H.; Shozu, M.; Segawa, T.; Murakami, K.; Yang, H.J.; Shimada, K.; Inoue, M. In Situ Estrogen Synthesized by Aromatase P450 in Uterine Leiomyoma Cells Promotes Cell Growth Probably via an Autocrine/Intracrine Mechanism. Endocrinology 2000, 141, 3852–3861. [Google Scholar] [CrossRef]
  39. Nelson, L.R.; Bulun, S.E. Estrogen Production and Action. J. Am. Acad. Dermatol. 2001, 45, S116–S124. [Google Scholar] [CrossRef]
  40. Yang, Y.; He, Y.; Zeng, Q.; Li, S. Association of Body Size and Body Fat Distribution with Uterine Fibroids among Chinese Women. J. Womens Health 2002 2014, 23, 619–626. [Google Scholar] [CrossRef]
  41. Ciebiera, M.; Włodarczyk, M.; Słabuszewska-Jóźwiak, A.; Nowicka, G.; Jakiel, G. Influence of Vitamin D and Transforming Growth Factor Β3 Serum Concentrations, Obesity, and Family History on the Risk for Uterine Fibroids. Fertil. Steril. 2016, 106, 1787–1792. [Google Scholar] [CrossRef] [Green Version]
  42. Sun, K.; Xie, Y.; Zhao, N.; Li, Z. A Case-Control Study of the Relationship between Visceral Fat and Development of Uterine Fibroids. Exp. Ther. Med. 2019, 18, 404–410. [Google Scholar] [CrossRef] [Green Version]
  43. Qin, H.; Lin, Z.; Vásquez, E.; Luan, X.; Guo, F.; Xu, L. Association between Obesity and the Risk of Uterine Fibroids: A Systematic Review and Meta-Analysis. J. Epidemiol. Community Health 2021, 75, 197–204. [Google Scholar] [CrossRef] [PubMed]
  44. Yang, Q.; Diamond, M.P.; Al-Hendy, A. Early Life Adverse Environmental Exposures Increase the Risk of Uterine Fibroid Development: Role of Epigenetic Regulation. Front. Pharmacol. 2016, 7, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Lamminen, S.; Rantala, I.; Helin, H.; Rorarius, M.; Tuimala, R. Proliferative Activity of Human Uterine Leiomyoma Cells as Measured by Automatic Image Analysis. Gynecol. Obstet. Investig. 1992, 34, 111–114. [Google Scholar] [CrossRef] [PubMed]
  46. Ishikawa, H.; Ishi, K.; Serna, V.A.; Kakazu, R.; Bulun, S.E.; Kurita, T. Progesterone Is Essential for Maintenance and Growth of Uterine Leiomyoma. Endocrinology 2010, 151, 2433–2442. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Kawaguchi, K.; Fujii, S.; Konishi, I.; Nanbu, Y.; Nonogaki, H.; Mori, T. Mitotic Activity in Uterine Leiomyomas during the Menstrual Cycle. Am. J. Obstet. Gynecol. 1989, 160, 637–641. [Google Scholar] [CrossRef]
  48. Wu, X.; Blanck, A.; Olovsson, M.; Möller, B.; Favini, R.; Lindblom, B. Apoptosis, Cellular Proliferation and Expression of P53 in Human Uterine Leiomyomas and Myometrium during the Menstrual Cycle and after Menopause. Acta Obstet. Gynecol. Scand. 2000, 79, 397–404. [Google Scholar]
  49. Palomba, S.; Sena, T.; Morelli, M.; Noia, R.; Zullo, F.; Mastrantonio, P. Effect of Different Doses of Progestin on Uterine Leiomyomas in Postmenopausal Women. Eur. J. Obstet. Gynecol. Reprod. Biol. 2002, 102, 199–201. [Google Scholar] [CrossRef]
  50. Patel, A.; Malik, M.; Britten, J.; Cox, J.; Catherino, W.H. Mifepristone Inhibits Extracellular Matrix Formation in Uterine Leiomyoma. Fertil. Steril. 2016, 105, 1102–1110. [Google Scholar] [CrossRef] [Green Version]
  51. Voronin, D.; Sotnikova, N.; Rukavishnikov, K.; Malyshkina, A.; Nagornii, S.; Antsiferova, Y. Differential Regulatory Effect of Progesterone on the Proliferation and Apoptosis of Uterine Leiomyoma Tissue Explants and Primary Leiomyoma Cell Cultures. JBRA Assist. Reprod. 2021, 25, 540–548. [Google Scholar] [CrossRef]
  52. Patel, B.; Elguero, S.; Thakore, S.; Dahoud, W.; Bedaiwy, M.; Mesiano, S. Role of Nuclear Progesterone Receptor Isoforms in Uterine Pathophysiology. Hum. Reprod. Update 2015, 21, 155–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kim, J.J.; Sefton, E.C. The Role of Progesterone Signaling in the Pathogenesis of Uterine Leiomyoma. Mol. Cell. Endocrinol. 2012, 358, 223–231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Elkafas, H.; Qiwei, Y.; Al-Hendy, A. Origin of Uterine Fibroids: Conversion of Myometrial Stem Cells to Tumor-Initiating Cells. Semin. Reprod. Med. 2017, 35, 481–486. [Google Scholar] [CrossRef] [PubMed]
  55. Hoekstra, A.V.; Sefton, E.C.; Berry, E.; Lu, Z.; Hardt, J.; Marsh, E.; Yin, P.; Clardy, J.; Chakravarti, D.; Bulun, S.; et al. Progestins Activate the AKT Pathway in Leiomyoma Cells and Promote Survival. J. Clin. Endocrinol. Metab. 2009, 94, 1768–1774. [Google Scholar] [CrossRef] [PubMed]
  56. Al-Hendy, A.; Laknaur, A.; Diamond, M.P.; Ismail, N.; Boyer, T.G.; Halder, S.K. Silencing Med12 Gene Reduces Proliferation of Human Leiomyoma Cells Mediated via Wnt/β-Catenin Signaling Pathway. Endocrinology 2017, 158, 592–603. [Google Scholar] [CrossRef]
  57. El Andaloussi, A.; Al-Hendy, A.; Ismail, N.; Boyer, T.G.; Halder, S.K. Introduction of Somatic Mutation in MED12 Induces Wnt4/β-Catenin and Disrupts Autophagy in Human Uterine Myometrial Cell. Reprod. Sci. Thousand Oaks Calif 2020, 27, 823–832. [Google Scholar] [CrossRef]
  58. Ono, M.; Yin, P.; Navarro, A.; Moravek, M.B.; Coon V, J.S.; Druschitz, S.A.; Gottardi, C.J.; Bulun, S.E. Inhibition of Canonical WNT Signaling Attenuates Human Leiomyoma Cell Growth. Fertil. Steril. 2014, 101, 1441–1449. [Google Scholar] [CrossRef]
  59. Al-Hendy, A.; Diamond, M.P.; El-Sohemy, A.; Halder, S.K. 1,25-Dihydroxyvitamin D3 Regulates Expression of Sex Steroid Receptors in Human Uterine Fibroid Cells. J. Clin. Endocrinol. Metab. 2015, 100, E572–E582. [Google Scholar] [CrossRef] [Green Version]
  60. Corachán, A.; Ferrero, H.; Aguilar, A.; Garcia, N.; Monleon, J.; Faus, A.; Cervelló, I.; Pellicer, A. Inhibition of Tumor Cell Proliferation in Human Uterine Leiomyomas by Vitamin D via Wnt/β-Catenin Pathway. Fertil. Steril. 2019, 111, 397–407. [Google Scholar] [CrossRef] [Green Version]
  61. El Sabeh, M.; Saha, S.K.; Afrin, S.; Borahay, M.A. Simvastatin Inhibits Wnt/β-Catenin Pathway in Uterine Leiomyoma. Endocrinology 2021, 162, bqab211. [Google Scholar] [CrossRef] [PubMed]
  62. Borahay, M.A.; Al-Hendy, A.; Kilic, G.S.; Boehning, D. Signaling Pathways in Leiomyoma: Understanding Pathobiology and Implications for Therapy. Mol. Med. Camb. Mass 2015, 21, 242–256. [Google Scholar] [CrossRef]
  63. Islam, M.S.; Afrin, S.; Singh, B.; Jayes, F.L.; Brennan, J.T.; Borahay, M.A.; Leppert, P.C.; Segars, J.H. Extracellular Matrix and Hippo Signaling as Therapeutic Targets of Antifibrotic Compounds for Uterine Fibroids. Clin. Transl. Med. 2021, 11, e475. [Google Scholar] [CrossRef] [PubMed]
  64. Ko, Y.-A.; Jamaluddin, M.F.B.; Adebayo, M.; Bajwa, P.; Scott, R.J.; Dharmarajan, A.M.; Nahar, P.; Tanwar, P.S. Extracellular Matrix (ECM) Activates β-Catenin Signaling in Uterine Fibroids. Reprod. Camb. Engl. 2018, 155, 61–71. [Google Scholar] [CrossRef] [PubMed]
  65. Wear, L.E. Uterine Myoma as a Hereditary Disease. Lancet Lond. Engl. 1957, 272, 25–26. [Google Scholar] [CrossRef] [PubMed]
  66. Vikhlyaeva, E.M.; Khodzhaeva, Z.S.; Fantschenko, N.D. Familial Predisposition to Uterine Leiomyomas. Int. J. Gynaecol. Obstet. Off. Organ Int. Fed. Gynaecol. Obstet. 1995, 51, 127–131. [Google Scholar] [CrossRef]
  67. Luoto, R.; Kaprio, J.; Rutanen, E.M.; Taipale, P.; Perola, M.; Koskenvuo, M. Heritability and Risk Factors of Uterine Fibroids--the Finnish Twin Cohort Study. Maturitas 2000, 37, 15–26. [Google Scholar] [CrossRef]
  68. Marshall, L.M.; Spiegelman, D.; Barbieri, R.L.; Goldman, M.B.; Manson, J.E.; Colditz, G.A.; Willett, W.C.; Hunter, D.J. Variation in the Incidence of Uterine Leiomyoma among Premenopausal Women by Age and Race. Obstet. Gynecol. 1997, 90, 967–973. [Google Scholar] [CrossRef]
  69. Keaton, J.M.; Jasper, E.A.; Hellwege, J.N.; Jones, S.H.; Torstenson, E.S.; Edwards, T.L.; Velez Edwards, D.R. Evidence That Geographic Variation in Genetic Ancestry Associates with Uterine Fibroids. Hum. Genet. 2021, 140, 1433–1440. [Google Scholar] [CrossRef]
  70. Stewart, E.A.; Nicholson, W.K.; Bradley, L.; Borah, B.J. The Burden of Uterine Fibroids for African-American Women: Results of a National Survey. J. Womens Health 2002 2013, 22, 807–816. [Google Scholar] [CrossRef]
  71. Cha, P.-C.; Takahashi, A.; Hosono, N.; Low, S.-K.; Kamatani, N.; Kubo, M.; Nakamura, Y. A Genome-Wide Association Study Identifies Three Loci Associated with Susceptibility to Uterine Fibroids. Nat. Genet. 2011, 43, 447–450. [Google Scholar] [CrossRef] [PubMed]
  72. Feng, Y.; Zhao, X.; Zhou, C.; Yang, L.; Liu, Y.; Bian, C.; Gou, J.; Lin, X.; Wang, Z.; Zhao, X. The Associations between the Val158Met in the Catechol-O-Methyltransferase (COMT) Gene and the Risk of Uterine Leiomyoma (ULM). Gene 2013, 529, 296–299. [Google Scholar] [CrossRef]
  73. Hellwege, J.N.; Jeff, J.M.; Wise, L.A.; Gallagher, C.S.; Wellons, M.; Hartmann, K.E.; Jones, S.F.; Torstenson, E.S.; Dickinson, S.; Ruiz-Narváez, E.A.; et al. A Multi-Stage Genome-Wide Association Study of Uterine Fibroids in African Americans. Hum. Genet. 2017, 136, 1363–1373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Rafnar, T.; Gunnarsson, B.; Stefansson, O.A.; Sulem, P.; Ingason, A.; Frigge, M.L.; Stefansdottir, L.; Sigurdsson, J.K.; Tragante, V.; Steinthorsdottir, V.; et al. Variants Associating with Uterine Leiomyoma Highlight Genetic Background Shared by Various Cancers and Hormone-Related Traits. Nat. Commun. 2018, 9, 3636. [Google Scholar] [CrossRef] [PubMed]
  75. Välimäki, N.; Kuisma, H.; Pasanen, A.; Heikinheimo, O.; Sjöberg, J.; Bützow, R.; Sarvilinna, N.; Heinonen, H.-R.; Tolvanen, J.; Bramante, S.; et al. Genetic Predisposition to Uterine Leiomyoma Is Determined by Loci for Genitourinary Development and Genome Stability. eLife 2018, 7, e37110. [Google Scholar] [CrossRef]
  76. Edwards, T.L.; Giri, A.; Hellwege, J.N.; Hartmann, K.E.; Stewart, E.A.; Jeff, J.M.; Bray, M.J.; Pendergrass, S.A.; Torstenson, E.S.; Keaton, J.M.; et al. A Trans-Ethnic Genome-Wide Association Study of Uterine Fibroids. Front. Genet. 2019, 10, 511. [Google Scholar] [CrossRef] [Green Version]
  77. Gallagher, C.S.; Mäkinen, N.; Harris, H.R.; Rahmioglu, N.; Uimari, O.; Cook, J.P.; Shigesi, N.; Ferreira, T.; Velez-Edwards, D.R.; Edwards, T.L.; et al. Genome-Wide Association and Epidemiological Analyses Reveal Common Genetic Origins between Uterine Leiomyomata and Endometriosis. Nat. Commun. 2019, 10, 4857. [Google Scholar] [CrossRef] [Green Version]
  78. Masuda, T.; Low, S.-K.; Akiyama, M.; Hirata, M.; Ueda, Y.; Matsuda, K.; Kimura, T.; Murakami, Y.; Kubo, M.; Kamatani, Y.; et al. GWAS of Five Gynecologic Diseases and Cross-Trait Analysis in Japanese. Eur. J. Hum. Genet. EJHG 2020, 28, 95–107. [Google Scholar] [CrossRef] [Green Version]
  79. Sakai, K.; Tanikawa, C.; Hirasawa, A.; Chiyoda, T.; Yamagami, W.; Kataoka, F.; Susumu, N.; Terao, C.; Kamatani, Y.; Takahashi, A.; et al. Identification of a Novel Uterine Leiomyoma GWAS Locus in a Japanese Population. Sci. Rep. 2020, 10, 1197. [Google Scholar] [CrossRef] [Green Version]
  80. Ponomarenko, I.; Reshetnikov, E.; Polonikov, A.; Verzilina, I.; Sorokina, I.; Yermachenko, A.; Dvornyk, V.; Churnosov, M. Candidate Genes for Age at Menarche Are Associated With Uterine Leiomyoma. Front. Genet. 2020, 11, 512940. [Google Scholar] [CrossRef]
  81. Alset, D.; Pokudina, I.O.; Butenko, E.V.; Shkurat, T.P. The Effect of Estrogen-Related Genetic Variants on the Development of Uterine Leiomyoma: Meta-Analysis. Reprod. Sci. Thousand Oaks Calif. 2022, 29, 1921–1929. [Google Scholar] [CrossRef] [PubMed]
  82. Kuznetsova, M.V.; Sogoyan, N.S.; Donnikov, A.J.; Trofimov, D.Y.; Adamyan, L.V.; Mishina, N.D.; Shubina, J.; Zelensky, D.V.; Sukhikh, G.T. Familial Predisposition to Leiomyomata: Searching for Protective Genetic Factors. Biomedicines 2022, 10, 508. [Google Scholar] [CrossRef] [PubMed]
  83. Edwards, T.L.; Hartmann, K.E.; Velez Edwards, D.R. Variants in BET1L and TNRC6B Associate with Increasing Fibroid Volume and Fibroid Type among European Americans. Hum. Genet. 2013, 132, 1361–1369. [Google Scholar] [CrossRef] [Green Version]
  84. Aissani, B.; Zhang, K.; Wiener, H. Follow-up to Genome-Wide Linkage and Admixture Mapping Studies Implicates Components of the Extracellular Matrix in Susceptibility to and Size of Uterine Fibroids. Fertil. Steril. 2015, 103, 528–534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Bray, M.J.; Edwards, T.L.; Wellons, M.F.; Jones, S.H.; Hartmann, K.E.; Velez Edwards, D.R. Admixture Mapping of Uterine Fibroid Size and Number in African American Women. Fertil. Steril. 2017, 108, 1034–1042. [Google Scholar] [CrossRef] [Green Version]
  86. Dzhemlikhanova, L.K.; Efimova, O.A.; Osinovskaya, N.S.; Parfenyev, S.E.; Niauri, D.A.; Sultanov, I.Y.; Malysheva, O.V.; Pendina, A.A.; Shved, N.Y.; Ivashchenko, T.E.; et al. Catechol-O-Methyltransferase Val158Met Polymorphism Is Associated with Increased Risk of Multiple Uterine Leiomyomas Either Positive or Negative for MED12 Exon 2 Mutations. J. Clin. Pathol. 2017, 70, 233–236. [Google Scholar] [CrossRef]
  87. Bray, M.J.; Wellons, M.F.; Jones, S.H.; Torstenson, E.S.; Edwards, T.L.; Velez Edwards, D.R. Transethnic and Race-Stratified Genome-Wide Association Study of Fibroid Characteristics in African American and European American Women. Fertil. Steril. 2018, 110, 737–745. [Google Scholar] [CrossRef] [Green Version]
  88. Launonen, V.; Vierimaa, O.; Kiuru, M.; Isola, J.; Roth, S.; Pukkala, E.; Sistonen, P.; Herva, R.; Aaltonen, L.A. Inherited Susceptibility to Uterine Leiomyomas and Renal Cell Cancer. Proc. Natl. Acad. Sci. USA 2001, 98, 3387–3392. [Google Scholar] [CrossRef] [Green Version]
  89. Tomlinson, I.P.M.; Alam, N.A.; Rowan, A.J.; Barclay, E.; Jaeger, E.E.M.; Kelsell, D.; Leigh, I.; Gorman, P.; Lamlum, H.; Rahman, S.; et al. Germline Mutations in FH Predispose to Dominantly Inherited Uterine Fibroids, Skin Leiomyomata and Papillary Renal Cell Cancer. Nat. Genet. 2002, 30, 406–410. [Google Scholar] [CrossRef]
  90. Lehtonen, R.; Kiuru, M.; Vanharanta, S.; Sjöberg, J.; Aaltonen, L.-M.; Aittomäki, K.; Arola, J.; Butzow, R.; Eng, C.; Husgafvel-Pursiainen, K.; et al. Biallelic Inactivation of Fumarate Hydratase (FH) Occurs in Nonsyndromic Uterine Leiomyomas but Is Rare in Other Tumors. Am. J. Pathol. 2004, 164, 17–22. [Google Scholar] [CrossRef] [Green Version]
  91. Miettinen, M.; Felisiak-Golabek, A.; Wasag, B.; Chmara, M.; Wang, Z.; Butzow, R.; Lasota, J. Fumarase-Deficient Uterine Leiomyomas: An Immunohistochemical, Molecular Genetic, and Clinicopathologic Study of 86 Cases. Am. J. Surg. Pathol. 2016, 40, 1661–1669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Harrison, W.J.; Andrici, J.; Maclean, F.; Madadi-Ghahan, R.; Farzin, M.; Sioson, L.; Toon, C.W.; Clarkson, A.; Watson, N.; Pickett, J.; et al. Fumarate Hydratase-Deficient Uterine Leiomyomas Occur in Both the Syndromic and Sporadic Settings. Am. J. Surg. Pathol. 2016, 40, 599–607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Popp, B.; Erber, R.; Kraus, C.; Vasileiou, G.; Hoyer, J.; Burghaus, S.; Hartmann, A.; Beckmann, M.W.; Reis, A.; Agaimy, A. Targeted Sequencing of FH-Deficient Uterine Leiomyomas Reveals Biallelic Inactivating Somatic Fumarase Variants and Allows Characterization of Missense Variants. Mod. Pathol. Off. J. U. S. Can. Acad. Pathol. Inc 2020, 33, 2341–2353. [Google Scholar] [CrossRef] [PubMed]
  94. Siegler, L.; Erber, R.; Burghaus, S.; Brodkorb, T.; Wachter, D.; Wilkinson, N.; Bolton, J.; Stringfellow, H.; Haller, F.; Beckmann, M.W.; et al. Fumarate Hydratase (FH) Deficiency in Uterine Leiomyomas: Recognition by Histological Features versus Blind Immunoscreening. Virchows Arch. Int. J. Pathol. 2018, 472, 789–796. [Google Scholar] [CrossRef]
  95. Kashtan, C.E. Alport Syndrome. An Inherited Disorder of Renal, Ocular, and Cochlear Basement Membranes. Medicine (Baltimore) 1999, 78, 338–360. [Google Scholar] [CrossRef]
  96. Liu, C.; Dillon, J.; Beavis, A.L.; Liu, Y.; Lombardo, K.; Fader, A.N.; Hung, C.-F.; Wu, T.-C.; Vang, R.; Garcia, J.E.; et al. Prevalence of Somatic and Germline Mutations of Fumarate Hydratase in Uterine Leiomyomas from Young Patients. Histopathology 2020, 76, 354–365. [Google Scholar] [CrossRef]
  97. Mehine, M.; Mäkinen, N.; Heinonen, H.-R.; Aaltonen, L.A.; Vahteristo, P. Genomics of Uterine Leiomyomas: Insights from High-Throughput Sequencing. Fertil. Steril. 2014, 102, 621–629. [Google Scholar] [CrossRef]
  98. Yatsenko, S.A.; Mittal, P.; Wood-Trageser, M.A.; Jones, M.W.; Surti, U.; Edwards, R.P.; Sood, A.K.; Rajkovic, A. Highly Heterogeneous Genomic Landscape of Uterine Leiomyomas by Whole Exome Sequencing and Genome-Wide Arrays. Fertil. Steril. 2017, 107, 457–466. [Google Scholar] [CrossRef] [Green Version]
  99. Shaik, N.A.; Lone, W.G.; Khan, I.A.; Vaidya, S.; Rao, K.P.; Kodati, V.L.; Hasan, Q. Detection of Somatic Mutations and Germline Polymorphisms in Mitochondrial DNA of Uterine Fibroids Patients. Genet. Test. Mol. Biomark. 2011, 15, 537–541. [Google Scholar] [CrossRef]
  100. Mäkinen, N.; Mehine, M.; Tolvanen, J.; Kaasinen, E.; Li, Y.; Lehtonen, H.J.; Gentile, M.; Yan, J.; Enge, M.; Taipale, M.; et al. MED12, the Mediator Complex Subunit 12 Gene, Is Mutated at High Frequency in Uterine Leiomyomas. Science 2011, 334, 252–255. [Google Scholar] [CrossRef]
  101. McGuire, M.M.; Yatsenko, A.; Hoffner, L.; Jones, M.; Surti, U.; Rajkovic, A. Whole Exome Sequencing in a Random Sample of North American Women with Leiomyomas Identifies MED12 Mutations in Majority of Uterine Leiomyomas. PLoS ONE 2012, 7, e33251. [Google Scholar] [CrossRef] [Green Version]
  102. Heinonen, H.-R.; Sarvilinna, N.S.; Sjöberg, J.; Kämpjärvi, K.; Pitkänen, E.; Vahteristo, P.; Mäkinen, N.; Aaltonen, L.A. MED12 Mutation Frequency in Unselected Sporadic Uterine Leiomyomas. Fertil. Steril. 2014, 102, 1137–1142. [Google Scholar] [CrossRef]
  103. Osinovskaya, N.S.; Malysheva, O.V.; Shved, N.Y.; Ivashchenko, T.E.; Sultanov, I.Y.; Efimova, O.A.; Yarmolinskaya, M.I.; Bezhenar, V.F.; Baranov, V.S. Frequency and Spectrum of MED12 Exon 2 Mutations in Multiple Versus Solitary Uterine Leiomyomas From Russian Patients. Int. J. Gynecol. Pathol. Off. J. Int. Soc. Gynecol. Pathol. 2016, 35, 509–515. [Google Scholar] [CrossRef] [PubMed]
  104. Lee, M.; Cheon, K.; Chae, B.; Hwang, H.; Kim, H.-K.; Chung, Y.-J.; Song, J.-Y.; Cho, H.-H.; Kim, J.-H.; Kim, M.-R. Analysis of MED12 Mutation in Multiple Uterine Leiomyomas in South Korean Patients. Int. J. Med. Sci. 2018, 15, 124–128. [Google Scholar] [CrossRef] [Green Version]
  105. He, C.; Nelson, W.; Li, H.; Xu, Y.-D.; Dai, X.-J.; Wang, Y.-X.; Ding, Y.-B.; Li, Y.-P.; Li, T. Frequency of MED12 Mutation in Relation to Tumor and Patient’s Clinical Characteristics: A Meta-Analysis. Reprod. Sci. Thousand Oaks Calif. 2022, 29, 357–365. [Google Scholar] [CrossRef] [PubMed]
  106. Heinonen, H.-R.; Pasanen, A.; Heikinheimo, O.; Tanskanen, T.; Palin, K.; Tolvanen, J.; Vahteristo, P.; Sjöberg, J.; Pitkänen, E.; Bützow, R.; et al. Multiple Clinical Characteristics Separate MED12-Mutation-Positive and -Negative Uterine Leiomyomas. Sci. Rep. 2017, 7, 1015. [Google Scholar] [CrossRef]
  107. Mäkinen, N.; Vahteristo, P.; Kämpjärvi, K.; Arola, J.; Bützow, R.; Aaltonen, L.A. MED12 Exon 2 Mutations in Histopathological Uterine Leiomyoma Variants. Eur. J. Hum. Genet. EJHG 2013, 21, 1300–1303. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Mäkinen, N.; Kämpjärvi, K.; Frizzell, N.; Bützow, R.; Vahteristo, P. Characterization of MED12, HMGA2, and FH Alterations Reveals Molecular Variability in Uterine Smooth Muscle Tumors. Mol. Cancer 2017, 16, 101. [Google Scholar] [CrossRef] [Green Version]
  109. Wu, X.; Serna, V.A.; Thomas, J.; Qiang, W.; Blumenfeld, M.L.; Kurita, T. Subtype-Specific Tumor-Associated Fibroblasts Contribute to the Pathogenesis of Uterine Leiomyoma. Cancer Res. 2017, 77, 6891–6901. [Google Scholar] [CrossRef] [Green Version]
  110. Äyräväinen, A.; Pasanen, A.; Ahvenainen, T.; Heikkinen, T.; Pakarinen, P.; Härkki, P.; Vahteristo, P. Systematic Molecular and Clinical Analysis of Uterine Leiomyomas from Fertile-Aged Women Undergoing Myomectomy. Hum. Reprod. Oxf. Engl. 2020, 35, 2237–2244. [Google Scholar] [CrossRef]
  111. Maekawa, R.; Sato, S.; Tamehisa, T.; Sakai, T.; Kajimura, T.; Sueoka, K.; Sugino, N. Different DNA Methylome, Transcriptome and Histological Features in Uterine Fibroids with and without MED12 Mutations. Sci. Rep. 2022, 12, 8912. [Google Scholar] [CrossRef]
  112. Kämpjärvi, K.; Park, M.J.; Mehine, M.; Kim, N.H.; Clark, A.D.; Bützow, R.; Böhling, T.; Böhm, J.; Mecklin, J.-P.; Järvinen, H.; et al. Mutations in Exon 1 Highlight the Role of MED12 in Uterine Leiomyomas. Hum. Mutat. 2014, 35, 1136–1141. [Google Scholar] [CrossRef]
  113. Di Tommaso, S.; Tinelli, A.; Malvasi, A.; Massari, S. Missense Mutations in Exon 2 of the MED12 Gene Are Involved in IGF-2 Overexpression in Uterine Leiomyoma. Mol. Hum. Reprod. 2014, 20, 1009–1015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Mittal, P.; Shin, Y.-H.; Yatsenko, S.A.; Castro, C.A.; Surti, U.; Rajkovic, A. Med12 Gain-of-Function Mutation Causes Leiomyomas and Genomic Instability. J. Clin. Investig. 2015, 125, 3280–3284. [Google Scholar] [CrossRef]
  115. Turunen, M.; Spaeth, J.M.; Keskitalo, S.; Park, M.J.; Kivioja, T.; Clark, A.D.; Mäkinen, N.; Gao, F.; Palin, K.; Nurkkala, H.; et al. Uterine Leiomyoma-Linked MED12 Mutations Disrupt Mediator-Associated CDK Activity. Cell Rep. 2014, 7, 654–660. [Google Scholar] [CrossRef] [Green Version]
  116. Park, M.J.; Shen, H.; Kim, N.H.; Gao, F.; Failor, C.; Knudtson, J.F.; McLaughlin, J.; Halder, S.K.; Heikkinen, T.A.; Vahteristo, P.; et al. Mediator Kinase Disruption in MED12-Mutant Uterine Fibroids From Hispanic Women of South Texas. J. Clin. Endocrinol. Metab. 2018, 103, 4283–4292. [Google Scholar] [CrossRef] [PubMed]
  117. Park, M.J.; Shen, H.; Spaeth, J.M.; Tolvanen, J.H.; Failor, C.; Knudtson, J.F.; McLaughlin, J.; Halder, S.K.; Yang, Q.; Bulun, S.E.; et al. Oncogenic Exon 2 Mutations in Mediator Subunit MED12 Disrupt Allosteric Activation of Cyclin C-CDK8/19. J. Biol. Chem. 2018, 293, 4870–4882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Mehine, M.; Kaasinen, E.; Heinonen, H.-R.; Mäkinen, N.; Kämpjärvi, K.; Sarvilinna, N.; Aavikko, M.; Vähärautio, A.; Pasanen, A.; Bützow, R.; et al. Integrated Data Analysis Reveals Uterine Leiomyoma Subtypes with Distinct Driver Pathways and Biomarkers. Proc. Natl. Acad. Sci. USA 2016, 113, 1315–1320. [Google Scholar] [CrossRef] [Green Version]
  119. Liu, S.; Yin, P.; Kujawa, S.A.; Coon, J.S.; Okeigwe, I.; Bulun, S.E. Progesterone Receptor Integrates the Effects of Mutated MED12 and Altered DNA Methylation to Stimulate RANKL Expression and Stem Cell Proliferation in Uterine Leiomyoma. Oncogene 2019, 38, 2722–2735. [Google Scholar] [CrossRef] [Green Version]
  120. Hutchinson, A.P.; Yin, P.; Neale, I.; Coon, J.S.; Kujawa, S.A.; Liu, S.; Bulun, S.E. Tryptophan 2,3-Dioxygenase-2 in Uterine Leiomyoma: Dysregulation by MED12 Mutation Status. Reprod. Sci. Thousand Oaks Calif. 2022, 29, 743–749. [Google Scholar] [CrossRef]
  121. Moyo, M.B.; Parker, J.B.; Chakravarti, D. Altered Chromatin Landscape and Enhancer Engagement Underlie Transcriptional Dysregulation in MED12 Mutant Uterine Leiomyomas. Nat. Commun. 2020, 11, 1019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Kol’tsova, A.S.; Pendina, A.A.; Efimova, O.A.; Kaminskaya, A.N.; Tikhonov, A.V.; Osinovskaya, N.S.; Sultanov, I.Y.; Shved, N.Y.; Kakhiani, M.I.; Baranov, V.S. Differential DNA Hydroxymethylation in Human Uterine Leiomyoma Cells Depending on the Phase of Menstrual Cycle and Presence of MED12 Gene Mutations. Bull. Exp. Biol. Med. 2017, 163, 646–649. [Google Scholar] [CrossRef] [PubMed]
  123. Muralimanoharan, S.; Shamby, R.; Stansbury, N.; Schenken, R.; de la Pena Avalos, B.; Javanmardi, S.; Dray, E.; Sung, P.; Boyer, T.G. Aberrant R-Loop-Induced Replication Stress in MED12-Mutant Uterine Fibroids. Sci. Rep. 2022, 12, 6169. [Google Scholar] [CrossRef] [PubMed]
  124. Nadine Markowski, D.; Tadayyon, M.; Bartnitzke, S.; Belge, G.; Maria Helmke, B.; Bullerdiek, J. Cell Cultures in Uterine Leiomyomas: Rapid Disappearance of Cells Carrying MED12 Mutations. Genes Chromosomes Cancer 2014, 53, 317–323. [Google Scholar] [CrossRef]
  125. Luo, N.; Guan, Q.; Zheng, L.; Qu, X.; Dai, H.; Cheng, Z. Estrogen-Mediated Activation of Fibroblasts and Its Effects on the Fibroid Cell Proliferation. Transl. Res. J. Lab. Clin. Med. 2014, 163, 232–241. [Google Scholar] [CrossRef]
  126. Okamoto, T.; Sato, J.D.; Barnes, D.W.; Sato, G.H. Biomedical Advances from Tissue Culture. Cytotechnology 2013, 65, 967–971. [Google Scholar] [CrossRef] [Green Version]
  127. Malik, M.; Britten, J.; Catherino, W.H. Development and Validation of Hormonal Impact of a Mouse Xenograft Model for Human Uterine Leiomyoma. Reprod. Sci. Thousand Oaks Calif. 2020, 27, 1304–1317. [Google Scholar] [CrossRef]
  128. Salas, A.; López, J.; Reyes, R.; Évora, C.; de Oca, F.M.; Báez, D.; Delgado, A.; Almeida, T.A. Organotypic Culture as a Research and Preclinical Model to Study Uterine Leiomyomas. Sci. Rep. 2020, 10, 5212. [Google Scholar] [CrossRef] [Green Version]
  129. Shved, N.; Egorova, A.; Osinovskaya, N.; Kiselev, A. Development of Primary Monolayer Cell Model and Organotypic Model of Uterine Leiomyoma. Methods Protoc. 2022, 5, 16. [Google Scholar] [CrossRef]
  130. Markowski, D.N.; Bartnitzke, S.; Löning, T.; Drieschner, N.; Helmke, B.M.; Bullerdiek, J. MED12 Mutations in Uterine Fibroids--Their Relationship to Cytogenetic Subgroups. Int. J. Cancer 2012, 131, 1528–1536. [Google Scholar] [CrossRef]
  131. Kämpjärvi, K.; Mäkinen, N.; Mehine, M.; Välipakka, S.; Uimari, O.; Pitkänen, E.; Heinonen, H.-R.; Heikkinen, T.; Tolvanen, J.; Ahtikoski, A.; et al. MED12 Mutations and FH Inactivation Are Mutually Exclusive in Uterine Leiomyomas. Br. J. Cancer 2016, 114, 1405–1411. [Google Scholar] [CrossRef] [Green Version]
  132. Galindo, L.J.; Hernández-Beeftink, T.; Salas, A.; Jung, Y.; Reyes, R.; de Oca, F.M.; Hernández, M.; Almeida, T.A. HMGA2 and MED12 Alterations Frequently Co-Occur in Uterine Leiomyomas. Gynecol. Oncol. 2018, 150, 562–568. [Google Scholar] [CrossRef] [PubMed]
  133. Mello, J.B.H.; Barros-Filho, M.C.; Abreu, F.B.; Cirilo, P.D.R.; Domingues, M.a.C.; Pontes, A.; Rogatto, S.R. MicroRNAs Involved in the HMGA2 Deregulation and Its Co-Occurrence with MED12 Mutation in Uterine Leiomyoma. Mol. Hum. Reprod. 2018, 24, 556–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Koltsova, A.S.; Efimova, O.A.; Malysheva, O.V.; Osinovskaya, N.S.; Liehr, T.; Al-Rikabi, A.; Shved, N.Y.; Sultanov, I.Y.; Chiryaeva, O.G.; Yarmolinskaya, M.I.; et al. Cytogenomic Profile of Uterine Leiomyoma: In Vivo vs. In Vitro Comparison. Biomedicines 2021, 9, 1777. [Google Scholar] [CrossRef]
  135. Koltsova, A.S.; Efimova, O.A.; Pendina, A.A.; Chiryaeva, O.G.; Osinovskaya, N.S.; Shved, N.Y.; Yarmolinskaya, M.I.; Polenov, N.I.; Kunitsa, V.V.; Sagurova, Y.M.; et al. Uterine Leiomyomas with an Apparently Normal Karyotype Comprise Minor Heteroploid Subpopulations Differently Represented in Vivo and in Vitro. Cytogenet. Genome Res. 2021, 161, 43–51. [Google Scholar] [CrossRef] [PubMed]
  136. Sandberg, A.A. Updates on the Cytogenetics and Molecular Genetics of Bone and Soft Tissue Tumors: Leiomyoma. Cancer Genet. Cytogenet. 2005, 158, 1–26. [Google Scholar] [CrossRef] [PubMed]
  137. Nilbert, M.; Heim, S.; Mandahl, N.; Flodérus, U.M.; Willén, H.; Mitelman, F. Characteristic Chromosome Abnormalities, Including Rearrangements of 6p, Del(7q), +12, and t(12;14), in 44 Uterine Leiomyomas. Hum. Genet. 1990, 85, 605–611. [Google Scholar] [CrossRef]
  138. Hu, J.; Surti, U. Subgroups of Uterine Leiomyomas Based on Cytogenetic Analysis. Hum. Pathol. 1991, 22, 1009–1016. [Google Scholar] [CrossRef]
  139. Mashal, R.D.; Fejzo, M.L.; Friedman, A.J.; Mitchner, N.; Nowak, R.A.; Rein, M.S.; Morton, C.C.; Sklar, J. Analysis of Androgen Receptor DNA Reveals the Independent Clonal Origins of Uterine Leiomyomata and the Secondary Nature of Cytogenetic Aberrations in the Development of Leiomyomata. Genes Chromosomes Cancer 1994, 11, 1–6. [Google Scholar] [CrossRef]
  140. Hayashi, S.; Miharu, N.; Okamoto, E.; Samura, O.; Hara, T.; Ohama, K. Detection of Chromosomal Abnormalities of Chromosome 12 in Uterine Leiomyoma Using Fluorescence in Situ Hybridization. Jpn. J. Hum. Genet. 1996, 41, 193–202. [Google Scholar] [CrossRef] [Green Version]
  141. Schoenmakers, E.F.; Mols, R.; Wanschura, S.; Kools, P.F.; Geurts, J.M.; Bartnitzke, S.; Bullerdiek, J.; van den Berghe, H.; Van de Ven, W.J. Identification, Molecular Cloning, and Characterization of the Chromosome 12 Breakpoint Cluster Region of Uterine Leiomyomas. Genes Chromosomes Cancer 1994, 11, 106–118. [Google Scholar] [CrossRef] [PubMed]
  142. Mehine, M.; Kaasinen, E.; Mäkinen, N.; Katainen, R.; Kämpjärvi, K.; Pitkänen, E.; Heinonen, H.-R.; Bützow, R.; Kilpivaara, O.; Kuosmanen, A.; et al. Characterization of Uterine Leiomyomas by Whole-Genome Sequencing. N. Engl. J. Med. 2013, 369, 43–53. [Google Scholar] [CrossRef] [PubMed]
  143. Holzmann, C.; Markowski, D.N.; Koczan, D.; Küpker, W.; Helmke, B.M.; Bullerdiek, J. Cytogenetically Normal Uterine Leiomyomas without MED12-Mutations—A Source to Identify Unknown Mechanisms of the Development of Uterine Smooth Muscle Tumors. Mol. Cytogenet. 2014, 7, 88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Heim, S.; Nilbert, M.; Vanni, R.; Floderus, U.M.; Mandahl, N.; Liedgren, S.; Lecca, U.; Mitelman, F. A Specific Translocation, t(12;14)(Q14-15;Q23-24), Characterizes a Subgroup of Uterine Leiomyomas. Cancer Genet. Cytogenet. 1988, 32, 13–17. [Google Scholar] [CrossRef] [PubMed]
  145. Hennig, Y.; Deichert, U.; Bonk, U.; Thode, B.; Bartnitzke, S.; Bullerdiek, J. Chromosomal Translocations Affecting 12q14-15 but Not Deletions of the Long Arm of Chromosome 7 Associated with a Growth Advantage of Uterine Smooth Muscle Cells. Mol. Hum. Reprod. 1999, 5, 1150–1154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Schoenberg Fejzo, M.; Ashar, H.R.; Krauter, K.S.; Powell, W.L.; Rein, M.S.; Weremowicz, S.; Yoon, S.J.; Kucherlapati, R.S.; Chada, K.; Morton, C.C. Translocation Breakpoints Upstream of the HMGIC Gene in Uterine Leiomyomata Suggest Dysregulation of This Gene by a Mechanism Different from That in Lipomas. Genes Chromosomes Cancer 1996, 17, 1–6. [Google Scholar] [CrossRef]
  147. Gattas, G.J.; Quade, B.J.; Nowak, R.A.; Morton, C.C. HMGIC Expression in Human Adult and Fetal Tissues and in Uterine Leiomyomata. Genes Chromosomes Cancer 1999, 25, 316–322. [Google Scholar] [CrossRef]
  148. Parisi, S.; Piscitelli, S.; Passaro, F.; Russo, T. HMGA Proteins in Stemness and Differentiation of Embryonic and Adult Stem Cells. Int. J. Mol. Sci. 2020, 21, 362. [Google Scholar] [CrossRef] [Green Version]
  149. Mansoori, B.; Mohammadi, A.; Ditzel, H.J.; Duijf, P.H.G.; Khaze, V.; Gjerstorff, M.F.; Baradaran, B. HMGA2 as a Critical Regulator in Cancer Development. Genes 2021, 12, 269. [Google Scholar] [CrossRef]
  150. Quade, B.J.; Weremowicz, S.; Neskey, D.M.; Vanni, R.; Ladd, C.; Dal Cin, P.; Morton, C.C. Fusion Transcripts Involving HMGA2 Are Not a Common Molecular Mechanism in Uterine Leiomyomata with Rearrangements in 12q15. Cancer Res. 2003, 63, 1351–1358. [Google Scholar]
  151. Pradhan, B.; Sarvilinna, N.; Matilainen, J.; Aska, E.; Sjöberg, J.; Kauppi, L. Detection and Screening of Chromosomal Rearrangements in Uterine Leiomyomas by Long-Distance Inverse PCR. Genes Chromosomes Cancer 2016, 55, 215–226. [Google Scholar] [CrossRef]
  152. Peng, Y.; Laser, J.; Shi, G.; Mittal, K.; Melamed, J.; Lee, P.; Wei, J.-J. Antiproliferative Effects by Let-7 Repression of High-Mobility Group A2 in Uterine Leiomyoma. Mol. Cancer Res. MCR 2008, 6, 663–673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Klemke, M.; Meyer, A.; Hashemi Nezhad, M.; Belge, G.; Bartnitzke, S.; Bullerdiek, J. Loss of Let-7 Binding Sites Resulting from Truncations of the 3’ Untranslated Region of HMGA2 MRNA in Uterine Leiomyomas. Cancer Genet. Cytogenet. 2010, 196, 119–123. [Google Scholar] [CrossRef]
  154. Klemke, M.; Meyer, A.; Nezhad, M.H.; Bartnitzke, S.; Drieschner, N.; Frantzen, C.; Schmidt, E.H.; Belge, G.; Bullerdiek, J. Overexpression of HMGA2 in Uterine Leiomyomas Points to Its General Role for the Pathogenesis of the Disease. Genes Chromosomes Cancer 2009, 48, 171–178. [Google Scholar] [CrossRef] [PubMed]
  155. George, J.W.; Fan, H.; Johnson, B.; Carpenter, T.J.; Foy, K.K.; Chatterjee, A.; Patterson, A.L.; Koeman, J.; Adams, M.; Madaj, Z.B.; et al. Integrated Epigenome, Exome, and Transcriptome Analyses Reveal Molecular Subtypes and Homeotic Transformation in Uterine Fibroids. Cell Rep. 2019, 29, 4069–4085. [Google Scholar] [CrossRef] [Green Version]
  156. Takahashi, K.; Kawamura, N.; Ishiko, O.; Ogita, S. Shrinkage Effect of Gonadotropin Releasing Hormone Agonist Treatment on Uterine Leiomyomas and t(12;14). Int. J. Oncol. 2002, 20, 279–283. [Google Scholar] [CrossRef]
  157. Markowski, D.N.; Bartnitzke, S.; Belge, G.; Drieschner, N.; Helmke, B.M.; Bullerdiek, J. Cell Culture and Senescence in Uterine Fibroids. Cancer Genet. Cytogenet. 2010, 202, 53–57. [Google Scholar] [CrossRef] [PubMed]
  158. Markowski, D.N.; von Ahsen, I.; Nezhad, M.H.; Wosniok, W.; Helmke, B.M.; Bullerdiek, J. HMGA2 and the P19Arf-TP53-CDKN1A Axis: A Delicate Balance in the Growth of Uterine Leiomyomas. Genes Chromosomes Cancer 2010, 49, 661–668. [Google Scholar] [CrossRef] [PubMed]
  159. Markowski, D.N.; Helmke, B.M.; Belge, G.; Nimzyk, R.; Bartnitzke, S.; Deichert, U.; Bullerdiek, J. HMGA2 and P14Arf: Major Roles in Cellular Senescence of Fibroids and Therapeutic Implications. Anticancer Res. 2011, 31, 753–761. [Google Scholar]
  160. Mas, A.; Cervelló, I.; Fernández-Álvarez, A.; Faus, A.; Díaz, A.; Burgués, O.; Casado, M.; Simón, C. Overexpression of the Truncated Form of High Mobility Group A Proteins (HMGA2) in Human Myometrial Cells Induces Leiomyoma-like Tissue Formation. Mol. Hum. Reprod. 2015, 21, 330–338. [Google Scholar] [CrossRef]
  161. Helmke, B.M.; Markowski, D.N.; Müller, M.H.; Sommer, A.; Müller, J.; Möller, C.; Bullerdiek, J. HMGA Proteins Regulate the Expression of FGF2 in Uterine Fibroids. Mol. Hum. Reprod. 2011, 17, 135–142. [Google Scholar] [CrossRef] [Green Version]
  162. Hodge, J.C.; Kim, T.-M.; Dreyfuss, J.M.; Somasundaram, P.; Christacos, N.C.; Rousselle, M.; Quade, B.J.; Park, P.J.; Stewart, E.A.; Morton, C.C. Expression Profiling of Uterine Leiomyomata Cytogenetic Subgroups Reveals Distinct Signatures in Matched Myometrium: Transcriptional Profilingof the t(12;14) and Evidence in Support of Predisposing Genetic Heterogeneity. Hum. Mol. Genet. 2012, 21, 2312–2329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Klemke, M.; Müller, M.H.; Wosniok, W.; Markowski, D.N.; Nimzyk, R.; Helmke, B.M.; Bullerdiek, J. Correlated Expression of HMGA2 and PLAG1 in Thyroid Tumors, Uterine Leiomyomas and Experimental Models. PLoS ONE 2014, 9, e88126. [Google Scholar] [CrossRef] [PubMed]
  164. Xie, J.; Ubango, J.; Ban, Y.; Chakravarti, D.; Kim, J.J.; Wei, J.-J. Comparative Analysis of AKT and the Related Biomarkers in Uterine Leiomyomas with MED12, HMGA2, and FH Mutations. Genes Chromosomes Cancer 2018, 57, 485–494. [Google Scholar] [CrossRef]
  165. Liu, B.; Chen, G.; He, Q.; Liu, M.; Gao, K.; Cai, B.; Qu, J.; Lin, S.; Geng, A.; Li, S.; et al. An HMGA2-P62-ERα Axis Regulates Uterine Leiomyomas Proliferation. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2020, 34, 10966–10983. [Google Scholar] [CrossRef]
  166. Ingraham, S.E.; Lynch, R.A.; Kathiresan, S.; Buckler, A.J.; Menon, A.G. HREC2, a RAD51-like Gene, Is Disrupted by t(12;14) (Q15;Q24.1) in a Uterine Leiomyoma. Cancer Genet. Cytogenet. 1999, 115, 56–61. [Google Scholar] [CrossRef]
  167. Schoenmakers, E.F.; Huysmans, C.; Van de Ven, W.J. Allelic Knockout of Novel Splice Variants of Human Recombination Repair Gene RAD51B in t(12;14) Uterine Leiomyomas. Cancer Res. 1999, 59, 19–23. [Google Scholar] [PubMed]
  168. Kurose, K.; Mine, N.; Doi, D.; Ota, Y.; Yoneyama, K.; Konishi, H.; Araki, T.; Emi, M. Novel Gene Fusion of COX6C at 8q22-23 to HMGIC at 12q15 in a Uterine Leiomyoma. Genes Chromosomes Cancer 2000, 27, 303–307. [Google Scholar] [CrossRef]
  169. Mine, N.; Kurose, K.; Konishi, H.; Araki, T.; Nagai, H.; Emi, M. Fusion of a Sequence from HEI10 (14q11) to the HMGIC Gene at 12q15 in a Uterine Leiomyoma. Jpn. J. Cancer Res. Gann 2001, 92, 135–139. [Google Scholar] [CrossRef]
  170. Mine, N.; Kurose, K.; Nagai, H.; Doi, D.; Ota, Y.; Yoneyama, K.; Konishi, H.; Araki, T.; Emi, M. Gene Fusion Involving HMGIC Is a Frequent Aberration in Uterine Leiomyomas. J. Hum. Genet. 2001, 46, 408–412. [Google Scholar] [CrossRef]
  171. Velagaleti, G.V.N.; Tonk, V.S.; Hakim, N.M.; Wang, X.; Zhang, H.; Erickson-Johnson, M.R.; Medeiros, F.; Oliveira, A.M. Fusion of HMGA2 to COG5 in Uterine Leiomyoma. Cancer Genet. Cytogenet. 2010, 202, 11–16. [Google Scholar] [CrossRef] [PubMed]
  172. Hodgson, A.; Swanson, D.; Tang, S.; Dickson, B.C.; Turashvili, G. Gene Fusions Characterize a Subset of Uterine Cellular Leiomyomas. Genes Chromosomes Cancer 2020. [Google Scholar] [CrossRef] [PubMed]
  173. Kazmierczak, B.; Bol, S.; Wanschura, S.; Bartnitzke, S.; Bullerdiek, J. PAC Clone Containing the HMGI(Y) Gene Spans the Breakpoint of a 6p21 Translocation in a Uterine Leiomyoma Cell Line. Genes Chromosomes Cancer 1996, 17, 191–193. [Google Scholar] [CrossRef]
  174. Nezhad, M.H.; Drieschner, N.; Helms, S.; Meyer, A.; Tadayyon, M.; Klemke, M.; Belge, G.; Bartnitzke, S.; Burchardt, K.; Frantzen, C.; et al. 6p21 Rearrangements in Uterine Leiomyomas Targeting HMGA1. Cancer Genet. Cytogenet. 2010, 203, 247–252. [Google Scholar] [CrossRef] [PubMed]
  175. Williams, A.J.; Powell, W.L.; Collins, T.; Morton, C.C. HMGI(Y) Expression in Human Uterine Leiomyomata. Involvement of Another High-Mobility Group Architectural Factor in a Benign Neoplasm. Am. J. Pathol. 1997, 150, 911–918. [Google Scholar] [PubMed]
  176. Tallini, G.; Vanni, R.; Manfioletti, G.; Kazmierczak, B.; Faa, G.; Pauwels, P.; Bullerdiek, J.; Giancotti, V.; Van Den Berghe, H.; Dal Cin, P. HMGI-C and HMGI(Y) Immunoreactivity Correlates with Cytogenetic Abnormalities in Lipomas, Pulmonary Chondroid Hamartomas, Endometrial Polyps, and Uterine Leiomyomas and Is Compatible with Rearrangement of the HMGI-C and HMGI(Y) Genes. Lab. Investig. J. Tech. Methods Pathol. 2000, 80, 359–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Ozisik, Y.Y.; Meloni, A.M.; Surti, U.; Sandberg, A.A. Involvement of 10q22 in Leiomyoma. Cancer Genet. Cytogenet. 1993, 69, 132–135. [Google Scholar] [CrossRef] [PubMed]
  178. Moore, S.D.P.; Herrick, S.R.; Ince, T.A.; Kleinman, M.S.; Dal Cin, P.; Morton, C.C.; Quade, B.J. Uterine Leiomyomata with t(10;17) Disrupt the Histone Acetyltransferase MORF. Cancer Res. 2004, 64, 5570–5577. [Google Scholar] [CrossRef] [Green Version]
  179. Ainsworth, A.J.; Dashti, N.K.; Mounajjed, T.; Fritchie, K.J.; Davila, J.; Mopuri, R.; Jackson, R.A.; Halling, K.C.; Bakkum-Gamez, J.N.; Schoolmeester, J.K. Leiomyoma with KAT6B-KANSL1 Fusion: Case Report of a Rapidly Enlarging Uterine Mass in a Postmenopausal Woman. Diagn. Pathol. 2019, 14, 32. [Google Scholar] [CrossRef] [Green Version]
  180. Agaimy, A.; Clarke, B.A.; Kolin, D.L.; Lee, C.-H.; Lee, J.-C.; McCluggage, W.G.; Pöschke, P.; Stoehr, R.; Swanson, D.; Turashvili, G.; et al. Recurrent KAT6B/A::KANSL1 Fusions Characterize a Potentially Aggressive Uterine Sarcoma Morphologically Overlapping With Low-Grade Endometrial Stromal Sarcoma. Am. J. Surg. Pathol. 2022, 46, 1298–1308. [Google Scholar] [CrossRef]
  181. Xing, Y.P.; Powell, W.L.; Morton, C.C. The Del(7q) Subgroup in Uterine Leiomyomata: Genetic and Biologic Characteristics. Further Evidence for the Secondary Nature of Cytogenetic Abnormalities in the Pathobiology of Uterine Leiomyomata. Cancer Genet. Cytogenet. 1997, 98, 69–74. [Google Scholar] [CrossRef]
  182. Ozisik, Y.Y.; Meloni, A.M.; Powell, M.; Surti, U.; Sandberg, A.A. Chromosome 7 Biclonality in Uterine Leiomyoma. Cancer Genet. Cytogenet. 1993, 67, 59–64. [Google Scholar] [CrossRef] [PubMed]
  183. Medikare, V.; Kandukuri, L.R.; Ananthapur, V.; Deenadayal, M.; Nallari, P. The Genetic Bases of Uterine Fibroids; a Review. J. Reprod. Infertil. 2011, 12, 181–191. [Google Scholar] [PubMed]
  184. Kingsley, K.L.; Meloni, A.M.; Peier, A.M.; Sandberg, A.A.; Surti, U. Deletion of Both Chromosome 7 Homologues in Leiomyoma. Cancer Genet. Cytogenet. 1995, 81, 99–100. [Google Scholar] [CrossRef]
  185. Sargent, M.S.; Weremowicz, S.; Rein, M.S.; Morton, C.C. Translocations in 7q22 Define a Critical Region in Uterine Leiomyomata. Cancer Genet. Cytogenet. 1994, 77, 65–68. [Google Scholar] [CrossRef] [PubMed]
  186. Ishwad, C.S.; Ferrell, R.E.; Davare, J.; Meloni, A.M.; Sandberg, A.A.; Surti, U. Molecular and Cytogenetic Analysis of Chromosome 7 in Uterine Leiomyomas. Genes Chromosomes Cancer 1995, 14, 51–55. [Google Scholar] [CrossRef] [PubMed]
  187. Ishwad, C.S.; Ferrell, R.E.; Hanley, K.; Davare, J.; Meloni, A.M.; Sandberg, A.A.; Surti, U. Two Discrete Regions of Deletion at 7q in Uterine Leiomyomas. Genes Chromosomes Cancer 1997, 19, 156–160. [Google Scholar] [CrossRef]
  188. van der Heijden, O.; Chiu, H.C.; Park, T.C.; Takahashi, H.; LiVolsi, V.A.; Risinger, J.I.; Barrett, J.C.; Berchuck, A.; Evans, A.C.; Behbakht, K.; et al. Allelotype Analysis of Uterine Leiomyoma: Localization of a Potential Tumor Suppressor Gene to a 4-CM Region of Chromosome 7q. Mol. Carcinog. 1998, 23, 243–247. [Google Scholar] [CrossRef]
  189. Vanharanta, S.; Wortham, N.C.; Langford, C.; El-Bahrawy, M.; van der Spuy, Z.; Sjöberg, J.; Lehtonen, R.; Karhu, A.; Tomlinson, I.P.M.; Aaltonen, L.A. Definition of a Minimal Region of Deletion of Chromosome 7 in Uterine Leiomyomas by Tiling-Path Microarray CGH and Mutation Analysis of Known Genes in This Region. Genes Chromosomes Cancer 2007, 46, 451–458. [Google Scholar] [CrossRef]
  190. Zeng, W.R.; Scherer, S.W.; Koutsilieris, M.; Huizenga, J.J.; Filteau, F.; Tsui, L.C.; Nepveu, A. Loss of Heterozygosity and Reduced Expression of the CUTL1 Gene in Uterine Leiomyomas. Oncogene 1997, 14, 2355–2365. [Google Scholar] [CrossRef] [Green Version]
  191. Schoenmakers, E.F.P.M.; Bunt, J.; Hermers, L.; Schepens, M.; Merkx, G.; Janssen, B.; Kersten, M.; Huys, E.; Pauwels, P.; Debiec-Rychter, M.; et al. Identification of CUX1 as the Recurrent Chromosomal Band 7q22 Target Gene in Human Uterine Leiomyoma. Genes Chromosomes Cancer 2013, 52, 11–23. [Google Scholar] [CrossRef]
  192. Quintana, D.G.; Thome, K.C.; Hou, Z.H.; Ligon, A.H.; Morton, C.C.; Dutta, A. ORC5L, a New Member of the Human Origin Recognition Complex, Is Deleted in Uterine Leiomyomas and Malignant Myeloid Diseases. J. Biol. Chem. 1998, 273, 27137–27145. [Google Scholar] [CrossRef] [Green Version]
  193. Sell, S.M.; Tullis, C.; Stracner, D.; Song, C.-Y.; Gewin, J. Minimal Interval Defined on 7q in Uterine Leiomyoma. Cancer Genet. Cytogenet. 2005, 157, 67–69. [Google Scholar] [CrossRef]
  194. Ptacek, T.; Song, C.; Walker, C.L.; Sell, S.M. Physical Mapping of Distinct 7q22 Deletions in Uterine Leiomyoma and Analysis of a Recently Annotated 7q22 Candidate Gene. Cancer Genet. Cytogenet. 2007, 174, 116–120. [Google Scholar] [CrossRef]
  195. Saito, E.; Okamoto, A.; Saito, M.; Shinozaki, H.; Takakura, S.; Yanaihara, N.; Ochiai, K.; Tanaka, T. Genes Associated with the Genesis of Leiomyoma of the Uterus in a Commonly Deleted Chromosomal Region at 7q22. Oncol. Rep. 2005, 13, 469–472. [Google Scholar] [CrossRef]
  196. Hodge, J.C.; Park, P.J.; Dreyfuss, J.M.; Assil-Kishawi, I.; Somasundaram, P.; Semere, L.G.; Quade, B.J.; Lynch, A.M.; Stewart, E.A.; Morton, C.C. Identifying the Molecular Signature of the Interstitial Deletion 7q Subgroup of Uterine Leiomyomata Using a Paired Analysis. Genes Chromosomes Cancer 2009, 48, 865–885. [Google Scholar] [CrossRef] [Green Version]
  197. Takahashi, K.; Kawamura, N.; Tsujimura, A.; Ichimura, T.; Ito, F.; Ishiko, O.; Ogita, S. Association of the Shrinkage of Uterine Leiomyoma Treated with GnRH Agonist and Deletion of Long Arm of Chromosome 7. Int. J. Oncol. 2001, 18, 1259–1263. [Google Scholar] [CrossRef]
  198. Nilbert, M.; Heim, S.; Mandahl, N.; Flodérus, U.M.; Willén, H.; Mitelman, F. Karyotypic Rearrangements in 20 Uterine Leiomyomas. Cytogenet. Cell Genet. 1988, 49, 300–304. [Google Scholar] [CrossRef]
  199. Nilbert, M.; Heim, S.; Mandahl, N.; Flodérus, U.M.; Willén, H.; Akerman, M.; Mitelman, F. Ring Formation and Structural Rearrangements of Chromosome 1 as Secondary Changes in Uterine Leiomyomas with t(12;14)(Q14-15;Q23-24). Cancer Genet. Cytogenet. 1988, 36, 183–190. [Google Scholar] [CrossRef]
  200. Pandis, N.; Heim, S.; Bardi, G.; Flodérus, U.M.; Willén, H.; Mandahl, N.; Mitelman, F. Parallel Karyotypic Evolution and Tumor Progression in Uterine Leiomyoma. Genes Chromosomes Cancer 1990, 2, 311–317. [Google Scholar] [CrossRef]
  201. Pandis, N.; Heim, S.; Bardi, G.; Flodérus, U.M.; Willén, H.; Mandahl, N.; Mitelman, F. Chromosome Analysis of 96 Uterine Leiomyomas. Cancer Genet. Cytogenet. 1991, 55, 11–18. [Google Scholar] [CrossRef]
  202. Levy, B.; Mukherjee, T.; Hirschhorn, K. Molecular Cytogenetic Analysis of Uterine Leiomyoma and Leiomyosarcoma by Comparative Genomic Hybridization. Cancer Genet. Cytogenet. 2000, 121, 1–8. [Google Scholar] [CrossRef]
  203. Christacos, N.C.; Quade, B.J.; Dal Cin, P.; Morton, C.C. Uterine Leiomyomata with Deletions of Ip Represent a Distinct Cytogenetic Subgroup Associated with Unusual Histologic Features. Genes Chromosomes Cancer 2006, 45, 304–312. [Google Scholar] [CrossRef]
  204. Hodge, J.C.; Pearce, K.E.; Clayton, A.C.; Taran, F.A.; Stewart, E.A. Uterine Cellular Leiomyomata with Chromosome 1p Deletions Represent a Distinct Entity. Am. J. Obstet. Gynecol. 2014, 210, 572.e1–572.e7. [Google Scholar] [CrossRef] [Green Version]
  205. Korbel, J.O.; Campbell, P.J. Criteria for Inference of Chromothripsis in Cancer Genomes. Cell 2013, 152, 1226–1236. [Google Scholar] [CrossRef] [Green Version]
  206. Stephens, P.J.; Greenman, C.D.; Fu, B.; Yang, F.; Bignell, G.R.; Mudie, L.J.; Pleasance, E.D.; Lau, K.W.; Beare, D.; Stebbings, L.A.; et al. Massive Genomic Rearrangement Acquired in a Single Catastrophic Event during Cancer Development. Cell 2011, 144, 27–40. [Google Scholar] [CrossRef]
  207. Kloosterman, W.P.; Koster, J.; Molenaar, J.J. Prevalence and Clinical Implications of Chromothripsis in Cancer Genomes. Curr. Opin. Oncol. 2014, 26, 64–72. [Google Scholar] [CrossRef]
  208. Koltsova, A.S.; Pendina, A.A.; Efimova, O.A.; Chiryaeva, O.G.; Kuznetzova, T.V.; Baranov, V.S. On the Complexity of Mechanisms and Consequences of Chromothripsis: An Update. Front. Genet. 2019, 10, 393. [Google Scholar] [CrossRef]
  209. Pendina, A.A.; Koltsova, A.S.; Efimova, O.A.; Malysheva, O.V.; Osinovskaya, N.S.; Sultanov, I.Y.; Tikhonov, A.V.; Shved, N.Y.; Chiryaeva, O.G.; Simareva, A.D.; et al. Case of Chromothripsis in a Large Solitary Non-Recurrent Uterine Leiomyoma. Eur. J. Obstet. Gynecol. Reprod. Biol. 2017, 219, 134–136. [Google Scholar] [CrossRef]
  210. Bowden, W.; Skorupski, J.; Kovanci, E.; Rajkovic, A. Detection of Novel Copy Number Variants in Uterine Leiomyomas Using High-Resolution SNP Arrays. Mol. Hum. Reprod. 2009, 15, 563–568. [Google Scholar] [CrossRef] [Green Version]
  211. Nagai, K.; Asano, R.; Sekiguchi, F.; Asai-Sato, M.; Miyagi, Y.; Miyagi, E. MED12 Mutations in Uterine Leiomyomas: Prediction of Volume Reduction by Gonadotropin-Releasing Hormone Agonists. Am. J. Obstet. Gynecol. 2022, S0002-9378(22)00748-7. [Google Scholar] [CrossRef] [PubMed]
  212. Kolterud, Å.; Välimäki, N.; Kuisma, H.; Patomo, J.; Ilves, S.T.; Mäkinen, N.; Kaukomaa, J.; Palin, K.; Kaasinen, E.; Karhu, A.; et al. Molecular Subclass of Uterine Fibroids Predicts Tumor Shrinkage in Response to Ulipristal Acetate. Hum. Mol. Genet. 2022, ddac217. [Google Scholar] [CrossRef] [PubMed]
  213. Nibert, M.; Heim, S. Uterine Leiomyoma Cytogenetics. Genes Chromosomes Cancer 1990, 2, 3–13. [Google Scholar] [CrossRef]
  214. Canevari, R.A.; Pontes, A.; Rosa, F.E.; Rainho, C.A.; Rogatto, S.R. Independent Clonal Origin of Multiple Uterine Leiomyomas That Was Determined by X Chromosome Inactivation and Microsatellite Analysis. Am. J. Obstet. Gynecol. 2005, 193, 1395–1403. [Google Scholar] [CrossRef]
  215. Zhang, P.; Zhang, C.; Hao, J.; Sung, C.J.; Quddus, M.R.; Steinhoff, M.M.; Lawrence, W.D. Use of X-Chromosome Inactivation Pattern to Determine the Clonal Origins of Uterine Leiomyoma and Leiomyosarcoma. Hum. Pathol. 2006, 37, 1350–1356. [Google Scholar] [CrossRef]
  216. Cai, Y.-R.; Diao, X.-L.; Wang, S.-F.; Zhang, W.; Zhang, H.-T.; Su, Q. X-Chromosomal Inactivation Analysis of Uterine Leiomyomas Reveals a Common Clonal Origin of Different Tumor Nodules in Some Multiple Leiomyomas. Int. J. Oncol. 2007, 31, 1379–1389. [Google Scholar] [CrossRef] [Green Version]
  217. Mehine, M.; Heinonen, H.-R.; Sarvilinna, N.; Pitkänen, E.; Mäkinen, N.; Katainen, R.; Tuupanen, S.; Bützow, R.; Sjöberg, J.; Aaltonen, L.A. Clonally Related Uterine Leiomyomas Are Common and Display Branched Tumor Evolution. Hum. Mol. Genet. 2015, 24, 4407–4416. [Google Scholar] [CrossRef] [Green Version]
  218. Fasih, N.; Prasad Shanbhogue, A.K.; Macdonald, D.B.; Fraser-Hill, M.A.; Papadatos, D.; Kielar, A.Z.; Doherty, G.P.; Walsh, C.; McInnes, M.; Atri, M. Leiomyomas beyond the Uterus: Unusual Locations, Rare Manifestations. Radiogr. Rev. Publ. Radiol. Soc. N. Am. Inc 2008, 28, 1931–1948. [Google Scholar] [CrossRef]
  219. Quade, B.J.; McLachlin, C.M.; Soto-Wright, V.; Zuckerman, J.; Mutter, G.L.; Morton, C.C. Disseminated Peritoneal Leiomyomatosis. Clonality Analysis by X Chromosome Inactivation and Cytogenetics of a Clinically Benign Smooth Muscle Proliferation. Am. J. Pathol. 1997, 150, 2153–2166. [Google Scholar]
  220. Dal Cin, P.; Quade, B.J.; Neskey, D.M.; Kleinman, M.S.; Weremowicz, S.; Morton, C.C. Intravenous Leiomyomatosis Is Characterized by a Der(14)t(12;14)(Q15;Q24). Genes Chromosomes Cancer 2003, 36, 205–206. [Google Scholar] [CrossRef]
  221. Bowen, J.M.; Cates, J.M.; Kash, S.; Itani, D.; Gonzalez, A.; Huang, D.; Oliveira, A.; Bridge, J.A. Genomic Imbalances in Benign Metastasizing Leiomyoma: Characterization by Conventional Karyotypic, Fluorescence in Situ Hybridization, and Whole Genome SNP Array Analysis. Cancer Genet. 2012, 205, 249–254. [Google Scholar] [CrossRef] [Green Version]
  222. Wu, R.-C.; Chao, A.-S.; Lee, L.-Y.; Lin, G.; Chen, S.-J.; Lu, Y.-J.; Huang, H.-J.; Yen, C.-F.; Han, C.M.; Lee, Y.-S.; et al. Massively Parallel Sequencing and Genome-Wide Copy Number Analysis Revealed a Clonal Relationship in Benign Metastasizing Leiomyoma. Oncotarget 2017, 8, 47547–47554. [Google Scholar] [CrossRef] [Green Version]
  223. Jiang, J.; He, M.; Hu, X.; Ni, C.; Yang, L. Deep Sequencing Reveals the Molecular Pathology Characteristics between Primary Uterine Leiomyoma and Pulmonary Benign Metastasizing Leiomyoma. Clin. Transl. Oncol. Off. Publ. Fed. Span. Oncol. Soc. Natl. Cancer Inst. Mex. 2018, 20, 1080–1086. [Google Scholar] [CrossRef]
  224. Ofori, K.; Fernandes, H.; Cummings, M.; Colby, T.; Saqi, A. Benign Metastasizing Leiomyoma Presenting with Miliary Pattern and Fatal Outcome: Case Report with Molecular Analysis & Review of the Literature. Respir. Med. Case Rep. 2019, 27, 100831. [Google Scholar] [CrossRef]
  225. Ma, Y.; Wang, S.; Liu, Q.; Lu, B. A Clinicopathological and Molecular Analysis in Uterine Leiomyomas and Concurrent/Metachronous Peritoneal Nodules: New Insights into Disseminated Peritoneal Leiomyomatosis. Pathol. Res. Pract. 2020, 216, 152938. [Google Scholar] [CrossRef]
  226. Nucci, M.R.; Drapkin, R.; Dal Cin, P.; Fletcher, C.D.M.; Fletcher, J.A. Distinctive Cytogenetic Profile in Benign Metastasizing Leiomyoma: Pathogenetic Implications. Am. J. Surg. Pathol. 2007, 31, 737–743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Ahvenainen, T.; Khamaiseh, S.; Alkodsi, A.; Mehine, M.; Nevala, R.; Äyräväinen, A.; Bützow, R.; Vahteristo, P. Lung Metastases and Subsequent Malignant Transformation of a Fumarate Hydratase -Deficient Uterine Leiomyoma. Exp. Mol. Pathol. 2022, 126, 104760. [Google Scholar] [CrossRef] [PubMed]
  228. Ahvenainen, T.V.; Mäkinen, N.M.; von Nandelstadh, P.; Vahteristo, M.E.A.; Pasanen, A.M.; Bützow, R.C.; Vahteristo, P.M. Loss of ATRX/DAXX Expression and Alternative Lengthening of Telomeres in Uterine Leiomyomas. Cancer 2018, 124, 4650–4656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Sanada, S.; Ushijima, K.; Yanai, H.; Mikami, Y.; Ohishi, Y.; Kobayashi, H.; Tashiro, H.; Mikami, M.; Miyamoto, S.; Katabuchi, H. A Critical Review of “Uterine Leiomyoma” with Subsequent Recurrence or Metastasis: A Multicenter Study of 62 Cases. J. Obstet. Gynaecol. Res. 2022, 48, 3242–3251. [Google Scholar] [CrossRef]
  230. Panesar, H.; Dhaliwal, H.S. Iatrogenic Parasitic Leiomyomas: A Late and Uncommon Complication After Laparoscopic Morcellation. Cureus 2022, 14, e24718. [Google Scholar] [CrossRef]
  231. Holzmann, C.; Kuepker, W.; Rommel, B.; Helmke, B.; Bullerdiek, J. Reasons to Reconsider Risk Associated With Power Morcellation of Uterine Fibroids. Vivo Athens Greece 2020, 34, 1–9. [Google Scholar] [CrossRef]
  232. Kyozuka, H.; Jin, T.; Sugeno, M.; Kuratsune, K.; Ando, H.; Ito, F.; Odajima, H.; Suzuki, D.; Nomura, Y. A Case of Spontaneous Parasitic Myoma in a Patient without a History of Myomectomy Treated Laparoscopically. Fukushima J. Med. Sci. 2022, 68, 123–127. [Google Scholar] [CrossRef]
  233. Mittal, K.; Joutovsky, A. Areas with Benign Morphologic and Immunohistochemical Features Are Associated with Some Uterine Leiomyosarcomas. Gynecol. Oncol. 2007, 104, 362–365. [Google Scholar] [CrossRef] [PubMed]
  234. Mittal, K.R.; Chen, F.; Wei, J.J.; Rijhvani, K.; Kurvathi, R.; Streck, D.; Dermody, J.; Toruner, G.A. Molecular and Immunohistochemical Evidence for the Origin of Uterine Leiomyosarcomas from Associated Leiomyoma and Symplastic Leiomyoma-like Areas. Mod. Pathol. Off. J. U. S. Can. Acad. Pathol. Inc 2009, 22, 1303–1311. [Google Scholar] [CrossRef] [Green Version]
  235. Packenham, J.P.; du Manoir, S.; Schrock, E.; Risinger, J.I.; Dixon, D.; Denz, D.N.; Evans, J.A.; Berchuck, A.; Barrett, J.C.; Devereux, T.R.; et al. Analysis of Genetic Alterations in Uterine Leiomyomas and Leiomyosarcomas by Comparative Genomic Hybridization. Mol. Carcinog. 1997, 19, 273–279. [Google Scholar] [CrossRef]
  236. Ravegnini, G.; Mariño-Enriquez, A.; Slater, J.; Eilers, G.; Wang, Y.; Zhu, M.; Nucci, M.R.; George, S.; Angelini, S.; Raut, C.P.; et al. MED12 Mutations in Leiomyosarcoma and Extrauterine Leiomyoma. Mod. Pathol. Off. J. U. S. Can. Acad. Pathol. Inc 2013, 26, 743–749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  237. Bertsch, E.; Qiang, W.; Zhang, Q.; Espona-Fiedler, M.; Druschitz, S.; Liu, Y.; Mittal, K.; Kong, B.; Kurita, T.; Wei, J.-J. MED12 and HMGA2 Mutations: Two Independent Genetic Events in Uterine Leiomyoma and Leiomyosarcoma. Mod. Pathol. Off. J. U. S. Can. Acad. Pathol. Inc 2014, 27, 1144–1153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Hodge, J.C.; Morton, C.C. Genetic Heterogeneity among Uterine Leiomyomata: Insights into Malignant Progression. Hum. Mol. Genet. 2007, 16 Spec No 1, R7–R13. [Google Scholar] [CrossRef]
  239. Zhang, Q.; Ubago, J.; Li, L.; Guo, H.; Liu, Y.; Qiang, W.; Kim, J.J.; Kong, B.; Wei, J.-J. Molecular Analyses of 6 Different Types of Uterine Smooth Muscle Tumors: Emphasis in Atypical Leiomyoma. Cancer 2014, 120, 3165–3177. [Google Scholar] [CrossRef] [PubMed]
  240. Dharajiya, N.G.; Namba, A.; Horiuchi, I.; Miyai, S.; Farkas, D.H.; Almasri, E.; Saldivar, J.-S.; Takagi, K.; Kamei, Y. Uterine Leiomyoma Confounding a Noninvasive Prenatal Test Result. Prenat. Diagn. 2015, 35, 990–993. [Google Scholar] [CrossRef]
  241. Dharajiya, N.G.; Grosu, D.S.; Farkas, D.H.; McCullough, R.M.; Almasri, E.; Sun, Y.; Kim, S.K.; Jensen, T.J.; Saldivar, J.-S.; Topol, E.J.; et al. Incidental Detection of Maternal Neoplasia in Noninvasive Prenatal Testing. Clin. Chem. 2018, 64, 329–335. [Google Scholar] [CrossRef] [PubMed]
  242. Ivashchenko, T.E.; Vashukova, E.S.; Kozyulina, P.Y.; Dvoynova, N.M.; Talantova, O.E.; Koroteev, A.L.; Pendina, A.A.; Tikhonov, A.V.; Chiryaeva, O.G.; Petrova, L.I.; et al. Noninvasive Prenatal Testing Using Next Generation Sequencing: Pilot Experience of the D.O. Ott Research Institute of Obstetrics, Gynecology and Reproductology. Russ. J. Genet. 2019, 55, 1208–1213. [Google Scholar] [CrossRef]
  243. Scott, F.; Menezes, M.; Smet, M.E.; Carey, K.; Hardy, T.; Fullston, T.; Rolnik, D.L.; McLennan, A. Influence of Fibroids on Cell-Free DNA Screening Accuracy. Ultrasound Obstet. Gynecol. Off. J. Int. Soc. Ultrasound Obstet. Gynecol. 2022, 59, 114–119. [Google Scholar] [CrossRef] [PubMed]
  244. Linder, D.; Gartler, S.M. Glucose-6-Phosphate Dehydrogenase Mosaicism: Utilization as a Cell Marker in the Study of Leiomyomas. Science 1965, 150, 67–69. [Google Scholar] [CrossRef] [PubMed]
  245. Swierczek, S.I.; Piterkova, L.; Jelinek, J.; Agarwal, N.; Hammoud, S.; Wilson, A.; Hickman, K.; Parker, C.J.; Cairns, B.R.; Prchal, J.T. Methylation of AR Locus Does Not Always Reflect X Chromosome Inactivation State. Blood 2012, 119, e100–e109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  246. Bullerdiek, J.; Rommel, B. Factors Targeting MED12 to Drive Tumorigenesis? F1000Research 2018, 7, 359. [Google Scholar] [CrossRef] [PubMed]
  247. Osborne, C.S. Molecular Pathways: Transcription Factories and Chromosomal Translocations. Clin. Cancer Res. Off. J. Am. Assoc. Cancer Res. 2014, 20, 296–300. [Google Scholar] [CrossRef] [Green Version]
  248. Fritz, A.J.; Sehgal, N.; Pliss, A.; Xu, J.; Berezney, R. Chromosome Territories and the Global Regulation of the Genome. Genes Chromosomes Cancer 2019, 58, 407–426. [Google Scholar] [CrossRef] [Green Version]
  249. Rogalla, P.; Rohen, C.; Hennig, Y.; Deichert, U.; Bonk, U.; Bullerdiek, J. Telomere Repeat Fragment Sizes Do Not Limit the Growth Potential of Uterine Leiomyomas. Biochem. Biophys. Res. Commun. 1995, 211, 175–182. [Google Scholar] [CrossRef]
  250. Bonatz, G.; Frahm, S.O.; Andreas, S.; Heidorn, K.; Jonat, W.; Parwaresch, R. Telomere Shortening in Uterine Leiomyomas. Am. J. Obstet. Gynecol. 1998, 179, 591–596. [Google Scholar] [CrossRef]
  251. Ulaner, G.A.; Hu, J.F.; Vu, T.H.; Oruganti, H.; Giudice, L.C.; Hoffman, A.R. Regulation of Telomerase by Alternate Splicing of Human Telomerase Reverse Transcriptase (HTERT) in Normal and Neoplastic Ovary, Endometrium and Myometrium. Int. J. Cancer 2000, 85, 330–335. [Google Scholar] [CrossRef]
  252. Oh, B.-K.; Choi, Y.; Choi, J.S. Telomere Shortening and Expression of TRF1 and TRF2 in Uterine Leiomyoma. Mol. Med. Rep. 2021, 24, 606. [Google Scholar] [CrossRef]
  253. Maciejowski, J.; Chatzipli, A.; Dananberg, A.; Chu, K.; Toufektchan, E.; Klimczak, L.J.; Gordenin, D.A.; Campbell, P.J.; de Lange, T. APOBEC3-Dependent Kataegis and TREX1-Driven Chromothripsis during Telomere Crisis. Nat. Genet. 2020, 52, 884–890. [Google Scholar] [CrossRef]
  254. Dewhurst, S.M.; Yao, X.; Rosiene, J.; Tian, H.; Behr, J.; Bosco, N.; Takai, K.K.; de Lange, T.; Imieliński, M. Structural Variant Evolution after Telomere Crisis. Nat. Commun. 2021, 12, 2093. [Google Scholar] [CrossRef]
  255. Kuisma, H.; Bramante, S.; Rajamäki, K.; Sipilä, L.J.; Kaasinen, E.; Kaukomaa, J.; Palin, K.; Mäkinen, N.; Sjöberg, J.; Sarvilinna, N.; et al. Parity Associates with Chromosomal Damage in Uterine Leiomyomas. Nat. Commun. 2021, 12, 5448. [Google Scholar] [CrossRef] [PubMed]
  256. Westhorpe, F.G.; Straight, A.F. The Centromere: Epigenetic Control of Chromosome Segregation during Mitosis. Cold Spring Harb. Perspect. Biol. 2014, 7, a015818. [Google Scholar] [CrossRef] [Green Version]
  257. Yang, Q.; Mas, A.; Diamond, M.P.; Al-Hendy, A. The Mechanism and Function of Epigenetics in Uterine Leiomyoma Development. Reprod. Sci. Thousand Oaks Calif. 2016, 23, 163–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  258. Strahl, B.D.; Allis, C.D. The Language of Covalent Histone Modifications. Nature 2000, 403, 41–45. [Google Scholar] [CrossRef] [PubMed]
  259. Zhang, Y.; Sun, Z.; Jia, J.; Du, T.; Zhang, N.; Tang, Y.; Fang, Y.; Fang, D. Overview of Histone Modification. Adv. Exp. Med. Biol. 2021, 1283, 1–16. [Google Scholar] [CrossRef] [PubMed]
  260. Wei, L.-H.; Torng, P.-L.; Hsiao, S.-M.; Jeng, Y.-M.; Chen, M.-W.; Chen, C.-A. Histone Deacetylase 6 Regulates Estrogen Receptor Alpha in Uterine Leiomyoma. Reprod. Sci. Thousand Oaks Calif. 2011, 18, 755–762. [Google Scholar] [CrossRef] [PubMed]
  261. Ali, M.; Shahin, S.M.; Sabri, N.A.; Al-Hendy, A.; Yang, Q. Activation of β-Catenin Signaling and Its Crosstalk With Estrogen and Histone Deacetylases in Human Uterine Fibroids. J. Clin. Endocrinol. Metab. 2020, 105, e1517–e1535. [Google Scholar] [CrossRef]
  262. Sant’Anna, G.D.S.; Brum, I.S.; Branchini, G.; Pizzolato, L.S.; Capp, E.; Corleta, H. von E. Ovarian Steroid Hormones Modulate the Expression of Progesterone Receptors and Histone Acetylation Patterns in Uterine Leiomyoma Cells. Gynecol. Endocrinol. Off. J. Int. Soc. Gynecol. Endocrinol. 2017, 33, 629–633. [Google Scholar] [CrossRef] [PubMed]
  263. Carbajo-García, M.C.; García-Alcázar, Z.; Corachán, A.; Monleón, J.; Trelis, A.; Faus, A.; Pellicer, A.; Ferrero, H. Histone Deacetylase Inhibition by Suberoylanilide Hydroxamic Acid: A Therapeutic Approach to Treat Human Uterine Leiomyoma. Fertil. Steril. 2022, 117, 433–443. [Google Scholar] [CrossRef]
  264. Yang, Q.; Nair, S.; Laknaur, A.; Ismail, N.; Diamond, M.P.; Al-Hendy, A. The Polycomb Group Protein EZH2 Impairs DNA Damage Repair Gene Expression in Human Uterine Fibroids. Biol. Reprod. 2016, 94, 69. [Google Scholar] [CrossRef] [PubMed]
  265. Yang, Q.; Laknaur, A.; Elam, L.; Ismail, N.; Gavrilova-Jordan, L.; Lue, J.; Diamond, M.P.; Al-Hendy, A. Identification of Polycomb Group Protein EZH2-Mediated DNA Mismatch Repair Gene MSH2 in Human Uterine Fibroids. Reprod. Sci. Thousand Oaks Calif. 2016, 23, 1314–1325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Greathouse, K.L.; Bredfeldt, T.; Everitt, J.I.; Lin, K.; Berry, T.; Kannan, K.; Mittelstadt, M.L.; Ho, S.; Walker, C.L. Environmental Estrogens Differentially Engage the Histone Methyltransferase EZH2 to Increase Risk of Uterine Tumorigenesis. Mol. Cancer Res. MCR 2012, 10, 546–557. [Google Scholar] [CrossRef] [Green Version]
  267. Yu, L.; Ham, K.; Gao, X.; Castro, L.; Yan, Y.; Kissling, G.E.; Tucker, C.J.; Flagler, N.; Dong, R.; Archer, T.K.; et al. Epigenetic Regulation of Transcription Factor Promoter Regions by Low-Dose Genistein through Mitogen-Activated Protein Kinase and Mitogen-and-Stress Activated Kinase 1 Nongenomic Signaling. Cell Commun. Signal. CCS 2016, 14, 18. [Google Scholar] [CrossRef] [Green Version]
  268. Yu, L.; Liu, J.; Yan, Y.; Burwell, A.; Castro, L.; Shi, M.; Dixon, D. “Metalloestrogenic” Effects of Cadmium Downstream of G Protein-Coupled Estrogen Receptor and Mitogen-Activated Protein Kinase Pathways in Human Uterine Fibroid Cells. Arch. Toxicol. 2021, 95, 1995–2006. [Google Scholar] [CrossRef]
  269. Berta, D.G.; Kuisma, H.; Välimäki, N.; Räisänen, M.; Jäntti, M.; Pasanen, A.; Karhu, A.; Kaukomaa, J.; Taira, A.; Cajuso, T.; et al. Deficient H2A.Z Deposition Is Associated with Genesis of Uterine Leiomyoma. Nature 2021, 596, 398–403. [Google Scholar] [CrossRef]
  270. Nan, X.; Ng, H.H.; Johnson, C.A.; Laherty, C.D.; Turner, B.M.; Eisenman, R.N.; Bird, A. Transcriptional Repression by the Methyl-CpG-Binding Protein MeCP2 Involves a Histone Deacetylase Complex. Nature 1998, 393, 386–389. [Google Scholar] [CrossRef]
  271. Moore, L.D.; Le, T.; Fan, G. DNA Methylation and Its Basic Function. Neuropsychopharmacol. Off. Publ. Am. Coll. Neuropsychopharmacol. 2013, 38, 23–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Bochtler, M.; Kolano, A.; Xu, G.-L. DNA Demethylation Pathways: Additional Players and Regulators. BioEssays News Rev. Mol. Cell. Dev. Biol. 2017, 39, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Tahiliani, M.; Koh, K.P.; Shen, Y.; Pastor, W.A.; Bandukwala, H.; Brudno, Y.; Agarwal, S.; Iyer, L.M.; Liu, D.R.; Aravind, L.; et al. Conversion of 5-Methylcytosine to 5-Hydroxymethylcytosine in Mammalian DNA by MLL Partner TET1. Science 2009, 324, 930–935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  274. Ito, S.; D’Alessio, A.C.; Taranova, O.V.; Hong, K.; Sowers, L.C.; Zhang, Y. Role of Tet Proteins in 5mC to 5hmC Conversion, ES-Cell Self-Renewal and Inner Cell Mass Specification. Nature 2010, 466, 1129–1133. [Google Scholar] [CrossRef] [Green Version]
  275. Ito, S.; Shen, L.; Dai, Q.; Wu, S.C.; Collins, L.B.; Swenberg, J.A.; He, C.; Zhang, Y. Tet Proteins Can Convert 5-Methylcytosine to 5-Formylcytosine and 5-Carboxylcytosine. Science 2011, 333, 1300–1303. [Google Scholar] [CrossRef] [Green Version]
  276. Xu, W.; Yang, H.; Liu, Y.; Yang, Y.; Wang, P.; Kim, S.-H.; Ito, S.; Yang, C.; Wang, P.; Xiao, M.-T.; et al. Oncometabolite 2-Hydroxyglutarate Is a Competitive Inhibitor of α-Ketoglutarate-Dependent Dioxygenases. Cancer Cell 2011, 19, 17–30. [Google Scholar] [CrossRef] [Green Version]
  277. Xiao, M.; Yang, H.; Xu, W.; Ma, S.; Lin, H.; Zhu, H.; Liu, L.; Liu, Y.; Yang, C.; Xu, Y.; et al. Inhibition of α-KG-Dependent Histone and DNA Demethylases by Fumarate and Succinate That Are Accumulated in Mutations of FH and SDH Tumor Suppressors. Genes Dev. 2012, 26, 1326–1338. [Google Scholar] [CrossRef] [Green Version]
  278. Efimova, O.A.; Koltsova, A.S.; Krapivin, M.I.; Tikhonov, A.V.; Pendina, A.A. Environmental Epigenetics and Genome Flexibility: Focus on 5-Hydroxymethylcytosine. Int. J. Mol. Sci. 2020, 21, 3223. [Google Scholar] [CrossRef]
  279. Nabel, C.S.; Jia, H.; Ye, Y.; Shen, L.; Goldschmidt, H.L.; Stivers, J.T.; Zhang, Y.; Kohli, R.M. AID/APOBEC Deaminases Disfavor Modified Cytosines Implicated in DNA Demethylation. Nat. Chem. Biol. 2012, 8, 751–758. [Google Scholar] [CrossRef] [Green Version]
  280. Boorstein, R.J.; Cummings, A.; Marenstein, D.R.; Chan, M.K.; Ma, Y.; Neubert, T.A.; Brown, S.M.; Teebor, G.W. Definitive Identification of Mammalian 5-Hydroxymethyluracil DNA N-Glycosylase Activity as SMUG1. J. Biol. Chem. 2001, 276, 41991–41997. [Google Scholar] [CrossRef] [Green Version]
  281. Wu, X.; Zhang, Y. TET-Mediated Active DNA Demethylation: Mechanism, Function and Beyond. Nat. Rev. Genet. 2017, 18, 517–534. [Google Scholar] [CrossRef]
  282. Ji, D.; Lin, K.; Song, J.; Wang, Y. Effects of Tet-Induced Oxidation Products of 5-Methylcytosine on Dnmt1- and DNMT3a-Mediated Cytosine Methylation. Mol. Biosyst. 2014, 10, 1749–1752. [Google Scholar] [CrossRef] [Green Version]
  283. Li, S.; Chiang, T.; Richard-Davis, G.; Barrett, J.C.; Mclachlan, J.A. DNA Hypomethylation and Imbalanced Expression of DNA Methyltransferases (DNMT1, 3A, and 3B) in Human Uterine Leiomyoma. Gynecol. Oncol. 2003, 90, 123–130. [Google Scholar] [CrossRef]
  284. Yamagata, Y.; Maekawa, R.; Asada, H.; Taketani, T.; Tamura, I.; Tamura, H.; Ogane, J.; Hattori, N.; Shiota, K.; Sugino, N. Aberrant DNA Methylation Status in Human Uterine Leiomyoma. Mol. Hum. Reprod. 2009, 15, 259–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Carbajo-García, M.C.; Corachán, A.; Segura-Benitez, M.; Monleón, J.; Escrig, J.; Faus, A.; Pellicer, A.; Cervelló, I.; Ferrero, H. 5-Aza-2′-Deoxycitidine Inhibits Cell Proliferation, Extracellular Matrix Formation and Wnt/β-Catenin Pathway in Human Uterine Leiomyomas. Reprod. Biol. Endocrinol. RBE 2021, 19, 106. [Google Scholar] [CrossRef] [PubMed]
  286. Maekawa, R.; Yagi, S.; Ohgane, J.; Yamagata, Y.; Asada, H.; Tamura, I.; Sugino, N.; Shiota, K. Disease-Dependent Differently Methylated Regions (D-DMRs) of DNA Are Enriched on the X Chromosome in Uterine Leiomyoma. J. Reprod. Dev. 2011, 57, 604–612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Asada, H.; Yamagata, Y.; Taketani, T.; Matsuoka, A.; Tamura, H.; Hattori, N.; Ohgane, J.; Hattori, N.; Shiota, K.; Sugino, N. Potential Link between Estrogen Receptor-Alpha Gene Hypomethylation and Uterine Fibroid Formation. Mol. Hum. Reprod. 2008, 14, 539–545. [Google Scholar] [CrossRef] [Green Version]
  288. Maekawa, R.; Sato, S.; Yamagata, Y.; Asada, H.; Tamura, I.; Lee, L.; Okada, M.; Tamura, H.; Takaki, E.; Nakai, A.; et al. Genome-Wide DNA Methylation Analysis Reveals a Potential Mechanism for the Pathogenesis and Development of Uterine Leiomyomas. PLoS ONE 2013, 8, e66632. [Google Scholar] [CrossRef] [Green Version]
  289. Carbajo-García, M.C.; Corachán, A.; Juárez-Barber, E.; Monleón, J.; Payá, V.; Trelis, A.; Quiñonero, A.; Pellicer, A.; Ferrero, H. Integrative Analysis of the DNA Methylome and Transcriptome in Uterine Leiomyoma Shows Altered Regulation of Genes Involved in Metabolism, Proliferation, Extracellular Matrix, and Vesicles. J. Pathol. 2022, 257, 663–673. [Google Scholar] [CrossRef]
  290. Navarro, A.; Yin, P.; Monsivais, D.; Lin, S.M.; Du, P.; Wei, J.-J.; Bulun, S.E. Genome-Wide DNA Methylation Indicates Silencing of Tumor Suppressor Genes in Uterine Leiomyoma. PLoS ONE 2012, 7, e33284. [Google Scholar] [CrossRef] [Green Version]
  291. Jiang, W.; Luo, H.; Shen, Z.; Zhang, W.; Chen, A.; Zhu, X. Down-Regulation and Gene Hypermethylation of the 14-3-3 Gamma in Uterine Leiomyoma. Front. Biosci. Landmark Ed. 2016, 21, 1286–1295. [Google Scholar] [CrossRef] [PubMed]
  292. Braný, D.; Dvorská, D.; Grendár, M.; Ňachajová, M.; Szépe, P.; Lasabová, Z.; Žúbor, P.; Višňovský, J.; Halášová, E. Different Methylation Levels in the KLF4, ATF3 and DLEC1 Genes in the Myometrium and in Corpus Uteri Mesenchymal Tumours as Assessed by MS-HRM. Pathol. Res. Pract. 2019, 215, 152465. [Google Scholar] [CrossRef] [PubMed]
  293. Sato, S.; Maekawa, R.; Tamura, I.; Shirafuta, Y.; Shinagawa, M.; Asada, H.; Taketani, T.; Tamura, H.; Sugino, N. SATB2 and NGR1: Potential Upstream Regulatory Factors in Uterine Leiomyomas. J. Assist. Reprod. Genet. 2019, 36, 2385–2397. [Google Scholar] [CrossRef] [PubMed]
  294. Sato, S.; Maekawa, R.; Yamagata, Y.; Tamura, I.; Lee, L.; Okada, M.; Jozaki, K.; Asada, H.; Tamura, H.; Sugino, N. Identification of Uterine Leiomyoma-Specific Marker Genes Based on DNA Methylation and Their Clinical Application. Sci. Rep. 2016, 6, 30652. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Liu, S.; Yin, P.; Xu, J.; Dotts, A.J.; Kujawa, S.A.; Coon V, J.S.; Zhao, H.; Shilatifard, A.; Dai, Y.; Bulun, S.E. Targeting DNA Methylation Depletes Uterine Leiomyoma Stem Cell-Enriched Population by Stimulating Their Differentiation. Endocrinology 2020, 161, bqaa143. [Google Scholar] [CrossRef]
  296. Liu, S.; Yin, P.; Xu, J.; Dotts, A.J.; Kujawa, S.A.; Coon V, J.S.; Zhao, H.; Dai, Y.; Bulun, S.E. Progesterone Receptor-DNA Methylation Crosstalk Regulates Depletion of Uterine Leiomyoma Stem Cells: A Potential Therapeutic Target. Stem Cell Rep. 2021, 16, 2099–2106. [Google Scholar] [CrossRef]
  297. Navarro, A.; Yin, P.; Ono, M.; Monsivais, D.; Moravek, M.B.; Coon, J.S.; Dyson, M.T.; Wei, J.-J.; Bulun, S.E. 5-Hydroxymethylcytosine Promotes Proliferation of Human Uterine Leiomyoma: A Biological Link to a New Epigenetic Modification in Benign Tumors. J. Clin. Endocrinol. Metab. 2014, 99, E2437–E2445. [Google Scholar] [CrossRef]
  298. Cao, T.; Jiang, Y.; Wang, Z.; Zhang, N.; Al-Hendy, A.; Mamillapalli, R.; Kallen, A.N.; Kodaman, P.; Taylor, H.S.; Li, D.; et al. H19 LncRNA Identified as a Master Regulator of Genes That Drive Uterine Leiomyomas. Oncogene 2019, 38, 5356–5366. [Google Scholar] [CrossRef] [Green Version]
  299. Zhan, X.; Zhou, H.; Sun, Y.; Shen, B.; Chou, D. Long Non-Coding Ribonucleic Acid H19 and Ten-Eleven Translocation Enzyme 1 Messenger RNA Expression Levels in Uterine Fibroids May Predict Their Postoperative Recurrence. Clin. Sao Paulo Braz. 2021, 76, e2671. [Google Scholar] [CrossRef]
  300. Efimova, O.A.; Pendina, A.A.; Tikhonov, A.V.; Baranov, V.S. The Evolution of Ideas on the Biological Role of 5-Methylcytosine Oxidative Derivatives in the Mammalian Genome. Russ. J. Genet. Appl. Res. 2018, 8, 11–21. [Google Scholar] [CrossRef]
  301. Wu, H.; D’Alessio, A.C.; Ito, S.; Wang, Z.; Cui, K.; Zhao, K.; Sun, Y.E.; Zhang, Y. Genome-Wide Analysis of 5-Hydroxymethylcytosine Distribution Reveals Its Dual Function in Transcriptional Regulation in Mouse Embryonic Stem Cells. Genes Dev. 2011, 25, 679–684. [Google Scholar] [CrossRef] [Green Version]
  302. Li, W.; Liu, M. Distribution of 5-Hydroxymethylcytosine in Different Human Tissues. J. Nucleic Acids 2011, 2011, 870726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Nestor, C.E.; Ottaviano, R.; Reddington, J.; Sproul, D.; Reinhardt, D.; Dunican, D.; Katz, E.; Dixon, J.M.; Harrison, D.J.; Meehan, R.R. Tissue Type Is a Major Modifier of the 5-Hydroxymethylcytosine Content of Human Genes. Genome Res. 2012, 22, 467–477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Efimova, O.A.; Pendina, A.A.; Tikhonov, A.V.; Fedorova, I.D.; Krapivin, M.I.; Chiryaeva, O.G.; Shilnikova, E.M.; Bogdanova, M.A.; Kogan, I.Y.; Kuznetzova, T.V.; et al. Chromosome Hydroxymethylation Patterns in Human Zygotes and Cleavage-Stage Embryos. Reprod. Camb. Engl. 2015, 149, 223–233. [Google Scholar] [CrossRef] [Green Version]
  305. Efimova, O.A.; Pendina, A.A.; Tikhonov, A.V.; Parfenyev, S.E.; Mekina, I.D.; Komarova, E.M.; Mazilina, M.A.; Daev, E.V.; Chiryaeva, O.G.; Galembo, I.A.; et al. Genome-Wide 5-Hydroxymethylcytosine Patterns in Human Spermatogenesis Are Associated with Semen Quality. Oncotarget 2017, 8, 88294–88307. [Google Scholar] [CrossRef] [Green Version]
  306. Efimova, O.A.; Pendina, A.A.; Krapivin, M.I.; Kopat, V.V.; Tikhonov, A.V.; Petrovskaia-Kaminskaia, A.V.; Navodnikova, P.M.; Talantova, O.E.; Glotov, O.S.; Baranov, V.S. Inter-Cell and Inter-Chromosome Variability of 5-Hydroxymethylcytosine Patterns in Noncultured Human Embryonic and Extraembryonic Cells. Cytogenet. Genome Res. 2018, 156, 150–157. [Google Scholar] [CrossRef]
  307. Kudo, Y.; Tateishi, K.; Yamamoto, K.; Yamamoto, S.; Asaoka, Y.; Ijichi, H.; Nagae, G.; Yoshida, H.; Aburatani, H.; Koike, K. Loss of 5-Hydroxymethylcytosine Is Accompanied with Malignant Cellular Transformation. Cancer Sci. 2012, 103, 670–676. [Google Scholar] [CrossRef] [PubMed]
  308. Wu, Y.-C.; Ling, Z.-Q. The Role of TET Family Proteins and 5-Hydroxymethylcytosine in Human Tumors. Histol. Histopathol. 2014, 29, 991–997. [Google Scholar] [CrossRef]
  309. Chen, Z.; Shi, X.; Guo, L.; Li, Y.; Luo, M.; He, J. Decreased 5-Hydroxymethylcytosine Levels Correlate with Cancer Progression and Poor Survival: A Systematic Review and Meta-Analysis. Oncotarget 2017, 8, 1944–1952. [Google Scholar] [CrossRef] [Green Version]
  310. Lu, J.-L.; Zhao, L.; Han, S.-C.; Bi, J.-L.; Liu, H.-X.; Yue, C.; Lin, L. MiR-129 Is Involved in the Occurrence of Uterine Fibroid through Inhibiting TET1. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 4419–4426. [Google Scholar] [CrossRef]
  311. Chuang, T.-D.; Khorram, O. Regulation of Cell Cycle Regulatory Proteins by MicroRNAs in Uterine Leiomyoma. Reprod. Sci. Thousand Oaks Calif 2019, 26, 250–258. [Google Scholar] [CrossRef]
  312. Chuang, T.-D.; Rehan, A.; Khorram, O. Tranilast Induces MiR-200c Expression through Blockade of RelA/P65 Activity in Leiomyoma Smooth Muscle Cells. Fertil. Steril. 2020, 113, 1308–1318. [Google Scholar] [CrossRef]
  313. Lazzarini, R.; Caffarini, M.; Delli Carpini, G.; Ciavattini, A.; Di Primio, R.; Orciani, M. From 2646 to 15: Differentially Regulated MicroRNAs between Progenitors from Normal Myometrium and Leiomyoma. Am. J. Obstet. Gynecol. 2020, 222, 596.e1–596.e9. [Google Scholar] [CrossRef]
  314. Zota, A.R.; Geller, R.J.; VanNoy, B.N.; Marfori, C.Q.; Tabbara, S.; Hu, L.Y.; Baccarelli, A.A.; Moawad, G.N. Phthalate Exposures and MicroRNA Expression in Uterine Fibroids: The FORGE Study. Epigenetics Insights 2020, 13, 2516865720904057. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  315. Kim, M.; Kang, D.; Kwon, M.Y.; Lee, H.J.; Kim, M.J. MicroRNAs as Potential Indicators of the Development and Progression of Uterine Leiomyoma. PLoS ONE 2022, 17, e0268793. [Google Scholar] [CrossRef] [PubMed]
  316. Chuang, T.-D.; Khorram, O. Cross-Talk between MiR-29c and Transforming Growth Factor-Β3 Is Mediated by an Epigenetic Mechanism in Leiomyoma. Fertil. Steril. 2019, 112, 1180–1189. [Google Scholar] [CrossRef] [PubMed]
  317. Huang, D.; Xue, H.; Shao, W.; Wang, X.; Liao, H.; Ye, Y. Inhibiting Effect of MiR-29 on Proliferation and Migration of Uterine Leiomyoma via the STAT3 Signaling Pathway. Aging 2022, 14, 1307–1320. [Google Scholar] [CrossRef] [PubMed]
  318. Ciebiera, M.; Włodarczyk, M.; Zgliczyński, S.; Łoziński, T.; Walczak, K.; Czekierdowski, A. The Role of MiRNA and Related Pathways in Pathophysiology of Uterine Fibroids-From Bench to Bedside. Int. J. Mol. Sci. 2020, 21, 3016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Wang, W.; Zhou, L.; Wang, J.; Zhang, X.; Liu, G. Circular RNA Expression Profiling Identifies Novel Biomarkers in Uterine Leiomyoma. Cell. Signal. 2020, 76, 109784. [Google Scholar] [CrossRef] [PubMed]
  320. Suo, M.; Lin, Z.; Guo, D.; Zhang, A. Hsa_circ_0056686, Derived from Cancer-Associated Fibroblasts, Promotes Cell Proliferation and Suppresses Apoptosis in Uterine Leiomyoma through Inhibiting Endoplasmic Reticulum Stress. PLoS ONE 2022, 17, e0266374. [Google Scholar] [CrossRef]
  321. Farzaneh, F.; Saravani, M.; Esmailpoor, M.; Mokhtari, M.; Teimoori, B.; Rezaei, M.; Salimi, S. Association of HOTAIR Gene Polymorphisms and Haplotypes with Uterine Leiomyoma Susceptibility in Southeast of Iran. Mol. Biol. Rep. 2019, 46, 4271–4277. [Google Scholar] [CrossRef] [PubMed]
  322. Falahati, Z.; Mohseni-Dargah, M.; Mirfakhraie, R. Emerging Roles of Long Non-Coding RNAs in Uterine Leiomyoma Pathogenesis: A Review. Reprod. Sci. Thousand Oaks Calif. 2022, 29, 1086–1101. [Google Scholar] [CrossRef] [PubMed]
  323. Zhou, W.; Wang, G.; Li, B.; Qu, J.; Zhang, Y. LncRNA APTR Promotes Uterine Leiomyoma Cell Proliferation by Targeting ERα to Activate the Wnt/β-Catenin Pathway. Front. Oncol. 2021, 11, 536346. [Google Scholar] [CrossRef] [PubMed]
  324. Chuang, T.-D.; Quintanilla, D.; Boos, D.; Khorram, O. Long Noncoding RNA MIAT Modulates the Extracellular Matrix Deposition in Leiomyomas by Sponging MiR-29 Family. Endocrinology 2021, 162, bqab186. [Google Scholar] [CrossRef] [PubMed]
  325. Yang, E.; Xue, L.; Li, Z.; Yi, T. Lnc-AL445665.1-4 May Be Involved in the Development of Multiple Uterine Leiomyoma through Interacting with MiR-146b-5p. BMC Cancer 2019, 19, 709. [Google Scholar] [CrossRef]
  326. Akbari, M.; Yassaee, F.; Aminbeidokhti, M.; Abedin-Do, A.; Mirfakhraie, R. LncRNA SRA1 May Play a Role in the Uterine Leiomyoma Tumor Growth Regarding the MED12 Mutation Pattern. Int. J. Womens Health 2019, 11, 495–500. [Google Scholar] [CrossRef] [Green Version]
  327. Zota, A.R.; Geller, R.J.; Calafat, A.M.; Marfori, C.Q.; Baccarelli, A.A.; Moawad, G.N. Phthalates Exposure and Uterine Fibroid Burden among Women Undergoing Surgical Treatment for Fibroids: A Preliminary Study. Fertil. Steril. 2019, 111, 112–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  328. Lee, G.; Kim, S.; Bastiaensen, M.; Malarvannan, G.; Poma, G.; Caballero Casero, N.; Gys, C.; Covaci, A.; Lee, S.; Lim, J.-E.; et al. Exposure to Organophosphate Esters, Phthalates, and Alternative Plasticizers in Association with Uterine Fibroids. Environ. Res. 2020, 189, 109874. [Google Scholar] [CrossRef]
  329. Lee, J.; Jeong, Y.; Mok, S.; Choi, K.; Park, J.; Moon, H.-B.; Choi, G.; Kim, H.-J.; Kim, S.Y.; Choi, S.R.; et al. Associations of Exposure to Phthalates and Environmental Phenols with Gynecological Disorders. Reprod. Toxicol. Elmsford N. Y. 2020, 95, 19–28. [Google Scholar] [CrossRef] [PubMed]
  330. Liu, J.; Yu, L.; Castro, L.; Yan, Y.; Clayton, N.P.; Bushel, P.; Flagler, N.D.; Scappini, E.; Dixon, D. Short-Term Tetrabromobisphenol A Exposure Promotes Fibrosis of Human Uterine Fibroid Cells in a 3D Culture System through TGF-Beta Signaling. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2022, 36, e22101. [Google Scholar] [CrossRef]
  331. Kim, J.H. Analysis of the in Vitro Effects of Di-(2-Ethylhexyl) Phthalate Exposure on Human Uterine Leiomyoma Cells. Exp. Ther. Med. 2018, 15, 4972–4978. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  332. Yu, L.; Das, P.; Vall, A.J.; Yan, Y.; Gao, X.; Sifre, M.I.; Bortner, C.D.; Castro, L.; Kissling, G.E.; Moore, A.B.; et al. Bisphenol A Induces Human Uterine Leiomyoma Cell Proliferation through Membrane-Associated ERα36 via Nongenomic Signaling Pathways. Mol. Cell. Endocrinol. 2019, 484, 59–68. [Google Scholar] [CrossRef] [PubMed]
  333. Li, Z.; Lu, Q.; Ding, B.; Xu, J.; Shen, Y. Bisphenol A Promotes the Proliferation of Leiomyoma Cells by GPR30-EGFR Signaling Pathway. J. Obstet. Gynaecol. Res. 2019, 45, 1277–1285. [Google Scholar] [CrossRef] [PubMed]
  334. Mahalingaiah, S.; Hart, J.E.; Wise, L.A.; Terry, K.L.; Boynton-Jarrett, R.; Missmer, S.A. Prenatal Diethylstilbestrol Exposure and Risk of Uterine Leiomyomata in the Nurses’ Health Study II. Am. J. Epidemiol. 2014, 179, 186–191. [Google Scholar] [CrossRef] [Green Version]
  335. Gore, A.C.; Chappell, V.A.; Fenton, S.E.; Flaws, J.A.; Nadal, A.; Prins, G.S.; Toppari, J.; Zoeller, R.T. EDC-2: The Endocrine Society’s Second Scientific Statement on Endocrine-Disrupting Chemicals. Endocr. Rev. 2015, 36, E1–E150. [Google Scholar] [CrossRef]
  336. Katz, T.A.; Yang, Q.; Treviño, L.S.; Walker, C.L.; Al-Hendy, A. Endocrine-Disrupting Chemicals and Uterine Fibroids. Fertil. Steril. 2016, 106, 967–977. [Google Scholar] [CrossRef] [Green Version]
  337. Prusinski Fernung, L.E.; Yang, Q.; Sakamuro, D.; Kumari, A.; Mas, A.; Al-Hendy, A. Endocrine Disruptor Exposure during Development Increases Incidence of Uterine Fibroids by Altering DNA Repair in Myometrial Stem Cells. Biol. Reprod. 2018, 99, 735–748. [Google Scholar] [CrossRef] [Green Version]
  338. Chen, M.; Guo, J.; Ruan, J.; Yang, Z.; He, C.; Zuo, Z. Neonatal Exposure to Environment-Relevant Levels of Tributyltin Leads to Uterine Dysplasia in Rats. Sci. Total Environ. 2020, 720, 137615. [Google Scholar] [CrossRef]
  339. Elkafas, H.; Ali, M.; Elmorsy, E.; Kamel, R.; Thompson, W.E.; Badary, O.; Al-Hendy, A.; Yang, Q. Vitamin D3 Ameliorates DNA Damage Caused by Developmental Exposure to Endocrine Disruptors in the Uterine Myometrial Stem Cells of Eker Rats. Cells 2020, 9, 1459. [Google Scholar] [CrossRef]
  340. Bulun, S.E.; Simpson, E.R.; Word, R.A. Expression of the CYP19 Gene and Its Product Aromatase Cytochrome P450 in Human Uterine Leiomyoma Tissues and Cells in Culture. J. Clin. Endocrinol. Metab. 1994, 78, 736–743. [Google Scholar] [CrossRef]
  341. Bulun, S.E.; Noble, L.S.; Takayama, K.; Michael, M.D.; Agarwal, V.; Fisher, C.; Zhao, Y.; Hinshelwood, M.M.; Ito, Y.; Simpson, E.R. Endocrine Disorders Associated with Inappropriately High Aromatase Expression. J. Steroid Biochem. Mol. Biol. 1997, 61, 133–139. [Google Scholar] [CrossRef] [PubMed]
  342. Ishihara, H.; Kitawaki, J.; Kado, N.; Koshiba, H.; Fushiki, S.; Honjo, H. Gonadotropin-Releasing Hormone Agonist and Danazol Normalize Aromatase Cytochrome P450 Expression in Eutopic Endometrium from Women with Endometriosis, Adenomyosis, or Leiomyomas. Fertil. Steril. 2003, 79 (Suppl. 1), 735–742. [Google Scholar] [CrossRef]
  343. Calzada-Mendoza, C.C.; Sánchez, E.C.; Campos, R.R.; Becerril, A.M.; Madrigal, E.B.; Sierra, A.R.; Mendez, E.B.; Ocharán, E.H.; Herrera, N.G.; Ceballos-Reyes, G. Differential Aromatase (CYP19) Expression in Human Arteries from Normal and Neoplasic Uterus: An Immunohistochemical and in Situ Hybridization Study. Front. Biosci. J. Virtual Libr. 2006, 11, 389–393. [Google Scholar] [CrossRef]
  344. Bulun, S.E.; Imir, G.; Utsunomiya, H.; Thung, S.; Gurates, B.; Tamura, M.; Lin, Z. Aromatase in Endometriosis and Uterine Leiomyomata. J. Steroid Biochem. Mol. Biol. 2005, 95, 57–62. [Google Scholar] [CrossRef]
  345. Hatok, J.; Zubor, P.; Galo, S.; Kirschnerova, R.; Dobrota, D.; Danko, J.; Racay, P. Endometrial Aromatase MRNA as a Possible Screening Tool for Advanced Endometriosis and Adenomyosis. Gynecol. Endocrinol. Off. J. Int. Soc. Gynecol. Endocrinol. 2011, 27, 331–336. [Google Scholar] [CrossRef] [PubMed]
  346. Noble, L.S.; Simpson, E.R.; Johns, A.; Bulun, S.E. Aromatase Expression in Endometriosis. J. Clin. Endocrinol. Metab. 1996, 81, 174–179. [Google Scholar] [CrossRef] [Green Version]
  347. Noble, L.S.; Takayama, K.; Zeitoun, K.M.; Putman, J.M.; Johns, D.A.; Hinshelwood, M.M.; Agarwal, V.R.; Zhao, Y.; Carr, B.R.; Bulun, S.E. Prostaglandin E2 Stimulates Aromatase Expression in Endometriosis-Derived Stromal Cells. J. Clin. Endocrinol. Metab. 1997, 82, 600–606. [Google Scholar] [CrossRef]
  348. Shozu, M.; Sumitani, H.; Segawa, T.; Yang, H.-J.; Murakami, K.; Kasai, T.; Inoue, M. Overexpression of Aromatase P450 in Leiomyoma Tissue Is Driven Primarily through Promoter I.4 of the Aromatase P450 Gene (CYP19). J. Clin. Endocrinol. Metab. 2002, 87, 2540–2548. [Google Scholar] [CrossRef]
  349. Imir, A.G.; Lin, Z.; Yin, P.; Deb, S.; Yilmaz, B.; Cetin, M.; Cetin, A.; Bulun, S.E. Aromatase Expression in Uterine Leiomyomata Is Regulated Primarily by Proximal Promoters I.3/II. J. Clin. Endocrinol. Metab. 2007, 92, 1979–1982. [Google Scholar] [CrossRef] [Green Version]
  350. Ishikawa, H.; Reierstad, S.; Demura, M.; Rademaker, A.W.; Kasai, T.; Inoue, M.; Usui, H.; Shozu, M.; Bulun, S.E. High Aromatase Expression in Uterine Leiomyoma Tissues of African-American Women. J. Clin. Endocrinol. Metab. 2009, 94, 1752–1756. [Google Scholar] [CrossRef]
  351. Thompson, P.A.; Khatami, M.; Baglole, C.J.; Sun, J.; Harris, S.A.; Moon, E.-Y.; Al-Mulla, F.; Al-Temaimi, R.; Brown, D.G.; Colacci, A.; et al. Environmental Immune Disruptors, Inflammation and Cancer Risk. Carcinogenesis 2015, 36 (Suppl. 1), S232–S253. [Google Scholar] [CrossRef] [Green Version]
  352. Sacco, K.; Portelli, M.; Pollacco, J.; Schembri-Wismayer, P.; Calleja-Agius, J. The Role of Prostaglandin E2 in Endometriosis. Gynecol. Endocrinol. Off. J. Int. Soc. Gynecol. Endocrinol. 2012, 28, 134–138. [Google Scholar] [CrossRef] [PubMed]
  353. Kim, H.G.; Jung, G.Y.; Park, S.B.; Cho, Y.J.; Han, M. Assessment of the Effects of Prostaglandins on Myometrial and Leiomyoma Cells in Vitro through MicroRNA Profiling. Mol. Med. Rep. 2018, 18, 2499–2505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  354. Barcikowska, Z.; Rajkowska-Labon, E.; Grzybowska, M.E.; Hansdorfer-Korzon, R.; Zorena, K. Inflammatory Markers in Dysmenorrhea and Therapeutic Options. Int. J. Environ. Res. Public Health 2020, 17, 1191. [Google Scholar] [CrossRef] [Green Version]
  355. Ye, Y.; Wang, X.; Jeschke, U.; von Schönfeldt, V. COX-2-PGE2-EPs in Gynecological Cancers. Arch. Gynecol. Obstet. 2020, 301, 1365–1375. [Google Scholar] [CrossRef] [PubMed]
  356. Baranov, V.S.; Ivaschenko, T.E.; Yarmolinskaya, M.I. Comparative Systems Genetics View of Endometriosis and Uterine Leiomyoma: Two Sides of the Same Coin? Syst. Biol. Reprod. Med. 2016, 62, 93–105. [Google Scholar] [CrossRef] [Green Version]
  357. Roy, D.; Liehr, J.G. Estrogen, DNA Damage and Mutations. Mutat. Res. 1999, 424, 107–115. [Google Scholar] [CrossRef]
  358. Yager, J.D. Mechanisms of Estrogen Carcinogenesis: The Role of E2/E1-Quinone Metabolites Suggests New Approaches to Preventive Intervention--A Review. Steroids 2015, 99, 56–60. [Google Scholar] [CrossRef] [Green Version]
  359. Tian, H.; Gao, Z.; Wang, G.; Li, H.; Zheng, J. Estrogen Potentiates Reactive Oxygen Species (ROS) Tolerance to Initiate Carcinogenesis and Promote Cancer Malignant Transformation. Tumour Biol. J. Int. Soc. Oncodevelopmental Biol. Med. 2016, 37, 141–150. [Google Scholar] [CrossRef]
  360. Bulun, S.E.; Moravek, M.B.; Yin, P.; Ono, M.; Coon, J.S.; Dyson, M.T.; Navarro, A.; Marsh, E.E.; Zhao, H.; Maruyama, T.; et al. Uterine Leiomyoma Stem Cells: Linking Progesterone to Growth. Semin. Reprod. Med. 2015, 33, 357–365. [Google Scholar] [CrossRef]
  361. Kastner, P.; Krust, A.; Turcotte, B.; Stropp, U.; Tora, L.; Gronemeyer, H.; Chambon, P. Two Distinct Estrogen-Regulated Promoters Generate Transcripts Encoding the Two Functionally Different Human Progesterone Receptor Forms A and B. EMBO J. 1990, 9, 1603–1614. [Google Scholar] [CrossRef] [PubMed]
  362. Omar, M.; Laknaur, A.; Al-Hendy, A.; Yang, Q. Myometrial Progesterone Hyper-Responsiveness Associated with Increased Risk of Human Uterine Fibroids. BMC Womens Health 2019, 19, 92. [Google Scholar] [CrossRef] [PubMed]
Figure 1. An illustration of predisposing factors intricacy and intratumoral heterogeneity complicating the search for key events of uterine leiomyoma (UL) genesis. UL formation is associated with a wide range of predisposing factors: family medical history, African or Latin American ethnicity, gene variants, epigenetic changes, early menarche, exposure to xenoestrogens, obesity, hypertension, vitamin D deficiency, alcohol intake, dietary habits, nulliparity and advanced reproductive age. During UL growth, each cell acquires a unique set of parameters (cell type, genetic and epigenetic alterations, characteristics of signalling and microenvironmental components), thus resulting in a pronounced intratumoral heterogeneity and complicating identification of the characteristics of initial UL cell(s).
Figure 1. An illustration of predisposing factors intricacy and intratumoral heterogeneity complicating the search for key events of uterine leiomyoma (UL) genesis. UL formation is associated with a wide range of predisposing factors: family medical history, African or Latin American ethnicity, gene variants, epigenetic changes, early menarche, exposure to xenoestrogens, obesity, hypertension, vitamin D deficiency, alcohol intake, dietary habits, nulliparity and advanced reproductive age. During UL growth, each cell acquires a unique set of parameters (cell type, genetic and epigenetic alterations, characteristics of signalling and microenvironmental components), thus resulting in a pronounced intratumoral heterogeneity and complicating identification of the characteristics of initial UL cell(s).
Ijms 24 05752 g001
Figure 2. The hypothetic timeline of uterine leiomyoma (UL) genesis. A prerequisite to UL formation may appear in parental gametogenesis and could be presented by altered epigenetic patterns in oocytes or spermatozoa. The accumulation of UL-associated negative effects during ontogenesis can lead to an acute imbalance of risk and protective factors and the beginning of UL formation. UL growth takes a long period of time, during which driver (MED12 and FH mutations, HMGA2 rearrangements and overexpression, COL4A5-COL4A6 rearrangements) and passenger (7q deletions, chromosome 1 rearrangements, heteroploidy, etc.) changes occur and contribute to the intratumoral heterogeneity.
Figure 2. The hypothetic timeline of uterine leiomyoma (UL) genesis. A prerequisite to UL formation may appear in parental gametogenesis and could be presented by altered epigenetic patterns in oocytes or spermatozoa. The accumulation of UL-associated negative effects during ontogenesis can lead to an acute imbalance of risk and protective factors and the beginning of UL formation. UL growth takes a long period of time, during which driver (MED12 and FH mutations, HMGA2 rearrangements and overexpression, COL4A5-COL4A6 rearrangements) and passenger (7q deletions, chromosome 1 rearrangements, heteroploidy, etc.) changes occur and contribute to the intratumoral heterogeneity.
Ijms 24 05752 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Koltsova, A.S.; Efimova, O.A.; Pendina, A.A. A View on Uterine Leiomyoma Genesis through the Prism of Genetic, Epigenetic and Cellular Heterogeneity. Int. J. Mol. Sci. 2023, 24, 5752. https://doi.org/10.3390/ijms24065752

AMA Style

Koltsova AS, Efimova OA, Pendina AA. A View on Uterine Leiomyoma Genesis through the Prism of Genetic, Epigenetic and Cellular Heterogeneity. International Journal of Molecular Sciences. 2023; 24(6):5752. https://doi.org/10.3390/ijms24065752

Chicago/Turabian Style

Koltsova, Alla S., Olga A. Efimova, and Anna A. Pendina. 2023. "A View on Uterine Leiomyoma Genesis through the Prism of Genetic, Epigenetic and Cellular Heterogeneity" International Journal of Molecular Sciences 24, no. 6: 5752. https://doi.org/10.3390/ijms24065752

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop