Next Article in Journal
Antiviral Activity of Quercetin Hydrate against Zika Virus
Next Article in Special Issue
Exploring the Impact of Endometrioma Aspiration and Dienogest Combination Therapy on Cyst Size, Inflammatory Cytokines in Follicular Fluid and Fertility Outcomes
Previous Article in Journal
Characterization of PcSTT3B as a Key Oligosaccharyltransferase Subunit Involved in N-glycosylation and Its Role in Development and Pathogenicity of Phytophthora capsici
Previous Article in Special Issue
Does the Use of the “Proseek® Multiplex Inflammation I Panel” Demonstrate a Difference in Local and Systemic Immune Responses in Endometriosis Patients with or without Deep-Infiltrating Lesions?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Lifelong Impact on Endometriosis: Pathophysiology and Pharmacological Treatment

1
Department of Obstetrics and Gynecology, Linkou Medical Center, Chang Gung Memorial Hospital, College of Medicine, Chang Gung University, Taoyuan 33305, Taiwan
2
Graduate Institute of Clinical Medical Sciences, College of Medicine, Chang Gung University, Taoyuan 33302, Taiwan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(8), 7503; https://doi.org/10.3390/ijms24087503
Submission received: 6 February 2023 / Revised: 6 April 2023 / Accepted: 16 April 2023 / Published: 19 April 2023

Abstract

:
Endometriosis is a chronic inflammatory disease associated with bothersome symptoms in premenopausal women and is complicated with long-term systemic impacts in the post-menopausal stage. It is generally defined by the presence of endometrial-like tissue outside the uterine cavity, which causes menstrual disorders, chronic pelvic pain, and infertility. Endometriotic lesions can also spread and grow in extra-pelvic sites; the chronic inflammatory status can cause systemic effects, including metabolic disorder, immune dysregulation, and cardiovascular diseases. The uncertain etiologies of endometriosis and their diverse presentations limit the treatment efficacy. High recurrence risk and intolerable side effects result in poor compliance. Current studies for endometriosis have paid attention to the advances in hormonal, neurological, and immunological approaches to the pathophysiology and their potential pharmacological intervention. Here we provide an overview of the lifelong impacts of endometriosis and summarize the updated consensus on therapeutic strategies.

1. Introduction

Endometriosis is a common disease arising in adolescents that affects about 6–10% of women of reproductive age [1]. It progresses throughout menstrual cycles and involves multiple organs, resulting in local gynecologic lesions and systemic inflammatory disorders. Endometriosis is associated with a wide range of presentations, and its common symptoms include dysmenorrhea, dyspareunia, pelvic pain, dyschezia, and hematochezia. Furthermore, asymptomatic endometriotic lesions can be detected in nearly half of the women seeking infertility treatment [2]. As a condition of chronic inflammation and immune dysregulation, women with endometriosis are at higher risk of developing cardiovascular disease, rheumatoid arthritis, asthma, melanoma, ovarian cancer, and breast cancer [3].
Endometriosis is defined by the presence of endometrial-like tissue (‘‘lesions’’) outside the uterine cavity confirmed during surgery, where the diagnosis is often delayed after the onset of symptoms and mistaken because of its nonspecific complaints [1,4]. Nowadays, the diagnosis of endometriosis can be accelerated by advanced imaging techniques and associated serum biomarkers [5]. The conventional treatment includes surgical removal of endometriotic lesions followed by hormonal suppression. Current pharmacological treatments have limited efficacy and unwanted side effects. Half of the women undergoing surgery without long-term medication control may have another procedure in 5 years, resulting in organ damage complicated with loss of function [6]. Current therapeutic strategies highlight enduring symptom relief and fertility preservation [4,7]. The theories address whether the endometriotic cells travel to an abnormal location in a consequent mechanism or whether the endometriotic cells pre-exist in a milieu of genetic or epigenetic changes. A better understanding of the pathobiology of endometriosis in hormonal, inflammatory, metabolic, and pain pathways helps develop novel pharmacologic targets for clinical trials. In the present review, we provide an overview of the lifelong impact of endometriosis and summarize the current practice strategies in pathophysiologic and pharmacological aspects.

2. The Pathophysiology of Endometriosis

2.1. Retrograde Menstruation, Coelomic Metaplasia, and Müllerian Remnants Theory

Endometriosis is an “ectopic” endometriotic lesion that resembles the phenotype of the “eutopic” endometrial lining of the uterus. The most widely accepted hypothesis is retrograde menstruation that the efflux of menstrual tissue fragments via the fallopian tubes introduces endometrial cells seeding and growing into the peritoneal cavity [8]. Retrograde menstruation is a physiological event that occurs during menses, but in women with endometriosis, this viable steroid-responsive endometrial tissue adheres to the peritoneum and invades the pelvic structures. Higher risks for developing endometriosis in women with menarche at a younger age, menopause at an older age, long duration, and heavy menstrual flow support this hypothesis [9,10]. However, it is not sufficient to explain hereditary changes or clonal aspects of endometriosis in men, and its associated diverse clinical presentations.
Coelomic metaplasia, another well-recognized theory for forming endometriotic lesions, is based on the peritoneal mesothelium transformation. Moreover, the Müllerian remnants hypothesis proposed that endometriosis can be differentiated from embryological remnants. The mechanism that endometriotic lesions originate “in-situ” by metaplasia or from Müllerian remnants further explains the endometriosis in adolescents shortly after menarche and in fetuses with the absence of menstruation [11,12,13].

2.2. Circular Dissemination and Stem Cell Theory

For uncommon extraperitoneal locations, Sampson suggests that the shedding cells enter the uterine vasculature or lymphatic system, disseminating to distant organs, including the lung, liver, spleen, and brain [8,14,15]. An alternative theory has become clear that stem cell contributes to the pathogenesis of endometriosis [16]. Bone marrow-derived stem cells (BMDSCs) travel to the uterine cavity and regenerate eutopic endometrium. During menstruation, women with endometriosis have more pluripotential cells and shed more progenitor cells than healthy women, further expanding the hypothesis of retrograde menstruation [17,18]. These BMDSCs can directly differentiate into endometrial cells at ectopic locations in the peritoneal cavity and distant sites [19,20].

2.3. Invagination Theory in Adenomyosis

Endometriosis of uterus is characterized by a junctional zone disease, resulting from altered endometrial basalis cells invading the uterine myometrium through disruption of the endometrial-myometrial interface (EMI), subsequently establishing ectopic endometriotic lesions [21,22]. Higher risks of adenomyosis are found in women with microtrauma of EMI, such as repeated endometrial curettage, cesarean delivery, and previous uterine surgery [23]. Furthermore, cycles of sustained uterine hyperperistalsis and repeated myocyte overstretching augment auto-traumatization, potentially promoting the invagination of endometrial basalis in the tissue healing process [24,25,26] (Figure 1).

2.4. Epigenomic and Genomic Alterations

Several studies have reported the relationship between heritability and endometriosis. The results of family aggregation studies suggest that the probability of developing endometriosis is 8% from a diseased mother and 6% from an affected sister [27,28]. The risk of developing the disease is less than 1% in both situations in the control population without a family history. The increased disease prevalence has been found in first-degree relatives of women with endometriosis, although the specific genetic origin of the association remains unknown. Genome-wide association studies (GWAS) have investigated samples from thousands of women in the United States, Australia, Europe, and Japan [29,30,31]. These new tools have identified genetic variations (Single Nucleotide Polymorphisms, SNPs) that are thought to be missense mutations in patients with endometriosis, among which the following genes involved in the development of endometriosis through cell proliferation, migration, and adhesion: VETZ, WNT-4, GREB1, CDKN2B-AS1 and ID4. VETZ participates in cell adhesion, migration and transmembrane cell junction, which is associated with epithelial mesenchymal transition; WNT-4 plays decisive roles during the development of female reproductive system; GREB1 is responsible for estrogen regulation; CDKN2B-AS1, a tumor suppressor gene and ID4, an ovarian oncogene have been implicated in molecular pathogenesis of endometriosis. The specific loci strongly correlate SNPs with advanced disease development [28,32,33,34]. Unsurprising, the common SNPs associated with endometriosis have been detected overlap compared to other gynecological disorders such as infertility, fibroids, and cancer [35,36,37]. The similar loci between SNP datasets implicate the shared pathogenesis of other gynecologic diseases that the genes significantly associated with endometriosis risk are close to genes involved in sex steroid hormone pathways [38], MAP kinase signaling cascade [39], interleukin 1A (a cytokine implicated in inflammatory responses) [31], WNT signaling [36], and steroid metabolism [38]. With regard to genetic markers obtained from GWAS to identify high-risk populations for developing endometriosis, each functional group involved in pathogenesis has at least one or more genes that link with endometriosis. Nevertheless, until now, no genetic tests could be considered reliable for clinical diagnosis; translating findings into validated tests should only be performed within a research setting [40].
Studies on the impact of epigenetic modifications expanded rapidly to date. Without altering the DNA sequence, these regulatory mechanisms modify gene expression through DNA methylation, histone modification, and noncoding RNA regulation [34]. The first discovered evidence showed that the promoter of the HOXA10 gene is hypermethylated in women with endometriosis compared to healthy individuals. Methylation of the Homeobox A (HOXA10) leads to gene silencing, which reduces the expression of E-cadherin, favoring the process of cell invasion from the weak intercellular junctions [41,42]. Additionally, progesterone resistance is generated by hypermethylation of the progesterone receptor-β (PR-β), contributing to compromised stromal epithelial crosstalk [43,44,45]. Some hypomethylated genes, including estrogen receptor-β (ER-β), steroidogenic factor-1 (SF-1), and aromatase, are overexpressed, increasing estrogen and its receptor, respectively [46].
Histone modification, characterized by changes in chromatin structure, provides or blocks access of effectors modulators and of transcription factors to their binding sequences in gene promoters, leading to gene deregulation and disease. Generally, histone methylation causes gene silencing while histone acetylation contributes to activation of gene expression. Post-translational histone modification within promoter regions of selected genes has been reported in endometriosis [47,48]. Hypoacetylation of H3/H4 within promotor regions of target genes known to be downregulated in endometriosis, such as HOXA10. Hyperacetylation of H3/H4 within promotor regions of SF-1, a transcription factor involved in estrogen biosynthesis, results in overexpression of aromatase and local estradiol. These data contribute to a better understanding of gene expression regulated by histone modification.
Noncoding RNAs (ncRNAs), which are not translated into a protein, serve to regulate chromosome structure, interact with messenger RNA, and usually inhibit gene expression at a post-transcriptional level [49]. MicroRNAs (miRNAs) are small single-stranded ncRNAs. The endometrium’s miRNA signatures change with the different phases of menstrual cycles [50,51]. Hundreds of dysregulated miRNAs have been found in paired eutopic and ectopic endometriotic lesions of women with and without endometriosis [52,53]. Extracellular miRNAs are found in circulation [54] and can potentially mediate intercellular communication between eutopic endometrium and ectopic endometriotic lesions [55]. Up- or down-regulation of miRNA plays a crucial role in the endometriotic implant establishment [56,57,58] and provides potential therapeutic targets in the future.

2.5. Estrogen and Progesterone Modulation

In both normal endometrium and ectopic endometriotic lesions, steroid hormones and their receptors regulate cell proliferation, angiogenesis, neurogenesis, and inflammatory pathways. Properly balanced concentrations of estrogen and progesterone regulate functional eutopic endometrium during menstrual cycles. Endometriosis is also defined as a ‘‘steroid-dependent’’ disorder, which depends on its cell-specific patterns of steroid receptor expression and menstrual cycle phase-dependent hormone metabolism [9,44]. Estradiol activates cyclooxygenase-2 (COX-2) within uterine endothelial cells, increasing prostaglandin E2 production in a feed-forward mechanism [59]. An increase in estradiol production by aromatase, the loss function of 17-hydroxysteroid dehydrogenase type 2 (17β-HSD2), and the overexpression of ER-β promote cell growth and perpetuated inflammation in ectopic endometriotic lesions [60,61,62]. The accumulation of estradiol activates mitogenic activity by stimulating a series of genes related to cell proliferation (GREB1, MYC, and CCND1), inhibiting the apoptosis induced by apoptosis signal-regulating kinase-1 (ASK-1) and tumor necrosis factor-α (TNF-α), leading to the development of endometriotic implants [63].
In contrast, downregulation of PR-β results in progesterone resistance, which causes a nonreceptive eutopic endometrium and a pro-inflammatory microenvironment limiting the effects of progestin therapy and further driving the systemic impact of endometriosis [34,64]. With regard to strong progesterone resistance in endometriotic lesions, the effect of hormone therapy on superficial endometriotic lesions seems to be a consequence of the decreased estrogen concentrations rather than a direct progestin effect [65]. The optimal solution would be partially reducing estrogen levels just enough to suppress survival and vascularization of endometriotic implants, while at the same time maintaining adequate concentrations to alleviate vasomotor menopausal symptoms and bone mineral density loss.

2.6. Inflammation, Angiogenesis, and Tissue Remodeling

Endometriotic implants are complex multicellular structures that ectopic endometrial cells migrate, adhere, and evade through a serial process of tissue remodeling, followed by the influx of pro-inflammatory cytokines and the growth of new blood vessels (angiogenesis) [66] (Figure 2). Peritoneal fluid in affected patients is also found to contain increased pro-inflammatory cytokines [67,68,69]. The aberrantly increased concentrations of interleukins (IL-1β, IL-6, IL-8, IL-33), tumor necrosis factor-alpha (TNF-α), insulin-like growth facto-1 (IGF-1), monocyte chemoattractant protein/C-C motif chemokine ligand (MCP-1 CCL2, CCL5), and vascular endothelial growth factor (VEGF) activate the inflammatory response by upregulating nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) in affected women [70,71,72,73,74,75,76,77]. Circulating cytokines and immune cells further create a widespread inflammatory environment which drives the systemic effect of endometriosis on immunologic, cardiovascular, neurological, and metabolic function [64,71,76,78,79]. VEGF/tyrosine kinase signaling pathway has been upregulated and involved in numerous mechanisms of vascularization, including de novo growth (angiogenesis), vasculogenesis, and the formation of interconnected networks. Furthermore, the link between the growth of new blood vessels and nerve fibers contributes to the “neuro-angiogenesis”, ectopic endometriotic lesions, and pain pathways [80,81].
Stromal fibroblasts, with the phenotype of clonogenic and multilineage potential, contribute to multicellular lesions at extra-uterine locations [45,82]. In women with an aberrant response to estrogen, the endometriotic lesions formed from tissue fragments regulates by transient hypoxia [83], the release of iron, and the activation of platelets [84,85].

2.7. Immune Dysregulations

Aberrant production of pro-inflammatory cytokines recruits a large pool of immune cell populations which alters the peritoneal environment in women with endometriosis [1,4,86]. The abundance of innate immune cells and different populations of adaptive immune cells has been detected in the peritoneal fluid of affected women or the endometriotic lesions from patients implicating a compromised immune system in endometriosis [1,87,88]. The neutrophil chemotactic factors such as IL-8, granulocyte colony stimulating factor (G-CSF), and chemokine ligands 1, 2, and 3 (CXCL-1, CXCL-2, and CXCL-3) gather the immune cells in a loop of positive feedback. In peritoneal fluid and eutopic endometrium of women with endometriosis, a significant increase of macrophages has found a decrease in the phagocytic activity [89,90], which promotes angiogenesis [91], lesion innervation [68,91] and pain symptoms [92]. Activated macrophages are divided into two phenotypes, among which M1 macrophages are dominant in pro-inflammatory responses and M2 macrophages are mainly involved in anti-inflammatory responses. Although the studies remain controversial, the affected women have both physiological endometria with M1 predominates and ectopic portions with M2 polarization, allowing angiogenesis, tissue remodeling, and thus the development of the disease [32,86,90,93,94]. Survival of ectopic lesions is also provided by the activation of T and B cells and the decreased cytotoxicity of natural killer (NK) cells [95,96]. The cytokines, including the platelet-derived transforming growth factor β (TGF-β), IL-6, and IL-15, inhibit the cytotoxicity of NK cells, thus contributing to the implantation, proliferation, and immune escape of ectopic endometrial cells [95,96,97].

3. Clinical Features of Endometriosis and Its Lifelong Impacts

Endometriosis causes heterogeneous presentation, varying from superficial peritoneal lesions, ovarian(endometrioma) and uterine(adenomyosis) tumors, and deep infiltrative endometriosis (DIE), which is often accompanied by scarring and adhesions [1,4]. These lesions are associated with gynecological disorders and the variability of pain symptoms. Endometriotic lesions can also spread and grow in extra-pelvic sites, including visceral organs in the upper abdomen, chest (thoracic endometriosis), brain, and nerve systems [98] (Figure 3). Although none of the biomarkers displayed enough accuracy, understanding the diffuse clinical presentations and diverse disease patterns of endometriosis helps early identification and intervention.

3.1. Menstrual Disorders in Adenomyosis

Gynecologic disorders such as dysmenorrhea, menorrhagia, and abnormal uterine bleeding are the main symptoms in women with adenomyosis. In adenomyosis-affected myometrium, higher expression of oxytocin receptors and altered membrane depolarization of uterine smooth muscle cells contribute to abnormal uterine contractility [99,100,101]. The increased expression of TNF-α promotes the production of IL-1β [102], activation of the NF-κB pathway [103], and engagement of IL-18/IL-18R complex [104,105], resulting in pain symptoms by mediating prostaglandins synthesis. The severity of dysmenorrhea also correlates with immunoreactivity, neuropathologic factor, and microvascular function [106]. Abnormal uterine contractility and high microvessel density further cause heavy menstrual bleeding.

3.2. Endometriosis-Associated Symptoms

The common symptoms of pelvic endometriosis are chronic pelvic pain (cyclical and non-cyclical) and other pain conditions, including painful sexual intercourse (dyspareunia), painful urination (dysuria), and painful defecation (dyschezia). The severity of pelvic endometriosis can be determined after surgical intervention using the revised scoring system of the American Society for Reproductive Medicine (ASRM) [107] (Supplementary Figure S1). The stage of pelvic endometriosis is not always correlated with patient-reported pain symptoms [108], and a lack of awareness of deep infiltrative endometriosis and extra-pelvic endometriosis may delay diagnosis. Some women experience recurrent, particularly perimenstrual, changes in bowel habits (diarrhea or constipation), irritable bowel syndromes, and bloody stool implicating the evidence of deep infiltrative endometriosis of the lower gastrointestinal tract. Other women suffering from cyclic dysuria and hematuria have been treated as refractory urinary tract infections or bladder pain syndrome, which can be caused by endometriosis [109]. Endometriosis of the diaphragm and pleura has been associated with chest and shoulder pain [110]. Neurological changes and chronic inflammation in endometriosis enhance pain perception, anxiety, fatigue, and depression [111,112].

3.3. Endometriosis-Associated Infertility

Endometriosis should be considered a cause of infertility in women with pain symptoms. Avoiding sexual intercourse due to severe dyspareunia and chronic pelvic pain limits the feasibility of natural conception [113,114]. Pelvic adhesion can also cause anatomical distortion, interrupting the conception process, including oocyte release from ovaries, ovum pickup, and transport through fallopian tubes [115]. Secondly, oxidative stress in an endometrioma causes damage to the adjacent healthy ovarian cortex, reducing ovarian reserve [116,117]. Alterations of the intrafollicular microenvironment and aberrant steroidogenesis impair folliculogenesis and oocyte competence [118,119]. Dysregulation of immune and inflammatory profiles plays an essential role in recurrent implantation failure and early pregnancy loss [49,120,121]. Adenomyosis, characterized by the defective junctional zone and perturbed uterine peristalsis, is strongly associated with primary infertility and adverse in vitro fertilization (IVF) outcomes [122,123].

3.4. Endometriosis-Associated Obstetric Complications

Emerging research has demonstrated the relationship between endometriosis and obstetric complications, including miscarriage, preterm birth, preterm premature rupture of membranes, antepartum hemorrhage, placental abruption, placenta previa, preeclampsia, gestational hypertensive and metabolic disorders (diabetes or cholestasis), and adverse neonatal outcomes (small for gestational age, low birth weight, admission to neonatal intensive care and neonatal death) [124,125]. Endometriosis and obstetric diseases share some molecular features and pathophysiologic mechanisms of the defective junctional zone, perturbed uterine peristalsis, and aberrant inflammation. Several differentially expressed genes involved in endometriosis are common in adverse pregnancy outcomes such as preeclampsia, small for gestational age, or preterm birth. Alterations of imprinted gene clusters (CDKN1C, DLX5, GATA3) in the link between endometriosis and abnormal decidualization are considered critical regulators of embryogenesis and placentation [126,127,128,129,130,131,132]. Adverse maternal environments can lead to placental genetic and epigenetic aberrant, which alters the placenta’s ability to modulate fetal exposure and response to maternal cortisol, causing infant neurobehavioral deficits [133]. In addition to suboptimal placentation, overexpression of COX-2 and prostaglandins secretion in chronic inflammation can lead to early cervical ripening and uterine hypercontractility in women with endometriosis, thus causing adverse fetal outcomes [134,135,136].

3.5. Malignancy Potential

A recent meta-analysis has reported that endometriosis is associated with an increased risk of certain cancers, such as ovarian, breast, and thyroid cancers [137]. Endometriosis induces the microenvironment with an aberrant immune response and altered hormonal milieu [60], favoring neoplastic transformation. Accumulations of oxidative stress and chronic inflammatory response contribute to the development and progression of endometriosis-associated malignancies. Somatic mutations in the genes of women with endometriosis have been recognized as a precursor of malignant transformation [138]. Several genetic studies have discovered that mutations or alterations in genes (PTEN, TP53, KRAS, and ARID1A) of endometriosis are directly related to neoplasms [139,140,141,142,143]. Atypical endometriosis, as a histologically borderline tumor corresponding to the features of hyperchromatic nuclei, cellular crowding, and the high nucleus-to-cytoplasm ratio [144,145], has a greater risk of malignant change to clear cell and endometrioid ovarian cancers [146]. Recent evidence of a link between endometriosis and malignant potential has raised concerns in the long-term management of patients with endometriosis through the lifetime from puberty to post-menopause [147].

3.6. Long-Term Systemic Diseases

Endometriosis, rather than a localized pelvic disease, has a detrimental effect on cardiovascular, neurological, metabolic, and immune function, stemming from circulating pro-inflammatory cytokine and shifts in immune cell populations [148]. Increasing research has reported that differential gene expression in endometriosis alters metabolism in the liver and adipose tissue leading to systemic inflammation [149,150,151] but also affects gene expression in the brain causing pain sensitization, anxiety, and depression [111,152,153]. These data suggest that a variety of metabolic phenotypes in endometriosis results in a life course of systemic effects. Early recognition and management of all aspects of the disease can relieve problematic symptoms and avoid long-term sequelae.

4. Pharmacologic Therapies in Current Clinical Practice

Clinical management of endometriosis-associated symptoms depends on the disease’s severity, extent, and location. The choice of treatments includes medication, surgery, or a combination of both. Pharmacological therapy for endometriosis aims to relieve symptoms, maintain long-term control, or prevent recurrence after surgical removal of lesions. However, therapeutic windows in the life course of endometriosis are challenging due to unwanted side effects and the desire to conceive.

4.1. Hormonal Manipulation

Hormonal targets on presumed altered steroidogenesis in endometriosis act by suppressing fluctuations in gonadotropic and ovarian hormones, establishing either a hypo-estrogenic or hyper-progestogenic milieu, resulting in the inhibition of ovulation and the reduction of menstrual bleeding [9,154] (Figure 4).

4.1.1. Progestin-Based Therapies

Combined oral contraceptives (COCs), containing estrogen and progestin, have been extensively prescribed in clinical practice due to their efficacy in managing dysmenorrhea. It is particularly beneficial in continuous rather than cyclic administration [1,40,156,157]. Progestin-dominant COCs can establish a hyper-progestogenic status, inducing decidualization and subsequent apoptosis of ectopic endometriotic implants [158]. The estrogen component induces central inhibition of gonadotropin secretion, inhibiting ovulation and reducing ovarian estrogen production, creating a hypo-estrogenic milieu [139]. However, the main concern about the off-labeled use of COCs in endometriosis is that the estrogen content of COCs potentially contributes to the progression of endometriosis [159,160].
Progestins [155], acting as natural progesterone, can induce anovulation and endometrial pseudo-decidualization, resulting in the atrophy of endometriotic implants by decreasing inflammation and angiogenesis [161,162]. Progestin-only pills, including dienogest, norethisterone, and medroxyprogesterone, are currently the first-line treatment for symptomatic endometriosis and aim to prevent recurrence after surgery [163,164,165]. Additionally, progestins can be administered by other routes such as intramuscular, subcutaneously (etonogestrel implant), or intrauterine (levonorgestrel-releasing intrauterine device, LNG-IUD) [166]. Dienogest (2 mg daily), a 19-nortestosterone derivative, can increase PRβ expression in endometriotic lesions, potentially overcoming progesterone resistance [167]. Several randomized controlled trials have proved its efficacy for endometriosis-associated pain regarding different phenotypes [165,168,169,170]. Dienogest may decrease the size of ovarian endometrioma without decreasing ovarian reserve [171,172,173] and reduce pain symptoms related to deep infiltrating endometriosis (dysmenorrhea, dyspareunia, dyschezia, and chronic pelvic pain), thus improving patients’ quality of life. Norethisterone acetate (NETA, 2.5–15 mg daily), a 19-nortestosterone derivative, was confirmed as effective as dienogest in reducing the size of ovarian endometrioma and endometriosis-related symptoms, whereas dienogest was superior in symptoms relief and tolerability [172,174,175].
Medroxyprogesterone acetate (MPA, 10–60 mg daily), a 17OH-progesterone derivative available as oral or depot formulation (administered every three months subcutaneously or intramuscularly), is as effective as Gonadotrophin releasing hormone analogs (GnRH agonist) and limited to treat refractory endometriosis due to long-term hypo-estrogenic status, consequently leading to bone loss [176,177]. LNG-IUD, a potent 19-nortesterone derivative released directly into the uterine cavity, improves menstrual disorders and pelvic pain symptoms related to endometriosis [178,179,180].
Commonly reported side effects of progestin-based therapies are abnormal uterine bleeding, which progressively improves with the continuation of treatment, headache, mood changes, and particularly for long-term use of depot MPA, as well as loss of bone marrow density [181].

4.1.2. GnRH Agonists

GnRH agonist, substituting a D-amino acid for the native L-amino acid at position six of the native GnRH peptides, initially stimulates gonadotrophins secretion (the flare effect) and subsequently reduces estrogen production by downregulation and desensitization of the pituitary GnRH receptors [182]. The induced hypo-estrogenic status can cause the regression of endometriotic lesions, but the prolonged receptor occupancy leads to vasomotor symptoms, vaginal dryness, reduced libido, sleep disturbance, mood disorder, and bone loss [183]. Add-back therapy (the addition of progestin alone or COCs) has been advocated to extend pain relief up to 10 years of treatment by minimizing the adverse effects of estrogen deprivation [184,185,186]. However, GnRH agonist remains a second-line treatment for endometriosis because of the high cost and the limitation of long-term maintenance [148].

4.1.3. GnRH Antagonists

GnRH antagonists, similar in structure to natural GnRH, suppress pituitary function through direct competition with GnRH receptors and thus induce a rapid drop of estrogen without provoking the flare effect [187]. Upcoming clinical trials have explored the efficacy and safety of oral nonpeptidic GnRH antagonists, including Elagolix, Relugolix, and Linzagolix, for treating endometriosis-associated pain [188]. Elagolix (150 mg daily to 200 mg twice daily), a uracil derivative, the first approved compound, is effective in reducing moderate and severe endometriosis-associated pain for six months of treatment in two phase III trials (Elaris Endometriosis I and II) [189] and improving quality of life for six additional months in phase III extension studies (Elaris EM-III and EM-IV) [190,191,192]. Relugolix (10–20–40 mg daily), a thieno [2,3-d]pyrimidine-2,4-one derivative, alleviates endometriosis-associated pain in a dose-response manner for the 12-week treatment in a Phase II trial [193] and with similar results for 24-week therapy in two other replicate phase III trials (SPIRIT-1 and 2) [194,195]. Linzagolix (75–100–200 mg daily), the newest compound of GnRH antagonists, has been investigated for 24-week treatment in a phase IIb trial [196] and in the extension study up to 52 weeks [197], which showed promising therapeutic effects on pain symptoms in women with endometriosis. Like GnRH agonists, the side effects of GnRH antagonists, such as hot flushes and bone loss, are related to the hypo-estrogenic status and proportional to the doses and duration of treatment [189,191,195,196].

4.1.4. Other Potential Hormonal Drugs

Other drugs currently under investigation include aromatase inhibitors (AI, e.g., Anastrozole, Letrozole), selective progesterone receptor modulators (SPRM, e.g., Mifepristone, Anoprisnil), or selective estrogen receptor modulators (SERM, e.g., Bezedoxifene, Raloxifene). Considering the high rate of adverse effects, guidelines from the European Society of Human Reproduction and Embryology (ESHRE) recommend that these drugs be used in a scientific setting or combined with other medicines for women refractories to other traditional hormonal treatments [40,156].
Hormonal suppressive therapies modify the endocrine environment in both eutopic endometrium and ectopic lesions, reduce menstrual bleeding that will decrease retrograde flow, and blunt the triggering of inflammatory pathways implicated in menstrual pain [198]. Although the endometriotic lesions histologically resemble the endometrial lining of the uterus, the biochemical differences between eutopic endometrium and endometriotic tissue can be interpreted as the consequence of cell-specific patterns of steroid receptors expression, the immunological microenvironment, questioning the similarity of endometrium to endometriotic lesions [65]. Most subtle superficial endometriosis lesions must have a strong progesterone resistance, considering high progesterone concentrations in peritoneal fluid [199]. The mechanisms of resistance include being implanted basal endometrium, an effect of peritoneal fluid, and epigenetic changes in endometriotic lesions [65]. However, epigenetic therapies are far from ready for clinical application in patients with endometriosis [200]. Peritoneal fluid might be important for understanding the initiation and growth of endometriosis lesions and the lack of synchronicity with the endometrium. Higher estrogen and progesterone concentrations are found in the peritoneal fluid compared with plasma both in women with and without endometriosis [201]. After ovarian suppression, the decreased estrogen concentrations in peritoneal fluid are probably comparable to plasma concentrations which are very low, thus reducing endometriotic lesions [65].

4.2. Analgesics and Neuromodulators

Hormonal therapy for the first-line treatment is typically accompanied by direct analgesia using nonsteroidal anti-inflammatory drugs (NSAIDs), paracetamol (acetaminophen), or various opioids [202]. NSAIDs, widely used in treating chronic inflammation, are effective in relieving pain symptoms. They inhibit the cyclooxygenase enzymes, thus reducing prostaglandin production and inflammation. Common side effects such as gastric ulcers, cardiovascular events, and acute kidney injury should be concerned for long-term use. Increased pain perception in the endometriotic lesions and the nervous system has been demonstrated in women with chronic pain symptoms [203]. The mechanisms that overexpressed nociceptive channels, including P2X3, small diameter sensory neurons, and increased neuropeptides, including calcitonin gene-related peptide, substance P, and neurokinin, play an essential role in generating hyperalgesic responses [204,205,206]. A new investigational drug inhibiting purinergic P2X3 receptor (BAY1817080) has been recruited in clinical trials (Table 1).

4.3. Targeting on Inflammation, Angiogenesis, and Immunomodulators

IL-1 cytokine family (including IL-33 and IL-1β) and their receptors show potential targets for inhibition of inflammation by downregulating the MyD88 pathway, reducing endometriotic lesions [69,207]. MT2990, a monoclonal antibody directly against IL-33, appears to have promising results in treating moderate to severe endometriosis-associated pain (NCT03840993). Another exploratory trial (NCT03991520) also shows the therapeutic potential of Anakinra, an IL-1 receptor antagonist in endometriosis. Anti-angiogenic agents (e.g., DLBS1442, Quinagolide, and Cabergoline) involved in inflammatory or pain pathways also have promising therapeutic effects on endometriosis. However, a previous study using anti-TNF-α therapy (Infliximab) failed to show a welcome impact on women with deep endometriosis [208].

4.4. Other Complementary Therapies

Dietary supplements and natural products, such as Omega-3 polyunsaturated fatty acids, resveratrol (derived from grapes/berries), curcumin (derived from the roots of Curcuma longa), and green tea (rich in catechins/flavonoids) are considered complementary therapies for endometriosis due to their antioxidative, antimitotic, anti-inflammatory, and anti-angiogenic properties [209,210,211,212]. These bioactive compounds reduce the expression of IL-6, IL-8, TNF-α, and COX-2, presenting a reduction of VEGF expression and matrix metalloproteinase-9 activity, thus inhibiting the development of endometriosis [213,214].

5. Surgical Considerations

While medical therapies are not always effective in specific subtypes of endometriosis, surgical excision of all visible lesions is considered the alternative treatment for refractory pain symptoms and intolerable adverse effects of medical therapy [202,215,216]. Surgical approach helps define the severity of endometriosis, restore the pelvic anatomy, and obtain the tissue specimen to rule out suspected malignancy. Surgical reduction of ovarian endometrioma has raised attention to the damage to ovarian reserve and addressed debated issues on managing ovarian endometrioma. DIE can cause irreversible fibrosis and pelvic adhesion, obstructing bowels, ureters, and fallopian tubes [108]. Completely surgical destruction of DIE has higher complication rates, particularly when colorectal and urological resection and/or anastomosis is concomitantly required [217,218,219,220]. Conservative surgery, known as fertility-sparing/preservation surgery (resection of endometriotic lesions without removal of the ovaries and the uterus), is preferred in women with a desire for pregnancy, retaining natural fertility, and augmenting assisted conception [156]. Long-term medical maintenance following conservative surgery reduces endometriosis recurrence, thus avoiding repetitive surgery [221,222,223,224]. In addition, there is increasing awareness of persistent endometriosis-associated pain following surgery [225], exacerbating pain symptoms after repeated surgery, and chronic post-surgical pain (CPSP), which occurs in the postoperative 3 to 6 months [226]. These sequelae imply concomitant adenomyosis, chronic central pain sensitization, or other non-endometriotic-related problems [227]. Accordingly, surgery should be considered only when the benefits overwhelm the drawbacks.

6. Management of Endometriosis-Related Infertility

Women with endometriosis often experience infertility either because of endometriosis itself or due to hormonal suppression [228]. Most of the medical therapies currently used for endometriosis prevent or compromise conception; therefore, reproductive counseling and fertility survey should be established prior to surgery and at diagnosis [40]. The therapeutic decision-making strategies for the complexity of endometriosis are presented in Figure 5.

6.1. Fertility Treatments and Assisted Reproductive Technology (ART)

For women who immediately attempt conception, expectant management is an option but is not recommended due to the low fecundity rate [229,230]. Ovulation induction combined with intrauterine insemination increases the pregnancy rate in women with mild endometriosis [231,232]. In vitro fertilization (IVF) remains the most effective treatment of endometriosis-associated infertility [233,234]. Hormonal suppression by GnRH agonist for 3–6 months or continuous COCs use for 6–8 weeks before IVF has been reported to increase the pregnancy rate, probably due to improving the endometriosis-associated hormonal disturbance [235,236,237,238,239]. Based on evidence from RCTs, a meta-analysis of studies comparing different GnRH agonist protocols (short, long, ultra-long) reported that, a GnRH agonist ultra-long protocol could improve clinical pregnancy rates, especially in patients with stages III/IV endometriosis. However, considering both RCTs and observational studies, the different downregulation protocols showed no significant difference in improving clinical outcomes in patients with endometriosis [240]. In a retrospective study including 151 patients with endometriosis and a previous failed IVF cycle, DNG pre-treatment for 3 months prior to IVF versus no pre-treatment significantly increased cumulative implantation, clinical pregnancy, and live birth rates [241]. This controversial evidence is insufficient to recommend the extended administration of GnRH agonist prior to ART treatment [40]. Current studies showed promising results in a segmented ART protocol that initiates an IVF with all embryos frozen and then administers GnRH agonist or progestin for 3–6 months, followed by deferred thawed embryo transfer [242,243] (Figure 6). Controlled ovarian stimulation may also accelerate endometrioma growth, causing mass effects that diminish oocyte maturation and ovum pickup. However, there are conflicting results regarding the benefits of endometrioma excision before ART [244,245,246].

6.2. Fertility Preservation in Patients with Endometriosis

For women who do not immediately desire pregnancy, endometriosis and ovarian surgery have an increased risk of diminished ovarian reserve [229,230]. Available procedures for fertility preservation include embryo or oocyte freezing and ovarian tissue cryopreservation, which is no longer considered experimental technology [247,248,249]. Generally providing fertility preservation could expose affected women to unnecessary iatrogenic risks, and the effectiveness remains controversial [250,251,252,253]. Women with bilateral endometrioma, repetitive ovarian surgery, or known diminished ovarian reserve are optimal candidates for fertility preservation after reproduction counseling [250,254].

7. Conclusions

Endometriosis was previously considered a gynecologic disease associated with menstrual disorders, pelvic pain symptoms, and infertility. Following the understanding of endometriosis pathogenesis, the systemic and long-term impacts of endometriosis have raised concern. Management of patients in whom endometriosis is found incidentally (without pain or infertility), adolescents and menopausal women with endometriosis depend on better understanding of the disease diversity and associated risks, such as malignant transformation, cardiovascular and metabolic disease. Hormonal regulation, angio-neurogenesis and inflammatory pathway play a crucial role in the complex pathogenesis of endometriosis. The mechanisms involved in endometriosis are not fully discovered, and managing long-term endometriosis-associated complications is still challenging. The current strategy of endometriosis aims to relieve associated symptoms, avoid repetitive surgery, preserve fertility, and potentially reduce the lifelong systemic impacts. Hormonal modification remains the mainstay of endometriosis treatment at initial diagnosis and for long-term maintenance. However, symptomatic women who desire to become pregnant have limited therapeutic options. In light of the limited efficacy and intolerable side effects of many commonly used drugs, future treatment modalities focusing on new molecular targets are most urgently required to enhance lifespan management of the full scope of endometriosis.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24087503/s1.

Author Contributions

L.-H.C., W.-C.L., H.-Y.H. and H.-M.W. all contributed to the concept, design, literature review, draft, and final approval of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Chang Gung Foundation (CMRPG3I0423; CMRPG3L0802; CMRPG3M0871).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zondervan, K.T.; Becker, C.M.; Missmer, S.A. Endometriosis. N. Engl. J. Med. 2020, 382, 1244–1256. [Google Scholar] [CrossRef] [PubMed]
  2. Meuleman, C.; Vandenabeele, B.; Fieuws, S.; Spiessens, C.; Timmerman, D.; D’Hooghe, T. High prevalence of endometriosis in infertile women with normal ovulation and normospermic partners. Fertil. Steril. 2009, 92, 68–74. [Google Scholar] [CrossRef] [PubMed]
  3. Kvaskoff, M.; Mu, F.; Terry, K.L.; Harris, H.R.; Poole, E.M.; Farland, L.; Missmer, S.A. Endometriosis: A high-risk population for major chronic diseases? Hum. Reprod. Update 2015, 21, 500–516. [Google Scholar] [CrossRef]
  4. Horne, A.W.; Saunders, P.T.K.; Abokhrais, I.M.; Hogg, L.; on behalf of theEndometriosis Priority Setting Partnership Steering Group. Top ten endometriosis research priorities in the UK and Ireland. Lancet 2017, 389, 2191–2192. [Google Scholar] [CrossRef]
  5. Rizner, T.L. Noninvasive biomarkers of endometriosis: Myth or reality? Expert Rev. Mol. Diagn. 2014, 14, 365–385. [Google Scholar] [CrossRef] [PubMed]
  6. Saraswat, L.; Ayansina, D.; Cooper, K.G.; Bhattacharya, S.; Horne, A.W.; Bhattacharya, S. Impact of endometriosis on risk of further gynaecological surgery and cancer: A national cohort study. BJOG 2018, 125, 64–72. [Google Scholar] [CrossRef]
  7. Rogers, P.A.; Adamson, G.D.; Al-Jefout, M.; Becker, C.M.; D’Hooghe, T.M.; Dunselman, G.A.; Fazleabas, A.; Giudice, L.C.; Horne, A.W.; Hull, M.L.; et al. Research Priorities for Endometriosis. Reprod. Sci. 2017, 24, 202–226. [Google Scholar] [CrossRef]
  8. Sampson, J.A. Metastatic or Embolic Endometriosis, due to the Menstrual Dissemination of Endometrial Tissue into the Venous Circulation. Am. J. Pathol. 1927, 3, 93–143. [Google Scholar]
  9. Gibson, D.A.; Simitsidellis, I.; Collins, F.; Saunders, P.T.K. Androgens, oestrogens and endometrium: A fine balance between perfection and pathology. J. Endocrinol. 2020, 246, R75–R93. [Google Scholar] [CrossRef]
  10. Wang, W.; Vilella, F.; Alama, P.; Moreno, I.; Mignardi, M.; Isakova, A.; Pan, W.; Simon, C.; Quake, S.R. Single-cell transcriptomic atlas of the human endometrium during the menstrual cycle. Nat. Med. 2020, 26, 1644–1653. [Google Scholar] [CrossRef]
  11. Signorile, P.G.; Baldi, F.; Bussani, R.; D’Armiento, M.; De Falco, M.; Baldi, A. Ectopic endometrium in human foetuses is a common event and sustains the theory of mullerianosis in the pathogenesis of endometriosis, a disease that predisposes to cancer. J. Exp. Clin. Cancer Res. 2009, 28, 49. [Google Scholar] [CrossRef]
  12. Batt, R.E.; Mitwally, M.F. Endometriosis from thelarche to midteens: Pathogenesis and prognosis, prevention and pedagogy. J. Pediatr. Adolesc. Gynecol. 2003, 16, 337–347. [Google Scholar] [CrossRef] [PubMed]
  13. Tandoi, I.; Somigliana, E.; Riparini, J.; Ronzoni, S.; Vigano, P.; Candiani, M. High rate of endometriosis recurrence in young women. J. Pediatr. Adolesc. Gynecol. 2011, 24, 376–379. [Google Scholar] [CrossRef]
  14. Vinatier, D.; Orazi, G.; Cosson, M.; Dufour, P. Theories of endometriosis. Eur. J. Obstet. Gynecol. Reprod. Biol. 2001, 96, 21–34. [Google Scholar] [CrossRef] [PubMed]
  15. Samani, E.N.; Mamillapalli, R.; Li, F.; Mutlu, L.; Hufnagel, D.; Krikun, G.; Taylor, H.S. Micrometastasis of endometriosis to distant organs in a murine model. Oncotarget 2019, 10, 2282–2291. [Google Scholar] [CrossRef] [PubMed]
  16. Taylor, H.S. Endometrial cells derived from donor stem cells in bone marrow transplant recipients. JAMA 2004, 292, 81–85. [Google Scholar] [CrossRef] [PubMed]
  17. Gargett, C.E.; Schwab, K.E.; Deane, J.A. Endometrial stem/progenitor cells: The first 10 years. Hum. Reprod. Update 2016, 22, 137–163. [Google Scholar] [CrossRef]
  18. Leyendecker, G.; Herbertz, M.; Kunz, G.; Mall, G. Endometriosis results from the dislocation of basal endometrium. Hum. Reprod. 2002, 17, 2725–2736. [Google Scholar] [CrossRef]
  19. Figueira, P.G.; Abrao, M.S.; Krikun, G.; Taylor, H.S. Stem cells in endometrium and their role in the pathogenesis of endometriosis. Ann. N. Y. Acad. Sci. 2011, 1221, 10–17. [Google Scholar] [CrossRef]
  20. Du, H.; Taylor, H.S. Contribution of bone marrow-derived stem cells to endometrium and endometriosis. Stem Cells 2007, 25, 2082–2086. [Google Scholar] [CrossRef]
  21. Bergeron, C.; Amant, F.; Ferenczy, A. Pathology and physiopathology of adenomyosis. Best Pract. Res. Clin. Obstet. Gynaecol. 2006, 20, 511–521. [Google Scholar] [CrossRef]
  22. Zhai, J.; Vannuccini, S.; Petraglia, F.; Giudice, L.C. Adenomyosis: Mechanisms and Pathogenesis. Semin. Reprod. Med. 2020, 38, 129–143. [Google Scholar] [CrossRef]
  23. Curtis, K.M.; Hillis, S.D.; Marchbanks, P.A.; Peterson, H.B. Disruption of the endometrial-myometrial border during pregnancy as a risk factor for adenomyosis. Am. J. Obstet. Gynecol. 2002, 187, 543–544. [Google Scholar] [CrossRef] [PubMed]
  24. Ibrahim, M.G.; Chiantera, V.; Frangini, S.; Younes, S.; Kohler, C.; Taube, E.T.; Plendl, J.; Mechsner, S. Ultramicro-trauma in the endometrial-myometrial junctional zone and pale cell migration in adenomyosis. Fertil. Steril. 2015, 104, 1475–1483.e3. [Google Scholar] [CrossRef] [PubMed]
  25. Leyendecker, G.; Wildt, L.; Mall, G. The pathophysiology of endometriosis and adenomyosis: Tissue injury and repair. Arch. Gynecol. Obstet. 2009, 280, 529–538. [Google Scholar] [CrossRef] [PubMed]
  26. Leyendecker, G.; Bilgicyildirim, A.; Inacker, M.; Stalf, T.; Huppert, P.; Mall, G.; Bottcher, B.; Wildt, L. Adenomyosis and endometriosis. Re-visiting their association and further insights into the mechanisms of auto-traumatisation. An MRI study. Arch. Gynecol. Obstet. 2015, 291, 917–932. [Google Scholar] [CrossRef]
  27. Simpson, J.L.; Bischoff, F.Z.; Kamat, A.; Buster, J.E.; Carson, S.A. Genetics of endometriosis. Obstet. Gynecol. Clin. N. Am. 2003, 30, 21–40, vii. [Google Scholar] [CrossRef]
  28. Deiana, D.; Gessa, S.; Anardu, M.; Daniilidis, A.; Nappi, L.; D’Alterio, M.N.; Pontis, A.; Angioni, S. Genetics of endometriosis: A comprehensive review. Gynecol. Endocrinol. 2019, 35, 553–558. [Google Scholar] [CrossRef]
  29. Nyholt, D.R.; Low, S.K.; Anderson, C.A.; Painter, J.N.; Uno, S.; Morris, A.P.; MacGregor, S.; Gordon, S.D.; Henders, A.K.; Martin, N.G.; et al. Genome-wide association meta-analysis identifies new endometriosis risk loci. Nat. Genet. 2012, 44, 1355–1359. [Google Scholar] [CrossRef]
  30. Rahmioglu, N.; Missmer, S.A.; Montgomery, G.W.; Zondervan, K.T. Insights into Assessing the Genetics of Endometriosis. Curr. Obstet. Gynecol. Rep. 2012, 1, 124–137. [Google Scholar] [CrossRef]
  31. Sapkota, Y.; Low, S.K.; Attia, J.; Gordon, S.D.; Henders, A.K.; Holliday, E.G.; MacGregor, S.; Martin, N.G.; McEvoy, M.; Morris, A.P.; et al. Association between endometriosis and the interleukin 1A (IL1A) locus. Hum. Reprod. 2015, 30, 239–248. [Google Scholar] [CrossRef] [PubMed]
  32. Lagana, A.S.; Garzon, S.; Gotte, M.; Vigano, P.; Franchi, M.; Ghezzi, F.; Martin, D.C. The Pathogenesis of Endometriosis: Molecular and Cell Biology Insights. Int. J. Mol. Sci. 2019, 20, 5615. [Google Scholar] [CrossRef] [PubMed]
  33. Vercellini, P.; Vigano, P.; Somigliana, E.; Fedele, L. Endometriosis: Pathogenesis and treatment. Nat. Rev. Endocrinol. 2014, 10, 261–275. [Google Scholar] [CrossRef] [PubMed]
  34. Bulun, S.E.; Yilmaz, B.D.; Sison, C.; Miyazaki, K.; Bernardi, L.; Liu, S.; Kohlmeier, A.; Yin, P.; Milad, M.; Wei, J. Endometriosis. Endocr. Rev. 2019, 40, 1048–1079. [Google Scholar] [CrossRef]
  35. Barban, N.; Jansen, R.; de Vlaming, R.; Vaez, A.; Mandemakers, J.J.; Tropf, F.C.; Shen, X.; Wilson, J.F.; Chasman, D.I.; Nolte, I.M.; et al. Genome-wide analysis identifies 12 loci influencing human reproductive behavior. Nat. Genet. 2016, 48, 1462–1472. [Google Scholar] [CrossRef]
  36. Gallagher, C.S.; Makinen, 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]
  37. Painter, J.N.; O’Mara, T.A.; Morris, A.P.; Cheng, T.H.T.; Gorman, M.; Martin, L.; Hodson, S.; Jones, A.; Martin, N.G.; Gordon, S.; et al. Genetic overlap between endometriosis and endometrial cancer: Evidence from cross-disease genetic correlation and GWAS meta-analyses. Cancer Med. 2018, 7, 1978–1987. [Google Scholar] [CrossRef]
  38. Sapkota, Y.; Steinthorsdottir, V.; Morris, A.P.; Fassbender, A.; Rahmioglu, N.; De Vivo, I.; Buring, J.E.; Zhang, F.; Edwards, T.L.; Jones, S.; et al. Meta-analysis identifies five novel loci associated with endometriosis highlighting key genes involved in hormone metabolism. Nat. Commun. 2017, 8, 15539. [Google Scholar] [CrossRef]
  39. Uimari, O.; Rahmioglu, N.; Nyholt, D.R.; Vincent, K.; Missmer, S.A.; Becker, C.; Morris, A.P.; Montgomery, G.W.; Zondervan, K.T. Genome-wide genetic analyses highlight mitogen-activated protein kinase (MAPK) signaling in the pathogenesis of endometriosis. Hum. Reprod. 2017, 32, 780–793. [Google Scholar] [CrossRef]
  40. Becker, C.M.; Bokor, A.; Heikinheimo, O.; Horne, A.; Jansen, F.; Kiesel, L.; King, K.; Kvaskoff, M.; Nap, A.; Petersen, K.; et al. ESHRE guideline: Endometriosis. Hum. Reprod. Open 2022, 2022, hoac009. [Google Scholar] [CrossRef]
  41. Signorile, P.G.; Severino, A.; Santoro, M.; Spyrou, M.; Viceconte, R.; Baldi, A. Methylation analysis of HOXA10 regulatory elements in patients with endometriosis. BMC Res. Notes 2018, 11, 722. [Google Scholar] [CrossRef]
  42. Wu, Y.; Halverson, G.; Basir, Z.; Strawn, E.; Yan, P.; Guo, S.W. Aberrant methylation at HOXA10 may be responsible for its aberrant expression in the endometrium of patients with endometriosis. Am. J. Obstet. Gynecol. 2005, 193, 371–380. [Google Scholar] [CrossRef]
  43. Wu, Y.; Strawn, E.; Basir, Z.; Halverson, G.; Guo, S.W. Promoter hypermethylation of progesterone receptor isoform B (PR-B) in endometriosis. Epigenetics 2006, 1, 106–111. [Google Scholar] [CrossRef] [PubMed]
  44. Yilmaz, B.D.; Bulun, S.E. Endometriosis and nuclear receptors. Hum. Reprod. Update 2019, 25, 473–485. [Google Scholar] [CrossRef] [PubMed]
  45. Houshdaran, S.; Oke, A.B.; Fung, J.C.; Vo, K.C.; Nezhat, C.; Giudice, L.C. Steroid hormones regulate genome-wide epigenetic programming and gene transcription in human endometrial cells with marked aberrancies in endometriosis. PLoS Genet. 2020, 16, e1008601. [Google Scholar] [CrossRef] [PubMed]
  46. Guo, S.W. Epigenetics of endometriosis. Mol. Hum. Reprod. 2009, 15, 587–607. [Google Scholar] [CrossRef] [PubMed]
  47. Monteiro, J.B.; Colon-Diaz, M.; Garcia, M.; Gutierrez, S.; Colon, M.; Seto, E.; Laboy, J.; Flores, I. Endometriosis is characterized by a distinct pattern of histone 3 and histone 4 lysine modifications. Reprod. Sci. 2014, 21, 305–318. [Google Scholar] [CrossRef]
  48. Xiaomeng, X.; Ming, Z.; Jiezhi, M.; Xiaoling, F. Aberrant histone acetylation and methylation levels in woman with endometriosis. Arch. Gynecol. Obstet. 2013, 287, 487–494. [Google Scholar] [CrossRef]
  49. Fehlmann, T.; Ludwig, N.; Backes, C.; Meese, E.; Keller, A. Distribution of microRNA biomarker candidates in solid tissues and body fluids. RNA Biol. 2016, 13, 1084–1088. [Google Scholar] [CrossRef] [PubMed]
  50. Kuokkanen, S.; Chen, B.; Ojalvo, L.; Benard, L.; Santoro, N.; Pollard, J.W. Genomic profiling of microRNAs and messenger RNAs reveals hormonal regulation in microRNA expression in human endometrium. Biol. Reprod. 2010, 82, 791–801. [Google Scholar] [CrossRef]
  51. Sha, A.G.; Liu, J.L.; Jiang, X.M.; Ren, J.Z.; Ma, C.H.; Lei, W.; Su, R.W.; Yang, Z.M. Genome-wide identification of micro-ribonucleic acids associated with human endometrial receptivity in natural and stimulated cycles by deep sequencing. Fertil. Steril. 2011, 96, 150–155.e5. [Google Scholar] [CrossRef]
  52. Bjorkman, S.; Taylor, H.S. MicroRNAs in endometriosis: Biological function and emerging biomarker candidatesdagger. Biol. Reprod. 2019, 100, 1135–1146. [Google Scholar] [CrossRef] [PubMed]
  53. Wei, S.; Xu, H.; Kuang, Y. Systematic enrichment analysis of microRNA expression profiling studies in endometriosis. Iran. J. Basic Med. Sci. 2015, 18, 423–429. [Google Scholar] [PubMed]
  54. Weber, J.A.; Baxter, D.H.; Zhang, S.; Huang, D.Y.; Huang, K.H.; Lee, M.J.; Galas, D.J.; Wang, K. The microRNA spectrum in 12 body fluids. Clin. Chem. 2010, 56, 1733–1741. [Google Scholar] [CrossRef] [PubMed]
  55. Panir, K.; Schjenken, J.E.; Robertson, S.A.; Hull, M.L. Non-coding RNAs in endometriosis: A narrative review. Hum. Reprod. Update 2018, 24, 497–515. [Google Scholar] [CrossRef] [PubMed]
  56. Yang, Y.M.; Yang, W.X. Epithelial-to-mesenchymal transition in the development of endometriosis. Oncotarget 2017, 8, 41679–41689. [Google Scholar] [CrossRef]
  57. Vigano, P.; Ottolina, J.; Bartiromo, L.; Bonavina, G.; Schimberni, M.; Villanacci, R.; Candiani, M. Cellular Components Contributing to Fibrosis in Endometriosis: A Literature Review. J. Minim. Invasive Gynecol. 2020, 27, 287–295. [Google Scholar] [CrossRef]
  58. Saare, M.; Rekker, K.; Laisk-Podar, T.; Rahmioglu, N.; Zondervan, K.; Salumets, A.; Gotte, M.; Peters, M. Challenges in endometriosis miRNA studies—From tissue heterogeneity to disease specific miRNAs. Biochim. Biophys. Acta Mol. Basis Dis. 2017, 1863, 2282–2292. [Google Scholar] [CrossRef]
  59. Gurates, B.; Bulun, S.E. Endometriosis: The ultimate hormonal disease. Semin. Reprod. Med. 2003, 21, 125–134. [Google Scholar] [CrossRef]
  60. Zondervan, K.T.; Becker, C.M.; Koga, K.; Missmer, S.A.; Taylor, R.N.; Vigano, P. Endometriosis. Nat. Rev. Dis. Primers 2018, 4, 9. [Google Scholar] [CrossRef]
  61. Burney, R.O.; Giudice, L.C. Pathogenesis and pathophysiology of endometriosis. Fertil. Steril. 2012, 98, 511–519. [Google Scholar] [CrossRef] [PubMed]
  62. Patel, B.G.; Rudnicki, M.; Yu, J.; Shu, Y.; Taylor, R.N. Progesterone resistance in endometriosis: Origins, consequences and interventions. Acta Obstet. Gynecol. Scand. 2017, 96, 623–632. [Google Scholar] [CrossRef]
  63. Madjid, T.H.; Jumadi; Judistiani, R.T.D.; Hernowo, B.S.; Faried, A. Detection of endometriosis using immunocytochemistry of P450 Aromatase expressions in eutopic endometrial cells obtained from menstrual sloughing: A diagnostic study. BMC Res. Notes 2020, 13, 233. [Google Scholar] [CrossRef] [PubMed]
  64. Konrad, L.; Dietze, R.; Riaz, M.A.; Scheiner-Bobis, G.; Behnke, J.; Horne, F.; Hoerscher, A.; Reising, C.; Meinhold-Heerlein, I. Epithelial-Mesenchymal Transition in Endometriosis-When Does It Happen? J. Clin. Med. 2020, 9, 1915. [Google Scholar] [CrossRef] [PubMed]
  65. Koninckx, P.R.; Ussia, A.; Adamyan, L.; Gomel, V.; Martin, D.C. Peritoneal fluid progesterone and progesterone resistance in superficial endometriosis lesions. Hum. Reprod. 2022, 37, 203–211. [Google Scholar] [CrossRef]
  66. Taylor, R.N.; Yu, J.; Torres, P.B.; Schickedanz, A.C.; Park, J.K.; Mueller, M.D.; Sidell, N. Mechanistic and therapeutic implications of angiogenesis in endometriosis. Reprod. Sci. 2009, 16, 140–146. [Google Scholar] [CrossRef]
  67. Cheong, Y.C.; Shelton, J.B.; Laird, S.M.; Richmond, M.; Kudesia, G.; Li, T.C.; Ledger, W.L. IL-1, IL-6 and TNF-alpha concentrations in the peritoneal fluid of women with pelvic adhesions. Hum. Reprod. 2002, 17, 69–75. [Google Scholar] [CrossRef]
  68. Forster, R.; Sarginson, A.; Velichkova, A.; Hogg, C.; Dorning, A.; Horne, A.W.; Saunders, P.T.K.; Greaves, E. Macrophage-derived insulin-like growth factor-1 is a key neurotrophic and nerve-sensitizing factor in pain associated with endometriosis. FASEB J. 2019, 33, 11210–11222. [Google Scholar] [CrossRef]
  69. Kato, T.; Yasuda, K.; Matsushita, K.; Ishii, K.J.; Hirota, S.; Yoshimoto, T.; Shibahara, H. Interleukin-1/-33 Signaling Pathways as Therapeutic Targets for Endometriosis. Front. Immunol. 2019, 10, 2021. [Google Scholar] [CrossRef]
  70. Othman Eel, D.; Hornung, D.; Salem, H.T.; Khalifa, E.A.; El-Metwally, T.H.; Al-Hendy, A. Serum cytokines as biomarkers for nonsurgical prediction of endometriosis. Eur. J. Obstet. Gynecol. Reprod. Biol. 2008, 137, 240–246. [Google Scholar] [CrossRef]
  71. Tseng, J.F.; Ryan, I.P.; Milam, T.D.; Murai, J.T.; Schriock, E.D.; Landers, D.V.; Taylor, R.N. Interleukin-6 secretion in vitro is up-regulated in ectopic and eutopic endometrial stromal cells from women with endometriosis. J. Clin. Endocrinol. Metab. 1996, 81, 1118–1122. [Google Scholar] [CrossRef] [PubMed]
  72. Mori, H.; Sawairi, M.; Nakagawa, M.; Itoh, N.; Wada, K.; Tamaya, T. Peritoneal fluid interleukin-1 beta and tumor necrosis factor in patients with benign gynecologic disease. Am. J. Reprod. Immunol. 1991, 26, 62–67. [Google Scholar] [CrossRef] [PubMed]
  73. Akoum, A.; Jolicoeur, C.; Boucher, A. Estradiol amplifies interleukin-1-induced monocyte chemotactic protein-1 expression by ectopic endometrial cells of women with endometriosis. J. Clin. Endocrinol. Metab. 2000, 85, 896–904. [Google Scholar] [CrossRef] [PubMed]
  74. Khorram, O.; Taylor, R.N.; Ryan, I.P.; Schall, T.J.; Landers, D.V. Peritoneal fluid concentrations of the cytokine RANTES correlate with the severity of endometriosis. Am. J. Obstet. Gynecol. 1993, 169, 1545–1549. [Google Scholar] [CrossRef]
  75. Arici, A.; Seli, E.; Zeyneloglu, H.B.; Senturk, L.M.; Oral, E.; Olive, D.L. Interleukin-8 induces proliferation of endometrial stromal cells: A potential autocrine growth factor. J. Clin. Endocrinol. Metab. 1998, 83, 1201–1205. [Google Scholar] [CrossRef]
  76. Arici, A.; Oral, E.; Attar, E.; Tazuke, S.I.; Olive, D.L. Monocyte chemotactic protein-1 concentration in peritoneal fluid of women with endometriosis and its modulation of expression in mesothelial cells. Fertil. Steril. 1997, 67, 1065–1072. [Google Scholar] [CrossRef]
  77. Gonzalez-Ramos, R.; Van Langendonckt, A.; Defrere, S.; Lousse, J.C.; Colette, S.; Devoto, L.; Donnez, J. Involvement of the nuclear factor-kappaB pathway in the pathogenesis of endometriosis. Fertil. Steril. 2010, 94, 1985–1994. [Google Scholar] [CrossRef]
  78. Mu, F.; Rich-Edwards, J.; Rimm, E.B.; Spiegelman, D.; Missmer, S.A. Endometriosis and Risk of Coronary Heart Disease. Circ. Cardiovasc. Qual. Outcomes 2016, 9, 257–264. [Google Scholar] [CrossRef]
  79. Alderman, M.H., 3rd; Yoder, N.; Taylor, H.S. The Systemic Effects of Endometriosis. Semin. Reprod. Med. 2017, 35, 263–270. [Google Scholar] [CrossRef]
  80. Asante, A.; Taylor, R.N. Endometriosis: The role of neuroangiogenesis. Annu. Rev. Physiol. 2011, 73, 163–182. [Google Scholar] [CrossRef]
  81. Laschke, M.W.; Menger, M.D. Basic mechanisms of vascularization in endometriosis and their clinical implications. Hum. Reprod. Update 2018, 24, 207–224. [Google Scholar] [CrossRef] [PubMed]
  82. Bozorgmehr, M.; Gurung, S.; Darzi, S.; Nikoo, S.; Kazemnejad, S.; Zarnani, A.H.; Gargett, C.E. Endometrial and Menstrual Blood Mesenchymal Stem/Stromal Cells: Biological Properties and Clinical Application. Front. Cell Dev. Biol. 2020, 8, 497. [Google Scholar] [CrossRef]
  83. Li, W.N.; Wu, M.H.; Tsai, S.J. HYPOXIA AND REPRODUCTIVE HEALTH: The role of hypoxia in the development and progression of endometriosis. Reproduction 2021, 161, F19–F31. [Google Scholar] [CrossRef] [PubMed]
  84. Ng, S.W.; Norwitz, S.G.; Taylor, H.S.; Norwitz, E.R. Endometriosis: The Role of Iron Overload and Ferroptosis. Reprod. Sci. 2020, 27, 1383–1390. [Google Scholar] [CrossRef] [PubMed]
  85. Yan, D.; Liu, X.; Xu, H.; Guo, S.W. Mesothelial Cells Participate in Endometriosis Fibrogenesis Through Platelet-Induced Mesothelial-Mesenchymal Transition. J. Clin. Endocrinol. Metab. 2020, 105, e4124–e4147. [Google Scholar] [CrossRef]
  86. Riccio, L.; Santulli, P.; Marcellin, L.; Abrao, M.S.; Batteux, F.; Chapron, C. Immunology of endometriosis. Best Pract. Res. Clin. Obstet. Gynaecol. 2018, 50, 39–49. [Google Scholar] [CrossRef]
  87. Guo, M.; Bafligil, C.; Tapmeier, T.; Hubbard, C.; Manek, S.; Shang, C.; Martinez, F.O.; Schmidt, N.; Obendorf, M.; Hess-Stumpp, H.; et al. Mass cytometry analysis reveals a distinct immune environment in peritoneal fluid in endometriosis: A characterisation study. BMC Med. 2020, 18, 3. [Google Scholar] [CrossRef]
  88. Xiao, F.; Liu, X.; Guo, S.W. Platelets and Regulatory T Cells May Induce a Type 2 Immunity That Is Conducive to the Progression and Fibrogenesis of Endometriosis. Front. Immunol. 2020, 11, 610963. [Google Scholar] [CrossRef]
  89. Symons, L.K.; Miller, J.E.; Kay, V.R.; Marks, R.M.; Liblik, K.; Koti, M.; Tayade, C. The Immunopathophysiology of Endometriosis. Trends Mol. Med. 2018, 24, 748–762. [Google Scholar] [CrossRef]
  90. Kralickova, M.; Fiala, L.; Losan, P.; Tomes, P.; Vetvicka, V. Altered Immunity in Endometriosis: What Came First? Immunol. Investig. 2018, 47, 569–582. [Google Scholar] [CrossRef]
  91. Bacci, M.; Capobianco, A.; Monno, A.; Cottone, L.; Di Puppo, F.; Camisa, B.; Mariani, M.; Brignole, C.; Ponzoni, M.; Ferrari, S.; et al. Macrophages are alternatively activated in patients with endometriosis and required for growth and vascularization of lesions in a mouse model of disease. Am. J. Pathol. 2009, 175, 547–556. [Google Scholar] [CrossRef] [PubMed]
  92. Hogg, C.; Panir, K.; Dhami, P.; Rosser, M.; Mack, M.; Soong, D.; Pollard, J.W.; Jenkins, S.J.; Horne, A.W.; Greaves, E. Macrophages inhibit and enhance endometriosis depending on their origin. Proc. Natl. Acad. Sci. USA 2021, 118, e2013776118. [Google Scholar] [CrossRef] [PubMed]
  93. Wang, Y.; Fu, Y.; Xue, S.; Ai, A.; Chen, H.; Lyu, Q.; Kuang, Y. The M2 polarization of macrophage induced by fractalkine in the endometriotic milieu enhances invasiveness of endometrial stromal cells. Int. J. Clin. Exp. Pathol. 2014, 7, 194–203. [Google Scholar] [PubMed]
  94. Miller, J.E.; Ahn, S.H.; Marks, R.M.; Monsanto, S.P.; Fazleabas, A.T.; Koti, M.; Tayade, C. IL-17A Modulates Peritoneal Macrophage Recruitment and M2 Polarization in Endometriosis. Front. Immunol. 2020, 11, 108. [Google Scholar] [CrossRef]
  95. Liang, Y.; Wu, J.; Wang, W.; Xie, H.; Yao, S. Pro-endometriotic niche in endometriosis. Reprod. Biomed. Online 2019, 38, 549–559. [Google Scholar] [CrossRef]
  96. Sciezynska, A.; Komorowski, M.; Soszynska, M.; Malejczyk, J. NK Cells as Potential Targets for Immunotherapy in Endometriosis. J. Clin. Med. 2019, 8, 1468. [Google Scholar] [CrossRef]
  97. Xu, H.; Zhao, J.; Lu, J.; Sun, X. Ovarian endometrioma infiltrating neutrophils orchestrate immunosuppressive microenvironment. J. Ovarian Res. 2020, 13, 44. [Google Scholar] [CrossRef]
  98. Andres, M.P.; Arcoverde, F.V.L.; Souza, C.C.C.; Fernandes, L.F.C.; Abrao, M.S.; Kho, R.M. Extrapelvic Endometriosis: A Systematic Review. J. Minim. Invasive Gynecol. 2020, 27, 373–389. [Google Scholar] [CrossRef]
  99. Guo, S.W.; Mao, X.; Ma, Q.; Liu, X. Dysmenorrhea and its severity are associated with increased uterine contractility and overexpression of oxytocin receptor (OTR) in women with symptomatic adenomyosis. Fertil. Steril. 2013, 99, 231–240. [Google Scholar] [CrossRef]
  100. Nie, J.; Liu, X.; Guo, S.W. Immunoreactivity of oxytocin receptor and transient receptor potential vanilloid type 1 and its correlation with dysmenorrhea in adenomyosis. Am. J. Obstet. Gynecol. 2010, 202, 346.e1–346.e8. [Google Scholar] [CrossRef]
  101. Brainard, A.M.; Korovkina, V.P.; England, S.K. Potassium channels and uterine function. Semin. Cell Dev. Biol. 2007, 18, 332–339. [Google Scholar] [CrossRef] [PubMed]
  102. Yang, Y.; Bin, W.; Aksoy, M.O.; Kelsen, S.G. Regulation of interleukin-1beta and interleukin-1beta inhibitor release by human airway epithelial cells. Eur. Respir. J. 2004, 24, 360–366. [Google Scholar] [CrossRef] [PubMed]
  103. Hayden, M.S.; Ghosh, S. Regulation of NF-kappaB by TNF family cytokines. Semin. Immunol. 2014, 26, 253–266. [Google Scholar] [CrossRef] [PubMed]
  104. Kojima, H.; Aizawa, Y.; Yanai, Y.; Nagaoka, K.; Takeuchi, M.; Ohta, T.; Ikegami, H.; Ikeda, M.; Kurimoto, M. An essential role for NF-kappa B in IL-18-induced IFN-gamma expression in KG-1 cells. J. Immunol. 1999, 162, 5063–5069. [Google Scholar] [CrossRef]
  105. Chen, L.H.; Chan, S.H.; Li, C.J.; Wu, H.M.; Huang, H.Y. Altered Expression of Interleukin-18 System mRNA at the Level of Endometrial Myometrial Interface in Women with Adenomyosis. Curr. Issues Mol. Biol. 2022, 44, 5550–5561. [Google Scholar] [CrossRef]
  106. Nie, J.; Liu, X.; Zheng, Y.; Geng, J.G.; Guo, S.W. Increased immunoreactivity to SLIT/ROBO1 and its correlation with severity of dysmenorrhea in adenomyosis. Fertil. Steril. 2011, 95, 1164–1167. [Google Scholar] [CrossRef]
  107. Canis, M.; Donnez, J.G.; Guzick, D.S.; Halme, J.K.; Rock, J.A.; Schenken, R.S.; Vernon, M.W. Revised American Society for Reproductive Medicine classification of endometriosis: 1996. Fertil. Steril. 1997, 67, 817–821. [Google Scholar] [CrossRef]
  108. Vercellini, P.; Buggio, L.; Somigliana, E. Role of medical therapy in the management of deep rectovaginal endometriosis. Fertil. Steril. 2017, 108, 913–930. [Google Scholar] [CrossRef]
  109. Chung, M.K.; Chung, R.R.; Gordon, D.; Jennings, C. The evil twins of chronic pelvic pain syndrome: Endometriosis and interstitial cystitis. JSLS 2002, 6, 311–314. [Google Scholar]
  110. Redwine, D.B. Diaphragmatic endometriosis: Diagnosis, surgical management, and long-term results of treatment. Fertil. Steril. 2002, 77, 288–296. [Google Scholar] [CrossRef]
  111. Li, T.; Mamillapalli, R.; Ding, S.; Chang, H.; Liu, Z.W.; Gao, X.B.; Taylor, H.S. Endometriosis alters brain electrophysiology, gene expression and increases pain sensitization, anxiety, and depression in female mice. Biol. Reprod. 2018, 99, 349–359. [Google Scholar] [CrossRef] [PubMed]
  112. As-Sanie, S.; Harris, R.E.; Napadow, V.; Kim, J.; Neshewat, G.; Kairys, A.; Williams, D.; Clauw, D.J.; Schmidt-Wilcke, T. Changes in regional gray matter volume in women with chronic pelvic pain: A voxel-based morphometry study. Pain 2012, 153, 1006–1014. [Google Scholar] [CrossRef] [PubMed]
  113. Mabrouk, M.; Del Forno, S.; Spezzano, A.; Raimondo, D.; Arena, A.; Zanello, M.; Leonardi, D.; Paradisi, R.; Seracchioli, R. Painful Love: Superficial Dyspareunia and Three Dimensional Transperineal Ultrasound Evaluation of Pelvic Floor Muscle in Women with Endometriosis. J. Sex Marital Ther. 2020, 46, 187–196. [Google Scholar] [CrossRef] [PubMed]
  114. Wahl, K.J.; Orr, N.L.; Lisonek, M.; Noga, H.; Bedaiwy, M.A.; Williams, C.; Allaire, C.; Albert, A.Y.; Smith, K.B.; Cox, S.; et al. Deep Dyspareunia, Superficial Dyspareunia, and Infertility Concerns Among Women with Endometriosis: A Cross-Sectional Study. Sex. Med. 2020, 8, 274–281. [Google Scholar] [CrossRef]
  115. Schenken, R.S.; Asch, R.H.; Williams, R.F.; Hodgen, G.D. Etiology of infertility in monkeys with endometriosis: Luteinized unruptured follicles, luteal phase defects, pelvic adhesions, and spontaneous abortions. Fertil. Steril. 1984, 41, 122–130. [Google Scholar] [CrossRef]
  116. Sanchez, A.M.; Papaleo, E.; Corti, L.; Santambrogio, P.; Levi, S.; Vigano, P.; Candiani, M.; Panina-Bordignon, P. Iron availability is increased in individual human ovarian follicles in close proximity to an endometrioma compared with distal ones. Hum. Reprod. 2014, 29, 577–583. [Google Scholar] [CrossRef]
  117. Sanchez, A.M.; Vigano, P.; Somigliana, E.; Panina-Bordignon, P.; Vercellini, P.; Candiani, M. The distinguishing cellular and molecular features of the endometriotic ovarian cyst: From pathophysiology to the potential endometrioma-mediated damage to the ovary. Hum. Reprod. Update 2014, 20, 217–230. [Google Scholar] [CrossRef]
  118. Gomes, F.M.; Navarro, P.A.; de Abreu, L.G.; Ferriani, R.A.; dos Reis, R.M.; de Moura, M.D. Effect of peritoneal fluid from patients with minimal/mild endometriosis on progesterone release by human granulosa-lutein cells obtained from infertile patients without endometriosis: A pilot study. Eur. J. Obstet. Gynecol. Reprod. Biol. 2008, 138, 60–65. [Google Scholar] [CrossRef]
  119. Skrzypczak, J. Morphology and steroidogenesis of cultured granulosa cells from endometrioidally changed ovaries. Exp. Clin. Endocrinol. Diabetes 1995, 103, 228–232. [Google Scholar] [CrossRef]
  120. Ferlita, A.; Battaglia, R.; Andronico, F.; Caruso, S.; Cianci, A.; Purrello, M.; Pietro, C.D. Non-Coding RNAs in Endometrial Physiopathology. Int. J. Mol. Sci. 2018, 19, 2120. [Google Scholar] [CrossRef]
  121. Wu, H.M.; Chen, L.H.; Hsu, L.T.; Lai, C.H. Immune Tolerance of Embryo Implantation and Pregnancy: The Role of Human Decidual Stromal Cell- and Embryonic-Derived Extracellular Vesicles. Int. J. Mol. Sci. 2022, 23, 13382. [Google Scholar] [CrossRef]
  122. Barrier, B.F.; Malinowski, M.J.; Dick, E.J., Jr.; Hubbard, G.B.; Bates, G.W. Adenomyosis in the baboon is associated with primary infertility. Fertil. Steril. 2004, 82 (Suppl. S3), 1091–1094. [Google Scholar] [CrossRef]
  123. Liang, T.; Zhang, W.; Pan, N.; Han, B.; Li, R.; Ma, C. Reproductive Outcomes of In Vitro Fertilization and Fresh Embryo Transfer in Infertile Women With Adenomyosis: A Retrospective Cohort Study. Front. Endocrinol. 2022, 13, 865358. [Google Scholar] [CrossRef] [PubMed]
  124. Kobayashi, H.; Kawahara, N.; Ogawa, K.; Yoshimoto, C. A Relationship Between Endometriosis and Obstetric Complications. Reprod. Sci. 2020, 27, 771–778. [Google Scholar] [CrossRef] [PubMed]
  125. Lalani, S.; Choudhry, A.J.; Firth, B.; Bacal, V.; Walker, M.; Wen, S.W.; Singh, S.; Amath, A.; Hodge, M.; Chen, I. Endometriosis and adverse maternal, fetal and neonatal outcomes, a systematic review and meta-analysis. Hum. Reprod. 2018, 33, 1854–1865. [Google Scholar] [CrossRef] [PubMed]
  126. Kobayashi, H.; Kawahara, N.; Ogawa, K.; Yoshimoto, C. Shared Molecular Features Linking Endometriosis and Obstetric Complications. Reprod. Sci. 2020, 27, 1089–1096. [Google Scholar] [CrossRef] [PubMed]
  127. Kobayashi, H. Imprinting genes associated with endometriosis. EXCLI J. 2014, 13, 252–264. [Google Scholar]
  128. Zadora, J.; Singh, M.; Herse, F.; Przybyl, L.; Haase, N.; Golic, M.; Yung, H.W.; Huppertz, B.; Cartwright, J.E.; Whitley, G.; et al. Disturbed Placental Imprinting in Preeclampsia Leads to Altered Expression of DLX5, a Human-Specific Early Trophoblast Marker. Circulation 2017, 136, 1824–1839. [Google Scholar] [CrossRef] [PubMed]
  129. Bellessort, B.; Le Cardinal, M.; Bachelot, A.; Narboux-Neme, N.; Garagnani, P.; Pirazzini, C.; Barbieri, O.; Mastracci, L.; Jonchere, V.; Duvernois-Berthet, E.; et al. Dlx5 and Dlx6 control uterine adenogenesis during post-natal maturation: Possible consequences for endometriosis. Hum. Mol. Genet. 2016, 25, 97–108. [Google Scholar] [CrossRef]
  130. Chen, P.; Wang, D.B.; Liang, Y.M. Evaluation of estrogen in endometriosis patients: Regulation of GATA-3 in endometrial cells and effects on Th2 cytokines. J. Obstet. Gynaecol. Res. 2016, 42, 669–677. [Google Scholar] [CrossRef]
  131. Cordeiro, A.; Neto, A.P.; Carvalho, F.; Ramalho, C.; Doria, S. Relevance of genomic imprinting in intrauterine human growth expression of CDKN1C, H19, IGF2, KCNQ1 and PHLDA2 imprinted genes. J. Assist. Reprod. Genet. 2014, 31, 1361–1368. [Google Scholar] [CrossRef] [PubMed]
  132. Burris, H.H.; Baccarelli, A.A.; Motta, V.; Byun, H.M.; Just, A.C.; Mercado-Garcia, A.; Schwartz, J.; Svensson, K.; Tellez-Rojo, M.M.; Wright, R.O. Association between length of gestation and cervical DNA methylation of PTGER2 and LINE 1-HS. Epigenetics 2014, 9, 1083–1091. [Google Scholar] [CrossRef] [PubMed]
  133. Paquette, A.G.; Lester, B.M.; Koestler, D.C.; Lesseur, C.; Armstrong, D.A.; Marsit, C.J. Placental FKBP5 genetic and epigenetic variation is associated with infant neurobehavioral outcomes in the RICHS cohort. PLoS ONE 2014, 9, e104913. [Google Scholar] [CrossRef] [PubMed]
  134. Brosens, I.A.; De Sutter, P.; Hamerlynck, T.; Imeraj, L.; Yao, Z.; Cloke, B.; Brosens, J.J.; Dhont, M. Endometriosis is associated with a decreased risk of pre-eclampsia. Hum. Reprod. 2007, 22, 1725–1729. [Google Scholar] [CrossRef]
  135. Harada, T.; Taniguchi, F.; Onishi, K.; Kurozawa, Y.; Hayashi, K.; Harada, T.; Japan, E.; Children’s Study, G. Obstetrical Complications in Women with Endometriosis: A Cohort Study in Japan. PLoS ONE 2016, 11, e0168476. [Google Scholar] [CrossRef]
  136. Leone Roberti Maggiore, U.; Ferrero, S.; Mangili, G.; Bergamini, A.; Inversetti, A.; Giorgione, V.; Vigano, P.; Candiani, M. A systematic review on endometriosis during pregnancy: Diagnosis, misdiagnosis, complications and outcomes. Hum. Reprod. Update 2016, 22, 70–103. [Google Scholar] [CrossRef]
  137. Kvaskoff, M.; Mahamat-Saleh, Y.; Farland, L.V.; Shigesi, N.; Terry, K.L.; Harris, H.R.; Roman, H.; Becker, C.M.; As-Sanie, S.; Zondervan, K.T.; et al. Endometriosis and cancer: A systematic review and meta-analysis. Hum. Reprod. Update 2021, 27, 393–420. [Google Scholar] [CrossRef]
  138. Anglesio, M.S.; Papadopoulos, N.; Ayhan, A.; Nazeran, T.M.; Noe, M.; Horlings, H.M.; Lum, A.; Jones, S.; Senz, J.; Seckin, T.; et al. Cancer-Associated Mutations in Endometriosis without Cancer. N. Engl. J. Med. 2017, 376, 1835–1848. [Google Scholar] [CrossRef]
  139. Akahane, T.; Sekizawa, A.; Purwosunu, Y.; Nagatsuka, M.; Okai, T. The role of p53 mutation in the carcinomas arising from endometriosis. Int. J. Gynecol. Pathol. 2007, 26, 345–351. [Google Scholar] [CrossRef]
  140. Amemiya, S.; Sekizawa, A.; Otsuka, J.; Tachikawa, T.; Saito, H.; Okai, T. Malignant transformation of endometriosis and genetic alterations of K-ras and microsatellite instability. Int. J. Gynaecol. Obstet. 2004, 86, 371–376. [Google Scholar] [CrossRef] [PubMed]
  141. Borrelli, G.M.; Abrao, M.S.; Taube, E.T.; Darb-Esfahani, S.; Kohler, C.; Chiantera, V.; Mechsner, S. (Partial) Loss of BAF250a (ARID1A) in rectovaginal deep-infiltrating endometriosis, endometriomas and involved pelvic sentinel lymph nodes. Mol. Hum. Reprod. 2016, 22, 329–337. [Google Scholar] [CrossRef] [PubMed]
  142. Er, T.K.; Su, Y.F.; Wu, C.C.; Chen, C.C.; Wang, J.; Hsieh, T.H.; Herreros-Villanueva, M.; Chen, W.T.; Chen, Y.T.; Liu, T.C.; et al. Targeted next-generation sequencing for molecular diagnosis of endometriosis-associated ovarian cancer. J. Mol. Med. 2016, 94, 835–847. [Google Scholar] [CrossRef]
  143. Siufi Neto, J.; Kho, R.M.; Siufi, D.F.; Baracat, E.C.; Anderson, K.S.; Abrao, M.S. Cellular, histologic, and molecular changes associated with endometriosis and ovarian cancer. J. Minim. Invasive Gynecol. 2014, 21, 55–63. [Google Scholar] [CrossRef]
  144. Soyama, H.; Miyamoto, M.; Takano, M.; Iwahashi, H.; Kato, K.; Sakamoto, T.; Kuwahara, M.; Ishibashi, H.; Matuura, H.; Yoshikawa, T.; et al. A Pathological Study Using 2014 WHO Criteria Reveals Poor Prognosis of Grade 3 Ovarian Endometrioid Carcinomas. In Vivo 2018, 32, 597–602. [Google Scholar] [CrossRef] [PubMed]
  145. Lim, D.; Murali, R.; Murray, M.P.; Veras, E.; Park, K.J.; Soslow, R.A. Morphological and Immunohistochemical Reevaluation of Tumors Initially Diagnosed as Ovarian Endometrioid Carcinoma With Emphasis on High-grade Tumors. Am. J. Surg. Pathol. 2016, 40, 302–312. [Google Scholar] [CrossRef] [PubMed]
  146. Vercellini, P.; Vigano, P.; Buggio, L.; Makieva, S.; Scarfone, G.; Cribiu, F.M.; Parazzini, F.; Somigliana, E. Perimenopausal management of ovarian endometriosis and associated cancer risk: When is medical or surgical treatment indicated? Best Pract. Res. Clin. Obstet. Gynaecol. 2018, 51, 151–168. [Google Scholar] [CrossRef]
  147. Huang, K.J.; Li, Y.X.; Wu, C.J.; Chang, W.C.; Wei, L.H.; Sheu, B.C. Sonographic features differentiating early-stage ovarian clear cell carcinoma from endometrioma with atypical features. J. Ovarian Res. 2022, 15, 84. [Google Scholar] [CrossRef]
  148. Taylor, H.S.; Kotlyar, A.M.; Flores, V.A. Endometriosis is a chronic systemic disease: Clinical challenges and novel innovations. Lancet 2021, 397, 839–852. [Google Scholar] [CrossRef]
  149. Ferrero, S.; Anserini, P.; Remorgida, V.; Ragni, N. Body mass index in endometriosis. Eur. J. Obstet. Gynecol. Reprod. Biol. 2005, 121, 94–98. [Google Scholar] [CrossRef]
  150. Goetz, T.G.; Mamillapalli, R.; Taylor, H.S. Low Body Mass Index in Endometriosis Is Promoted by Hepatic Metabolic Gene Dysregulation in Mice. Biol. Reprod. 2016, 95, 115. [Google Scholar] [CrossRef]
  151. Zolbin, M.M.; Mamillapalli, R.; Nematian, S.E.; Goetz, T.G.; Taylor, H.S. Adipocyte alterations in endometriosis: Reduced numbers of stem cells and microRNA induced alterations in adipocyte metabolic gene expression. Reprod. Biol. Endocrinol. 2019, 17, 36. [Google Scholar] [CrossRef] [PubMed]
  152. Bajaj, P.; Bajaj, P.; Madsen, H.; Arendt-Nielsen, L. Endometriosis is associated with central sensitization: A psychophysical controlled study. J. Pain 2003, 4, 372–380. [Google Scholar] [CrossRef] [PubMed]
  153. Lagana, A.S.; Condemi, I.; Retto, G.; Muscatello, M.R.; Bruno, A.; Zoccali, R.A.; Triolo, O.; Cedro, C. Analysis of psychopathological comorbidity behind the common symptoms and signs of endometriosis. Eur. J. Obstet. Gynecol. Reprod. Biol. 2015, 194, 30–33. [Google Scholar] [CrossRef] [PubMed]
  154. Capezzuoli, T.; Rossi, M.; La Torre, F.; Vannuccini, S.; Petraglia, F. Hormonal drugs for the treatment of endometriosis. Curr. Opin. Pharmacol. 2022, 67, 102311. [Google Scholar] [CrossRef]
  155. Houshdaran, S.; Chen, J.C.; Vallve-Juanico, J.; Balayan, S.; Vo, K.C.; Smith-McCune, K.; Greenblatt, R.M.; Irwin, J.C.; Giudice, L.C. Progestins Related to Progesterone and Testosterone Elicit Divergent Human Endometrial Transcriptomes and Biofunctions. Int. J. Mol. Sci. 2020, 21, 2625. [Google Scholar] [CrossRef] [PubMed]
  156. Dunselman, G.A.; Vermeulen, N.; Becker, C.; Calhaz-Jorge, C.; D’Hooghe, T.; De Bie, B.; Heikinheimo, O.; Horne, A.W.; Kiesel, L.; Nap, A.; et al. ESHRE guideline: Management of women with endometriosis. Hum. Reprod. 2014, 29, 400–412. [Google Scholar] [CrossRef]
  157. Brown, J.; Crawford, T.J.; Datta, S.; Prentice, A. Oral contraceptives for pain associated with endometriosis. Cochrane Database Syst. Rev. 2018, 5, CD001019. [Google Scholar] [CrossRef]
  158. Meresman, G.F.; Auge, L.; Baranao, R.I.; Lombardi, E.; Tesone, M.; Sueldo, C. Oral contraceptives suppress cell proliferation and enhance apoptosis of eutopic endometrial tissue from patients with endometriosis. Fertil. Steril. 2002, 77, 1141–1147. [Google Scholar] [CrossRef]
  159. Caruso, S.; Cianci, A.; Iraci, M.; Fava, V.; Di Pasqua, S.; Cianci, S. Does Nomegestrol Acetate Plus 17beta-Estradiol Oral Contraceptive Improve Endometriosis-Associated Chronic Pelvic Pain in Women? J. Womens Health 2020, 29, 1184–1191. [Google Scholar] [CrossRef]
  160. Caruso, S.; Cianci, A.; Iraci Sareri, M.; Panella, M.; Caruso, G.; Cianci, S. Randomized study on the effectiveness of nomegestrol acetate plus 17beta-estradiol oral contraceptive versus dienogest oral pill in women with suspected endometriosis-associated chronic pelvic pain. BMC Womens Health 2022, 22, 146. [Google Scholar] [CrossRef]
  161. Reis, F.M.; Coutinho, L.M.; Vannuccini, S.; Batteux, F.; Chapron, C.; Petraglia, F. Progesterone receptor ligands for the treatment of endometriosis: The mechanisms behind therapeutic success and failure. Hum. Reprod. Update 2020, 26, 565–585. [Google Scholar] [CrossRef] [PubMed]
  162. Buggio, L.; Somigliana, E.; Barbara, G.; Frattaruolo, M.P.; Vercellini, P. Oral and depot progestin therapy for endometriosis: Towards a personalized medicine. Expert Opin. Pharmacother. 2017, 18, 1569–1581. [Google Scholar] [CrossRef] [PubMed]
  163. Casper, R.F. Progestin-only pills may be a better first-line treatment for endometriosis than combined estrogen-progestin contraceptive pills. Fertil. Steril. 2017, 107, 533–536. [Google Scholar] [CrossRef]
  164. Vercellini, P.; Pietropaolo, G.; De Giorgi, O.; Pasin, R.; Chiodini, A.; Crosignani, P.G. Treatment of symptomatic rectovaginal endometriosis with an estrogen-progestogen combination versus low-dose norethindrone acetate. Fertil. Steril. 2005, 84, 1375–1387. [Google Scholar] [CrossRef] [PubMed]
  165. Harada, T.; Momoeda, M.; Taketani, Y.; Aso, T.; Fukunaga, M.; Hagino, H.; Terakawa, N. Dienogest is as effective as intranasal buserelin acetate for the relief of pain symptoms associated with endometriosis—A randomized, double-blind, multicenter, controlled trial. Fertil. Steril. 2009, 91, 675–681. [Google Scholar] [CrossRef]
  166. Abou-Setta, A.M.; Houston, B.; Al-Inany, H.G.; Farquhar, C. Levonorgestrel-releasing intrauterine device (LNG-IUD) for symptomatic endometriosis following surgery. Cochrane Database Syst. Rev. 2013, 1, CD005072. [Google Scholar] [CrossRef]
  167. Hayashi, A.; Tanabe, A.; Kawabe, S.; Hayashi, M.; Yuguchi, H.; Yamashita, Y.; Okuda, K.; Ohmichi, M. Dienogest increases the progesterone receptor isoform B/A ratio in patients with ovarian endometriosis. J. Ovarian Res. 2012, 5, 31. [Google Scholar] [CrossRef]
  168. Strowitzki, T.; Marr, J.; Gerlinger, C.; Faustmann, T.; Seitz, C. Dienogest is as effective as leuprolide acetate in treating the painful symptoms of endometriosis: A 24-week, randomized, multicentre, open-label trial. Hum. Reprod. 2010, 25, 633–641. [Google Scholar] [CrossRef]
  169. Strowitzki, T.; Marr, J.; Gerlinger, C.; Faustmann, T.; Seitz, C. Detailed analysis of a randomized, multicenter, comparative trial of dienogest versus leuprolide acetate in endometriosis. Int. J. Gynaecol. Obstet. 2012, 117, 228–233. [Google Scholar] [CrossRef]
  170. Strowitzki, T.; Faustmann, T.; Gerlinger, C.; Seitz, C. Dienogest in the treatment of endometriosis-associated pelvic pain: A 12-week, randomized, double-blind, placebo-controlled study. Eur. J. Obstet. Gynecol. Reprod. Biol. 2010, 151, 193–198. [Google Scholar] [CrossRef]
  171. Muzii, L.; Galati, G.; Di Tucci, C.; Di Feliciantonio, M.; Perniola, G.; Di Donato, V.; Benedetti Panici, P.; Vignali, M. Medical treatment of ovarian endometriomas: A prospective evaluation of the effect of dienogest on ovarian reserve, cyst diameter, and associated pain. Gynecol. Endocrinol. 2020, 36, 81–83. [Google Scholar] [CrossRef] [PubMed]
  172. Del Forno, S.; Mabrouk, M.; Arena, A.; Mattioli, G.; Giaquinto, I.; Paradisi, R.; Seracchioli, R. Dienogest or Norethindrone acetate for the treatment of ovarian endometriomas: Can we avoid surgery? Eur. J. Obstet. Gynecol. Reprod. Biol. 2019, 238, 120–124. [Google Scholar] [CrossRef] [PubMed]
  173. Angioni, S.; Pontis, A.; Malune, M.E.; Cela, V.; Luisi, S.; Litta, P.; Vignali, M.; Nappi, L. Is dienogest the best medical treatment for ovarian endometriomas? Results of a multicentric case control study. Gynecol. Endocrinol. 2020, 36, 84–86. [Google Scholar] [CrossRef] [PubMed]
  174. Vercellini, P.; Bracco, B.; Mosconi, P.; Roberto, A.; Alberico, D.; Dhouha, D.; Somigliana, E. Norethindrone acetate or dienogest for the treatment of symptomatic endometriosis: A before and after study. Fertil. Steril. 2016, 105, 734–743.e3. [Google Scholar] [CrossRef]
  175. Morotti, M.; Venturini, P.L.; Biscaldi, E.; Racca, A.; Calanni, L.; Vellone, V.G.; Stabilini, C.; Ferrero, S. Efficacy and acceptability of long-term norethindrone acetate for the treatment of rectovaginal endometriosis. Eur. J. Obstet. Gynecol. Reprod. Biol. 2017, 213, 4–10. [Google Scholar] [CrossRef]
  176. Brown, J.; Kives, S.; Akhtar, M. Progestagens and anti-progestagens for pain associated with endometriosis. Cochrane Database Syst. Rev. 2012, 2012, CD002122. [Google Scholar] [CrossRef]
  177. Committee on Adolescent Health Care Committee. Committee Opinion No. 602: Depot medroxyprogesterone acetate and bone effects. Obstet. Gynecol. 2014, 123, 1398–1402. [Google Scholar] [CrossRef]
  178. Vigano, P.; Somigliana, E.; Vercellini, P. Levonorgestrel-releasing intrauterine system for the treatment of endometriosis: Biological and clinical evidence. Womens Health 2007, 3, 207–214. [Google Scholar] [CrossRef]
  179. Gibbons, T.; Georgiou, E.X.; Cheong, Y.C.; Wise, M.R. Levonorgestrel-releasing intrauterine device (LNG-IUD) for symptomatic endometriosis following surgery. Cochrane Database Syst. Rev. 2021, 12, CD005072. [Google Scholar] [CrossRef]
  180. Chen, Y.J.; Hsu, T.F.; Huang, B.S.; Tsai, H.W.; Chang, Y.H.; Wang, P.H. Postoperative maintenance levonorgestrel-releasing intrauterine system and endometrioma recurrence: A randomized controlled study. Am. J. Obstet. Gynecol. 2017, 216, 582.e1–582.e9. [Google Scholar] [CrossRef]
  181. Romer, T. Long-term treatment of endometriosis with dienogest: Retrospective analysis of efficacy and safety in clinical practice. Arch. Gynecol. Obstet. 2018, 298, 747–753. [Google Scholar] [CrossRef] [PubMed]
  182. Brown, J.; Pan, A.; Hart, R.J. Gonadotrophin-releasing hormone analogues for pain associated with endometriosis. Cochrane Database Syst. Rev. 2010, 2010, CD008475. [Google Scholar] [CrossRef] [PubMed]
  183. Sagsveen, M.; Farmer, J.E.; Prentice, A.; Breeze, A. Gonadotrophin-releasing hormone analogues for endometriosis: Bone mineral density. Cochrane Database Syst. Rev. 2003, 2003, CD001297. [Google Scholar] [CrossRef] [PubMed]
  184. Tosti, C.; Biscione, A.; Morgante, G.; Bifulco, G.; Luisi, S.; Petraglia, F. Hormonal therapy for endometriosis: From molecular research to bedside. Eur. J. Obstet. Gynecol. Reprod. Biol. 2017, 209, 61–66. [Google Scholar] [CrossRef]
  185. Sauerbrun-Cutler, M.T.; Alvero, R. Short- and long-term impact of gonadotropin-releasing hormone analogue treatment on bone loss and fracture. Fertil. Steril. 2019, 112, 799–803. [Google Scholar] [CrossRef]
  186. Bedaiwy, M.A.; Alfaraj, S.; Yong, P.; Casper, R. New developments in the medical treatment of endometriosis. Fertil. Steril. 2017, 107, 555–565. [Google Scholar] [CrossRef]
  187. Donnez, J.; Taylor, R.N.; Taylor, H.S. Partial suppression of estradiol: A new strategy in endometriosis management? Fertil. Steril. 2017, 107, 568–570. [Google Scholar] [CrossRef]
  188. Tukun, F.L.; Olberg, D.E.; Riss, P.J.; Haraldsen, I.; Kaass, A.; Klaveness, J. Recent Development of Non-Peptide GnRH Antagonists. Molecules 2017, 22, 2188. [Google Scholar] [CrossRef]
  189. Taylor, H.S.; Giudice, L.C.; Lessey, B.A.; Abrao, M.S.; Kotarski, J.; Archer, D.F.; Diamond, M.P.; Surrey, E.; Johnson, N.P.; Watts, N.B.; et al. Treatment of Endometriosis-Associated Pain with Elagolix, an Oral GnRH Antagonist. N. Engl. J. Med. 2017, 377, 28–40. [Google Scholar] [CrossRef]
  190. Hornstein, M.D. An Oral GnRH Antagonist for Endometriosis—A New Drug for an Old Disease. N. Engl. J. Med. 2017, 377, 81–83. [Google Scholar] [CrossRef]
  191. Surrey, E.; Taylor, H.S.; Giudice, L.; Lessey, B.A.; Abrao, M.S.; Archer, D.F.; Diamond, M.P.; Johnson, N.P.; Watts, N.B.; Gallagher, J.C.; et al. Long-Term Outcomes of Elagolix in Women With Endometriosis: Results From Two Extension Studies. Obstet. Gynecol. 2018, 132, 147–160. [Google Scholar] [CrossRef] [PubMed]
  192. Taylor, H.S.; Soliman, A.M.; Johns, B.; Pokrzywinski, R.M.; Snabes, M.; Coyne, K.S. Health-Related Quality of Life Improvements in Patients With Endometriosis Treated With Elagolix. Obstet. Gynecol. 2020, 136, 501–509. [Google Scholar] [CrossRef] [PubMed]
  193. Osuga, Y.; Seki, Y.; Tanimoto, M.; Kusumoto, T.; Kudou, K.; Terakawa, N. Relugolix, an oral gonadotropin-releasing hormone receptor antagonist, reduces endometriosis-associated pain in a dose-response manner: A randomized, double-blind, placebo-controlled study. Fertil. Steril. 2021, 115, 397–405. [Google Scholar] [CrossRef] [PubMed]
  194. Harada, T.; Osuga, Y.; Suzuki, Y.; Fujisawa, M.; Fukui, M.; Kitawaki, J. Relugolix, an oral gonadotropin-releasing hormone receptor antagonist, reduces endometriosis-associated pain compared with leuprorelin in Japanese women: A phase 3, randomized, double-blind, noninferiority study. Fertil. Steril. 2022, 117, 583–592. [Google Scholar] [CrossRef]
  195. Giudice, L.C.; As-Sanie, S.; Arjona Ferreira, J.C.; Becker, C.M.; Abrao, M.S.; Lessey, B.A.; Brown, E.; Dynowski, K.; Wilk, K.; Li, Y.; et al. Once daily oral relugolix combination therapy versus placebo in patients with endometriosis-associated pain: Two replicate phase 3, randomised, double-blind, studies (SPIRIT 1 and 2). Lancet 2022, 399, 2267–2279. [Google Scholar] [CrossRef]
  196. Donnez, J.; Taylor, H.S.; Taylor, R.N.; Akin, M.D.; Tatarchuk, T.F.; Wilk, K.; Gotteland, J.P.; Lecomte, V.; Bestel, E. Treatment of endometriosis-associated pain with linzagolix, an oral gonadotropin-releasing hormone-antagonist: A randomized clinical trial. Fertil. Steril. 2020, 114, 44–55. [Google Scholar] [CrossRef]
  197. Taylor, R.N.; Bestel, E.; Gotteland, J.P.; Lecomte, V.; Dubouloz, R.; Terrill, P.; Humberstone, A.; Loumaye, E. Long term treatment of endometriosis associated pain (EAP) with linzagolix: Efficacy and safety after 12 months of treatment. Fertil. Steril. 2019, 112, e323. [Google Scholar] [CrossRef]
  198. Laux-Biehlmann, A.; d’Hooghe, T.; Zollner, T.M. Menstruation pulls the trigger for inflammation and pain in endometriosis. Trends Pharmacol. Sci. 2015, 36, 270–276. [Google Scholar] [CrossRef]
  199. Donnez, J.; Nisolle, M.; Casanas-Roux, F.; Brion, P.; Da Costa Ferreira, N. Stereometric evaluation of peritoneal endometriosis and endometriotic nodules of the rectovaginal septum. Hum. Reprod. 1996, 11, 224–228. [Google Scholar] [CrossRef]
  200. Dolmans, M.M.; Donnez, J. Emerging Drug Targets for Endometriosis. Biomolecules 2022, 12, 1654. [Google Scholar] [CrossRef]
  201. Koninckx, P.R.; Kennedy, S.H.; Barlow, D.H. Endometriotic disease: The role of peritoneal fluid. Hum. Reprod. Update 1998, 4, 741–751. [Google Scholar] [CrossRef] [PubMed]
  202. Kuznetsov, L.; Dworzynski, K.; Davies, M.; Overton, C.; Guideline, C. Diagnosis and management of endometriosis: Summary of NICE guidance. BMJ 2017, 358, j3935. [Google Scholar] [CrossRef] [PubMed]
  203. Arnold, J.; Barcena de Arellano, M.L.; Ruster, C.; Vercellino, G.F.; Chiantera, V.; Schneider, A.; Mechsner, S. Imbalance between sympathetic and sensory innervation in peritoneal endometriosis. Brain Behav. Immun. 2012, 26, 132–141. [Google Scholar] [CrossRef] [PubMed]
  204. Tokushige, N.; Markham, R.; Russell, P.; Fraser, I.S. Nerve fibres in peritoneal endometriosis. Hum. Reprod. 2006, 21, 3001–3007. [Google Scholar] [CrossRef] [PubMed]
  205. Greaves, E.; Grieve, K.; Horne, A.W.; Saunders, P.T. Elevated peritoneal expression and estrogen regulation of nociceptive ion channels in endometriosis. J. Clin. Endocrinol. Metab. 2014, 99, E1738–E1743. [Google Scholar] [CrossRef] [PubMed]
  206. Trapero, C.; Martin-Satue, M. Purinergic Signaling in Endometriosis-Associated Pain. Int. J. Mol. Sci. 2020, 21, 8512. [Google Scholar] [CrossRef]
  207. Chan, B.C.L.; Lam, C.W.K.; Tam, L.S.; Wong, C.K. IL33: Roles in Allergic Inflammation and Therapeutic Perspectives. Front. Immunol. 2019, 10, 364. [Google Scholar] [CrossRef]
  208. Lu, D.; Song, H.; Shi, G. Anti-TNF-alpha treatment for pelvic pain associated with endometriosis. Cochrane Database Syst. Rev. 2013, 3, CD008088. [Google Scholar] [CrossRef]
  209. Abokhrais, I.M.; Denison, F.C.; Whitaker, L.H.R.; Saunders, P.T.K.; Doust, A.; Williams, L.J.; Horne, A.W. A two-arm parallel double-blind randomised controlled pilot trial of the efficacy of Omega-3 polyunsaturated fatty acids for the treatment of women with endometriosis-associated pain (PurFECT1). PLoS ONE 2020, 15, e0227695. [Google Scholar] [CrossRef]
  210. Bruner-Tran, K.L.; Osteen, K.G.; Taylor, H.S.; Sokalska, A.; Haines, K.; Duleba, A.J. Resveratrol inhibits development of experimental endometriosis in vivo and reduces endometrial stromal cell invasiveness in vitro. Biol. Reprod. 2011, 84, 106–112. [Google Scholar] [CrossRef]
  211. Vallee, A.; Lecarpentier, Y. Curcumin and Endometriosis. Int. J. Mol. Sci. 2020, 21, 2440. [Google Scholar] [CrossRef] [PubMed]
  212. Xu, H.; Becker, C.M.; Lui, W.T.; Chu, C.Y.; Davis, T.N.; Kung, A.L.; Birsner, A.E.; D’Amato, R.J.; Wai Man, G.C.; Wang, C.C. Green tea epigallocatechin-3-gallate inhibits angiogenesis and suppresses vascular endothelial growth factor C/vascular endothelial growth factor receptor 2 expression and signaling in experimental endometriosis in vivo. Fertil. Steril. 2011, 96, 1021–1028. [Google Scholar] [CrossRef] [PubMed]
  213. Harlev, A.; Gupta, S.; Agarwal, A. Targeting oxidative stress to treat endometriosis. Expert Opin. Ther. Targets 2015, 19, 1447–1464. [Google Scholar] [CrossRef]
  214. Saunders, P.T.K.; Horne, A.W. Endometriosis: Etiology, pathobiology, and therapeutic prospects. Cell 2021, 184, 2807–2824. [Google Scholar] [CrossRef] [PubMed]
  215. Langebrekke, A.; Johannessen, H.O.; Qvigstad, E. Surgical treatment of endometriosis. Tidsskr. Nor. Laegeforen. 2008, 128, 1515–1518. [Google Scholar]
  216. Zanelotti, A.; Decherney, A.H. Surgery and Endometriosis. Clin. Obstet. Gynecol. 2017, 60, 477–484. [Google Scholar] [CrossRef]
  217. Donnez, J.; Squifflet, J. Complications, pregnancy and recurrence in a prospective series of 500 patients operated on by the shaving technique for deep rectovaginal endometriotic nodules. Hum. Reprod. 2010, 25, 1949–1958. [Google Scholar] [CrossRef]
  218. De Cicco, C.; Corona, R.; Schonman, R.; Mailova, K.; Ussia, A.; Koninckx, P. Bowel resection for deep endometriosis: A systematic review. BJOG 2011, 118, 285–291. [Google Scholar] [CrossRef]
  219. Meuleman, C.; Tomassetti, C.; D’Hoore, A.; Van Cleynenbreugel, B.; Penninckx, F.; Vergote, I.; D’Hooghe, T. Surgical treatment of deeply infiltrating endometriosis with colorectal involvement. Hum. Reprod. Update 2011, 17, 311–326. [Google Scholar] [CrossRef]
  220. Kent, A.; Shakir, F.; Rockall, T.; Haines, P.; Pearson, C.; Rae-Mitchell, W.; Jan, H. Laparoscopic Surgery for Severe Rectovaginal Endometriosis Compromising the Bowel: A Prospective Cohort Study. J. Minim. Invasive Gynecol. 2016, 23, 526–534. [Google Scholar] [CrossRef]
  221. Practice Committee of the American Society for Reproductive Medicine. Treatment of pelvic pain associated with endometriosis. Fertil. Steril. 2008, 90, S260–S269. [Google Scholar] [CrossRef] [PubMed]
  222. Koga, K.; Takamura, M.; Fujii, T.; Osuga, Y. Prevention of the recurrence of symptom and lesions after conservative surgery for endometriosis. Fertil. Steril. 2015, 104, 793–801. [Google Scholar] [CrossRef] [PubMed]
  223. Capezzuoli, T.; Vannuccini, S.; Mautone, D.; Sorbi, F.; Chen, H.; Reis, F.M.; Ceccaroni, M.; Petraglia, F. Long-term hormonal treatment reduces repetitive surgery for endometriosis recurrence. Reprod. Biomed. Online 2021, 42, 451–456. [Google Scholar] [CrossRef] [PubMed]
  224. Ceccaroni, M.; Clarizia, R.; Liverani, S.; Donati, A.; Ceccarello, M.; Manzone, M.; Roviglione, G.; Ferrero, S. Dienogest vs. GnRH agonists as postoperative therapy after laparoscopic eradication of deep infiltrating endometriosis with bowel and parametrial surgery: A randomized controlled trial. Gynecol. Endocrinol. 2021, 37, 930–933. [Google Scholar] [CrossRef]
  225. Guo, S.W. Recurrence of endometriosis and its control. Hum. Reprod. Update 2009, 15, 441–461. [Google Scholar] [CrossRef] [PubMed]
  226. Simanski, C.J.; Althaus, A.; Hoederath, S.; Kreutz, K.W.; Hoederath, P.; Lefering, R.; Pape-Kohler, C.; Neugebauer, E.A. Incidence of chronic postsurgical pain (CPSP) after general surgery. Pain Med. 2014, 15, 1222–1229. [Google Scholar] [CrossRef] [PubMed]
  227. Althaus, A.; Hinrichs-Rocker, A.; Chapman, R.; Arranz Becker, O.; Lefering, R.; Simanski, C.; Weber, F.; Moser, K.H.; Joppich, R.; Trojan, S.; et al. Development of a risk index for the prediction of chronic post-surgical pain. Eur. J. Pain 2012, 16, 901–910. [Google Scholar] [CrossRef]
  228. Hughes, E.; Brown, J.; Collins, J.J.; Farquhar, C.; Fedorkow, D.M.; Vandekerckhove, P. Ovulation suppression for endometriosis. Cochrane Database Syst. Rev. 2007, 2007, CD000155. [Google Scholar] [CrossRef]
  229. Horton, J.; Sterrenburg, M.; Lane, S.; Maheshwari, A.; Li, T.C.; Cheong, Y. Reproductive, obstetric, and perinatal outcomes of women with adenomyosis and endometriosis: A systematic review and meta-analysis. Hum. Reprod. Update 2019, 25, 592–632. [Google Scholar] [CrossRef]
  230. Bonavina, G.; Taylor, H.S. Endometriosis-associated infertility: From pathophysiology to tailored treatment. Front. Endocrinol. 2022, 13, 1020827. [Google Scholar] [CrossRef]
  231. Vesali, S.; Razavi, M.; Rezaeinejad, M.; Maleki-Hajiagha, A.; Maroufizadeh, S.; Sepidarkish, M. Endometriosis fertility index for predicting non-assisted reproductive technology pregnancy after endometriosis surgery: A systematic review and meta-analysis. BJOG 2020, 127, 800–809. [Google Scholar] [CrossRef] [PubMed]
  232. Tomassetti, C.; Bafort, C.; Meuleman, C.; Welkenhuysen, M.; Fieuws, S.; D’Hooghe, T. Reproducibility of the Endometriosis Fertility Index: A prospective inter-/intra-rater agreement study. BJOG 2020, 127, 107–114. [Google Scholar] [CrossRef] [PubMed]
  233. Senapati, S.; Sammel, M.D.; Morse, C.; Barnhart, K.T. Impact of endometriosis on in vitro fertilization outcomes: An evaluation of the Society for Assisted Reproductive Technologies Database. Fertil. Steril. 2016, 106, 164–171.e1. [Google Scholar] [CrossRef] [PubMed]
  234. Practice Committee of the American Society for Reproductive Medicine. Fertility evaluation of infertile women: A committee opinion. Fertil. Steril. 2021, 116, 1255–1265. [Google Scholar] [CrossRef]
  235. Surrey, E.S.; Silverberg, K.M.; Surrey, M.W.; Schoolcraft, W.B. Effect of prolonged gonadotropin-releasing hormone agonist therapy on the outcome of in vitro fertilization-embryo transfer in patients with endometriosis. Fertil. Steril. 2002, 78, 699–704. [Google Scholar] [CrossRef]
  236. Sallam, H.N.; Garcia-Velasco, J.A.; Dias, S.; Arici, A. Long-term pituitary down-regulation before in vitro fertilization (IVF) for women with endometriosis. Cochrane Database Syst. Rev. 2006, 2006, CD004635. [Google Scholar] [CrossRef]
  237. de Ziegler, D.; Gayet, V.; Aubriot, F.X.; Fauque, P.; Streuli, I.; Wolf, J.P.; de Mouzon, J.; Chapron, C. Use of oral contraceptives in women with endometriosis before assisted reproduction treatment improves outcomes. Fertil. Steril. 2010, 94, 2796–2799. [Google Scholar] [CrossRef]
  238. Tamura, H.; Yoshida, H.; Kikuchi, H.; Josaki, M.; Mihara, Y.; Shirafuta, Y.; Shinagawa, M.; Tamura, I.; Taketani, T.; Takasaki, A.; et al. The clinical outcome of Dienogest treatment followed by in vitro fertilization and embryo transfer in infertile women with endometriosis. J. Ovarian Res. 2019, 12, 123. [Google Scholar] [CrossRef]
  239. Khalifa, E.; Mohammad, H.; Abdullah, A.; Abdel-Rasheed, M.; Khairy, M.; Hosni, M. Role of suppression of endometriosis with progestins before IVF-ET: A non-inferiority randomized controlled trial. BMC Pregnancy Childbirth 2021, 21, 264. [Google Scholar] [CrossRef]
  240. Cao, X.; Chang, H.Y.; Xu, J.Y.; Zheng, Y.; Xiang, Y.G.; Xiao, B.; Geng, X.J.; Ni, L.L.; Chu, X.Y.; Tao, S.B.; et al. The effectiveness of different down-regulating protocols on in vitro fertilization-embryo transfer in endometriosis: A meta-analysis. Reprod. Biol. Endocrinol. 2020, 18, 16. [Google Scholar] [CrossRef]
  241. Barra, F.; Lagana, A.S.; Scala, C.; Garzon, S.; Ghezzi, F.; Ferrero, S. Pretreatment with dienogest in women with endometriosis undergoing IVF after a previous failed cycle. Reprod. Biomed. Online 2020, 41, 859–868. [Google Scholar] [CrossRef] [PubMed]
  242. Ozgur, K.; Humaidan, P.; Coetzee, K. Segmented ART—The new era in ART? Reprod. Biol. 2016, 16, 91–103. [Google Scholar] [CrossRef] [PubMed]
  243. Feng, M.; Chen, L.H.; Hsu, L.T.; Wu, H.M. Segmented in vitro fertilization and frozen embryo transfer in dienogest-treated patient with adenomyosis: A case report and lierature review. Taiwan. J. Obstet. Gynecol. 2022, 61, 906–908. [Google Scholar] [CrossRef] [PubMed]
  244. Khine, Y.M.; Taniguchi, F.; Harada, T. Clinical management of endometriosis-associated infertility. Reprod. Med. Biol. 2016, 15, 217–225. [Google Scholar] [CrossRef] [PubMed]
  245. Vercellini, P.; Somigliana, E.; Vigano, P.; Abbiati, A.; Barbara, G.; Crosignani, P.G. Surgery for endometriosis-associated infertility: A pragmatic approach. Hum. Reprod. 2009, 24, 254–269. [Google Scholar] [CrossRef]
  246. Kuroda, K.; Ikemoto, Y.; Ochiai, A.; Ozaki, R.; Matsumura, Y.; Nojiri, S.; Nakagawa, K.; Sugiyama, R. Combination Treatment of Preoperative Embryo Cryopreservation and Endoscopic Surgery (Surgery-ART Hybrid Therapy) in Infertile Women with Diminished Ovarian Reserve and Uterine Myomas or Ovarian Endometriomas. J. Minim. Invasive Gynecol. 2019, 26, 1369–1375. [Google Scholar] [CrossRef]
  247. Practice Committee of the American Society for Reproductive Medicine. Fertility preservation in patients undergoing gonadotoxic therapy or gonadectomy: A committee opinion. Fertil. Steril. 2019, 112, 1022–1033. [Google Scholar] [CrossRef]
  248. Cobo, A.; Giles, J.; Paolelli, S.; Pellicer, A.; Remohi, J.; Garcia-Velasco, J.A. Oocyte vitrification for fertility preservation in women with endometriosis: An observational study. Fertil. Steril. 2020, 113, 836–844. [Google Scholar] [CrossRef]
  249. Kim, S.J.; Kim, S.K.; Lee, J.R.; Suh, C.S.; Kim, S.H. Oocyte cryopreservation for fertility preservation in women with ovarian endometriosis. Reprod. Biomed. Online 2020, 40, 827–834. [Google Scholar] [CrossRef]
  250. Calagna, G.; Della Corte, L.; Giampaolino, P.; Maranto, M.; Perino, A. Endometriosis and strategies of fertility preservation: A systematic review of the literature. Eur. J. Obstet. Gynecol. Reprod. Biol. 2020, 254, 218–225. [Google Scholar] [CrossRef]
  251. Somigliana, E.; Vigano, P.; Filippi, F.; Papaleo, E.; Benaglia, L.; Candiani, M.; Vercellini, P. Fertility preservation in women with endometriosis: For all, for some, for none? Hum. Reprod. 2015, 30, 1280–1286. [Google Scholar] [CrossRef] [PubMed]
  252. Hirsch, M.; Becker, C.; Davies, M. AGAINST: Fertility preservation for women with ovarian endometriosis: It is time to adopt this as routine practice. BJOG 2022, 129, 1937–1938. [Google Scholar] [CrossRef] [PubMed]
  253. Latif, S.; Saridogan, E.; Yasmin, E. FOR: Fertility preservation for women with ovarian endometriosis: It is time to adopt it as routine practice. BJOG 2022, 129, 1935–1936. [Google Scholar] [CrossRef] [PubMed]
  254. Streuli, I.; Benard, J.; Hugon-Rodin, J.; Chapron, C.; Santulli, P.; Pluchino, N. Shedding light on the fertility preservation debate in women with endometriosis: A swot analysis. Eur. J. Obstet. Gynecol. Reprod. Biol. 2018, 229, 172–178. [Google Scholar] [CrossRef]
Figure 1. Theories of endometriosis pathogenesis. The potential origins of endometriotic lesions include traveling of endometrial tissue through retrograde menstruation and in situ by coelomic metaplasia of the peritoneal lining or from Müllerian remnants. Lymphovascular dissemination contributes to extra-pelvic lesions. Bone marrow-derived stem cells (BMDSCs) can directly differentiate into endometrial cells at ectopic locations in the peritoneal cavity and distant sites, further expanding the hypothesis of retrograde menstruation and lymphovascular dissemination. Invagination theory is characterized by altered endometrial basalis cells invading the uterine myometrium through disruption of the endometrial-myometrial interface (EMI), resulting in endometriosis of uterus. Epigenomic and genomic alterations further explain the aberrant gene expression in endometriotic lesions.
Figure 1. Theories of endometriosis pathogenesis. The potential origins of endometriotic lesions include traveling of endometrial tissue through retrograde menstruation and in situ by coelomic metaplasia of the peritoneal lining or from Müllerian remnants. Lymphovascular dissemination contributes to extra-pelvic lesions. Bone marrow-derived stem cells (BMDSCs) can directly differentiate into endometrial cells at ectopic locations in the peritoneal cavity and distant sites, further expanding the hypothesis of retrograde menstruation and lymphovascular dissemination. Invagination theory is characterized by altered endometrial basalis cells invading the uterine myometrium through disruption of the endometrial-myometrial interface (EMI), resulting in endometriosis of uterus. Epigenomic and genomic alterations further explain the aberrant gene expression in endometriotic lesions.
Ijms 24 07503 g001
Figure 2. Pathophysiological processes of endometriosis. Endometriotic lesions are established through interacting molecular mechanisms in a micro-environment of hyperestrogenism and progesterone resistance that promote cell survival and invasion, systemic and localized steroidogenesis, inflammatory response, immune dysregulation, and neuro-angiogenesis. The upward arrows represent overexpression; the downward arrows represent down-regulation.
Figure 2. Pathophysiological processes of endometriosis. Endometriotic lesions are established through interacting molecular mechanisms in a micro-environment of hyperestrogenism and progesterone resistance that promote cell survival and invasion, systemic and localized steroidogenesis, inflammatory response, immune dysregulation, and neuro-angiogenesis. The upward arrows represent overexpression; the downward arrows represent down-regulation.
Ijms 24 07503 g002
Figure 3. Endometriosis across the life course. Summary of the symptoms associated with endometriosis, including pain, infertility, and systemic effects. They have a wide range of spectrum and can overlap with these other conditions. The upward arrows represent an increase, and the downward arrows represent a decrease.
Figure 3. Endometriosis across the life course. Summary of the symptoms associated with endometriosis, including pain, infertility, and systemic effects. They have a wide range of spectrum and can overlap with these other conditions. The upward arrows represent an increase, and the downward arrows represent a decrease.
Ijms 24 07503 g003
Figure 4. Medical treatments for endometriosis [155].
Figure 4. Medical treatments for endometriosis [155].
Ijms 24 07503 g004
Figure 5. Treatment algorithm of endometriosis.
Figure 5. Treatment algorithm of endometriosis.
Ijms 24 07503 g005
Figure 6. Assisted reproductive technology for endometriosis.
Figure 6. Assisted reproductive technology for endometriosis.
Ijms 24 07503 g006
Table 1. Investigational drugs for endometriosis.
Table 1. Investigational drugs for endometriosis.
MechanismTargetsInvestigational DrugsClinical TrialsStatus
Anti-angiogenic drugsSelective dopamine D2-receptor agonistQuinagolide (vaginal rings)
Cabergoline (oral)
NCT03749109
NCT03692403
NCT03928288
Phase II completed
Phase II completed
Phase II recruiting
ImmunomodulatorsSuppression NF-ᴋB and COX-2DLBS1442 (oral)NCT01942122Phase II completed
AntioxidantsDownregulates inflammatory cytokines (e.g., IL-6, IL-1β, MCP-1)Resveratrol (oral)NCT02475564Phase IV completed
Anti-neurogenic and
Anti-inflammatory drugs
Purinergic P2X3 receptors
antagonist
BAY1817080 (oral)NCT04487431
NCT04471337
NCT04454424
NCT04423744
Phase I competed
Monoclonal antibodies of IL-33MT-2990 (intravenous)NCT03840993Phase II completed
An IL-1 receptor antagonistAnakinra (subcutaneous)NCT03991520Phase II recruiting
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

Chen, L.-H.; Lo, W.-C.; Huang, H.-Y.; Wu, H.-M. A Lifelong Impact on Endometriosis: Pathophysiology and Pharmacological Treatment. Int. J. Mol. Sci. 2023, 24, 7503. https://doi.org/10.3390/ijms24087503

AMA Style

Chen L-H, Lo W-C, Huang H-Y, Wu H-M. A Lifelong Impact on Endometriosis: Pathophysiology and Pharmacological Treatment. International Journal of Molecular Sciences. 2023; 24(8):7503. https://doi.org/10.3390/ijms24087503

Chicago/Turabian Style

Chen, Liang-Hsuan, Wei-Che Lo, Hong-Yuan Huang, and Hsien-Ming Wu. 2023. "A Lifelong Impact on Endometriosis: Pathophysiology and Pharmacological Treatment" International Journal of Molecular Sciences 24, no. 8: 7503. https://doi.org/10.3390/ijms24087503

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