Next Article in Journal
Exploring the Molecular Landscape of Myelofibrosis, with a Focus on Ras and Mitogen-Activated Protein (MAP) Kinase Signaling
Previous Article in Journal
Using Single-Case Experimental Design and Patient-Reported Outcome Measures to Evaluate the Treatment of Cancer-Related Cognitive Impairment in Clinical Practice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nuclear Estrogen Receptors in Prostate Cancer: From Genes to Function

1
Department of Life Sciences, University of Modena and Reggio Emilia, 41125 Modena, Italy
2
Unit of Endocrinology, Department of Biomedical, Metabolic and Neural Sciences, University of Modena and Reggio Emilia, Ospedale di Baggiovara, 41126 Modena, Italy
*
Author to whom correspondence should be addressed.
Cancers 2023, 15(18), 4653; https://doi.org/10.3390/cancers15184653
Submission received: 16 August 2023 / Revised: 1 September 2023 / Accepted: 18 September 2023 / Published: 20 September 2023
(This article belongs to the Section Cancer Therapy)

Abstract

:

Simple Summary

Estrogens are steroid hormones that interact with nuclear receptors (ERα and ERβ) and a membrane G-protein-coupled receptor (GPER) to regulate multiple physiological processes. Abnormal ERs and GPER signaling can lead to different disorders, including cancer, making them attractive drug targets. Estrogen-related pathways are implicated not only in breast cancer but also in prostate cancer, providing potential treatment opportunities. New compounds targeting ERs have led to therapeutic advancements, but cancer drug resistance remains a challenge. Genetic and biological mechanisms regulating the expression and activity of nuclear estrogen receptors in prostate cancer are discussed in this review. A comprehensive characterization of specific splice variants in prostate cancer subtypes might lead to new stratification and therapeutic opportunities.

Abstract

Estrogens are almost ubiquitous steroid hormones that are essential for development, metabolism, and reproduction. They exert both genomic and non-genomic action through two nuclear receptors (ERα and ERβ), which are transcription factors with disregulated functions and/or expression in pathological processes. In the 1990s, the discovery of an additional membrane estrogen G-protein-coupled receptor augmented the complexity of this picture. Increasing evidence elucidating the specific molecular mechanisms of action and opposing effects of ERα and Erβ was reported in the context of prostate cancer treatment, where these issues are increasingly investigated. Although new approaches improved the efficacy of clinical therapies thanks to the development of new molecules targeting specifically estrogen receptors and used in combination with immunotherapy, more efforts are needed to overcome the main drawbacks, and resistance events will be a challenge in the coming years. This review summarizes the state-of-the-art on ERα and ERβ mechanisms of action in prostate cancer and promising future therapies.

1. Introduction

Estrogens are steroid hormones primarily produced by the ovaries and placenta in females and by the adrenal cortex and testis in males. They play important modulatory and almost ubiquitary roles in physiological and pathophysiological processes by interacting with their two nuclear receptors (ERα and ERβ) and a membrane G-protein-coupled receptor (GPER, also referred to as GPR30) [1]. Nuclear estrogen receptors (ERs) are ligand-dependent transcription factors that regulate gene transcription through estrogen response elements (EREs), facilitating normal biological functions of the ligands, while GPER modulates mainly estrogen non-genomic effects. Since abnormal ERs and GPER signaling results in multiple disorders, including cancer, these molecules are interesting targets for pharmacological therapies.
In this review, genetic aspects regulating the role of nuclear estrogen receptors in prostate cancer are discussed.

2. Physiology of Estrogens and Their Roles

2.1. Estrogen Biosynthesis and Physiological Functions

Estrogens are molecules known since several decades ago [2]. They are synthesized upon aromatization of androgens, and catalyzed by the cytochrome P450 aromatase (P450Aro) that is encoded by the CYP19A1 gene [3,4]. The aromatase coding gene is located at position 15q21.1 and spans about 75 kilobases. It has multiple transcriptional start sites with individual promoters, leading to an extremely tissue-specific regulation of the transcription modulated by several hormones and other molecules [5,6,7]. Relatively high expression levels of aromatase transcripts are found in ovarian granulosa cells from women of a fertile age and, to a lesser extent, in the testis of men [8]. Moreover, human extragonadal sites of estrogen production were described, such as the placenta, adrenocortical cells, adipocytes and stromal cells, osteoblasts, fibroblasts and keratinocytes, smooth muscle cells, hippocampus, and hypothalamus [8]. Aromatase expression is assumed to be a rate-limiting step for estrogen synthesis, although weak androgen aromatization depending on the steroidogenic acute regulatory protein (StAR) enzyme was detected in adolescent females [9].
Estrogens are commonly known to induce the feminization of girls during adolescence, occurring under the stimuli of gonadotropins and in the absence of effective opposing action from androgens. Moreover, during the fertile age of women, estrogens are known to regulate the ovarian follicle development and growth, acting as proliferative and anti-apoptotic factors supporting oocyte maturation [10,11]. These hormones are synthesized in ovarian granulosa cells in response to follicle-stimulating hormone (FSH) stimulation, which triggers aromatase activity [12,13], and supported by luteinizing hormone (LH)-induced androgen production [14,15]. Effective estrogen production is ensured with the collaborative interaction between ovarian granulosa cells, which express FSH receptors (FSHR) and surround the oocyte, together with theca cells, which have androgenic capability and LH/choriogonadotropin (hCG) receptors (LHCGR) [16]. Finally, the placenta produces estrogens during pregnancy, primarily estriol and estradiol, by aromatizing fetal steroids of an adrenal origin [17,18]. A deficit of estrogens may lead to a range of symptoms [19,20] at different grades that, at the level of the ovary, may be identified as an irregular menstrual cycle or amenorrhea [21,22,23]. Also, a deficit of estrogens is classically associated with an impaired bone metabolism and increased risk of osteoporosis [24], since these hormones regulate bone cell functions and the remodeling of this tissue [25,26]. Interestingly, these effects occur even in men, since estrogen replacement therapy restores normal biochemical measures of bone metabolism in men affected by aromatase deficiency [27].
The adrenal gland is another site for estrogen production in both sexes [28]. Here, these steroids could have a role in the regulation of the renin–angiotensin–aldosterone system, contributing to the physiological maintenance of blood pressure and hydromineral balance [29,30]. Moreover, estrogens may support the development of the adrenal cortex during fetal life, thanks to the interplay between this structure and the placenta [31]. The adrenal contribution to circulating levels of these hormones has been debated for a long time [28], given that estrogens synthesized within extragonadal compartments do not achieve relatively high concentrations and are likely only active locally through paracrine or intracrine action.
Other sites of estrogen synthesis could produce a minor contribution to the overall serum levels of these hormones. Among all, the estrogenic potential of adipose tissue cells and of the brain is well known [32,33,34], since they have aromatase expression [35]. On one side, the adipose tissue works as the main storage site for these sex steroids [36,37]. On the other side, estrogens impact adipose tissue distribution via paracrine action [38], modulate adipocyte functions and metabolism [39], and are responsible for an increased breast cancer risk [40,41,42,43,44,45].
In the brain, estrogens are involved in the regulation of cognitive functions and stimulate spinogenesis and synaptogenesis in regions such as the prefrontal cortex, hippocampus, basal forebrain, thalamus, cerebellum, and brainstem [46,47]. These actions vary according to age and, in women, to the menstrual cycle stage [46]. The role of estrogen in the central nervous system is not limited to sexual dimorphic differences. Rather, these molecules are involved in the processing of sensory information, regulation of affective behavior, synaptic plasticity, learning and memory [48], and neuroprotection against inflammatory damages [49]. During the fetal life, these molecules play a key role in determining sexual differentiation of the brain and subsequent sex differences in behavior, according to the aromatase expression pattern [50,51]. Taken together, these data reveal the central role played by estrogens in regulating brain functions, outlining the need to explore their potential as molecules with neuroprotective properties [52] and in the prevention of brain pathologies such as Alzheimer’s disease [53]. However, the clinical relevance of estrogen action in the brain is still largely unclear and needs further research.

2.2. Role of Estrogens in the Prostate

In men, germ cells, sperm, and Leydig and Sertoli cells display aromatase expression, and the testis is indeed one of the main estrogen-producing glands in males [54]. These cells serve as the major source of estrogens in the male reproductive tract. Therefore, estrogens are also found in the semen [55]. However, evidence demonstrating that estrogens have a major direct role in adult testicular functions is poor [56]. Pioneering data were provided by the study of a very rare case of male estrogen deficiency, which described an adult man with defects in bone and glucose metabolism, but normal pubertal development, masculinization, and sexual function [57,58]. Serum estrogen levels were elevated while testosterone concentration was normal, as an indication of an ER-coding gene disruptive mutation confirmed with a genetic analysis. Other informative results were obtained from men affected by an aromatase deficit caused by CYP19A1 gene loss-of-function mutations who, again, had normal pubertal development and masculinization in addition to metabolic and bone clinical issues [27,59]. Therefore, the role of estrogens in spermatogenesis and testicular functions is still largely unknown, although parameters of sperm quality in infertile men might be sensitive to administration of selective estrogen receptor modulators (SERMs) [60] or estrogenic disruptors [61]. A synergistic effect with androgens, likely exerted through membrane receptors, has been proposed [62], but this matter is still unclear.
Over the years, high relevance has been given to the role of estrogens in males [63]. For instance, these hormones have a role in modulating tissue homeostasis and cell proliferation in the prostate, where they can stimulate as well as inhibit growth, depending on the predominant ERα- or ERβ-mediated effects [64,65]. In general, it is assumed that excessive ERα activation is linked to aberrant proliferation, inflammation, and development of premalignant lesions, while these signals are counterbalanced by anti-proliferative effects exerted through ERβ [64]. Therefore, these steroids contribute to the maintenance of prostate functions, such as the production of the seminal liquid. Several studies agree with the view that estrogens exert their action through both direct receptor-mediated action and indirect effects, by altering the whole endocrine status [66]. Informative results were provided using genetically modified mouse models, such as aromatase-knock-out (KO) or over-expressing mice (ArKO and AROM+), and the ER-KO (αERKO and βERKO). Only prostatic enlargement occurring with age was found in αERKO mice, while βERKO and ArKO displayed prostate hyperplasia. AROM+ even developed squamous metaplasia and had elevated serum estradiol levels, although these results were not univocally confirmed [66]. In men, aromatase is physiologically expressed in the stroma of the prostate, which is capable of local estrogen biosynthesis [67]. However, aberrant expression of the aromatase protein was found in other prostate cells, confirmed in prostate cancer (PCa) cell lines and even in cases of benign prostatic hyperplasia [7,68]. In these tissues, disregulation of aromatase expression would be due to alteration of CYP19A1 promoter usage [67]. Given the clear estrogenic nature of certain prostate tumors, the use of aromatase inhibitors and SERMs as an endocrine therapy for prostate cancer has been proposed since a long time ago [69,70]. The role of nuclear estrogen receptors in PCa is discussed in the next sections.

3. Nuclear Estrogen Receptors

Estrogens regulate several physiological functions mediated by their two nuclear receptors. In humans, the ESR1 gene is located at locus 6q25.1 on chromosome 6 and encodes for the full-length ERα protein, consisting of 595 amino acids (66/67 kDa) [71,72]. The ERβ protein of 530 amino acids (59 kDa) is encoded by the ESR2 gene, which is located on chromosome 14 at locus 14q23.2 [73,74]. These molecules have similar binding affinity for the two main ligands, estradiol and estrone, and act as ligand-activated transcription factors for target genes in the cell nucleus.
As other nuclear receptors, ERs have a modular structure where multiple functional domains have specific functions, namely an N-terminal transactivation domain (NTD/AF1), a DNA-binding domain (DBD), a hinge region, a C-terminal cofactor interaction domain, and a ligand-binding domain (LBD). The NTD/AF1 contains zinc-finger domains that allow the binding and transactivation of target genes. The NTD is one of the most significant structural differences between ERα and ERβ, having approximately 15% sequence homology [75]. The DBD allows ERs to bind estrogen response elements (EREs) in target genes, while the hinge region contains nuclear localization signals and can bind to molecular chaperones. The LBD/AF2 carboxyl terminus encompasses binding sites for co-activators, co-repressors, and the estrogen-binding region, together with a ligand-dependent activation domain (AF2). In particular, natural ligands bind a cavity-shaped domain of the receptor, which may be targeted by several other molecules with agonistic and antagonistic behaviors [76].
New insights into the molecular structure of ERα were provided using the resolution of its structure in a complex with DNA and co-activators using cryo-electron microscopy (EM) [77]. These data provided large advancements compared to what was carried out with the crystallographic structure resolved previously [78]. Cryo-EM depicted the molecular mechanism of action with which ERα binds the ERE DNA as a dimer. Then, the steroid receptor co-activator 3 (SRC-3) protein is independently recruited, via the transactivation domain, by each of the two ligand-bound ERα monomers and together with activation function 1 (AF-1). In turn, the two SRC-3s interact with different regions of one secondary co-activator (p300) protein through multiple contacts [77]. These data are of high relevance to elucidate the mechanisms with which active nuclear receptors upregulate target gene transcription.
ERs carry highly polymorphic regions that may impact the activity of estrogens. This is not surprising, since ER-coding genes are evolutionarily old and unstable, predisposing these sequences to repeated mutations over time [79], and leading to aspecificity for their ligands [80] and capability to bind even plant-derived molecules, such as flavonoids [81]. For instance, the ESR1 gene carries single-nucleotide polymorphisms (SNPs) linked to estrogen resistance [82], breast cancer [83,84], mineral bone density [85], and cardiovascular risk factors [86]. CA nucleotide repeats were found in the genomic region containing the ESR2 gene, suggesting it could be used as a marker of endocrine functions [87] and could be associated with blood pressure [88], Alzheimer’s disease [89], risk of hypospadias [90], and osteoarthritis [91]. Five more receptor variants were identified upon screening of ESR2 5′ and 3′ untranslated regions, although no association with specific phenotypes or pathologies was found [92]. Other ESR2 SNPs that are linked with increased transcriptional activation were related to disorders of sexual and pubertal development [93,94]. In summary, ER-coding genes are highly unstable, leading to high susceptibility to mutations and subsequent modulation or impairment of receptor functionality.
In the body, tissue-specific expression patterns (Figure 1) and the ratio between the two receptors determine the sensitivity of the target tissue to estrogens and their local effects [95]. Although these receptors are expressed almost ubiquitously, ERα is preferentially expressed in the cardiovascular and adipose tissues, thymus, mammary gland, and uterus. In contrast, ERβ is prevalently expressed in the ovaries, prostate, testis, and bone, while similar expression levels could be detected in the brain, hypothalamus, pituitary, and liver [95,96,97]. The tissue distribution of ERs is one of the determinants to understand their differential response to agonists triggered upon interaction with the ligand-binding cavity [76]. In fact, it is well known that ERα and ERβ mediate different, or even opposite, estrogen-induced conformational changes [98] that lead to distinct effects at the cellular level, depending on the site of action and receptor expression levels and sub-type [99].

4. Expression and Splice Isoforms of Estrogen Receptors in Prostate Cancer

4.1. Estrogen Receptor Expression in Prostate Cancer

Prostate tissues express both ERα and ERβ estrogen receptors. ERβ is predominantly expressed in prostate epithelial cells, particularly luminal cells [100], whereas ERα is mostly expressed in stromal cells and basal prostate epithelial cells [74,101].
Until now, it has been hypothesized that ERβ has a primarily protective effect in PCa, whereas ERα could be potentially oncogenic. However, the function of ERs in PCa is still debatable, as contradictory findings regarding their behavior in PCa progression are still emerging.
Using publicly available genomic datasets, the prevalence of ESR1 and ESR2 gene mutations or copy number variations in patients with metastatic or advanced PCa was 4% (5/150) compared to 2% (11/492) in patients with early PCa [102]. Interestingly, the prevalence of ER aberrations in patients with neuroendocrine PCa (NEPC) was considerably higher than in patients with early PCa (17% vs. 2%; p < 0.05). The most prevalent abnormality was an increased copy number. In patients with early PCa, an altered ER copy number or mutation was associated with shorter overall survival (p = 0.01), but there was no significant difference in progression-free survival [102]. In addition, a large population-based case–control investigation revealed an association between an SNP in the promoter region of the ESR2 gene and the risk of developing PCa [103].
In high-grade prostatic intraepithelial neoplasia (hg-PIN), ERα protein and mRNA levels are upregulated, and its expression spreads from basal to luminal cells. Immunohistochemistry and mRNA in situ hybridization experiments indicate that ERα is upregulated in high-grade Gleason score (GS) tumors compared with those with low GS, and in recurrent carcinomas, likely mediating the carcinogenic effects of estradiol. Nevertheless, contrasting results were obtained about ERα protein and mRNA expression in castration-resistant prostate cancers (CRPC) [101,104,105,106]. In addition, inconsistent findings were obtained concerning the transcriptional regulation of ERα. Indeed, the ESR1 gene promoter was found to be extensively methylated in PCa cell lines and tissues, with a positive correlation between methylation levels and tumor pathological grade [107,108]. In contrast, another study found no direct correlation between promoter methylation and mRNA levels of ESR1 in PCa data from The Cancer Genome Atlas (TCGA) [109]. High expression levels of ERα protein and mRNA were associated with shorter progression-free interval and shorter biochemical recurrence-free survival, respectively. Yet, the presence of ERα protein, as determined with IHC, was also associated with the efficacy of endocrine therapy [109,110,111,112,113,114,115] and high ERα expression has been identified in tumor stroma cells, and this correlates with a better prognosis for PCa patients [116,117,118]. These contradictory results make it difficult to draw conclusions regarding ERα expression and relevance in advanced tumors.
In contrast, ERβ is partially lost in hg-PIN and downregulated in approximately 50% of localized and CRPC tumors. Decreased ERβ expression may affect both stromal and epithelial cancer cell development, and PCa tissue staining revealed that ERβ expression is inversely correlated with the progression of PCa to a high Gleason’s grade. High levels of CpG methylation in the proximal promoter region of the ESR2 gene contribute to its downregulation in PCa [119]. It is widely accepted that ERβ may inhibit the development of PCa by acting as a tumor suppressor. However, many findings have questioned this latter concept [120]. To depict a clear picture of the role of ERs in the progression of PCa, it is necessary to consider the experimental pitfalls caused by the available ER antibodies and the significance of various ER splicing variants [121,122,123].

4.2. ER Isoforms in Prostate Cancer

It is well known that specific splice variants of transcription factors modulate the balance between cellular processes that enhance or inhibit tumorigenesis and cancer progression [124]. This was also observed in PCa [125,126,127]. Human ER variants were identified in testis and prostate cells (Figure 2), as well as in breast and other cancer cell lines. In addition to the full-length ERα-66, several isoforms of ERα resulting from alternative gene splicing have been identified, including ERα-46 and ERα-36. A 46-kDa ERα isoform was cloned and described as a receptor splicing variant lacking 173 N-terminal amino acids, including AF-1 [128,129]. It resulted in a ligand-independent transactivation domain, which retains both hormone- and DNA-binding capability, and may be targeted at the cell membrane and work as a dominant-negative inhibitor of ERα activity.
The ERα-36 transcript is initiated by an alternative promoter within the first intron of the ESR1 gene [130,131,132]. ERα-36 differs from ERα-66 in that it lacks both transcriptional activation domains (AF-1 and AF-2) while retaining the DNA-binding domain, partial dimerization domain, and ligand-binding domain. It features a unique 27-amino acid domain that replaces the final 138 amino acids encoded by exons 7 and 8 of the ERα-66 gene. ERα-36 is primarily expressed on the plasma membrane and in the cytoplasm, where it mediates membrane-initiated estrogen signaling effects, such as activation of the mitogen-activated protein kinase/extracellular-signal-regulated kinase (MAPK/ERK) signaling pathway, and stimulates cell growth. ERα-36 has been found in the prostate tissue of both healthy and PCa-affected individuals [102].
Regarding ERβ, five splicing isoforms (ERβ1, -2, -3, -4, and -5) were identified three decades ago [133,134]. Human ERβ splice variants originate from the alternative splicing of exon 8, resulting in sequence variations or protein truncations in the C-terminal regions. The full-length ERβ1 (often referred to as ERβ) and the truncated ERβ2 and ERβ3 have a length ranging between 495 and 530 amino acids and mainly differ from each other in helix 10 of the ligand-binding domain (LBD). ERβ4 and ERβ5 are characterized by a portion of the C-terminal LBD upstream of unique sequences.
These variations lead to the shortening of the LBD and the loss of the AF2 function. Consequently, ERβ1 is the only isoform with ligand-binding ability, whereas truncated ERβ cannot bind estrogens. However, ERβ1 may form heterodimers with ERβ4 or ERβ5 and these molecular complexes have greater transcriptional transactivation activity than the ERβ1 homodimer [135,136]. All these variants are expressed almost ubiquitously, except for ERβ3, which is relatively rare and is mainly restricted to the testis [134], although all of them have a tissue-specific preferential pattern. ERβ2, ERβ3, ERβ4, and ERβ5 have been detected in prostate and PCa samples. While ERβ2 is expressed in both basal and luminal epithelial cells, ERβ5 is almost exclusively found in the basal compartment.
Drug resistance and apoptosis in PCa have been linked to ERβ splice variants [102]. In contrast to ERβ1, which has tumor growth-suppressive effects, ERβ2 and ERβ5 promote tumor growth, inducing stem cell properties and the development of chemoresistance in PCa, whereas ERβ1 might improve cancer cell responsiveness to chemotherapeutic agents [137]. Generally, multiple splice variants are co-expressed in PCa cells; hence, the expression ratio between splice variants is an important indicator of clinical treatment success.

5. Functions of Estrogen Receptors in Prostate Cancer Cells

Estrogen-regulated gene products physiologically control proper autophagy, proliferation, apoptosis, survival, differentiation, and vasodilation. Although the androgen receptor (AR) is a leading player in PCa pathogenesis and exerts its functional effect mainly through transcriptional pathways (recently reviewed in [138,139,140]), ERs have been linked to PCa occurrence, development, and prognoses in several studies (Table 1) and PCa risk is correlated with serum estrogen levels. This suggests that estrogen and ERs may be risk factors. Previous studies showed that ERα is upregulated during malignant transformation of the prostatic epithelium, high-grade and metastatic PCa, and CRPC, where androgen-deprivation therapy increases ERα expression, suggesting its oncogenic role [101]. ERα, expressed in prostate stromal tissues, increases prostatic epithelial proliferation via growth factors like basic fibroblast growth factor (bFGF), epidermal growth factor (EGF), and insulin-like growth factor 1 (IGF-1) [141]. In experimental carcinogenesis with testosterone and estradiol, ERα-knockout mice (KO) did not develop hg-PIN or PCa [142], suggesting that ERα promotes prostate epithelial tumors.
Unlike ERα, ERβ is mostly expressed in epithelial cells; therefore, estrogens directly affect the prostate epithelium and protect it against malignancy via ERβ activity [101]. Hyperplastic alterations decrease ERβ expression in luminal prostatic epithelial cells. ERβ is decreased in high-grade PIN and lost in high-grade PCa and after androgen-deprivation therapy (ADT), suggesting that it interferes with tumors. Grade 4/5 carcinomas have lower ERβ levels than grade 3 carcinomas [100,148]. Exogenous ERβ overexpression in PCa cells showed anti-proliferative, anti-invasive, and pro-apoptotic effects [149]. PCa xenograft studies showed that 17beta-estradiol supplements given to ovariectomized female mice restrict tumor establishment and growth with androgen-independent mechanisms [150], suggesting that estrogen therapy could be beneficial for CRPC. Moreover, hg-PIN occurs in ERβ-knockout (βERKO) mice due to increased proliferation, reduced apoptosis, and accumulation of poorly differentiated cells [151,152]. The above findings suggest that ERβ1 is crucial for prostate health and tumor suppression. Although it is diminished in PCa above Gleason grade 3, ERβ is a possible therapeutic target in the early stages of the disease. As further evidence, ERα activation in βERKO mice triggers abnormal proliferation, inflammation, and premalignant lesions, while ERβ activation in αERKO mice is essential for prostatic stromal–epithelial cell signaling and mediates anti-proliferative effects that counteract the proliferative effects of androgens on the epithelium.
Intensive research on the expression of ER and its variants in the human prostate has revealed that ERβ1 is lost during cancer progression, although its splice variant ERβ2 is expressed in advanced PCa [144]. In PCa PC3 cells stably expressing ERβ1 or ERβ2, ERβ1 decreased proliferation and bone-metastasis-associated factors, whereas ERβ2 boosted proliferation and upregulated them [145]. Consequently, in PCa cells, ERβ2 possesses oncogenic properties that are in opposition to ERβ1 tumor-suppressing effects.

5.1. Transcriptional Activity of ERs

Many, but not all, estrogen-responsive genes include an ERE sequence in their regulatory regions that can be directly bound by ligand-associated ERs as homodimers or heterodimers [153,154,155]. To mediate transcription regulation, ERs undergo conformational changes upon ligand binding and then interact with numerous co-activators and co-repressors of transcription, such as members of the general transcription factor apparatus and chromatin-remodeling proteins [156]. The transcriptional responses can be fine-tuned to meet specific physiological requirements by recruiting certain groups of co-activators or co-repressors. In addition, ER proteins can indirectly regulate genes without ERE-like sequences by interacting with other transcription factors. Activator protein-1 (AP-1), nuclear factor-κB (NF-κB), stimulating protein-1 (Sp-1), activating transcription factor (ATF)-2/c-jun, ATF-2/cAMP response element-binding protein (CREB), ATF-1/CREB, and nuclear transcription factor Y (NF-Y) are examples of intermediate factors that can mediate ER recruitment to promoters or enhancers. About 35% of the annotated human estrogen-responsive genes are regulated with indirect ER-DNA interactions [157].
According to transcriptional studies [158,159], estrogen stimulation both activates and inhibits the transcription of ER target genes. Positive regulators of cell proliferation, such as growth factors and cell cycle regulators, are often upregulated genes. Many of the downregulated genes work in the opposite direction, inhibiting the cell cycle or inducing apoptosis [158]. This gene expression signature is consistent with the concept that estrogen stimulates cell survival. Moreover, recent analyses pointed out that estrogen receptor signaling could be associated with neuroendocrine-like tumors [160].
Estrogen signaling via ERα increases with PCa progression and can induce crucial oncogenic events. While the effects of ERα signaling on breast cancer have been thoroughly studied (recently reviewed in [161,162,163]), the function of this nuclear receptor in prostate pathophysiology is less well understood. Evidence for the presence of a functional ERα-signaling network in PCa includes ERα-mediated regulation of the oncogenic transmembrane protease, serine 2 (TMPRSS2)-Ets-related gene (ERG) fusion protein. ERα is preferentially recruited to intergenic regions of the prostate genome, as determined with a ChIP-seq analysis of global ERα binding in PCa cells [106]. Transcriptome sequencing data and binding profile comparisons revealed that ERα may control the expression of non-coding RNAs. Expression of the long noncoding RNA (lncRNA) nuclear paraspeckle assembly transcript 1 (NEAT1) is regulated with functional ERα signaling, and cancers of the prostate that originate in NEAT1-positive epithelial cells are resistant to androgen inhibitors and androgen deprivation.
Differently from ERα, several studies characterized the role of ERβ as a major tumor regulator in PCa, with a role in controlling cell proliferation and metabolism through both direct and indirect mechanisms. Dey and colleagues [164] found that ERβ regulates apoptosis in PCa cells by upregulating the transcription of forkhead box O3 (FOXO3a), which in turn increases the p53-upregulated modulator of apoptosis (PUMA) in a p53-independent manner and triggers apoptosis via the intrinsic pathway and caspase-9. In advanced CRPC cell lines, the upregulation of ERβ levels can significantly suppress transforming growth factor β 1 (TGF-β1) and IGF-1 expression which, in turn, reduces the expression of downstream anti-apoptotic proteins B-cell lymphoma 2 (Bcl-2) and Survivin, and induces apoptosis [165].
Epithelial hyperplasia and enhanced expression of AR-regulated genes are among the traits observed in the prostates of mice with ERβ inactivation. These genes are also increased in PCa. AR and ERβ are crucial co-regulatory receptor proteins in PCa, and their interactions and crosstalk in signaling have a major influence on PCa pathogenesis. Repression of AR activity by ERβ is associated with tumor suppression in PCa [166]. Therefore, it is essential for PCa therapy to preserve the dynamic equilibrium between AR and ER. ERβ downregulates AR signaling, inducible nitric oxide synthase (NOS), antioxidant glutathione peroxidase 3 (GPX3), and interleukin (IL)-6, thus reducing inflammation and cell proliferation. It has been established that ERβ regulates AR signaling by upregulating the AR co-repressors Dachshund family transcription factor 1 and 2 (DACH1/2) and downregulating the nuclear receptor RORc, which recruits co-activators to the AR promoter and activates AR expression [167]. In line with these observations, it has been demonstrated that ERβ agonists suppress AR expression in androgen-dependent metastatic vertebral cancer of the prostate (VCaP) cells, leading to lower cell survival and increased apoptosis [168].
Moreover, ERβ induces anti-tumoral activity in PCa cells by increasing the expression of anti-proliferative genes like phosphatase and tensin homolog (PTEN), FOXO3a, Kruppel-like factor 5 (KLF5), and cyclin-dependent kinase inhibitor 1A (CDKN1A, p21) and 1B (CDKN1B, p27), as well as decreasing the expression of genes like phosphatidylinositol 3-kinase (PI3K), F-box protein p45 (SKP2), c-MYC, and cyclin E (CCNE), or the oncogenic TMPRSS2-ERG fusion. In PCa cells, ERβ activation can also suppress the effects of ERα and induce cell apoptosis [101,167,169].
In PCa, maintenance of the epithelial phenotype and repression of mesenchymal traits is an important function of ERβ and its ligand 5alpha-androstane-3beta,17beta-diol (3beta-adiol), a dihydrotestosterone metabolite that does not bind AR, but efficiently binds ERβ [170,171]. Epithelial–mesenchymal transition (EMT) inducing stimuli, such as TGF-β and hypoxia, reduce ERβ expression, and ERβ loss is sufficient to establish an EMT. ERβ1 destabilizes hypoxia-inducible factor (HIF)-1 protein through proteasomal degradation and represses HIF-1α-mediated transcription of vascular endothelial growth factor (VEGF)-A [170]. In addition, under hypoxic circumstances, ERβ1 directly represses VEGF-A transcription via the ERE. In contrast, the ERβ variants ERβ2 and ERβ5 can interact with and stabilize the HIF-1α protein and induce hypoxic gene expression under normoxic conditions [172]. Moreover, the expression of metalloproteinase-2 (MMP-2), VEGF, and other important proteins for invasion and migration is increased, together with the reduction in ERβ expression, resulting in higher invasion and migratory capacity of PCa cells. This finding was further confirmed in a mouse model, where the volume of transplanted tumors and their capacity to metastasize were both considerably higher in the ERβ-silenced nude mouse model than in the control group [146]. In the prostate, ERβ expression correlates with E-cadherin levels [173], and ERβ is a negative regulator of inflammation, which are well-recognized factors in carcinogenesis and metastasis. When ERβ is oxidized by reactive oxygen species, DNA binding is lost and production of E-cadherin is decreased, suggesting that ERβ transcriptional activity is vulnerable to oxidation arising from tissue inflammatory processes [174,175].

5.2. Non-Transcriptional ER Signaling

Estrogens control a non-transcriptional signaling pathway by binding to membrane-bound ERα and ERβ, which rapidly regulate ion channel opening or the activation of related enzymes including Ca2+ mobilization, PI3K, and mitogen-activated protein kinase (MAPK). This process does not rely on gene regulation and occurs within seconds to minutes, resulting in a rapid non-genomic effect.
Multiple estrogen receptors, including ERα, ERα-36, and ERβ, have been implicated in the non-genomic responses to estrogens [176]. Extensive evidence indicates that a small reservoir of ERα exists outside of the nucleus, at or near the plasma membrane of tumor cells. A small proportion of ERα is palmitoylated in the absence of hormone stimulation, triggering ERα signaling at the plasma membrane via its interaction with caveolin-1. ERα is depalmitoylated in response to estrogen stimulation and initiates downstream phosphorylation cascades via direct interactions with multiple proteins, including Src tyrosine kinases and PI3K. Other regulatory proteins, such as the adaptor protein Crk-associated substrate (CAS, p130Cas) and the focal adhesion kinase (FAK), have been identified as components of the non-genomic complex. This complex activates the Src/MAPK and PI3K/AKT serine–threonine kinase (AKT) pathways, which regulate cell proliferation and survival. Moreover, membrane/cytoplasm-localized ERs may activate signaling cascades via Galectin-3 and non-phosphorylated β-catenin in androgen-independent prostate cancer cell lines (PC3, DU145), thus modulating proliferation, migration, invasion, and anchorage-independent growth of these cells [177].
During the cell cycle, ERβ1 and ERβ2 isoforms are differentially regulated in LNCaP cells. Unlike ERβ2, which is predominantly expressed in the G2/M phase, ERβ1 induces a cell cycle arrest in the early G1 phase in response to estradiol via a non-genomic pathway involving c-Jun N-terminal kinases (JNK). Specifically, the interaction of ERβ1 with JNK is associated with a decrease in c-Jun phosphorylation and, as a result, an inhibition of the c-Jun/AP-1 complex activity that controls cyclin D1 expression [178]. In addition, in several studies on CRPC cells, it was observed that silencing of ERβ leads to phosphorylation of extracellular signal-regulated kinases 1 and 2 (ERK1/2), activates the ERK1/2 signaling pathway, and promotes cell proliferation, while decreasing the ratio of G0/G1 phase cells and apoptosis [74,179,180].
Rapid activation signaling of ERs in the extranuclear region activates multiple signaling pathways, including MAPK/ERK, making endocrine therapy a promising treatment for PCa by blocking pathways associated with extra-nuclear ERβ. ER-mediated transcriptional and non-transcriptional activities in PCa are summarized in Figure 3.

6. Molecules Targeting Nuclear Estrogen Receptors

Estrogens and their receptors are recognized targets for therapeutic strategies against ER-positive tumors. Several drugs capable of disrupting the molecular machinery for estrogen synthesis were developed, such as the aromatase inhibitors letrozole, anastrozole, and exemestane [181,182]. ERα-specific drugs, such as selective ER modulators (SERMs), selective ER degraders (SERDs), complete ER antagonists (CERANs), selective ER covalent antagonists (SERCAs), and proteolysis-targeting chimera (PROTAC) ER degraders, are used to counteract the effects of estrogens, especially in breast and uterus cancers, where this receptor triggers mainly proliferative signals [162]. Tamoxifen was the first nonsteroidal anti-estrogen [183]. It belongs to the SERM family and was initially developed for contraception purposes, before becoming a largely used anti-cancer agent for ERα-positive tumors of the breast and uterus. In these tissues, tamoxifen displayed mixed pharmacological behavior, acting differentially as an antagonist and an agonist, depending on ER expression levels [184,185]. In early-stage clinical trials, the anti-estrogen fulvestrant and ERα antagonist toremifene showed potential anti-tumor activity in PCa [186]. Fulvestrant acts as a pure ERα antagonist and belongs to the broad class of SERDs [187], enhancing the degradation of the receptor by blocking the interaction between ERα and the co-regulator AF-2, although other molecular mechanisms of action were described [188,189]. This drug has greater efficacy than tamoxifen and higher binding affinity to ERα than estradiol [190]. Interestingly, fulvestrant and other anti-estrogens demonstrated potential anti-tumor activity in PCa also through ERβ-dependent mechanisms [191]. This effect was exacerbated in those cells positive to ERβ, but not to Erα [192], likely via combined downregulation of ESR2 and AR transcripts [193,194]. These data should be interpreted carefully, since fulvestrant lead to contradicting results in clinical studies of men with castration-resistant PCa [194,195]. Moreover, increasing interest is given to GPER as a target for anti-cancer drugs [196,197], since it may drive cell proliferation via molecular mechanisms depending on nuclear ER co-expression [198,199,200].
Regarding the endocrine treatment of PCa, other drugs were developed or are under study, such as anti-androgens [201,202], anti-AR [203,204,205], AR modulators [206,207] and degraders [208,209], glucocorticoid [210] and progesterone [211] receptor modulators, gonadotropin-releasing hormone agonists/antagonists [212], and steroid 17α-monooxygenase enzyme inhibitors [213,214]. This section will describe nuclear ER-directed molecules used for clinical treatment of PCa (Figure 4).
The use of tamoxifen for the treatment of PCa has been suggested since a long time ago [215,216]. A few preclinical studies showed that tamoxifen inhibits the growth of PCa cells and patients may benefit from the use of this drug in combination with immunotherapy, TGFβ, or Wingless and Int-1 (Wnt) antagonists [217]. Moreover, high-dose tamoxifen therapy was well tolerated in a heavily pretreated patient cohort with castratation-resistant PCa, where the drug demonstrated an inhibitory effect on cancer cell proliferation via suppression of phosphatidylinositol-4-phosphate 5-kinase-α/AKT and matrix metalloproteinase (MMP)−9/VEGF signaling pathways [218]. Optimistic results also arose from a phase II clinical trial testing the effect of another SERM, raloxifene, in combination with the anti-androgen bicalutamide [219]. Raloxifene has a similar half-maximal inhibitory concentration compared to tamoxifen (IC50: 2.9–5.7 nM and 3.0 nM, respectively [69]), hinting that the former could be suitable for clinical purposes. Another SERM, toremifene, led to a reduced fracture risk in PCa patients co-treated with denosumab, extending the list of compounds of this class potentially effective in PCa therapy, although it has a relatively high IC50 (1.0 µM) [69]. These data suggest that the clinical use of SERMs may be optimized for personalized PCa therapy, although complete regression does not occur when they are used alone. Despite these apparently positive results, clinical considerations should be carried out cautiously, since most of our knowledge about the use of SERM against PCa is based on in vitro experiments. Similar conclusions may be extended to the clinical use of fulvestrant as a treatment for PCa, taking into consideration that, in addition to ERα, inhibition of ERβ and agonistic activity at GPER were demonstrated as well [69,220,221]. This drug displays higher inhibitory activity on ERα than ERβ, as demonstrated with a lower IC50, which is 0.47 nM for ERα and 3.8 nM for ERβ [69]. Moreover, an in vitro study performed using LNCaP PCa cells suggested that fulvestrant could inhibit AR transcripts, decreasing the response to androgens [193]. Based on these premises, the drug was tested in clinical trials in men affected by androgen-independent PCa, with contrasting results, animating the debate about its real benefits [195,195,217,222,223]. Interestingly, encouraging data came from the combinatorial use of fulvestrant and immunotherapy, although in the context of breast cancer [224,225,226,227]. However, data from the treatment of PCa are still missing. An intriguing molecular mechanism of resistance to fulvestrant may be explained with the agonistic action on GPER [228]. This receptor is overexpressed in a series of tumors, including PCa, modulating proliferative and anti-apoptotic signals [229]. The use of fulvestrant could be linked to disregulation of GPER expression and signaling, providing the basis for tumor resistance.
PROTACs are synthetic heterobifunctional molecules forming complexes between target proteins and an E3 ubiquitin ligase, leading to ubiquitination and degradation of the target via the proteasome [230]. Among the various types of PROTACs available [231], compounds against ERα were recently described as a strategy possibly effective to degrade both wild-type and mutant receptor isoforms in breast cancer [232]. These molecules displayed ERα ubiquitination and degradation within the 0.1–10 µM range, as well as a “hook effect” at higher concentrations [233]. Interesting results were provided with in vitro studies, mainly from breast cancer cell lines, where PROTACs showed anti-proliferative effects and overall low toxicity [234]. These data encourage the use of PROTACs against PCa, given the consistent percentage of castration-resistant cases and positivity to ERα [235]. Another strategy for wild-type and mutant ERα targeting is provided by SERCAs [236,237]. These molecules exert antagonistic activity by binding covalently to an ERα-specific cysteine residue (C530) absent in other steroid hormone receptors. In preclinical studies, the SERCA H3B-6545 had consistent activity against endocrine-therapy-resistant tumors expressing wild-type or mutant ESR1 transcripts [237], stimulating its testing into an ongoing clinical trial on breast cancer patients (NCT03250676). Moreover, in ERα-positive breast cell lines, SERCAs demonstrated an increased anti-proliferative effect via synergistic activity with cyclin-dependent kinases 4 and 6 (CDK4/6) or the mechanistic target of rapamycin (mTOR) inhibitors [236]. SERCAs are molecules developed recently. However, most of the endocrine-resistant tumors display ERα-dependent growth, suggesting that this novel class of antagonists could be interesting tools to be optimized for PCa clinical treatment [238]. Finally, a class of blockers of both ligand-independent transcriptional activation function 1 and 2 (AF1 and AF2) domains of nuclear receptors are CERANs. Studies on molecules with AF1 and AF2 inhibitory activity have been performed since several decades ago [239]. More recently, the CERAN OP-1250 has been available for oral administration. OP-1250 was tested in ESR1-positive preclinical models, where it induced complete ER degradation and inactivation, blocking gene transcription and cell growth, and demonstrated the capability to shrink brain metastases (reviewed in [240]). OP-1250 is under testing in a clinical trial for advanced and/or metastatic hormone-receptor-positive, HER2-negative breast cancer patients (NCT04505826). Finally, dual-mechanism ER inhibitors (DMERIs) were recently developed [241]. These molecules feature two distinct ER-targeting moieties, inducing noncanonical perturbations of the receptor structure. As a result, DMERIs alter the ER ligand-binding domain and stabilize multiple, dimeric antagonist substrates, effectively antagonizing the proliferation of ER-positive breast cancer cells. This tool provides a new, exciting strategy to potentially counteract breast cancer and PCa in the future.
New frontiers for the treatment of PCa consist in the recent development of ERβ agonists [242,243,244,245]. Oppositely to ERα, this receptor exerts anti-proliferative effects in the prostate, suggesting that ERβ agonists may be useful tools to counteract PCa, avoiding side effects of SERM [246]. The recently developed compound 8β-VE2 demonstrated efficacy in prolonging the life of breast cancer patients, when used in combination with tamoxifen [247]. The anti-proliferative role of this agonist was also tested in xenografted benign prostate hyperplasia and PCa tissues from men, where 8β-VE2 induced apoptosis of stromal and epithelial cells, leading to potential clinical opportunities [169]. Another ERβ agonist is erteberel, which has been tested in preclinical models of glioblastoma, revealing chemotactic potential exerted through mechanisms involving IL-1β, useful to promote innate immunity against cancer [248]. These results suggest that erteberel could be an interesting compound to be tested against PCa, but no data are still available in this regard. Finally, a recent study described the use of androgen synthesis blockers in combination with ERβ agonist soy isoflavones, in men affected by PCa [249]. The combination of these drugs was suggested as potentially beneficial to sustain ERβ-dependent signals, via a mechanism counteracting the EGFR migration to the nucleus, as a stimulus supporting cell proliferation. In this case, further studies are needed to understand how phytoestrogens targeting ERβ might prevent the development of tyrosine-kinase-driven cancer.

7. Conclusions

Growing evidence suggests that estrogen-related pathways are implicated not only in the development of breast cancer but also in PCa. Therefore, in addition to androgen-signaling pathways, transcriptional and non-transcriptional estrogen signaling may represent an opportunity for the treatment of advanced PCa. New compounds targeting specifically ERs led to advancements in the efficacy of therapies, although there is still a long way to go and resistance in cancer therapeutics is a problem to be faced in the next few years. Immunotherapies seem to lead to intriguing perspectives, although most PCa patients fail to respond to the clinical treatment. These approaches could find benefits from the combinatorial use of molecules targeting ERs, which may improve the efficacy of immunotherapies. Moreover, a deep characterization of the expression and role of specific splice variants in PCa subtypes may represent a new opportunity for the stratification of PCa patients and the development of splice-switching oligonucleotide (SSO) therapies.

Author Contributions

Conceptualization and writing—original draft preparation, S.B., C.I. and L.C. All authors have read and agreed to the published version of the manuscript.

Funding

The laboratory of C.I. was supported by AIRC (Fondazione Italiana per la Ricerca sul Cancro) under IG 2018—I.D. 21323 project—P.I. Carol Imbriano.

Acknowledgments

L.C. was supported by Le Studium Loire Valley Institute for Advanced Studies, Orléans and Tours, France through the ARD CVL Biopharmaceuticals program, funded by the Centre-Val de Loire region.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rochira, V.; Fabbi, M.; Valassi, E.; Madeo, B.; Carani, C. Estrogens, Male Reproduction and Beyond. Andrologie 2000, 13, 51–61. [Google Scholar] [CrossRef]
  2. Selye, H. The General Adaptation Syndrome and the Diseases of Adaptation. J. Clin. Endocrinol. Metab. 1946, 6, 117–230. [Google Scholar] [CrossRef] [PubMed]
  3. Simpson, E.R.; Mahendroo, M.S.; Means, G.D.; Kilgore, M.W.; Hinshelwood, M.M.; Graham-Lorence, S.; Amarneh, B.; Ito, Y.; Fisher, C.R.; Michael, M.D.; et al. Aromatase Cytochrome P450, the Enzyme Responsible for Estrogen Biosynthesis. Endocr. Rev. 1994, 15, 342–355. [Google Scholar] [CrossRef] [PubMed]
  4. Miller, W.L.; Auchus, R.J. The Molecular Biology, Biochemistry, and Physiology of Human Steroidogenesis and Its Disorders. Endocr. Rev. 2011, 32, 81–151. [Google Scholar] [CrossRef] [PubMed]
  5. Mahendroo, M.S.; Means, G.D.; Mendelson, C.R.; Simpson, E.R. Tissue-Specific Expression of Human P-450AROM. The Promoter Responsible for Expression in Adipose Tissue Is Different from That Utilized in Placenta. J. Biol. Chem. 1991, 266, 11276–11281. [Google Scholar] [CrossRef] [PubMed]
  6. Mahendroo, M.S.; Mendelson, C.R.; Simpson, E.R. Tissue-Specific and Hormonally Controlled Alternative Promoters Regulate Aromatase Cytochrome P450 Gene Expression in Human Adipose Tissue. J. Biol. Chem. 1993, 268, 19463–19470. [Google Scholar] [CrossRef]
  7. Stocco, C. Tissue Physiology and Pathology of Aromatase. Steroids 2012, 77, 27–35. [Google Scholar] [CrossRef] [PubMed]
  8. Barakat, R.; Oakley, O.; Kim, H.; Jin, J.; Ko, C.M.J. Extra-Gonadal Sites of Estrogen Biosynthesis and Function. BMB Rep. 2016, 49, 488–496. [Google Scholar] [CrossRef]
  9. Bose, H.S.; Pescovitz, O.H.; Miller, W.L. Spontaneous Feminization in a 46,XX Female Patient with Congenital Lipoid Adrenal Hyperplasia Due to a Homozygous Frameshift Mutation in the Steroidogenic Acute Regulatory Protein. J. Clin. Endocrinol. Metab. 1997, 82, 1511–1515. [Google Scholar] [CrossRef]
  10. Casarini, L.; Lazzaretti, C.; Paradiso, E.; Limoncella, S.; Riccetti, L.; Sperduti, S.; Melli, B.; Marcozzi, S.; Anzivino, C.; Sayers, N.S.; et al. Membrane Estrogen Receptor (GPER) and Follicle-Stimulating Hormone Receptor (FSHR) Heteromeric Complexes Promote Human Ovarian Follicle Survival. iScience 2020, 23, 101812. [Google Scholar] [CrossRef]
  11. Casarini, L.; Santi, D.; Brigante, G.; Simoni, M. Two Hormones for One Receptor: Evolution, Biochemistry, Actions, and Pathophysiology of LH and HCG. Endocr. Rev. 2018, 39, 549–592. [Google Scholar] [CrossRef] [PubMed]
  12. McDonald, C.A.; Millena, A.C.; Reddy, S.; Finlay, S.; Vizcarra, J.; Khan, S.A.; Davis, J.S. Follicle-Stimulating Hormone-Induced Aromatase in Immature Rat Sertoli Cells Requires an Active Phosphatidylinositol 3-Kinase Pathway and Is Inhibited via the Mitogen-Activated Protein Kinase Signaling Pathway. Mol. Endocrinol. 2006, 20, 608–618. [Google Scholar] [CrossRef] [PubMed]
  13. Erickson, G.F.; Hsueh, A.J.W. Stimulation of Aromatase Activity by Follicle Stimulating Hormone in Rat Granulosa Cells In Vivo and In Vitro. Endocrinology 1978, 102, 1275–1282. [Google Scholar] [CrossRef]
  14. Liu, Y.X.; Hsueh, A.J.W. Synergism between Granulosa and Theca-Interstitial Cells in Estrogen Biosynthesis by Gonadotropin-Treated Rat Ovaries: Studies on the Two-Cell, Two-Gonadotropin Hypothesis Using Steroid Antisera. Biol. Reprod. 1986, 35, 27–36. [Google Scholar] [CrossRef]
  15. Wada, Y.; Tsuiki, A.; Fukaya, T.; Shinkawa, O.; Satoh, S.; Horiguchi, M.; Hoshiai, H.; Yajima, A. Effects of Androgen on 17 Beta-Estradiol Production by Cultured Human Granulosa Cells. Tohoku J. Exp. Med. 1988, 154, 253–260. [Google Scholar] [CrossRef]
  16. Millier, S.G.; Whitelaw, P.F.; Smyth, C.D. Follicular Oestrogen Synthesis: The “two-Cell, Two-Gonadotrophin” Model Revisited. Mol. Cell Endocrinol. 1994, 100, 51–54. [Google Scholar] [CrossRef]
  17. Miller, W.L. Steroid Hormone Biosynthesis and Actions in the Materno-Feto-Placental Unit. Clin. Perinatol. 1998, 25, 799–817. [Google Scholar] [CrossRef]
  18. Ishimoto, H.; Jaffe, R.B. Development and Function of the Human Fetal Adrenal Cortex: A Key Component in the Feto-Placental Unit. Endocr. Rev. 2011, 32, 317–355. [Google Scholar] [CrossRef] [PubMed]
  19. Shufelt, C.L.; Torbati, T.; Dutra, E. Hypothalamic Amenorrhea and the Long-Term Health Consequences. Semin. Reprod. Med. 2017, 35, 256–262. [Google Scholar] [CrossRef]
  20. Chen, M.; Jiang, H.; Zhang, C. Selected Genetic Factors Associated with Primary Ovarian Insufficiency. Int. J. Mol. Sci. 2023, 24, 4423. [Google Scholar] [CrossRef]
  21. Stuenkel, C.A.; Gompel, A.; Davis, S.R.; Pinkerton, J.A.V.; Lumsden, M.A.; Santen, R.J. Approach to the Patient With New-Onset Secondary Amenorrhea: Is This Primary Ovarian Insufficiency? J. Clin. Endocrinol. Metab. 2022, 107, 825–835. [Google Scholar] [CrossRef]
  22. Kaplan, J.R.; Manuck, S.B. Ovarian Dysfunction, Stress, and Disease: A Primate Continuum. ILAR J. 2004, 45, 89–115. [Google Scholar] [CrossRef] [PubMed]
  23. Golden, N.H.; Carlson, J.L. The Pathophysiology of Amenorrhea in the Adolescent. Ann. N. Y. Acad. Sci. 2008, 1135, 163–178. [Google Scholar] [CrossRef]
  24. Riggs, B.L.; Khosla, S.; Melton, L.J. A Unitary Model for Involutional Osteoporosis: Estrogen Deficiency Causes Both Type I and Type II Osteoporosis in Postmenopausal Women and Contributes to Bone Loss in Aging Men. J. Bone Miner. Res. 1998, 13, 763–773. [Google Scholar] [CrossRef] [PubMed]
  25. Riggs, B.L. The Mechanisms of Estrogen Regulation of Bone Resorption. J. Clin. Investig. 2000, 106, 1203–1204. [Google Scholar] [CrossRef] [PubMed]
  26. McNamara, L.M. Osteocytes and Estrogen Deficiency. Curr. Osteoporos. Rep. 2021, 19, 592–603. [Google Scholar] [CrossRef] [PubMed]
  27. Carani, C.; Qin, K.; Simoni, M.; Faustini-Fustini, M.; Serpente, S.; Boyd, J.; Korach, K.S.; Simpson, E.R. Effect of Testosterone and Estradiol in a Man with Aromatase Deficiency. N. Engl. J. Med. 1997, 337, 91–95. [Google Scholar] [CrossRef] [PubMed]
  28. Wasada, T.; Akamine, Y.; Kato, K.I.; Ibayashi, H.; Nomura, Y. Adrenal Contribution to Circulating Estrogens in Woman. Endocrinol. Jpn. 1978, 25, 123–128. [Google Scholar] [CrossRef]
  29. Almeida-Pereira, G.; Rorato, R.; Reis, L.C.; Elias, L.L.K.; Antunes-Rodrigues, J. The Role of Estradiol in Adrenal Insufficiency and Its Interaction with Corticosterone on Hydromineral Balance. Horm. Behav. 2013, 64, 847–855. [Google Scholar] [CrossRef]
  30. Caroccia, B.; Seccia, T.M.; Barton, M.; Rossi, G.P. Estrogen Signaling in the Adrenal Cortex: Implications for Blood Pressure Sex Differences. Hypertension 2016, 68, 840–848. [Google Scholar] [CrossRef]
  31. Kaludjerovic, J.; Ward, W.E. The Interplay between Estrogen and Fetal Adrenal Cortex. J. Nutr. Metab. 2012, 2012, 837901. [Google Scholar] [CrossRef]
  32. Meseguer, A.; Puche, C.; Cabero, A. Sex Steroid Biosynthesis in White Adipose Tissue. Horm. Metab. Res. 2002, 34, 731–736. [Google Scholar] [CrossRef] [PubMed]
  33. Simpson, E.R.; Merrill, J.C.; Hollub, A.J.; Graham-Lorence, S.; Mendelson, C.R. Regulation of Estrogen Biosynthesis by Human Adipose Cells. Endocr. Rev. 1989, 10, 136–148. [Google Scholar] [CrossRef]
  34. Mendelson, C.R.; Simpson, E.R. Regulation of Estrogen Biosynthesis by Human Adipose Cells in Vitro. Mol. Cell. Endocrinol. 1987, 52, 169–176. [Google Scholar] [CrossRef]
  35. Simpson, E.R.; Mahendroo, M.S.; Means, G.D.; Kilgore, M.W.; Jo Corbin, C.; Mendelson, C.R. Tissue-Specific Promoters Regulate Aromatase Cytochrome P450 Expression. J. Steroid Biochem. Mol. Biol. 1993, 44, 321–330. [Google Scholar] [CrossRef] [PubMed]
  36. Mair, K.M.; Gaw, R.; MacLean, M.R. Obesity, Estrogens and Adipose Tissue Dysfunction—Implications for Pulmonary Arterial Hypertension. Pulm. Circ. 2020, 10, 1–21. [Google Scholar] [CrossRef]
  37. Steiner, B.M.; Berry, D.C. The Regulation of Adipose Tissue Health by Estrogens. Front. Endocrinol. 2022, 13, 889923. [Google Scholar] [CrossRef]
  38. Cooke, P.S.; Heine, P.A.; Taylor, J.A.; Lubahn, D.B. The Role of Estrogen and Estrogen Receptor-α in Male Adipose Tissue. Mol. Cell. Endocrinol. 2001, 178, 147–154. [Google Scholar] [CrossRef]
  39. Kim, J.H.; Cho, H.T.; Kim, Y.J. The Role of Estrogen in Adipose Tissue Metabolism: Insights into Glucose Homeostasis Regulation. Endocr. J. 2014, 61, 1055–1067. [Google Scholar] [CrossRef]
  40. Simpson, E.R.; Mahendroo, M.S.; Nichols, J.E.; Bulun, S.E. Aromatase Gene Expression in Adipose Tissue: Relationship to Breast Cancer. Int. J. Fertil. Menopausal Stud. 1994, 39 (Suppl. 2), 75–83. [Google Scholar] [PubMed]
  41. Simpson, E.R. Sources of Estrogen and Their Importance. J. Steroid Biochem. Mol. Biol. 2003, 86, 225–230. [Google Scholar] [CrossRef] [PubMed]
  42. Zhao, H.; Zhou, L.; Shangguan, A.J.; Bulun, S.E. Aromatase Expression and Regulation in Breast and Endometrial Cancer. J. Mol. Endocrinol. 2016, 57, R19–R33. [Google Scholar] [CrossRef]
  43. Bulun, S.E.; Price, T.M.; Aitkens, J.; Mahendroos, M.S.; Simpso, E.R. A Link between Breast Cancer and Local Estrogen Biosynthesis Suggested by Quantification of Breast Adipose Tissue Aromatase Cytochrome P450 Transcripts Using Competitive Polymerase Chain Reaction after Reverse Transcription. J. Clin. Endocrinol. Metab. 1993, 77, 1622–1628. [Google Scholar] [CrossRef]
  44. Wang, X.; Simpson, E.R.; Brown, K.A. Aromatase Overexpression in Dysfunctional Adipose Tissue Links Obesity to Postmenopausal Breast Cancer. J. Steroid Biochem. Mol. Biol. 2015, 153, 35–44. [Google Scholar] [CrossRef] [PubMed]
  45. Simpson, E.R. Biology of Aromatase in the Mammary Gland. J. Mammary Gland Biol. Neoplasia 2000, 5, 251–258. [Google Scholar] [CrossRef] [PubMed]
  46. Hara, Y.; Waters, E.M.; McEwen, B.S.; Morrison, J.H. Estrogen Effects on Cognitive and Synaptic Health Over the Lifecourse. Physiol. Rev. 2015, 95, 785–807. [Google Scholar] [CrossRef]
  47. Azcoitia, I.; Yague, J.G.; Garcia-Segura, L.M. Estradiol Synthesis within the Human Brain. Neuroscience 2011, 191, 139–147. [Google Scholar] [CrossRef]
  48. Azcoitia, I.; Mendez, P.; Garcia-Segura, L.M. Aromatase in the Human Brain. Androg. Clin. Res. Ther. 2021, 2, 189–202. [Google Scholar] [CrossRef]
  49. Duncan, K.A.; Saldanha, C.J. Central Aromatization: A Dramatic and Responsive Defense against Threat and Trauma to the Vertebrate Brain. Front. Neuroendocrinol. 2020, 56, 100816. [Google Scholar] [CrossRef]
  50. Brann, D.W.; Lu, Y.; Wang, J.; Zhang, Q.; Thakkar, R.; Sareddy, G.R.; Pratap, U.P.; Tekmal, R.R.; Vadlamudi, R.K. Brain-Derived Estrogen and Neural Function. Neurosci. Biobehav. Rev. 2022, 132, 793–817. [Google Scholar] [CrossRef]
  51. Spool, J.A.; Bergan, J.F.; Remage-Healey, L. A Neural Circuit Perspective on Brain Aromatase. Front. Neuroendocrinol. 2022, 65, 100973. [Google Scholar] [CrossRef]
  52. Loucks, T.L.; Berga, S.L. Does Postmenopausal Estrogen Use Confer Neuroprotection? Semin. Reprod. Med. 2009, 27, 260–274. [Google Scholar] [CrossRef] [PubMed]
  53. Hiltunen, M.; Livonen, S.; Soininen, H. Aromatase Enzyme and Alzheimer’s Disease. Minerva Endocrinol. 2006, 31, 61–73. [Google Scholar] [PubMed]
  54. Hess, R.A. Estrogen in the Adult Male Reproductive Tract: A Review. Reprod. Biol. Endocrinol. 2003, 1, 52. [Google Scholar] [CrossRef]
  55. Adamopoulos, D.; Lawrence, D.M.; Vassilopoulos, P.; Kapolla, N.; Kontogeorgos, L.; McGarrigle, H.H.G. Hormone Levels in the Reproductive System of Normospermic Men and Patients with Oligospermia and Varicocele. J. Clin. Endocrinol. Metab. 1984, 59, 447–452. [Google Scholar] [CrossRef]
  56. O’Donnell, L.; Robertson, K.M.; Jones, M.E.; Simpson, E.R. Estrogen and Spermatogenesis. Endocr. Rev. 2001, 22, 289–318. [Google Scholar] [CrossRef]
  57. Smith, E.P.; Boyd, J.; Frank, G.R.; Takahashi, H.; Cohen, R.M.; Specker, B.; Williams, T.C.; Lubahn, D.B.; Korach, K.S. Estrogen Resistance Caused by a Mutation in the Estrogen-Receptor Gene in a Man. N. Engl. J. Med. 1994, 331, 1056–1061. [Google Scholar] [CrossRef] [PubMed]
  58. Fukami, M.; Ogata, T. Congenital Disorders of Estrogen Biosynthesis and Action. Best Pract. Res. Clin. Endocrinol. Metab. 2022, 3, 101580. [Google Scholar] [CrossRef]
  59. Morishima, A.; Grumbach, M.M.; Simpson, E.R.; Fisher, C.; Qin, K. Aromatase Deficiency in Male and Female Siblings Caused by a Novel Mutation and the Physiological Role of Estrogens. J. Clin. Endocrinol. Metab. 1995, 80, 3689–3698. [Google Scholar] [CrossRef]
  60. Huijben, M.; Huijsmans, R.L.N.; Lock, M.T.W.T.; de Kemp, V.F.; de Kort, L.M.O.; van Breda, J.H.M.K. Clomiphene Citrate for Male Infertility: A Systematic Review and Meta-Analysis. Andrology 2023, 11, 987–996. [Google Scholar] [CrossRef]
  61. Balló, A.; Busznyákné Székvári, K.; Czétány, P.; Márk, L.; Török, A.; Szántó, Á.; Máté, G. Estrogenic and Non-Estrogenic Disruptor Effect of Zearalenone on Male Reproduction: A Review. Int. J. Mol. Sci. 2023, 24, 1578. [Google Scholar] [CrossRef] [PubMed]
  62. Cooke, P.S.; Walker, W.H. Nonclassical Androgen and Estrogen Signaling Is Essential for Normal Spermatogenesis. Semin. Cell Dev. Biol. 2022, 121, 71–81. [Google Scholar] [CrossRef] [PubMed]
  63. Russell, N.; Grossmann, M. Mechanisms in Endocrinology: Estradiol as a Male Hormone. Eur. J. Endocrinol. 2019, 181, R23–R43. [Google Scholar] [CrossRef] [PubMed]
  64. Ellem, S.J.; Risbridger, G.P. The Dual, Opposing Roles of Estrogen in the Prostate. Ann. N. Y. Acad. Sci. 2009, 1155, 174–186. [Google Scholar] [CrossRef] [PubMed]
  65. Nelles, J.L.; Hu, W.Y.; Prins, G.S. Estrogen Action and Prostate Cancer. Expert Rev. Endocrinol. Metab. 2011, 6, 437–451. [Google Scholar] [CrossRef] [PubMed]
  66. Jarred, R.A.; McPherson, S.J.; Bianco, J.J.; Couse, J.F.; Korach, K.S.; Risbridger, G.P. Prostate Phenotypes in Estrogen-Modulated Transgenic Mice. Trends Endocrinol. Metab. 2002, 13, 163–168. [Google Scholar] [CrossRef] [PubMed]
  67. Ellem, S.J.; Schmitt, J.F.; Pedersen, J.S.; Frydenberg, M.; Risbridger, G.P. Local Aromatase Expression in Human Prostate Is Altered in Malignancy. J. Clin. Endocrinol. Metab. 2004, 89, 2434–2441. [Google Scholar] [CrossRef]
  68. Wu, Q.; Zhou, Y.; Chen, L.; Shi, J.; Wang, C.Y.; Miao, L.; Klocker, H.; Park, I.; Lee, C.; Zhang, J. Benign Prostatic Hyperplasia (BPH) Epithelial Cell Line BPH-1 Induces Aromatase Expression in Prostatic Stromal Cells via Prostaglandin E2. J. Endocrinol. 2007, 195, 89–94. [Google Scholar] [CrossRef]
  69. Lafront, C.; Germain, L.; Weidmann, C.; Audet-Walsh, É. A Systematic Study of the Impact of Estrogens and Selective Estrogen Receptor Modulators on Prostate Cancer Cell Proliferation. Sci. Rep. 2020, 10, 4024. [Google Scholar] [CrossRef]
  70. Sciarra, F. Anti-Estrogens and Aromatase Inhibitors: Tamoxifen and Testolactone. J. Endocrinol. Investig. 1988, 11, 755–762. [Google Scholar] [CrossRef]
  71. Greene, G.L.; Gilna, P.; Waterfield, M.; Baker, A.; Hort, Y.; Shine, J. Sequence and Expression of Human Estrogen Receptor Complementary DNA. Science 1986, 231, 1150–1154. [Google Scholar] [CrossRef]
  72. Walter, P.; Green, S.; Greene, G.; Krust, A.; Bornert, J.M.; Jeltsch, J.M.; Staub, A.; Jensen, E.; Scrace, G.; Waterfield, M. Cloning of the Human Estrogen Receptor CDNA. Proc. Natl. Acad. Sci. USA 1985, 82, 7889–7893. [Google Scholar] [CrossRef] [PubMed]
  73. Mosselman, S.; Polman, J.; Dijkema, R. ER Beta: Identification and Characterization of a Novel Human Estrogen Receptor. FEBS Lett. 1996, 392, 49–53. [Google Scholar] [CrossRef]
  74. Li, J.; Liu, Q.; Jiang, C. Signal Crosstalk and the Role of Estrogen Receptor Beta (ERβ) in Prostate Cancer. Med. Sci. Monit. 2022, 28, e935599-1. [Google Scholar] [CrossRef] [PubMed]
  75. Kumar, R.; Zakharov, M.N.; Khan, S.H.; Miki, R.; Jang, H.; Toraldo, G.; Singh, R.; Bhasin, S.; Jasuja, R. The dynamic structure of the estrogen receptor. Res. J. Amino Acids 2011, 2011, 812540. [Google Scholar] [CrossRef] [PubMed]
  76. Barkhem, T.; Carlsson, B.; Nilsson, Y.; Enmark, E.; Gustafsson, J.Å.; Nilsson, S. Differential Response of Estrogen Receptor Alpha and Estrogen Receptor Beta to Partial Estrogen Agonists/Antagonists. Mol. Pharmacol. 1998, 54, 105–112. [Google Scholar] [CrossRef]
  77. Yi, P.; Wang, Z.; Feng, Q.; Pintilie, G.D.; Foulds, C.E.; Lanz, R.B.; Ludtke, S.J.; Schmid, M.F.; Chiu, W.; O’Malley, B.W. Structure of a Biologically Active Estrogen Receptor-Coactivator Complex on DNA. Mol. Cell 2015, 57, 1047–1058. [Google Scholar] [CrossRef] [PubMed]
  78. Shiau, A.K.; Barstad, D.; Loria, P.M.; Cheng, L.; Kushner, P.J.; Agard, D.A.; Greene, G.L. The Structural Basis of Estrogen Receptor/Coactivator Recognition and the Antagonism of This Interaction by Tamoxifen. Cell 1998, 95, 927–937. [Google Scholar] [CrossRef]
  79. Harms, M.J.; Eick, G.N.; Goswami, D.; Colucci, J.K.; Griffin, P.R.; Ortlund, E.A.; Thornton, J.W. Biophysical Mechanisms for Large-Effect Mutations in the Evolution of Steroid Hormone Receptors. Proc. Natl. Acad. Sci. USA 2013, 110, 11475–11480. [Google Scholar] [CrossRef]
  80. Eick, G.N.; Colucci, J.K.; Harms, M.J.; Ortlund, E.A.; Thornton, J.W. Evolution of Minimal Specificity and Promiscuity in Steroid Hormone Receptors. PLoS Genet. 2012, 8, e1003072. [Google Scholar] [CrossRef]
  81. D’arrigo, G.; Gianquinto, E.; Rossetti, G.; Cruciani, G.; Lorenzetti, S.; Spyrakis, F. Binding of Androgen- and Estrogen-Like Flavonoids to Their Cognate (Non)Nuclear Receptors: A Comparison by Computational Prediction. Molecules 2021, 26, 1613. [Google Scholar] [CrossRef] [PubMed]
  82. McInerney, E.M.; Ince, B.A.; Shapiro, D.J.; Katzenellenbogen, B.S. A Transcriptionally Active Estrogen Receptor Mutant Is a Novel Type of Dominant Negative Inhibitor of Estrogen Action. Mol. Endocrinol. 1996, 10, 1519–1526. [Google Scholar] [CrossRef] [PubMed]
  83. Toy, W.; Shen, Y.; Won, H.; Green, B.; Sakr, R.A.; Will, M.; Li, Z.; Gala, K.; Fanning, S.; King, T.A.; et al. ESR1 Ligand-Binding Domain Mutations in Hormone-Resistant Breast Cancer. Nat. Genet. 2013, 45, 1439–1445. [Google Scholar] [CrossRef] [PubMed]
  84. Skolnick, M.H.; Thompson, E.A.; Bishop, D.T.; Cannon, L.A. Possible Linkage of a Breast Cancer-Susceptibility Locus to the ABO Locus: Sensitivity of LOD Scores to a Single New Recombinant Observation. Genet. Epidemiol. 1984, 1, 363–373. [Google Scholar] [CrossRef] [PubMed]
  85. Styrkarsdottir, U.; Halldorsson, B.V.; Gretarsdottir, S.; Gudbjartsson, D.F.; Walters, G.B.; Ingvarsson, T.; Jonsdottir, T.; Saemundsdottir, J.; Center, J.R.; Nguyen, T.V.; et al. Multiple Genetic Loci for Bone Mineral Density and Fractures. N. Engl. J. Med. 2008, 358, 2355–2365. [Google Scholar] [CrossRef]
  86. Herrington, D.M.; Howard, T.D.; Hawkins, G.A.; Reboussin, D.M.; Xu, J.; Zheng, S.L.; Brosnihan, K.B.; Meyers, D.A.; Bleecker, E.R. Estrogen-Receptor Polymorphisms and Effects of Estrogen Replacement on High-Density Lipoprotein Cholesterol in Women with Coronary Disease. N. Engl. J. Med. 2002, 346, 967–974. [Google Scholar] [CrossRef]
  87. Tsukamoto, K.; Inoue, S.; Hosoi, T.; Orimo, H.; Emi, M. Isolation and Radiation Hybrid Mapping of Dinucleotide Repeat Polymorphism at the Human Estrogen Receptor Beta Locus. J. Hum. Genet. 1998, 43, 73–74. [Google Scholar] [CrossRef]
  88. Ogawa, S.; Emi, M.; Shiraki, M.; Hosoi, T.; Ouchi, Y.; Inoue, S. Association of Estrogen Receptor Beta (ESR2) Gene Polymorphism with Blood Pressure. J. Hum. Genet. 2000, 45, 327–330. [Google Scholar] [CrossRef]
  89. Forsell, C.; Enmark, E.; Axelman, K.; Blomberg, M.; Wahlund, L.O.; Gustafsson, J.Å.; Lannfelt, L. Investigations of a CA Repeat in the Oestrogen Receptor Beta Gene in Patients with Alzheimer’s Disease. Eur. J. Hum. Genet. 2001, 9, 802–804. [Google Scholar] [CrossRef]
  90. Beleza-Meireles, A.; Kockum, I.; Lundberg, F.; Söderhäll, C.; Nordenskjöld, A. Risk Factors for Hypospadias in the Estrogen Receptor 2 Gene. J. Clin. Endocrinol. Metab. 2007, 92, 3712–3718. [Google Scholar] [CrossRef]
  91. Fytili, P.; Giannatou, E.; Papanikolaou, V.; Stripeli, F.; Karachalios, T.; Malizos, K.; Tsezou, A. Association of Repeat Polymorphisms in the Estrogen Receptors Alpha, Beta, and Androgen Receptor Genes with Knee Osteoarthritis. Clin. Genet. 2005, 68, 268–277. [Google Scholar] [CrossRef]
  92. Rosenkranz, K.; Hinney, A.; Ziegler, A.; Hermann, H.; Fichter, M.; Mayer, H.; Siegfried, W.; Young, J.K.; Remschmidt, H.; Hebebrand, J. Systematic Mutation Screening of the Estrogen Receptor Beta Gene in Probands of Different Weight Extremes: Identification of Several Genetic Variants. J. Clin. Endocrinol. Metab. 1998, 83, 4524–4527. [Google Scholar] [CrossRef] [PubMed]
  93. Lang-Muritano, M.; Sproll, P.; Wyss, S.; Kolly, A.; Hürlimann, R.; Konrad, D.; Biason-Lauber, A. Early-Onset Complete Ovarian Failure and Lack of Puberty in a Woman With Mutated Estrogen Receptor β (ESR2). J. Clin. Endocrinol. Metab. 2018, 103, 3748–3756. [Google Scholar] [CrossRef] [PubMed]
  94. Baetens, D.; Güran, T.; Mendonca, B.B.; Gomes, N.L.; De Cauwer, L.; Peelman, F.; Verdin, H.; Vuylsteke, M.; Van Der Linden, M.; Atay, Z.; et al. Biallelic and Monoallelic ESR2 Variants Associated with 46,XY Disorders of Sex Development. Genet. Med. 2018, 20, 717–727. [Google Scholar] [CrossRef] [PubMed]
  95. Biason-Lauber, A.; Lang-Muritano, M. Estrogens: Two Nuclear Receptors, Multiple Possibilities. Mol. Cell. Endocrinol. 2022, 554, 111710. [Google Scholar] [CrossRef] [PubMed]
  96. Chen, P.; Li, B.; Ou-Yang, L. Role of Estrogen Receptors in Health and Disease. Front. Endocrinol. 2022, 13, 839005. [Google Scholar] [CrossRef]
  97. Paterni, I.; Granchi, C.; Katzenellenbogen, J.A.; Minutolo, F. Estrogen Receptors Alpha (ERα) and Beta (ERβ): Subtype-Selective Ligands and Clinical Potential. Steroids 2014, 90, 13–29. [Google Scholar] [CrossRef]
  98. Brzozowski, A.M.; Pike, A.C.W.; Dauter, Z.; Hubbard, R.E.; Bonn, T.; Engström, O.; Öhman, L.; Greene, G.L.; Gustafsson, J.Å.; Carlquist, M. Molecular Basis of Agonism and Antagonism in the Oestrogen Receptor. Nature 1997, 389, 753–758. [Google Scholar] [CrossRef]
  99. Paech, K.; Webb, P.; Kuiper, G.G.J.M.; Nilsson, S.; Gustafsson, J.Å.; Kushner, P.J.; Scanlan, T.S. Differential Ligand Activation of Estrogen Receptors ERalpha and ERbeta at AP1 Sites. Science 1997, 277, 1508–1510. [Google Scholar] [CrossRef]
  100. Fixemer, T.; Remberger, K.; Bonkhoff, H. Differential Expression of the Estrogen Receptor Beta (ERβ) in Human Prostate Tissue, Premalignant Changes, and in Primary, Metastatic, and Recurrent Prostatic Adenocarcinoma. Prostate 2003, 54, 79–87. [Google Scholar] [CrossRef]
  101. Bonkhoff, H. Estrogen Receptor Signaling in Prostate Cancer: Implications for Carcinogenesis and Tumor Progression. Prostate 2018, 78, 2–10. [Google Scholar] [CrossRef]
  102. Qu, L.G.; Wardan, H.; Davis, I.D.; Iddawela, M.; Sluka, P.; Pezaro, C.J. Circulating Oestrogen Receptor Mutations and Splice Variants in Advanced Prostate Cancer. BJU Int. 2019, 124 (Suppl. S1), 50–56. [Google Scholar] [CrossRef] [PubMed]
  103. Thellenberg-Karlsson, C.; Lindström, S.; Malmer, B.; Wiklund, F.; Augustsson-Bälter, K.; Adami, H.O.; Stattin, P.; Nilsson, M.; Dahlman-Wright, K.; Gustafsson, J.Å.; et al. Estrogen Receptor Beta Polymorphism Is Associated with Prostate Cancer Risk. Clin. Cancer Res. 2006, 12, 1936–1941. [Google Scholar] [CrossRef]
  104. Latil, A.; Biè, I.; Vidaud, D.; Lidereau, R.; Berthon, P.; Cussenot, O.; Vidaud, M. Evaluation of Androgen, Estrogen (ER and ER), and Progesterone Receptor Expression in Human Prostate Cancer by Real-Time Quantitative Reverse Transcription-Polymerase Chain Reaction Assays. Cancer Res. 2001, 61, 1919–1926. [Google Scholar] [PubMed]
  105. Sehgal, P.D.; Bauman, T.M.; Nicholson, T.M.; Vellky, J.E.; Ricke, E.A.; Tang, W.; Xu, W.; Huang, W.; Ricke, W.A. Tissue-Specific Quantification and Localization of Androgen and Estrogen Receptors in Prostate Cancer. Hum. Pathol. 2019, 89, 99–108. [Google Scholar] [CrossRef] [PubMed]
  106. Chakravarty, D.; Sboner, A.; Nair, S.S.; Giannopoulou, E.; Li, R.; Hennig, S.; Mosquera, J.M.; Pauwels, J.; Park, K.; Kossai, M.; et al. The Oestrogen Receptor Alpha-Regulated LncRNA NEAT1 Is a Critical Modulator of Prostate Cancer. Nat. Commun. 2014, 5, 5383. [Google Scholar] [CrossRef] [PubMed]
  107. Li, L.C.; Shiina, H.; Deguchi, M.; Zhao, H.; Okino, S.T.; Kane, C.J.; Carroll, P.R.; Igawa, M.; Dahiya, R. Age-Dependent Methylation of ESR1 Gene in Prostate Cancer. Biochem. Biophys. Res. Commun. 2004, 321, 455–461. [Google Scholar] [CrossRef]
  108. Li, L.C.; Chui, R.; Nakajima, K.; Oh, B.R.; Au, H.C.; Dahiya, R. Frequent Methylation of Estrogen Receptor in Prostate Cancer: Correlation with Tumor Progression. Cancer Res. 2000, 60, 702–706. [Google Scholar]
  109. Hu, C.; Liu, Y.; Jiang, S.; Chen, H.; Xu, H.; Hu, J.; Li, C.; Xia, H. The Variable Association between Expression and Methylation of Estrogen Receptors and the Survival of Patients with Different Tumors. Clin. Transl. Med. 2020, 10, e49. [Google Scholar] [CrossRef]
  110. Clemons, J.; Michael Glodé, L.; Gao, D.; Flaig, T.W. Low-Dose Diethylstilbestrol for the Treatment of Advanced Prostate Cancer. Urol. Oncol. Semin. Orig. Investig. 2013, 31, 198–204. [Google Scholar] [CrossRef]
  111. Shamash, J.; Powles, T.; Sarker, S.J.; Protheroe, A.; Mithal, N.; Mills, R.; Beard, R.; Wilson, P.; Tranter, N.; O’Brien, N.; et al. A Multi-Centre Randomised Phase III Trial of Dexamethasone vs Dexamethasone and Diethylstilbestrol in Castration-Resistant Prostate Cancer: Immediate vs Deferred Diethylstilbestrol. Br. J. Cancer 2011, 104, 620–628. [Google Scholar] [CrossRef] [PubMed]
  112. Grenader, T.; Plotkin, Y.; Gips, M.; Cherny, N.; Gabizon, A. Diethylstilbestrol for the Treatment of Patients with Castration-Resistant Prostate Cancer: Retrospective Analysis of a Single Institution Experience. Oncol. Rep. 2014, 31, 428–434. [Google Scholar] [CrossRef] [PubMed]
  113. Stein, M.; Goodin, S.; Doyle-Lindrud, S.; Silberberg, J.; Kane, M.; Metzger, D.; Eddy, S.; Shih, W.; DiPaola, R.S. Transdermal Estradiol in Castrate and Chemotherapy Resistant Prostate Cancer. Med. Sci. Monit. 2012, 18, CR260–CR264. [Google Scholar] [CrossRef] [PubMed]
  114. Konishi, N.; Nakaoka, S.; Hiasa, Y.; Kitahori, Y.; Ohshima, M.; Samma, S.; Okajima, E. Immunohistochemical Evaluation of Estrogen Receptor Status in Benign Prostatic Hypertrophy and in Prostate Carcinoma and the Relationship to Efficacy of Endocrine Therapy. Oncology 1993, 50, 259–263. [Google Scholar] [CrossRef] [PubMed]
  115. Olczak, M.; Orzechowska, M.J.; Bednarek, A.K.; Lipiński, M. The Transcriptomic Profiles of ESR1 and MMP3 Stratify the Risk of Biochemical Recurrence in Primary Prostate Cancer beyond Clinical Features. Int. J. Mol. Sci. 2023, 24, 8399. [Google Scholar] [CrossRef]
  116. Yeh, C.R.; Slavin, S.; Da, J.; Hsu, I.; Luo, J.; Xiao, G.Q.; Ding, J.; Chou, F.J.; Yeh, S. Estrogen Receptor α in Cancer Associated Fibroblasts Suppresses Prostate Cancer Invasion via Reducing CCL5, IL6 and Macrophage Infiltration in the Tumor Microenvironment. Mol. Cancer 2016, 15, 7. [Google Scholar] [CrossRef]
  117. Slavin, S.; Yeh, C.-R.; Da, J.; Yu, S.; Miyamoto, H.; Messing, E.M.; Guancial, E.; Yeh, S. Estrogen Receptor α in Cancer-Associated Fibroblasts Suppresses Prostate Cancer Invasion via Modulation of Thrombospondin 2 and Matrix Metalloproteinase 3. Carcinogenesis 2014, 35, 1301–1309. [Google Scholar] [CrossRef]
  118. Grindstad, T.; Skjefstad, K.; Andersen, S.; Ness, N.; Nordby, Y.; Al-Saad, S.; Fismen, S.; Donnem, T.; Khanehkenari, M.R.; Busund, L.T.; et al. Estrogen Receptors α and β and Aromatase as Independent Predictors for Prostate Cancer Outcome. Sci. Rep. 2016, 6, 33114. [Google Scholar] [CrossRef]
  119. Nojima, D.; Li, L.-C.; Dharia, A.; Perinchery, G.; Ribeiro-Filho, L.; Yen, T.-S.B.; Dahiya, R. CpG Hypermethylation of the Promoter Region Inactivates the Estrogen Receptor-Gene in Patients with Prostate Carcinoma. Cancer 2001, 92, 2076–2083. [Google Scholar] [CrossRef]
  120. Božović, A.; Mandušić, V.; Todorović, L.; Krajnović, M. Estrogen Receptor Beta: The Promising Biomarker and Potential Target in Metastases. Int. J. Mol. Sci. 2021, 22, 1656. [Google Scholar] [CrossRef]
  121. Di Zazzo, E.; Galasso, G.; Giovannelli, P.; Di Donato, M.; Bilancio, A.; Perillo, B.; Sinisi, A.A.; Migliaccio, A.; Castoria, G. Estrogen Receptors in Epithelial-Mesenchymal Transition of Prostate Cancer. Cancers 2019, 11, 1418. [Google Scholar] [CrossRef]
  122. Andersson, S.; Sundberg, M.; Pristovsek, N.; Ibrahim, A.; Jonsson, P.; Katona, B.; Clausson, C.M.; Zieba, A.; Ramström, M.; Söderberg, O.; et al. Insufficient Antibody Validation Challenges Oestrogen Receptor Beta Research. Nat. Commun. 2017, 8, 15840. [Google Scholar] [CrossRef]
  123. Gustafsson, J.A.; Strom, A.; Warner, M. Update on ERbeta. J. Steroid Biochem. Mol. Biol. 2019, 191, 105312. [Google Scholar] [CrossRef]
  124. Belluti, S.; Rigillo, G.; Imbriano, C. Transcription Factors in Cancer: When Alternative Splicing Determines Opposite Cell Fates. Cells 2020, 9, 760. [Google Scholar] [CrossRef]
  125. Belluti, S.; Semeghini, V.; Rigillo, G.; Ronzio, M.; Benati, D.; Torricelli, F.; Reggiani Bonetti, L.; Carnevale, G.; Grisendi, G.; Ciarrocchi, A.; et al. Alternative Splicing of NF-YA Promotes Prostate Cancer Aggressiveness and Represents a New Molecular Marker for Clinical Stratification of Patients. J. Exp. Clin. Cancer Res. 2021, 40, 362. [Google Scholar] [CrossRef]
  126. Munkley, J.; Livermore, K.; Rajan, P.; Elliott, D.J. RNA Splicing and Splicing Regulator Changes in Prostate Cancer Pathology. Hum. Genet. 2017, 136, 1143. [Google Scholar] [CrossRef]
  127. Del Giudice, M.; Foster, J.G.; Peirone, S.; Rissone, A.; Caizzi, L.; Gaudino, F.; Parlato, C.; Anselmi, F.; Arkell, R.; Guarrera, S.; et al. FOXA1 Regulates Alternative Splicing in Prostate Cancer. Cell Rep. 2022, 40, 111404. [Google Scholar] [CrossRef]
  128. Flouriot, G.; Brand, H.; Denger, S.; Metivier, R.; Kos, M.; Reid, G.; Sonntag-Buck, V.; Gannon, F. Identification of a New Isoform of the Human Estrogen Receptor-Alpha (HER-α) That Is Encoded by Distinct Transcripts and That Is Able to Repress HER-α Activation Function 1. EMBO J. 2000, 19, 4688–4700. [Google Scholar] [CrossRef]
  129. Denger, S.; Reid, G.; Koš, M.; Flouriot, G.; Parsch, D.; Brand, H.; Korach, K.S.; Sonntag-Buck, V.; Gannon, F. ERα Gene Expression in Human Primary Osteoblasts: Evidence for the Expression of Two Receptor Proteins. Mol. Endocrinol. 2001, 15, 2064–2077. [Google Scholar] [CrossRef]
  130. Wang, Z.Y.; Zhang, X.; Shen, P.; Loggie, B.W.; Chang, Y.; Deuel, T.F. Identification, Cloning, and Expression of Human Estrogen Receptor-A36, a Novel Variant of Human Estrogen Receptor-A66. Biochem. Biophys. Res. Commun. 2005, 336, 1023–1027. [Google Scholar] [CrossRef]
  131. Lee, L.M.J.; Cao, J.; Deng, H.; Chen, P.; Gatalica, Z.; Wang, Z.Y. ER-A36, a Novel Variant of ER-α, Is Expressed in ER-Positive and -Negative Human Breast Carcinomas. Anticancer Res. 2008, 28, 479–484. [Google Scholar]
  132. Wang, Z.Y.; Zhang, X.T.; Shen, P.; Loggie, B.W.; Chang, Y.C.; Deuel, T.F. A Variant of Estrogen Receptor-{alpha}, HER-{alpha}36: Transduction of Estrogen- and Antiestrogen-Dependent Membrane-Initiated Mitogenic Signaling. Proc. Natl. Acad. Sci. USA 2006, 103, 9063–9068. [Google Scholar] [CrossRef]
  133. Kim, C.K.; Torcaso, A.; Asimes, A.; Chung, W.C.J.; Pak, T.R. Structural and Functional Characteristics of Estrogen Receptor Beta (ERβ) Splice Variants: Implications for the Aging Brain. J. Neuroendocrinol. 2018, 30, e12488. [Google Scholar] [CrossRef]
  134. Moore, J.T.; McKee, D.D.; Slentz-Kesler, K.; Moore, L.B.; Jones, S.A.; Horne, E.L.; Su, J.L.; Kliewer, S.A.; Lehmann, J.M.; Willson, T.M. Cloning and Characterization of Human Estrogen Receptor β Isoforms. Biochem. Biophys. Res. Commun. 1998, 247, 75–78. [Google Scholar] [CrossRef]
  135. Leung, Y.K.; Mak, P.; Hassan, S.; Ho, S.M. Estrogen Receptor (ER)-Beta Isoforms: A Key to Understanding ER-Beta Signaling. Proc. Natl. Acad. Sci. USA 2006, 103, 13162–13167. [Google Scholar] [CrossRef]
  136. Zhao, C.; Dahlman-Wright, K.; Gustafsson, J.A. Estrogen Receptor Beta: An Overview and Update. Nucl. Recept. Signal. 2008, 6, e003. [Google Scholar] [CrossRef]
  137. Faria, M.; Shepherd, P.; Pan, Y.; Chatterjee, S.S.; Navone, N.; Gustafsson, J.-Å.; Strom, A.; Faria, M.; Shepherd, P.; Pan, Y.; et al. The Estrogen Receptor Variants Β2 and Β5 Induce Stem Cell Characteristics and Chemotherapy Resistance in Prostate Cancer through Activation of Hypoxic Signaling. Oncotarget 2018, 9, 36273–36288. [Google Scholar] [CrossRef]
  138. Formaggio, N.; Rubin, M.A.; Theurillat, J.P. Loss and Revival of Androgen Receptor Signaling in Advanced Prostate Cancer. Oncogene 2021, 40, 1205–1216. [Google Scholar] [CrossRef]
  139. Leach, D.A.; Fernandes, R.C.; Bevan, C.L. Cellular Specificity of Androgen Receptor, Coregulators, and Pioneer Factors in Prostate Cancer. Endocr. Oncol. 2022, 2, R112–R131. [Google Scholar] [CrossRef]
  140. Özturan, D.; Morova, T.; Lack, N.A. Androgen Receptor-Mediated Transcription in Prostate Cancer. Cells 2022, 11, 898. [Google Scholar] [CrossRef]
  141. Lau, K.M.; To, K.F. Importance of Estrogenic Signaling and Its Mediated Receptors in Prostate Cancer. Int. J. Mol. Sci. 2016, 17, 1434. [Google Scholar] [CrossRef]
  142. Ricke, W.A.; McPherson, S.J.; Bianco, J.J.; Cunha, G.R.; Wang, Y.; Risbridger, G.P. Prostatic Hormonal Carcinogenesis Is Mediated by in Situ Estrogen Production and Estrogen Receptor Alpha Signaling. FASEB J. 2008, 22, 1512–1520. [Google Scholar] [CrossRef]
  143. Leung, Y.K.; Lam, H.M.; Wu, S.; Song, D.; Levin, L.; Cheng, L.; Wu, C.L.; Ho, S.M. Estrogen Receptor Β2 and Β5 Are Associated with Poor Prognosis in Prostate Cancer, and Promote Cancer Cell Migration and Invasion. Endocr. Relat. Cancer 2010, 17, 675. [Google Scholar] [CrossRef] [PubMed]
  144. Omoto, Y.; Iwase, H. Clinical Significance of Estrogen Receptor β in Breast and Prostate Cancer from Biological Aspects. Cancer Sci. 2015, 106, 337. [Google Scholar] [CrossRef] [PubMed]
  145. Dey, P.; Jonsson, P.; Hartman, J.; Williams, C.; Ström, A.; Gustafsson, J.Å. Estrogen Receptors Β1 and Β2 Have Opposing Roles in Regulating Proliferation and Bone Metastasis Genes in the Prostate Cancer Cell Line PC3. Mol. Endocrinol. 2012, 26, 1991–2003. [Google Scholar] [CrossRef]
  146. Lombardi, A.P.G.; Vicente, C.M.; Porto, C.S. Estrogen Receptors Promote Migration, Invasion and Colony Formation of the Androgen-Independent Prostate Cancer Cells PC-3 Through β-Catenin Pathway. Front. Endocrinol. 2020, 11, 527553. [Google Scholar] [CrossRef]
  147. Stevens, J.H.; Bano, A.; Bensaoula, L.; Strom, A.M.; Gustafsson, J.-Å. Estrogen Receptor β Isoforms Regulate Chemotherapy Resistance and the Cancer Stem Cell Population in Prostate Cancer Cells. Receptors 2023, 2, 176–190. [Google Scholar] [CrossRef]
  148. Leav, I.; Lau, K.M.; Adams, J.Y.; McNeal, J.E.; Taplin, M.E.; Wang, J.; Singh, H.; Ho, S.M. Comparative Studies of the Estrogen Receptors β and α and the Androgen Receptor in Normal Human Prostate Glands, Dysplasia, and in Primary and Metastatic Carcinoma. Am. J. Pathol. 2001, 159, 79–92. [Google Scholar] [CrossRef]
  149. Cheng, J.; Lee, E.J.; Madison, L.D.; Lazennec, G. Expression of Estrogen Receptor β in Prostate Carcinoma Cells Inhibits Invasion and Proliferation and Triggers Apoptosis. FEBS Lett. 2004, 566, 169–172. [Google Scholar] [CrossRef]
  150. Corey, E.; Quinn, J.E.; Emond, M.J.; Buhler, K.R.; Brown, L.G.; Vessella, R.L. Inhibition of Androgen-Independent Growth of Prostate Cancer Xenografts by 17beta-Estradiol. Clin. Cancer Res. 2002, 8, 1003–1007. [Google Scholar]
  151. Weihua, Z.; Mäkelä, S.; Andersson, L.C.; Salmi, S.; Saji, S.; Webster, J.I.; Jensen, E.V.; Nilsson, S.; Warner, M.; Gustafsson, J.Å. A Role for Estrogen Receptor β in the Regulation of Growth of the Ventral Prostate. Proc. Natl. Acad. Sci. USA 2001, 98, 6330–6335. [Google Scholar] [CrossRef]
  152. Imamov, O.; Morani, A.; Shim, G.J.; Omoto, Y.; Thulin-Andersson, C.; Warner, M.; Gustafsson, J.Å. Estrogen Receptor β Regulates Epithelial Cellular Differentiation in the Mouse Ventral Prostate. Proc. Natl. Acad. Sci. USA 2004, 101, 9375–9380. [Google Scholar] [CrossRef]
  153. Kumar, V.; Chambon, P. The Estrogen Receptor Binds Tightly to Its Responsive Element as a Ligand-Induced Homodimer. Cell 1988, 55, 145–156. [Google Scholar] [CrossRef]
  154. Cowley, S.M.; Hoare, S.; Mosselman, S.; Parker, M.G. Estrogen Receptors a and b Form Heterodimers on DNA. J. Biol. Chem. 1997, 272, 19858–19862. [Google Scholar] [CrossRef] [PubMed]
  155. Jakacka, M.; Ito, M.; Weiss, J.; Chien, P.-Y.; Gehm, B.D.; Jameson, J.L. Estrogen Receptor Binding to DNA Is Not Required for Its Activity through the Nonclassical AP1 Pathway. Pediatrics 2001, 276, 13615–13621. [Google Scholar] [CrossRef] [PubMed]
  156. McKenna, N.J.; Lanz, R.B.; O’Malley, B.W. Nuclear Receptor Coregulators: Cellular and Molecular Biology. Endocr. Rev. 1999, 20, 321–344. [Google Scholar] [CrossRef] [PubMed]
  157. O’Lone, R.; Frith, M.C.; Karlsson, E.K.; Hansen, U. Genomic Targets of Nuclear Estrogen Receptors. Mol. Endocrinol. 2004, 18, 1859–1875. [Google Scholar] [CrossRef]
  158. Frasor, J.; Danes, J.M.; Komm, B.; Chang, K.C.N.; Richard Lyttle, C.; Katzenellenbogen, B.S. Profiling of Estrogen Up- and down-Regulated Gene Expression in Human Breast Cancer Cells: Insights into Gene Networks and Pathways Underlying Estrogenic Control of Proliferation and Cell Phenotype. Endocrinology 2003, 144, 4562–4574. [Google Scholar] [CrossRef]
  159. Sheikh, M.S.; Shao, Z.-M.; Chen, J.-C.; Li, X.-S.; Hussain, A.; Fontana, J.A. Expression of Estrogen Receptors in Estrogen Receptor–Negative Human Breast Carcinoma Cells: Modulation of Epidermal Growth Factor-receptor (EGF-R) and Transforming Growth Factor α (TGFα) Gene Expression. J. Cell. Biochem. 1994, 54, 289–298. [Google Scholar] [CrossRef]
  160. Ostano, P.; Mello-grand, M.; Sesia, D.; Gregnanin, I.; Peraldo-neia, C.; Guana, F.; Jachetti, E.; Farsetti, A.; Chiorino, G. Gene Expression Signature Predictive of Neuroendocrine Transformation in Prostate Adenocarcinoma. Int. J. Mol. Sci. 2020, 21, 1078. [Google Scholar] [CrossRef]
  161. Khan, M.Z.I.; Uzair, M.; Nazli, A.; Chen, J.Z. An Overview on Estrogen Receptors Signaling and Its Ligands in Breast Cancer. Eur. J. Med. Chem. 2022, 241, 114658. [Google Scholar] [CrossRef] [PubMed]
  162. Rej, R.K.; Thomas, J.E.; Acharyya, R.K.; Rae, J.M.; Wang, S. Targeting the Estrogen Receptor for the Treatment of Breast Cancer: Recent Advances and Challenges. J. Med. Chem. 2023, 66, 8339–8381. [Google Scholar] [CrossRef] [PubMed]
  163. Tecalco-Cruz, A.C.; Macías-Silva, M.; Ramírez-Jarquín, J.O.; Ramírez-Jarquín, U.N. Decoding the Therapeutic Implications of the ERα Stability and Subcellular Distribution in Breast Cancer. Front. Endocrinol. 2022, 13, 867448. [Google Scholar] [CrossRef]
  164. Dey, P.; Ström, A.; Gustafsson, J.A. Estrogen Receptor β Upregulates FOXO3a and Causes Induction of Apoptosis through PUMA in Prostate Cancer. Oncogene 2013, 33, 4213–4225. [Google Scholar] [CrossRef] [PubMed]
  165. Xiao, L.; Xiao, M.; Zou, M.; Xu, W. Estrogen Receptor β Inhibits Prostate Cancer Cell Proliferation through Downregulating TGF-Β1/IGF-1 Signaling. Int. J. Clin. Exp. Pathol. 2017, 10, 8569. [Google Scholar] [PubMed]
  166. Chaurasiya, S.; Widmann, S.; Botero, C.; Lin, C.Y.; Gustafsson, J.Å.; Strom, A.M. Estrogen Receptor β Exerts Tumor Suppressive Effects in Prostate Cancer through Repression of Androgen Receptor Activity. PLoS ONE 2020, 15, e0226057. [Google Scholar] [CrossRef]
  167. Wu, W.F.; Maneix, L.; Insunza, J.; Nalvarte, I.; Antonson, P.; Kere, J.; Yu, N.Y.L.; Tohonen, V.; Katayama, S.; Einarsdottir, E.; et al. Estrogen Receptor β, a Regulator of Androgen Receptor Signaling in the Mouse Ventral Prostate. Proc. Natl. Acad. Sci. USA 2017, 114, E3816–E3822. [Google Scholar] [CrossRef]
  168. Gehrig, J.; Kaulfuß, S.; Jarry, H.; Bremmer, F.; Stettner, M.; Burfeind, P.; Thelen, P.; Gehrig, J.; Kaulfuß, S.; Jarry, H.; et al. Prospects of Estrogen Receptor β Activation in the Treatment of Castration-Resistant Prostate Cancer. Oncotarget 2017, 8, 34971–34979. [Google Scholar] [CrossRef]
  169. McPherson, S.J.; Hussain, S.; Balanathan, P.; Hedwards, S.L.; Niranjan, B.; Grant, M.; Chandrasiri, U.P.; Toivanen, R.; Wang, Y.; Taylor, R.A.; et al. Estrogen Receptor–β Activated Apoptosis in Benign Hyperplasia and Cancer of the Prostate Is Androgen Independent and TNFα Mediated. Proc. Natl. Acad. Sci. USA 2010, 107, 3123–3128. [Google Scholar] [CrossRef]
  170. Mak, P.; Leav, I.; Pursell, B.; Bae, D.; Yang, X.; Taglienti, C.A.; Gouvin, L.M.; Sharma, V.M.; Mercurio, A.M. ERβ Impedes Prostate Cancer EMT by Destabilizing HIF-1α and Inhibiting VEGF-Mediated Snail Nuclear Localization: Implications for Gleason Grading. Cancer Cell 2010, 17, 319–332. [Google Scholar] [CrossRef]
  171. Guerini, V.; Sau, D.; Scaccianoce, E.; Rusmini, P.; Ciana, P.; Maggi, A.; Martini, P.G.V.; Katzenellenbogen, B.S.; Martini, L.; Motta, M.; et al. The Androgen Derivative 5alpha-Androstane-3beta,17beta-Diol Inhibits Prostate Cancer Cell Migration through Activation of the Estrogen Receptor Beta Subtype. Cancer Res. 2005, 65, 5445–5453. [Google Scholar] [CrossRef]
  172. Dey, P.; Velazquez-Villegas, L.A.; Faria, M.; Turner, A.; Jonsson, P.; Webb, P.; Williams, C.; Gustafsson, J.-Å.; Ström, A.M. Estrogen Receptor Β2 Induces Hypoxia Signature of Gene Expression by Stabilizing HIF-1α in Prostate Cancer. PLoS ONE 2015, 10, e0128239. [Google Scholar] [CrossRef]
  173. Nelson, A.W.; Tilley, W.D.; Neal, D.E.; Carroll, J.S. Estrogen Receptor Beta in Prostate Cancer: Friend or Foe? Endocr. Relat. Cancer 2014, 21, T219–T234. [Google Scholar] [CrossRef] [PubMed]
  174. Grubisha, M.J.; Defranco, D.B. Local Endocrine, Paracrine and Redox Signaling Networks Impact Estrogen and Androgen Crosstalk in the Prostate Cancer Microenvironment. Steroids 2013, 78, 538–541. [Google Scholar] [CrossRef] [PubMed]
  175. Grubisha, M.J.; Cifuentes, M.E.; Hammes, S.R.; DeFranco, D.B. A Local Paracrine and Endocrine Network Involving TGFβ, Cox-2, ROS, and Estrogen Receptor β Influences Reactive Stromal Cell Regulation of Prostate Cancer Cell Motility. Mol. Endocrinol. 2012, 26, 940–954. [Google Scholar] [CrossRef] [PubMed]
  176. Thiebaut, C.; Vlaeminck-Guillem, V.; Trédan, O.; Poulard, C.; Le Romancer, M. Non-Genomic Signaling of Steroid Receptors in Cancer. Mol. Cell. Endocrinol. 2021, 538, 111453. [Google Scholar] [CrossRef]
  177. Souza, D.S.; Macheroni, C.; Pereira, G.J.S.; Vicente, C.M.; Porto, C.S. Molecular Regulation of Prostate Cancer by Galectin-3 and Estrogen Receptor. Front. Endocrinol. 2023, 14, 1124111. [Google Scholar] [CrossRef]
  178. Hurtado, A.; Pinós, T.; Barbosa-Desongles, A.; López-Avilés, S.; Barquinero, J.; Petriz, J.; Santamaria-Martínez, A.; Morote, J.; De Torres, I.; Bellmunt, J.; et al. Estrogen Receptor Beta Displays Cell Cycle-Dependent Expression and Regulates the G1 Phase through a Non-Genomic Mechanism in Prostate Carcinoma Cells. Anal. Cell. Pathol. 2008, 30, 349–365. [Google Scholar] [CrossRef]
  179. Zhao, Z.; Yu, H.; Kong, Q.; Liu, C.; Tian, Y.; Zeng, X.; Li, D. Effect of ERβ-Regulated ERK1/2 Signaling on Biological Behaviors of Prostate Cancer Cells. Am. J. Transl. Res. 2017, 9, 2775. [Google Scholar]
  180. Silva, R.D.S.; Lombardi, A.P.G.; de Souza, D.S.; Vicente, C.M.; Porto, C.S. Activation of Estrogen Receptor Beta (ERβ) Regulates the Expression of N-Cadherin, E-Cadherin and β-Catenin in Androgen-Independent Prostate Cancer Cells. Int. J. Biochem. Cell Biol. 2018, 96, 40–50. [Google Scholar] [CrossRef]
  181. Augusto, T.V.; Georgina, C.D.S.; Rodrigues, C.M.P.; Teixeira, N.; Amaral, C. Acquired Resistance to Aromatase Inhibitors: Where We Stand! Endocr. Relat. Cancer 2018, 25, R283–R301. [Google Scholar] [CrossRef] [PubMed]
  182. Ma, C.X.; Reinert, T.; Chmielewska, I.; Ellis, M.J. Mechanisms of Aromatase Inhibitor Resistance. Nat. Rev. Cancer 2015, 15, 261–275. [Google Scholar] [CrossRef] [PubMed]
  183. Jordan, V.C. Tamoxifen: A Most Unlikely Pioneering Medicine. Nat. Rev. Drug Discov. 2003, 2, 205–213. [Google Scholar] [CrossRef] [PubMed]
  184. Zwart, W.; Griekspoor, A.; Berno, V.; Lakeman, K.; Jalink, K.; Mancini, M.; Neefjes, J.; Michalides, R. PKA-Induced Resistance to Tamoxifen Is Associated with an Altered Orientation of ERalpha towards Co-Activator SRC-1. EMBO J. 2007, 26, 3534–3544. [Google Scholar] [CrossRef]
  185. Ali, S.; Buluwela, L.; Coombes, R.C. Antiestrogens and Their Therapeutic Applications in Breast Cancer and Other Diseases. Annu. Rev. Med. 2011, 62, 217–232. [Google Scholar] [CrossRef]
  186. Shiota, M.; Fujimoto, N.; Kashiwagi, E.; Eto, M. The Role of Nuclear Receptors in Prostate Cancer. Cells 2019, 8, 602. [Google Scholar] [CrossRef]
  187. Wakeling, A.E.; Dukes, M.; Bowler, J. A Potent Specific Pure Antiestrogen with Clinical Potential. Cancer Res. 1991, 51, 3867–3873. [Google Scholar]
  188. Movérare-Skrtic, S.; Börjesson, A.E.; Farman, H.H.; Sjögren, K.; Windahl, S.H.; Lagerquist, M.K.; Andersson, A.; Stubelius, A.; Carlsten, H.; Gustafsson, J.-Å.; et al. The Estrogen Receptor Antagonist ICI 182,780 Can Act Both as an Agonist and an Inverse Agonist When Estrogen Receptor α AF-2 Is Modified. Proc. Natl. Acad. Sci. USA 2014, 111, 1180–1185. [Google Scholar] [CrossRef]
  189. Zhao, L.; Zhou, S.; Gustafsson, J.Å. Nuclear Receptors: Recent Drug Discovery for Cancer Therapies. Endocr. Rev. 2019, 40, 1207–1249. [Google Scholar] [CrossRef]
  190. Burris, T.P.; Solt, L.A.; Wang, Y.; Crumbley, C.; Banerjee, S.; Griffett, K.; Lundasen, T.; Hughes, T.; Kojetin, D.J. Nuclear Receptors and Their Selective Pharmacologic Modulators. Pharmacol. Rev. 2013, 65, 710–778. [Google Scholar] [CrossRef]
  191. Raza, S.; Meyer, M.; Goodyear, C.; Hammer, K.D.P.; Guo, B.; Ghribi, O. The Cholesterol Metabolite 27-Hydroxycholesterol Stimulates Cell Proliferation via ERβ in Prostate Cancer Cells. Cancer Cell Int. 2017, 17, 52. [Google Scholar] [CrossRef] [PubMed]
  192. Lau, K.-M.; Laspina, M.; Long, J.; Ho, S.-M. Expression of Estrogen Receptor (ER)-and ER-in Normal and Malignant Prostatic Epithelial Cells: Regulation by Methylation and Involvement in Growth Regulation 1. Cancer Res. 2000, 60, 3175–3182. [Google Scholar] [PubMed]
  193. Bhattacharyya, R.S.; Krishnan, A.V.; Swami, S.; Feldman, D. Fulvestrant (ICI 182,780) down-Regulates Androgen Receptor Expression and Diminishes Androgenic Responses in LNCaP Human Prostate Cancer Cells. Mol. Cancer Ther. 2006, 5, 1539–1549. [Google Scholar] [CrossRef]
  194. Gasent Blesa, J.M.; Alberola Candel, V.; Giner Marco, V.; Giner-Bosch, V.; Provencio Pulla, M.; Laforga Canales, J.B. Experience with Fulvestrant Acetate in Castration-Resistant Prostate Cancer Patients. Ann. Oncol. 2010, 21, 1131–1132. [Google Scholar] [CrossRef] [PubMed]
  195. Chadha, M.K.; Ashraf, U.; Lawrence, D.; Tian, L.; Levine, E.; Silliman, C.; Escott, P.; Payne, V.; Trump, D.L. Phase II Study of Fulvestrant (Faslodex) in Castration Resistant Prostate Cancer. Prostate 2008, 68, 1461–1466. [Google Scholar] [CrossRef] [PubMed]
  196. Prossnitz, E.R.; Barton, M. The G Protein-Coupled Oestrogen Receptor GPER in Health and Disease: An Update. Nat. Rev. Endocrinol. 2023, 19, 407–424. [Google Scholar] [CrossRef]
  197. Rekha, P.; Gupta, A.; Goud, K.S.; Biswas, B.; Bhattar, S.; Vijayakumar, G.; Selvaraju, S. GPER Induces Mitochondrial Fission through P44/42 MAPK—Drp1 Pathway in Breast Cancer Cells. Biochem. Biophys. Res. Commun. 2023, 643, 16–23. [Google Scholar] [CrossRef]
  198. Kampa, M.; Lappano, R.; Grande, F.; Rizzuti, B.; Maggiolini, M.; Castanas, E.; Jacquot, Y. Promising Perspectives of the Antiproliferative GPER Inverse Agonist ERα17p in Breast Cancer. Cells 2023, 12, 653. [Google Scholar] [CrossRef]
  199. Pal, U.; Manjegowda, M.C.; Singh, N.; Saikia, S.; Philip, B.S.; Jyoti Kalita, D.; Kumar Rai, A.; Sarma, A.; Raphael, V.; Modi, D.; et al. The G-Protein-Coupled Estrogen Receptor, a Gene Co-Expressed with ERα in Breast Tumors, Is Regulated by Estrogen-ERα Signalling in ERα Positive Breast Cancer Cells. Gene 2023, 877, 147548. [Google Scholar] [CrossRef]
  200. Czogalla, B.; Partenheimer, A.; Jeschke, U.; von Schönfeldt, V.; Mayr, D.; Mahner, S.; Burges, A.; Simoni, M.; Melli, B.; Benevelli, R.; et al. β-Arrestin 2 Is a Prognostic Factor for Survival of Ovarian Cancer Patients Upregulating Cell Proliferation. Front. Endocrinol. 2020, 11, 554733. [Google Scholar] [CrossRef]
  201. Tran, C.; Ouk, S.; Clegg, N.J.; Chen, Y.; Watson, P.A.; Arora, V.; Wongvipat, J.; Smith-Jones, P.M.; Yoo, D.; Kwon, A.; et al. Development of a Second-Generation Antiandrogen for Treatment of Advanced Prostate Cancer. Science 2009, 324, 787–790. [Google Scholar] [CrossRef]
  202. Ricci, F.; Buzzatti, G.; Rubagotti, A.; Boccardo, F. Safety of Antiandrogen Therapy for Treating Prostate Cancer. Expert. Opin. Drug Saf. 2014, 13, 1483–1499. [Google Scholar] [CrossRef]
  203. Fizazi, K.; Massard, C.; Bono, P.; Jones, R.; Kataja, V.; James, N.; Garcia, J.A.; Protheroe, A.; Tammela, T.L.; Elliott, T.; et al. Activity and Safety of ODM-201 in Patients with Progressive Metastatic Castration-Resistant Prostate Cancer (ARADES): An Open-Label Phase 1 Dose-Escalation and Randomised Phase 2 Dose Expansion Trial. Lancet Oncol. 2014, 15, 975–985. [Google Scholar] [CrossRef] [PubMed]
  204. Wong, Y.N.S.; Ferraldeschi, R.; Attard, G.; De Bono, J. Evolution of Androgen Receptor Targeted Therapy for Advanced Prostate Cancer. Nat. Rev. Clin. Oncol. 2014, 11, 365–376. [Google Scholar] [CrossRef]
  205. Culig, Z. Targeting the Androgen Receptor in Prostate Cancer. Expert. Opin. Pharmacother. 2014, 15, 1427–1437. [Google Scholar] [CrossRef]
  206. Dalton, J.T.; Mukherjee, A.; Zhu, Z.; Kirkovsky, L.; Miller, D.D. Discovery of Nonsteroidal Androgens. Biochem. Biophys. Res. Commun. 1998, 244, 1–4. [Google Scholar] [CrossRef]
  207. Schmidt, A.; Meissner, R.S.; Gentile, M.A.; Chisamore, M.J.; Opas, E.E.; Scafonas, A.; Cusick, T.E.; Gambone, C.; Pennypacker, B.; Hodor, P.; et al. Identification of an Anabolic Selective Androgen Receptor Modulator That Actively Induces Death of Androgen-Independent Prostate Cancer Cells. J. Steroid Biochem. Mol. Biol. 2014, 143, 29–39. [Google Scholar] [CrossRef] [PubMed]
  208. Loddick, S.A.; Ross, S.J.; Thomason, A.G.; Robinson, D.M.; Walker, G.E.; Dunkley, T.P.J.; Brave, S.R.; Broadbent, N.; Stratton, N.C.; Trueman, D.; et al. AZD3514: A Small Molecule That Modulates Androgen Receptor Signaling and Function in Vitro and in Vivo. Mol. Cancer Ther. 2013, 12, 1715–1727. [Google Scholar] [CrossRef]
  209. Yu, Z.; Cai, C.; Gao, S.; Simon, N.I.; Shen, H.C.; Balk, S.P. Galeterone Prevents Androgen Receptor Binding to Chromatin and Enhances Degradation of Mutant Androgen Receptor. Clin. Cancer Res. 2014, 20, 4075–4085. [Google Scholar] [CrossRef] [PubMed]
  210. Kach, J.; Long, T.M.; Selman, P.; Tonsing-Carter, E.Y.; Bacalao, M.A.; Lastra, R.R.; De Wet, L.; Comiskey, S.; Gillard, M.; VanOpstall, C.; et al. Selective Glucocorticoid Receptor Modulators (SGRMs) Delay Castrate-Resistant Prostate Cancer Growth. Mol. Cancer Ther. 2017, 16, 1680–1692. [Google Scholar] [CrossRef]
  211. Taplin, M.E.; Manola, J.; Oh, W.K.; Kantoff, P.W.; Bubley, G.J.; Smith, M.; Barb, D.; Mantzoros, C.; Gelmann, E.P.; Balk, S.P. A Phase II Study of Mifepristone (RU-486) in Castration-Resistant Prostate Cancer, with a Correlative Assessment of Androgen-Related Hormones. BJU Int. 2008, 101, 1084–1089. [Google Scholar] [CrossRef] [PubMed]
  212. Labrie, F.; Bélanger, A.; Luu-The, V.; Labrie, C.; Simard, J.; Cusan, L.; Gomez, J.; Candas, B. Gonadotropin-Releasing Hormone Agonists in the Treatment of Prostate Cancer. Endocr. Rev. 2005, 26, 361–379. [Google Scholar] [CrossRef]
  213. Thakur, A.; Roy, A.; Ghosh, A.; Chhabra, M.; Banerjee, S. Abiraterone Acetate in the Treatment of Prostate Cancer. Biomed. Pharmacother. 2018, 101, 211–218. [Google Scholar] [CrossRef] [PubMed]
  214. Norris, J.D.; Ellison, S.J.; Baker, J.G.; Stagg, D.B.; Wardell, S.E.; Park, S.; Alley, H.M.; Baldi, R.M.; Yllanes, A.; Andreano, K.J.; et al. Androgen Receptor Antagonism Drives Cytochrome P450 17A1 Inhibitor Efficacy in Prostate Cancer. J. Clin. Investig. 2017, 127, 2326–2338. [Google Scholar] [CrossRef]
  215. Horton, J.; Rosenbaum, C.; Cummings, F.J. Tamoxifen in Advanced Prostate Cancer: An ECOG Pilot Study. Prostate 1988, 12, 173–177. [Google Scholar] [CrossRef] [PubMed]
  216. El-Arini, M.O. Response to Tamoxifen in Drug-Resistant Prostatic Carcinoma. Lancet 1979, 2, 588. [Google Scholar] [CrossRef]
  217. Tong, D. Selective Estrogen Receptor Modulators Contribute to Prostate Cancer Treatment by Regulating the Tumor Immune Microenvironment. J. Immunother. Cancer 2022, 10, e002944. [Google Scholar] [CrossRef]
  218. Semenas, J.; Wang, T.; Sajid Syed Khaja, A.; Firoj Mahmud, A.K.M.; Simoulis, A.; Grundström, T.; Fällman, M.; Persson, J.L. Targeted Inhibition of ERα Signaling and PIP5K1α/Akt Pathways in Castration-Resistant Prostate Cancer. Mol. Oncol. 2021, 15, 968–986. [Google Scholar] [CrossRef]
  219. Ho, T.H.; Nunez-Nateras, R.; Hou, Y.X.; Bryce, A.H.; Northfelt, D.W.; Dueck, A.C.; Wong, B.; Stanton, M.L.; Joseph, R.W.; Castle, E.P. A Study of Combination Bicalutamide and Raloxifene for Patients With Castration-Resistant Prostate Cancer. Clin. Genitourin. Cancer 2017, 15, 196–202.e1. [Google Scholar] [CrossRef] [PubMed]
  220. Lai, A.C.; Crews, C.M. Induced Protein Degradation: An Emerging Drug Discovery Paradigm. Nat. Rev. Drug Discov. 2017, 16, 101–114. [Google Scholar] [CrossRef]
  221. Barton, M. Position Paper: The Membrane Estrogen Receptor GPER--Clues and Questions. Steroids 2012, 77, 935–942. [Google Scholar] [CrossRef]
  222. Gasent Blesa, J.M.; Alberola Candel, V. PSA Decrease with Fulvestrant Acetate in a Hormone-Resistant Metastatic Prostate Cancer Patient. Onkologie 2010, 33, 57–59. [Google Scholar] [CrossRef] [PubMed]
  223. Féchon, A.; Droz, J.P. Do We Really Need New Trials on Fulvestrant in Prostate Cancer? Onkologie 2010, 33, 12–13. [Google Scholar] [CrossRef] [PubMed]
  224. Turner, N.C.; Slamon, D.J.; Ro, J.; Bondarenko, I.; Im, S.-A.; Masuda, N.; Colleoni, M.; DeMichele, A.; Loi, S.; Verma, S.; et al. Overall Survival with Palbociclib and Fulvestrant in Advanced Breast Cancer. N. Engl. J. Med. 2018, 379, 1926–1936. [Google Scholar] [CrossRef]
  225. Cristofanilli, M.; Turner, N.C.; Bondarenko, I.; Ro, J.; Im, S.A.; Masuda, N.; Colleoni, M.; DeMichele, A.; Loi, S.; Verma, S.; et al. Fulvestrant plus Palbociclib versus Fulvestrant plus Placebo for Treatment of Hormone-Receptor-Positive, HER2-Negative Metastatic Breast Cancer That Progressed on Previous Endocrine Therapy (PALOMA-3): Final Analysis of the Multicentre, Double-Blind, Phase 3 Randomised Controlled Trial. Lancet Oncol. 2016, 17, 425–439. [Google Scholar] [CrossRef] [PubMed]
  226. Márquez-Garbán, D.C.; Deng, G.; Comin-Anduix, B.; Garcia, A.J.; Xing, Y.; Chen, H.W.; Cheung-Lau, G.; Hamilton, N.; Jung, M.E.; Pietras, R.J. Antiestrogens in Combination with Immune Checkpoint Inhibitors in Breast Cancer Immunotherapy. J. Steroid Biochem. Mol. Biol. 2019, 193, 105415. [Google Scholar] [CrossRef]
  227. Slamon, D.J.; Neven, P.; Chia, S.; Fasching, P.A.; De Laurentiis, M.; Im, S.A.; Petrakova, K.; Val Bianchi, G.; Esteva, F.J.; Martín, M.; et al. Phase III Randomized Study of Ribociclib and Fulvestrant in Hormone Receptor-Positive, Human Epidermal Growth Factor Receptor 2-Negative Advanced Breast Cancer: MONALEESA-3. J. Clin. Oncol. 2018, 36, 2465–2472. [Google Scholar] [CrossRef]
  228. Giessrigl, B.; Schmidt, W.M.; Kalipciyan, M.; Jeitler, M.; Bilban, M.; Gollinger, M.; Krieger, S.; Jäger, W.; Mader, R.M.; Krupitza, G. Fulvestrant Induces Resistance by Modulating GPER and CDK6 Expression: Implication of Methyltransferases, Deacetylases and the HSWI/SNF Chromatin Remodelling Complex. Br. J. Cancer 2013, 109, 2751–2762. [Google Scholar] [CrossRef]
  229. Jung, J. Role of G Protein-Coupled Estrogen Receptor in Cancer Progression. Toxicol. Res. 2019, 35, 209–214. [Google Scholar] [CrossRef]
  230. Burslem, G.M.; Crews, C.M. Small-Molecule Modulation of Protein Homeostasis. Chem. Rev. 2017, 117, 11269–11301. [Google Scholar] [CrossRef]
  231. Negi, A.; Kesari, K.K.; Voisin-Chiret, A.S. Estrogen Receptor-α Targeting: PROTACs, SNIPERs, Peptide-PROTACs, Antibody Conjugated PROTACs and SNIPERs. Pharmaceutics 2022, 14, 2523. [Google Scholar] [CrossRef] [PubMed]
  232. Lin, X.; Xiang, H.; Luo, G. Targeting Estrogen Receptor α for Degradation with PROTACs: A Promising Approach to Overcome Endocrine Resistance. Eur. J. Med. Chem. 2020, 206, 112689. [Google Scholar] [CrossRef]
  233. Sakamoto, K.M.; Kim, K.B.; Verma, R.; Ransick, A.; Stein, B.; Crews, C.M.; Deshaies, R.J. Development of Protacs to Target Cancer-Promoting Proteins for Ubiquitination and Degradation. Mol. Cell. Proteom. 2003, 2, 1350–1358. [Google Scholar] [CrossRef] [PubMed]
  234. Gonzalez, T.L.; Hancock, M.; Sun, S.; Gersch, C.L.; Larios, J.M.; David, W.; Hu, J.; Hayes, D.F.; Wang, S.; Rae, J.M. Targeted Degradation of Activating Estrogen Receptor α Ligand-Binding Domain Mutations in Human Breast Cancer. Breast Cancer Res. Treat. 2020, 180, 611–622. [Google Scholar] [CrossRef]
  235. Savarese, D.M.; Halabi, S.; Hars, V.; Akerley, W.L.; Taplin, M.E.; Godley, P.A.; Hussain, A.; Small, E.J.; Vogelzang, N.J. Phase II Study of Docetaxel, Estramustine, and Low-Dose Hydrocortisone in Men with Hormone-Refractory Prostate Cancer: A Final Report of CALGB 9780. Cancer and Leukemia Group B. J. Clin. Oncol. 2001, 19, 2509–2516. [Google Scholar] [CrossRef]
  236. Puyang, X.; Furman, C.; Zheng, G.Z.; Wu, Z.J.; Banka, D.; Aithal, K.; Agoulnik, S.; Bolduc, D.M.; Buonamici, S.; Caleb, B.; et al. Discovery of Selective Estrogen Receptor Covalent Antagonists for the Treatment of ERαWT and ERαMUT Breast Cancer. Cancer Discov. 2018, 8, 1176–1193. [Google Scholar] [CrossRef] [PubMed]
  237. Furman, C.; Puyang, X.; Zhang, Z.; Wu, Z.J.; Banka, D.; Aithal, K.B.; Albacker, L.A.; Hao, M.H.; Irwin, S.; Kim, A.; et al. Covalent ERα Antagonist H3B-6545 Demonstrates Encouraging Preclinical Activity in Therapy-Resistant Breast Cancer. Mol. Cancer Ther. 2022, 21, 890–902. [Google Scholar] [CrossRef]
  238. Furman, C.; Hao, M.H.; Prajapati, S.; Reynolds, D.; Rimkunas, V.; Zheng, G.Z.; Zhu, P.; Korpal, M. Estrogen Receptor Covalent Antagonists: The Best Is Yet to Come. Cancer Res. 2019, 79, 1740–1745. [Google Scholar] [CrossRef]
  239. Berry, M.; Metzger, D.; Chambon, P. Role of the Two Activating Domains of the Oestrogen Receptor in the Cell-Type and Promoter-Context Dependent Agonistic Activity of the Anti-Oestrogen 4-Hydroxytamoxifen. EMBO J. 1990, 9, 2811–2818. [Google Scholar] [CrossRef]
  240. Pagliuca, M.; Donato, M.; D’Amato, A.L.; Rosanova, M.; Russo, A.O.M.; Scafetta, R.; De Angelis, C.; Trivedi, M.V.; André, F.; Arpino, G.; et al. New Steps on an Old Path: Novel Estrogen Receptor Inhibitors in Breast Cancer. Crit. Rev. Oncol. Hematol. 2022, 180, 103861. [Google Scholar] [CrossRef]
  241. Min, J.; Nwachukwu, J.C.; Min, C.K.; Njeri, J.W.; Srinivasan, S.; Rangarajan, E.S.; Nettles, C.C.; Guillen, V.S.; Ziegler, Y.; Yan, S.; et al. Dual-Mechanism Estrogen Receptor Inhibitors. Proc. Natl. Acad. Sci. USA 2021, 118, e2101657118. [Google Scholar] [CrossRef] [PubMed]
  242. Sareddy, G.R.; Nair, B.C.; Gonugunta, V.K.; Zhang, Q.G.; Brenner, A.; Brann, D.W.; Tekmal, R.R.; Vadlamudi, R.K. Therapeutic Significance of Estrogen Receptor β Agonists in Gliomas. Mol. Cancer Ther. 2012, 11, 1174–1182. [Google Scholar] [CrossRef] [PubMed]
  243. Schüler-Toprak, S.; Häring, J.; Inwald, E.C.; Moehle, C.; Ortmann, O.; Treeck, O. Agonists and Knockdown of Estrogen Receptor β Differentially Affect Invasion of Triple-Negative Breast Cancer Cells in Vitro. BMC Cancer 2016, 16, 951. [Google Scholar] [CrossRef]
  244. Ruddy, S.C.; Lau, R.; Cabrita, M.A.; McGregor, C.; McKay, B.C.; Murphy, L.C.; Wright, J.S.; Durst, T.; Pratt, M.A.C. Preferential Estrogen Receptor β Ligands Reduce Bcl-2 Expression in Hormone-Resistant Breast Cancer Cells to Increase Autophagy. Mol. Cancer Ther. 2014, 13, 1882–1893. [Google Scholar] [CrossRef]
  245. Schüler-Toprak, S.; Moehle, C.; Skrzypczak, M.; Ortmann, O.; Treeck, O. Effect of Estrogen Receptor β Agonists on Proliferation and Gene Expression of Ovarian Cancer Cells. BMC Cancer 2017, 17, 319. [Google Scholar] [CrossRef]
  246. Warner, M.; Huang, B.; Gustafsson, J.A. Estrogen Receptor β as a Pharmaceutical Target. Trends Pharmacol. Sci. 2017, 38, 92–99. [Google Scholar] [CrossRef]
  247. Honma, N.; Horii, R.; Iwase, T.; Saji, S.; Younes, M.; Takubo, K.; Matsuura, M.; Ito, Y.; Akiyama, F.; Sakamoto, G. Clinical Importance of Estrogen Receptor-Beta Evaluation in Breast Cancer Patients Treated with Adjuvant Tamoxifen Therapy. J. Clin. Oncol. 2008, 26, 3727–3734. [Google Scholar] [CrossRef]
  248. Sareddy, G.R.; Li, X.; Liu, J.; Viswanadhapalli, S.; Garcia, L.; Gruslova, A.; Cavazos, D.; Garcia, M.; Strom, A.M.; Gustafsson, J.A.; et al. Selective Estrogen Receptor β Agonist LY500307 as a Novel Therapeutic Agent for Glioblastoma. Sci. Rep. 2016, 6, 24185. [Google Scholar] [CrossRef] [PubMed]
  249. Wu, W.F.; Wang, L.; Spetsieris, N.; Boukovala, M.; Efstathiou, E.; Brössner, C.; Warner, M.; Gustafsson, J.A. Estrogen Receptor β and Treatment with a Phytoestrogen Are Associated with Inhibition of Nuclear Translocation of EGFR in the Prostate. Proc. Natl. Acad. Sci. USA 2021, 118, e2011269118. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Representation of ERα and ERβ sex-specific tissue distribution. The picture is not indicative of absolute receptor expression levels.
Figure 1. Representation of ERα and ERβ sex-specific tissue distribution. The picture is not indicative of absolute receptor expression levels.
Cancers 15 04653 g001
Figure 2. Schematic structures of estrogen receptors’ (ERs) isoforms. Different functional domains are highlighted: the N-terminal transactivation domain NTD/AF-1 (A/B), the DNA-binding domain (C), a hinge region (D), and the C-terminal cofactor interaction domain AF-2 and a ligand-binding domain (LBD) (E/F).
Figure 2. Schematic structures of estrogen receptors’ (ERs) isoforms. Different functional domains are highlighted: the N-terminal transactivation domain NTD/AF-1 (A/B), the DNA-binding domain (C), a hinge region (D), and the C-terminal cofactor interaction domain AF-2 and a ligand-binding domain (LBD) (E/F).
Cancers 15 04653 g002
Figure 3. ERs can act through genomic and non-genomic pathways in PCa. In the genomic pathway, estrogen triggers the dimerization of ERs and their nuclear translocation to regulate the transcriptional activity of genes. ERs can directly control the transcription of genes bearing the estrogen response element (ERE) in their promoter region. In addition, ERs can modulate gene transcription through protein−protein interactions with other transcription factors (TF), which can mediate the recruitment of ERs at promoters without EREs. In the non-genomic pathway, membrane-associated ERs can participate in signal transduction and modulate the activation of key enzymes and molecular pathways, thus resulting in the indirect regulation of transcriptional programs. Known ER-regulated transcriptional programs in PCa: (I) ERα: transactivation of proliferative, apoptotic, survival, differentiation, vasodilation, autophagy, and inflammation targets; transactivation of oncogenes and non-coding RNAs. (II) ERβ1 transactivation of apoptotic, anti-proliferative, anti-invasive, and anti-inflammatory targets; repression of AR, oncogenes, and hypoxic genes. (III) ERβ2 and ERβ5—stabilization of the HIF-1α protein and induction of hypoxic genes. Created with BioRender.com, accessed on 10 July 2023.
Figure 3. ERs can act through genomic and non-genomic pathways in PCa. In the genomic pathway, estrogen triggers the dimerization of ERs and their nuclear translocation to regulate the transcriptional activity of genes. ERs can directly control the transcription of genes bearing the estrogen response element (ERE) in their promoter region. In addition, ERs can modulate gene transcription through protein−protein interactions with other transcription factors (TF), which can mediate the recruitment of ERs at promoters without EREs. In the non-genomic pathway, membrane-associated ERs can participate in signal transduction and modulate the activation of key enzymes and molecular pathways, thus resulting in the indirect regulation of transcriptional programs. Known ER-regulated transcriptional programs in PCa: (I) ERα: transactivation of proliferative, apoptotic, survival, differentiation, vasodilation, autophagy, and inflammation targets; transactivation of oncogenes and non-coding RNAs. (II) ERβ1 transactivation of apoptotic, anti-proliferative, anti-invasive, and anti-inflammatory targets; repression of AR, oncogenes, and hypoxic genes. (III) ERβ2 and ERβ5—stabilization of the HIF-1α protein and induction of hypoxic genes. Created with BioRender.com, accessed on 10 July 2023.
Cancers 15 04653 g003
Figure 4. Anti-estrogen molecules’ mode of action. Androgens are the substrate for estrogen production, mediated by the aromatase (CYP19A1) enzyme. Estrogens bind and activate ERs, thus regulating target genes’ transcription. The synthesis of estrogens may be blocked by specific inhibitors. SERDs are ER antagonists that bind the receptor, block its migration to the nucleus, and inhibit receptor-ERE interaction. SERD-ER complexes are degraded by the proteosome. PROTACs are bispecific ligands binding both ER and the E3 ubiquitin ligase, which, in turn, mediates the ubiquitination of the complex, leading to its degradation. SERM competitively binds ERs. The SERM-ER complex can interact with ERE sequences and co-repressors, inhibiting the transcription of target genes. However, this effect is tissue-specific and co-activators may be recruited in certain tissues, such as the endometrium and bone. CERANs exert antagonistic action by blocking AF1 and AF2 transcriptional activation domains via co-repressor recruitment. SERCAs covalently bind to an ER cysteine residue (C530), inhibiting the receptor activation and gene transcription. Selective ERβ agonists may be used to upregulate anti-proliferative signals in prostate cancer cells.
Figure 4. Anti-estrogen molecules’ mode of action. Androgens are the substrate for estrogen production, mediated by the aromatase (CYP19A1) enzyme. Estrogens bind and activate ERs, thus regulating target genes’ transcription. The synthesis of estrogens may be blocked by specific inhibitors. SERDs are ER antagonists that bind the receptor, block its migration to the nucleus, and inhibit receptor-ERE interaction. SERD-ER complexes are degraded by the proteosome. PROTACs are bispecific ligands binding both ER and the E3 ubiquitin ligase, which, in turn, mediates the ubiquitination of the complex, leading to its degradation. SERM competitively binds ERs. The SERM-ER complex can interact with ERE sequences and co-repressors, inhibiting the transcription of target genes. However, this effect is tissue-specific and co-activators may be recruited in certain tissues, such as the endometrium and bone. CERANs exert antagonistic action by blocking AF1 and AF2 transcriptional activation domains via co-repressor recruitment. SERCAs covalently bind to an ER cysteine residue (C530), inhibiting the receptor activation and gene transcription. Selective ERβ agonists may be used to upregulate anti-proliferative signals in prostate cancer cells.
Cancers 15 04653 g004
Table 1. The expression of ER isoforms and their roles in PCa.
Table 1. The expression of ER isoforms and their roles in PCa.
IsoformKDaExpression in ProstateFunction in Prostate CancerRefs.
ESR1 gene: Estrogen Receptor-α
ERα-66
(full-length, often referred to
as ERα)
66–67Highly expressed in tumor stroma;
↑ expression in epithelium in PCa;
↑ hg-PIN;
↑ high Gleason score (GS) tumors;
↑↓ (?) CRPC.
Tumor-promoting.[100,101,104,105,106,116,117,118]
ERα-4646Expressed in normal and malignant
prostate tissue.
Not described in PCa.[128,129]
ERα-3636Expressed in normal and malignant
prostate tissue.
Not described in PCa.[102,130,131,132]
ESR2 gene: Estrogen Receptor-β
ERβ1
(full-length, often referred to
as ERβ)
59–60Mostly expressed in prostate epithelial cells;
↓ hg-PIN;
↓ localized tumors;
↓ CRPC;
↓ high Gleason score (GS) tumors.
Tumor-suppressive.[74,101,137,143,144,145]
ERβ255–56Predominantly in the cytoplasm of prostate epithelial cells;
↑ in PCa, especially in the nucleus;
↑ metastatic PCa.
Tumor-promoting;
promotes stem cell properties and the development of chemoresistance.
[137,143,144,145,146]
ERβ356Not expressed in normal and malignant prostate cells.Not applicable.[134,143]
ERβ454Expressed in normal and malignant prostate cells.Not described in PCa;
heterodimerizes with ERβ1 and enhances its transcriptional activity (from yeast two-hybrid and promoter luciferase assays).
[135,147]
ERβ553Expressed in basal epithelial cells in benign prostate glands;
↑ metastatic PCa.
Tumor-promoting;
promotes stem cell properties and the development of chemoresistance.
[135,137,143,145,146]
↑ increased expression; ↓ decreased expression; (?) contrasting results.
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

Belluti, S.; Imbriano, C.; Casarini, L. Nuclear Estrogen Receptors in Prostate Cancer: From Genes to Function. Cancers 2023, 15, 4653. https://doi.org/10.3390/cancers15184653

AMA Style

Belluti S, Imbriano C, Casarini L. Nuclear Estrogen Receptors in Prostate Cancer: From Genes to Function. Cancers. 2023; 15(18):4653. https://doi.org/10.3390/cancers15184653

Chicago/Turabian Style

Belluti, Silvia, Carol Imbriano, and Livio Casarini. 2023. "Nuclear Estrogen Receptors in Prostate Cancer: From Genes to Function" Cancers 15, no. 18: 4653. https://doi.org/10.3390/cancers15184653

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